You are on page 1of 7

Removal of Chromium (VI) from wastewater using

bentonite-supported nanoscale zero-valent iron


Li-na Shi
a
, Xin Zhang
b
, Zu-liang Chen
a,
*
a
School of Chemistry and Material Sciences, Fujian Normal University, Fuzhou 350007, Fujian Province, China
b
School of Medicine, Shanxi University of Chinese Medicine, Xianyang 712000, Shanxi Province, China
a r t i c l e i n f o
Article history:
Received 8 February 2010
Received in revised form
14 September 2010
Accepted 18 September 2010
Available online 1 October 2010
Keywords:
Bentonite
Nanoscale zero-valent iron
Cr(VI)
Wastewater
a b s t r a c t
Bentonite-supported nanoscale zero-valent iron (B-nZVI) was synthesized using liquid-
phase reduction. The orthogonal method was used to evaluate the factors impacting Cr(VI)
removal and this showed that the initial concentration of Cr(VI), pH, temperature, and
B-nZVI loading were all importance factors. Characterization with scanning electron
microscopy (SEM) validated the hypothesis that the presence of bentonite led to a decrease
inaggregationof ironnanoparticles anda corresponding increase inthe specic surface area
(SSA) of the iron particles. B-nZVI with a 50% bentonite mass fraction had a SSA of 39.94 m
2
/
g, while the SSA of nZVI and bentonite was 54.04 and 6.03 m
2
/g, respectively. X-ray
diffraction (XRD) conrmed the existence of Fe
0
before the reaction and the presence of Fe
(II), Fe(III) and Cr(III) after the reaction. Batch experiments revealed that the removal of Cr
(VI) using B-nZVI was consistent with pseudo rst-order reaction kinetics. Finally, B-nZVI
was used to remediate electroplating wastewater with removal efciencies for Cr, Pb and Cu
> 90%. Reuse of B-nZVI after washing with ethylenediaminetetraacetic acid (EDTA) solution
was possible but the capacity of B-nZVI for Cr(VI) removal decreased by approximately 70%.
2010 Elsevier Ltd. All rights reserved.
1. Introduction
Chromium (VI) is an industrial contaminant in both soil and
groundwater andis alsoawell-knownhumancarcinogen(Katz
and Salem, 1994). Due to its toxicity, Cr(VI) must be removed
from wastewaters prior to discharge into aquatic environ-
ments (Ju-Nam and Lead, 2008). Conventional remediation
techniques typically involve reduction of Cr(VI) to Cr(III) which
precipitates as chromium hydroxide or chromium iron
hydroxide at high pH, followed by disposal of the resulting
dewatered sludge (Ngomsik et al., 2005). Other treatments,
including phytoextraction, reverse osmosis, electrodialysis,
ion exchange, membrane ltration and adsorption, have
also been developed to remove metals from industrial waste-
waters. While these methods are useful in removing Cr(VI)
fromaqueous solution, theyhavesomelimitationsandit isstill
necessary to develop new and effective remediation tech-
niques (Mohan and Pittman, 2006). In recent years, nanoscale
zero-valent iron (nZVI) has been used to remove various
groundwater contaminants. Theadvantages of nZVI over zero-
valent iron (ZVI) include higher reactive surface area, faster
and more complete reactions, and better injectability into
aquifers (Li et al., 2006). However, there are still some technical
challenges associated with practical applications, such as the
aggregation of nZVI particles and limitations imposed by high
reactivity and low stability (Liu et al., 2007). Furthermore, the
agglomeration of iron particles is often unavoidable due to the
extremely high-pressure drops occurring in conventional
systems, which along with its lack of durability and mechan-
ical strengthlimits theapplicationof nZVI (Cumbal et al., 2003).
* Corresponding author. Tel./fax: 86 591 83465689.
E-mail address: zlchen@fjnu.edu.cn (Z.-l. Chen).
Avai l abl e at www. sci encedi r ect . com
j our nal homepage: www. el sevi er . com/ l ocat e/ wat r es
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2
0043-1354/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2010.09.025
Recently, technologies have been developed using porous
materials as mechanical supports to enhance the dis-
persibility of nZVI particles (U

zu m et al., 2009). Bentonite is


a traditional low-cost efcient adsorbent, which has high
potential for heavy metal removal from wastewaters due to
its abundance, chemical and mechanical stability, high
adsorption capability and unique structural properties
(Bhattacharyya and Gupta, 2008). Removal of metal ions using
bentonite is based on ion exchange and adsorption mecha-
nisms due to the materials relatively high cation exchange
capacity (CEC) and specic surface area (Bhattacharyya and
Gupta, 2008). In this paper, bentonite was used as a porous-
based support material for synthesized nZVI. More recently,
nZVI supported by zeolite (Li et al., 2007) and stabilized by
chitosan (Geng et al., 2009) has been reported to increase the
durability and mechanical strength of nZVI. However, only
a few studies have reported using natural clays as support
materials for nZVI (U

zu m et al., 2009).
In this paper the removal of Cr(VI) from an aqueous solu-
tion was investigated using B-nZVI and the objectives were: (1)
synthesis of bentonite-supported nanoscale zero-valent
(B-nZVI) by reduction of Fe
3
ions with NaBH
4
, and charac-
terization of the produced material with SEM, XRDand BET-N
2
technology; (2) evaluation of the factors impacting on Cr(VI)
removal using an orthogonal method; nZVI and bentonite
were used for Cr(VI) removal individually as a control, and the
kinetics of Cr(VI) reductionby B-nZVI were also evaluated; and
(3) remediation of electroplating wastewater including some
heavy metal ions using B-nZVI and evaluation of reuse.
2. Materials and methods
2.1. Materials and chemicals
Bentonite was provided by Fenghong Co. Ltd, Anji, Zhejiang,
China, primarily as Na-Mt montmorillonite (>90%), the
chemical composition was 62.5% SiO
2
, 18.5% Al
2
O
3
, 1.75%
Fe
2
O
3
, 4.25% MgO, 0.95% CaO, and 2.75% Na
2
O. The cation
exchange capacity (CEC) was 75.4 meq/100g. After drying
overnight at 80

C, the raw bentonite was ground and sieved
through a 200 mesh screen prior to use in experiments.
All the reagents were analytical grade (Shanghai Nanxiang
Reagent Co., Ltd., China) and distilled water was used in all
preparations. A stock solution containing potassium dichro-
mate (K
2
Cr
2
O
7
) was prepared by dissolving K
2
Cr
2
O
7
with
deionized water and a series of solutions used during the
experiment were prepared by diluting the stock to the desired
concentrations.
2.2. Synthesis of nZVI and supported nZVI
The nZVI and B-nZVI were prepared using conventional
liquid-phase methods via the reduction of ferric iron by
borohydride without or with bentonite as a support material
(Celebi et al., 2007). Bentonite (2.00 g) was initially placed into
a three-necked open ask, and a ferric solution produced by
dissolving ferric chloride hexa-hydrate (9.66 g) in an ethanol-
water solution (50 mL, 4:1 v/v) was added and stirred for
10 min. Subsequently, a freshly prepared NaBH
4
solution
(3.54 g of NaBH
4
in 100 mL) was added drop-wise into the
mixture with constant stirring for 20 min after addition. The
whole process described above was performed under a N
2
atmosphere with vigorous stirring to avoid the oxidization of
B-nZVI. The formed suspension was ltered and the black
nanoscale precipitate was washed three times with pure
ethanol and dried overnight at 75

C under vacuum (Celebi
et al., 2007; U

zu m et al., 2009). The theoretical mass fraction


of bentonite in synthesized B-nZVI was 50%, and nZVI was
prepared under identical conditions but with bentonite
omitted. The nZVI and supported nZVI samples were stored in
brown, sealed bottles under dry conditions and were not
acidied prior to use.
2.3. Characterizations and measurements
Scanning electron microscopy (SEM) was performed using
a Philips-FEI XL30 ESEM-TMP (Philips Electronics Co., Eind-
hoven, The Netherlands). Images of various materials were
obtained at an operating voltage of 30 kV. The SSA of nZVI,
B-nZVI, and bentonite was measured via the BET adsorption
method (U

zu m et al., 2009) using Micromeritics ASAP 2020


Accelerated Surface Area and Porosimetry Analyzer (Micro-
meritics Instrument Corp.,USA). The specic surface areas of
nZVI, B-nZVI and bentonite were 54.04, 39.94 and 6.03 m
2
/g,
respectively. X-ray diffraction (XRD) patterns of B-nZVI before
and after contacting Cr(VI) were performed using a Philips-
XPert Pro MPD (Netherlands) with a high-power Cu- Ka
radioactive source (l 0.154 nm) at 40 kV/40 mA.
The concentration of total Cr in solution and the concen-
trations of different heavy metal ions in the wastewater were
determined using a ame atomic absorbance spectrometer
(VARIAN AA 240FS, USA), and the Cr(VI) concentration was
determined using the 1,5-diphenylcarbazide method (Geng
et al., 2009) on a 722N visible spectrophotometer (Shanghai
Precision & Scientic Instrument Co., Ltd, China).
2.4. Batch experiments
The orthogonal method was used to test the effects of
various factors on the reaction, and to optimize the condi-
tions for Cr(VI) removal using B-nZVI. The experimental
design was developed with the aid of the Orthogonal Design
Assistant, where the initial concentration of Cr(VI), B-nZVI
loading, temperature and pH were chosen as variables. Cr(VI)
solutions (25 mL) with a known mass of B-nZVI were sealed
in 50 mL centrifuge tubes and mixed for 4h before being
centrifuged prior to analysis of the aqueous phase for
residual Cr(VI).
In order to investigate the role that bentonite and Fe
0
played in the B-nZVI system, nZVI, B-nZVI and bentonite were
all used in batch experiments examining Cr(VI) removal from
aqueous solutions at an initial concentration of 50 mg/L at
35

C and 250 r/min. As the mass ratio of Fe
0
:bentonite was 1:1
in the B-nZVI system, the dosages of nZVI and bentonite were
both set at 1.5 g/L, which was half the B-nZVI dosage. The
mixtures were ltered through 0.45 mm mixed cellulose
ester (MCE) membranes prior to determining the residual
concentrations of Cr(VI) after contacting for 3 h. In order to
investigate the effects of the different factors mentioned in
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 887
the orthogonal experiments further, the mixtures of Cr(VI)
solution and B-nZVI were mixed in 50 mL centrifugal tubes in
a rotary shaker for determined periods of time, the conditions
of which were initially set at 25 mL of 50 mg/L Cr(VI) solution,
3 g/L of B-nZVI, 35

C and 250 r/min. At selected timed inter-
vals, the suspension was ltered through 0.45 mm MCE
membranes, and the concentration of Cr(VI) in the ltrate was
determined.
To explore the feasibility of removing heavy metal ions
from wastewater, B-nZVI was used to remediate electro-
plating wastewater collected from an electroplating factorys
sewage outfall (Fuzhou, China). The wastewater was centri-
fuged at 3000 r/min for 10 min to remove any insoluble
impurities, prior to determining the initial pH and concen-
trations of total Cu, Cr, Pb and Zn. A batch of 50 mL bottles
containing wastewater (10 mL) and B-nZVI (0.10 g) were mixed
on a rotary shaker at 35

C and 250 r/min for 4 h. Then the
mixtures were centrifuged at 3000 r/min for 10 min and the
upper aliquot collected to determine pH and the concentra-
tion of each heavy metal ion.
The potential to reuse B-nZVI for removing Cr(VI) from
aqueous solution was also evaluated. B-nZVI (0.1 g) was added
to 50 mg/L Cr(VI) solution (25 mL) and the mixture was shaken
on a rotary shaker (35

C and 250 r/min). After 3 h, centrifu-
gations at 3000 r/min were performed for 10 min to obtain
solid-liquid separation. The supernatant was decanted care-
fully and used to determine the concentration of Cr(VI)
remaining in solution while the used B-nZVI was mixed with
different concentrations of EDTA. The residual B-nZVI-EDTA
solution was washed with distilled water three times and
shaken for another 3 h under identical conditions. The B-nZVI
treated with 50 mg/L and 10 mg/L of EDTAwas used to remove
Cr(VI) for 3 times in succession to test the efcacy of reuse.
In order to ascertain the accuracy, reliability and repro-
ducibility of the data, orthogonal experiments were conducted
in quadruplicate (n 4) and other batch experiments were
carried out in triplicate (n 3) to minimize any experimental
errors. The average values of the parallel measurements were
used in all analysis and together with the standard deviations
of these means were listed in Tables 1 and 3.
3. Results and discussion
3.1. Characterization
The SEM images of nZVI and B-nZVI showed the morphology
and nanoparticle distribution of nZVI in the absence or
presence of bentonite (Fig. 1). The synthesized nZVI without
bentonite as a support material showed that nZVI particles
were roughly globular and aggregated into a chain-like
conformation (Fig. 1a). The diameters of the nanoscale zero-
valent iron particles were in the range of 20e90 nm when
bentonite was introduced as a support material. Compared
with Fig. 1a, the aggregation of nZVI particles seemed to
decrease and their dispersity increase in Fig. 1b, where the
mass fraction of bentonite was 50%. A similar conclusion has
been drawn using kaolin as a support material to synthesize
kaolinsupportednZVI, whichwas usedtoremoveCu(II) andCo
(II) fromanaqueous solution(U

zu met al., 2009). As indicatedin


Fig. 1c, the sizes of iron nanoparticles increase prominently
after reacting with Cr(VI). This phenomenon could be attrib-
uted to the co-precipitation of Cr(III) and Fe(III) on the surface
of the nanoparticles (Ponder et al., 2000; Manning et al., 2007),
which occurs due to a redox reaction between Cr(VI) and Fe
0
(Ponder et al., 2000; Manning et al., 2007).
The XRD patterns of synthesized materials were compared
with the XRD patterns obtained from standard materials, to
identify the apparent peaks attributable to different iron and
chromium compounds. The XRD patterns of B-nZVI before
reaction (Fig. 2a) with Cr(VI) showed an apparent peak of Fe
0
(2q z 44.90), which weakened signicantly after the reaction
Table 1 eOrthogonal experimental designand the results
obtained from the full 2
4
factorial experiment matrix.
T
(

C)
Cr(VI)
ina
(mg/L)
pH
ina
B-nZVI
ina
(mg/L)
Cr(VI)
res
(mg/L)
Removal
efciency (%)
1 25 20 2.0 2 0.11
a
99.5
2 25 50 5.0 3 29 1 42.0
3 25 70 8.0 4 45.0 0.6 35.7
4 25 100 10 5 73 2 26.6
5 30 20 5.0 4 0
a
100
6 30 50 2.0 5 0.06
a
99.9
7 30 70 10.0 2 55.3 0.9 21.0
8 30 100 8.0 3 75 2 25.3
9 35 20 8.0 5 0.04
a
99.8
10 35 50 10.0 4 31.1 0.7 55.6
11 35 70 2.0 3 0.04
a
99.9
12 35 100 5.0 2 71 2 29.4
13 40 20 10.0 3 0.10
a
99.5
14 40 50 8.0 2 6.75
a
86.5
15 40 70 5.0 5 32.0 0.9 54.3
16 40 100 2.0 4 0.24
a
99.8
ina - initial; res - residual; a means the standard deviations are too
low to be listed.
Table 2 e Range analysis and variance analysis of the orthogonal test.
Factors Range Analysis Variance Analysis
k
1
k
2
k
3
k
4
Ranges SSE DOF F-value F critical values
Temperature (

C) 50.9 61.6 71.2 85.0 34.1 2518 3 0.63 3.49


Cr(VI)
ina
(mg/L) 99.7 71.0 52.7 45.3 54.4 7039 3 1.77 3.49
pH
ina
99.7 56.4 61.8 50.7 49.1 5912 3 1.49 3.49
B-nZVI
ina
(mg/L) 59.1 66.7 72.8 70.1 13.7 422 3 0.11 3.49
ina - initial; SSE - the Square Sum of Errors; DOF - Degree of Freedom.
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 888
(Fig. 2b) (U

zu m et al., 2009). The XRD patterns of B-nZVI after


reaction (Fig. 2b) indicated the presence of g-Fe
2
O
3
(2q 35.68),
Fe
3
O
4
(2q 35.45) and Cr
2
FeO
4
(2q 35.50), which were not
detected in the sample before reaction (Chen et al., 2008). The
appearance of Fe(II), Fe(III) and Cr(III) in B-nZVI after reaction
demonstrated the occurrence of redox reactions between Fe
0
and Cr(VI) where nZVI particles were acting as reductants,
which was consistent with previous literature (Ponder et al.,
2000).
3.2. Orthogonal test
The designed complex conditions and the nal removal ef-
ciencies of Cr(VI) from aqueous solution by B-nZVI were listed
in Table 1. Results were processed using the Orthogonal
Design Assistant software (El Hajjouji et al., 2008) in Range
Analysis and Variance Analysis (Table 2). The higher the range
and the F-value, the more signicant the factor was and
the greater the inuence of the factor on Cr(VI) removal
Comparing the ranges and F-values in Table 2, factors inu-
encing Cr(VI) removal were (in order of decreasing inu-
encing): initial Cr(VI) concentration > pH > temperature > B-
nZVI loading. The K values from K
1
to K
4
represented each
level of each factor, from the lowest to the highest. The higher
the K value, the higher the removal efciency, and the better
the level of the factor. Take temperature for example, the
highest K was K
4
, which represented the level 40

C, and this
made 40

C the optimumtemperature. The change in K values
indicated that chromiumremoval increased withtemperature
and decreased as initial Cr(VI) concentration and pH rose. The
optimal conditions for chromiumremoval were 40

C, 20 mg/L
of initial Cr(VI) concentration, 4 g/L B-nZVI loading and pH 2.0.
3.3. Conditions affecting Cr(VI) removal
After contacting for 3 h under identical conditions, the
removal efciencies of Cr(VI) were 5.5, 60.0 and 100.0%
respectively when bentonite, nZVI and B-nZVI were added
individually. Bentonite generally has poor adsorption of Cr(VI)
due to its negatively charged surface and the predominant
existence of Cr(VI) as anions (Bhattacharyya and Gupta, 2008).
The activity of nanoscale zero-valent iron particles was
enhanced signicantly when bentonite was introduced as
a support material, which conrmed the role bentonite played
as a dispersant and stabilizer in B-nZVI (Ponder et al., 2000).
Kinetics studies of Cr(VI) reduction using B-nZVI suggested
that the reactivity of nZVI particles supported on bentonite
were enhanced signicantly due to an increase in SSA and
a decrease in aggregation. Reduction kinetics of Cr(VI) by
B-nZVI were described by a pseudo rst-order reaction (Ponder
et al., 2000):
v
dc
dt
k
SA
a
s
r
m
c (1)
Where c was the concentration (mg/L) of contaminant, k
SA
was the specic reaction rate constant associated with the
SSA of the materials (L/h m
2
), a
s
was the specic surface area
(m
2
/g), and r
m
was the mass concentration (g/L). For k
SA
, a
s
,
and r
m
are constant for a specic reaction, the product of the
three can be expressed with one parameter k
obs,
which is
called the observed rate constant of a pseudo rst-order
reaction (h
1
). Therefore Eq. (1) can be integrated into:
ln
c
c
0
k
obs
t (2)
The k
obs
values under different conditions are equal to the
slope of the line achieved by plotting lnc=c
o
versus time
under various conditions.
In this study, the plots of lnc=c
o
versus time produced
linear plots with correlation coefcients (R
2
) higher than 0.9
(Fig. 3). This indicated that the rate of Cr(VI) reduction by
B-nZVI tted well the pseudo rst-order model under various
conditions. Additionally, the reduction of Cr(VI) by B-nZVI
represented a solid-liquid inter-phase reaction, which agreed
with the pseudo rst-order kinetics model.
Table 3 e Remediation of actual electroplating
wastewater by B-nZVI.
Wastewater Total
Cr
Pb
2
Cu
2
Zn
2
pH
c
0
(mg/L) 73 2 13.1 0.6 33 1 284 4 1.9
c (mg/L) after reaction
a a
2.4 0.1 115 3 4.5
Removal amount
(mg/g B-nZVI)
7.3 1.3 3.0 16.8 e
Removal percentage (%) 100 100 92.7 59.4 e
a means the concentration of the heavy metal was under the limit
of detection.
Fig. 1 e SEM images of laboratory synthesized iron particles with and without a support material. a. NZVI; b. B-nZVI before
reaction with Cr(VI) solution; c. B-nZVI after reaction with Cr(VI) solution. The scale bar in the gure is 500 nm.
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 889
3.3.1. Effect of initial Cr(VI) concentration
The effect of initial Cr(VI) concentration on removal efciency
was investigated in the range 20e100 mg/L. The plot tted the
pseudo rst-order model well (Fig. 3a), where the observed
rate constant decreased signicantly as the initial Cr(VI)
concentration increased, which agreed with the orthogonal
test results. The equilibrium time became longer and the nal
removal efciency of Cr(VI) decreased as the initial Cr(VI)
concentration increased, so that while the percentage of Cr(VI)
removed within 20 min at a Cr(VI) concentration of 20 mg/L
was nearly 100%, it was only 30.4% within 60 min at a Cr(VI)
concentration of 100 mg/L. Generally, the slower rate and
lower efciency of Cr(VI) removal from aqueous solution were
found at higher concentrations of Cr(VI). Based on the SEM
analysis and previous research, Cr(VI) reduction by nZVI could
be dened as a surface-mediated process (Ponder et al., 2000;
Rivero-Huguet and Marshall, 2009). The more the Cr(VI) ion
approached the surface of nZVI dispersed on the bentonite,
the faster Fe
0
was oxidized into Fe(III) and the faster the co-
precipitation of Cr(III) and Fe(III) oxides/hydroxides occurred.
This reduced the reactivity of nZVI and subsequently resulted
Fig. 2 e X-ray diffractogram of B-nZVI. a. before the
reaction with Cr(VI) solution; b. after the reaction with Cr
(VI) solution.
Fig. 3 e Effects of various factors on Cr(VI) removal by tting to the pseudo rst-order model. a. initial Cr(VI) concentration;
b. B-nZVI loading; c. pH value; d. temperature.
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 890
in a decrease in k
obs
(Geng et al., 2009; Yuan et al., 2009). On the
other hand, the highest removal amount was obtained at an
initial Cr(VI) concentration of 70 mg/L, which could be
ascribed to the limited capacity of B-nZVI for Cr(VI) removal
determined by SSA.
3.3.2. Effect of B-nZVI loading
The initial loadings of B-nZVI in Cr(VI) solution were 1, 2, 3 and
4 g/L. The Cr(VI) removal percentage increased as the B-nZVI
concentration increased (Fig. 3b). The removal percentage of
Cr(VI) was 54.6% using B-nZVI at 1 g/L for 120 min, but was
nearly 100% when the B-nZVI loading was over 3 g/L. Mean-
while, k
obs
increased as the B-nZVI loading increased. These
phenomena can be attributed to the increase in the available
active sites resulting from the elevation in B-nZVI loading,
where the reduction of Cr(VI) occurred (Geng et al., 2009; Yuan
et al., 2009). However, the concentration of Cr(VI) decreased
dramatically in the initial 10 min, then slightly declined in the
later reduction. A few researchers (Ponder et al., 2000;
Manning et al., 2007) reported a sorption phase during the
reaction which could also be supported by our SEM images,
suggesting that the overall mechanismwas more complicated
than a simple chemical reaction.
3.3.3. Effect of the pH value
The two dominant forms of Cr(VI) in aqueous solution were
HCrO
4

, between pH 1.0 to 6.0 and CrO


4
2
above pH 6.0 (Mohan
and Pittman, 2006). The dependence of the reaction rate
constant on pH was investigated by adjusting the solution pH
to 2.0, 4.0, 6.0 and 8.0 with either 0.1 M HCl or NaOH (Yuan
et al., 2009). Except for pH 2.0, the reduction of Cr(VI) can be
described using the pseudo rst-order model well (Fig. 3c). A
remarkable increase in the removal rate occurred at pH 2.0,
where equilibrium was achieved within 1 min and the
residual Cr(VI) was below the detection limit, which made
kinetic tting infeasible. The Cr(VI) removal percentage
decreased signicantly with increases in the initial pH, so that
only 27.2% Cr(VI) was reduced at pH 8.0 in 20 min while nearly
100% Cr(VI) was removed in 1 min at pH 2.0. In addition, k
obs
was respectively 0.0275, 0.0163, and 0.0083/min, when the
initial pH value was 4.0, 6.0 and 8.0, indicating that the
reduction rate increased as pH decreased, which had also
been reported in other studies (Geng et al., 2009; Yuan et al.,
2009). These results demonstrated that a lower pH favored
Cr(VI) reduction, since at lower pH corrosion of nZVI was
accelerated and the precipitation of Cr(III) and Fe(III) hydrox-
ides on the surface of iron was consequently not as favorable,
which led to an increase in the reaction rate (Lee et al., 2003).
Furthermore, the increase in H

concentration left the surface


of bentonite less negatively charged, which reduced the
electrostatic repulsion between bentonite and Cr(VI) anions.
This consequently promoted the electron transfer between
zero-valent iron and Cr(VI) (Yuan et al., 2009).
3.3.4. Effect of temperature
To assess the effect of different temperatures, batch experi-
ments were conducted at 25, 30, 35 and 40

C. The results
showed that 82.4% Cr(VI) was removed at 40

C while only
73.4% Cr(VI) was reduced at 25

C in 60 min. Thus zero-valent
iron could have a positive effect on Cr(VI) reduction even at
ambient temperature. Fig. 3d highlights the relationship
between lnc=c
o
and time, where the linearity suggested that
the reduction of Cr(VI) at different temperatures in the pres-
ence of B-nZVI tted pseudo rst-order dynamics (Ponder
et al., 2000; Manning et al., 2007). The k
obs
was 0.020, 0.023,
0.026 and 0.030/min at four temperatures (25, 30, 35 and 40

C),
showing that an increase in the reaction temperature resulted
in an improved reaction rate. The apparent activation energy
(Ea) of Cr(VI) reduction by B-nZVI was 24.9 kJ/mol, demon-
strating that it is a chemically controlled adsorption process
having an Ea value higher than 21 kJ/mol (Geng et al., 2009).
3.4. B-nZVI used to remove Cr(VI) from electroplating
wastewater and B-nZVI reuse
The data obtained from batch experiments where B-nZVI was
used to remove Cr(VI) and other metals from electroplating
wastewater are presented in Table 3, which indicated that B-
nZVI had the capacity to remove various heavy metals and
was a potential promising candidate for applications to in situ
environmental remediation. After reacting 10 mL of the
wastewater with 0.1 g of B-nZVI for 4 h, the residual concen-
tration of each metal ion showed that 100% total Cr, 100% Pb
(II), 92.7% Cu(II), and 59.4% Zn(II) were removed, following
treatment with B-nZVI. Total Cr, Pb(II) and Cu(II) received
higher removal percentages due to their higher standard
reduction potentials compared to Fe(II) (f
q
FeII=Fe
0
0:44V). In
contrast, a lower removal percentage of Zn(II) was obtained
because the standard reduction potential of Zn(II)
(f
q
ZnII=Zn
0
0:762V) was more negative than Fe(II) (Ladd,
2004).
The amount of Cr(VI) removed when using B-nZVI treated
with different concentrations of EDTA after being used four
times was calculated and it was shown that B-nZVIs ability to
remove Cr(VI) was dramatically reduced after being used only
once with an initial Cr(VI) concentration of 50 mg/L (Fig. 4).
The rapid deterioration of B-nZVI was ascribed to the inability
of the redox reaction between Cr(VI) and Fe
0
to proceed
Fig. 4 e The variation of Cr(VI) removal amount by B-nZVI
after reusing four different times. The solutions of EDTA
used for treatment of B-nZVI were 50 mg/L and 10 mg/L
respectively as marked in the gure.
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 891
further since this was a chemical controlled and irreversible
process. As conrmed by XRD analysis (Fig. 2), reaction
products were deposited on the surface of B-nZVI in the form
of oxide-hydroxide co-precipitation of Fe(II), Fe(III) and Cr(III),
which consequently decreased the activity of Fe
0
(Chen et al.,
2008). This phenomenon conrmed that the active ingredient
of B-nZVI was Fe
0
which acted as a reductant, while bentonite
only played a role as a dispersant and stabilizer.
4. Conclusions
In this study, nZVI particles became more effective when
bentonite was introduced as a support material due to
reduction of aggregation and increased SSA. Batch experi-
ments indicated that the removal rate increased as the
temperature and B-nZVI loading (4 g/L) increased, and fell as
the initial Cr(VI) concentration and pH increased, which
agreed with the result obtained from the orthogonal test.
Under the various operational conditions considered, reduc-
tion of Cr(VI) using B-nZVI was in accordance with a pseudo
rst-order model. B-nZVI was effective in removing Cr(VI) and
other heavy metals, including Pb(II), Cu(II) and Zn(II) from
electroplating wastewater. Since bentonite is a stable and low-
cost clay mineral, B-nZVI could be an efcient and promising
remediation material to remove Cr(VI) and other metals from
wastewater. However, further research must be carried out to
slow and control the degree of nZVI oxidation in the atmo-
sphere and as a consequence a more effective regeneration
method may emerge from such studies.
Acknowledgements
This work is supported by the Fujian Minjiang Fellowship
from Fujian Normal University (gs1). The authors also
gratefully acknowledge the signicant contributions of Dr
Gary Owens in editing and improving the many revisions of
this manuscript, his suggestions and corrections have
undoubtedly signicantly improved the quality of the nal
manuscript.
r e f e r e n c e s
Bhattacharyya, K.G., Gupta, S.S., 2008. Adsorption of a few heavy
metals on natural and modied kaolinite and
montmorillonite: a review. Adv. Colloid Interface Sci. 140,
114e131.
Celebi, O., U

zu m, C ., Shahwan, T., Erten, H.N., 2007. A radiotracer


study of the adsorption behavior of aqueous Ba
2
ions on
nanoparticles of zero-valent iron. J. Hazard. Mater. 148,
761e767.
Chen, S.S., Hsu, B.C., Hung, L.W., 2008. Chromate reduction by
waste iron from electroplating wastewater using plug ow
reactor. J. Hazard. Mater. 152, 1092e1097.
Cumbal, L., Greenleaf, J., Leun, D., SenGupta, A.K., 2003. Polymer
supported inorganic nanoparticles: characterization and
environmental applications. React. Funct. Polym. 54, 167e180.
El Hajjouji, H., Ait Baddi, G., Yaacoubi, A., Hamdi, H., Winterton, P.,
Revel, J.C., Adi, M., 2008. Optimisation of biodegradation
conditions for the treatment of olive mill wastewater.
Bioresour. Technol. 99, 5505e5510.
Geng, B., Jin, Z., Li, T., Qi, X., 2009. Kinetics of hexavalent
chromium removal from water by chitosan-Fe
0
nanoparticles.
Chemosphere 75, 825e830.
Ju-Nam, Y., Lead, J., 2008. Manufactured nanoparticles: an
overview of their chemistry, interactions and potential
environmental implications. Sci. Total Environ. 400, 396e414.
Katz, S., Salem, H., 1994. The Biological and Environmental
Chemistry of Chromium. VCH Publishers, New York.
Ladd, M.F.C., 2004. Introduction to Physical Chemistry, third ed.
Cambridge University Press, Cambridge.
Lee, T., Lim, H., Lee, Y., Park, J.W., 2003. Use of waste iron metal
for removal of Cr(VI) from water. Chemosphere 53, 479e485.
Li, X., Elliott, D.W., Zhang, W.X., 2006. Zero-valent iron
nanoparticles for abatement of environmental pollutants:
materials and engineering aspects. Crit. Rev. Solid State Mater.
Sci. 31, 111e122.
Li, Z., Kirk Jones, H., Zhang, P., Bowman, R.S., 2007. Chromate
transport through columns packed with surfactant-modied
zeolite/zero valent iron pellets. Chemosphere 68, 1861e1866.
Liu, Y., Phenrat, T., Lowry, G.V., 2007. Effect of TCE concentration
and dissolved groundwater solutes on NZVI-promoted TCE
dechlorination and H
2
evolution. Environ. Sci. Technol. 41,
7881e7887.
Manning, B.A., Kiser, J.R., Kwon, H., Kanels, S.R., 2007.
Spectroscopic investigation of Cr (III)-and Cr (VI)-treated
nanoscale zerovalent iron. Environ. Sci. Technol. 41, 86e592.
Mohan, D., Pittman, C.U., 2006. Activated carbons and low cost
adsorbents for remediation of tri-and hexavalent chromium
from water. J. Hazard. Mater. 137, 62e811.
Ngomsik, A., Bee, A., Draye, M., Cote, G., Cabuil, V., 2005. Magnetic
nano-and microparticles for metal removal and
environmental applications: a review. Comptes Rendus-
Chimie 8, 963e970.
Ponder, S.M., Darab, J.G., Mallouk, T.E., 2000. Remediation of Cr
(VI) and Pb (II) aqueous solutions using supported, nanoscale
zero-valent iron. Environ. Sci. Technol. 4, 2564e2569.
Rivero-Huguet, M., Marshall, W.D., 2009. Reduction of hexavalent
chromium mediated by micron-and nano-scale zero-valent
metallic particles. J. Environ. Monit. 11, 1072e1079.
U

zu m, C ., Shahwan, T., Ero lu, A.E., Hallam, K.R., Scott, T.B.,


Lieberwirth, I., 2009. Synthesis and characterization of
kaolinite-supported zero-valent iron nanoparticles and their
application for the removal of aqueous Cu
2
and Co
2
ions.
Appl. Clay Sci. 43, 172e181.
Yuan, P., Fan, M., Yang, D., He, H., Liu, D., Yuan, A., Zhu, J.,
Chen, T., 2009. Montmorillonite supported magnetite
nanoparticles for the removal of hexavalent chromium [Cr
(VI)] from aqueous solutions. J. Hazard. Mater. 166, 821e882.
wa t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 8 8 6 e8 9 2 892

You might also like