Vous êtes sur la page 1sur 5

Ecological Engineering 37 (2011) 989993

Contents lists available at ScienceDirect

Ecological Engineering
journal homepage: www.elsevier.com/locate/ecoleng

Short communication

Benthic transfer and speciation of mercury in wetland sediments downstream


from a sewage outfall
Hieu Van Duong, Seunghee Han
School of Environmental Science and Engineering, Gwangju Institute of Science and Technology (GIST), Gwangju 500-712, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 18 August 2010
Received in revised form 6 January 2011
Accepted 23 January 2011
Available online 4 March 2011
Keywords:
Benthic ux
Mercury
Monomethylmercury
Elemental mercury
Wetland
Organic matter

a b s t r a c t
In order to evaluate the role of hypoxic conditions of overlying water in the benthic ux and speciation of Hg, we analyzed sediment cores from hypoxic or oxic sites downstream from a sewage outfall in
the Damyang Riverine Wetland, Korea. Each core was analyzed for total Hg (THg), monomethylmercury
(MMHg), and elemental Hg (Hg0 ) from sediment, and for THg and MMHg from pore water. Hypoxic conditions of the overlying water near the sewage outfall were associated with a peak production of Hg0 , but
the lowest production of MMHg, in the upper 2 cm sediments. The benthic uxes of THg and MMHg were
estimated at 1302109 ng m2 day1 and 12 to 260 ng m2 day1 , respectively. The order of MMHg ux
from sediment to overlying water at each site did not follow the order of MMHg concentration in sediment, but was highest in hypoxic water conditions. The results suggest that maintaining oxic conditions
in wetland water is important for decreasing the transfer of MMHg from sediment into overlying water.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Riverine and coastal wetlands are rich in oral and faunal
species and provide nourishing habitats for aquatic ecosystems
(Trebitz et al., 2009). In wetland habitats, pollutants such as heavy
metals can be removed from surface waters by several different
processes including particle sedimentation, ltration by macrophytes, adsorption, and biological assimilation (El-Sheikh et al.,
2010; Sasmaz et al., 2008). Of these processes, sedimentation seems
to be dominant with respect to metal removal (Lung and Light,
1996; Walker and Hurl, 2002). Therefore, understanding the key
factors affecting metal transfer from sediment to water is important for prediction of the remobilization and bioaccumulation of
heavy metals deposited in wetland sediments.
Monomethylmercury (MMHg) is a toxic form of mercury (Hg)
that bioaccumulates and biomagnies within aquatic food webs.
In stagnant freshwater environments, such as at banks or wetlands, most MMHg is produced by biotic processes occurring in
anoxic sediments (Gilmour et al., 1992; King et al., 2002). In particular, sulfate-reducing bacteria are capable of converting inorganic
Hg(II) into MMHg using methyltransferase enzymes that are normal components of the acetate metabolic pathway (Choi et al.,
1994). Given the high biological activities of wetland microbial

Corresponding author. Tel.: +82 62 715 2438; fax: +82 62 715 2434.
E-mail address: shan@gist.ac.kr (S. Han).
0925-8574/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.ecoleng.2011.01.011

populations, wetland sediments can be major sources of MMHg


for downstream river water. Indeed, the proportion of wetland
area within a drainage basin has been shown to correlate with
MMHg levels in downstream lakes and streams (Driscoll et al.,
1994; Guentzel, 2009; Warner et al., 2005).
At the sediment surface, divalent Hg is reduced to elemental
mercury (Hg0 ), which is then released back to the water column
or adsorbed onto sediment particles. Although Hg0 may play a key
role in the exchange of Hg at the sediment-water interface, the
mechanism responsible for the reduction of divalent Hg in freshwater sediments is largely unknown (Bouffard and Amyot, 2009; Shi
et al., 2005). Possible reduction processes include Hg reduction by
UV near the sediment surface (Horvth and Vogler, 1998; Lalonde
et al., 2001), bacterial Hg detoxication using the mer operon
(Andra and Edmar, 2003), and Hg reduction by organic matter,
magnetite, and dissimilatory metal-reducing bacteria (Allard and
Arsenie, 1991; Wiatrowski et al., 2006, 2009). Recently, Bouffard
and Amyot (2009) reported that Hg0 constitutes a signicant portion (728%) of the total Hg (THg) in lake sediments, and that Hg0
adsorption to sediment is a fast reaction associated with the solid
organic matter content of the sediment. Nevertheless, the dominant Hg reduction processes in wetland sediments and the role of
hypoxic conditions of the water column have not yet been fully
established.
In the present study, we examined the sediment proles of THg,
Hg0 , and MMHg in cores collected from wetland sediments close to
a sewage outfall in the Damyang Riverine Wetland. Our aim was to

990

H.V. Duong, S. Han / Ecological Engineering 37 (2011) 989993

elucidate the effects of hypoxic conditions of the water column on


the benthic speciation of Hg. In addition, THg and MMHg concentrations in the overlying and the surface pore waters were measured
to determine the effects of hypoxia of the overlying water on the
benthic transfer of THg and MMHg.

were double-bagged and transported to the laboratory in a cooler.


In the laboratory, the surface water samples were acidied to 0.4%
(v/v) HCl solution and stored at 4 C until analysis.

2. Materials and methods

For the analysis of THg in the surface and pore water samples, each acidied water sample was oxidized with 0.22 M
BrCl solution (0.5% v/v), then subsequently reduced with 4.3 M
NH2 OHHCl (0.2% v/v) solution (US EPA, 2002). For sediment THg
analysis, approximately 0.5 g of freeze-dried sediment was oxidized overnight in PTFE bottles with 4 mL of 12 N HCl and 1 mL
of 14 N HNO3 . The THg in water samples and sediment digests
was detected using cold vapor atomic uorescence spectroscopy
(CVAFS; Choe et al., 2004). For quality control of the THg in the
water and sediment samples, BCR-579 (coastal seawater, IRMM)
and ERM-CC580 (estuarine sediment, IRMM), respectively, were
used as certied reference materials (CRM). The recoveries of the
water CRM were averaged to 98 9.7% (mean SD, n = 3), and those
of the matrix spike were averaged to 95% (n = 2). The recoveries of
the sediment CRM were averaged to 103 12% (n = 3), and those
of the matrix spike were averaged to 105 11% (n = 3). The relative percent difference for duplicate analyses of THg in water and
sediment were 11% (n = 7) and 14% (n = 22), respectively. Elemental mercury in the sediment samples was measured following the
methods described by Bouffard and Amyot (2009).
Monomethylmercury in the freeze-dried sediment samples
(0.51 g) was extracted into the organic phase following a reaction
with 5 mL of acidic KBr solution, 1 mL of 1 mol L1 CuSO4 solution,
and 10 mL of CH2 Cl2 (Bloom et al., 1999). After organic extraction,
MMHg was back-extracted from the organic phase into the aqueous phase using N2 gas, as described by Bloom et al. (1999). The
sediment extracts and water samples were analyzed for MMHg
by aqueous phase ethylation, trapping on a Tenax column, gas
chromatography separation, thermal decomposition, and detection by CVAFS (Bloom et al., 1999). The average recovery of the CRM
(ERM-CE464, tuna sh, IRMM) was 110 2% (n = 4), and that of the
matrix spike was 102 13% (n = 3). The relative percent difference
for duplicate analyses of MMHg in water and sediment averaged
19% (n = 7) and 18% (n = 22), respectively.

2.1. Study area


The Damyang Riverine Wetland (DRW), total area of 0.98 km2 ,
is located in the upper stream area of the Yeongsan River in Korea,
which belongs to Damyang-gun in Chollanam-do. The DRW is the
largest river marsh in Korea inhabited by protected wild animals
(NWC, 2010). Compared to other major rivers in Korea, the average
bed slope of the Yeongsan River is very gentle (0.0011); therefore,
sedimentation is active even in upstream areas, including the DRW.
The geomorphic landscape of this area consists of streambed wetlands, oodplains, and piedmont hills (NWC, 2010). The sampling
site was Namsan Wetland (from 35 17 45 to 35 18 56 and from
126 57 15 to 126 58 09 ), one of six sub-wetlands in the DRW,
that has a total area of 0.26 km2 (NWC, 2010).
Sediment and overlying water were sampled on November 11,
2009. Because Korea is in the Asian Monsoon Climate region, the
sampling season was a dry season (late autumn to spring) associated with a relatively low ow rate of 179,000 m3 day1 , measured
at a location lose to the sampling sites (KWRC, 2010). The sampling
sites (A, B, C, B1, and C1) were located downstream of the Damyang
sewage treatment plant (STP), which has a sewage outow rate of
6200 m3 day1 . The control site was located approximately 1 km
upstream of the STP. Among the ve sampling sites, Site A was
located closest to the STP (<10 m) and Site C was the furthest downstream. The distances between Sites A and B, and between Sites B
and C, were approximately 1 km. The water depth of the sampling
sites was approximately 50100 cm. A variety of woody plants,
including cattails (Typha), giant reed (Phragmites), and sweet ag
(Acorus), grew at Sites A, B, and C. The dominant species growing during the sampling season was Phragmites japonica, a large
perennial grass. Sites B1 and C1 were located at the same distance from the STP as Sites B and C, respectively, but had no woody
plant populations. Biolms were visible at Sites B and C, and abundant benthic macrofauna were not noticeable from the sampling
sites.
2.2. Sediment core and surface-water sampling
Sediment cores were collected from Sites A, B, C, B1, and C1 using
acid-cleaned acrylic corers (12 cm in diameter, 30 cm in length).
The cores were transported to the laboratory in a cooler and sliced
into segments of 02, 24, 46, 68, and 810 cm from the sediment surface. Slicing was done in a glove bag lled with N2 . Pore
water was extracted by centrifugation (3000 rpm for 20 min) after
the cores had been sectioned, and then approximately 2030 mL
of each ltered pore water sample was acidied (0.4%, v/v HCl) for
analysis of dissolved THg and MMHg. The remaining sediment was
stored freeze-dried until analysis. The sediment depth of core C1
was approximately 6 cm; therefore, core C1 was sliced into two
segments: 02 and 24 cm.
Surface water samples were collected from Sites A, B, C, B1, C1,
and at a sewage discharge pipe using an ultra-clean sampling protocol (Gill and Bruland, 1990). Water samples were collected at a
depth of approximately 20 cm using a peristaltic pump and acidcleaned Teon tubing with a PES (polyethersulfone) cartridge lter
(0.45 m pore size) connected to the tubing exit. Water samples

2.3. Analysis of THg and MMHg

2.4. Analysis of ancillary parameters


Water quality parameters (i.e., pH, dissolved oxygen concentration, and temperature) were measured at a water depth of
20 cm using a multi-function water quality meter. Organic matter
in sediment was estimated by %LOI (loss on ignition): freeze-dried
sediment samples were heated in a furnace at 450500 C for 12 h,
and weight losses were measured using a microbalance.
Correlations between factors were tted with linear regression
model established and tested for 95% signicance using SigmaPlot
10.0. Unless otherwise stated, values represent means with one
standard deviation.
3. Results and discussion
3.1. Organic matter content controls the sediment distribution of
THg
Total Hg (THg) in the top 2 cm of sediment decreased from
1450 ng g1 (Site A) to 144 (B1), 111 (B), 98.9 (C) and 65.9 (C1)
ng g1 (Table 1). These THg values, except for Site C1, were higher
than the background concentration found from the control site
(61 3.4 ng g1 , n = 3) and that for uncontaminated river sediments

H.V. Duong, S. Han / Ecological Engineering 37 (2011) 989993

991

Table 1
Concentrations of THg, loss on ignition (LOI), MMHg, %MMHg/THg, Hg0 , and %Hg0 /THg in sediment cores from the Damyang Riverine Wetlands. Temperature of the overlying
water was 17 C.

B1

C1

02
24
46
68
810
02
24
46
68
810
02
24
46
68
810
02
24
46
68
810
02
24

THg (ng g1 )
1450
314
123
331
174
111
138
684
180
114
98.9
83.7
139
59.2
33.2
144
139
65.2
182
115
65.9
39.9

LOI (%)

MMHg (pg g1 )

MMHg/THg (%)

Hg0 (ng g1 )

Hg0 /THg (%)

2.0
0.72
1.1
1.1
0.74
2.8
1.7
1.4
1.3
1.7
2.2
2.2
1.3
0.96
0.27
4.6
3.5
2.0
1.4
0.89
1.0
0.80

407
46.9
53.9
352
56.3
186
594
163
96.6
133
828
432
263
23.8
11.1
454
623
364
274
53.0
757
181

0.028
0.015
0.044
0.11
0.032
0.17
0.43
0.024
0.054
0.12
0.84
0.52
0.19
0.040
0.033
0.32
0.45
0.56
0.15
0.046
1.1
0.45

985
269
80.8
201
32.4
28.8
15.4
591
62.3
7.80
2.90
0.700
38.0
34.1
11.9
45.5
3.30
3.30
94.5
55.4
26.4
3.70

68
86
67
61
19
26
11
86
34
6.9
2.9
0.83
27
57
36
31
2.3
4.9
52
48
40
9.3

in Korea (50 ng g1 , Park et al., 2009). The THg concentration measured near the STP discharge point (Site A) was similar to that
of sediments contaminated with Hg from anthropogenic activities (5008000 ng g1 ; Biester et al., 2002; Gray et al., 2006). The
sediment concentrations of THg at Site A showed a surface peak,
suggesting higher Hg deposition in recent years, which agrees with
the discharge record of the STP: the STP began releasing sewage in
1999.
In general, the LOI concentrations showed a decreasing trend
with depth (Table 1), which is typical for LOI in freshwater sediments, and indicates active carbon mineralization near the surface
of the sediment (Bouffard and Amyot, 2009). While the LOI for
Site B1 also decreased with depth, the surface concentration of
LOI was higher at Site B1 than at Sites A, B, C, or C1 which is in
agreement with the sediment composition at each site (A and C1:
sand, B and C: silty sand, and B1: sandy silt; Shepards classication). Organic matter content is known to inuence the surface
distribution of sediment THg due to a strong association between
organic matter and THg (Warner et al., 2005; Delongchamp et al.,
2009). Indeed, the surface distribution of THg showed a positive
linear relationship with %LOI for Sites B, C, B1 and C1 (Fig. 1). Site
A showed higher THg concentration (1450 ng g1 ) than expected
from the %LOI (2.0%) due to the treated sewage discharge. For other
depths, no signicant correlation was found between THg and %LOI
(r2 = 0.11, p = 0.14).
3.2. Enhanced %MMHg/THg in sediment under oxic water
conditions

the MMHg proles were typical of wetland sediments: higher at


the surface and subsurface, and decreasing with depth (King et al.,
2001).
In the upper 2 cm of sediment, the proportion of MMHg in THg
(%MMHg/THg) showed an increasing trend with distance from the
STP, with the highest measurements at Site C and C1, and the lowest
at Site A (Table 1). The lowest %MMHg/THg at Site A was associated
with hypoxic conditions of the overlying water A. It is possible that
the low %MMHg/THg at Site A is caused by sulde inhibition of
Hg methylation potential (Benoit et al., 2001; Drott et al., 2007;
Han et al., 2007, 2008). An inverse correlation between the sulde
concentration and Hg methylation rates has been observed in Everglades sediments (FL, USA), which indicated that sulde produced
from in situ sulfate reduction inhibited microbial Hg methylation
(Gilmour et al., 1998). Indeed, a strong hydrogen sulde odor was
noted when we sliced core A.

0-2cm
2-4cm
4-6cm
6-8cm
8-10cm

r2=0.99
p<0.001

%LOI

Depth (cm)

Due to the treated sewage discharges from the STP close to Site
A, overlying waters at this site were hypoxic on the sampling date:
DO concentrations were 51 M (Site A), 250 M (B), 260 M (C),
240 M (B1), and 240 M (C1). Monomethylmercury in the upper
2 cm of sediment ranged from 186 to 828 pg g1 (Table 1). The
highest MMHg content was detected at the Site C, which was furthest downstream from the STP, and the lowest MMHg content
was detected at Site B, which is in contrast to the THg distribution. According to this order, MMHg discharged from the STP is
not the major source of the MMHg deposited in the surface sediment, which indicates the importance of in situ production within
the sediment (Gilmour et al., 1998; King et al., 2002). Vertically,

20

40

60

80

100

120

140

160

180

200

THg (ng g -1)


Fig. 1. Relationship between THg and loss on ignition (LOI) in sediment cores of
Sites A, B, C, B1, and C1. The linear regression model for 02 cm (%LOI = 0.046 [THg]
2.2) includes Sites B, C, B1, and C1.

992

H.V. Duong, S. Han / Ecological Engineering 37 (2011) 989993

700

A
B
B1
C
C1

600

-1

400

Hg (ng g )

500

300

r2=0.95
p<0.001
100

200

50

100
0

200

400

50

600

100

150

800

200

1000

THg (ng g-1)

Fig. 3. Benthic uxes of THg and MMHg in the Damyang Riverine Wetlands, Korea.

Fig. 2. Linear regression between the THg and Hg0 in sediment cores of Sites A, B,
C, B1, and C1 ([Hg0 ] = 0.92 [THg] 69).

3.3. Enhanced %Hg0 /THg in sediment under hypoxic water


conditions
Elemental Hg concentrations in surface sediment (02 cm) were
signicantly higher at Site A than at the other sites (Table 1).
The range of Hg0 concentrations at Sites B, B1, C, and C1
(2.9045.5 ng g1 ) was similar to that found in pristine lake sediments (660 ng g1 , Bouffard and Amyot, 2009), whereas the Hg0
concentration of Site A was similar to that found in Hg mines in
Texas (Gray et al., 2006). The positive linear relationship between
Hg0 and THg in sediments, shown in Fig. 2, suggested that the
THg loading was the major factor in the overall control of the Hg0
content in wetland sediment.
The sediment proles of %Hg0 /THg showed various distributions: mean 60% for Site A, 33% for Site B, 28% for Site B1, 25% for
Sites C and C1 (Table 1). In the literature, the %Hg0 /THg found in
pristine lake sediment was between 7% and 25%, a range more or
less similar to Sites B, B1, C, and C1 (Bouffard and Amyot, 2009).
The highest mean fraction of Hg0 for Site A appeared to be associated with the hypoxic conditions of the overlying water of Site
A. Recently, Wiatrowski et al. (2006, 2009) reported the reduction
of Hg(II) to Hg0 by metal-reducing bacteria and by the Fe(II)/Fe(III)
mixed valence iron oxide mineral, magnetite. For example, reaction of Hg(II) with magnetite resulted in the formation of Hg0 due
to the oxidation of Fe(II) to Fe(III), and the Hg(II) reduction rate
increased with increasing magnetite surface area. These previous
results, together with those from the current study, suggest that
Hg0 production occurs in situ in wetland sediment and may be
associated with an iron redox process.
3.4. Increased diffusion ux of MMHg under hypoxic water
conditions
In the absence of bioirrigation, the diffusive ux of solutes in
pore water is determined using Ficks rst law (1):
F =

Dw C

x
2

(1)

where F (ng m2 day1 ) is ux of a solute with concentration C


(ng L1 ) at x cm depth of pore water, and  is sediment porosity ( = (Mw /w )/[(Ms /s ) + (Mw /w )]), where Mw is the amount of
water loss through drying, Ms is dry sediment weight, w is water
density, and s is sediment density (Covelli et al., 1999). Sediment

porosities were experimentally calculated, and  is a tortuosity


term that can be calculated by (1 ln(2 )) (Boudreau, 1996). For the
diffusion coefcient, 5.0 106 cm2 s1 was used for THg (Gobeil
and Cossa, 1993; Mason et al., 1993) and 1.2 105 cm2 s1 was
used for MMHg, which is a value for CH3 HgSH (Covelli et al., 2008).
The nal diffusive ux of THg ranged between 130 and
460 ng m2 day1 among the vegetative sites (A, B, and C) and
between 790 and 2100 ng m2 day1 among the non-vegetative
sites (B1 and C1; Fig. 3). The range of THg ux for these
sites are lower than THg ux found at a coastal sh farm
(690041,000 ng m2 day1 , Covelli et al., 2008), but is much
higher than the ux in natural estuarine and coastal sediments
(241 ng m2 day1 , Covelli et al., 1999; Gill et al., 1999). The range
of benthic ux of MMHg (12 to 260 ng day1 m2 ) shown in Fig. 3
agrees with the range of MMHg ux in freshwater wetland sediments (420 to 290 ng day1 m2 ; Holmes and Lean, 2006; Goulet
et al., 2007). Considering that MMHg concentration in sediments
was highest at Site C, the extent of MMHg ux demonstrates that
mobility of MMHg from sediment to overlying water is associated
with the geochemical conditions of sediment and overlying water
(Gill et al., 1999; Holmes and Lean, 2006). Sediment composition, as
well as organic matter content, was not radically different between
each site; therefore, redox conditions of the overlying water may
be related to the degree of MMHg ux. A negative correlation
between benthic ux of THg (and MMHg) and DO concentration
in overlying water has been suggested previously (Covelli et al.,
2008; Ramalhosa et al., 2006; Rothenberg et al., 2008). While the
mechanism of inhibition by DO is not yet clearly understood, it is
likely that DO in overlying water enhances co-precipitation of Hg
and MMHg with Fe(III)- and Mn(IV)-oxyhydroxide in the sediment
surface, thereby preventing diffusion of Hg and MMHg from pore
water to overlying water (Canrio et al., 2003; Qumerais et al.,
1998).

Acknowledgments
We thank Dr. Sun-Baek Bang for technical support. We appreciate the support of Hyunji Kim, Jiyi Jang, and Choonho Lee during
sample collection and preparation. This study was supported by
the Ministry of Education, Science and Technology, Korea, through
the Institute of Science and Technology for Sustainability (UNU &
GIST Joint Program). The research was also supported by a National
Research Foundation of Korea Grant, funded by the Korean government (2009-0074529).

H.V. Duong, S. Han / Ecological Engineering 37 (2011) 989993

References
Allard, B., Arsenie, I., 1991. Abiotic reduction of mercury by humic substances in
aquatic systeman important process for the mercury cycle. Water Air Soil
Pollut. 56, 457464.
Andra, M.A.N., Edmar, C.S., 2003. Operon mer: Bacterial resistance to mercury and
potential for bioremediation of contaminated environments. Genet. Mol. Res. 2,
92101.
Benoit, J.M., Gilmour, C.C., Mason, R.P., 2001. Aspects of bioavailability of mercury for
methylation in pure cultures of desulfobulbus propionicus (1pr3). Appl. Environ.
Microbiol. 67, 5158.
Biester, H., Mller, G., Schler, H.F., 2002. Binding and mobility of mercury in soils
contaminated by emissions from chlor-alkali plants. Sci. Total Environ. 284,
191203.
Bloom, N.S., Gill, G.A., Cappellino, S., Dobbs, C., McShea, L., Driscoll, C., Mason, R.,
Rudd, J., 1999. Speciation and cycling of mercury in Lavaca Bay, Texas, Sediments.
Environ. Sci. Technol. 33, 713.
Boudreau, B.P., 1996. The diffusive tortuosity of ne-grained unlithied sediments.
Geochim. Cosmochim. Acta 60, 31393142.
Bouffard, A., Amyot, M., 2009. Importance of elemental mercury in lake sediments.
Chemosphere 74, 10981103.
Canrio, J., Vale, C., Caetano, M., Madureira, M.J., 2003. Mercury in contaminated sediments and pore waters enriched in sulphate (Tagus Estuary Portugal). Environ.
Pollut. 126, 425433.
Choe, K.Y., Gill, G.A., Lehman, R.D., Han, S., Heim, W.A., Coale, K.H., 2004. Sedimentwater exchange of total mercury and monomethyl mercury in the San Francisco
Bay-Delta. Limnol. Oceanogr. 49, 15121527.
Choi, S.-C., Chase Jr., T., Bartha, R., 1994. Metabolic pathways leading to mercury methylation in Desulfovibrio desulfuricans LS. Appl. Environ. Microbiol. 60,
40724077.
Covelli, S., Faganeli, J., Horvat, M., Brambati, A., 1999. Porewater distribution and
benthic ux measurements of mercury and methylmercury in the Gulf of Trieste
(Northern Adriatic Sea). Estuarine Coast. Shelf Sci. 48, 415428.
Covelli, S., Faganeli, J., De Vittor, C., Predonzani, S., Acquavita, A., Horvat, M., 2008.
Benthic uxes of mercury species in a lagoon environment (Grado Lagoon,
Northern Adriatic Sea Italy). Appl. Geochem. 23, 529546.
Delongchamp, T.M., Lean, D.R.S., Ridal, J.J., Blais, J.M., 2009. Sediment mercury
dynamics and historical trends of mercury deposition in the St. Lawrence
River area of concern near Cornwall, Ontario Canada. Sci. Total Environ. 407,
40954104.
Driscoll, C.T., Yan, C., Schoeld, C.L., Munson, R., Holsapple, J., 1994. The mercury
cycle and sh in the Adirondack lakes. Environ. Sci. Technol. 28, 136A143A.
Drott, A., Lambertsson, L., Bjrn, E., Skyllberg, U., 2007. Importance of dissolved
neutral mercury suldes for methyl mercury production in contaminated sediments. Environ. Sci. Technol. 41, 22702276.
El-Sheikh, M.A., Saleh, H.I., El-Quosy, D.E., Mahmoud, A.A., 2010. Improving water
quality in polluted drains with free water surface constructed wetlands. Ecol.
Eng. 36, 14781484.
Gill, G.A., Bruland, K.W., 1990. Mercury speciation in surface freshwater systems in
California and other areas. Environ. Sci. Technol. 24, 13921400.
Gill, G.A., Bloom, N.S., Cappellino, S., Driscoll, C.T., Dobbs, C., McShea, L., Mason,
R., Rudd, J.W.M., 1999. Sediment-water uxes of mercury in Lavaca Bay Texas.
Environ. Sci. Technol. 33, 663669.
Gilmour, C.C., Henry, E.A., Mitchell, R., 1992. Sulfate stimulation of mercury methylation in freshwater sediments. Environ. Sci. Technol. 26, 22812287.
Gilmour, C.C., Riedel, G.S., Ederington, M.C., Bell, J.T., Benoit, J.M., Gill, G.A., Stordal,
M.C., 1998. Methylmercury concentrations and production rates across a trophic
gradient in the northern Everglades. Biogeochemistry 40, 327345.
Gobeil, C., Cossa, D., 1993. Mercury in sediments and sediment pore water in the
Laurentian Trough. Can. J. Fish. Aquat. Sci. 50, 17941800.
Gray, J.E., Hines, M.E., Biester, H., 2006. Mercury methylation inuenced by areas
of past mercury mining in the Terlingua district, Southwest Texas USA. Appl.
Geochem. 21, 19401954.
Guentzel, J.L., 2009. Wetland inuences on mercury transport and bioaccumulation
in South Carolina. Sci. Total Environ. 407, 13441353.

993

Goulet, R.R., Holmes, J., Page, B., Poissant, L., Siciliano, S.D., Lean, D.R.S., et al.,
2007. Mercury transformations and uxes in sediments of a riverine wetland.
Geochim. Cosmochim. Acta 71, 33933406.
Han, S., Obraztsova, A., Pretto, P., Choe, K.Y., Gieskes, J., Deheyn, D.D., Tebo, B.M.,
2007. Biogeochemical factors affecting mercury methylation in sediments of
the Venice Lagoon Italy. Environ. Toxicol. Chem. 26, 655663.
Han, S., Obraztsova, A., Pretto, P., Deheyn, D.D., Gieskes, J., Tebo, B.M., 2008. Sulde
and iron control on mercury speciation in anoxic estuarine sediment slurries.
Mar. Chem. 111, 214220.
Holmes, J., Lean, D., 2006. Factors that inuence methylmercury ux rates from
wetland sediments. Sci. Total Environ. 368, 306319.
Horvth, O., Vogler, A., 1998. Photoreduction of mercury(II) in aqueous solution
in the presence of cyclohexene hydroxomercuration and two-stage photolysis.
Inorg. Chem. Commun. 1, 270272.
King, J.K., Kostka, J.E., Frischer, M.E., Saunders, F.M., Jahnke, R.A., 2001. A quantitative
relationship that demonstrates mercury methylation rates in marine sediments
are based on the community composition and activity of sulfate-reducing bacteria. Environ. Sci. Technol. 35, 24912496.
King, J.K., Harmon, S.M., Fu, T.T., Gladden, J.B., 2002. Mercury removal, methylmercury formation, and sulfate-reducing bacteria proles in wetland mesocosms.
Chemosphere 46, 859870.
KWRC (Korea Water Resources Corporation), 2010, http://english.kwater.or.kr/.
Lalonde, J.D., Amyot, M., Kraepiel, A.M.L., Morel, F.M.M., 2001. Photooxidation of
Hg(0) in articial and natural waters. Environ. Sci. Technol. 35, 13671372.
Lung, W.S., Light, R.N., 1996. Modelling copper removal in wetland ecosystems. Ecol.
Model. 93, 89100.
Mason, R.P., Fitzgerald, W.F., Hurley, J., Hanson Jr., A.K., Donaghay, P.L., Sieburth, J.M.,
1993. Mercury biogeochemical cycling in a stratied estuary. Limnol. Oceanogr.
38, 12271241.
NWC (National Wetland Center), 2010, http://www.wetland.go.kr/en/main.html.
Park, J.S., Lee, J.S., Kim, G.B., Cha, J.S., Shin, S., Kang, H.G., Hong, E.J., Chung, G.T.,
Kim, Y.H., 2009. Mercury and methylmercury in freshwater sh and sediments
in South Korea using newly adopted purge and trap GCMS detection method.
Water Air Soil Pollut. 207, 391401.
Qumerais, B., Cossa, D., Rondeau, B., Pham, T.T., Fortin, B., 1998. Mercury distribution in relation to iron and manganese in the waters of the St Lawrence river.
Sci. Total Environ. 213, 193201.
Ramalhosa, E., Segade, S.R., Pereira, E., Vale, C., Duarte, A., 2006. Mercury cycling
between the water column and surface sediments in a contaminated area. Water
Res. 40, 28932900.
Rothenberg, S.E., Ambrose, R.F., Jay, J.A., 2008. Mercury cycling in surface water, pore
water and sediments of Mugu Lagoon CA, USA. Environ. Pollut. 154, 3245.
Sasmaz, A., Obek, E., Hasar, H., 2008. The accumulation of heavy metals in Typha
latifolia L. grown in a stream carrying secondary efuent. Ecol. Eng. 33, 278284.
Shi, J., Liang, L., Jiang, G., Jin, X., 2005. The speciation and bioavailability of mercury
in sediments of Haihe River. Chin. Environ. Int. 3, 357365.
Trebitz, A.S., Brazner, J.C., Pearson, M.S., Peterson, G.S., Tanner, D.K., Taylor, D.L.,
2009. Patterns in habitat and sh assemblages within Great Lakes coastal
wetlands and implications for sampling design. Can. J. Fish. Aquat. Sci. 66,
13431354.
United States Environmental Protection Agency, 2002. Method 1631, Revision E:
Mercury in Water by Oxidation, Purge and Trap and Cold Vapor Atomic Fluorescence Spectrometry. USEPA 821-R-02-019, Washington, DC.
Walker, D.J., Hurl, S., 2002. The reduction of heavy metals in a stormwater wetland.
Ecol. Eng. 18, 407414.
Warner, K.A., Bonzongo, J.C.J., Roden, E.E., Ward, G.M., Green, A.C., Chaubey, I., Lyons,
W.B., Arrington, D.A., 2005. Effect of watershed parameters on mercury distribution in different environmental compartments in the Mobile Alabama River
Basin, USA. Sci. Total Environ. 347, 187207.
Wiatrowski, H.A., Ward, P.M., Barkay, T., 2006. Novel reduction of mercury(II) by
mercury-sensitive dissimilatory metal reducing bacteria. Environ. Sci. Technol.
40, 66906696.
Wiatrowski, H.A., Das, S., Kukkadapu, R., Ilton, E.S., Barkay, T., Yee, N., 2009. Reduction of Hg(II) to Hg(0) by magnetite. Environ. Sci. Technol. 43, 53075313.

Vous aimerez peut-être aussi