Вы находитесь на странице: 1из 12

Journal of Contaminant Hydrology 107 (2009) 128–139

Contents lists available at ScienceDirect

Journal of Contaminant Hydrology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / j c o n h y d

Biogeochemistry of two types of permeable reactive barriers, organic carbon


and iron-bearing organic carbon for mine drainage treatment:
Column experiments
Qiang Guo, David W. Blowes ⁎
Department of Earth Sciences, University of Waterloo, 200 Univ. Ave. W., Waterloo, Ontario, Canada N2L 3G1

a r t i c l e i n f o a b s t r a c t

Article history: Permeable reactive barriers (PRBs) are an alternative technology to treat mine drainage
Received 15 October 2007 containing sulfate and heavy metals. Two column experiments were conducted to assess the
Received in revised form 18 April 2009 suitability of an organic carbon (OC) based reactive mixture and an Fe0-bearing organic carbon
Accepted 24 April 2009
(FeOC) based reactive mixture, under controlled groundwater flow conditions. The organic
Available online 3 May 2009
carbon mixture contains about 30% (volume) organic carbon (composted leaf mulch) and 70%
(volume) sand and gravel. The Fe0-bearing organic carbon mixture contains 10% (volume)
Keywords:
zero-valent iron, 20% (volume) organic carbon, 10% (volume) limestone, and 60% (volume)
Permeable reactive barrier (PRB)
sand and gravel. Simulated groundwater containing 380 ppm sulfate, 5 ppm As, and 0.5 ppm Sb
Zero-valent iron
Organic carbon was passed through the columns at flow rates of 64 (the OC column) and 62 (the FeOC column)
Sulfate reducing bacteria (SRB) ml d− 1, which are equivalent to 0.79 (the OC column) and 0.78 (the FeOC column) pore
δ34S volumes (PVs) per week or 0.046 m d− 1 for both columns. The OC column showed an initial
Cathodic hydrogen sulfate reduction rate of 0.4 µmol g (OC)− 1 d− 1 and exhausted its capacity to promote sulfate
reduction after 30 PVs, or 9 months of flow. The FeOC column sustained a relatively constant
sulfate reduction rate of 0.9 µmol g (OC)− 1 d− 1 for at least 65 PVs (17 months). In the FeOC
column, the δ34S values increase with the decreasing sulfate concentration. The δ34S
fractionation follows a Rayleigh fractionation model with an enrichment factor of 21.6‰. The
performance decline of the OC column was caused by the depletion of substrate or electron
donor. The cathodic production of H2 by anaerobic corrosion of Fe probably sustained a higher
level of SRB activity in the FeOC column. These results suggest that zero-valent iron can be used
to provide an electron donor in sulfate reducing PRBs. A sharp increase in the δ13C value of the
dissolved inorganic carbon and a decrease in the concentration of HCO− 3 indicate that
hydrogenotrophic methanogenesis is occurring in the first 15 cm of the FeOC column.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction component of the reactive medium in many PRBs. About


three-quarters of the current full-scale PRBs use iron metal as
Permeable reactive barriers (PRBs) are a promising the reactant medium (Jambor et al., 2005). PRBs containing
alternative technology for the in situ remediation of both zero-valent iron can effectively remediate a variety of
organic and inorganic contaminants in groundwater, com- contaminants in groundwater including chlorinated solvents
pared to the traditional pump and treat method (Scherer and inorganic contaminants (such as Cr, Se, U, As and nitrate).
et al., 2000; Jambor et al., 2005). Zero-valent iron is a Organic carbon (wood chips, paper mill pulp, municipal
compost, etc.) has been used to treat acidic drainage at mine
sites and groundwater contamination at industrial sites
⁎ Corresponding author. Tel.: +1 519 888 4878; fax: +1 519 746 3882.
which is rich in sulfate and heavy metals (Benner et al.,
E-mail addresses: q5guo@sciborg.uwaterloo.ca (Q. Guo), 1997, 1999, 2000). Under favourable conditions, sulfate
blowes@sciborg.uwaterloo.ca (D.W. Blowes). reducing bacteria (SRB) catalyze the oxidation of organic

0169-7722/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jconhyd.2009.04.008
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 129

Table 1 carbon and iron filings increased both the extent and the rate
Composition of reactive mixture in columns. of sulfate reduction and metal removal, compared to the
Composition The OC column The FeOC column organic material alone (Mountjoy and Blowes, 2002). How-
ever, biogeochemical and isotopic changes coupled with
Dry wt.% Dry wt.%
microbial activity were not determined in those experiments.
Top layer (4 cm)
Ottawa sand 100 100 The objectives of this study are: 1) to examine the micro-
biology and isotope geochemistry of an organic carbon reactive
Middle layer (33 cm) mixture, 2) to examine the sulfate reduction rate, geochemistry,
Composted leaf mulch 14 8 isotope geochemistry and microbiology of an Fe0-bearing
Cornnelly zero-valent iron 0 17
Limestone 0 3
organic carbon mixture, and 3) to compare the biogeochemical
Ottawa sand 33 30 characteristics of the two reactive mixtures.
Pea gravel 53 42
2. Experimental and analytical methods
River sediments layer (2 cm)
Fresh river sediments 7 4
Composted leaf mulch 7 4 2.1. Column design and experimental setup
Cornnelly zero-valent iron 0 17
Limestone 0 3 The experimental setup and procedures were similar to
Ottawa sand 33 30 those used by Waybrant et al. (2002). Two acrylic columns
Pea gravel 53 42
(7.6 cm inner diameter, 41 cm long) were packed with silica
Bottom layer (2 cm) sand, gravel, and reactive material (Table 1). The OC column
Silica sand 100 100 contains 30% (volume) organic carbon and the FeOC column
contains 10% (volume) zero-valent iron and 20% (volume)
organic carbon. The organic carbon is composted leaf mulch
obtained from a municipal recycling program. The zero-valent
iron is from Connelly GPM Inc., in Chicago, Il. Ottawa sand is
carbon coupled with the reduction of sulfate to sulfide coarse pure silica sand. The silica sand was added as an inert
through the following reaction: material, intended to maintain the permeability of the
reactive mixture, and to retain the reactive material within
2− −
SO4 þ 2CH2 O→H2 S þ 2HCO3 ð1Þ the column. Sediment obtained from a stream flowing
through the University of Waterloo campus was added into
where CH2O represents simple organic compounds. Increased both columns to provide a source of sulfate reducing bacteria.
H2S concentrations enhance the precipitation of metals as This sediment, which was obtained from an area where
metal sulfides: sulfate reducing conditions were indicated by a strong odour
of H2S was mixed thoroughly with other materials in the first
2þ 2−
M þS →MS ð2Þ approximately 2 cm from the bottom in both columns
(Table 1).
where M2+ denotes a divalent metal such as Fe, Cd, Ni, Cu, Co, Carbon dioxide gas was bubbled through the columns for
and Zn. Rates of sulfate reduction and metal removal capacity 12 h in order to displace the less soluble air from the pore
in organic carbon reactive barriers have been studied spaces, and thereby increase the level of saturation of the
previously, as have changes in groundwater and solid-phase column materials. The reactive mixtures were saturated with
geochemistry, and microbial communities in these systems a solution, which was made with distilled water, containing
(Benner et al., 1997; Herbert et al., 1998; Waybrant et al., 1998; 1000 mg/l SO4 as CaSO4 and 5% sodium lactate to help
Benner et al., 2000; Herbert et al., 2000; Waybrant et al., promote sulfate reducing bacteria activity. The columns were
2002; Ludwig et al., 2002; Jambor et al., 2005; Logan et al., transferred into an anaerobic glovebox. The anaerobicity of
2005). the glovebox was monitored regularly using methylene blue
Changes in sulfur isotope ratio can be used as an indicator anaerobic strip indicators (GasPak).
or the extent of bacterially mediated sulfate reduction The same simulated groundwater as the input solution
reactions. Fractionation of sulfur isotopes occurs because was used for both columns. The chemical composition
less energy is required to break 32S–O bonds than 34S–O (Table 2) of the simulated groundwater was based on the
bonds (Clark and Fritz, 1997). As a consequence, the heavier water chemistry analysis of groundwater collected in January
isotope, 34S, accumulates in the residual dissolved sulfate. 1997 from monitoring well MMW5 at Getchell gold mine,
Changes in isotopic ratios of carbon and sulfur have been Nevada (Tempel et al., 2000).
demonstrated to be sensitive indicators of biogeochemical
processes occurring in permeable reactive barrier systems
Table 2
that promote biologically mediated sulfate reduction (Benner Chemical composition of simulated groundwater (mg/l).
et al., 2000; Waybrant et al., 2002).
A previous study combined zero-valent iron and organic Na K Mg Ca
carbon together to treat groundwater contaminated with 23.00 2.00 12.00 191.00
sulfate and heavy metals at an industrial site near Vancouver
HCO3 SO4 Sb As
British Columbia. The lab tests prior to the installation of the
195.20 380.00 0.50 5.00
full-scale barrier showed that a mixture containing organic
130 Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139

The pore volumes were 568 ml for the OC column and Electrode calibration and performance were checked before
556 ml for the FeOC column. A variable speed, 4-channel and after each sample measurement. The pH electrode (Orion
peristaltic pump (ISMATEC CP78016-30 from Cole-Parmer pH 9107BN) was calibrated using standard 7.0 and 10.0
Canada Inc.) was used to pump the influent through each of buffers (J.T. Baker), and the Eh electrode (Orion 9678BN) was
the columns. The column effluent solutions were collected in checked using Zobell's solution.
flow-through sample cells within the glovebox, and overflow Samples were filtered through a 0.45-µm cellulose acetate
was directed to waste jugs outside of the glovebox. Flow rates filter, and both acidified and nonacidified samples were
were determined by measuring the overflow. The average collected and stored at 5 °C until analyzed. Alkalinity was
flow rate of the OC column was 64 ml d− 1 or 17 m a− 1 and determined on the filtered samples using standardized H2SO4
that of the FeOC column was 62 ml d− 1 or 17 m a− 1, which are and a Hach digital titrator. Concentrations of cations (Ca, Fe, K,
equivalent to 0.79 (the OC column) and 0.78 (the FeOC Mg, Na, and Mn) and total S were determined using inductively
column) pore volumes (PVs) per week or 0.046 m d− 1 for coupled plasma atomic emission spectroscopy (ICP-AES), and
both columns. SO4 analyses were conducted using ion chromatography (IC).
Analyses were completed within 60 d unless otherwise
2.2. Geochemical analysis specified. Dissolved S2− was measured using a spectrophot-
ometer (HACH DR/2010). Selected samples were also analyzed
Each column has 15 sampling ports. Ports 1 and 15 are for δ34S–SO4 and δ13C–CO3 in the Environmental Isotope
3 cm from the bottom (influent) and the top (effluent) ends Laboratory at the University of Waterloo.
of the column, respectively, and every two ports are separated
by 2 cm. Samples from influent, effluent and various ports 2.3. Solid-phase analysis
were collected within the glovebox for Eh, pH, anion, cation,
and isotope analyses. Measurements of pH, Eh and alkalinity At about PV 71, the experiment was terminated and solid
were determined immediately after sampling. The pH and Eh samples were taken from both columns and dried in a vacuum
were measured on unfiltered samples in sealed cells outside desiccator inside the glovebox for X-ray photoelectron spectro-
of the glovebox. Some measurements were made inside the scopy (XPS) and X-ray absorption near edge spectroscopy
glovebox, but no difference was found between measure- (XANES) analyses. The XPS analysis was conducted using a
ments inside the glovebox and those outside the glovebox. Kratos Axis Ultra X-ray photoelectron spectrometer. XPS

Fig. 1. Temporal changes in sulfate and sulfide of the OC and the FeOC columns. Port 2 is at 5 cm from the bottom of the columns. The error bars of some data were
smaller than the symbols.
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 131

analyzes a surface area of 300 × 700 μm and probes the surface


of the sample to a depth of 7–10 nm. All high-resolution spectra
were charge corrected to the main peak of the carbon 1s
spectrum set to 285.0 eV. The XANES analysis was carried out at
beamline X15B, National Synchrotron Light Source (NSLS) at
Brookhaven National Laboratory (Upton, NY). Analysis condi-
tions were similar to those of Einsiedl et al. (2007).

2.4. Microbiological analysis

Microbial enumerations were conducted to assess the


populations of sulfate reducing bacteria (SRB), iron reducing
bacteria (IRB), and fermentative acid producing bacteria
(APB). To enumerate bacteria, 1 g of reactive mixture is
added to each of five replicate serum bottles or tubes, each of
which contains 9 ml of the medium. The inoculated serum
bottles or tubes are then diluted tenfold using 1 ml syringes
into another series of 5 serum bottles or tubes. This is
repeated until there are 10 serial dilutions. The bottles or
tubes are then incubated for certain number of days (inside an
anaerobic chamber for SRB and IRB). The positive results are
usually indicated by color change or precipitation of black
sulfides (Gould et al., 2003). The most probable number
(MPN) technique (Cochran, 1950) was used to estimate the
density of bacteria in the reactive mixture.
The procedures for enumerating SRB, IRB, and APB have
been described in detail in Benner et al. (2000), Gould et al.
(2003), and Hulshof et al. (2003), respectively. Spectro-
photometric determination of the hydrolysis of fluorescein Fig. 2. The two profiles on the top are sulfate and sulfide concentrations
diacetate (FDA) (Schnurer and Rosswall, 1982) also was used during PVs 25–29. The two profiles on the bottom are δ34S and sulfate
to measure the total microbial activity in columns. Enumera- concentrations during PVs 43–52 (for the OC column) or PVs 43–50 (for the
FeOC column). The error bars of some data were smaller than the symbols.
tions were conducted on samples of the column material
immediately following the termination of the column
experiments. For each column, 7 samples (Ports 2, 4, 6, 8, period (Fig. 2). The sulfate removal capacity of the OC column
10, 12 and 14) were extracted. is 42 µmol g− 1(dry weight of OC). The sulfate removal rate of
the FeOC column is 0.9 µmol g (dry weight of OC)− 1 d− 1 or
2.5. Geochemical modelling 0.3 µmol g (dry weight of FeOC)− 1 d− 1. Since the removal
rate is relatively stable from the start to PV 65, the actual
The geochemical-speciation mass-transfer model MIN- removal capacity of the FeOC column was not attained.
TEQA2 (Allison et al., 1990) was used to assist in the However, the total sulfate removed from influent to Port 2
interpretation of the water chemistry data and to indicate was used to calculate the minimum estimate of the removal
chemical equilibrium reactions potentially controlling the capacity (Figs. 1 and 2). The minimum sulfate removal
concentrations of dissolved constituents in column water. The capacity of the FeOC column is 816 µmol g− 1 (dry weight of
MINTEQA2 thermodynamic database was modified to be OC), almost 20 times of the removal capacity of the OC
consistent with the WATEQ4F database (Ball and Nordstrom, column material, or 297 µmol g− 1 (dry weight of FeOC).
1991). The possible sulfate sinks are precipitation, adsorption and
reduction. Geochemical calculations, conducted with MIN-
3. Results and discussion TEQA2, show undersaturation with respect to all of the sulfate
minerals in the database, indicating that there is no tendency for
3.1. Sulfate reduction and sulfur isotopes sulfate minerals to precipitate (Fig. 3). Although not measured
in this study, goethite commonly exists in zero-valent iron
Sulfate concentration changes are quite different in the (Ritter et al., 2002) and it can adsorb sulfate significantly (Liu
two columns. In the OC column, about 10% to 50% sulfate was et al., 1999). However, sulfate adsorption capacity of goethite is
removed in the first 20 PVs, after 30 PVs no removal was relatively low at pH N 7 (Liu et al.,1999). Many indicators suggest
observed (Fig. 1). Sulfate concentration in the FeOC column that sulfate is reduced to sulfide. Dissolved hydrogen sulfide was
decreased by about 50% consistently and the sulfate removal found in both columns. Black Fe sulfides were visible in both
rate did not diminish after 65 PVs (Fig. 1). The sulfate profile columns. XANES and XPS analysis also found reduced sulfur
of the OC column during PVs 25–29 shows no removal from species in solid samples (Figs. 4 and 5).
the bottom to the top of the column, whereas the profile of Isotopic measurements indicate that the sulfate reduction
the FeOC column shows that sulfate concentration contin- was mediated by sulfate reducing bacteria (SRB) (Fig. 2). The
uously decreases from influent to effluent over the same time stable isotope composition of the sulfate–sulfur is reported as
132 Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139

Fig. 3. Saturation indices for selected mineral phases for effluent samples from the OC and the FeOC columns.

δ34S, ratio of 34S/32S in per mil relative to the standard Canyon concentration decreases. The Rayleigh equation can be used
Diablo Troilite (CDT). to describe this fractionation:
  h  i
34 34 3 ðα − 1Þ 3
34
δ Sðin xÞ =
34
S=
32
Ssample =
34
S=
32
Sstandard − 1 × 1000: δ S= δ S0 + 10 f − 10 ð4Þ

ð3Þ where δ34S0 and δ34S refer to the δ34S values of the aqueous
Sulfate reducing bacteria preferentially remove the light sulfate of the influent and the samples at different sampling
isotope 32S from the sulfate in the solution, causing isotopic
enrichment of 34S in the residual sulfate and the accumulation
of light isotopes 32S in the sulfides (Strebel et al., 1990). The
δ34S values of the residual sulfate increase with decreasing
sulfate concentrations in a manner that is consistent with
bacterial reduction of the aqueous sulfate. From the influent
to the effluent in the FeOC column, δ34S increases from 4.37‰
to 20.33‰ and then decreases to 14.81‰. The OC column has
a much smaller increase in δ34S (from 4.37‰ to 5.73‰),
which is consistent with the observation that the OC column
lost most of its sulfate removal capacity after 25 PVs. Sulfur
isotope data suggest that bacterially sulfate reduction is more
likely than sulfate-mineral precipitation.
The profiles of SO4 and δ34S during PVs 43–50 show a close
relationship between sulfate removal and sulfur isotope
composition (Fig. 2): the δ34S increases as the sulfate

Fig. 5. High resolution scan of S(2p) XPS spectra. (a) the OC column Port 2;
(b) the FeOC column Port 2. Fit curves include 2p3/2 (solid line) and 2p1/2
Fig. 4. Sulfur XANES spectra. (dashed line) components.
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 133

ports, respectively, and ƒ is the fraction of the remaining Table 3


sulfate, and α is the instantaneous isotope fractionation factor Sulfur isotopic fractionation of reactive materials.

(Kendall and Doctor, 2003). α is defined by Reactive material αa ε b (‰)


FeOC 0.9784 − 21.6
α = Rprod = Rreact ð5Þ
Column 1 c 0.9616 − 38.4
Column 2 c 0.9579 − 42.1
where Rprod and Rreact are the 34S/32S ratios of the sulfide and a
Calculated from best-fit curves (see Fig. 3.) using δ34S and ƒ data of this
sulfate at an instant during sulfate reduction. The extent of study and of Waybrant et al. (2002).
b
the fractionation or the 34S enrichment in the residual sulfate Calculated using Eq. (3) instead of ε = (δs − δm) / lnƒ used in Waybrant
et al. (2002), which is not accurate when |δ| N 10.
also can be indicated by the enrichment factor ε, which is c
From Waybrant et al. (2002).
defined by

e = ðα − 1Þ × 1000: ð6Þ
SRB population of the OC column was inactive. The SRB
If the fractionation factor α is found to be constant throughout population of FeOC still actively participated in sulfate
the column, then the Rayleigh equation and the δ34S values reduction. Population size is not a good indicator of activity.
can be used to indicate how much sulfate is reduced (Kendall A large population of SRB does not necessarily indicate
and Doctor, 2003). significant sulfate reducing activity. For example, Hines et al.
Values of α and ε were calculated using best-fit curves of (1999) observed that the relative abundance SRB in a salt
the data from the FeOC column of this study and from marsh did not vary much throughout the year, despite the
Columns 1 and 2 of Waybrant et al. (2002) (Fig. 6 and large temporal change of the sulfate reduction rate.
Table 3). In all cases, the data fit the Rayleigh fractionation PRB systems usually perform very well in the first months
model within the uncertainties of the measurements. The two after bacterial activity is established (Waybrant et al., 1998;
columns of Waybrant et al. (2002) have similar enrichment Hulshof et al., 2003), but longer-term performance in both
factors (− 38.4‰ and − 42.1‰), but the FeOC column has a field and laboratory-scale systems gradually declines well
smaller fractionation (− 21.6‰). The results suggest that before the theoretical capacities of the barriers have been
within a PRB δ34S can be used to determine the extent of reached (Blowes et al., 2000). Sulfate reduction in the OC
sulfate reduction, but δ34S fractionation is not consistent column stopped after 9 months, while the FeOC column was
between reactive materials. showing a much longer and better performance (Fig. 1).

3.2. SRB and H2

Microbial enumerations show that both FeOC and OC have


5.2 × 103 to 3.2 × 105 MPN/g SRB throughout the column,
except there are two large peaks (107 MPN/g) of SRB at Port 2
and Port 10 of FeOC (Fig. 7). Because of the two peaks, the
average SRB population of the FeOC column (1.4 × 107 MPN/
g) is two orders of magnitude higher than the average SRB
population of the OC column (1.4 × 105 MPN/g). Because the
rate of sulfate reduction had declined in the OC column prior
to the sampling for the enumerations, it is possible that the

Fig. 6. (δ34S + 103)/(δ34S0 + 103) vs. fraction of remaining SO4. δ34S0 and
δ34S are the δ34S values of the aqueous sulfate of the influent and the samples
at different sampling ports, respectively. Best-fit curves are used to calculate Fig. 7. Profiles of SRB, IRB, APB, and FDA (immediately after the termination of
instantaneous fractionation factor α (listed in Table 3). both columns).
134 Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139

Potential causes of the decline in performance include 1999). Although the APB population (~ 20 MPN/g) is also
exposure of the reactive mixture to oxidized conditions, much less than that of SRB (1.4 × 107 MPN/g) in the FeOC
depletion of a reactive component of a mixture, a decline of column, cathodic H2 is able to sustain a high sulfate reduction
surface area resulting from precipitation of secondary miner- rate and high SRB population. A similar situation is also found
als on reactive surfaces, clogging and preferential flow in temperate salt marches, where sulfate reduction rate is
(Blowes et al., 2000). The decline of surface area is not likely high and SRB populations on roots can account for over 30% of
the cause of the performance decline of the OC column, the total microbial community, because SRB can use acids and
because microscopic analysis of the solid samples shows that alcohols released directly by plant roots and do not depend on
FeOC has much more carbonate and sulfide precipitation than fermentation (Rooney-Varga et al., 1997; Hines et al., 1999).
OC. Reactive components in PRB usually include a variety of Hydrogen might also play a role in sulfur isotope
cellulosic organic substrates which act as the source of carbon fractionation. Electron donor effect on sulfur isotope frac-
and nutrients, such as N, P, etc. (Logan et al., 2005). Both tionation of Desulfovibrio desulfuricans was studied and the
columns had the same organic material as the carbon source fractionation was significantly lower when H2 was the
and nutrients sources, and the OC column maintained a stable electron donor than when other donors such as lactate,
neutral pH. Nutrients, sulfate and pH were similar and so are acetate, and ethanol were consumed (Kaplan and Riggenberg,
not considered the main contributors of the performance 1964). This difference might explain why the two columns of
differences between the two columns. Waybrant et al. (2002), where the substrate is likely lactate
SRB utilize simple compounds such as H2, ethanol, acetate, and acetate, have larger fractionation ( ε = − 38.4‰
and lactate, and require upstream cellulolytic microbes and and −42.1‰) than the FeOC column (ε = − 21.6‰) (Table 3
fermenters to break down complex polymers such as and Fig. 6).
cellulose and lignin to simple compounds. Decomposition Populations of IRB were also enumerated. IRB of the OC
and fermentation can limit microbial activity in anaerobic column (average 1.1 × 105 MPN/g) is slightly larger than that of
ecosystem (Megonigal et al., 2003). Cellulolysis is considered the FeOC column (average 3.6 × 104 MPN/g) (Fig. 7). Measure-
the limiting factor in PRB performance (Logan et al., 2005). So ments of the hydrolysis of fluorescein diacetate (FDA), which
the depletion of substrate or energy source could be the cause indicate the total microbial activity, indicated higher rates of
of the performance decline observed in the OC column. FDA hydrolysis in the OC column (4.3× 10− 3 nmol/h/g) than in
In PRB systems, only a small fraction of the organic matter the FeOC column (1.8 × 10− 3 nmol/h/g) (Fig. 7). Because the
originally in the reactive mixture is immediately available to sum of SRB, IRB and APB of the FeOC column is greater than that
SRB (Benner et al., 1999), and after it is consumed SRB depend of the OC column, the higher rates of FDA hydrolysis in the OC
on cellulolytic and fermentative organisms to provide the column indicate that there were other microbial groups besides
substrate. Cellulolytic and fermentative bacteria can greatly the three bacteria groups in both columns and those groups
outnumber SRB in environments where the SRB are totally were more active in the OC column than in the FeOC column.
dependent on the fermentative products. For example, in
13
sulfate reducing sediments, SRB typically account for ~5% of 3.3. Methanogenesis and C in the FeOC column
the all bacteria present, while most bacteria are involved in
polymer hydrolysis and fermentation (Devereux et al., 1996). Carbon isotopic data indicate methanogenesis in the first
The microbial enumeration results show that the populations 5.5 cm (from the influent to Port 2) of the FeOC column. The
of APB are about 4 orders of magnitude less than those of SRB very negative δ13C value − 33.2‰ of the influent (Table 4 and
in the OC column (Fig. 7). The small population of APB and the Fig. 8) is probably derived from the CO2 gas (δ13C value
decline of sulfate reduction can be explained by: 1) SRB was ranging from −30‰ to − 35‰), which was used to dissolve
initially thriving on the immediately available organic matter CaCO3 and CaSO4 when the simulated groundwater was
and created relatively large population compared to that of prepared. From influent to Port 2, δ13C values increase
APB; 2) after the immediately available organic matter was dramatically from −33.2‰ to −7.1‰ in the FeOC column. In
consumed the small population of APB could not provide natural waters, very positive values of δ13C of dissolved
sufficient substrate to sustain the much larger population of inorganic carbon (DIC) are a useful indicator of methanogenesis
SRB, resulting in a decline in the activity of the SRB population (Kendall and Doctor, 2003). Carbonate dissolution has a much
and the observed decline in the rate of sulfate reduction. smaller fractionation factor (δ13CCaCO3 =δ13CHCO−3 + 2; Buchardt
In the FeOC column, another substrate is available, H2, and Fritz, 1980). In the FeOC column, the increase in δ13C is also
produced by anaerobic Fe corrosion: accompanied by loss of DIC (or alkalinity) from influent to
Port 2 (Fig. 9). The isotopic mass balance calculation shows that
2þ −
Fe þ 2H2 O→Fe þ H2 þ 2OH : ð7Þ the lost DIC has a δ13C value of −67.4‰ (Table 4), which is
within the typical δ13C range −52‰ to −80‰ of the methane
H2 is the most common electron donor in the terminal an- produced by fermentation of organic matter and subsequent
aerobic metabolism (Megonigal et al., 2003). Methanogens, methanogenesis (Kendall and Doctor, 2003). It is likely, there-
SRB, iron reducers and other hydrogen-oxidizing bacteria fore, that methanogenesis causes the increase in the δ13C of the
consume cathodic hydrogen and accelerate anaerobic iron DIC.
corrosion. SRB are regarded as the chief agent of biocorrosion Methanogenic bacteria use a more limited variety of
in anaerobic environments (Rajagopal and LeGall, 1989). simple energy sources compared to SRB, iron reducers and
Hydrogen generated by Fe corrosion also stimulated the nitrate reducers. The most important energy sources for
growth of microbial populations and increased sulfate methanogens are H2 and acetate (Megonigal et al., 2003).
reduction in zero-valent iron packed columns (Gu et al., Hydrogenotrophic methanogenesis consumes hydrogen and
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 135

Table 4 rinsed with deionised water prior to drying. Dissolved sulfate


C isotopic composition of FeOC influent and Port 2 water. in porewater probably precipitated during drying. Geochem-
Concentration δ13C ical modelling shows no sulfate minerals reach saturation in
−3 either column (Fig. 3). XANES spectra show that the samples
10 mol/l ‰
of both columns have reduced species (Fig. 4). S(2p) XPS
Influent DIC 4.99 a − 33.2
Influent Ca 5.88 spectra collected from the OC column Port 2 and the FeOC
Port 2 DIC 1.36 a − 7.1 column Port 2 samples are similar. Both have a high energy
Port 2 Ca 3.43 and a low energy broad peaks. The high energy peak is
CaCO3 ppt 2.45 b − 31.2 c attribute to sulfate with a binding energy of 168.5 eV and the
Lost DIC 1.18 d − 67.4 e
low energy peak is best fit with monosulfide (161.7–161.8 eV)
a
Calculated from pH and alkalinity using MINTEQA2 (Allison et al., 1990). and polysulfide (163.4–163.8 eV) (Fig. 5, Tables 5 and 6). It is
b
CaCO3 precipitation = Influent Ca − Port 2 Ca.
c generally proposed that formation of pyrite proceeds via the
δ13CCaCO3 ppt = δ13CHCO3− + 2 (Buchardt and Fritz, 1980).
d
Lost DIC = Influent DIC − Port 2 DIC − CaCO3 ppt. following pathway: amorphous FeS or disordered mackina-
e
δ13Clost DIC = (Influent DIC × δ13CInfluentDIC − Port 2 DIC × δ13CPort 2DIC − wite–mackinawite (FeS)–greigite (Fe 2+ Fe 3+ 2 S 4 )–pyrite
CaCO3 ppt × δ13CCaCO3 ppt) / lost DIC. (FeS2), and polysulfide suggests that mackinawite is being
oxidized to form greigite (Wilkin and Barnes, 1997; Jambor et
al., 2005). The observed distribution of sulfur species is
DIC to generate methane (H2 is the source of electrons and consistent with the early stages of this sequence.
DIC is both an electron acceptor and the source of cellular
carbon; Megonigal et al., 2003): 3.5. pH, Eh, alkalinity, Ca, and Fe
− −
4H2 þ HCO3 →CH4 þ 2H2 O þ OH : ð8Þ Because CO2 gas was used to dissolve CaCO3 and CaSO4
when the simulated groundwater was prepared, the pH at the
This pathway is consistent with the characteristics of the beginning of each batch of input solution was ~ 6. Then it
first 5.5 cm of the FeOC column: that the cathodic H2 is more gradually rose to 8.3 as CO2 degassing proceeded. But both
abundant than labile organic carbon, the large increase in columns have strong pH buffering capacity and the pH values
δ13C value of DIC (Fig. 8), and the decrease in alkalinity (DIC) in both columns (except that of Port 2 of the OC column)
concentration (Table 4). Thus, the methanogenesis in the first generally are not affected by the oscillating pH value of the
5.5 cm of the FeOC column is most likely hydrogenotrophic. input solution (Figs. 9 and 10). The pH values in the OC
The small increase in δ13C value of DIC from Port 2 to Port 6 column are relatively stable between 7.0 and 7.4, which is
might indicate the methanogenesis is slowing down and the
small decrease in δ13C value of DIC from Port 6 to Port 10
might be the result of the anaerobic oxidation of methane by
bacteria such as SRB. Anaerobic methane oxidation coupled to
sulfate reduction is thermodynamically favourable under
conditions found in marine sediments (Martens and Berner,
1977).

3.4. Sulfur speciation on solid surface

Both sulfate and sulfide phases were found in XANES and


XPS results (Figs. 4 and 5). Sulfate was not one of the original
solid phases. To protect surface properties, samples were not

Fig. 8. Profiles of δ13C of the FeOC column during PVs 43–50. The error bars Fig. 9. Profiles of Fe, Ca, Eh, and pH during PVs 25–29. The error bars of some
were smaller than the symbols. data were smaller than the symbols.
136 Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139

Table 5 oversaturated with respect to FeS precipitation, which is


Binding energies for S(2p) in various references. consistent with the results of the solid-phase analysis (Figs. 4
Species BE (eV) References
and 5). The effluent water from both columns was under-
saturated with respect to gypsum and other sulfate minerals
Monosulfide S2− 160.9 Herbert et al. (1998)
161.3 Pratt et al. (1994a) (not shown here), suggesting that it is unlikely that secondary
161.6 Boursiquot et al. (2001) sulfate minerals precipitated in the columns. This is most
Bisulfide S2−
2 162.2 Herbert et al. (1998) obvious in the OC column because Ca and sulfate did not
162.4 Thomas et al. (1998)
change from influent to effluent after 25 PVs. In the FeOC
Polysulfide S2−
n 163.3 Pratt et al. (1994b)
163.4 Thomas et al. (1998) column Ca decreased with decreasing alkalinity, which
Sulfate SO2−
4 168.5 Jones et al. (1992) suggests calcite precipitation rather than gypsum precip-
169.1 Pratt et al. (1994b) itation. The SI values for siderite indicated supersaturation in
the OC column and for the FeOC column at early time (b17
PVs). Siderite was proposed as a minor Fe sink in areas of high
similar to other organic carbon PRBs (Waybrant et al., 1998, Fe flux in a PRB system (Herbert et al., 2000). Fe concentra-
2002). The pH value increases rapidly from ~7.5 at Port 2 up tions in both columns were very low, so siderite was not
to ~9.5 at Port 4 and stays ~ 9 in the rest of the FeOC column, expected to control either Fe or alkalinity.
which is lower than the typical pH values associated with
iron-bearing PRBs (Gu et al., 1999). This increase in pH is 4. Implications
caused by anaerobic corrosion of Fe (Eq. (7)).
Generally Eh values decrease from 400 mV in the input The Fe0-bearing organic carbon reactive mixture has the
solution to b200 mV in both columns, which reflects the potential to provide a comprehensive method to remove
change from aerobic to anaerobic environments (Figs. 9 and metals (such as Cr, U, Cu, Zn, As, Sb, Se, etc.) and other
10). But the Eh measurements are unstable and inconsistent. inorganic contaminants (such as sulfate and nitrate) in acid
This level of variability probably results from the sampling mine drainage and groundwater. 1) Fe can immobilize metals
procedures. by reduction, such as Cr6+ → Cr3+ (Blowes et al., 1997) and
Alkalinity shows quite different trends between the two U6+ → U4+ (Gu et al., 1998). 2) Hydrous Fe oxides produced
columns (Fig. 10). The alkalinity increased in the OC column, by Fe corrosion is a strong absorbent to remove As (McRae
although the extent of the increase diminished from PV 1 to et al., 1999; Bain et al., 2003). 3) Cathodic H2 produced by
PV 40. In the FeOC column alkalinity decreased and the extent anaerobic Fe corrosion is an electron donor for SRB and can
of the decrease is greater during the same period of time. The sustain a high sulfate reduction rate. 4) Cathodic H2 can also
causes for these changes were discussed above. be used by denitrifying bacteria (Megonigal et al., 2003) and
Calcium concentrations were relatively uniform in the OC has a potential to enhance denitrification. 5) Anaerobic Fe
column after 15 PVs, there was not much change in concentra- corrosion also produces OH− and helps to neutralize the acid
tion between influent and effluent (Figs. 9 and 10). This in mine drainage.
observation is consistent with the observation that no Several processes can affect the concentration of sulfate in
secondary calcium-bearing minerals were detected in solid mine drainage or groundwater, such as dilution, sulfate-mineral
samples. In the FeOC column 60% to 70% Ca was constantly dissolution and precipitation, and sulfate reduction. So sulfate
removed from the solution and after 25 PVs much of the concentration alone cannot truly show how much sulfate has
removal happened before Port 2 (the first 5.5 cm). Geochemical been reduced. Bacterial reduction of sulfate is the primary
modelling shows that conditions favour calcite precipitation source of the variability of sulfate isotopic compositions in
(Fig. 3) and mineralogical observations confirmed the presence
of significant carbonate minerals in solid samples. The
precipitation of carbonate minerals could affect the long term Table 6
performance of FeOC based PRB system by clogging up the pore Binding energies, peak full width at half maximum (FWHM), and peak areas
space or coating the reactive materials. for S(2p) XPS spectra.
Iron concentrations were very low in both columns (Figs. 9
BE (eV) a FWHM (eV) Area (%) b Species
and 10). Because there is no Fe in the input solution, the
Wood chips of the OC column Port 2
dissolved Fe is attributed to the reductive dissolution of Fe 161.72 1.47 54.8 Monosulfide
oxides in solid material (iron oxide coating of sand and gravel 163.81 1.47 45.2 Polysulfide
in the OC column and on the zero-valent iron in the FeOC 168.53 2.00 0c Sulfate
column) and from anaerobic corrosion of Fe. Fe concentra-
Wood chips of the FeOC column Port 2
tions are controlled by precipitation of Fe monosulfides.
161.81 1.56 32.0 Monosulfide
Poorly crystalline Fe monosulfides or disordered mackinawite 163.77 1.56 68.0 Polysulfide
are the primary sink for Fe and S in a PRB at Nickel Rim mine 168.49 2.11 0c Sulfate
site (Herbert et al., 2000) and black Fe sulfides were visible in
Iron chips of the FeOC column Port 2
the FeOC column.
163.42 1.38 100.0 Polysulfide
168.27 2.07 0c Sulfate
3.6. Geochemical modelling a
All binding energy peak locations were charge corrected to the main
peak of the carbon 1s spectrum set to 285.0 eV.
Saturation indices (SI) were calculated for the effluent b
Area includes 2p1/2 and 2p3/2 contributions.
samples of both columns (Fig. 3). Both columns were c
Sulfate is not included in area calculation, see text.
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 137

Fig. 10. Temporal changes of Fe, Ca, Eh, pH and alkalinity. The error bars of some data were smaller than the symbols.
138 Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139

natural aquatic systems (Kendall and Doctor, 2003). Dissolution Benner, S.G., Blowes, D.W., Gould, W.D., Herbert Jr., R.B., Ptacek, C.J., 1999.
Geochemistry of a permeable reactive barrier for metals and acid mine
of sulfate minerals does not significantly fractionate sulfate drainage. Environmental Science Technology 33, 2793–2799.
isotopic compositions and precipitation of sulfate minerals is Benner, S.G., Gould, W.D., Blowes, D.W., 2000. Microbial populations
accompanied by only slight isotopic fractionation (fractionation associated with the generation and treatment of acid mine drainage.
Chemical Geology 169, 435–448.
factor b2%, Holser and Kaplan, 1966). The δ34S measurements Blowes, D.W., Ptacek, C.J., Jambor, J.L., 1997. In-situ remediation of Cr(VI)-
from this study and from Waybrant et al. (2002) confirm that contaminated groundwater using permeable reactive walls: laboratory
δ34S provides a good indicator of the extent of sulfate reduction. studies. Environmental Science and Technology 31, 3348–3357.
Blowes, D.W., Ptacek, C.J., Benner, S.G., McRae, C.W.T., Bennett, T.A., Puls, R.W.,
Thus δ34S measurements provide a useful tool for evaluating 2000. Treatment of inorganic contaminants using permeable reactive
the performance of remediation methods based on bacterial barriers. Journal of Contaminant Hydrology 45, 123–137.
sulfate reduction, such as organic carbon PRBs. Boursiquot, S., Mullet, M., Abdelmoula, M., Genin, J.M., Ehrhardt, J.J., 2001. The
dry oxidation of tetragonal FeS1 − x mackinawite. Physics and Chemistry
of Minerals 28, 600–611.
5. Conclusions Buchardt, B., Fritz, P., 1980. Environmental isotopes as environmental and
climatological indicators. In: Fritz, P., Fontes, J.Ch. (Eds.), Handbook of
The results of this study demonstrate that: Environmental Isotope Geochemistry, pp. 473–504.
Clark, I.D., Fritz, P., 1997. Environmental Isotopes in Hydrogeology. Lewis
Publishers, New York, USA. 328 pp.
• The Fe0-bearing organic carbon reactive mixture sustained a Cochran, W.G., 1950. Estimation of bacterial densities by means of the “most
higher sulfate reduction rate for a longer period of time than probable number”. Biometrics 6, 105–116.
Devereux, R., Hines, M.E., Stahl, D.A., 1996. S cycling: characterization of
the organic carbon reactive mixture. natural communities of sulfate-reducing bacteria by 16S rRNA sequence
• The microbial enumerations and isotopic measurements comparisons. Microbial Ecology 32, 283–292.
indicate that the sulfate reduction was mediated by SRB. Einsiedl, F., Schafer, T., Northrup, P., 2007. Combined sulfur K-edge XANES
spectroscopy and stable isotope analysis of fulvic acids and groundwater
• The cathodic production of H2 by anaerobic corrosion of Fe sulfate identify sulfur cycling in a karstic catchment area. Chemical Geology
is most likely the cause of the difference in sulfate reduction 238, 268–276.
rates between the two reactive mixtures. Gould, W.D., Stichbury, M., Francis, M., Lortie, L., Blowes, D.W., 2003. An MPN
Method for the Enumeration of Iron-reducing Bacteria, Mining and
• Zero-valent iron can be used to provide an electron donor in Environment Conference III, Sudbury, Ontario.
sulfate reducing PRBs and Fe0-bearing organic carbon Gu, B., Liang, L., Dickey, M.J., Yin, X., Dai, S., 1998. Reductive precipitation of
reactive mixture has a potential to improve the perfor- uranium (VI) by zero-valent iron. Environmental Science and Technology
32, 3366–3373.
mance of organic carbon PRBs.
Gu, B., Phelps, T.J., Liang, L., Dickey, M.J., Roh, Y., Kinsall, B.L., Palumbo, A.V.,
• The δ34S values could be used to determine the extent of Jacobs, G.K., 1999. Biogeochemical dynamics in zero-valent iron columns:
sulfate reduction, but more research has to be done to make implications for permeable reactive barriers. Environmental Science and
it a practical field method. Technology 33, 2170–2177.
Herbert Jr., R.B., Benner, S.G., Pratt, A.R., Blowes, D.W., 1998. Surface chemistry
• The δ13C values indicate that methanogenesis is occurring and morphology of poorly crystalline iron sulfides precipitated in media
in the front part of the FeOC column. containing sulfate-reducing bacteria. Chemical Geology 144, 87–97.
Herbert Jr., R.B., Benner, S.G., Blowes, D.W., 2000. Solid phase iron–sulfur
geochemistry of a reactive barrier for treatment of mine drainage.
Applied Geochemistry 15, 1331–1343.
Acknowledgements Hines, M.E., Evans, R.S., Genthner, B.R.S., Willis, S.G., Friedman, F., Rooney-
Varga, J.N., Devereux, R., 1999. Molecular phylogenetic and biogeochem-
ical studies of sulfate-reducing bacteria in the rhizosphere of Spartina
We thank Laura Groza, Jeff Bain, and Matt Lindsay for their alterniflora. Applied and Environmental Microbiology 65 (5), 2209–2216.
assistance with lab work and helpful comments. Mark Holser, W.T., Kaplan, I.R., 1966. Isotope geochemistry of sedimentary sulfates.
Biesinger's help in XPS analysis and interpretation and Paul Chemical Geology 1, 93–135.
Hulshof, A.H., Blowes, D.W., Ptacek, C.J., Gould, W.D., 2003. Microbial and
Northrup's support in XANES analysis are greatly appreciated.
nutrient investigations into the use of in situ layers for treatment of
Funding for this research was provided by the Natural tailings effluent. Environmental Science and Technology 37, 5027–5033.
Sciences and Engineering Research Council of Canada, Jambor, J.L., Raudsepp, M., Mountjoy, K., 2005. Mineralogy of permeable
reactive barriers for the attenuation of subsurface contaminants.
through a grant provided to D.W. Blowes. Use of the National
Canadian Mineralogist 43, 2117–2140.
Synchrotron Light Source, Brookhaven National Laboratory, Jones, C.F., LeCount, S., Smart, R.C., 1992. Compositional and structural
was supported by the U.S. Department of Energy, Office of alteration of pyrrhotite surfaces in solution: XPS and XRD studies.
Science, Office of Basic Energy Sciences, under Contract No. Applied Surface Science 55, 65–85.
Kaplan, I.R., Riggenberg, S.C., 1964. Microbiological fractionation of sulfur
DE-AC02-98CH10886. isotopes. Journal of General Microbiology 34, 195–212.
Kendall, C., Doctor, D.H., 2003. Stable isotope applications in hydrologic
studies. Treaties on Geochemistry, vol. 5. Elsevier, pp. 319–364.
References Liu, F., He, J., Colombo, C., Violante, A., 1999. Competitive adsorption of sulfate
and oxalate on goethite in the absence of presence of phosphate. Soil
Allison, J.D., Brown, D.S., Novo-Gradac, K.J., 1990. A Geochemical Assessment Science 164 (3), 180–189.
Model for Environmental Systems: Version 3.0 User's Manual, Environ- Logan, M.V., Reardon, K.F., Figueroa, L.A., McLain, J.E.T., Ahmann, D.M., 2005.
mental Research Laboratory, U.S. Environmental Protection Agency, Microbial community activities during establishment, performance, and
Athens, GA. 106 p. decline of bench-scale passive treatment systems for mine drainage.
Bain, J., Spink, L., Blowes, D.W., Smyth, D., 2003. The removal of arsenic from Water Research 39, 4537–4551.
groundwater using permeable reactive materials. Sudbury '03 — Mining Ludwig, R.D., McGregor, R.G., Blowes, D.W., Benner, S.G., Mountjoy, K., 2002. A
and the Environment III, Sudbury, Ontario. permeable reactive barrier for the treatment of dissolved metals. Ground
Ball, J.W., Nordstrom, D.K., 1991. User's manual for WATEQ4F, with revised Water 40 (1), 59–66.
thermodynamic data base and test cases for calculating speciation of Martens, C.S., Berner, R.A., 1977. Interstitial water chemistry of anoxic Long Island
major, trace, and redox elements in natural waters, US Geological Survey. Sound sediments: 1. Dissolved gases. Limnology and Oceanography 22,10–25.
Open-File Report 91-183. 192p, 2 tab, 103 ref, 6 append. McRae, C.W.T., Blowes, D.W., Ptacek, C.J., 1999. In Situ Removal of Arsenic
Benner, S.G., Blowes, D.W., Ptacek, C.J., 1997. A full-scale porous reactive wall from Groundwater Using Permeable Reactive Barriers: a Laboratory
for prevention of acid mine drainage. Ground Water Monitoring and Study. Sudbury '99 — Mining and the Environment II Conference,
Remediation 17 (4), 99–107. September 13–17, Sudbury, Ontario, pp. 601–609.
Q. Guo, D.W. Blowes / Journal of Contaminant Hydrology 107 (2009) 128–139 139

Megonigal, J.P., Hines, M.E., Visscher, P.T., 2003. Anaerobic metabolism: Scherer, M.M., Richter, S., Valentine, R.L., Alvarez, P.J.J., 2000. Chemistry and
linkages to trace gases and aerobic processes. Treaties on Geochemistry, microbiology of permeable reactive barriers for in situ groundwater
vol. 8. Elsevier, pp. 317–424. clean up. Critical Reviews in Microbiology 26 (4), 221–264.
Mountjoy, K.J., Blowes, D.W., 2002. Installation of a full scale permeable Schnurer, J., Rosswall, T., 1982. Fluorescein diacetate hydrolysis as a measure
reactive barrier for the treatment of metal contaminated groundwater. of total microbial activity in soil and litter. Applied and Environmental
Proceedings of the Third International Conference on Remediation of Microbiology 43, 1256–1261.
Chlorinated and Recalcitrant Compounds 2. InBattlle Press, Columbus, Strebel, O., Bottcher, J., Fritz, P., 1990. Use of isotope fractionation of sulfate–
Ohio, pp. 281–289. sulfur and sulfate–oxygen to assess bacterial desulfurication in a sandy
Pratt, A.R., Muir, I.J., Nesbitt, H.W., 1994a. X-ray photoelectron and Auger aquifer. Journal of Hydrology 121, 155–172.
electron spectroscopic studies pyrrhotite and mechanism of air oxida- Tempel, R.N., Shevenell, L.A., Lechler, P., Price, J., 2000. Geochemical modeling
tion. Geochimica et Cosmochimica Acta 58, 327–841. approach to predicting arsenic concentrations in a mine pit lake. Applied
Pratt, A.R., Nesbitt, H.W., Mycroft, J.R., 1994b. The increased reactivity of Geochemistry 15, 475–492.
pyrrhotite and magnetite phases in sulfide mine tailings. Journal of Thomas, J.E., Jones, C.F., Skinner, W.M., Smart, R.C., 1998. The role of surface
Geochemical Exploration 56, 1–11. sulfur species in the inhibition of pyrrhotite dissolution in acid
Rajagopal, B.S., LeGall, J., 1989. Utilization of cathodic hydrogen by hydrogen- conditions. Geochimica et Cosmochimica Acta 62, 1555–1565.
oxidizing bacteria. Applied Microbiology and Biotechnology 31, 406–412. Waybrant, K.R., Blowes, D.W., Ptacek, C.J., 1998. Selection of reactive mixtures
Ritter, K., Odziemkowski, M.S., Gillham, R.W., 2002. An in situ study of the role of for use in permeable reactive walls for treatment of mine drainage.
surface films on granular iron in the permeable iron wall technology. Environmental Science and Technology 32, 1972–1979.
Journal of Contaminant Hydrology 55, 87–111. Waybrant, K.R., Ptacek, C.J., Blowes, D.W., 2002. Treatment of mine drainage
Rooney-Varga, J.N., Devereux, R., Evans, R.S., Hines, M.E., 1997. Seasonal using permeable reactive barriers: column experiments. Environmental
changes in the relative abundance of uncultivated sulfate-reducing Science and Technology 36, 1346–1356.
bacteria in a salt marsh sediment and in the rhizosphere of Spartina Wilkin, R.T., Barnes, H.L., 1997. Formation processes of framboidal pyrite.
alterniflora. Applied and Environmental Microbiology 63, 3895–3901. Geochimica et Cosmochimica Acta 61, 323–339.

Вам также может понравиться