You are on page 1of 8

Journal of Contaminant Hydrology 114 (2010) 35–42

Contents lists available at ScienceDirect

Journal of Contaminant Hydrology


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

Reduction of hexavalent chromium by carboxymethyl cellulose-stabilized


zero-valent iron nanoparticles
Qian Wang a, Huijing Qian a, Yueping Yang a,b, Zhen Zhang a, Cissoko Naman a, Xinhua Xu a,⁎
a
Department of Environmental Engineering, Zhejiang University, Hangzhou 310027, PR China
b
Zhejiang Chimey Environment Science & Technology Co., Ltd, Hangzhou 310030, PR China

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

Article history: The reduction of hexavalent chromium or Cr(VI) by zero-valent iron (Fe0) nanoparticles has
Received 5 March 2009 received increasing attention in recent years. However, Fe0 nanoparticles prepared using
Received in revised form 24 February 2010
conventional methods suffered several drawbacks due to their high reactivity towards
Accepted 24 February 2010
Available online 2 March 2010
surrounding media, which led to the formation of much larger flocs and significant loss in
reactivity. To overcome these problems, we synthesized Fe0 nanoparticles by applying water-
soluble carboxymethyl cellulose (CMC) as a stabilizer. CMC-stabilized Fe0 nanoparticles
Keywords:
Hexavalent chromium displayed much less agglomeration but greater Cr(VI) reduced power than those prepared
Iron nanoparticles without a stabilizer. At a dose of 0.15 g L−1, CMC-stabilized Fe0 nanoparticles were able to
Carboxymethyl cellulose reduce 100% of 10 mg L−1 Cr(VI) in minutes. Several factors that may affect the efficiency of Cr
Remediation (VI) removal were investigated. These included the concentration of CMC, the concentration of
Fe0 nanoparticles, the initial Cr(VI) concentration, the pH value, the reaction temperature and
the concentration of the calcium cation in the reaction mixture. Our study suggested that the
introduction of an innocuous stabilizer such as CMC could significantly improve the
performance of Fe0 nanoparticles for environmental remediation applications.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction much less toxic. Indeed, Cr(III) is a required nutrient for


humans and animals (Kotas and Stasicka, 2000; Vainshtein et
Hexavalent chromium, Cr(VI), is a common contaminant in al., 2003; Costa and Klein, 2006). Hence, the reduction of Cr(VI)
soil and groundwater (Vainshtein et al., 2003). It is commonly to Cr(III) is environmentally favorable, and can be used to
found in wastewater from electroplating and metal finishing remediate Cr(VI) contaminated sites.
processes, pigment manufacturing, tannery facilities, and Much of the research has focused on the remediation of Cr
chromium mining operations. Cr(VI) is carcinogenic to (VI) and many treatment processes have been developed. For
human and animals. It can cause dermatitis, rhinitis even lung example, the remediation of Cr(VI) by physicochemical
cancer and nasopharynx cancer (Fendorf et al., 2000; Costa and adsorption has long been investigated (Hideaki et al., 2002;
Klein, 2006). In an aqueous environment, Cr(VI) is highly Bowman, 2003), but its operational cost is high. Moreover, Cr
mobile and can impact large-volume aquifers. The mobile form (VI) is often just transferred, but not removed. Bioremedia-
of Cr(VI) has the potential for increased exposure and harm to tion by strains of bacteria (Chen and Hao, 1998) is a cost-
human health and the environment (Vainshtein et al., 2003; effective alternative to reduce Cr(VI) to Cr(III), but a variety of
Costa and Klein, 2006). In contrast, the reduced form of bactericidal toxicants (e. g. hydrocarbon) at many waste sites
chromium, Cr(III), is less mobile in the environment and may limit their growth and effectiveness (Hua and Deng,
2003). Chemical reduction is also known to transform Cr(VI)
⁎ Corresponding author. Tel.: +86 571 87951239; fax: +86 571
to Cr(III) rapidly and effectively. Many reducing agents have
87952771. been reported, these include H2S, Fe2+, Fe0, etc (Buerge and
E-mail address: xuxinhua@zju.edu.cn (X. Xu). Hug, 1999; Hua and Deng, 2003). Cr(VI) reduction by Fe0

0169-7722/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jconhyd.2010.02.006
36 Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42

(Powell et al., 1995; Powell and Puls, 1997; Pratt et al., 1997; stabilizer for the removal of Cr(VI). Specifically, water-soluble
Ponder et al., 2000; Alowitz and Scherer, 2002; Gheju and CMC was used as the stabilizer during the preparation of the
Iovi, 2006) appears to be one of the most promising Fe0 nanoparticles. The specific aims were (1) to prepare water
technologies. Powell et al. (1995) suggested that the dispersed Fe0 nanoparticles stabilized by CMC under inert
mechanism of Cr(VI) reduction by Fe0 is a cyclic, multiple- conditions; and (2) to characterize the resulting CMC-
step electrochemical corrosion process and confirmed that stabilized Fe0 nanoparticles with regard to their physical
aluminosilicate minerals could enhance the rate of Cr(VI) stability and chemical reactivity for reductive removal of Cr
reduction. Alowitz and Scherer (2002) evaluated the effect of (VI) from water.
the different type of iron metal, the Fe0 area concentration
and the pH value on the removal efficiency of Cr(VI) 2. Experiments and methods
reduction by Fe0. Fe0 nanoparticles, due to their extremely
high surface areas, can enhance the reduction efficiencies 2.1. Chemicals
remarkably. Typically, Fe0 nanoparticles are prepared by
reducing Fe(II) or Fe(III) in an aqueous solution using a strong All chemicals, such as potassium dichromate (K2CrO4),
reducing agent (e.g., NaBH4). However, because of their high water-soluble carboxymethyl cellulose (CMC), ferrous sulfate
surface area and high reactivity, Fe0 nanoparticles prepared (FeSO4·7H2O), sulfuric acid (H2SO4) and acetone were of
by this approach tend to react rapidly with their surrounding analytical reagent grade. Waste iron filings (0.55–0.85 mm)
media (e.g., dissolved oxygen, or water) or agglomerate were obtained from the Machine Factory of Zhejiang Univer-
rapidly, resulting in the formation of much larger particles sity. The commercial Fe0 powder (N98%, size passing through
and rapid loss in reactivity (He and Zhao, 2005; Xu and Zhao, a mesh screen to maintain their diameter less than 0.075 mm)
2007). As a result, new processes are being investigated to was purchased from Jinshan Metallurgical Factory. Both types
prepare stable Fe0 nanoparticles using stabilizers such as of iron were pretreated by washing with acetone and a sulfuric
cetylpyridinium chloride (CPC) (Chen et al., 2004), starch (He acid solution (pH = 2), which created fresh sites for oxidation
and Zhao, 2005), hydrophilic carbon, poly (acrylic acid) and made the surface more active. Deionized water was used
(Schrick et al., 2002, 2004), as well as other surfactants and for preparing all solutions.
polymers (Shen et al., 1999; Sousa et al., 2001; Wu et al., The Fe0 and CMC-stabilized Fe0 nanoparticles were freshly
2005; Lu et al., 2007; Sun et al., 2007; Saleh et al., 2008; synthesized before use. They were prepared by following the
Tiraferri and Sethi, 2008; Tiraferri et al., 2008; Bai et al., 2009; methods reported by Wei et al. (2004) and He and Zhao
Parshetti and Doong, 2009). Recently, carboxymethyl cellu- (2005). In brief, the iron nanoparticles were synthesized by
lose (CMC) was used as a stabilizing agent for preparing adding stoichiometric amounts of NaBH4 aqueous solution at
bimetallic nanoparticles in aqueous media (Raveendran et al., a BH4−/Fe2+ molar ratio of 2.0 dropwise to a 1000 mL three-
2003; Xu and Zhao, 2007). necked flask containing FeSO4·7H2O aqueous solution and
A range of Fe(0) and Fe(II)-bearing materials can promote 0.2% (w/w) CMC with electrical stirring at ambient temper-
the reduction and precipitation of Cr(VI). The net reactions of ature. Before use, deionized water and the CMC solution were
Cr(VI) reduction with Fe0 and co-precipitation of Cr(III) and purged with purified N2 to remove dissolved oxygen. The
Fe(III) can be written as below (Lee et al., 2003): ferrous iron was reduced to zero-valent iron according to the
following reaction:
2− 0 þ 3þ 3þ
CrO4 þ Fe þ 8H →Fe þ Cr þ 4H2 O ð1Þ
2þ − 0
FeðH2 OÞ6 þ 2BH4 →Fe ↓ þ 2BðOHÞ3 þ 7H2 ↑ ð3Þ
Cellulose is a carbohydrate consisting of a series of
3þ 3þ þ hydroglucose units interconnected by an oxygen linkage
ð1−xÞFe þ ðxÞCr þ 2H2 O→Feð1 − xÞ Crx OOH↓ þ 3H ð2Þ
(known as a beta linkage) to form a linear molecular chain
To prepare physically more stable and chemically more structure. Cellulose can easily be modified to become CMC by
reactive Fe0-based nanoparticles, Mallouk et al. employed replacing the native CH2OH group in the glucose unit with a
carbon nanoparticles and poly (acrylic acid) (PAA) as carboxymethyl group (He et al., 2007), the molecular
supports for iron and bimetallic nanoparticles (Ponder et al., structure of CMC is shown below:
2000, 2001; Schrick et al., 2002; Zhu et al., 2006). These
supports prevented iron particles from agglomerating and
thereby prolonged the reactivity of the particles. Generally,
polymers such as CMC, guar gum, chitosan, and PAA provide
steric stabilization that exhibit a larger repulsion force than
electrostatic repulsion (Geng et al., 2009), hence they can
help to stabilize Fe0 nanoparticles (Kanel et al., 2008; Tiraferri
and Sethi, 2008; Tiraferri et al., 2008) and superparamagnetic
ferrofluid (Lin et al., 2005) via carboxylate binding. CMC is a
commercial, environmentally friendly material and has
already been applied for stabilizing Fe0 nanoparticles (Xu
and Zhao, 2007; He et al., 2009).
The primary objective of this work was to prepare a new Another important reason for selecting CMC is that CMC is
class of Fe0 nanoparticles involving the use of an innocuous a food-grade ingredient, nontoxic and biodegradable, likely
Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42 37

due to the presence of highly biodegradable –OH, –CO–, and


–COOH groups.

2.2. Batch experiments

The batch experiments for the reduction of Cr(VI) were


performed in the same three-necked flask to which Fe0
nanoparticles were introduced, followed by the addition of
CrO2−
4 aqueous solution. The reaction solution was stirred
under a nitrogen atmosphere to maintain reducing conditions
and periodically sampled by a glass syringe. The samples
were filtered immediately through 0.22 µm membrane filters
before further analyses.
The experimental conditions were as follows: 4.0–10.0 mL
of 1.0 g L−1 Cr(VI) stock solution and deoxygenated deionized
water were added into the flask containing 0.03–0.075 g of
freshly prepared CMC-stabilized Fe0 nanoparticles, the total
reaction volume was 500 mL. The reaction solution was stirred
under a nitrogen atmosphere to simulate an anaerobic
environment in groundwater at 25 °C, the initial pH was 5.5,
and the hydraulic retention time was 1 h. The stirring speed
was 400 rpm. Samples were periodically collected at 0, 2, 5, 15,
and 60 min with glass syringes and the reaction was stopped by
passing the aliquots through 0.22 µm membrane filters, all the
samples were analyzed within 4 h. Conditional experiments
were c onduc ted under the si milar experimen tal
conditions. Table 1 provides the details of the experiment
parameters for the conditional experiments. Fe0 nanoparticles
of different sizes were compared at appropriate concentrations.

2.3. Characterization and analytical methods

Fresh Fe0 nanoparticles (including CMC-stabilized Fe0


nanoparticles) were examined under a JEOL JEM 200CX
transmission electron microscope (TEM) at 180 kV. Prior to
TEM analysis, the particles were dispersed by an ultrasonicator.
Cr(VI) concentrations were determined via the diphenyl
Fig. 1. TEM images of Fe0 nanoparticles prepared (a) without CMC, and (b)
carbazide colorimetric method by measuring the absorbance
with CMC.
at 540 nm using a UV–VIS spectrophotometer. The pH value
was measured by a JENCO-6173 pH meter.

stabilizer. Fig. 1a shows that the resulting particles without


3. Results and discussion
the stabilizer do not appear as discrete nanoscale particles,
rather, they form much bulkier dendritic flocs with varying
3.1. Characterization of both Fe0 nanoparticles
optical density. The size of some denser flocs can be much
larger than 1 μm. This type of aggregation could be attributed
Fig. 1 depicts transmission electron micrographs of the Fe0
to the electrostatic forces between nanoscale particles and
nanoparticles prepared (a) with and (b) without a CMC
small particles, as well as their surface tension interactions
(Cushing et al., 2004; Manning et al., 2007). As a result, a lower
surface area is likely and a poorer reactivity is expected. In
Table 1 contrast, the CMC-stabilized Fe0 nanoparticles shown in
Experimental parameters values. Fig. 1b appear to be clearly discrete and well-dispersed, the
particles are close to spherical in shape, and their sizes range
CMC Fe0 Cr initial pH Temperature Ca2+
concentration nanoparticles concentration from 20 to 100 nm in diameter. Evidently, the presence of
concentration CMC minimized agglomeration of the resulting iron particles
(g L−1) (g L−1) (mg L−1) (°C) (mM) and thus maintained the high surface area and potential
0 / / / / 0 reactivity of the particles. Fig. 1b shows the volume weighted
0.1 0.06 8 4.5 10 1 mean hydrodynamic diameter and size distribution of ZVI
0.3 0.08 10 5.5 15 2 nanoparticles synthesized in the presence of CMC with a fixed
0.5 0.10 15 8.5 25 5 Fe2+ concentration of 0.1 g L−1 and CMC:Fe2+ = 5:1, the
1.0 0.15 20 10.5 35
volume weighted distribution consisted primarily (about
38 Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42

80.0%) of particles 50 nm in diameter, and dispersed as a concentrations of 0, 0.1, 0.3, 0.5 and 1.0 g L−1, respectively.
colloidal suspension in solution. Conversely, in pristine Fe0 as The addition of CMC to Fe0 nanoparticles significantly improved
Fig. 1a, more than 95% of the ZVI particles formed aggregates the Cr(VI) removal efficiency. The removal efficiency is defined
that were greater than 300 nm in length. as the fraction of Cr(VI) in the solution during the reaction time
To further examine the stability of CMC-stablized Fe0 as a fraction of the initial concentration. At a concentration of
nanoparticles, precipitation experiments were carried out, in 1.0 g L−1 of CMC added, a lower removal efficiency was
which 20 mL of nanoparticle suspensions (i.e., bare- and observed. It is likely that the CMC occupied a part of the surface
CMC-Fe0 nanoparticles) were transferred from the flask of the Fe0 nanoparticles at that concentration, thus inhibiting
reactor into 25-mL tubes with tapered bottoms as shown in the reaction.
Fig. 2. The Fe0 nanoparticles prepared without the CMC
settled at the bottom of the tube in less than 30 min while the 3.3. Effect of Fe0 nanoparticle concentrations on Cr(VI) removal
CMC-stabilized Fe0 nanoparticles remained in suspension
over an extended period of time (N15 d) with no noticeable Four CMC–Fe0 nanoparticle concentrations were employed
sedimentation or flocculation (i.e., with stable particle sizes in this study. Fig. 3b shows that increasing the concentration of
and solution density). These results further confirmed that Fe0 nanoparticles greatly enhanced the Cr(VI) removal
the presence of CMC prevented agglomeration of the efficiency. With Fe0 nanoparticles concentration increasing
resulting Fe0 nanoparticles and thus maintained the high from 0.06 to 0.08, 0.10 and 0.15 g L−1, the ratio of Fe0:Cr
surface area and potential reactivity of the particles. changed from 6:1 to 8:1, 10:1 and 15:1. 100% of Cr(VI) was
removed when the Fe0 nanoparticles mass concentration was
3.2. Effect of CMC concentration on Cr(VI) removal 0.15 g L−1, but only 67% was removed when the mass
concentration was 0.06 g L−1 over a 60 min period. When the
Four different CMC concentrations were investigated in reaction proceeded, the pH quickly rose from 5.5 to between 6.0
this study (while other parameters were fixed and not and 6.3, and remained basically unchanged in the next reaction.
changed). Fig. 3a shows that the increasing concentration of It is believed that the Cr(VI) reductive reaction occurs on the Fe0
CMC greatly enhanced the removal efficiency when the nanoparticles' surfaces. As the Fe0 nanoparticles mass concen-
concentration of CMC was less than 0.5 g L−1. The removal tration increases, the reactive Fe sites increase proportionally,
percentages of Cr(VI) at different CMC concentrations over a which lead to the increase of Cr(VI) removal efficiency.
period of 60 min were 52, 69, 77, 94 and 88%, for CMC
3.4. Effect of initial Cr(VI) concentration on Cr(VI) removal

Aqueous Cr(VI) concentrations in the batch reactors at


room temperature (25 °C) as a function of elapsed time, with
different initial Cr(VI) concentrations, are shown in Fig. 3c.
The results show that for initial Cr(VI) concentrations
between 8 and 20 mg L−1, Cr(VI) removal efficiencies
decreased with increasing initial Cr(VI) concentration. The
removed fraction was 41% at a concentration of 20 mg L−1,
and 98% at the lowest initial concentration of 8 mg L−1. The
optimal mass ratio of Fe0 nanoparticles to Cr(VI) is about
10:1. The final removal rate of Cr(VI) decreased with the
increase of initial Cr(VI) concentration with the same amount
of the addition of CMC-stabilized Fe0 nanoparticles, but the
absolute removal amount increased with increasing initial Cr
(VI) concentration.

3.5. Effect of the initial pH values on Cr(VI) removal

Preliminary experiments indicated that Cr(VI) reduction


was very fast under acidic conditions (e.g., pH 4.5), but
proceeded slowly when the pH was increased up to 10.5. As
shown in Fig. 3d, the Cr(VI) removal efficiency increased
significantly with decreasing pH. When Cr(VI) was intro-
duced, the pH was changed quickly with the different intial
pH, and remained basically unchanged in the next reaction.
The final pH after the 60 min reaction were 5.9, 6.2, 8.6 and
10.5, for initial pH of 4.5, 5.5, 8.5 and 10.5, respectively. Acidic
conditions would accelerate the corrosion of CMC–Fe0, thus
enhancing Cr(VI) reduction. In the initial 5 min, Cr(VI)
Fig. 2. Sedimentation experiments with two different types of Fe0
removal was rapid for pH values below 5.0. It indicates that
nanoparticles. (a) Common Fe0 nanoparticles, and (b) CMC-stabilized Fe0 the Fe0 nanoparticles have a high reactivity. In contrast, the
nanoparticles, CMC = 0.5 g L−1, Fe0 = 0.1 g L−1. plots for pH N 8 show a less rapid removal. This may be
Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42 39

Fig. 3. Factors influencing the Cr(VI) removal efficiency by CMC-stabilized Fe0 nanoparticles. (a) CMC concentration changes, Fe0 = 0.1 g L−1, CMC = 0–1.0 g L−1,
Cr(VI) = 10 mg L−1, pH = 5.5, T = 25 °C, ω = 400 rpm; (b) Fe0 nanoparticles concentration changes, Fe0 = 0.06–0.15 g L−1, CMC:Fe0 = 5:1, Cr(VI) = 10 mg L−1,
pH = 5.5, T = 25 °C, ω = 400 rpm; (c) initial Cr(VI) concentration changes, Fe0 = 0.1 g L−1, CMC = 0.5 g L−1, Cr(VI) = 8–20 mg L−1, pH = 5.5, T = 25 °C,
ω = 400 rpm; (d) pH changes, Fe0 = 0.1 g L−1, CMC = 0.5 g L−1, Cr(VI) = 10 mg L−1, pH= 4.5–10.5, T = 25 °C, ω = 400 rpm; (e) temperature changes, Fe0 = 0.1 g L−1,
CMC = 0.5 g L−1, Cr(VI)= 10 mg L−1, pH= 5.5, T = 10–35 °C, ω = 400 rpm.

because of the formation of mixed Fe and Cr oxyhydroxides at requires 8 mol of hydrogen ions for each mol of Cr(VI) and is
high pH values on the iron surfaces (Powell et al., 1995; Lee highly dependent on H+ concentration.
et al., 2003). The near plateau segment of the plots after
60 min was probably due to the passivation of the surface 3.6. Effect of the reaction temperature on Cr(VI) removal
with the consequent loss of reactivity (Rivero-Huguet, and
Marshall, 2009a,b; Wu et al., 2009). This is also in accordance The effect of the reaction temperature on the reduction of
with the stoichiometry of this reaction (Eq. (1)) which Cr(VI) by CMC-stabilized Fe0 nanoparticles was examined at
40 Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42

pH 5.5 and an initial Cr(VI) concentration of 10 mg L−1, for 3.8. Effect of Ca2+ on Cr(VI) removal
temperature values ranging from 10 to 35 °C, using a
temperature-controlled water bath. The results, presented Four different Ca2+ concentrations were studied in this
in Fig. 3e, show an important dependence of the Cr(VI) study. Fig. 5 shows that the increased Ca2+ concentration has
removal efficiencies on the reaction temperature. The Cr(VI) an obvious inhibitory effect on Cr(VI) removal. The removal
removal efficiency is 97% at 35 °C, but just 75% at 10 °C. efficiencies of Cr(VI) at different Ca2+ concentrations in
60 min were determined to be 94, 47, 56 and 61%, for Ca2+
concentrations of 0, 1, 2 and 5 mM, respectively. The probable
3.7. Comparison of different types of Fe0 nanoparticles reason was with the increasing of Ca2+ concentration in the
solution, Na+ ions are replaced by the divalent cations,
Fig. 4 shows that the removal of Cr(VI), especially for Fe0 resulting in CMC-metal precipitates, which accordingly
nanoparticles and CMC-stabilized Fe0 nanoparticles, was rapid blocked the Cr(VI) reduction (He and Zhao, 2007). The
in the first minutes. This indicates that the predominant concentration of Ca2+ further increased, thus may lead to the
mechanism for the removal of Cr(VI) is most likely due to the formation of Ca(OH)2 and the absorption of Cr(VI) by Ca
adsorption to the Fe0 nanoparticles surface rather than the (OH)2 may also take place.
reductive transformation (Geng et al., 2009). After the first
minutes till the end of the experiment, the Cr(VI) concentration 3.9. Electrochemical analysis of the reaction process
decreased gradually. In aqueous solutions, Fe0 nanoparticles
will react with water to generate Fe(II) which then reacts with The standard reduction potentials of E Fe(II)/Fe(0) and E Cr
Cr(VI) rapidly (Kieber and Helz, 1992; Kendelewicz et al., 2000) are −0.44 and 1.33 V, respectively. When mixed
(VI)/Cr(III)
resulting in the disappearance of Cr(VI) from water in the with Cr2O2− 0
7 , Fe would be corroded and dissolved, causing
second phase(Geng et al., 2009). The removal efficiency was the reduction of Cr(VI) to Cr(III). The standard reduction
just 8% and 9% for Fe0 filings and powder after 60 min, but was potential of E Cr(III)/Cr(0) is −0.74 V, which is lower than that of
22% and 94% for Fe0 nanoparticles and CMC-stabilized Fe0 Fe0. This suggests that the Cr(III) or Cr(III) hydroxide might be
nanoparticles, respectively. CMC-stabilized Fe0 nanoparticles the steady product in the reaction. In fact, several investiga-
had a greater reactivity than Fe0 nanoparticles, and Fe0 tors have reported that the removal mechanism involves the
nanoparticles had a greater reactivity than Fe0 powder. Fe0 reduction of highly soluble Cr2O2− 7 to sparingly soluble Cr(III)
filings had the least reactivity. The comparison shows that Fe0 compounds at the surface of the iron (Powell et al., 1995; Lee
nanoparticles had a greater removal efficiency than Fe0 filings et al., 2003). Spectroscopic data have also shown that Cr(OH)3
or powder, but the removal of Cr(VI) from the solution ceased and mixed Cr(III)/Fe(III) hydroxides are precipitated on the
quickly, this may be due to aggregation. The CMC-stabilized Fe0 iron surface (Powell et al., 1995; Pratt et al., 1997; Ponder
nanoparticles apparently did not aggregate as much, yielding a et al., 2000). The net reactions of Cr(VI) reduction with Fe0
much increased removal efficiency. It has been proposed that and the co-precipitation of Cr(III) and Fe(III) could be
Fe0 nanoparticles reduce Cr(VI) primarily to Cr(III). Then Cr(III) expressed by Eqs. (1) and (2). No dissolved Cr(III) was
can also precipitate as Fe(III)–Cr(III) hydroxides on the Fe0 detected during the reaction, which suggests that all the Cr
nanoparticles surface. The formation of this surface passivation (III) was co-precipitated with Fe(III). This observation is
layer on Fe0 nanoparticles accordingly blocked Cr(VI) reduction consistent with those reported previously (Pratt et al., 1997;
(Alowitz and Scherer, 2002). For the CMC-stabilized Fe0 Williams and Scherer, 2001; He and Trainas, 2005).
nanoparticles system, CMC, can inhibit aggregation of the iron Insight into the process of Cr(VI) removal can also be
nanoparticles and possibly inhibit the formation of Fe(III)–Cr evaluated by determining the reduction potential E and pH
(III) precipitation, as was reported for chitosan by Geng et al. values in solution. The E and pH values of the Cr2O2− 7 solution
2009. (0.1 mM) were 0.239 V and 7.00, respectively. The reduction

Fig. 4. Comparison of different Fe0 types on the Cr(VI) removal efficiency, Fig. 5. Effect of Ca2+ on the Cr(VI) removal efficiency, CMC = 0.5 g L−1,
CMC = 0.5 g L−1, Fe0 = 0.1 g L−1, CMC:Fe0 = 5:1, Cr(VI) = 10 mg L−1, Fe0 = 0.1 g L−1, CMC:Fe0 = 5:1, Cr(VI) = 10 mg L−1, pH = 5.5, T = 25 °C,
pH = 5.5, T = 25 °C, ω = 400 rpm. ω = 400 rpm.
Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42 41

potential E decreased with increasing initial pH in the Fendorf, S., Wielinga, B.W., Hansel, C.M., 2000. Chromium transformations in
natural environments: the role of biological and abiological processes in
reaction, which indicates that Cr(VI) would be converted to chromium(VI) reduction. Int. Geol. Rev. 42, 691–701.
Cr(OH)3 gradually. After achieving its minimum, E increased Geng, B., Jin, Z.H., Li, T.L., Qi, X.H., 2009. Preparation of chitosan-stabilized Fe0
with the decreasing pH due to the likely formation of Cr(III)/ nanoparticles for removal of hexavalent chromium in water. Sci. Total
Environ. 407, 4994–5000.
Fe(III) hydroxides. These would cover the surface of Fe0 Gheju, M., Iovi, A., 2006. Kinetics of hexavalent chromium reduction by scrap
nanoparticles and reduce their activity. The reaction was in iron. J. Hazard. Mater. B135, 66–73.
the sample area where the Cr(OH)3 was dominant. Based on He, Y.T., Trainas, S.J., 2005. Cr(VI) reduction and immobilization by magnetite
under alkaline pH conditions: the role of passivation. Environ. Sci.
the afore-mentioned observation and previously documented Technol. 39, 4499–4504.
results (Pratt et al., 1997; Williams and Scherer, 2001; He and He, F., Zhao, D., 2005. Preparation and characterization of a new class of
Trainas, 2005), it can be concluded that iron hydroxide and starch-stabilized bimetallic nanoparticles for degradation of chlorinated
hydrocarbons in water. Environ. Sci. Technol. 39, 3314–3320.
chromium hydroxide could be the final and predominant
He, F., Zhao, D., 2007. Manipulating the size and dispersibility of zerovalent
products of Cr(VI) reduction by CMC-stabilized Fe 0 iron nanoparticles by use of carboxymethyl cellulose stabilizers. Environ.
nanoparticles. Sci. Technol. 41, 6216–6221.
He, F., Zhao, D., Liu, J., Roberts, C.B., 2007. Stabilization of Fe–Pd nanoparticles
with sodium carboxymethyl cellulose for enhanced transport and
4. Conclusions dechlorination of trichloroethylene in soil and groundwater. Ind. Eng.
Chem. Res. 46, 29–34.
He, F., Zhang, M., Qian, T.W., Zhao, D.Y., 2009. Transport of carboxymethyl
The reduction of Cr(VI) by CMC-stabilized Fe0 nanoparti- cellulose stabilized iron nanoparticles in porous media: column experi-
cles was studied using potassium dichromate (K2CrO4) as the ments and modeling. J Colloid. Interface Sci. 334, 96–102.
model contaminant. The concentration of CMC-stabilized Fe0 Hideaki, Y., Toshiiyuki, Y., Takashi, T., 2002. Adsorption of chromate and arsenate
by amino-functionalized MCM-41 and SBA-1. Chem. Mater. 14, 4603–4610.
nanoparticles had a significant effect on the reduction of Cr
Hua, B., Deng, B., 2003. Infulences of water vapor on Cr(VI) reduction by
(VI). When the mass ratio of Fe0 to Cr(VI) was about 10:1, a gaseous hydrogen sulfide. Environ. Sci. Technol. 37, 4771–4777.
maximum Cr(VI) removal efficiency of 100% was achieved. Kanel, S.R., Goswami, R.R., Clement, T.P., Barnett, M.O., Zhao, D., 2008. Two
The reaction occurred within a broad pH range and the dimensional transport characteristics of surface stabilized zero-valent
iron nanoparticles in porous media. Environ. Sci. Technol. 42, 896–900.
reaction efficiency increased significantly with decreasing Kendelewicz, T., Liu, P., Doyle, C.S., Brown, G.E., Nelson, E.J., Chambers, S.A.,
initial pH. The CMC-stabilized Fe0 nanoparticles exhibited a 2000. Spectroscopic study of the reaction of aqueous Cr(VI) with Fe3O4
higher removal efficiency than those prepared without a (111) surfaces. Surf. Sci. 469, 144–163.
Kieber, R.J., Helz, G.R., 1992. Indirect photoreduction of aqueous chromium
stabilizer because the CMC acted as a good dispersant to prevent (VI). Environ. Sci. Technol. 26, 307–312.
Fe0 nanoparticles from agglomerating. The electrochemical Kotas, J., Stasicka, Z., 2000. Chromium occurrence in the environment and
analysis of the reaction process suggested that iron hydroxide methods of its speciation. Environ. Pollut. 107, 263–283.
Lee, T., Lim, H., Lee, Y., Park, J., 2003. Use of waste iron metal for removal of Cr
and chromium hydroxide should be the final and predominant (VI) from water. Chemosphere 53, 479–485.
products of Cr(VI) reduction by Fe0 nanoparticles. This study Lin, C.L., Lee, C.F., Chiu, W.Y., 2005. Preparation and properties of poly(acrylic
demonstrated that Fe0 nanoparticles, especially those stabilized acid) oligomer stabilized superparamagnetic ferrofluid. J. Colloid
Interface Sci. 291, 411–420.
by CMC, could lead to a high removal efficiency for applications Lu, A.H., Salabas, E.L., Schuth, F., 2007. Magnetic nanoparticles: synthesis,
such as in-situ remediation of Cr(VI)-contaminated soils and protection, functionalization, and application. Angew. Chem. Int. Edit. 46,
groundwater. 1222–1244.
Manning, B.A., Kiser, J.R., Kwon, H., Kanel, S.R., 2007. Spectroscopic
investigation of Cr(III)- and Cr(VI)-treated nanoscale zerovalent iron.
Acknowledgements Environ. Sci. Technol. 41, 586–592.
Parshetti, G.K., Doong, R., 2009. Dechlorination of trichloroethylene by Ni/Fe
nanoparticles immobilized in PEG/PVDF and PEG/nylon 66 membrane.
The authors are grateful for the financial support of the Water Res. 43, 3086–3094.
Zhejiang Provincial Natural Science Foundation of China (No. Ponder, S.M., Darab, J.G., Mallouk, T.E., 2000. Remediation of Cr(VI) and Pb(II)
R5090033), the Zhejiang Provincial Water Pollution Control aqueous solutions using supported, nanoscale zero-valent iron. Environ.
Sci. Technol. 34, 2564–2569.
and Management Project of China (2008C13007-1) and the Ponder, S.M., Darab, J.G., Bucher, J., Caulder, D., Craig, I., Davis, L., Edelstein, N.,
National Water Pollution Control and Management Project of Lukens, W., Nitsche, H., Rao, L., Shuh, D.K., Mallouk, T.E., 2001. Surface
China (2008ZX07101-006). chemistry and electrochemistry of aqueous metal contaminants. Chem.
Mater. 13, 479–486.
Powell, R.M., Puls, R.W., 1997. Proton generation by dissolution of intrinsic or
References augmented aluminosilicate minerals for in situ contaminant remediation
by zero-valence-state iron. Environ. Sci. Technol. 31, 2244–2251.
Alowitz, M.J., Scherer, M.M., 2002. Kinetics of nitrate, nitrite, and Cr(VI) Powell, R.M., Puls, R.W., Hightower, S.K., Sabatini, D.A., 1995. Coupled iron
reduction by iron metal. Environ. Sci. Technol. 36, 299–306. corrosion and chromate reduction: mechanisms for subsurface remedi-
Bai, X., Ye, Z.F., Qu, Y.Z., Li, Y.F., Wang, Z.Y., 2009. Immobilization of nanoscale ation. Environ. Sci. Technol. 29, 1913–1922.
Fe0 in and on PVA microspheres for nitrobenzene reduction. J. Hazard. Pratt, A.R., Blowes, D.W., Ptacek, C.J., 1997. Products of chromate reduction on
Mater. 172, 1357–1364. proposed subsurface remediation material. Environ. Sci. Technol. 31,
Bowman, R.S., 2003. Applications of surfactant-modified zeolites to envi- 2492–2498.
ronmental remediation. Micropor. Mesopor. Mat. 61, 43–56. Raveendran, P., Fu, J., Wallen, S.L., 2003. Complete ‘green’ synthesis and
Buerge, I.J., Hug, S.J., 1999. Influence of mineral surfaces on chromium(VI) stabilization of metal nanoparticles. J. Am. Chem. Soc. 125, 13940–13941.
reduction by iron(II). Environ. Sci. Technol. 33, 4285–4291. Rivero-Huguet, M., Marshall, W.D., 2009a. Reduction of hexavalent chromi-
Chen, J.M., Hao, O.J., 1998. Microbial chromium(VI) reduction. Critical Rev. um mediated by micro- and nano-sized mixed metallic particles. J.
Environ. Sci. Teechnol. 28, 219–251. Hazard. Mater. 169, 1081–1087.
Chen, S.S., Hsu, H.D., Li, C.W., 2004. A new method to produce nanoscale iron Rivero-Huguet, M., Marshall, W.D., 2009b. Reduction of hexavalent chromium
for nitrate removal. J. Nanopart Res. 6, 639–647. mediated by micron- and nano-scale zero-valent metallic particles. J.
Costa, M., Klein, C.B., 2006. Toxicity and carcinogenicity of chromium Environ. Monit. 11, 1072–1079.
compounds in humans. Crit. Rev. Toxicol. 36, 155–163. Saleh, N., Kim, H.J., Phenrat, T., Matyjaszewski, K., Tilton, R.D., Lowry, G.V.,
Cushing, B.L., Kolesnichenko, V.L., O'Connor, C.J., 2004. Recent advances in the 2008. Ionic strength and composition affect the mobility of surface-
liquid-phase syntheses of inorganic nanoparticles. J. Chem. Rev. 104 (9), modified Fe0 nanoparticles in water-saturated sand columns. Environ.
3893–3946. Sci. Technol. 42, 3349–3355.
42 Q. Wang et al. / Journal of Contaminant Hydrology 114 (2010) 35–42

Schrick, B., Blough, J.L., Jones, A.D., Mallouk, T.E., 2002. Hydrodechlorination Vainshtein, M., Kuschk, P., Mattusch, J., Vatsourina, A., Wiessner, A., 2003.
of trichloroethylene to hydrocarbons using bimetallic nickel-iron Model experiments on the microbial removal of chromium from
nanoparticles. Chem. Mater. 14, 5140–5147. contaminated groundwater. Water Res. 37, 1401–1405.
Schrick, B., Hydutsky, B.W., Blough, J.L., Mallouk, T.E., 2004. Delivery vehicles Wei, J.J., Xu, X.H., Liu, Y., 2004. Kinetics and mechanism of dechlorination of
for zerovalent metal nanoparticles in soil and groundwater. Chem. o-chlorophenol by nanoscale Pd/Fe. Chem. Res. Chin. Univ. 20, 73–76.
Mater. 16, 2187–2193. Williams, A.G.B., Scherer, M.M., 2001. Kinetics of Cr(VI) reduction by
Shen, L., Laibinis, P.E., Hatton, T.A., 1999. Bilayer surfactant stabilized magnetic carbonate green rust. Environ. Sci. Technol. 35, 3488–3494.
fluids: synthesis and interactions at interfaces. Langmuir 15, 447–453. Wu, L., Shamsuzzoha, M., Ritchie, S.M.C., 2005. Preparation of cellulose
Sousa, M.H., Tourinho, F.A., Depeyrot, J., da Silva, G.J., Lara, M.C.F.L., 2001. acetate supported zero-valent iron nanoparticles for the dechlorination
New electric double-layered magnetic fluids based on copper, nickel, and of trichloroethylene in water. J. Nanopart. Res. 7, 469–476.
zinc ferrite nanostructures. J. Phys. Chem. B 105, 1168–1175. Wu, Y.J., Zhang, J.H., Tong, Y.F., Xu, X.H., 2009. Chromium (VI) reduction in
Sun, Y.P., Li, X.Q., Zhang, W.X., Wang, H.P., 2007. A method for the aqueous solutions by Fe3O4-stabilized Fe0 nanoparticles. J. Hazard.
preparation of stable dispersion of zero-valent iron nanoparticles. Mater. 172, 1640–1645.
Colloids Surf., A Physicochem. Eng. Aspects 308, 60–66. Xu, Y.H., Zhao, D.Y., 2007. Reductive immobilization of chromate in water
Tiraferri, A., Sethi, R., 2008. Enhanced transport of zerovalent iron and soil using stabilized iron nanoparticles. Water Res. 41, 2101–2108.
nanoparticles in saturated porous media by guar gum. J. Nanopart. Res. Zhu, B.W., Lim, T.T., Feng, J., 2006. Reductive dechlorination of 1, 2, 4-
11, 635–645. trichlorobenzene with palladized nanoscale Fe0 particles supported
Tiraferri, A., Chen, K.L., Sethi, R., Elimelech, M., 2008. Reduced aggregation on chitosan and silica. Chemosphere 65, 1137–1146.
and sedimentation of zero-valent iron nanoparticles in the presence of
guar gum. J. Colloid Interface Sci. 324, 71–79.

You might also like