You are on page 1of 29

In situ sequenced bioremediation of mixed

contaminants in groundwater
J.F. Devlin
a,
*
, D. Katic
b
, J.F. Barker
c
a
Geology Department, University of Kansas, 120 Lindley Hall, Lawrence, KS 66045-7613, USA
b
Golder Associates, Mississauga, Ontario, Canada
c
Department of Earth Sciences, University of Waterloo, Waterloo, Ontario, Canada
Received 29 August 2002; accepted 9 July 2003
Abstract
A mixture of chlorinated solvents (about 0.510 mg/l), including tetrachloroethene (PCE) and
carbon tetrachloride (CT), together with a petroleum hydrocarbon, toluene (TOL), were introduced
into a 24 m long 2 m wide 3 m deep isolated section (henceforth called a gate) of the Borden
aquifer and subjected to sequential in situ treatment. An identical section of aquifer was similarly
contaminated and allowed to self-remediate by natural attenuation, thus serving as a control. The
control presents a rare opportunity to critically assess the performance of the treatment systems, and
represents the first such study for sequenced in situ remediation. The first treatment step was
anaerobic bioremediation. This was accomplished using a modified nutrient injection wall (NIW) to
pulse benzoate and a nutrient solution into the aquifer, maximizing mixing by dispersion and
minimizing fouling near the injection wells. In the anaerobic bioactive zone that developed, PCE, CT
and chloroform (CF), a degradation product of CT, degraded with a half-lives of about 59, 5.9 and
1.7 days, respectively. The second step was aerobic bioremediation, using a biosparge system. TOL
and cis-1,2 dichloroethene (cDCE), from PCE degradation, were found to degrade aerobically with
half-lives of 17 and 15 days, respectively. Compared to natural attenuation, PCE and TOL removal
rates were significantly better in the sequenced treatment gate. However, CT and CF were similarly
and completely attenuated in both gates. It is believed that the presence of TOL helped sustain the
reducing environment needed for the reduction of these two compounds.
D 2003 Elsevier B.V. All rights reserved.
Keywords: Sequenced treatment; Bioremediation; In situ remediation; Natural attenuation; Chlorinated solvents;
Hydrocarbons; Anaerobicaerobic bioremediation
0169-7722/$ - see front matter D 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0169-7722(03)00156-6
* Corresponding author. Fax: +1-785-864-5276.
E-mail address: jfdevlin@ku.edu (J.F. Devlin).
www.elsevier.com/locate/jconhyd
Journal of Contaminant Hydrology 69 (2004) 233261
1. Introduction
The treatment of mixtures of organic contaminants in water has been the subject of study
and practice for many years (Fogel et al., 1982). With the growing acceptance and
implementation of in situ technologies for groundwater remediation, there is an opportunity
to further develop and adapt these systems to the problem of treating mixed contaminant
plumes in the ground. The primary goal of this research was to evaluate the performance of
an in situ groundwater treatment system consisting of two sequenced technologies, the first
anaerobic and the second aerobic. Toward this end, an artificial plume of carbon
tetrachloride (CT, 12 mg/l), tetrachloroethene (PCE, 12 mg/l) and toluene (TOL, 1
10 mg/l) was created and treated inside an isolated section of aquifer subjected to a constant
rate of groundwater flow. The treatment system was evaluated against an identical,
adjacently located section of aquifer, in which natural attenuation was permitted to occur.
Numerous investigators have suggested that sequenced in situ treatment of groundwater
should be beneficial (Zitomer and Speece, 1993; Guest et al., 1993; Dickel et al., 1993;
Nelson et al., 1994; Knackmuss, 1996; Agrawal and Tratnyek, 1996; Hofstetter et al.,
1999). Attempts have been made, or are in progress, to implement systems of this kind
(Tratnyek and Johnson, 1998; Lizotte et al., 1999; Barcelona and Xie, 2001). To our
knowledge, there was only one complete and controlled field test of such a system in the
literature at the time of writing (Morkin et al., 2000). In that experiment, an anaerobic
aerobic treatment sequence was engineered by arranging a granular iron reactive barrier
upgradient from a biosparge unit, within a Funnel-and-Gate structure (Starr and Cherry,
1994). The contaminants of concern included trichloroethene (TCE, 0.09 mg/l), cis-
dichloroethene (cDCE, 198 mg/l), vinyl chloride (VC, 48 mg/l), benzene (0.21 mg/l), TOL
(6.3 mg/l) and traces ( < 0.5 mg/l) of other chlorinated and aromatic volatile organic
substances. The granular iron barrier reduced the concentrations of chlorinated compounds
between 91% and >99%, where cDCE and TCE were the respective end-members of this
range. The biosparge zone removed about 66% of the remaining cDCE and VC from the
groundwater. At the end of 8 months of operation, the system successfully lowered
contaminant concentrations below the target levels, which were based on the Californias
maximum contaminant levels (MCLs). Thus, there is evidence that sequenced in situ
treatment is viable, but additional work is needed to explore alternative technologies in
such sequences.
Avariety of in situ technologies for groundwater cleanup have been introduced over the
past two decades. These fall into two classes: passive and active technologies. The passive
technologies, such as permeable reactive barriers (PRB), attempt to maximize the benefit
of low operation and maintenance costs for long-term operation. In these systems, there is
little or no pumping, and treatment is ensured either by the installation of a suitably large
PRB, or through the use of a passive flow-control structure, such as a Funnel-and-Gate
(Starr and Cherry, 1994) or a Trench-and-Gate (Bowles et al., 2000). An example of a PRB
is the granular iron barrier, first proposed by Gillham and OHannesin (1992) for the
treatment of chlorinated organics, and later used for the treatment of a variety of other
contaminants (chromium, Pratt et al., 1997; uranium, Morrison et al., 2001; arsenic, Su
and Puls, 2001; nitroaromatics, Devlin et al., 1998; nitrate, Alowitz and Scherer, 2002) and
used in the sequenced system mentioned above (Morkin et al., 2000).
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 234
The active technologies place more emphasis on lower capital costs and, arguably,
greater control over the system hydraulics (achieved by pumping). Enhanced bioremedi-
ation has generally fallen into the category of active in situ technologies, since it often
involves continuous pumping or recirculation of water. Traditionally, the recirculation has
taken place between conventional wells dedicated to either extraction or injection (Norris
et al., 1994). However, recirculation systems using single boreholes have also been tested
and found to be promising in at least one stratified aquifer (McCarty et al., 1998).
Although both passive and active methods have advantages in different situations, it was
decided that this research would focus on the passive technologies. In order to extend our
experience with treatment sequences beyond granular iron PRBs for anaerobic treatment,
anaerobic bioremediation was chosen for inclusion in this experiment.
An approach to bioremediation that minimizes pumping while maximizing dispersive
mixing was tested in the CFB Borden aquifer and reported by Devlin and Barker (1994,
1996, 1999) and Devlin and Muller (1999). Briefly, the method involves the injection of a
nutrient or primary substrate solution from a nutrient injection wall (NIW) at regular
intervals determined by the average linear groundwater velocity. The solution pulses are
timed to that they overlap due to hydrodynamic dispersion at some predetermined distance
(on the order of a few metres) from the injection wells. Where the overlap occurs,
subsurface bacteria receive a continuous supply of nutrients/substrate, and a bioactive zone
is sustained. Although the NIW reported by Devlin and Barker (1996) was successful in
creating and sustaining a modified anaerobic redox zone in which CT was transformed, it
was unsuccessful at stimulating the reduction of TCE (Devlin, 1994; Devlin and Muller,
1999). A challenge for this study was to achieve chlorinated ethene reduction with the
NIW technique.
The biodegradation of aromatic hydrocarbons (e.g. TOL) and lesser chlorinated ethenes
(e.g. cDCE, VC) in aerobic environments is known to occur relatively rapidly (Norris et
al., 1994; Bradley and Chapelle, 1998). Moreover, there are several methods currently
available for introducing oxygen to the subsurface to support aerobic biodegradation.
These include air sparging, sometimes called biosparging (Norris et al., 1994; Morkin et
al., 1998), the use of an oxygen releasing metal peroxide compound (ORC; Regenesis
Bioremediation Products, CA) (Borden et al., 1997; Chapman et al., 1997; Landmeyer et
al., 2001) and oxygen introduction through diffusive emitters (Wilson et al., 2002). At the
time of this experiment, the oxygen diffusive emitters were under development and not
available for use. ORC was tested as part of a sister project (Barker et al., 2000). Therefore
a variation of biosparging was selected as the means of aerating the aquifer downgradient
of the NIW (Barker et al., 1993). Sparging was conducted only to emplace or renew
residual oxygen bubbles in the porous medium, with minimal excess gas delivery. The
residual oxygen bubbles dissolved into the flowing groundwater, providing a readily
available electron acceptor.
An inevitable consequence of sequenced anaerobic and aerobic treatment systems is
that a distinct geochemical boundary is created at the transition between the two systems.
Morkin et al. (2000) did not observe any performance limiting effects at the boundary.
However, their sequence included a limestone gravel buffer zone separating the two
extreme environments. The current design contained no such buffer zone. A further goal of
this work, then, was to evaluate any effects of the anaerobicaerobic boundary on the
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 235
performance of the system, with emphasis on the possible flow interferences that might
result.
2. Site description
The study site has been described previously (Devlin et al., 2002). Briefly, the
experiment was conducted in a pristine, well-sorted, sandy aquifer located at the Canadian
Forces Base Borden (CFB Borden). The aquifer was 3 m deep and underlain by a 6-m-
thick aquitard which was determined to be an effective hydraulic barrier beneath the
aquifer. Borden sand contains a significant carbonate component that tends to buffer the
pH of the groundwater to near neutral values. It contains low organic carbon
(foc = 0.0002), and a porosity of about 0.33 (Ball et al., 1990).
The water table varied between 0.3 and 1.5 m below ground surface (bgs), depending
on time of year and location. A sheet pile wall bounded the test section on three sides,
creating an alleyway measuring 2 m wide 24 m long, and hereafter referred to as the gate
(Fig. 1). An identical alleyway was constructed immediately next to the current one, and
served as a control; no treatment measures were attempted in that section of the aquifer
(Devlin et al., 2002). The gates were covered with a 30 m long 9.2 m wide greenhouse
structure consisting of steel tube framing and a flame retardant, vinyl-coated polyester
cover to protect the equipment and instrumentation from the elements. The structure also
facilitated sampling throughout the year and prevented direct local recharge due to rainfall
or snowmelt, which might have influenced the hydraulics.
The groundwater flow inside the treatment gate was maintained by continuous
pumping of a fully screened 5.08 cm (2 in.) well, located at the closed end of the
structure. The pumping rate was maintained at 130 ml/min. The uniformity of the flow
was assessed by a bromide tracer experiment tracked with fences of multilevel monitoring
Fig. 1. Schematic of the treatment gate.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 236
wells (see Section 4). Average linear groundwater velocities were determined to be in the
range of 415 cm/day, expressed as fence averages. Avalue of 14 cm/day was considered
most representative of conditions in the gate and was used in all calculations.
3. Experimental design
As alluded to above, the test section of aquifer was hydraulically separated from the bulk
of the aquifer by sheet piling (Fig. 1). The gate was enclosed on three sides with a pumping
well at the closed end; the fourth side was left open to permit native groundwater to enter
the test section. By isolating the test section in this way, consistency of flow could be
assured so that contaminant transformations could be interpreted with minimal ambiguity.
Three test compounds, CT, PCE and TOL, were introduced to the groundwater stream at the
open end of the gate using diffusive emitters (see below). This design permitted a long-term
source that could be discontinued without excavation at the conclusion of the experiment.
The target concentrations for the contaminants were 110 mg/l for TOL, and 12 mg/l for
PCE and CT. Previous work established that the retardation factors for these compounds in
the Borden aquifer were about 1.2, 2.7 and 1.2, respectively (Pankow and Cherry, 1996;
Devlin, 1994; Hubbard, 1992). In addition, chloroform (CF) formed from the transforma-
tion of CT; it was previously found to have a retardation factor of about 1.1 in the Borden
aquifer (Devlin, 1994). Also, cDCE was produced from PCE degradation and was
estimated to have a retardation factor of 1.1, based on its reported Koc, and the retardation
behaviour of PCE (Mercer and Cohen, 1990).
A NIW configured to introduce benzoate and inorganic nutrient pulses to the aquifer
was installed immediately downgradient from the source to create an 8-m-long anaerobic
zone for reductive dechlorination. The amendment pulses were added at intervals of 20
30 days to support a bioactive zone about 4 m downgradient of the NIW wells. At the
outset of the experiment, benzoate was the sole ingredient in the amendment solution. It
was prepared as a concentrated solution of sodium benzoate, 140 g benzoate/kg solution,
and added directly to the recirculating water from the NIW, in-line. Once mixed with the
groundwater, the benzoate in the NIW was diluted to about 250 mg/l. On day 151 (after
contaminant additions began), the benzoate concentrations were reduced to 125160 mg/
l without adversely affecting the anaerobic zone. On day 207 and for each injection
afterwards, the amendment solution was supplemented with nutrients consisting of a
mixture of approximately 10 mg/l of K
2
HPO
4
, KH
2
PO
4
and NH
4
NO
3
, as well as 2 mg/l of
MgSO
4
7H
2
O.
The interval between amendment pulses was determined using velocity and dispersion
parameters estimated from a bromide tracer pulse included in one of the benzoate pulses.
Using these transport parameters, a 1D solution to the advectiondispersion equation
(Devlin and Barker, 1996) was used to estimate the dimensions of a pulse that had been
transported 5 m from the NIW. The 5-m distance corresponded to the target location for
the beginning of the anaerobic bioactive zone. A pulsing period of 28 days was determined
appropriate from this analysis.
A residual oxygen barrier (ROB) was installed downgradient of the NIW. By placing
the aerobic treatment zone after the anaerobic treatment, partially dechlorinated inter-
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 237
mediates of PCE, such as cDCE and VC, could be further degraded at the same time as
TOL. Moreover, the aesthetics of the water emerging from the gate would be better after
aeration.
Sampling was conducted using multilevel wells to provide detailed vertical and lateral
data for the assessment of contaminant fate. Details of the installation and configuration
are provided below. The experiment was active for a period of about 1 year.
4. Methods
4.1. Site instrumentation
Waterloo BarrierR sealable joint sheet piling was driven into the ground and keyed into
the aquitard to break any hydraulic connection between the inside of the gate and the
surrounding aquifer. The joints were filled with grout, following installation, as a further
precaution against contaminant leakage out of the test volume. The integrity of the seals
was assessed throughout the experiment, and found to be satisfactory, by regular sampling
of monitoring wells located immediately outside the gate.
The organic test compounds, CT, PCE and TOL, were released from diffusive emitter
devices (Barker et al., 2000; Arildskov and Devlin, 2000). These devices have the
advantage of creating continuous sources for whatever period of time is required. When
the source is no longer needed, it can be stopped with little time and trouble. The emitters
were installed in four fully screened 20-cm-diameter PVC wells located at the open end of
the gate. Briefly, they were designed to continuously release about 1 mg/l of each
contaminant (CT, PCE, TOL) into the aquifer.
The emitters performed adequately for the test, although during the time period
between days 75 and 150, PCE, CT and TOL input concentrations (as fence 302 averages)
rose unexpectedly. In general, PCE was introduced at approximately 200300 Ag/l, except
briefly near day 150 when inputs rose to 600 Ag/l. CT was introduced at about 200600
Ag/l, except between days 100 and 150 when the concentration rose to 2000 Ag/l. TOL was
introduced at about 4001000 Ag/l, except between days 75 and 110 when concentrations
rose to about 10,000 Ag/l. Despite these difficulties, concentrations further along the flow
path (fence 306), upgradient of any transformations, remained quite steady (PCEf100
140 Ag/l; TOLf70140 Ag/l; CT not assessed at 306 due to transformation to CF),
indicating that the source anomalies were of minor importance to the plume. Further
details of the source emitter performance can be found in Barker et al. (2000) and Devlin
et al. (2002). Improvements to the emitter design were reported by Arildskov and Devlin
(2000).
Multilevel monitors were installed across the gate, in fences, at distances of 0.8, 9.29,
11.37, 13.08, 17.29, 18.66, 19.3, 20.4 and 22.65 m from the source wells, and were
designated G302, G310, G312, G315, G318, G320, G321, G322 and G324, respectively
(Fig. 1). Each fence consisted of three monitors designated L, M or R depending on
whether the device was located on the left middle or right side of the gate (looking
downgradient). Each multilevel was constructed from four 0.3-cm outer diameter (OD)
stainless steel (ss) tubes strapped to a central 1.25-cm OD PVC stalk. The ss tubes
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 238
terminated at depths of 1.2, 1.9, 2.6 and 3.3 m bgs, designated levels 1, 2, 3 and 4,
respectively. No filter packs were used in the completion of the multilevel monitors. Each
monitoring point was completed with a fine ss mesh wrapped around the ends to prevent
invasion by the native sand.
About 5 m into the gate, the NIW was installed using the materials and methods
described in Barker et al. (2000) and Devlin and Barker (1999); the NIW section was
excavated, instrumented and backfilled with a coarse filter sand. Three injection and
withdrawal wells were installed in two rows (Fig. 2). All were fully screened PVC
construction, and 5.08 cm in diameter. Nutrient amendments were accomplished using a
closed loop dosing system as described in Devlin and Barker (1996). However, in contrast
to the previous NIW, nutrient pulses were emplaced by flushing the barrier in the direction
of ambient groundwater flow, rather than perpendicular to it. In a gate setting (as opposed
to an open aquifer), this configuration was preferred since it reduced the effects of
asymmetric nutrient distribution, due to flow convergence at the pumping wells.
The ROB, located about 16 m into the gate, consisted of six horizontal sparge wells
installed at the base of the aquifer. Each device was perforated in six evenly spaced
Fig. 2. Schematic of the NIW.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 239
locations along the horizontal tubes. The wells were oriented perpendicular to the flow
direction and were equally spaced over a 2.8-m-long section of the gate (Fig. 3). Each
sparge well contained four 0.62-cm-diameter polyethylene tubes terminating at different
distances along the horizontal pipe. The tubes were pressurized simultaneously to ensure
that oxygen delivery occurred as uniformly as possible. Additional details of the
construction of the sparge system are given in Barker et al. (2000). The sparge wells
were overlain by a gravel fill with a porosity of 0.46. The upper 30 cm of the ROB
section was backfilled with pea gravel (grain diameters 0.51.4 cm) along the central
axis of the gate and filter sand along the sides. The unsaturated zone, to a depth of 22
cm, was monitored with four 0.3-cm inside diameter (ID) ss soil gas sampling tubes
equipped with mininert caps. These were installed in the pea gravel along the central
axis of the gate. The ROB was covered at surface with a plastic tarp to minimize the
escape of volatile compounds during sparging. Escape laterally was controlled by
installing partially penetrating sheet piles (extending to 1 m bgs) on either end of the
ROB.
The horizontal sparge wells were constructed from 5.1-cm outer diameter (OD) PVC
pipes installed vertically, three on each side of the gate, from the surface to the bottom of
the ROB where they were connected to same sized horizontal pipes extending across the
gate with six vent holes each. Inside each pipe were four 0.62-cm OD polyethylene tubes
running from surface to corresponding vent holes. At surface, the tubes were connected to
a manifold that was itself connected to a cylinder of oxygen. Gas injections were
performed at regular intervals using a household digital timer from Noma (model
Fig. 3. Schematic of the ROB.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 240
N1507) to control a two-way, direct lift diaphragm solenoid valve (from ColePalmer)
that was installed between the oxygen cylinder and the manifold.
Test sparging was completed once the ROB was backfilled with gravel up to the level
of the water table. These tests indicated that a 1-min sparge at 45 psi of gauge pressure
was sufficient to establish a bubble residual over about 60% of the barrier and to begin gas
transfer to the unsaturated zone. Based on this, and a gas filled porosity of 4% of the
sparged aquifer volume (determined in laboratory experiments (Barker et al., 2000)), it
was estimated that each 1-min sparge released a maximum of about 400 l of oxygen (3.0 m
deep 2.0 m wide 2.8 m long 0.04 porosity 0.6 coverage; note: 3 m deep corre-
sponds to the water table at surfacerepresenting the maximum saturated aquifer
thickness), corresponding to 22 mol. With a benzoate loading of about 0.46 mol/day (a
maximum of 300 mg/l entering the gate on average at 130 ml/min) and the following
stoichiometry for aerobic benzoate consumption
C
7
H
5
O

2 aq

15
2
O
2 aq
H

aq
!7CO
2 aq
3H
2
O
l
it was calculated that (22 mol O
2
per sparge/0.46 mol benzoate per day/(15/2)=) 6.5 days
between sparges would be the interval necessary to meet the oxygen demand. A
conservative semiweekly sparging schedule (3.5-day interval) was adopted at the outset
of the experiment; as discussed below, this was later modified to daily, minute-long
sparges.
Immediately next to the gate was a second three-sided sheet pile alleyway of identical
dimensions and similarly instrumented with monitoring wells. However, this structure
contained no NIW or ROB, but served instead as a control so that degradation resulting
from bioremediation could be compared to degradation due to natural attenuation
processes. Details of the control gate construction and performance are described
elsewhere (Devlin et al., 2002).
4.2. Sampling methods
Sampling was conducted for a variety of solutes and geochemical parameters.
Complete details of the protocols and QAQC are reported in Barker et al. (2000). Briefly,
all aqueous samples were collected from the multilevel points using either a peristaltic
pump or a 60-ml syringe for suction; the shallow water table and permeable aquifer
material contributed to minimal degassing of samples. The monitors were purged of a
volume equivalent to three times the standing water volume of the deepest point (57 ml)
prior to collecting samples. Dissolved organic solute samples, including volatile organics
compounds (VOCs) and organic acids, were collected in 40-ml screw cap glass vials with
Teflon-lined septa. The vials were placed in line before the pump to avoid contact with
pump-head tubing.
Inorganic samples were stored in polypropylene bottles; cations were filtered with 0.45-
Am membranes and acidified with nitric acid (10 drops per 60 ml bottle) to achieve a pH of
about 2. Anions were collected unfiltered and unacidified. Dissolved oxygen concen-
trations were measured either with an Orion model 835 meter and a model 083010 probe
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 241
(for concentrations >0.5 mg/l), or using the Chemetrics colorimetric method (for concen-
trations < 0.5 mg/l).
Hydrogen was measured using the method of Chapelle et al. (1997). Three samples
were collected from any given location and the results averaged for later interpretation.
4.3. Analytical methods
Nonchlorinated hydrocarbon gases were analysed using an HP 5840A gas chromato-
graph (GC) equipped with a 2-ml sample loop, a 30-m Megabore GS-Q capillary column
with helium flowing at 12 ml/min serving as the carrier gas and an oven temperature of
100 jC. The injector temperature was 100 jC and the detector, a flame ionization detector,
was kept at 200 jC. VOCs were analysed using an HP 5890 GC equipped with a
photoionization detector and a Varian Genesis headspace autosampler.
Inorganic anions were analyzed commercially by Water Technology International
(WTI), Burlington, Ontario, using a Waters Ion Chromatograph equipped with a Waters
Series 590 pump, a WISP UK Rheodyne Sample Injector, an IC-Pak A HR column, a 430
conductivity detector and a Spectra Physics SP4290 integrator. Cations were also analyzed
at WTI, by inductively coupled plasma (ICP).
Method detection limits for VOCs were estimated from calibration curves as
described by Devlin (1996). They were under 10 Ag/l for all compounds identified in
this study. Precision was estimated by comparing the analyses of duplicate pairs. In most
cases, precision was in the range of F1525% (including methane). The exceptions
were PCE ( F30%) and carbon tetrachloride ( F75%). In the latter case, the high
percentage uncertainty is reflective of the low concentrations in the duplicates collected
during the QAQC program, rather than the overall quality of the analyses. Even so,
since carbon tetrachloride transformed rapidly and was not present in most of the gate,
the uncertainty experienced in this compounds analysis was not an obstacle to
interpretation.
Inorganic anions and major cations all had method detection limits of 1 mg/l or less,
and were analyzed with precision of F10% or less, except for chloride ( F15%) and
dissolved oxygen (DO, F40% at concentrations less than 0.5 mg/l). Since the
concentrations of chlorinated compounds introduced to the gate were quite low in
comparison to background chloride, there was little chance for a chloride mass balance
even with a very sensitive and precise analytical technique. Therefore, chloride was
examined for site characterisation purposes only, and the precision experienced was
considered acceptable. DO levels below 0.5 mg/l were considered indicative of bulk
anaerobic conditions; no attempt was made to use the measured DO levels for more
quantitative purposes.
4.4. Data reduction
Mass balance calculations were performed using the FORTRAN program FENCE
(Barker et al., 2000) to integrate breakthrough curves for point specific breakthrough
masses, and then adding the masses to obtain totals for each fence. Flux estimates were
obtained by dividing these mass totals by the cross-sectional area (6 m
2
) and the number of
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 242
days over which the masses accumulated. Average linear velocities were estimated from
the arrival times of the centres of mass of the breakthrough curves.
5. Results and discussion
5.1. Tracer test
The tracer mass at the various monitoring fences (Fig. 1) was found to balance to
within F30%, in agreement with a similar tracer test conducted in the control gate (Devlin
et al., 2002). Velocities were estimated to be in range of 414 cm/day depending on the
sampling point examined. A subsequent tracer test, in which bromide was released from
the NIW, indicated that the average velocity was closer to 15 cm/day during the period
data were collected from the anaerobic treatment zone. A value of 14 cm/day was used in
the calculations presented here.
The bromide tracer test was used to characterise longitudinal dispersion and velocity
(14 cm/day) downgradient of the NIW. This was accomplished by averaging instantaneous
concentrations at fence G312, and fitting the resulting averaged breakthrough curve with a
1D solution of the advectiondispersion equation (Devlin and Barker, 1996). The fence-
averaged dispersivity was found to be 0.21 m, corresponding to relatively low dispersion
in the gate. In fact, degradation rate constants calculated from equations that included
dispersion and those from equations that neglected dispersion were less than 10% apart
(Devlin et al., 2002). This justified a simplified analysis of the data and the omission of
dispersion terms where convenient. Although, in this case, dispersion could be neglected
in the estimation of degradation rate constants, it was still occurring and still important for
the proper operation of the NIW.
5.2. Redox control using the NIW and ROB
Changes in redox conditions in various parts of the gate were tracked by monitoring
several redox sensitive parameters including DO, Fe(II), SO
4
2
, H
2 (aq)
and Eh. Prior to
beginning the experiment, the aquifer was characterised as being aerobic, with background
DO of 27 mg/l (Devlin et al., 2002). Nitrate was not detected in the groundwater and iron
concentrations were below 0.5 mg/l. Sulphate was present at concentrations of 1520 mg/
l and Eh measurements were in the range of 50300 mV. However, traces of methane
were detected in samples (maximum of 36 Ag/l) suggesting the presence of microenviron-
ments that were more reducing than the bulk of the aquifer. In addition, hydrogen
concentrations measured in background wells yielded values of 0.2 nM, suggesting that
iron reducing conditions prevailed.
By day 220, there were signs in the deepest parts of the aquifer, corresponding to
multilevel depths 3 and 4 (2.53.3 m bgs, respectively), that conditions had become more
reducing as a result of the benzoate additions. Sulphate was no longer detectable at these
lower depths and iron concentrations had increased from < 0.5 mg/l to an average of about
1 mg/l at G312. Moreover, methane concentrations increased to between 1000 and 2000
Ag/l in selected locations (e.g., G312M2, G312M4, G315M4) between G312 and G315.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 243
Additional sampling on day 339 indicated that the trend toward increasingly reducing
conditions between G310 and G315 was continuing. Fence-averaged iron concentrations
rose to between 1 and 2.5 mg/l while corresponding sulphate concentrations declined to
< 1 mg/l (averaging all depths). Methane production increased noticeably with concen-
trations between 6000 and 8000 Ag/l in the deepest parts of the aquifer (3.3 m bgs) and
10002000 Ag/l at level 2 (1.9 m bgs). Between days 332 and 334, hydrogen measure-
ments were made yielding concentrations between 0.4 and 5.3 nM, the highest measured
during the experiment. The mean hydrogen concentration between G310 and G315 was
3.0 nM, suggestive of sulphate reducing conditions, but the increasing methane concen-
trations indicated that methanogenesis dominated some zones within the aquifer, partic-
ularly at depth.
The groundwater in the ROB was initially aerobic, as described above. Between days 5
and 240 the ROB was sparged semiweekly. During this time, DO levels in the gravel fill
varied from below detection to 6 mg/l. On day 238, gas flow rate measurements were
made between the oxygen tank and the wells. It was discovered that each sparge was only
introducing 35 l of oxygen to the gravel fill, far less than the maximum 400 l anticipated in
the design. Increasing sparge pressure was ruled out as a remedy because it was feared that
this might cause large channels to form in the gravel, reducing the volume of porous
medium receiving oxygen bubbles (Ji et al., 1993; Austrins, 1997). Instead, it was decided
to increase the sparging frequency to once daily, for 1 min. Once this procedure was
adopted, on day 240, dissolved oxygen concentrations in the groundwater increased to
between 10 and 40 mg/l at most monitored points in the ROB. Exceptions included the
deepest monitoring points (levels 4) where DO was between 2 and 10 mg/l (note that these
points were located below the elevation of the sparge nozzles in the horizontal wells) and
at monitor G316R, all depths, where DO levels were found in the range 112 mg/l. On the
basis of these data, the ROB section was determine to be aerobic for the last 100 days of
the experiment.
Downgradient of the ROB, oxygen was rapidly depleted due to the presence of
benzoate from the NIW, TOL in the plume, and acetate produced in the anaerobic
treatment zone through fermentation of the two preceding compounds. Hydrogen levels in
the aquifer downgradient of the ROB were consistently 0.10.7 nM, suggesting an iron
reducing environment. This finding was consistent with an increase in Fe(II) concen-
trations (from < 0.5 to about 1 mg/l) in the groundwater sampled from G318 to G320
between days 220 and 238.
5.3. NIW treatment of VOCs
The test contaminants, CT, PCE and TOL, were released from the source wells and
permitted to migrate through the NIW into the anaerobic treatment zone. CT was rapidly
transformed to CF before the NIW, so for all practical purposes, CF became a source
compound rather than CT. This rapid transformation was also observed in the control gate
(Devlin et al., 2002).
Benzoate and, beginning on day 207, inorganic nutrient additions were pulsed into the
aquifer an average of once every 25 days (Katic, 1999) and the procedure required about
5 h to complete (0.21 days). So, the NIW system operated passively 99% of the time.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 244
Residence times in the NIW varied from 22 days (R= 1.2, estimated for TOL and cDCE)
to 50 days (R=2.7, estimated for PCE), assuming a groundwater velocity of 14 cm/day.
Thus, the more mobile compounds, TOL and CF, were likely recirculated at least once
during pulsing before passing into the anaerobic treatment zone while PCE may have
been recirculated twice, assuming similar mobility in the filter sand and the aquifer sand.
The recirculation apparatus was fashioned in a closed loop, so losses due to volatilisation
were considered unlikely; DO concentrations remained unchanged in the NIW before and
after injections, further indicating limited contact between the groundwater and the
atmosphere. Sorptive losses to pipes and flowmeters are likely to have been minor after
the first injection because the flow rates were high (several gallons per minute) and
sorptive uptake by the plastics used in construction would be expected to slow with
increased contact time (Parker and Ranney, 2000; Arildskov and Devlin, 2000). The
effect of recirculation on maximum contaminant concentrations is thought to be primarily
one of dilution due to mixing with water from less contaminated zones, with the result
being an increased uniformity of contaminant concentrations throughout the NIW. For
this reason, the analyses that follow will be limited to data from and downgradient of the
NIW.
In order to anticipate the progress of the contaminants as they were transported
through the gate, a 1D solution to the advectiondispersion equation was used that
accounted for linear equilibrium retardation; reactions were not considered in these
simulations (Fig. 4). Transport parameters were taken from the results of the preceding
tracer tests. This preliminary modeling indicated that by day 238, the mobile compounds,
TOL and CF, should have arrived at the pumping well, while PCE would have arrived at
the ROB. By day 339, TOL and CF were predicted to be present throughout the gate at
input levels, while PCE was anticipated to be near input levels in the ROB, and just
beginning to break through at the pumping well. Comparing these predictions to the
fence-averaged data from the experiment, it can be seen that TOL levels ( f0.9 mg/l)
were indeed quite consistent between the NIW and ROB on day 238 (samples were not
yet routinely collected beyond the ROB at this time) and PCE appeared to arrive at the
ROB at this time, in accordance with predictions (Fig. 5). No PCE degradation products
(e.g., cDCE, TCE) were detected in the groundwater between days 1 and 238, suggesting
that up to this time PCE degradation was minimal. On the other hand, CF concentrations
in the experimental gate were considerably less extensive than predicted. On day 112,
CF had been detected at the monitoring fence 11.5 m from the source wells in
concentrations in excess of 100 Ag/l, but by day 238, CF was absent from the anaerobic
zone. Since the source wells remained active during this time, these data indicate that CF
underwent degradation in the anaerobic treatment zone. No chlorinated methanes other
than CF or CT were identified in the groundwater, suggesting that complete dechlori-
nation was occurring.
By day 339, PCE had not measurably advanced, and may even have declined in
concentration within the anaerobic treatment zone (Fig. 5). At the same time, cDCE
appeared upgradient and within the ROB, and TCE was detected in trace quantities ( < 5
Ag/l) upgradient of the ROB (data not shown). TOL levels declined slightly in the
anaerobic zone between days 238 and 339, and both CT and CF remained below detection
at all but one sampling point (G310R, 184 Ag/l CF). These data suggest that while TOL
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 245
degradation was minimal in response to the NIW amendments, both CF and PCE were
being reductively dechlorinated at the end of the experiment.
The first day cDCE was detected in the treatment gate was day 316. At that time, it was
present in highest concentrations (4070 Ag/l at several monitoring points) at G315,
immediately upgradient of the ROB. The compound was also detected at monitoring
fences downgradient of the ROB, including G318 ( f30 Ag/l at several points). Since
cDCE is a reduction product of PCE, it must have been formed in the anaerobic section of
the gate and migrated into the aerobic section. G315 and G318 are about 6 m apart,
Fig. 4. Calculated relative positions of PCE and TOL on days 238 and 339, assuming an average groundwater
velocity of 15 cm/day, and a dispersivity of 0.2 m. Retardation factors were 1.2 for TOL and cDCE and 2.7 for
PCE.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 246
corresponding to a travel time of about 48 days (assuming R
cDCE
=1.2, as in the NIW
section). So, if cDCE first appeared at G318 on day 316, then it must have been present at
G315 (316 48 = 268) on about day 270. Given that PCE first arrived at G315 around day
196, without cDCE, the lag time for reductive dechlorination was on the order of
270 196 =74 days. This must be considered a minimum since PCE may have been
present at G315 before day 196, at some depths and locations in the aquifer.
5.4. Anaerobic degradation rates
In order to quantify the disappearance rates of CT, CF, TOL and PCE, concentration
profiles and flux calculations were performed. Because CT and CF transformed so quickly
Fig. 5. Fence-averaged concentrations of TOL, PCE and cDCE in the treatment gate on days 238 and 339.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 247
in the gate, there were insufficient data available from which to calculate time integrated
fluxes. Therefore, degradation rate estimates of these two compounds were based on
concentration profiles. CT and CF concentrations at fences G302, G310 and G312 were fit
to the following equations which neglect dispersion and assume pseudo-first-order
transformation kinetics (Devlin et al., 2002) (Fig. 6).
C C
o
exp
k
v
x

1
D
FkC
o
k k
exp
k
v
x

exp
k
v
x

2
where C= concentration of CT (M/L
3
), C
o
= input concentration of CT (M/L
3
), D= con-
centration of CF (M/L
3
), k = pseudo-first-order rate constant for CT degradation (T
1
),
k = pseudo-first-order rate constant for CF degradation (T
1
), F= conversion factor that
accounts for reaction stoichiometry (assumed 1 for these calculations) and differences in
molecular weights between parent and daughter compounds when data are reported in mg
rather than mmol ( = W
CF
/W
CT
= 119.377 g/mol/153.822 g/mol = 0.77) (dimensionless) and
x = distance from source wells (L). The transformations were largely complete by fence
G310, so the profile data used to fit the equations were sparse. The actual profiles may be
less advanced than represented here and therefore the actual rate constants may be higher
than the estimates. Eq. (1) was first fit to the CT data to obtain C
o
and k and then Eq. (2)
Fig. 6. Profiles of CTand CF on day 339. Solid circles are observed, fence-averaged CT concentrations, and open
diamonds are observed, fence-averaged CF concentrations. Lines represent modelled fits using Eqs. (1) and (2)
(see text).
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 248
was fit to CF data to obtain k (Table 1). The estimated rate constants correspond to half-
lives of 5.9 and 1.7 days for CT and CF, respectively. The CT degradation rate was
comparable with that observed by Devlin and Muller (1999) (t
1/2
=37 days) in a separate
experiment at the same site. However, Devlin and Muller did not observe CF degradation
in their experiment. This difference in CF persistence was likely due to the differences in
experimental design. Devlin and Muller introduced CT (which rapidly transformed to CF
and carbon disulphide) into the aquifer as a single pulse. The CF was not in contact with a
given section of aquifer long enough for microbial adaptation to occur and so it was not
degraded (note: there were a small number of locations, corresponding to finer grained
lenses, where methylene chloride was detected in that experiment. Thus, some CF
transformation did occur, but it was minimal). In the current test, the contaminants were
injected continuously. We speculate that this led to more adaptation time and a more
uniformly distributed and complete degradation of CF.
To assess transformation rates for PCE and TOL, fluxes of these compounds in the
anaerobic treatment zone were calculated for fences G306, G310, G312 and G315 for days
154339 (G306) and 196339 (all other fences). Previous work in the control gate
demonstrated the utility of this method, and indicated that flux-based rate constant
estimates were likely to be as representative, or more so than those from profile analysis
(Devlin et al., 2002). As described in Barker et al. (2000) and Devlin et al. (2001, 2002),
the equations to estimate a pseudo-first-order rate constant from flux data, neglecting
dispersion, are
lnJ lnJ
o

k
v
x 3
J
daughter
J
parent
o
F
1
k k
kexp
k
v
x

kexp
k
v
x

exp
k
v
x

4
where J = flux of solute (M/L
2
/T) at a point with distance x from the source, and J
o
= input
flux of solute (M/L
2
/T). Eq. (3) was applied to TOL and PCE data (Fig. 7), while Eq. (4)
Table 1
Summary of properties rate constants and half-lives for contaminant degradation in the NIW, ROB and by natural
attenuation
Compound H NIW k NIW t
1/2
x ROB
profile k
ROB
profile t
1/2
ROB
flux k
ROB
flux t
1/2
Nat.
Att. k
Nat.
Att. t
1/2
CT 0.12 5.9 0.063 11
CF 0.41 1.7 0.069 10
PCE 0.012 59
TOL 0.26 1.75e 03 3.38e 02 20.5 4.08e 02 17.0 0.011 63
cDCE 0.074 5.05e 04 6.97e 02 9.9 4.51e 02 15.4
METH 22.27 5.29e 02 0
H reported for approximately 10 jC, cDCE value from Gossett (1987), TOL value from Mackay et al. (1999), and
METH value from Wilhem et al. (1977).
Sparging rate was 3.58 10
2
m
3
day
1
, the gas filled porosity was estimated to be 3.6 10
2
, and the volume
of the ROB was 12.3 m
3
in the calculations. = k not distinguishable from 0.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 249
was modified, as described below, and used with cDCE data. No significant degradation of
TOL was observed in the anaerobic zone, while PCE was found to be degrading with a
half-life of between 42 and 104 days (range represents F1 standard error on the rate
constant) with a best fit value of 59 days (Table 1).
The expected degradation sequence for PCE is PCE!TCE!cDCE!VC!non-
chlorinated hydrocarbons (VC refers to vinyl chloride). However, the application of Eq.
(4) to cDCE assumes a degradation sequence of PCE!cDCE!other unidentified
products. Since TCE was only present in trace quantities ( < 5% of the total mass), it
seemed reasonable to treat it as a short-lived intermediate, so the simplified sequence is
justified. Also, since VC was not detected at any time in the gate, it may be
conservatively assumed that cDCE did not further transform, i.e., k = 0. With this
simplification, Eq. (4) becomes
J
daughter
J
parent
o
F 1 exp
k
v
x

5
where F is the stoichiometric factor 1 multiplied by the value W
cDCE
/W
PCE
=(96.943 g/
mol)/(165.833 g/mol) = 0.585, if PCE degrades along the assumed sequence. It should
be noted that PCE is known to degrade along at least one additional sequence that
involves the formation of chloroacetylene and acetylene as intermediates (Roberts et
al., 1996; Semadeni et al., 1998). Therefore, F could be used as a fitting parameter to
assess the likelihood that PCE was degrading according to a branched pathway as
opposed to a linear one. This possibility is explored below.
The cDCE flux calculated from measured data came to 1.3 mg/m
2
/day (Fig. 8). From
Fig. 7, J
o
PCE
was estimated to be 3.72 mg/m
2
/day at G306 (marking the beginning of the
Fig. 7. TOL and PCE flux as a function of distance in the anaerobic section of the treatment gate. Note
logarithmic scale. Field flux values were determined in mg/m
2
/day. Solid lines are best fits and broken lines
indicate standard error intervals.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 250
anaerobic treatment zone) and k/v = 7.78 10
2
m
1
. At G315, the only fence where
fluxes of cDCE could be calculated, x = 8.15 m from G306. If it is assumed that F= 0.585
(linear pathway), then from Eq. (5), J
cDCE
= 1.0 mg/m
2
/day. The agreement between the
calculated value and the independently determined value, plotted in Fig. 8, is considered
quite good, and supports the hypothesis that PCE degraded primarily to cDCE up to day
339 when the experiment ended.
5.5. ROB treatment of VOCs
Although the DO concentrations were not maintained at the levels desired between
days 1 and 240, as described in Section 5.2, an examination of the TOL profile for
day 238 (Fig. 5) suggested that some aerobic degradation was taking place. This
process was apparently enhanced by day 339, since TOL concentrations in and
downgradient of the ROB declined between days 238 and 339. TOL was in
competition for DO with benzoate, left over from the NIW amendments, and acetate
which formed from benzoate fermentation (Barker et al., 2000). It is believed that a
reduction in benzoate inputs at the NIW could have improved TOL degradation
without seriously affecting the anaerobic treatment. Further work is needed to assess
this possibility.
Aerobic activity was also indicated by the limited penetration of cDCE into the
aerobic treatment zone (Fig. 9). However, since the ROB was sparged regularly, there
remains some question about the role of volatilisation in the declining concentrations of
volatile compounds. In order to assess the rate of mass loss due to volatilisation during
sparging, gas samples were collected from the soil gas monitors installed in the
unsaturated gravel of the ROB. No contaminants were ever detected in these samples.
To further assess the importance of volatilisation, theoretical calculations were performed
Fig. 8. Predicted and calculated fluxes for PCE and cDCE. Open circles are observed values of PCE fluxes; solid
diamond is the observed cDCE flux at G315. Lines represent model fits.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 251
beginning with the following mass balance which applies to a continuously sparged,
well-mixed batch reactor:
M
w;t
M
g;t
M
w;tDt
M
g;tDt
M
g out
6
where M
w,t
=mass of compound in the groundwater at time t, M
w,t Dt
=mass in the
water at time t Dt, M
g,t
= mass in the residual gas phase at time t, M
g,t Dt
= mass in the
residual gas phase at t Dt, M
g out
= mass volatilized due to sparging. Expanding each
term in Eq. (6), invoking Henrys law (C
g
= HC
w
), and expressing the result in
differential form,
dC
w
dt

Q
g
H
h
w
V
tot
Hh
g
V
tot
C
w
7
where C
w
= concentration at time t (M/L
3
), H= Henrys law constant (dimension-
less = C
g
/C
w
), C
g
= concentration in the gas (M/L
3
), V
w
= h
w
V
tot
= volume of water
(L
3
), V
g
= h
g
V
tot
= volume of gas (L
3
), h
w
= water filled porosity (dimensionless), h
g
= gas
gas filled porosity (dimensionless), V
tot
= aquifer volume (L
3
) and Q
g
= gas flow rate (L
3
/
T). Eq. (7) shows that the loss of volatile compounds due to sparging can be
approximated as a first-order kinetic process with rate constant, x, equal to
x
Q
g
H
h
w
V
tot
Hh
g
V
tot
:
Fig. 9. Fence-averaged concentrations of cDCE at G315, immediately upgradient of the ROB, and G318,
immediately downgradient of the ROB. The dissimilar trends suggest attenuation across the ROB.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 252
This first-order decay process can be added to the steady-state advectiondispersion
reaction equation to account for water flow through the ROB.
BC
Bt
0 D
B
2
C
Bx
2
v
BC
Bx
kC xC 8
If dispersion is once again neglected, the solution to Eq. (8) is a modified version of
Eq. (1).
C C
o
exp
k x
v
x

9
Theoretical values for x are presented in Table 1, and model profiles were fit to the
fence-averaged data for day 339 of the experiment (Fig. 10). C
o
was assumed to be the
fence-averaged concentration at G315, and x was measured from G315. The sole fitting
parameter for this exercise was k, the pseudo-first-order biodegradation rate constant.
Profiles were fit to three compounds: TOL, cDCE and methane (METH). The latter
substance was a by-product of the NIW treatment and was included in this analysis
because of its high volatility. If losses due to sparging were significant for any of the
compounds studied, it should be most significant for METH. The magnitudes of k for
the various compounds are discussed in the next section. For our purposes, it is
sufficient to note that when k = 0 in Eq. (9), neither TOL nor cDCE showed significant
concentration declines of importance through the ROB (Fig. 10).
On the other hand, METH was predicted to decline noticeably due to volatilisation, and
the METH profile can be entirely explained by volatilisation alone. However, the
modeling assumed perfect stripping efficiency and continuous gas flow; it seems unlikely
that volatilisation could have occurred in so ideal a fashion. Therefore, the mass losses (of
all three compounds) due to volatilisation are likely to be over-predicted. This has little
effect on the interpretation of the TOL or cDCE data, but some component of the METH
loss may in fact be due to consumption by bacteria.
5.6. Aerobic degradation rates
The profiles were fit with Eq. (9) and pseudo-first-order biodegradation rate constants
(k) were calculated (Table 1). TOL and cDCE were found to be degrading with half-lives
of about 20 and 10 days, respectively. Due to the high Henrys law constant of METH and
moderate scatter in the profile data (Fig. 10), biodegradation could not be quantitatively
distinguished from volatilisation for this compound.
In addition to the profile data, biodegradation rates were determined from contaminant
fluxes through the gate (Table 1). Fluxes of TOL and cDCE were calculated for the period
after day 238 (once the sparging was occurring daily) and fit with Eq. (3) (Fig. 11). The
TOL degradation rates compare quite well between the two methods used, with half-lives
in the range 1720 days, suggesting that near steady-state conditions did exist for this
compound. The agreement for cDCE degradation rates is somewhat less remarkable, with
a flux derived half-life of 15 days compared with a profile half-life of 10 days. Since the
flux values are time-integrated, from day 238 to 339, while the profile data correspond to
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 253
Fig. 10. Profiles of TOL, cDCE and METH across and downgradient of the ROB on day 339. Broken line
indicates predicted profile if volatilisation were the dominant attenuation process. Solid lines represent best fits
using Eq. (9). Symbols are fence-averaged concentrations.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 254
day 339 only, the disparity might be an indication that cDCE was being degraded more
quickly at the end of the experiment than at earlier times.
5.7. Performance assessment and comparison to natural attenuation
This experiment was conducted immediately next to an identical gate in which no
treatment measures were undertaken. This provided a rare opportunity to compare an in
situ treatment scheme against natural attenuation under virtually identical conditions. A
glance at the distributions of contaminants on day 339 of the experiment reveals that in
both cases CT and CF were attenuated in relatively short travel distances. While CT
degradation occurred almost immediately in both gates, the onset of CF degradation was a
few months earlier in the treatment gate. Differences between experimental and control
half-life estimates are probably not significant given the limited number of data points
available for the treatment gate estimates (Fig. 12).
Perhaps the biggest differences between the two gates was apparent for PCE and TOL,
both of which migrated the length of the control gate, but which were significantly
attenuated in the treatment gate. For example, on day 339, the flux of TOL at the end of
the control gate was 13 mg/m
2
/day compared to only 2.8 mg/m
2
/day in the treatment gate.
PCE underwent degradation in both gates, as indicated by the presence of its daughter
products, TCE and cDCE. However, in the control, both TCE and cDCE were only
detected in trace quantities ( < 30 Ag/l), and the extent of degradation process was too
limited to quantify (Devlin et al., 2002). In contrast, PCE was found to degrade
anaerobically in the treatment gate with a half-life of about 60 days, and there were
Fig. 11. TOL and cDCE flux as a function of distance from G315 (near the beginning of the ROB). Symbols
represent observed data; solid lines are best fits using Eq. (3) and broken lines indicate the standard error range.
Fluxes were calculated with the units mg/m
2
/day.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 255
indications that the degradation rate had increased with time (based on ever increasing
levels of cDCE at G315). A mass balance calculation, and supporting flux calculations,
indicated that TOL also degraded in the control, with a half-life of about 60 days (Devlin
et al., 2002). In the treatment gate, TOL was found to degrade aerobically with a half-life
of about 20 days. These results represent notable improvements in TOL and PCE
degradation compared with natural attentuation.
Two degradation products of PCE were identified in the gates: TCE and cDCE. The
former compound never exceeded 15 Ag/l in either gate, and the latter compound never
exceeded 30 Ag/l in the control gate. However, cDCE was detected at 289 Ag/l at one point
(G312M1) in the treatment gate, and the fence-averaged concentrations were increasing in
the anaerobic treatment zone, at G315 (Fig. 9), at the end of the experiment. Despite the
build-up of cDCE in the anaerobic part of the treatment gate, there was significant
degradation of the substance in the aerobic section, as planned. The result was that very
little cDCE arrived at the end of the gate, and indications were that the degradation rate for
this compound was increasing at the time the experiment ended.
Despite the promising outcome of this experiment, additional work is needed to
improve the efficiency of removal of PCE and TOL. To illustrate this, it can be shown
that for 99% PCE removal within the treatment gate (between G306 and G315), a
degradation half-life of about 8.5 days would be required, and for TOL to similarly
degrade (between G316 and G324) would require a half-life of about 8 days. The
Fig. 12. Snapshots of the control (natural attenuation) and treatment gates on day 339. Maximum concentrations
at each monitoring well are contoured.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 256
estimated half-lives here were about 60 and 20 days, respectively. It should be noted that
the PCE degradation rate estimate may be conservatively low, since no PCE advance was
noted in the treatment gate between days 238 and 339, an interval over which there should
have been a clear invasion of the ROB by PCE. Also, TOL degradation may have been
more rapid if benzoate inputs had been scaled back. It should also be noted when the
experiment ended, there were indications that PCE and cDCE degradation rates were
increasing, so the goals of 8 day half-lives may yet be attainable.
5.8. Anaerobicaerobic boundary
By sequencing anaerobic and aerobic treatment systems, possible interferences might
be expected. In particular, chemical precipitates, such as iron oxides, might be expected to
accumulate at the transition boundary. There is also the possibility of biofouling at the
boundary, due to the accumulation of aerobic bacterial biomass in response to the chemical
disequilibrium that exists there. If pore clogging occurred and compromised the treatment
Fig. 13. Hydraulic gradients through the treatment gate at the beginning and end of the experiment. The
consistently linear gradient indicates that no hydraulically significant fouling occurred in the anaerobicaerobic
transition zone over the duration of the experiment.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 257
gate performance, it would most likely manifest itself as a hydraulic barrier. In a passive
system (no pumping well) such a barrier would divert water around the gate. In the system
investigated here, where a pumping well ensured flow through the gate, the effect of the
barrier would be to create an hydraulic head drop across the redox transition zone.
Hydraulic head measurements were made routinely during the experiment (Fig. 13).
The hydraulic gradients were generally quite reproducible and the trends in hydraulic head
along the treatment gate were remarkably consistent over the entire 339 days of the
experiment. There is no indication in these data that clogged pores at the anaerobic
aerobic boundary caused flow problems. Nevertheless, these data represent coarse
measurements made over a short duration experiment ( f1 year), and so they must be
considered preliminary. Additional work is planned to investigate this potential problem at
the pore scale. Further work is also needed to assess the longer term effects on
performance.
6. Conclusions
Based on the work presented above, it may be concluded that sequenced bioremedi-
ation is an advantageous and technically viable method of treating mixed contaminant
plumes. The validity of this conclusion is particularly well illustrated by the formation of
cDCE in the anaerobic zone, followed by its degradation in the aerobic zone. However, it
is also significant that the benzoate additions, with nutrient supplements, stimulated the
degradation of PCE, and TCE to form cDCE at all. Previous work in the same aquifer
failed to achieve chlorinated ethene transformations.
The treatment of mixed contaminants was successfully demonstrated with TOL. TOL
was found to degrade with a half-life of about 60 days by anaerobic natural attenuation
(Devlin et al., 2002). This process appeared to be inhibited in the anaerobic environment
created here. However, TOL degradation rates were tripled, compared to natural attenu-
ation, in the aerobic treatment zone.
Brief daily sparging for the emplacement of residual oxygen bubbles had no significant
effect on the aqueous concentrations of TOL or cDCE. The same conclusion is expected to
apply to other volatile compounds with Henrys constants of similar magnitude. However,
this amount of sparging appears to be sufficient to create and sustain an aerobic
environment in which the target compounds can be degraded. Further effort is needed
to properly balance the supply of carbon in the NIW section with the oxygen demand in
the ROB section.
There was little advantage of the sequenced treatment over natural attenuation for the
chlorinated methanes. Both CT and CF were completely attenuated over about the same
distances in the treatment and control gates. These transformations occurred upstream of
the benzoate additions in the NIW, and were also observed in the control gate where
benzoate was altogether absent. This result suggests that TOL was serving as an electron
donor for the degradation of CT and possibly CF. However, it was noted that the lag time
before CF degradation began was several weeks shorter in the treatment gate than the
control, suggesting that the added benzoate was the dominant electron acceptor where it
was available.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 258
Both TOL and PCE degraded more rapidly in the treatment gate than by natural
attenuation. Also, in both cases, the degradation rates were apparently increasing when the
experiment ended, so the maximum rates have yet to be determined.
Finally, there was no evidence of hydraulic interferences due to pore clogging, despite a
sharp transition form highly anaerobic conditions to aerobic conditions. Further work is
needed to examine the longer term performance of these coupled technologies and the
transition zone between them.
Acknowledgements
Financial support for this research was provided by the Advanced Alternative
Technology Demonstration Facility (AATDF), funded by the Department of Defense and
administered by Rice University. Additional support was provided by the NSERC/
Motorola/ETI Industrial Research Chair in Groundwater Remediation, and the Solvents in
Groundwater Consortium, University of Waterloo. T. Anderson of the University of
Massachusetts assisted with hydrogen measurements. The authors gratefully acknowledge
the comments, discussions and technical support given by M. McMaster, S. Froud, J.
Rosen, B. Butler, R.W. Gillham and S. Fiorenza.
References
Agrawal, A., Tratnyek, P.G., 1996. Reduction of nitro aromatic compounds in Fe
0
CO
2
H
2
O systems: impli-
cations for groundwater remediation with iron metal. Environmental Science and Technology 30 (1), 153160.
Alowitz, M.J., Scherer, M.M., 2002. Kinetics of nitrate, nitrite, and Cr(VI) reduction by iron metal. Environ-
mental Science and Technology 36 (3), 299306.
Arildskov, N.P., Devlin, J.F., 2000. Field and laboratory evaluation of a diffusive emitter for semipassive release
of PCE to an aquifer. Ground Water 38 (1), 129138.
Austrins, C.D., 1997. Enhanced in situ bioremediation of hydrocarbon contaminated groundwater using oxygen
and ammonia gas injection in a funnel and gate scheme. MSc Thesis, Department of Earth Sciences, Uni-
versity of Waterloo, Waterloo, Ontario, Canada.
Ball, W.P., Buehler, C.H., Harmon, T.C., Mackay, D.M., Roberts, P.V., 1990. Characterization of a sandy aquifer
material at the grain scale. Journal of Contaminant Hydrology 5, 253295.
Barcelona, M.J., Xie, G., 2001. In situ lifetimes and kinetics of a reductive whey barrier and an oxidative ORC
barrier in the subsurface. Environmental Science and Technology 35 (16), 33783385.
Barker, J.F., Weber, J., Fyfe, J.S., Devlin, J.F., Mackay, D.M., Cherry, J.A., 1993. Efficient addition technology
for in situ bioremediation. Proceedings of the GasRep Symposium, September, Quebec City. 12 pp.
Barker, J.F., Butler, B.J., Cox, E., Devlin, J.F., Focht, R., Froud, S.M., Katic, D.J., McMaster, M., Morkin, M.,
Vogan, J., 2000. In: Fiorenza, S., Oubre, C., Ward, H. (Eds.), Sequenced Reactive Barriers for Groundwater
Remediation. Lewis Publishers, Boca Raton, FL. 730 pp.
Borden, R.C., Goin, R.T., Kao, C.M., 1997. Control of BTEX migration using a biologically enhanced permeable
barrier. Ground Water Monitoring and Remediation 17 (1), 7080.
Bowles, M.W., Bentley, L.R., Hoyne, B., Thomas, D.A., 2000. In situ ground water remediation using the trench
and gate system. Ground Water 38 (2), 172181.
Bradley, P.M., Chapelle, F.H., 1998. Effect of contaminant concentration on aerobic microbial mineralization of
DCE and VC in stream-bed sediments. Environmental Science and Technology 32 (5), 553557.
Chapman, S.W., Byerley, B.T., Smyth, D.J.A., Mackay, D.M., 1997. A pilot test of passive oxygen release for
enhancement of in situ bioremediation of BTEX-contaminated ground water. Ground Water Monitoring and
Remediation 17 (2), 93105.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 259
Chapelle, F.H., Vroblesky, D.A., Woodward, J.C., Lovley, D.R., 1997. Practical considerations for measuring
hydrogen concentrations in groundwater. Environmental Science and Technology 31 (10), 28732877.
Devlin, J.F., 1994. Enhanced in situ biodegradation of carbon tetrachloride and trichloroethene using a permeable
wall injection system. PhD Dissertation, Department of Earth Sciences, University of Waterloo, Waterloo,
Ontario, Canada.
Devlin, J.F., 1996. A method to assess analytical uncertainties over large concentration ranges with reference to
volatile organics in water. Ground Water Monitoring and Remediation 16 (3), 179185.
Devlin, J.F., Barker, J.F., 1994. A semipassive nutrient injection scheme for enhanced in situ bioremediation.
Ground Water 32 (3), 374380.
Devlin, J.F., Barker, J.F., 1996. Field investigation of nutrient pulse mixing in an in situ biostimulation experi-
ment. Water Resources Research 32 (9), 28692877.
Devlin, J.F., Barker, J.F., 1999. Field demonstration of permeable wall flushing for biostimulation of a shallow
sandy aquifer. Ground Water Monitoring and Remediation 19 (1), 7583.
Devlin, J.F., Muller, D., 1999. Field and laboratory studies of carbon tetrachloride transformation in a sandy
aquifer under sulfate reducing conditions. Environmental Science and Technology 33 (7), 10211027.
Devlin, J.F., Klausen, J., Schwarzenbach, R.P., 1998. Kinetics of nitroaromatic reduction on granular iron in
recirculating batch experiments. Environmental Science and Technology 32 (13), 19411947.
Devlin, J.F., McMaster, M., Katic, D., Barker, J.F., 2001. Evaluating Natural Attenuation in a Controlled
Field Experiment by Mass Balances, Flux Fences and Snapshots: A Comparison of Results. Papers
from the Third International Conference on Groundwater Quality, IAHS, University of Sheffield, UK,
June 1821.
Devlin, J.F., McMaster, M., Barker, J.F., 2002. Hydrogeologic assessment of in situ natural attenuation in a
controlled field experiment. Water Resources Research 38 (1), 3-13-11.
Dickel, O., Haug, W., Knackmuss, H.J., 1993. Biodegradation of nitrobenzene by a sequential anaerobicaerobic
process. Biodegradation 4, 187194.
Fogel, S., Lancione, R.L., Sewall, A.E., 1982. Enhanced biodegradation of methoxychlor in soil under sequential
environmental conditions. Applied and Environmental Microbiology 44 (1), 113120.
Gillham, R.W., OHannesin, S.F., 1992. Metal-catalysed abiotic degradation of halogenated organic compounds.
Presented at Modern Trends in Hydrogeology, IAH, Hamilton, Ontario, May.
Gossett, J.M., 1987. Measurement of Henrys Law constants for C1 and C2 chlorinated hydrocarbons. Environ-
mental Science and Technology 21 (2), 202208.
Guest, P.R., Friesen, K.A., Blicker, B.R., 1993. An integrated approach to remediating a no. 2 fuel oil release in
soil and groundwater. In: Calabrese, E.J., Kostecki, P.T. (Eds.), Perspectives, Analysis/Site Assessment,
Human Health Risk Assessment, Environmental Fate and Exposure, Regulatory. Hydrocarbon Contaminated
Soils, vol. 3. Lewis Publishers, Boca Raton, FL, pp. 363378.
Hofstetter, T.B., Heijman, C.G., Haderlein, S.B., Holliger, C., Schwarzenbach, R.P., 1999. Complete reduction of
TNT and other (poly)nitroaromatic compounds under iron reducing subsurface conditions. Environmental
Science and Technology 33 (9), 14791487.
Hubbard, C.E., 1992. Transport and fate of dissolved methanol, methyl-tertiary-butyl ether and monoaromatic
hydrocarbons in a shallow sand aquifer. MSc Thesis, Department of Earth Sciences, University of Waterloo,
Waterloo, Ontario, Canada.
Ji, W., Dahmani, A., Ahlfeld, D.P., Lin, J.D., Hill III, E., 1993. Laboratory study of air flow visualization. Ground
Water Monitoring and Remediation 13 (4), 115126.
Katic, D.J., 1999. Field trial of an in situ anaerobic/aerobic bioremediation sequence. MSc Thesis, Department of
Earth Sciences, University of Waterloo, Waterloo, Ontario, Canada.
Knackmuss, H.-J., 1996. Basic knowledge and perspectives of bioelimination of xenobiotic compounds. Journal
of Biotechnology 51, 287295.
Landmeyer, J.E., Chapelle, G.H., Herlong, H.H., Bradley, P.M., 2001. Methyl tert-butyl ether biodegradation by
indigenous aquifer microorganisms under natural and artificial oxic conditions. Environmental Science and
Technology 35 (6), 11181126.
Lizotte, C.C., Marley, M.C., Crawford, S.C., Lee, A.M., Sreffan, R.J., 1999. Demonstration of a sequential
anaerobic/aerobic in-situ treatment system at a Superfund site. In Situ and On-Site Bioremediation, The Fifth
International Symposium, April 1922, San Diego, CA.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 260
Mackay, D., Shiu, W.Y., Ma, K.C., 1999. Physical chemical properties and environmental fate handbook.
CRCnetBase. CRC Press LLC, Boca Raton, FL.
McCarty, P.L., Goltz, M.N., Hopkins, G.D., Dolan, M.E., Allan, J.P., Kawakami, B.T., Carrothers, T.J., 1998.
Full-scale evaluation of in situ cometabolic degradation of trichloroethylene in groundwater through toluene
injection. Environmental Science and Technology 32 (1), 88100.
Mercer, J.W., Cohen, R.M., 1990. A review of immiscible fluids in the subsurface: properties, models, character-
ization and remediation. Journal of Contaminant Hydrology 6, 107163.
Morkin, M., Barker, J.F., Devlin, J.F., McMaster, M., 1998. In situ sequential treatment of a mixed organic plume
using granular iron, O2 and CO2 sparging. Presented at the First International Conference on Remediation of
Chlorinated and Recalcitrant Compounds, Battelle, Monterey, CA, May 1821.
Morkin, M., Devlin, J.F., Barker, J.F., Butler, B.J., 2000. In situ sequential treatment of a mixed contaminant
plume. Journal of Contaminant Hydrology 45, 283302.
Morrison, S.J., Metzler, D.R., Carpenter, C.E., 2001. Uranium precipitation in a permeable reactive barrier by
progressive irreversible dissolution of zerovalent iron. Environmental Science and Technology 35 (2),
385390.
Nelson, C.H., Hicks, R.J., Andrews, S.D., 1994. An integrated system approach for in situ bioremediation of
petroleum hydrocarbon contaminated soil and groundwater. In: Flathman, P.E., Jerger, D.E., Exner, J.H.
(Eds.), Bioremediation. Field Experience. Lewis Publishers, Boca Raton, FL, pp. 429456.
Norris, R.D., Hinchee, R.E., Brown, R.A., McCarty, P.L., Semprini, L., Wilson, J.T., Kampbell, D.H., Bouwer, E.,
Borden, R., Vogel, T.M., Thomas, J.M., Ward, C., 1994. Bioremediation. Lewis Publishers, Boca Raton, FL.
Pankow, J.F., Cherry, J.A., 1996. Dense Chlorinated Solvents and Other DNAPLs in Groundwater. Waterloo
Press, Portland, OR.
Parker, L.V., Ranney, T.A., 2000. Decontaminating materials used in ground water sampling devices: organic
contaminants. Ground Water Monitoring and Remediation 20 (1), 5668.
Pratt, A.R., Blowes, D.W., Ptacek, C.J., 1997. Products of chromate reduction on proposed subsurface remedia-
tion material. Environmental Science and Technology 31 (9), 24922498.
Roberts, A.L., Totten, L.A., Arnold, W.A., Burris, D.R., Campbell, T.J., 1996. Reductive elimination of chlori-
nated ethylenes by zero-valent metals. Environmental Science and Technology 30 (8), 26542659.
Semadeni, M., Chiu, P.C., Reinhard, M., 1998. Reductive transformation of trichloroethylene by cobalamin:
reactivities of the intermediates acetylene, chloroacetylene, and the DCE isomers. Environmental Science and
Technology 32 (9), 12071213.
Starr, R.C., Cherry, J.A., 1994. In situ remediation of contaminated ground water: the funnel-and-gate system.
Ground Water 32 (3), 465476.
Su, C., Puls, R.W., 2001. Arsenate and arsenite removal by zerovalent iron: effects of phosphate, silicate,
carbonate, borate, sulfate, chromate, molybdate, and nitrate, relative to chloride. Environmental Science
and Technology 35 (22), 45624568.
Tratnyek, P.G., Johnson, T.L., 1998. In situ treatment of organics by sequential reduction and oxidation. The
First International Conference on Remediation of Chlorinated and Recalcitrant Compounds, Monterey, CA,
May 1821.
Wilhelm, E., Battino, R., Wilcock, R.J., 1977. Low-pressure solubility of gases in liquid water. Chemical Re-
views 77 (2), 219262.
Wilson, R.D., Mackay, D.M., Scow, K.M., 2002. In situ MTBE biodegradation supported by diffusive oxygen
release. Environmental Science and Technology 36 (2), 190199.
Zitomer, D.H., Speece, R.E., 1993. Sequential environments for enhanced biotransformation of aqueous con-
taminants. Environmental Science and Technology 27 (2), 226244.
J.F. Devlin et al. / Journal of Contaminant Hydrology 69 (2004) 233261 261

You might also like