Sunteți pe pagina 1din 13

Ecological Engineering 69 (2014) 93–105

Contents lists available at ScienceDirect

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

Removal and fate of arsenic in the rhizosphere of Juncus effusus


treating artificial wastewater in laboratory-scale constructed
wetlands
Khaja Zillur Rahman a,∗ , Arndt Wiessner b , Peter Kuschk b , Manfred van Afferden a ,
Jürgen Mattusch c , Roland Arno Müller a
a
Centre for Environmental Biotechnology, UFZ−Helmholtz Centre for Environmental Research, Permoserstrasse 15, 04318 Leipzig, Germany
b
Department of Environmental Biotechnology, UFZ−Helmholtz Centre for Environmental Research, Permoserstrasse 15, 04318 Leipzig, Germany
c
Department of Analytical Chemistry, UFZ−Helmholtz Centre for Environmental Research, Permoserstrasse 15, 04318 Leipzig, Germany

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

Article history: The deposition, fate and distribution of arsenic (As) under dynamic redox conditions within the rhi-
Received 24 October 2013 zosphere of helophytes in treatment wetlands are still poorly understood. For this purpose, long-term
Received in revised form 30 January 2014 experiments were carried out in specially designed laboratory-scale constructed wetland reactors treat-
Accepted 29 March 2014
ing artificial domestic wastewater containing As (200 ␮g As l−1 ) in order to investigate the key aspects
of As immobilization, to identify the main As removal pathway by using a mass balance approach and
Keywords:
to assess the role of different sulfate (SO4 2− ) concentrations on As mass retention. The results with a
Arsenic removal
highly efficient As mass retention (>92%) indicated a better performance under C-deficient and oxidized
Artificial wastewater
Constructed wetland
conditions (Eh ∼324–795 mV) regardless to the SO4 2− concentration in the inflow wastewater. An ele-
Juncus effusus vated SO4 2− concentration (25 mg S l−1 in the inflow) facilitated high As-retention (>90%) under C-surplus
Mass balance and microbial dissimilatory SO4 2− reducing condition (Eh ∼−225–−149 mV) within the root-near envi-
Rhizosphere ronment of the rhizosphere in constructed wetlands. Mean pH in a range of 6.6–7.7 might be favoring
the immobilization of As but a comparatively low pH (3.9–5.9) within the root vicinity might enhance
plant uptake. In general, higher As concentrations were exhibited by the plant roots (90–315 mg As kg−1
dry wt) as compared to the shoots (3.5–3.8 mg As kg−1 dry wt). Nearly 3.5-fold higher As concentrations
within the roots from the experimental reactor as compared to the roots collected from control reactor
clearly indicated that a higher amount of As was retained, accumulated, adsorbed, metabolized to other
forms on root surface and/or translocated into the roots of Juncus effusus, where organic C and SO4 2− were
abundant. Based on As mass balance calculation, the reactor with the highest SO4 2− loading was found
to be retained nearly 85% of the total As mass input. Out of which only <1% of the total inflow As mass
was sequestered or translocated into the plant shoots, 42.2% was accumulated/recovered within the plant
roots, 17.2% was entrapped or deposited within the sediments of the gravel bed, 16.2% was recovered in
the pore water and 15.3% was flushed out as outflow. The remaining 9% was considered as unaccountable,
which might be released due to volatilizations or lost due to various unknown reasons. A 5-fold higher
SO4 2− concentration within the reactor might facilitate lower pH (3.9–5.9) and consequent remobiliza-
tion caused a higher amount of free or exchangeable As in the pore water (16.2%), that probably resulted
in a higher As uptake (42.2%) by the plant roots as compared to the roots from the control reactor (only
13%). The findings demonstrate the deposition and fate of As within the rhizosphere, which are of high
importance for an efficient treatment of wastewater containing As under constructed wetland conditions.

© 2014 Elsevier B.V. All rights reserved.

∗ Corresponding author. Tel.: +49 341 235 1019/179 940 9506; fax: +49 341 235 1830.
E-mail address: khaja.rahman@ufz.de (K.Z. Rahman).

http://dx.doi.org/10.1016/j.ecoleng.2014.03.050
0925-8574/© 2014 Elsevier B.V. All rights reserved.
94 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

1. Introduction developed (Rai et al., 1995). Fitz and Wenzel (2002) proposed
that hyperaccumulators may enhance metal solubility in the rhi-
Arsenic (As) is a toxic metalloid which can pollute water, soil, zosphere via root exudation, consequently increasing plant metal
crops and the environment at large, ultimately affecting human uptake. Larios et al. (2012) showed that the plants accumulated
health (Zhao et al., 2010). More than 245 minerals contain As, and extremely high amounts of total As in their tissues which varied
the principal source of As is geological. However, human activities depending on the part of the plant, with roots accumulating the
such as mining, pesticide application, and burning of fossil fuels most As in all the studied plants (up to 1400 mg kg−1 dry wt). In the
also cause As pollution (Sharma and Sohn, 2009). In recent years, context of constructed wetlands, García et al. (2010) reported that
there has been an increasing contamination of water, soil and crops the direct uptake and accumulation of As in plants appears to play a
by this metalloid in many regions of the world (Tripathi et al., 2007), very minor role in As removal. The same conclusion was drawn by
particularly in some countries of southern Asia (Meharg, 2004). It is Singhakant et al. (2009), who reported that only 0.5–1% of the total
therefore very important to choose appropriate methods to control As input was accumulated in plant tissues. However, there are also
As in the environment. Several treatment technologies have been studies indicating that wetland plants have a remarkable effect on
applied for the removal of As from contaminated waters, such as As retention (Rahman et al., 2011; Sasmaza and Obek, 2009).
coagulation/filtration, ion exchange, lime softening, adsorption on First results of laboratory-scale investigations showed the trans-
iron oxides or activated alumina, reverse osmosis etc. (Zouboulis formation processes and redox dynamics of As-species particularly
and Katsoyiannis, 2002). in the near-root environment of the Juncus effusus in model con-
Constructed wetlands are low-energy based ‘green’ technolo- structed wetlands. Changes in dynamic redox conditions and
gies that have been increasingly applied in wastewater treatment re-oxidation of reduced sulfur into other S species (e.g. S0 , SO4 2– )
since the mid-1980s (Sun and Saeed, 2009) and have consid- caused a total sulfur enrichment and a consequent As remobiliza-
erable potential to remove metals and metalloids, including As tion within the rhizosphere in this study (Rahman et al., 2008b).
(Buddhawong et al., 2005; Rahman et al., 2008a; Ye et al., 2003). But it is necessary to investigate the role of organic C, S and pH
Wetland plants have been shown to play important roles in con- on As retention within the root vicinity and enhanced plant uptake
structed wetlands to remove As from wastewater (Rahman et al., under different redox conditions. Differences between correspond-
2008a, 2011; Singhakant et al., 2009). Complex interactions of ing inflow and outflow data of total As indicated remarkable
As under redox gradient (both micro- and macro) conditions amounts of As immobilization within the rhizosphere. Therefore, in
have already been investigated in different laboratory-scale hor- addition to the investigations of As-removal efficiency and dynam-
izontal subsurface-flow constructed wetlands treating an artificial ics of As-species (Rahman et al., 2008b), it is necessary to deepen
wastewater (Rahman et al., 2008a). the understanding of the deposition, fate and mass balance of
The rhizosphere of constructed wetlands offers specific macro- As within the micro-scale root zone environment of the rhizo-
and micro gradients of redox conditions enabling the development sphere in treatment wetlands. However, Rahman et al. (2011) in
of highly diverse microbial consortia capable of different beneficial another study showed the fate and distribution of As along the
redox reactions (Bezbaruah and Zhang, 2004; Liesack et al., 2000; flow path of a laboratory-scale horizontal subsurface-flow con-
Wiessner et al., 2005b). Particularly due to the release of oxygen structed wetland. But the knowledge regarding the accumulation
and organic carbon at the same time by the roots of helophytes and mass balance of As under the micro-scale gradient of redox
into the rhizosphere, spatial and temporal micro-scale gradients conditions and the role of C, S and pH within the rhizosphere
of oxygen concentrations and redox states are established close of helophytes is still insufficient. In the past, little attention has
to the root surfaces. These conditions enable the development of been paid and virtually no information is available until now that
microbial mats and layers of functionally different microorgan- directly addresses the removal, fate and plant uptake of As under
isms which simultaneously realize multiple interactive processes dynamic redox conditions within the rhizosphere of constructed
like nitrification, denitrification, mineralization of organic carbon, wetland.
methanogenesis, reduction and oxidation of several sulfur and In the context of our research work in this field, several inves-
As compounds on a small spatial scale (Bezbaruah and Zhang, tigations like fundamental aspects and mechanisms of As fixation,
2004; Darrah et al., 2006; Rahman et al., 2008b; Wiessner et al., influences of dynamic redox conditions on As biotransformation
2005b). Recently, the application of a specially designed laboratory- processes, stability of As within the planted and unplanted wet-
scale constructed wetland (Kappelmeyer et al., 2002) in order to lands, bioaccumulation and uptake of As in plant biomass, sludge
evaluate micro-gradient processes within the near-root environ- sediment analysis, mass balance of As and S, etc., were carried out
ment of the rhizosphere was shown to be useful (Wiessner et al., in different laboratory-scale constructed wetlands (Rahman et al.,
2005a,b). 2008a,b; Rahman et al., 2011). In principle, the major objectives
The behavior of metals in aquatic systems is complex and may of this study were i) to investigate the fate, accumulation and dis-
include interactions among or between the major wetland com- tribution of As under redox dynamic conditions on a micro-scale
partments, above-ground plant parts, roots, litter, biofilms, soil, gradient within the rhizosphere; ii) to use a mass balance approach
and water (Kadlec and Knight, 1996). Volatilization of metals into to identify the main As removal pathway under “ideal flow” con-
a gaseous phase occurs with mercury, selenium, and arsenic to a ditions; and iii) to assess the role of organic C, pH and different
lesser degree. Dissolved metals can adsorb onto particles, or exist SO4 2− concentrations on As mass retention within the rhizo-
complexes to inorganic and organic ligands, or be present in solu- sphere of helophytes in treatment wetlands. By using Juncus effusus,
tion in the free-ion state. The adsorption and co-precipitation of long-term experiments in a specially designed laboratory-scale
As on hydrous oxides of Fe, Mn or Al oxides and Fe sulfides is an constructed wetland treating an artificial wastewater containing
important sink for As immobilization (Jacks et al., 2002). Moreover, As (200 ␮g l−1 ) were performed to evaluate all these processes. The
dissimilatory reduction caused by iron- and sulfate-reducing bacte- results of this study may help to better understand the key aspects
ria is widely considered the primary mechanism responsible for the of As immobilization within the plant root-zone of the rhizosphere
rapid As reduction and release observed in anaerobic environments under constructed wetland conditions and to optimize manage-
(Islam et al., 2004; Kirk et al., 2004). ment practices for maximum As retention in different wetland
Based on the characteristic of metal hyperaccumulation in compartments (shoots, roots, sediments etc.) through a complete
plants, suitable and sustainable remediation strategies could be mass balance analysis.
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 95

compared with the control reactor R1, which was constantly fed
with limited or very low inflow SO4 2− concentration.

2.2. Experimental conditions

An artificial wastewater (Wiessner et al., 2005a) simulated a


typical secondary effluent of domestic wastewater was used in
this investigation. The inflow concentrations of the used ingre-
dients were (in mg l−1 ): 204.9 CH3 COONa, 107.1 C6 H5 COONa,
117.8 NH4 Cl, 28 K2 HPO4 , 7 NaCl, 3.4 MgCl2 ·6H2 O, 4 CaCl2 ·2H2 O,
0.89/22.2/44.4/110.9 Na2 SO4 , 0.2 As(V) (Titrisol® As2 O5 in H2 O,
Merck, Germany) and 1 ml l−1 of trace mineral solution which
was adapted from Buddhawong et al. (2005). The resulting con-
centrations of the parameters were (in mg l−1 ): 340 chemical
oxygen demand (COD), 122.3 total organic carbon (TOC), 30.8
ammonia-N, 5 phosphate-P, 0.2/5/10/25 sulfate-S and 0.2 As(V).
These compounds were dissolved in deionized water and fed to
the corresponding reactors in different experimental phases.
Artificial wastewater was freshly prepared in every 3 days
to prevent microbial degradation during storage and operation
Fig. 1. Schematic diagram of a specially designed laboratory-scale constructed wet- (Fig. 1). N2 gas was vigorously purged through and bubbled out of
land (planted fixed bed reactor–PFBR): (1) feeding storage tank; (2) pump; (3) glass
the liquid phase of the wastewater for approximately 20−25 min
vessel; (4) distribution chamber; (5) suction cylinder; (6) gravel bed; (7) recir-
culation pump; (8) magnetic valve; (9) on-line measurement; (10) outflow; (11) after each preparation of fresh feeding solution in order to remove
plants. any traces of dissolved oxygen. Each reactor was fed separately
(Adapted from Kappelmeyer et al., 2002). from a feeding glass bottle with a volume of 10 l. In order to keep
an anoxic environment inside the feeding glass bottle contain-
ing the artificial wastewater, a continuous purging of nitrogen gas
(N2 ) through the headspace of the feeding bottles was maintained
2. Materials and methods throughout the whole operation period (Fig. 1).
The reactors were operated for several months (nearly 4–5
2.1. Experimental design: Planted fixed-bed reactors months) using similar artificial wastewater (Wiessner et al., 2005a)
but without any SO4 2− or As to establish biofilms and microbiolog-
The experiments were carried out in three laboratory-scale ical activity. Within these 4–5 months of normal operation, new
planted-fixed bed reactors (R1, R2 and R3), which were estab- shoots were grown from those five plants in each reactor. Prior to
lished under conditions of complete mixing of the pore water by start of our experiment by using simulated wastewater with As and
a permanent circulation. Since the internal flow conditions were SO4 2− , the total numbers of healthy green shoots were counted as
comparable to an ideal mixed vessel, and therefore, macro-scale 231, 245 and 241 in R1, R2 and R3, respectively. The duration of
gradients of concentrations were equalized and the effects of the the main experimental period was 340 days for all three reactors
micro-gradient changes could be determined. The design and the in this study. At the end of the experimental period of 340 days, the
principles of operation of the reactor (PFR – planted fixed bed reac- total shoot numbers were decreased down to a minimum of 180,
tor) have been described previously in detail (Kappelmeyer et al., 211 and 198 in R1, R2 and R3, respectively.
2002; Wiessner et al., 2005a,b). The reactors were fed and run under five different experimental
Briefly, the rhizosphere of the reactors was represented by a phases (phase I, II, III, IV and V) realized by varying SO4 2− concen-
rooted gravel bed (particle size 2–4 mm) in a glass vessel of 28 cm trations in the simulated wastewater inflow solution. The operation
diameter and a height of 30 cm. Each reactor was planted with five conditions in all the phases within the reactors R1, R2 and R3
rush plants (Juncus effusus) with an initial total shoot number of are listed in Table 1. The hypothesis behind different experimen-
72, 84 and 80 in the reactors R1, R2 and R3, respectively. The reac- tal phases was to investigate how the redox dynamics of As was
tors were closed tightly with a Teflon lid containing five circular influenced under C-deficient and oxidized condition with simulta-
openings through which the plants were grown in the gravel bed neous increasing of SO4 2− concentrations in each reactors (phase I),
(Fig. 1). The glass reactors were covered to prevent algal growth. under C-surplus and reducing condition with both constant (in R1)
The free pore water volume in the planted beds amounted to 10 and simultaneous increasing and/or decreasing of SO4 2− concen-
l and the hydraulic retention time was adjusted to 5 days. The trations (phase II, III, IV and V) in the reactors R2 and R3. Each phase
reactors were placed in a greenhouse and operated under defined had a sufficient duration to guarantee a representative number of
environmental conditions to simulate an average summer day in a samples that could be taken from each reactor.
moderate climate (Kappelmeyer et al., 2002; Wiessner et al., 2005b, In experimental phase I (duration of first 83 days), the reac-
2008). The temperature was set to 22 ◦ C from 6 a.m. to 9 p.m. to tors were fed by a continuous inflow of As-contaminated artificial
simulate daytime and to 16 ◦ C at night. One lamp (Master SON- wastewater (200 ␮g As l−1 ) associated with SO4 2− concentrations
PIA 400 W, Phillips, Belgium) was switched on during daytime as of 0.2, 5 and 25 mg S l−1 but without organic C in the reactor R1,
an additional artificial light source whenever the natural light fell R2 and R3, respectively. The resulted molecular ratio of SO4 2− -S to
below 1110 ␮mol m−2 s−1 (Wiessner et al., 2013). As(V) varied as 1:1, 25:1 and 125:1 in the inflow wastewater of the
The main reason to use three model reactors operating in par- reactors R1, R2 and R3, respectively.
allel was to investigate the influences of simultaneously increasing In experimental phase II (from day 83 to day144), the reactors
inflow feeding SO4 2− concentrations on the As mass retention, were fed and operated under conditions of surplus C by adding
both under C-deficient and C-surplus conditions in the respective organic C-sources (resulting a COD of ∼340 mg l−1 ) along with
reactors. Obtained results from the reactors R2 and R3 were then As(V) and only traces of SO4 2− (0.2 mg S l−1 ) in the simulated
96 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

Table 1
Operation conditions within the model reactors R1, R2 and R3 accomplished by defined arsenic, organic carbon and sulfate inflow concentrations of the artificial wastewater
(phase I–V).

Reactor Parameter Unit Experimental phases

I n II n III n IV n V n

R1 As (V) ␮g l−1 202 ± 8 12 205 ± 6 10 196 ± 4 13 202 ± 5 15 203 ± 6 6


SO4 2− -S mg l−1 0.2 ± 0.1 12 0.2 ± 0.1 10 0.2 ± 0.1 13 0.2 ± 0.1 15 0.2 ± 0.1 6
TOC mg l−1 bdl 12 120 ± 4 10 118 ± 10 13 120 ± 9 15 128 ± 4 6
R2 As (V) ␮g l−1 204 ± 5 12 201 ± 5 10 199 ± 8 13 201 ± 6 15 201 ± 4 6
SO4 2− -S mg l−1 5±1 12 0.2 ± 0.1 10 5 ± 1 13 5 ± 1 15 bdl 6
TOC mg l−1 bdl 12 124 ± 5 10 120 ± 9 13 125 ± 8 15 125 ± 5 6
R3 As (V) ␮g l−1 198 ± 10 12 199 ± 10 10 201 ± 7 13 199 ± 7 15 204 ± 7 6
SO4 2− -S mg l−1 25 ± 3 12 0.2 ± 0.1 10 25 ± 5 13 10 ± 2 15 bdl 6
TOC mg l−1 bdl 12 126 ± 7 10 118 ± 11 13 123 ± 6 15 126 ± 6 6

bdl: below the detection limit (<1 mg TOC l−1 ; <0.1 mg SO4 2− -S l−1 ).

wastewater inflow solution of all three reactors. Similar running SO4 2− was analyzed by ion chromatography (DIONEX 100,
conditions were prevailing at the end of this phase and the resulted columns AS4A-SC/AG4A-SC and CS12A/CG12A; Idstein, Germany)
molecular ratio of SO4 2− -S to As(V) was 1:1 in the reactors. and a conductivity detector. TOC was analyzed by using a TOC ana-
In phase III (from day 144 to 235), the same C-dosage and SO4 2− lyzer (Shimadzu, TOC 600, Duisburg, Germany).
concentration as like in phase II was maintained in the control reac- After the termination of the experiment, plant biomass sam-
tor R1 and only the SO4 2− concentration was increased to 5 and ples (shoots and roots) and sludge sediments were collected from
25 mg S l−1 in the inflow feeding solutions of the reactors R2 and each reactor in order to investigate the potential for As removal
R3, respectively. During experimental phase IV (corresponding to efficiency. Plant samples were sectioned into their shoot and root
day 235 to 315), conditions in all three reactors were similar to components after collecting them from each reactor. The roots
the inflow feeding solution of phase III operation, but only excep- were first thoroughly washed with tap water and then with deion-
tion was the concentration of SO4 2− in reactor R3. Instead of a ized water to remove any gravel aggregate or sludge sediment. The
SO4 2− concentration of 25 mg S l−1 (as in phase III), the amount was plant shoots and roots collected from each reactor were freshly
decreased to 10 mg S l−1 in this particular phase IV in the reactor R3 weighed, dried at 105–108 ◦ C for 3 days, allowed to cool, and then
(Table 1). The newly resulting molecular ratio of SO4 2− -S to As(V) the dry weights were determined and the water content was cal-
was established as 50:1 under this experimental phase in reactor culated. These dry weights are used throughout the text unless
R3. Maintaining the same operating conditions as like in Phase IV, otherwise specified. The dried samples were ground to a fine pow-
the supply of the inflow SO4 2− was completely stopped in reactor der using a mortar and pestle under liquid nitrogen in order to
R2 and R3 (phase V) until the termination of the experiment on day obtain a homogeneous sample and then preserved in sealed plastic
340. A comparative analysis was carried out in between the reac- bottles for analysis.
tors under each circumstance and also within different operation For the analysis of the total As concentration in plant
phases. biomass (shoots and roots), the homogenized powdered sam-
Plant transpiration represents 98% of the total water loss ples were digested by microwave extraction (PE Anton Paar
(Wiessner et al., 2005a) and was measured by balancing the inflow GmbH, Graz, Austria). For digestion, 2.0 ml of digestion mixture
and the outflow amounts of water once a week. The flow balances (HNO3 :HCl = 4:1) were added to 0.5 g powder in a Teflon pres-
were also used to control and adjust the inflow rate. All green shoots sure bomb and heated to 260 ◦ C for 1 h. After the digest cooled
from a length of at least approximately 2 cm were counted once a down, it was filled with deionized water to a total volume of 10 ml,
month. mixed and filtered using a 0.45 ␮m syringe filter (Satorius AG,
Goettingen, Germany). The filtrate solution was analyzed for total
As by using hydride generation atomic absorption spectrometry
2.3. Sample collection and analysis (HG-AAS) with a detection limit of 0.3 ␮g As l−1 . Acid blanks were
analyzed in order to assess possible contamination. All analyses
Samples for total As, total organic carbon (TOC) and sulfate were performed in duplicate.
(SO4 2− ) analysis were taken once a week. The preservation tech- Analysis of the sludge sediment samples was performed by the
nique of the collected samples for measuring the total As content energy dispersive X-ray fluorescence (EDXRF) spectrometer XLAB
and the analytical procedure of hydride generation atomic adsorp- 2000 (SPECTRO Instruments) running with the software package
tion spectrometry (HG-AAS) have already been described by Daus XLAB Pro 2.2. Collected sludge sediments from each reactor were
et al. (2002) and Schmidt et al. (2004). dried at 105 ◦ C for 24 h using oven MA4O (Satorius, Germany) and
The volatile As species were analyzed by gas chromatog- were ground by means of an agate ball mill (Retsch). Well-ground
raphy/mass spectrometry (GC–MS, Shimadzu-GC-17A-Shimadzu sample material (1 g) was mixed with stearine wax (Hoechst,
GC-MS-Qp5000) with electron ionization and quadrupole analyzer, Germany) for XRF analysis as a binder in a ratio of 80:20 (w/w)
using the method described by Pantsar-Kallio and Korpela (2000). and subsequently pressed at 150 MPa to pellets (with an internal
The analyses were performed isothermally at 50 ◦ C, and helium was diameter of 32 mm) and analyzed by means of energy dispersive X-
used as a carrier gas. ray fluorescence analysis (EDXRF). The mean value of two replicates
Due to the reactor design, the circulation flow represents the was calculated. The relative error of the method was 2–3%.
actual concentration of the pore water inside of the reactor and Experimental results were evaluated statistically (e.g. mean,
thus the pH and redox potential (Eh ) were controlled continuously standard deviation, bar graphs error bar) by using spread-
in the circulation flow and the data were recorded online twice per sheet program Microsoft Excel (Microsoft Corporation) in this
hour (Fig. 1). study.
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 97

2.4. As mass balance calculation

After 340 days of the experiment, a complete mass balance of As


was investigated in each reactor by considering the total As mass
input, the total As mass output, and the total As retained in the plant
biomass (the shoots and the roots), in the pore water and in the
sludge sediments. The remaining (loss or gain) of the As mass from
the mass balance calculation was considered to be unaccountable.
The total As mass input and output in each reactor was calculated
from the cumulative total As mass inflow and outflow during the
whole operation time period (Fig. 2). A simple As mass balance
was computed according to the following Eq. (1) (Singhakant et al.,
2009):

Asin = Asout + Asplant + Assed. + Aspw + Asunaccount (1)

Where Asin is total As mass in the inflow (g); Asout is total As


mass in the outflow (g); Asplant is total As mass in the plant biomass
(g), including plant roots and shoots; Assed. is total As mass retained
in the sludge sediment (g); Aspw is total As mass retained in the
pore water or standing water (g); Asunaccount. is the total As mass
that was unaccountable (g), including the loss or gain of As from
the mass balance calculation.

3. Results

3.1. As removal efficiency


Fig. 2. Cumulative total As mass inflow and retention in the three laboratory-scale
Arsenic removal efficiencies in terms of total amount of the As reactors (R1, R2 and R3) during the whole experimental period of 340 days.
mass inflow and the total As retention in all three corresponding
reactors during the whole study period of 340 days is shown in
Fig. 2. Under highly oxidized conditions in phase I within the first 83
A total inflow mass of 140.0, 135.1 and 123.3 mg As was fed and a days of operation, the corresponding mean removal rate was cal-
total of 72.8, 27.7 and 18.9 mg As mass was flushed out through the culated as 0.388 ± 0.012, 0.384 ± 0.015 and 0.347 ± 0.06 mg As per
outlet of reactors R1, R2 and R3, respectively. Therefore, a cumula- day and thus contributed to a mean As retention of 94%, 94% and
tive total mass of 67.2, 107.4 and 104.4 mg As was retained, which 92%, in the reactors R1, R2 and R3, respectively (Table 2). Addition
resulted in nearly 48%, 80% and 85% of the total As mass retention of organic C-sources in phase II (day 83–144) showed an immedi-
in the corresponding reactors R1, R2 and R3, respectively (Fig. 3). ate effect with a considerably high As concentration in the outflow,
During the stoppage period of SO4 2− -S supply (in phase V), no which resulted in a lower As retention of only 15, 28 and 61% in
remobilization or loss of the total As mass was observed in the the reactors R1, R2 and R3, respectively. During the experimen-
reactors, R2 and R3. tal phase III (from day 144 to day 235) under anaerobic condition

Table 2
Summary of the treatment performance in the reactors R1, R2 and R3 during the whole operational period of 340 days (phase I–V).

Phase Duration (day) Parameter Unit Reactor

R1 R2 R3

I 0–83 Inflow rate mg/day 0.411 ± 0.002 0.408 ± 0.002 0.379 ± 0.065
Outflow rate mg/day 0.023 ± 0.012 0.023 ± 0.016 0.032 ± 0.014
Removal % 94 94 92
n – 12 12 12
II 83–144 Inflow rate mg/day 0.415 ± 0.001 0.408 ± 0.001 0.364 ± 0.056
Outflow rate mg/day 0.352 ± 0.093 0.294 ± 0.091 0.146 ± 0.066
Removal % 15 28 61
n – 10 10 10
III 144–235 Inflow rate mg/day 0.411 ± 0.009 0.404 ± 0.009 0.326 ± 0.052
Outflow rate mg/day 0.278 ± 0.028 0.045 ± 0.025 0.033 ± 0.010
Removal % 32 89 90
n – 13 13 13
IV 235–315 Inflow rate mg/day 0.410 ± 0.002 0.404 ± 0.002 0.393 ± 0.009
Outflow rate mg/day 0.243 ± 0.021 0.017 ± 0.011 0.021 ± 0.013
Removal % 41 96 95
n – 15 15 15
V 297–340 Inflow rate mg/day 0.412 ± 0.001 0.403 ± 0.001 0.397 ± 0.001
Outflow rate mg/day 0.276 ± 0.038 0.037 ± 0.020 0.019 ± 0.003
Removal % 33 91 95
n – 6 6 6

n: number of samples.
98 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

All three reactors showed an efficient TOC removal (>74%) dur-


ing the experimental phases with the addition of organic C-sources.
Interestingly, a highly efficient TOC removal in the range of 86–94%
was observed in the reactors R2 and R3, as compared to control
reactor R1 (with a range of 74–86%).

3.3. Redox (Eh ) and pH

Table 4 shows the summary of the mean pH and redox val-


ues (also the ranges) in different experimental phases within the
reactors R1, R2 and R3. No addition of organic C-sources ensured a
Fig. 3. As mass retention capacity of the model reactors R1, R2 and R3 calculated persistent aerobic condition (with an Eh ∼ 324–795 mV) inside the
as a percentage of the inflow total As mass that retained inside the reactors during reactors during experimental phase I (first 83 days).
the whole experimental period of 340 days.
Reducing conditions were established immediately after the
addition of organic C-sources (122.3 mg TOC l−1 ) with a drop
down of Eh value from an aerobic (721–795 mV) to anaerobic
Table 3 (−217–−144 mV) condition within all three corresponding reac-
Removal efficiency of SO4 2− -S and TOC within the model reactors R1, R2 and R3 in
different experimental phases (phase I–V).
tors in experimental phase II (Table 4). But a simultaneous addition
of SO4 2− with constant C-dosage resulted in a low but relatively
Reactor Parameter Removal efficiency (%) stable redox potential value (Eh ∼ −225–−149 mV) within all three
Phase I Phase II Phase III Phase IV Phase IV reactors in phase III–V.
R1 SO4 2− -S − − − − −
Mean pH value of 3.1 ± 1.2, 3.6 ± 0.6 and 3.2 ± 0.4 was observed
TOC − 75 86 74 75 at the beginning (in phase I) and after addition of C-dosage in phase
R2 SO4 2− -S − − 73 68 − II, the pH remarkably raised to a mean value 4.3 ± 0.4, 6.6 ± 0.7
TOC − 86 94 93 92 and 5.9 ± 0.5 in the pore water of the corresponding reactors R1,
R3 SO4 2− -S − − 84 71 −
R2 and R3, respectively (Table 4). Afterwards, a relatively steady
TOC − 89 88 89 93
pH-value was observed in the reactors during the remaining exper-
imental phases. Interestingly, a decrease of nearly 2–3 pH units was
recorded in R3 than in the reactor R2 during the phases III–V.
along with varying SO4 2− concentration, the data demonstrated a
comparatively much higher As retention than the previous phase II
3.4. Concentration of total As in plant biomass (shoots and roots)
with a mean value 89% and 90% in R2 and R3, respectively and only
32% was retained in the control reactor R1 (Table 2). Traces (2–3 ␮g
After the termination of all experimental phases, the sampling
As l−1 ) of volatile As compound [gaseous arsine (AsH3 )] was found
and measurement of plant biomass (shoots and roots) in terms of
in this experimental phase but only in the reactor R3. Data in Table 2
fresh and dry weight, water content (%) and thereby the dry weight
also shows that at least 5% more As was retained within reactor R3
(in kg) of total shoots and roots in each reactor were carried out and
when SO4 2− concentration was 10 mg S l−1 (in phase IV) instead of
all these measurements are shown in Table 5. We measured a rela-
25 mg S l−1 (in phase III).
tively lower total root biomass of 0.93 kg (fresh wt) collected from
reactor R3, as compared to the other two reactors R1 and R2 with
3.2. SO4 2− -S and TOC removal efficiency 1.10 and 1.16 kg (fresh wt), respectively. Based on the water con-
tent analysis data from the plant biomass samples collected from
Table 3 shows the removal efficiency of SO4 2− -S and TOC in the reactors R1, R2 and R3, the dry weight of the total shoots were
different experimental phases within the reactors R1, R2 and R3. 0.16, 0.18, and 0.14 kg, and the corresponding dry weights of the
Overall, a comparatively higher SO4 2− -S removal was observed in total roots were calculated as 0.20, 0.21, and 0.16 kg, respectively
R3 (71–84%) in comparison to the reactor R2 (68–73%) under C- (Table 5). The analytical results of the total As concentrations in
surplus conditions (phase II–V). As expected, no SO4 2− reduction plant biomass are presented in Fig. 4. The mean total As concentra-
was observed under C-deficient and oxidized conditions in exper- tion was measured as 3.88 ± 0.13, 3.62 ± 0.07 and 3.57 ± 0.08 5 mg
imental phase I. As kg−1 (dry wt) in the shoots and was considerably very high in

Table 4
Summary of the average pH, redox (Eh ) values (mean ± SD) and ranges of the values (from minimum to maximum) in each reactor during the whole operational period
(phase I–V).

Experimental phases

Reactor Parameter I II III IV V

R1 pH 3.1 ± 1.2 4.3 ± 0.4 4.9 ± 1.4 6.7 ± 0.1 6.9 ± 0.1
(1.1–5.0) (3.4–4.8) (3.5–6.6) (6.6–6.8) (6.8–7.1)
Eh (mV) 499 ± 203 28 ± 162 −187 ± 6 −182 ± 6 −191 ± 4
(324–740) (−168–307) (−196–−179) (−189–−169) (−196–−188)
R2 pH 3.6 ± 0.6 6.6 ± 0.7 7.4 ± 0.1 7.6 ± 0.1 7.7 ± 0.1
(3.2–5.1) (5.1–7.4) (7.2–7.5) (7.4–7.8) (7.6–7.8)
Eh (mV) 597 ± 88 −97 ± 176 −217 ± 4 −219 ± 5 −216 ± 10
(396–721) (−217–277) (−225–−213) (−225–−210) (−224–−202)
R3 pH 3.2 ± 0.4 5.9 ± 0.5 5.8 ± 0.2 5.2 ± 0.5 3.9 ± 0.1
(2.8–4.1) (5.1–6.5) (5.5–6.1) (4.4–5.6) (3.8–4.1)
Eh (mV) 670 ± 88 −52 ± 139 −195 ± 17 −201 ± 13 −199 ± 23
(495–795) (−144–259) (−211–−149) (−215–−172) (−220–−166)
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 99

Table 5
Plant analysis data in terms of fresh and dry weight of shoots and roots samples, water content and total dry weight of collected shoots and roots biomass (in kg) from each
reactor after the termination of the experiment.

Reactor Shoots Roots

Sample Sample dry Water Total fresh Total dry Sample Sample dry Water Total fresh Total dry
fresh wt (g) wt (g) content (%) wt (kg) wt (kg) fresh wt (g) wt (g) content (%) wt (kg) wt (kg)

R1 1.393 0.512 63 0.421 0.156 11.440 2.029 82 1.100 0.195


R2 2.130 0.689 68 0.575 0.184 7.620 1.352 82 1.160 0.207
R3 2.236 0.860 62 0.361 0.137 9.817 1.728 82 0.930 0.164

Table 6
Distribution of total As in each reactor, which includes the total inflow and outflow mass of As and the distribution in different compartments (shoots, roots, sediment, etc.)
over the whole operation period of 340 days. Data are given in mg As, and the values in parentheses are percentile amounts of total inflow As.

Reactor Total inflow (mg As) Shoots (mg As) Roots (mg As) Sediment (mg As) Pore water (mg As) Unaccountable (mg As) Outflow (mg As)

R1 140.0 (100%) 0.6 (0.4%) 17.6 (12.6%) 6.2 (4.4%) 6.5 (4.6%) 36.5 (26.1%) 72.8 (52.0%)
R2 135.1 (100%) 0.5 (0.4%) 22.9 (17.0%) 30.9 (22.9%) 13.9 (10.3%) 39.1 (28.9%) 27.7 (20.5%)
R3 123.3 (100%) 0.4 (0.3%) 51.6 (41.9%) 21.2 (17.2%) 20.0 (16.2%) 11.1 (9.0%) 18.9 (15.3%)

the roots as 90 ± 2.83, 110.5 ± 3.54, 315 ± 4.24 mg As kg−1 (dry wt) than the concentration found in the sediments collected from the
within the reactors R1, R2 and R3, respectively. On the contrary, reactors R2 and R3.
the mean As concentration was found to be only 0.5 ± 0.14 and
2.25 ± 0.2 mg As kg−1 (dry wt) in the control shoot and root sam- 3.6. Data for As mass balance
ples of J. effusus (collected from natural, non-contaminated source),
respectively. Table 6 represents a summary of the total As mass deposition
and distribution in the different wetland compartments (shoots,
roots, sediments etc.) and a complete mass balance calculation
3.5. Concentration of the total As in the sediment
within the reactors over the whole operation period of 340 days.
Considering the inflowing As mass within the reactors as 100%, the
The results from the collected sediments showed a mean total As
percentile value of the outflowing As mass and retention was esti-
concentration of 206 ± 5, 806 ± 9 and 941 ± 4 mg As kg−1 (dry wt) in
mated. For instance, a total mass of 140.0 mg As (100%) was fed
the corresponding three reactors R1, R2 and R3, respectively (Fig. 4).
as an inflow, and a total mass of 72.8 mg As (52%) was flushed out
In comparison, a mean As concentration of 3.2 ± 0.4 mg As kg−1 (dry
of the reactor R1 as outflow. The recovered As masses in the pore
wt) in the control sediment sample (collected from a natural, non-
water, within the shoots, the roots and in the sediments collected
contaminated source) was shown to be nearly 251–294 times lower
from this reactor (R1) were measured as 6.5, 0.6, 17.6 and 6.2 mg
As, which resulted in a nearly 4.6%, 0.4%, 12.6% and 4.4% of the total
inflowing As mass retention in these compartments, respectively.
The remaining 36.5 mg As (nearly 26% of the total inflow As mass)
was considered to be unaccountable or retained in uncounted sinks.
Similarly, after calculating the total As mass in the plant shoots,
roots and sediments of other two reactors, it was observed that
nearly 17% and 42% of the total inflowing As mass were accumu-
lated/recovered/concentrated within the roots and nearly 23% and
17% were entrapped or deposited within the gravel bed (as sedi-
ments) of the reactors R2 and R3, respectively (Fig. 5). Only 9% of the
total inflow As mass was calculated as unaccountable within R3 in
comparison with the reactor R2, where nearly 29% was considered
to be retained in uncounted sink or termed as unaccountable. In all
three reactors, only a very little amount (<1% of the inflow As mass)
was translocated into the shoots of J. effusus in this experiment.

4. Discussion

4.1. As mass retention capacities of the model reactors

Overall, the total As mass retention capacities of the reactors


R2 and R3 were much higher (80% and 85%, respectively) than
the control reactor R1 (only 48%). In general, SO4 2− loading was
much higher in the reactors R2 and R3, as compared to the reac-
tor R1. Therefore, amount of SO4 2− is playing an important role in
case of As mass retention within the root-near environment of the
rhizosphere in treatment wetlands.
A highly efficient As mass retention (>92%) indicated a
Fig. 4. Concentrations of total As in plant shoots (A), roots (B) and sediments (C)
collected from the reactors R1, R2 and R3 (the values are the mean of two replicates better performance under C-deficient and oxidized conditions
and the error bars are standard deviations). (phase I) regardless to the concentration of SO4 2− in the inflow
100 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

This is also in agreement with the fact that under reducing condi-
tions, Fe(III) is reduced to Fe(II) and resulting in the mobilization of
some of the adsorbed As, particularly from sediments and the plant
root-zone (Kneebone et al., 2002). Moreover, reduction of As(V)
to more mobile As(III) and subsequent As(III) enrichment due to
microbial activity might also facilitated As remobilization (Rahman
et al., 2008b). Therefore, it was evident in this study that the addi-
tion of electron donor has a huge impact on As mobility within
the rhizosphere of helophytes in treatment wetlands. Presumably
there were several competing reactions including dissolution or
desorption, precipitation or adsorption and remobilization occur-
ring simultaneously under such conditions.
The occurrence of sulfide (S2– ) under anoxic conditions may
immobilize As due to its high affinity to sulfide minerals or the
formation of As-sulfide minerals (Bostick and Fendorf, 2003). This
might be the reason of highly efficient As mass retention (>89%)
under anaerobic condition by varying SO4 2− concentrations in R2
and R3 (phase III; Table 2). Other authors also suggested that under
reducing environments and in the presence of S and Fe, As can form
insoluble sulfide compounds (Buddhawong et al., 2005; Singhakant
et al., 2009a), such as orpiment As2 S3 , in which As is present as
As(III) and arsenopyrite (FeAsS). Dissimilatory reduction caused by
iron- and sulfate-reducing bacteria is widely considered as the pri-
mary mechanism responsible for the rapid As reduction and release
observed in anaerobic environments (Islam et al., 2004). Recently,
Mattes et al. (2010) provided more details of the wetland system
described in Duncan et al. (2004), highlighting that not only sulfate-
reducing bacteria played a role in As removal, but iron-oxidizing
bacteria also making a significant contribution.
Addition of relatively lower SO4 2− concentration (10 mg S l−1 in
phase IV) contributed to 5% more As retention within the reactor R3
as compared to the SO4 2− concentration in phase III (25 mg S l−1 ).
From this, it can also be concluded that a very high SO4 2− loading
does not necessarily improve As removal performances within the
rhizosphere of constructed wetlands.

4.2. Correlation of Redox (Eh ), pH and As mass retention


Fig. 5. Mass balance of As estimated as a percentage of inflow total As mass that
retained in different wetland compartments investigated in the model reactors R1,
R2 and R3. The mobility and plant availability of many trace and toxic
metals and metalloids in wetland soils are often governed by
oxidation-reduction (redox) potential and associated pH in the
artificial wastewater of all three reactors (see Fig. 2 and Table 2). rhizosphere (Gambrell, 1994), while the oxygen released from
Highly oxidized conditions (Eh ∼ 324–795 mV) were not favor- the roots probably influences the redox and pH of rhizosphere
able for microbial dissimilatory SO4 2− reduction and therefore no (Yang et al., 2010). Flooding conditions induce an enrichment of
SO4 2− reductions or removal were found in the corresponding metals in soils surrounding the roots of wetland plants (Wright
reactors (see phase I; Table 3). Hence, it can be concluded that and Otte, 1999). Under C-deficient condition in this study, a very
the immobilization of As within the rhizosphere of helophytes high redox potential (with an Eh ∼ 324–795 mV) were obtained
was accomplished by mechanisms other than arsenic-sulfide inside the reactors, which clearly indicated an aerobic condition in
precipitation (likely as As2 S3 ). Adsorption and/or concomitant co- the root-near environment of the rhizosphere. Aerobic conditions
precipitation of As, specifically with Fe(III) oxyhydroxides was the were unfavorable for SO4 2− reducing bacteria and therefore it can
most probable reasons for As immobilization under such condi- be concluded that adsorption and co-precipitation of As, specifi-
tions (Bednar et al., 2005; Rahman et al., 2008b). Highly oxidized cally predominant and thermodynamically more stable As(V) with
conditions due to plant root-mediated O2 release and re-oxidation iron(oxy)hydroxide favored As retention under such oxic condi-
of reduced As-species probably influenced higher sorption and pre- tions.
cipitation reactions within the rhizosphere of Juncus effusus. Reducing conditions play an important role for SO4 2− reduc-
Addition of organic C-sources probably caused an immediate tion, which requires a reducing environment and an electron donor.
mobilization of As within the reactors under limited SO4 2− and Under C-surplus condition along with available SO4 2− (in R2 and
reducing conditions (Rahman et al., 2008b), and hence a compara- R3), the Eh values varied in between −149 and −225 that resulted
tively low As removal efficiency was observed during experimental in a higher SO4 2− removal (68–84%) and a highly efficient As mass
phase II in all three reactors (Table 2). Under such conditions, there retention (>89%) within the reactors in this study (phase III–IV;
was probably a shift in terminal electron acceptors from O2 to Tables 2–4). Different authors have reported different Eh values
Fe(III) oxyhydroxides. Thereby, reductive dissolution of Fe(III) oxy- required by sulfate-reducing bacteria to thrive: less than −200 mV
hydroxides mediated by microbial activities probably contributed (Cabrera et al., 2006), less than −100 mV (Willow and Cohen,
to a release of immobilized As fraction in the aqueous phase of 2003), or between −150 and −200 mV (Tuttle, 1969). Rahman
all the corresponding reactors (Smedley and Kinniburgh, 2002). et al. (2008a) noted that microbial SO4 2− reduction was greater
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 101

under redox potential values between −160 and −190 mV, and might be favoring the immobilization of liable and exchangeable As
that higher SO4 2− removal contributed to higher removal of As. fractions in the rhizosphere but a lower pH within the root vicinity
However, these conclusions were drawn based on their measured of constructed wetlands might consequently enhance plant uptake.
water quality parameters (Eh , TOC), without monitoring microbial Therefore, a pH change may affect the bioavailability of As and may
community composition or function directly. Microorganisms can also play a vital role for plant As uptake within the rhizosphere.
enhance the removal of As by mediating redox and precipitation For example, Wells and Richardson (1985) reported a decrease in
processes (Lizama et al., 2011). Talukder et al. (2012) showed that arsenate uptake in the moss Hylocomium splendens with increas-
under aerobic water management (Eh ∼ 135–138 mV), As uptake ing pH. In this moss, arsenate uptake was optimal at pH 5, where
by the rice plant parts was significantly less as compared to anaer- H2 AsO4 − was the dominant form in solution. As the pH increased
obic environment (Eh ∼ −41–−76 mV). The study has confirmed to pH 8, where HAsO4 2− was the dominant anion, arsenate uptake
that anaerobic water management is the main reason for the high decreased. Mukherjee and Kumar (2005) observed in the aquatic
enhanced As uptake in rice. This was also in agreement in this study plant Pistia stratiotes that the maximum As uptake rates occurred
with a remarkably higher As uptake in the plant biomass (mainly at pH 6.5.
roots) under a persistent anaerobic environment, even in the roots
from the control reactor R1 with a concentration of 90 ± 2.83 mg
As kg−1 (dry wt) (Tables 4 and 6; Fig. 4B). 4.3. Accumulation of As in plant biomass
Several studies have shown that many factors affect the bioac-
cumulation (biosorption) of metals and metalloids in aquatic Terrestrial plants are able to accumulate As to a substantial
ecosystems. Among the physico-chemical factors, pH is possibly extent (Visoottiviseth et al., 2002) but survive the stress to dif-
the most important (Gadd, 2009; Lizama et al., 2011) and plays fering degrees of vitality. During the whole experimental period
a vital role for As mass retention within the rhizosphere. Persis- (341 days), As was accumulated in different compartments of the
tent low pH value in all corresponding reactors might referred to plant biomass, mainly into plant shoots and roots. No remarkable
as a rhizosphere acidification under oxidized conditions, which differences can be found in terms of As translocation into the plant
could have resulted from the effect of K+ uptake and release of shoots of the corresponding reactors. In general, the mean As con-
H+ under conditions of low redox buffer capacity (Rahman et al., centrations (dry wt) within the plant shoots of the reactors were
2008b). Arsenic-induced root exudation might also a probable rea- 7-fold higher than the As concentration in the control shoot sam-
son for a reduced pH in the rhizosphere. One study showed that ples (Fig. 4A).
the reduction in rhizosphere pH increases the chemical activity of The concentrations of total As in the plant roots were extremely
most metals, thereby increasing As uptake by the ferns (Tu and Ma, higher than in the shoots in this study (Fig. 4). In general, total As
2004). Nevertheless, As removal is strongly dependent on the pH concentrations exhibited in the roots were 23–88 times higher than
and these low pH in the reactors remarkably favored As removal to their shoots in all reactors. A vast majority of As were found to be
(>90%, predominantly as arsenate) under this C-deficient oxic con- fixed in/on the roots and only a very limited amount was translo-
dition. In general, sulfate-reducing microorganisms do not grow cated to the shoots. Buddhawong (2005) and other authors also
well at pH values below 5.5 and prefer higher levels of alkalin- reported similar facts when dealing with As and heavy metals in
ity, with 6.6 being optimal (Govind et al., 1999). Therefore, rapid constructed wetlands. Roots continuously remained under direct
As removal processes took place within the rhizosphere by other exposure to As in both oxic and anoxic environment and transloca-
processes than dissimilatory SO4 2− reduction. tion rate of As into shoots presumably depends on other factors and
Addition of organic C-sources caused simultaneously a reducing varies within different plant species. Carbonell et al. (1998) studied
condition inside the reactors and an increment of rhizosphere pH the As content in Spartina alterniflora and found As in the range of
level due to buffering effect (phase II; Table 4). Rapid SO4 2− reduc- 0.80–1.77 mg kg−1 in shoots and 6.87–86.60 mg kg−1 in the roots.
tion by dissimilatory sulfate-reducers utilizing already enriched The much higher accumulation of As in the plant roots compared
S-pool might produce alkalinity and triggered the increment of to the above-ground biomass (shoots) corresponded to the studies
rhizosphere pH under such conditions. Better As mass retention for Typha latifolia, Equisetum fluviatile, Triglochin palustre, and Spar-
in these experimental phases were observed in a highly consistent ganium sp. (Dushenko et al., 1995). However, Rahman et al. (2011)
pH with an average pH value of 7.0 ± 0.5 and 5.7 ± 0.5 in R2 and showed a translocation with a range of 12–20 times from the roots
R3, respectively (phase III–V, Tables 2 and 4). Changes in alkalinity to the shoots when investigating the fate of As in a laboratory-
can indicate changes in the speciation of As and sulfate-reducing scale horizontal subsurface-flow constructed wetland. The low As
bacteria activity, since these bacteria can also provide alkalinity to translocation from roots to shoots is probably due to the fact that
the water and affect its pH (Cohen, 2006). Lizama et al. (2011) in arsenate is rapidly reduced to arsenite in the roots, followed by
a review paper stated that the changes in the speciation of As can complexation with thiols and sequestration in the root vacuoles
affect pH, as the oxidation of As(III) to As(V) decreases the pH value, (Zhao et al., 2009). Iron plaque on root surfaces has been shown
whereas the precipitation of arseno-sulfides increases it. This was to control the uptake and transfer of As by rice (Liu et al., 2005).
also demonstrated within the rhizosphere of Juncus effusus in this Several other studies have shown that roots accumulate more As
study. than do shoots (Adhikari et al., 2011; Hozhina et al., 2001). Dif-
From a very high to a relatively lower or even limited SO4 2− ferent reasons may explain why As remains mostly in plant roots,
loading might also be resulting in a decreasing tendency of rhizo- such as limited translocation of As from roots to shoots (Wang et al.,
sphere pH, even though under C-surplus conditions (phase III–V in 2002), the presence of Fe and S (Zhao et al., 2010), the effect of As
R3; Tables 1 and 4). In fact, it was very interesting to observe a com- speciation in the mechanism of translocation and its relationship
paratively lower mean pH in all thorough the experimental phases to the phosphate transporter (Dhankher, 2005), and the formation
(I–V) of R3, specifically a decrease of nearly 2–3 pH units during of As(III)-phytochelation (PC) complexes in roots and subsequent
the phases III–V as compared to the reactor R2, This lower pH in sequestration in root vacuoles (Xue and Yan, 2011). However, the
the rhizosphere might be a convincing reason for higher amount of form of As that is translocated to shoots or how this transloca-
re-mobilized As in the pore water (16.2%) as well as higher As con- tion occurs is not known (Dhankher, 2005; Lizama et al., 2011).
centration in the roots collected from the reactor R3 than the roots Further studies on these As species are therefore needed to investi-
from R1 and R2 (Figs. 4 and 5). So, an increase of rhizosphere pH gate the mechanism of As uptake, translocation and accumulation,
102 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

considering both the water and the sediment and taking into the higher As concentration within the roots. The ability to carry
account the relationship with Fe, S and phosphate. oxygen from the air down to its stem and discharge it in the rhizo-
Our results with a mean concentration of 90 ± 2.8, 110.5 ± 3.5 sphere through the roots (Brammer and Ravenscroft, 2009) creates
and 315 ± 4.3 mg As kg−1 within the roots of Juncus effusus col- oxidized micro-environments around the roots in which iron is
lected from the three corresponding reactors were definitely not oxidized and precipitated to form a coating (Liu et al., 2006). This
surprising under an ideal-flow condition, also depending on var- oxidation in the rhizosphere may also lead to As precipitating on
ious loading conditions and exposure time. Nearly 3-fold higher root surfaces. The presence of available sulfur due to a high SO4 2−
As concentrations as compared to the roots from the reactor R2 loading might also be a reason for the formation of iron plaque on
and 3.5-fold higher from the control reactor R1 clearly indicated the root surfaces of Juncus effusus in this study. Nevertheless, more
a higher amount of As retention, accumulation, adsorbed, metab- investigations on the effects of O2 release through the plant-roots,
olized to other forms and/or translocated into the root vicinity of changing of redox and pH for the formation of Fe plaque on root sur-
Juncus effusus in the reactor R3, where organic carbon and SO4 2− faces and hence altering the bioavailability of As in the rhizosphere
loading were abundant. Nearly 5-fold lower C/S ratio (TOC: SO4 2− - of constructed wetlands are needed.
S) within the reactor R3 might facilitate lower pH and consequent
remobilization caused free or exchangeable As in the pore water 4.4. Distribution of As in sediment
and hence a higher concentration of As in roots as compared to
the reactor R2 was found. Higher enrichment of total sulfur due A big variety of suspended and biofilm-fixed microorganisms on
to the re-oxidation of reduced S2– to oxidized S-species like ele- plant-root surfaces along with surplus organic C produced sludge
mental sulfur (S0 ), SO4 2− etc. might also cause remobilization of As sediments, which were collected from the reactors after the ter-
(Rahman et al., 2008b) and hence a higher As-concentration in the mination of the experiments. Analytical results of the collected
roots of the reactor R3 was established. But interestingly, the dry and dried sediments showed a wide variation with a mean con-
weight of the total root biomass in reactor R3 was comparatively centration of 206 ± 5, 806 ± 9 and 941 ± 4 mg As kg−1 (dry wt) in
lower than the dry weight of the roots collected from other two R1, R2 and R3, respectively (Fig. 4C). Clear evidence of intensive
reactors R1 and R2 (Table 5). Therefore, a higher As accumulation dissimilatory SO4 2− reduction, highly efficient TOC removal, high
or uptake by the plant roots might have some effect on plant root As adsorption and/or precipitation in the sludge sediment of the
growth under such long-term investigation. reactor R3 suggested higher microbial activities within the rhizo-
The results of the study by Favas et al. (2012) revealed high sphere in presence of higher SO4 2− loading as compared to other
bioaccumulation levels of As in several species at a magnitude two reactors R1 and R2.
much higher than the concentration in the surrounding water. The
highest concentrations of As were found in the submerged species 4.5. As mass balance
Callitriche lusitanica (2346 mg As kg−1 dry wt), but the measured
concentrations in the other emergent plants were significantly Calculation of total As mass balance was carried out for all the
lower, even in the rhizomes/roots, such as T. latifolia (4.17 mg experimental reactors at the end of our investigations. Since the
As kg−1 dry wt) and J. effusus (14.4 mg As kg−1 dry wt). The sorp- reactors were closed tightly enough and well-controlled systems,
tion and bio-concentration of As were measured in Schoenoplectus we were able to obtain a quantitative mass balance of total As by
californicus and Typha angustifolia in a pilot-scale constructed wet- expressing its distribution as percentage of the total mass loaded
land receiving wastewater inflows containing As at potentially into the reactors. It was observed that the reactors R2 and R3 were
hazardous levels (Sundberg-Jones and Hassan, 2007). Moreover, performing with a much better As deposition than the control reac-
it has already been shown that roots and shoots As concentrations tor R1 with limited SO4 2− loading.
significantly increased with increasing As application rates to the A substantially higher amount of As mass was accumulated in
rooting medium of a wetland ecosystem (Mkandawire and Dudel, the roots than sludge sediments in reactor R3 with a higher SO4 2−
2005; Sundberg-Jones and Hassan, 2007). The root vicinity of Juncus loading. Based on the concentration of As in the plant biomass
effusus was continuously exposed to an artificial wastewater con- (shoots and roots) and sludge sediments of the reactor R3, we calcu-
taminated with As under idealized flow conditions, which might lated a 42% of total inflow As mass were accumulated in the roots
enhance bioaccumulation of As and the uptake by the roots in this of Juncus effusus and 17.2% were deposited within the sediment
investigation. (Fig. 5). A high level of As in the roots of R3 might be related to a
The increase of As uptake attributed to iron plaque on root higher SO4 2− loading (i.e. lower C/S ratio), consequently a lower pH
surfaces has been reported in rice and Spirodela polyrhiza (Liu and formation of elemental sulfur (S0 ) within this reactor, hence a
et al., 2005; Rahman et al., 2008c). At the root plaque interface, greater remobilization of As in the solution for plant-root uptake.
siderophores or phytosiderophores exuded by microbes or roots This was demonstrated by a comparatively higher As retention in
may complex with Fe(III) and mobilize Fe-bound arsenate, taken the pore water (16.2%) and lower retention (9%) in unaccountable
up through phosphate co-transporters, which may lead to simulta- sink than the reactors R1 and R2 (Table 4; Fig. 5).
neous uptake of Fe and arsenate (Liu et al., 2005). This phenomenon The accumulation or deposition of As in the sediments occurred
has also been observed for other metals such as Cu and Zn (Ye et al., probably due to the formation of insoluble precipitates like As2 S3
2001). Fitz and Wenzel (2002) reported that root-induced changes via abiotic processes or, more probably, driven by the high level of
in the rhizosphere altered its chemical composition and facilitated microbial activity and biotic processes incorporated to the organic
As uptake by Pteris vittata L (Chinese brake fern), the first-known matter content within the rhizosphere. Along with sludge sedi-
arsenic hyperaccumulator. Root architecture and physiology, root- ments, plant roots were also therefore considered to be the primary
induced changes in water and nutrient availability, root exudates, sink for the sequestration of As in the rhizosphere, which agrees
and fungal and bacterial associations (Gahoonia and Nielsen, 2003) with an As accumulation range of 44–49% within the roots of
are all components of the dynamic rhizosphere system and prob- Juncus effusus investigated in the planted horizontal subsurface-
ably facilitated higher As retention within the roots in this study. flow constructed wetlands (Rahman et al., 2011) and partially
The formation of iron plaque on the macrophyte roots has a high agrees with many other studies of treatment wetland systems
affinity for As, tending to have a higher affinity for As(V) than (Sundaravadivel and Vigneswaran, 2001). Several studies have
As(III) (Chen et al., 2005), and might also be a probable reason for identified aquatic plants with high As content: Lagarosiphon major
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 103

(300 mg As kg−1 dry wt) (Brooks and Robinson, 1998), Egeria densa 5. Conclusions
(>1000 mg As kg−1 dry wt) (Robinson et al., 1995), C. demersum
(>1000 mg As kg−1 dry wt) (Robinson et al., 1995), and Lemna This study reveals that wetland plants possess high As con-
gibba (1021.7 ± 250.8 mg As kg−1 dry wt) (Mkandawire et al., 2004). centrations on roots and a very low As translocation from roots
Therefore, the accumulation of As depends on the type of plant to the shoots. Redox potential and pH within the root-zone are
(Lizama et al., 2011; Zhao et al., 2010), and the potential of some greatly influencing As deposition and plant uptake. An increase
aquatic plants to accumulate As has already been well demon- of rhizosphere pH (alkalinity) might be favoring the immobiliza-
strated. This potential of aquatic plants to accumulate As supports tion of liable and exchangeable As within the root vicinity and
their possible use in phytoremediation of As-contaminated water a decrease of pH might cause remobilization of As and enhance
(Xue and Yan, 2011). Li et al. (2011) suggested that most of As plant (root) uptake. Sulfate loading with an elevated SO4 2− con-
uptake was accumulated in 9 wetland species root tissues rather centration is also playing an important role, which facilitates an
than on root surfaces or in shoot tissues. The root tissue is there- efficient SO4 2− reduction and TOC removal as well as a higher
fore the main barrier for As transport in wetland plants. These As retention under microbial dissimilatory SO4 2− reducing con-
might be critical considerations for improving the efficiency of dition within the rhizosphere. Consequently, high enrichment of
As removal from wastewater in full-scale constructed wetland total sulfur due to the re-oxidation might cause remobilization
systems. of As and enhance plant-root mediated As uptake. Nevertheless,
Only <1% of total inflow As mass were accumulated into the the higher the SO4 2− loading, the better will be the As binding
plant shoots (Fig. 5). This result clearly indicated the minor role capacity within the root-near environment of the rhizosphere in
by the plant shoots in terms of As uptake and showed a very little constructed wetlands. Plant shoots (Juncus effusus) are playing a
contribution to the overall As mass balance. Standing or pore water minor role in terms of As uptake and contributing very little to the
accounted for nearly 5%, 10% and 16% of total inflow As mass in the overall As mass balance. Roots and sludge sediments are consid-
three respective reactors with an increasing SO4 2− concentrations. ered to be the primary sinks for As retention in the rhizosphere
Enrichment of total sulfur due to the re-oxidation of reduced S2– to of constructed wetlands. Accumulation of a small portion of As to
oxidized S-species (S0 , SO4 2− etc.) might cause remobilization of unaccountable sinks and/or a probable loss due to volatilization
As and hence a higher percentile As mass can be found in the pore or other unknown reasons demands further research under such
water of the reactor R3 with a higher SO4 2− loading than the other strict conditions on a priority basis in order to protect the surround-
two reactors R1 and R2. ing environment from any toxic effects of volatile As compounds.
Nearly a total of 52%, 21% and 15% of the inflow As mass were Moreover, different physico-chemical properties on the root sur-
passed through the reactors R1, R2, and R3, respectively and col- face, storage characteristics in root tissues of wetland plants and
lected at the outlet. The larger amounts of As (52% of the total precipitation/adsorption/retention of different As species clearly
inflow As mass) were flushed out through the outlet of the control requires further investigation. Future studies should also intend
reactor R1 in comparison to the reactor R3 (15%), which indicated to focus on exploring the patterns of microbial ecology and their
a highly stable bonding of As with the attached biofilms within interactions, associated As toxicity on plants and microbial biomass
plant root zone of the rhizosphere associated with higher SO4 2− and volatilization of As under dynamic redox conditions within the
loading. Therefore, the higher the SO4 2− loading as well as an rhizosphere of constructed wetland system.
enhanced dissimilatory SO4 2− reduction in the reactor (Rahman
et al., 2008b), the better will be the As binding capacity within Acknowledgments
the root-near environment of the rhizosphere in constructed
wetlands. The work was funded by a grant from the German Fed-
Only traces (2–3 ␮g l−1 ) of inorganic volatile arsine (AsH3 ) were eral Ministry of Education and Research under the International
measured in this study, which resulted in a very small amount Postgraduate Study in Water Technology (BMBF-IPSWaT) pro-
of As mass that was left out of the systems due to volatilization. gram and by grants from the Helmholtz-Centre for Environmental
Nearly 26%, 29% and 9% of total As mass in these three reactors were Research–UFZ, Leipzig, Germany. The authors would like to thank
accumulated into unaccountable sink, which could be due to other Kerstin Puschendorf, Ines Mäusezahl, Karsten Marien, Jürgen
microbial reactions, adsorption/precipitation to other unknown Steffen, Reinhard Schumann, and Uwe Kappelmeyer for their out-
sink or even lost due to volatilization. Some of the retained or standing technical and analytical support.
trapped As might have been bio-transformed by plant-root activity
and associated microbes to other unidentified organic compounds,
which prevented the calculation of a complete As mass balance in References
this study.
Adhikari, A.R., Acharya, K., Shanahan, S.A., Zhou, X., 2011. Removal of nutrients and
From the previous results (Buddhawong, 2005), it was found metals by constructed and naturally created wetlands in the Las Vegas Valley,
that the gravel material could not absorb As mass in substantial Nevada. Environ. Monit. Assess. 180, 97–113.
amounts from the solution. Therefore, this kind of gravel itself had Bednar, A.J., Garbarino, J.R., Ranville, J.F., Wildeman, T.R., 2005. Effects of iron on
arsenic speciation and redox chemistry in acid mine water. J. Geochem. Explor.
no remarkable impact on the mass balance of As in constructed 85, 55–62.
wetland systems. Fractional analysis of deposited sediment might Bezbaruah, A.N., Zhang, T.C., 2004. pH, redox, and oxygen microprofiles in rhizo-
also be necessary to investigate different forms of As that were sphere of bulrush (Scirpus validus) in a constructed wetland treating municipal
wastewater. Biotechnol. Bioeng. 88 (1), 60–70.
retaining within the rhizosphere of helophytes in treatment wet- Bostick, B.C., Fendorf, S., 2003. Arsenite sorption on troilite (FeS) and pyrite (FeS2 ).
lands. Singhakant et al. (2009) reported about the forms of the Geochim. Cosmochim. Acta 67, 909–921.
retained As by using sequential fractionation, which could indi- Brammer, H., Ravenscroft, P., 2009. Arsenic in groundwater: a threat to sustainable
agriculture in South and South-east Asia. Environ. Int. 35, 647–654.
cate As complexation with Fe and Mn on the media surface of Brooks, R.R., Robinson, B.H., 1998. Aquatic phytoremediation by accumulator plants.
31–38% and As trapping into the media of 42–52% of the total In: Brooks, R.R. (Ed.), Plants that Hyperaccumulate Heavy Metals: Their Role in
As. However, more attention should be given to the accumulation Archaeology, Microbiology, Mineral Exploration, Phytomining and Phytoreme-
diation. CAB Int., Wallingford, pp. 203–226.
of As in different wetland compartments and to avoid poten-
Buddhawong, S., 2005. Constructed Wetlands and their Performance for Treatment
tial toxic effects the accumulated As could pose to the wetland of Water Contaminated with Arsenic and Heavy Metals. University of Leipzig,
plants. Germany, Ph.D. Thesis.
104 K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105

Buddhawong, S., Kuschk, P., Mattusch, J., Wießner, A., Stottmeister, U., 2005. Liu, W.J., Zhu, Y.G., Smith, F.A., 2005. Effects of iron and manganese plaques on
Removal of arsenic and zinc using different model wetlands systems. Eng. Life arsenic uptake by rice seedlings (Oryza sativa L.) grown in solution culture
Sci. 5 (3), 247–252. supplied with arsenate and arsenite. Plant Soil 277, 127–138.
Cabrera, G., Pérez, R., Gómez, J.M., Ábalos, A., Cantero, D., 2006. Toxic effects of Lizama, K.A., Fletcher, T.D., Sun, G., 2011. Removal processes for arsenic in con-
dissolved heavy metals on Desulfovibrio vulgaris and Desulfovibrio sp. strains. structed wetlands. Chemosphere 84, 1032–1043.
J. Hazard. Mater. 135, 40–46. Mattes, A.G., Evans, L., Glasauer, S., 2010. Biologically based treatment system for
Carbonell, A.A., Aarabi, M.A., DeLaune, R.D., Gambrell, R.P., Patrick Jr., W.H., 1998. removal of high concentrations of heavy metal and arsenic contaminants from
Arsenic in wetland vegetation: availability, phytotoxicity, uptake and effects on landfill leachate–a look at a system after 10 years of full time operation. In:
plant growth and nutrition. Sci. Total Environ. 217, 189–199. 12th International Conference on Wetland Systems for Water Pollution Control,
Chen, Z., Zhu, Y.G., Liu, W.J., Meharg, A.A., 2005. Direct evidence showing the effect of Venice, Italy.
root surface iron plaque on arsenite and arsenate uptake into rice (Oryza sativa) Meharg, A.A., 2004. Arsenic in rice—understanding a new disaster for South-East
roots. New Phytol. 165, 91–97. Asia. Trends Plant Sci. 9, 415–417.
Cohen, R.R.H., 2006. Use of microbes for cost reduction of metal removal Mkandawire, M., Dudel, E.G., 2005. Accumulation of arsenic in Lemna gibba L. (duck-
from metals and mining industry waste streams. J. Clean. Prod. 14, weed) in tailing waters of two abandoned uranium mining sites in Saxony,
1146–1157. Germany. Sci. Total Environ. 305, 81–89.
Darrah, P.R., Jones, D.L., Kirk, G.J.D., Roose, T., 2006. Modelling the rhizosphere: a Mkandawire, M., Taubert, B., Dudel, E.G., 2004. Capacity of Lemna gibba L. (duck-
review of methods for ‘upscaling’ to the whole-plant scale. Eur. J. Soil Sci. 57, weed) for uranium and arsenic phytoremediation in mine tailing waters. Int. J.
13–25. Phytorem. 6, 347–362.
Daus, B., Mattusch, J., Wennrich, R., Weiß, H., 2002. Investigation on stability Mukherjee, S., Kumar, S., 2005. Arsenic uptake potential of water lettuce (Pistia
and preservation of arsenic species in iron rich water samples. Talanta 58, Stratiotes L.). Int. J. Environ. Stud. 62, 249–258.
57–65. Pantsar-Kallio, M., Korpela, A., 2000. Analysis of gaseous arsenic species and stability
Dhankher, O.P., 2005. Arsenic metabolism in plants: an inside story. New Phytol. studies of arsine and trimethylarsine by gas-chromatography–mass spectrom-
168, 503–505. etry. Anal. Chim. Acta 410, 6570.
Duncan, W.F.A., Mattes, A.G., Gould, W.D., Goodazi, F., 2004. Multi-stage bio- Rahman, M.A., Hasegawa, H., Ueda, K., Maki, T., Rahman, M.M., 2008a. Arsenic uptake
logical treatment system for removal of heavy metal contaminants. In: by aquatic macrophyte Spirodela polyrhiza L.: interactions with phosphate and
Rao, S.R., Harrison, F.W., Kosinski, J.A., Amaratunga, L.M., Cheng, T.C., iron. J. Hazard. Mater. 160, 356–361.
Richards, G.G. (Eds.), The Fifth International Symposium on Waste Processing Rahman, K.Z., Wiessner, A., Kuschk, P., Mattusch, J., Kästner, M., Müller, R.A., 2008b.
and Recycling in Mineral and Metallurgical Industries. Canadian Insti- Dynamics of arsenic species in laboratory-scale horizontal subsurface-flow con-
tute of Mining, Metallurgy and Petroleum, Hamilton, Ontario, Canada, structed wetlands treating an artificial wastewater. Eng. Life Sci. 8, 603–611.
pp. 469–483. Rahman, K.Z., Wiessner, A., Kuschk, P., Mattusch, J., Offelder, A., Kästner, M., Müller,
Dushenko, W.T., Bright, D.A., Reimer, K.J., 1995. Arsenic bioaccumulation and tox- R.A., 2008c. Redox dynamics of arsenic species in the root-near environment of
icity in aquatic macrophytes exposed to gold-mine effluent: relationships with Juncus effusus investigated in a macro-gradient-free rooted gravel bed reactor.
environmental partitioning. Aquat. Bot. 50, 141–158. Eng. Life Sci. 8 (6), 612–621.
Favas, P.J.C., Pratas, J., Prasad, M.N.V., 2012. Accumulation of arsenic by aquatic Rahman, K.Z., Wiessner, A., Kuschk, P., van Afferden, M., Mattusch, J., Müller, R.A.,
plants in large-scale field conditions: opportunities for phytoremediation and 2011. Fate and distribution of arsenic in laboratory-scale subsurface horizontal-
bioindication. Sci. Total Environ. 433, 390–397. flow constructed wetlands treating an artificial wastewater. Ecol. Eng. 37,
Fitz, W.J., Wenzel, W.W., 2002. Arsenic transformations in the 1214–1224.
soil–rhizosphere–plant system: fundamentals and potential application Rai, U.N., Sinha, S., Tripathi, R.D., Chandra, P., 1995. Wastewater treatability potential
to phytoremediation. J. Biotechnol. 99, 259–278. of some aquatic macrophytes: removal of heavy metals. Ecol. Eng. 5, 5–12.
Gadd, G.M., 2009. Biosorption: critical review of scientific rationale, environmental Robinson, B.H., Brooks, R.R., Outred, H.A., Kirkman, J.H., 1995. The distribution and
importance and significance for pollution treatment. J. Chem. Technol. Bio- fate of arsenic in the Waikato river system, North Island, New Zealand. Chem.
technol. 84, 13–28. Spec. Bioavail. 7, 89–96.
Gahoonia, T.S., Nielsen, N.E., 2003. Phosphorus (P) uptake and growth of a root hair- Sasmaza, A., Obek, E., 2009. The accumulation of arsenic, uranium, and boron in
less barley mutant (bald root barley, brb) and wild type in low- and high-P soils. Lemna gibba L. exposed to secondary effluents. Ecol. Eng. 35, 1564–1567.
Plant Cell Environ. 26, 1759–1766. Schmidt, A.C., Reisser, W., Mattusch, J., Wennrich, R., Jung, K., 2004. Analysis of
Gambrell, R.P., 1994. Trace and toxic metals in wetlands – a review. J. Environ. Qual. arsenic species accumulation by plants and the influence on their nitrogen
23, 883–891. uptake. J. Anal. At. Spectrom. 19, 172–177.
García, J., Rousseau, D.P.L., Morató, J., Lesage, E., Matamoros, V., Bayona, J.M., 2010. Sharma, V.K., Sohn, M., 2009. Aquatic arsenic: toxicity, speciation, transformations,
Contaminant removal processes in subsurface-flow constructed wetlands: a and remediation. Environ. Int. 35, 743–759.
review. Crit. Rev. Environ. Sci. Technol. 40, 561–661. Singhakant, C., Koottatep, T., Satayavivad, J., 2009. Fractional analysis of arsenic
Govind, R., Yong, W., Tabak, H., 1999. Studies on bio-recovery of metals from acid in subsurface-flow constructed wetlands with different length to depth ratios.
mine drainage. In: Leeson, A., Allemans, B. (Eds.), Bioremediation of Metals and Water Sci. Technol. 60, 1771–1778.
Inorganic Compounds. Battelle Press, Columbus, OH, p. 37. Singhakant, C., Koottatep, T., Satayavivad, J., 2009a. Enhanced arsenic removals
Hozhina, E.I., Khramov, A.A., Gerasimov, P.A., Kumarkov, A.A., 2001. Uptake of heavy through plant interactions in subsurface-flow constructed wetlands. J. Environ.
metals, arsenic, and antimony by aquatic plants in the vicinity of ore mining and Sci. Health A 44, 163–169.
processing industries. J. Geochem. Explor. 74, 153–162. Smedley, P.L., Kinniburgh, D.G., 2002. A review of the source, behavior and distribu-
Islam, F.S., Gault, A.G., Boothman, C., Polya, D.A., Chatterjee, D., Lloyd, J.R., 2004. tion of arsenic in natural waters. Appl. Geochem. 17, 517–568.
Direct evidence of arsenic release from Bengali sediments due to metal-reducing Sun, G., Saeed, T., 2009. Kinetic modelling of organic matter removal in hori-
bacteria. Nature 430, 68–71. zontal flow reed beds for domestic sewage treatment. Process Biochem. 44,
Jacks, G., Bhattacharya, P., Routh, J., Martin, M.T., 2002. Arsenic cycling in a cov- 717–722.
ered mine tailings deposit. In: Schulz, H.D., Hadeler, A. (Eds.), Geo. Proc. Wiley Sundaravadivel, M., Vigneswaran, S., 2001. Constructed wetland for wastewater
Publication, Northern Sweden, pp. 303–309. treatment. Environ. Sci. Technol. 31 (4), 351–409.
Kadlec, R.H., Knight, R.L., 1996. Treatment Wetlands. CRC Press, Boca Raton, pp. 893. Sundberg-Jones, S.E., Hassan, S.M., 2007. Macrophyte sorption and bioconcentra-
Kappelmeyer, U., Wießner, A., Kuschk, P., Kästner, M., 2002. Operation of a univer- tion of elements in a pilot constructed wetland for flue gas desulfurization
sal test unit for planted soil filters–planted fixed bed reactor. Eng. Life Sci. 2, wastewater treatment. Water Air Soil Pollut. 183 (1-4), 187–200.
311–315. Talukder, A.S.M.H.M., Meisner, C.A., Sarkar, M.A.R., Islam, M.S., Sayre, K.D., Duxbury,
Kirk, M.F., Holm, T.R., Park, J., Jin, Q., Sanford, R.A., Fouke, B.W., Bethke, C.M., 2004. J.M., Lauren, J.G., 2012. Effect of water management, arsenic and phosphorus
Bacterial sulfate reduction limits natural arsenic contamination in groundwater. levels on rice in a high-arsenic soil–water system: II arsenic uptake. Ecotoxicol.
Geology 32, 953–956. Environ. Saf. 80, 145–151.
Kneebone, P.E., O’Day, P.A., Jones, N., Hering, J.G., 2002. Deposition and fate of arsenic Tripathi, R.D., Srivastava, S., Mishra, S., Singh, N., Tuli, R., Gupta, D.K., Maathuis,
in iron- and arsenic-enriched reservoir sediments. Environ. Sci. Technol. 36, F.J.M., 2007. Arsenic hazards: strategies for tolerance and remediation by plants.
381–386. Trends Biotechnol. 25, 158–165.
Larios, R., Martínez, R.F., LeHecho, I., Rucandio, I., 2012. A methodological Tu, S., Ma, L.Q., 2004. Comparison of arsenic uptake and distribution in arsenic hyper-
approach to evaluate arsenic speciation and bioaccumulation in different accumulator Pteris vittata L. and non-hyperaccumulator Nephrolepis exaltata L.
plant species from two highly polluted mining areas. Sci. Total Environ. 414, J. Plant Nutr. 27, 1227–1242.
600–607. Tuttle, J.H., 1969. Microbial sulfate reduction and its potential utility as an acid mine
Li, H., Ye, Z.H., Wei, Z.J., Wong, M.H., 2011. Root porosity and radial oxygen loss water pollution abatement procedure. Appl. Microbiol. 17, 297–302.
related to arsenic tolerance and uptake in wetland plants. Environ. Pollut. 159, Visoottiviseth, P., Francesconi, K., Sridokchan, W., 2002. The potential of Thai indige-
30–37. nous plant species for the phytoremediation of arsenic contaminated land.
Liesack, W., Schnell, S., Revsbech, N.P., 2000. Microbiology of flooded rice paddies. Environ. Pollut. 118, 453–461.
FEMS Microbiol. Rev. 24, 625–645. Wang, J., Zhao, F.J., Meharg, A.A., Raab, A., Feldmann, J., McGrath, S.P., 2002.
Liu, W.J., Zhu, Y.G., Hu, Y., Williams, P.N., Gault, A.G., Meharg, A.A., 2006. Arsenic Mechanisms of arsenic hyperaccumulation in Pteris vittata. Uptake kinet-
sequestration in iron plaque, its accumulation and speciation in mature rice ics, interactions with phosphate, and arsenic speciation. Plant Physiol. 130,
plants (Oryza sativa L.). Environ. Sci. Technol. 40, 5730–5736. 1552–1561.
K.Z. Rahman et al. / Ecological Engineering 69 (2014) 93–105 105

Wells, J.M., Richardson, D.H.S., 1985. Anion accumulation by the moss Hylocomium Xue, P.Y., Yan, C.Z., 2011. Arsenic accumulation and translocation in the
splendens: uptake and competition studies involving arsenate, selenate, selenite, submerged macrophyte Hydrilla verticillata (L.f.) Royle. Chemosphere 85,
phosphate, sulphate and sulphite. New Phytol. 101, 571–583. 1176–1181.
Wiessner, A., Gonzalias, A.E., Kästner, M., Kuschk, P., 2008. Effects of sulphur cycle Yang, J.X., Ma, Z.L., Ye, Z.H., Guo, X.Y., Qiu, R.L., 2010. Heavy metal (Pb, Zn) uptake
processes on ammonia removal in a laboratory-scale constructed wetland and chemical changes in rhizosphere soils of four wetland plants with different
planted with Juncus effusus. Ecol. Eng. 34, 162–167. radial oxygen loss. J. Environ. Sci. (China) 22, 696–702.
Wiessner, A., Kappelmeyer, U., Kaestner, M., Schultze-Nobre, L., Kuschk, P., 2013. Ye, Z.H., Lin, Z.Q., Whiting, S.N., de Souza, M.P., Terry, N., 2003. Possible use of con-
Response of ammonium removal to growth and transpiration of Juncus effusus structed wetland to remove selenocyanate, arsenic, and boron from electric
during the treatment of artificial sewage in laboratory-scale wetlands. Water utility wastewater. Chemosphere 52, 1571–1579.
Res. 47, 4265–4273. Ye, Z.H., Whiting, S., Qian, J.H., Lytle, C.M., Lin, Z.Q., Terry, N., 2001. Trace element
Wiessner, A., Kappelmeyer, U., Kuschk, P., Kästner, M., 2005a. Sulphate reduction and removal from coal pile leachate by an Alabama 10-year old constructed wetland.
the removal of carbon and ammonia in a laboratory-scale constructed wetland. J. Environ. Qual. 30, 1710–1719.
Water Res. 39 (19), 4643–4650. Zhao, F.J., Ma, J.F., Meharg, A.A., McGrath, S.P., 2009. Arsenic uptake and metabolism
Wiessner, A., Kappelmeyer, U., Kuschk, P., Kästner, M., 2005b. Influence of the redox in plants. New Phytol. 181, 777–794.
condition dynamics on the removal efficiency of a laboratory-scale constructed Zhao, F.J., McGrath, S.P., Meharg, A.A., 2010. Arsenic as a food chain contaminant:
wetland. Water Res. 39 (1), 248–256. mechanisms of plant uptake and metabolism and mitigation strategies. Annu.
Willow, M.A., Cohen, R.R.H., 2003. PH, dissolved oxygen, and adsorption effects on Rev. Plant Biol. 61, 535–559.
metal removal in anaerobic bioreactors. J. Environ. Qual. 32, 1212–1221. Zouboulis, A.I., Katsoyiannis, I.A., 2002. Removal of arsenates from contaminated
Wright, D.J., Otte, M.L., 1999. Wetland plant effects on the biogeochemistry of metals water by coagulation-direct filtration. Sep. Sci. Technol. 37 (12), 2859–2873.
beyond the rhizosphere. Biol. Environ. Proc. R. Ir. Acad. 99B, 3–10.

S-ar putea să vă placă și