ArticlePDF Available

A Review on Synthesis, Characterization and Applications of nano-Zero Valent Iron (nZVI) for Environmental Remediation

Taylor & Francis
Critical Reviews In Environmental Science and Technology
Authors:
  • Swiss Agency for Development and Cooperation

Abstract and Figures

The recent spark in the interest for the usage of nano zero valent iron (nZVI) as a remediation tool for contaminated land and groundwater is mainly due to its higher reactivity in comparison to micro zero valent iron, cost-effectiveness and potential to treat a broad range of contaminants. This paper reviews the recent developments and approaches made on synthesis on nZVI, strucuture and characterisation of nZVI, the challenges faced in the transport of nZVI in the subsurface environment and the augmentation of the motility of nZVI. The effective use of nZVI in remediating organic pollutants (halogenated organic compounds, pharmaceutical waste and azo dyes) and inorganic pollutants (Ni2+, PO43−, Co2+, Cu2+) carried out in recent studies have been discussed. The potential risks and limitations of this emerging nanotechnology are also addressed.
Content may be subject to copyright.
Full Terms & Conditions of access and use can be found at
http://www.tandfonline.com/action/journalInformation?journalCode=best20
Download by: [University of Auckland Library] Date: 23 March 2016, At: 20:06
Critical Reviews in Environmental Science and
Technology
ISSN: 1064-3389 (Print) 1547-6537 (Online) Journal homepage: http://www.tandfonline.com/loi/best20
A review on synthesis, characterization, and
applications of nano zero valent iron (nZVI) for
environmental remediation
Ritika Mukherjee, Rahul Kumar, Alok Sinha, Yangdup Lama & Amal Krishna
Saha
To cite this article: Ritika Mukherjee, Rahul Kumar, Alok Sinha, Yangdup Lama & Amal Krishna
Saha (2016) A review on synthesis, characterization, and applications of nano zero valent
iron (nZVI) for environmental remediation, Critical Reviews in Environmental Science and
Technology, 46:5, 443-466, DOI: 10.1080/10643389.2015.1103832
To link to this article: http://dx.doi.org/10.1080/10643389.2015.1103832
Accepted author version posted online: 07
Oct 2015.
Published online: 07 Oct 2015.
Submit your article to this journal
Article views: 245
View related articles
A review on synthesis, characterization, and applications of
nano zero valent iron (nZVI) for environmental remediation
Ritika Mukherjee, Rahul Kumar, Alok Sinha, Yangdup Lama, and Amal Krishna Saha
Department of Environmental Science and Engineering, Indian School of Mines, Dhanbad, Jharkhand,
India
ABSTRACT
The recent spark in the interest for the usage of nano zero
valent iron (nZVI) as a remediation tool for contaminated land
and groundwater is mainly due to its higher reactivity in
comparison to micro ZVI, cost effectiveness, and potential to
treat a broad range of contaminants. The authors review the
recent developments and approaches made on synthesis on
nZVI, strucuture and characterization of nZVI, the challenges
faced in the transport of nZVI in the subsurface environment,
and the augmentation of the motility of nZVI. The effective use
of nZVI in remediating organic pollutants (halogenated organic
compounds, pharmaceutical waste, and azo dyes) and
inorganic pollutants (Ni
2C
,PO
4
3¡
,Co
2C
,Cu
2C
) carried out in
recent studies is discussed. The potential risks and limitations
of this emerging nanotechnology are also addressed.
KEYWORDS
Nano zero valent iron (nZVI);
remediation; groundwater
1. Introduction
The need for a sustainable and safe water supply is compelling developing coun-
tries to develop innovative and economical methods for purication and treatment
of water and wastewater. The multidisciplinary nano boom has led to the evolu-
tion of a wide array of novel technologies for both domestic and industrial applica-
tions, including improved drug delivery and new methods for the treatment of
contaminated water (Bystrzejewska-Piotrowska et al., 2009). Nanomaterials have
been shown to possess unique chemical, catalytic, electronic, magnetic, mechani-
cal, and optical properties owing to their small size (Jortner and Rao, 2002). As the
particle size decreases, proportion of atoms at the surface increases, raising its ten-
dency to adsorb, interact, and react with other atoms, molecules, and complexes to
achieve charge stabilization. The potential use of nanoparticles for the treatment of
contaminated groundwater has sparked a great deal of interest (Fu et al., 2014).
CONTACT Alok Sinha aloksinha11@yahoo.com Department of Environmental Science and Engineering,
Indian School of Mines, Dhanbad, Jharkhand, 826 004, India.
Color versions of one or more of the gures in the article can be found online at www.tandfonline.com/best.
© 2016 Taylor & Francis Group, LLC
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY
2016, VOL. 46, NO. 5, 443466
http://dx.doi.org/10.1080/106 43389.2015.1103832
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
These particles can be used for in situ remediation of groundwater and can replace
costly technology such as permeable reactive barriers, currently in usage.
The key properties essential for the use of any engineered nanoparticle for in
situ remediation includes (a) high reactivity for contaminant removal, (b) suf-
cient mobility within porous media, (c) sufcient reactive longevity, and (d) low
toxicity. However, it should be noted that the process involved is at a sensible cost
and competitive with other existing technology (Zhang, 2003; Crane and Scott,
2012). The use of nanometals, such as iron and bimetallic particles (OCarroll,
2013), for subsurface remediation of halogenated organic compounds and heavy
metal contaminated sites has received signicant attention in the past; however,
nanoscale zero valent iron is most commonly used. The adaptation of nano zero
valent iron (nZVI) provides several advantages compared to microscale ZVI,
which is used as reactive material in permeable reactive barriers, including (a) an
increase in the reductive degradation reaction rate, (b) a decrease of the reductant
dosage, (c) control over the risk of release of toxic intermediates, and (d) the gener-
ation of a nontoxic end product (Orth and Gillham, 1996; Wang and Zhang, 1997;
Liu et al., 2005; Li et al., 2006; Carroll et al., 2013).
Metallic or ZVI is a moderate reducing agent that can react with dissolved oxy-
gen and to some extent with water. This corrosion reaction can be accelerated or
inhibited depending on solution chemistry and this is put into productive use in
the treatment of hazardous and toxic materials as per the discovery of Gillham and
coworkers (Gillham and OHannesin, 1994; Orth and Gillham, 1996;OHannesin
and Gillham, 1998).
The practical applicability of ZVI lies in the fact that it can easily get oxidized to
C2 and C3 oxidation states and in the process reduce other organic as well as inor-
ganic impurities. Metallic iron easily acts as an electron donor:
Fe
0
! Fe
2 C
C 2e
¡
(1)
These electrons are in turn accepted by chlorinated hydrocarbons undergoing
reductive dechlorination (Vogel et al., 1987; Matheson and Tratnyek, 1994).
RCl C H
C
C 2e
¡
! RH C Cl
¡
(2)
From thermodynamic point of view coupling of reactions 1 and 2 are energeti-
cally favorable as the standard reduction potential (E
0
) of ZVI (Fe
2C
/Fe) is
0.44 V, which is lower than many chlorinated organic and other hazardous metals
such as Pb, Cd, Ni, and Cr, thus are prone to reduction by ZVI (Li et al., 2006).
Iron has the ability to reduce and adsorb redox sensitive elements such as chro-
mium as well as dehalogenate the halogenated organic compounds as demon-
strated in laboratory as well as eld studies at ambient temperatures (Liang et al.,
1996; Puls et al., 1999; Deng and Hu, 2001; Sinha and Bose, 2006; Kim et al., 2007;
Ludwig et al., 2007).
444 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Nanoscale iron particles are in a process to replace microiron particles and have
proven to be quite effective reductant and catalyst for a wide variety of common
environmental contaminants including chlorinated organic compounds and metal
ions (Lien and Zhang, 2001). Halogenated hydrocarbons such as chlorinated meth-
ane, ethane, ethylene, and benzene can be reduced to benign hydrocarbons by
nanoiron particles (Zhang, 2003) in comparison to the production of toxic daugh-
ter products such as vinyl chloride by microiron particles (Arnold and Roberts,
2000). A number of eld trials to measure the effectiveness of iron-based nanopar-
ticles have been developed (Elliott and Zhang, 2001; Zhang, 2003; Interstate Tech-
nology and Regulatory Council, 2005; Quinn et al., 2005). The nanoparticles are
injected as slurry into the subsurface environment, to remediate contaminated
groundwater plumes or source zones, avoiding trenching methods. The particles
are suspended in a hydrophobic uid, as an emulsion, to prevent particle agglom-
eration and enhance reactivity and mobility. OHara et al. (2006) and Quinn et al.
(2005) reported substantial reductions of trichloroethylene (TCE) in soil (greater
than 80% reduction) and groundwater (60% to 100% reduction) during a eld-
scale demonstration at Cape Canaveral Air Force Station (Florida), by injecting
emulsied nZVI particles. nZVI particles have also been used for the removal and
stabilization of metals or metalloid contaminants in groundwater such as arsenic
(Kanel et al., 2005, 2006) and removal of Cr(VI) from wastewater (Hu et al., 2004).
Here we review the synthesis, characterization, and applications of nZVI for reme-
diation of contaminated groundwater.
2. Synthesis of nZVI particles
Synthesis of iron nanoparticles can be done using two approaches: top-down and
bottom-up. In former approach large size materials are converted to nZVI with
the aid of mechanical and chemical processes such as milling, etching, and/or
machining (U.S. Environmental Protection Agency, 2010; Crane and Scott, 2012).
The latter approach is based on the growth of nanostructures atom-by-atom or
molecule-by-molecule via chemical synthesis, self-assembling, positional assem-
bling, and so on (Li et al., 2006). Three distinct synthetic schemes were utilized at
Lehigh University to prepare the nZVI. All involved the reduction and precipita-
tion of ZVI from aqueous iron salts using sodium borohydride as the reductant.
2.1 Synthesis of nZVI using ferric chloride (Type I nZVI)
In this synthesis, 0.25 M sodium borohydride is slowly added to 0.045 M ferric
chloride hexa-hydrate, in aqueous solution, in vigorously mixed conditions; such
that the volumes of both the borohydride and ferric salt solutions are approxi-
mately equal (i.e., 1:1 v/v). The mixing time generally kept is 1 hr. The resulting
nano-iron particles were then washed successively with a large excess of distilled
water, typically > 100 mL g
¡1
. The solid nanoparticle mass was recovered by vac-
uum ltration and washed with ethanol. If bimetallic particles were desired, the
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 445
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
ethanol-wet nZVI mass is soaked in an ethanol solution containing approximately
1% palladium acetate. Figure 1a shows the SEM image of Type I nZVI particles
(Lien et al., 2006).
2.2 Synthesis of nZVI using ferrous sulfate (Type II and III nZVI)
This method was developed because of two concerns. First, ferric chloride salt is
highly acidic and hygroscopic, so safety was a major issue. Another concern was
the potential deleterious effects of excessive chloride levels from the nZVI matrix
in batch degradation tests where chlorinated hydrocarbons are the contaminant of
concern. Type II nZVI was prepared by pouring equal volumes of 0.50 M sodium
borohydride at 0.15 l min
¡1
into 0.28 M ferrous sulfate. The nished nanoparticles
were washed with distilled water (>100 mL g
¡1
), then by ethanol, purged with
nitrogen, and refrigerated in a sealed polyethylene container. The Type III nZVI
was also synthesized using the sulfate method. It was the latest generation of nZVI
with the average particle size of 5070 nm. The moisture content of Type III nZVI
was lower (i.e., 2030% in comparison to 4060% for Types I and II; Lien et al.,
2006).
2.3 Synthesis using top-down approach
In recent years, Golder Associates Inc. has emerged as a market leader for large-
scale eld deployment of nZVI. They produce nZVI in large quantities by top-
down approach by means of mechanical grinding of macroscale iron in planetary
ball mill systems. Though the method of production is very simple, the nanopar-
ticles so produced exhibit a very high surface energy and are thus prone to aggrega-
tion (Crane and Scott, 2012).
3. Structure of NZVI particles
The nZVI, produced by bottom-up approa ch, exhibit a typical core shell struc-
ture where the core consists of zero valent or metallic iron and a mixed valent
Figure 1. (a) SEM image of type I nZVI, (b) TEM of type II iron nanoparticles, and (c) TEM image of type
III iron nanoparticles (Lien et al., 2006). The scale bar represents 0.1 mm in (b) and (c) respectively.
© Sustainable Environment Research (SER). Reproduced by permission of Sustainable Environment
Research (SER). Permission to reuse must be obtained from the rightsholder.
446 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
oxide shell of Fe(II) and Fe(III) is formed as a result of oxidation of the core
shell (Fig ure 2). Thus this ZVI is actually a manufactured material not found
naturally such as Fe(II) and Fe(III). nZVI is fairly reactive in water and pos-
sess excellent elec tron donating properties, which makes it a versatile remedia-
tion material (Stumm and Morgan, 1996). It is the core that provides the
reducing power (the electron source) for the reactions whereas the she ll pro-
vides the site for chemical complex reactions (chemisorptions ) and electro-
static interactions. The core shell structure has important implications on the
chemical proper ties of nZVI. The defective and disordered nature of the oxide
shell renders it potentially more reactive compared to a simple passive oxide
layer formed on bulk iron material (Wang et al., 2009). The shell exhibits less
contrast compared to the dense interior core. When the nanoparticles agglom-
erate, they have a continuous oxide shell but the m etallic c ores are separated
by a thinner interfacial oxide layer (»1 nm). The oxid e layer is amorphous
and disorder ed, owing to the e xtremely small radii of the nanoparticles, which
hinders crystalline formation. Moreover the presence o f boron as precursor
also contribute to defective sites and the oxide layer (Carpenter et al., 2003;
Ponder et al., 2001).Theoxidelayerhassemiconductorpropertiesaswell
(Balko and Tra tnyek, 1998;Wang,2009) and so charge transfer is relatively
facile due to small thickness and pre sence of defecti ve sites and allow re duc-
tion of contaminants to occur. Apart from reductive dehalogenation of organic
compounds and inorganic contaminants, nZVI is capable of b eing applied to a
broader spectrum of contaminants amenable to red uction, surfa ce sorp tion,
precipitation, or indeed combinations of these (Yan et al., 2010). According to
thecoreshellmodel,themixedvalenceoxide shell is large ly insoluble under
neutral pH and thereby protects the core from rapid oxidation. Composition
Figure 2. The core-shell model of nZVI and schematic representations of the reaction mechanisms
for the removal of Hg(II), Ni(II), Zn(II), and H
2
S (Yan et al., 2010).
© Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 447
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
of oxide shell depends on the fabrication process and environmental condi-
tions (Li et al., 2006).
4. Characterization of NZVI
Morphological analysis of the iron nanoparticles prepared with the method of fer-
ric iron reduction by sodium borohydride have been done with the help of trans-
mission electron microscopy (TEM) and scanning electron microscopy (SEM; Sun
et al., 2006; Lin et al., 2010)(Figure 3). Preparation of the sample is done by
depositing 23 drops of dilute ethanol solution of the sample onto a carbon lm
supported on a 300-mesh copper grid. This was followed by placing the sample in
a vacuum hood till ethanol evaporated completely (Sun et al., 2006; Yan et al.,
2010). The results obtained from TEM analysis has shown that the nanoparticles
appear agglomerated forming chain-like formations due to magnetic and electro-
static interactions. It can be clearly inferred from the image that a single particle
comprises a dense core surrounded by a thin shell exhibiting markedly less con-
trast than the interior core (Yan et al., 2010).
Figure 3. Micrographs of (a) a single particle and (bd) aggregates of iron particles (Sun et al., 2006).
© Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.
448 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
The particle size and size distribution of nZVI is measured with an Acoustic Spec-
trometer that utilizes the sound pulses transmitted through a particle suspension to
measure the properties of suspended particles (Dukhin a nd Goetz, 2002). The instru-
ment is connected to a four-necked conical ask containing a suspension of iron nano-
particles in deionized water. The advantage of working with this instrument is that it is
exible with no limitation on opaque materials (Dukhin and Goetz, 2002;Morrison
and Ross, 2002). The specic surface area of the nanoparticles is determined with the
classic BET method (the BrunauerEmmettTeller isotherm). The BET isotherm is the
basis for determining the extent of nitrogen adsorption on a given surface. Sample prep-
aration is done by predrying at room temperature (22 § 1
C) in a vacuum desiccator,
and degassing at 90
Cfor1handthenat200
Cfor4h(Sunetal.,2006;Hwangetal.,
2011). X-ray diffraction is used to investigate the material structure of nanoparticles
(Sun et al., 2006). It can be inferred from the diffraction pattern that the particles pres-
ent were mainly in the zero valent state and all iron are in closed phase cubic structure
(Chatterjee et al., 2010). The surface composition for iron nanoparticle can be deter-
mined by X-ray photoelectron spectroscopy (XPS) up to a depth of less than 10 nm.
The iron nanoparticle samples are prepared by drying in a small nitrogen-purged hood
at room temperature and packed into the sample cell directly, to prevent oxidation of
samples. X-ray absorption near edge structure (XANES) uses synchrotron radiation to
photoionize the core electrons of Fe atom and provides useful information on the
valenceofiron(Sunetal.,2006). The surface charge of iron nanoparticles is often char-
acterized by the zeta (z) potential, which is dened as the electric potential at the surface
of shear relative to that in the distant bulk medium. It is the major factor determining
the mobility of particles in an electrical eld (Sun et al., 2006).
A detailed knowledge of the surface properties is essential for understanding the
reaction mechanisms, kinetics, and intermediate/product proles. The transport,
distribution, and fate of nanoparticles in the environment also depend on these
surface properties. The average particle size of the chloride method nZVI varies
between 50 and 200 nm whereas for sulfate method nZVI size varies between 50
and 70 nm (Lien et al., 2006). The average BET surface area of both type of nZVI
is in the range of 35 § 3m
2
/g (Lien et al., 2006).
5. Mobility and transport of nZVI in subsurface
The mobility of nanomaterials (NMs) in porous media is inuenced by their ability
to attach to mineral surfaces to form aggregates. NMs that readily attach to the
mineral surfaces may be less mobile in ground water aquifers (Wiesner et al.,
2006); smaller NMs that can t into the interlayer spaces between soil particles
may travel longer distances before becoming trapped in the soil matrix and soils
with high clay content tend to stabilize NMs and allow greater dispersal (U.S. Envi-
ronmental Protection Agency, 2008a). The surface chemistry and therefore the
mobility of NMs in porous media may be affected through the addition of surface
coatings. Although nZVI is widely used in site remediation, information is limited
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 449
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
on its fate and transport in the environment. While increased mobility due to the
smaller size may allow for efcient remediation, there are insufcient data regard-
ing whether such NMs could migrate beyond the contaminated plume area and
persist in drinking water aquifers or surface water (U.S. Environmental Protection
Agency, 2008a). Issues affecting the fate and transport includes the contaminant
concentration, the nZVI synthesis process, particle agglomeration, age of the nZVI
particles, improper handling during application, particle density, the soil matrix,
ionic strength of the groundwater, hydraulic properties of the aquifer, depth to the
water table, presence of organic matter, and other geochemical properties, such as
pH, dissolved oxygen (DO), oxidation reduction potential, and concentrations of
competing oxidants (U.S. Environmental Protection Agency, 2008b). Aggregation
of bare nZVI particles decreases the surface area of the nZVI, which decreases the
mobility and reactivity, thereby limiting the radius of inuence (He et al., 2007).
Several conditions causing the nZVI particles to agglomerate, includes the nZVI
particle concentration, the magnetism of the particles, size distribution, and zeta
(z) potential (Phenrat et al., 2007; Phenrat et al., 2009). Application of the nZVI
particles at too high of a concentration may increase the chances of particle
agglomeration (Saleh et al., 2007). Recent research shows that particles having con-
centrations of 30 mg/l or less are mobile regardless of the size distribution and
magnetic forces (Phenrat et al., 2009). Larger particles with higher ZVI content are
more magnetically attracted to each other and soil grains, which may also increase
the nZVI deposition potential. Smaller particles with low ZVI content travel far-
ther than those with high nZVI content because they are less likely to agglomerate
(Phenrat et al., 2009). Another condition that can cause agglomeration is the
z -potential of the particles. Particles with z-potential values greater than C30 mV
and less than 30 mV are considered stable, with maximum instability, or agglom-
eration, taking place at zero (Zhang and Elliott, 2006). Field-scale studies have
revealed 57-ft radius of inuence of nZVI particles (ESTCP, 2010) and they can
travel up to a distance of 20 m in groundwater and can remain reactive for period
of 48 weeks (Lien et al., 2006).
5.1 Modication of nZVI
Magnetic attraction between nanoiron particles causes the rapid aggregation of
particles (Phenrat et al., 2007) so various organic coatings are used nowadays, such
as emulsions, polymers, and polyelectrolytes, that limit reactivity and increase the
mobility of nanoiron in the subsurface (Ponder et al., 2000). Bare and pure nZVI
are more prone to react with dissolved oxygen and oxygen-rich compounds so, to
prevent it from non target oxidation, it is essential to coat with surfactants or poly-
electrolytes (Krajangpan et al., 2008). Particles enhanced with certain polyelectro-
lytes have proven to be more mobile than bare nZVI even after aging. Particles
coated with certain polyelectrolytes (e.g., polyasarate, carboxymethyl cellulose
[CMC], polystyrene sulfonate) can remain mobile for at least eight months after
450 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
the original injection depending on the hydrochemistry and geochemistry found at
a site (Kim et al., 2009). Selvarani and Prema (2012) used starch stabilized nano-
iron particles for Cr (VI) removal for better reactivity and mobility. The stabiliza-
tion of the nZVI particles can be enhanced by using various surface coatings. Some
of them are the following:
Hydrophilic biopolymers such as starch, guar gum, alginate, and aspartame
(He and Zhao, 2005; Tiraferri et al., 2008; Saleh et al., 2008; Tiraferri and
Sethi, 2009; Bezbaruah et al., 2009)
CMC (He et al., 2007)
Chitosan (Zhu et al., 2006; Geng et al., 2009)
Natural organic matter such as humic acid (Xie and Shang, 2005; Zhu et al.,
2008)
Polyelectrolytes such as polyacrylic acid, ion-exchange resins, and block
copolymers (Kanel and Choi, 2007; Zhao et al., 2008; Sirk et al., 2009)
Amphiphiles including various surfactants, which can be anionic, cationic, or
non-ionic (Hydutsky et al., 2007; Kanel et al., 2007; Zhu et al., 2008) and
block copolymers (Saleh et al., 2005; Saleh et al., 2007; Saleh et al., 2008)
Various oil-based microemulsions (Quinn et al., 2005)
Polyvinyl alcohol-co-vinyl acetate-co-itaconic acid, used during the synthesis of
the nZVI, leads to the formation of stable nZVI dispersion, which is stable for over
six months (Sun et al., 2007). Wu et al. (2005) used cellulose acetate to coat nZVI.
Effectiveness of various biodegradable dispersants was tested for nanoiron stability.
PAA (poly acrylic acid) was found to be best among various dispersants (Yang
et al., 2007). Allabaksh et al. (2010) synthesized stable nZVI using different chelat-
ing agents. They used ethylenediaminetetraacetic acid (EDTA), diethylenetriamine
pentacetic acid (DTPA), nitriloacetic acid (NTA), trans-1,2-diaminocyclohexane-
N,N,N,
0
N
0
-tetraacetic acid (CDTA), hydroxyethylenediaminetetraacetic acid
(HEDTA), triethylene tetraamine (TRTA), and N-cetyl-N,N,N-trimethyl ammo-
nium bromide (CTAB) chelating agents. The chelating effect was found to be the
best for EDTA, NTA and HEDTA, but the least for CDTA and CTAB. Another
study carried out by Krajangpan et al. (2008) was a landmark in the eld of pro-
duction of stabilized iron nanoparticles. They synthesized a series of amphiphilic
polysiloxane graft copolymers (APGCs). Treatment of nZVI with APGCs was
assessed. The APGC possessing the highest concentration of carboxylic acid
anchoring groups provided the highest colloidal stability.
The mobility of nZVI is inuenced by the chemistry of the aqueous phase (e.g.,
ionic strength, pH, natural organic matter [NOM]; Bian et al., 2011). Research
shows that the NOM enhances the mobility of nZVI. Even at lower concentration
of natural organic matter (2 mg/l), the mobility of nanoiron particles was higher
than no-NOM conditions. The sorption of natural organic matter on nZVI causes
reduction in sticking coefcient and hence accelerating the mobility (Johnson
et al., 2009). Also it has been found that bactericidal activity of nZVI toward gram-
negative Escherichia coli and gram-positive Bacillus subtilis can be mitigated in the
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 451
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
presence of NOM (Chen et al., 2011). The pre-existence of NOM in transport
medium apparently has scavenging effect on attachment (i.e., sticking) of nZVI
aggregates to the immobile surface of transport medium (Giasuddin et al., 2007;
Zhang et al., 2011). Organic matter (humic acid [HA]) also has negative effects on
nZVI reactivity, as it introduces morphological changes to carbonate green rust
(Hwang and Shin, 2013). Dong and Lo (2013) investigated the inuence of humic
acid (HA) on the surface-modied nZVI (SM-nZVI) and found HA to have pro-
found implication for nZVI stability. HA reacts with the SM-nZVI (PAA modied)
and complexes with SM-nZVI to enhance the electrostatic repulsion effect (i.e., it
increased the stability of the particles). On the other hand HA forms interparticle
bridging linkage with Tween-20 and starch modied nZVI, which induces agglom-
eration of nZVI. This suggests that the presence of NOM and HA in ground water
should be taken into consideration in choosing the proper SM-nZVI for particular
application (Dong & Lo, 2013).
6. Application of nZVI for remediation of organic compounds
Use of nZVI particles in advanced oxidation processes (AOPs) such as the Fenton
Process have shown an advantage over conventional method which requires
around 4080 ppm of ferrous ion in the solution and this value is above the stand-
ards. In addition, the application of the homogeneous AOPs to large water ow
rate may produce large amount of sludge in the nal neutralization step (Iurascu
et al., 2009). To avoid these disadvantages recently nZVI was used to degrade wide
range of organics at much lower operational cost (Pradhan and Gogate, 2010). In a
recent study the performance of nZVI in different AOPs such as F (Fenton),
electro Fenton (EF), and photo electro Fenton (PEF) processes on the degradation
of phenol was studied. The optimum dosage of nZVI and H
2
O
2
were found to be
0.5 g/l and 400 mg/l, respectively, and the complete removal of phenol was
observed within 30 min at these conditions at an initial pH and initial phenol con-
centration of 6.2 and 200 mg/l, respectively, and the removal process followed
pseudo rst-order kinetics (Babuponnusami and Muthukumar, 2012). In an
another study the t8rinitroglycerin (TNG) C
3
H
5
(ONO
2
)
3
, a versatile chemical
widely used in the manufacture of dynamite, was degraded by using nZVI sup-
ported on the surface of an inert polymer such as nanostructured silica SBA-15
(Santa Barbara Amorphous No. 15). Both nZVI and nZVI/SBA-15 degraded TNG
(100%) in water to eventually produce glycerol and ammonium. The reaction fol-
lowed pseudo rst-order kinetics and was faster with nZVI/SBA-15 (k
1
0.83
min
¡1
) than with ZVINs (k
1
0.228 min
¡1
; Saad, 2010).
6.1 Remediation of halogenated organic compounds
Chlorinated nitroaromatic compounds are widespread in the environment due to
the manufacturing and processing of a variety of industrial products, such as phar-
maceuticals, explosives, dyes, and pesticides (Lin et al., 2010; Liu et al., 2011) They
452 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
are of interest to environmental scientists as they are known or suspected to be
human carcinogens, mutagens or toxins (Lin et al., 2010; Shen et al., 2008). nZVI
and palladium-doped nZVI immobilized on alginate bed were shown to degrade
TCE. The removal efciency of TCE was > 99.8% when 50 g/l of Fe/Pd-alginate
(3.7 g Fe/l) was introduced to the aqueous solution. The removal of TCE by Fe/Pd-
alginate followed pseudo rst-order kinetics and the major nal products were
identied as ethane and butane. At low TCE concentration (4.4 mg/l) and excess
nZVI (»1.9 g/l), nal reaction products were mainly composed of 80% of ethane
and »20% of C
3
C
6
molecules. When limited amount of nZVI (»0.36 g/l) was
added to high TCE concentration (290 mg/l), even-numbered saturated alkanes
(C
2
,C
4
,C
6
) and other saturated and unsaturated hydrocarbons were detected
(Hojeong, 2010). Although vinyl chloride and cis-DCE were detected as intermedi-
ates, during reaction of nZVI with TCE, they were at a very low level and disap-
peared rapidly (Liu et al 2005). Following the model for TCE dechlorination by
nZVI presented by Liu et al. (2007) and a model for tetrachloroethylene (PCE)
dechlorination by surface modied nZVI (MRNIP2) presented by Kim (2009)it
was assumed that the reaction pathways could be described by reductive beta elim-
ination of PCE to acetylene followed by transformation of acetylene to ethene and
ethane. The use of nZVI to remediate dichlorophenyltrichloroethane (DDT) from
contaminated water has been studied by Poursaberi et al. (2012). Mixing an aque-
ous solution of 3 mg l
¡1
DDT with 30 mg l
¡1
Fe
0
resulted in 99.2% loss of DDT
within 4 hr. GC/MS analysis conrmed the formation of completely dechlorinated
hydrocarbon skeleton of DDT, namely diphenylethane (DPE), as the end product
in this treatment; thereby implying the removal of all ve chlorine atoms (three
alkyl and two aryl) of DDT. It was concluded that Nanoscale Fe
0
was successful at
dechlorinating DDT, 1,1-dichloro-2,2-bis(4-chlorophenyl)ethane (DDD) and 1-
chloro-2,2-bis (p-chlorophenyl)-ethane (DDMS). The use of nZVI with hydrogen
peroxide to treat pentachlorophenol contaminated soil offered new and effective
solution for treatment. By adding 5% of calcium carbonate for 40 h, the decay rate
of pentachlorophenol in Chengchun (Cf) series soil increased from 37% to 81%
and that in Pianchen (Pc) series increased from 41% to 75%. This treatment, add-
ing calcium carbonate in pentachlorophenol contaminated soil, was a new and rev-
olutionary method and served as the reference for on-site application (Liao et al.,
2007). Dehalogentaion of hexachlorobenzene (HCB) and polychlorinatedbiphenyl
(PCB) by nZVI showed sequential dehalogenation trend (Shih et al., 2011; Lowry
and Johnson, 2004). The end product of HCB reductive dehalogenation was 1,2-
DCB. In another study the reductive dechlorination of TCE sorbed in two model
soils (a potting soil and Smith Farm soil) using CMC-stabilized Fe-Pd bimetallic
nanoparticles was reported. The nanoparticles could effectively degrade TCE but
the soil organic matter (around 8.2 %) in the potting soil inhibited the degradation
kinetics. Four prototype surfactants were tested for their effects on TCE desorption
and degradation rates, including two anionic surfactants known as SDS (sodium
dodecyl sulfate) and SDBS (sodium dodecyl benzene sulfonate), a cationic
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 453
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
surfactant hexadecyltrimethylammonium (HDTMA) bromide, and a nonionic sur-
factant Tween-80. All four surfactants were observed to enhance TCE desorption
at concentrations below or above the critical micelle concentration with the
anionic surfactant SDS being most effective (Zhang et al., 2011).
6.2 Remediation of pharmaceuticals
The interactions of tetracycline (TC) with nZVI modied by polyvinylpyrrolidone
(PVP-K30) was investigated using batch experiments as a function of reactant con-
centration, pH, temperature, and competitive anions. Degradation of TC was
strongly dependent on pH and temperature. The presence of silicate and phos-
phate strongly inhibited the removal of TC, whereas acetate and sulfate only
caused slight inhibition. LCMS analysis of the treated solution showed that the
degradation products from TC resulted from the removal of functional groups
from the TC ring. The degradation products were detected both in the treated solu-
tion and on the surface of PVP-nZVI after 4-hr interaction, indicating that PVP-
nZVI could adsorb both TC and its degradation products (Chen et al., 2011). In
another study using ZVI, complete removal of amoxicillin (AMX) and ampicillin
(AMP) upon contact with Fe
0
and nFe
0
was obtained. The removal process was
attributed to three different mechanisms: (a) a rapid rupture of the b-lactam ring
(reduction), (b) an adsorption of AMX and AMP onto iron corrosion products,
and (c) sequestration of AMX and AMP in the matrix of precipitating iron hydrox-
ides (coprecipitation with iron corrosion products; Ghauch, 2009). Metronidazole
(MNZ) solution at 80 mg l
¡1
was rapidly removed by nZVI within 5 min, at initial
solution pH 5.60 and nZVI dose of 0.1 g l
¡1
. The removal process of MNZ fol-
lowed a pseudo-rst order kinetics model. The removal rate in the nZVI/air pro-
cess was slightly higher than that in the nZVI/N
2
process emphasizing that oxygen
was a key factor for the removal of MNZ. MNZ removal efciency by nZVI was
about 49 times higher than that by commercial iron powder under the same dos-
ages. The MNZ removal followed two processes: adsorption and degradation
(Fang et al., 2011). The pickling waste liquor discharged from steel industry was
utilized in preparation of nano zero valent metal (nZVM). nZVM was used to
degrade antibiotic metronidazole. The surface areanormalized rate coefcient
(k
SA
) for nZVM (0.254 l min
¡1
m
¡2
) was 375.2 times larger than that for commer-
cial iron powder (6.67£10¡4 l min
¡1
m
¡2
; Fang et al., 2010).
6.3 Remediation of azo dyes
Modern azo dyes are very resistant in natural environment; even their biological
degradation is not very easy. Zero valent iron is known to reduce the azo bond
and their products can be treated biologically. Fan et al. (2009) found that rapid
decolorization of azo dye methyl orange occurred by nZVI particles. Decoloriza-
tion was found to be very rapid, being completed in only 10 min. The products
were analyzed by GC-MS and they were sulfanilic acid, N,N-dimethyl-p-
454 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
phenylenediamine and N-methyl-p-phenylenediamine. Satapanajaru et al. (2011)
studied decolorization of Reactive Black 5 (RB5) and Reactive Red 198 (RR198)
using nZVI technology. Lower pH, pitting corrosion and bimetallic system (add-
ing Pd and TiO
2
/light on the surface of n ZVI) enhanced the reduction efciency.
The removal efciency of both RB5 and RR198 was found to be greater using n
ZVI supported on sand, silica and biological sludge. Cation exchange resin sup-
ported nZVI was used by Shu et al. (2010) for decoloration of Acid Blue 113 azo
dye solution. The removal efciency of AB 113 concentration and TOC was almost
100% and 12.6% within 30 min with 50.8 mg g
¡1
nZVI load, 20 gl
¡1
of resin load
and initial dye concentration of 100 mg l
¡1
. The color removal efciency followed
modied pseudo rst-order reaction. Lin et al. (2008) studied effective removal of
AB24 dye and factors affecting the rate of removal using nZVI. The reaction fol-
lowed pseudo rst-order kinetics. They showed that at pH < 6, the removal of dye
was mainly due to reduction while at pH > 6 the adsorption phenomena domi-
nates. The orange (II) decolorization by n ZVI supported on bentonite was studied
by Xi et al. (2011). The Fe0
¡
clay material showed better reactivity toward the dye
than pure nZVI. In both the cases, the amount of dye reduced was comparable
that is, initial concentration of 96.5 mg l
¡1
of orange (II) reduced to 3 mg/l. Zhao
et al. (2008) found the rapid decolorization of water-soluble azo dye by nZVI
immobilized on cation exchange resin.
7. Remediation of inorganic contaminants
The adsorption of Ni
2C
by iron nanoparticles was studied by Nazlı (2008).
The n ZVI had shown high potential toward uptake of N i
2C
ion. Li and
Zhang (2006) also studied Ni (II) sequestration us ing iron nanoparticles. They
suggested that nZVI particles act as sorbent as well as reductant for Ni (II)
removal from water. In the reactor having 100 mg/l of Ni(II) and nZVI con-
centration of 5 g /l, the complete removal was achieved within 3 h. NZVI has
also been used in phosphate removal. Batch study conducted by Almeelbi and
Bezbaruah (2012)conrmed that when different concentrations of phosphates
(1,5,and10mgPO
4
3¡
-P/l), reacted with 400 mg nZVI/l, 8895% phosphate
was removed in rst 10 min, and 96100% removal was ac hieved after
30 min. The maximum phosphate recovery of »78% was found to be at pH
12 within 30 min. The mechanism of phosphate remova l was determined to
be simultaneous adsorption and chemical precipitation (Wu et al., 2013;Wen
et al., 2014). The XRD patterns of the nZVI after the reactions indicated the
formation of crystalline vivianite (Fe
3
(PO
4
)
2
8H
2
O; Wu et al., 2 013). The
adsorption of PO
4
3¡
on nZVI followed both Langmuir and Freundlich model
and the maximum adsorption capacity was 245.65 mg/g (Wen et al., 2014).
Wu et al. (2013) and Wen et al . (2014) reported that the removal of phos-
phate decreased with an increasing pH due to the isoelectric point (IEP) of
nZVI, which contradicts the results by Almeelbi and Bezbaruah (2012). The
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 455
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
phosphate removal efciency increased with increasing dosage of nanoscale
iron particles, while it decreased with increasing initial phosphate concentra-
tion (Wu et al., 2013 ). The nZVI synthesized in a CMCwater solution
enhanced phosphate removal compared to nZVI synthesized in an ethanol-
water solution (Wu et al., 2013). The phosphate removal efciency by nZVI
was considerably higher than micro ZVI (Almeelbi and Bezb aruah, 2012 ;Wu
et al., 2013). Hence, nZVI may be successfully used for remediation of eutro-
phic waters. Uzum et al. (2008) used nZVI particles with size 2080 nm for
fast uptake of aqueous Co
2C
ions. This uptake wa s directly proportional to
increase in pH. Uzum et al. (2009) sho wed tha t kaolinite red uced th e aggre-
gartion of nZVI and showed higher uptake of Cu
2C
compared to Co
2C
.The
main mechanism for removal of Co
2C
was suggested to be adsorption and
precipitation while for Cu
2C
, it w as just a redox mecha nism. nZVI has also
been used as an adsorbent for removing cadmium from water ( Boparai et al.,
2011). The adsorption kinetic s followed pseudo second- order model and the
process can be better described by Langmuir isotherm compared to Freundlich
isotherm. nZVI are used for removal of both arsenic (III) and a rsenic (V)
from contaminated water (Kanel et al., 2005; Konstantina et al., 2007;Bang
et al., 2 005). 99.9% of arsenic (III) was found to be removed within 10 m in
by nZVI dosage of 1 g/l and pH 7. The efciency of removal decreased with
increase in pH and arsenic concentration. In another study performed by Tan-
boonchuy et al. (2011) conclu ded that the favorab le cond ition for Arsenic (V)
removal can be acidic condition and high dissolved oxygen. Zhu et al. (2009)
studied removal of arsenate and arsenite by adsorption to nZVI supported on
activated carbon. The common m etal cation such as Ca
2C
,Mg
2C
was found to
increase arsenate adsorption and Fe
2C
lowering the arsenite adsorption. The
removal rate of both arsenate and arsenite was r eported to decrease in the
presence of phosphate and silicate while the effect of humic acid was almost
negligible. Morgada et al. (2009) studied the effect of UV light and humic
acid on arsenic (V) removal using nZVI. The presence of humic acid
decreased the removal rate by 50% while UV doubled the removal rate.
Though, humic acid improved the reaction rate in the presence of light.
Nitrate removal from water was studied by Kassaee et al. (2011) using iron
nanoparticles produced by arc discharge and reduction method. The nanopar-
ticles fabricated by arc discharge method are l arger and pure, have high er dis-
persity, and have higher efciency in comparison with the nanoparticles
synthesized by reduction method. Low pH and high ionic strength of the solu-
tion increased nitrate removal. In another study performed by Hsu et al.
(2011) revealed that the best nitrate remova l by nZVI was achieved at around
neutral pH. At initia l pH of 10, the end product of reduction was mainly
dominated by ammonium ions but at initial pH of 6 and 8, other nitrogen
containing species are also formed besides ammonium and nitrite. Ziajahromi
et al. (2012) reported that the r eduction of nitrate b y nZVI inc reased with
456 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
dosage of nZVI and the rate of reduction decreased with increase in pH. nZVI
supported by polystyrene resins reduce the particle size and hence increases
the reduction of nit rate by nZVI (Jiang et al., 2011 ). Reduction of nitrate by
nZVI prepared from hydr ogen reduction of natural goethite (nZVI-N), hydro-
thermal goethite (nZVI-H) and ordinary ZVI (OZVI) was studied by Liu
et al. (2012). The reactivity toward nitrate removal followed the trend of
nZVI-N> nZVI-H> OZVI. Ryu et al. (2011) studied the reduction of high
concentration of nitrate and concluded that aggregation and presence of cata-
lyst both affect the reduction rate but aggregation effect is more pronounced
than catalytic effect. Decreasing the precursor (FeCl
3
.6H
2
O) concentr ation,
during the manufacture of nZVI, reduced the particle size and enhanced the
reactivity of nZVI toward nitrat e r eduction (Liou et al., 2006). The nZVI-
mediated microbial reduction was found to b e more effective for nitrate
removal as nZVI acts as source of electron for biological nitrogen reduction
(Shin and Cha, 2008). Olegario et al. (2010) studied the reduction of soluble
selenium (VI) to insoluble selenium (-II) by nano iron sus pensions, at a reac-
tion r ate four times higher than Fe
0
powder. Uranium (VI) was shown to
reduce to U(IV) by iron nanoparticles, for 48 hrs, af ter which it reoxidized to
U(VI), as Fe(II) oxidized to Fe(III) (Dickinson and Scott, 2010). Treatment of
acid mine drainage (AMD) with nZVI lead to decrease in concentration of
various contaminants such as Al, U, V, Cr, Cu, Ni, Cd, Zn, and As) due to
precipitation of reduced cations and copre cipitation w ith ir on oxy-hydroxides
(Klimkova et al., 2011 ). Nanoparticles of iron are able to reduce dimethyl sul-
de at neutral and acidic pH to produce methane and iron sulde. The reac-
tion does not occur at basic pH (Calderon et al., 20 12). Removal of Pb(II)
from aqueous solution by using PAA stabilized iron nanoparticles (Esfahani
et al., 20 14), kaolin-supporte d nZVI (Zhang et al., 2010), and zeolite-nZVI
composite (Kim et al., 2013) was also studied. Singh et al. (2011) used v aried
doses of nZVI on Cr(VI) containing soil samples for remediation of Cr(VI).
The Cr (VI) removal efciency increased with decreasing initial pH and the
removal efciency was four ti mes higher compared to ir on powder (Nhung
and Thuong, 2008). nZVIFe
3
O
4
nanocomposites, prepared by an in situ
reduction method, were successfully used fo r chromium (VI) removal in aque-
ous environment (Lv et al., 2012). Bentonite-supported nZVI decreased nZVI
aggregation and thus increased Cr (VI) removal efciency (Shi et al., 2 011).
CMC-stabilized iron nanoparticles (CMC-Fe
0
) resulted in 100% degradation of
Cr(VI) in comparison to 50% reduction pr oceeded with unstabilize d Fe
0
nano-
particles (Madhavi et al., 201 2).
8. Environmental impacts of nZVI
In addition to concerns regarding the reactivity and mobility of nZVI there is little
literature regarding the risks that this technology may pose on human and
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 457
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
ecological health (Tratnyek and Johnson, 2006). Studies have demonstrated that
the toxicity effects of nZVI are limited compared to other nanoparticles (Reijnders,
2006). The ecological impacts of nanoparticles on the environment can be summa-
rized as described subsequently.
8.1 Toxic effect to mammalian nerve cells
Different forms of nZVI (i.e., fresh, aged, and surface modied) are differentially
toxic to the rodent nerve cells (Phenrat et al., 2009). Fresh nZVI produced mor-
phological evidence of mitochondrial swelling and apoptosis. The results revealed
that partial or complete oxidation of nZVI reduces its redox activity, agglomera-
tion, sedimentation rate, and toxicity to mammalian cells. Surface modication of
nZVI also reduces its toxicity. In the presence of a polymeric coating, toxicity
effects are very limited or even absent (Li et al., 2010). All forms of nZVI aggre-
gated in soil and water in presence of a high concentration of calcium ions and
thus addition of calcium salts may help in reducing the toxicity of groundwater
due to nZVI (Keller et al., 2012).
8.2 Toxic effect to aquatic life
The effects of nZVI on medaka sh (Oryzias latipes) and their embryos were inves-
tigated recently. A dose- and time-dependent decrease in superoxide dismutase
(SOD) and malondialdehyde (MDA) activities was observed in the embryos. A sig-
nicant decrease of SOD and glutathione (GSH) activity was observed in liver and
brain samples taken from the adults, but as the exposure time increased, the adults
appeared to recover from the exposure by adjusting the levels of antioxidant
enzymes (Li et al., 2009).
8.3 Toxic effect to microorganisms
nZVI is found to be capable of removing viruses (e.g., ; X174 and MS-2) from
water by inactivating them and/or irreversibly adsorbing the viruses to the iron
(You et al., 2005). A study done by Lee et al. (2008) says that nZVI particles exhib-
ited a bactericidal effect on Escherichia coli that was not observed with other types
of iron-based compounds, such as iron oxide nanoparticles, microscale ZVI, and
Fe
3C
ions. nZVI coated with humic acid showed minor toxicity to E. coli (Li et al.,
2010).
9. Conclusions
The insight into the different methods of synthesis of nZVI, its structure and char-
acterization and its transport phenomenon establishes its role in the environmental
remediation. Aggregation of nZVI has been reported to be the major drawback for
its application. The modication of nZVI in order to overcome the challenges
faced in the transport of nZVI through the soil has been advocated by many
458 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
researches. With more research work focusing on developing proper techniques to
improve the nZVI motility the function of nZVI can be enhanced. Hence, the
application of nZVI as a remediation tool appears to be more promising than con-
ventional ZVI (microscale) or other in situ remediation methods. However, con-
tinued research effort toward modifying this technology is required to minimize
the unexpected adverse environmental impact or potential risks.
References
Allabaksh, M. B., Mandal, B. K., Kesarla, M. K., Kumar, K. S., and Pamanji, S. R. (2010). Prepa-
ration of stable zero valent iron nanoparticles using different chelating agents. J. Chem.
Pharm. Res. 2(5), 6774.
Almeelbi, T., and Bezbaruah, A. (2012). Aqueous phosphate removal using nanoscale zero-val-
ent iron. J. Nanopart. Res. 14, 900.
Arnold, W. A., and Roberts, A. L. (2000). Pathways and kinetics of chlorinated ethylene and
chlorinated acetylene reaction with Fe(0) particles. Environ. Sci. Technol. 34, 17941805.
Babuponnusami, A., and Muthukumar, K. (2012). Removal of phenol by heterogenous photo
electro Fenton-like process using nano-zero valent iron. Separ. Purif. Technol. 98, 130135.
Balko, B. A., and Tratnyek, P. G. (1998). Photoeffects on the reduction of carbon tetrachloride
by zero-valent iron. J. Phys. Chem. B 102, 14591465.
Bang, S., Johnson, M. D., Koratis, G. P., and Meng, X. (2005). Chemical reaction between arse-
nic and zero-valent iron. Water Res. J. 39, 763770.
Bezbaruah, A. N., Krajangpan, S., Chisholm, B. J., Khan, E., and Bermudez, J. J. (2009). Entrap-
ment of iron nanoparticles in calcium alginate beads for groundwater remediation applica-
tions. J. Hazard. Mater. 166, 13391343.
Bian, S. W., Mudunkotuwa, I. A., Rupasinghe, T., and Grassian, V. H. (2011). Aggregation and
dissolution of 4 nm ZnO nanoparticles in aqueous environments: inuence of pH, ionic
strength, size, and adsorption of humic acid. Langmuir 27, 60596068.
Boparai, H. K., Meera Joseph, M., and OCarroll, D. M. (2011). Kinetics and thermodynamics of
cadmium ion removal by adsorption onto nano zerovalent iron particles. J. Hazard. Mater.
186, 458465.
Bystrzejewska-Piotrowska,G.,Golimowski,J.,andUrban,P.L.(2009).Nanoparticles:Theirpotential
toxicity, waste and environmental management. Waste Management 29, 25872595.
Calderon, B., Aracil, I., and Fullana, A. (2012). Deodorization of a gas stream containing
dimethyl disulde with zero-valent iron nanoparticles. Chem. Eng. J. 183, 325331.
Carpenter, E. E., Calvin, S., Stroud, R. M., and Harris, V. G. (2003). Passivated iron as core-shell
nanoparticles. Chem. Mater. 15, 32453246.
Carroll,D.,Sleep,B.,Krol,M.,Boparai,H.,andKocur, C. (2013). Nanoscale zero valent iron and
bimetallic particles for contaminated site remediation. Adv. Water Resources 51, 104122.
Chatterjee, S., Lim, S.-R., Woo, S. H. (2010). Removal of reactive black 5 by zero-valent iron
modied with various surfactants. Chem. Eng. J. 160, 2732.
Chen, H., Luo, H., Lan, Y., Dong, T., Hu, B., and Wang, Y. (2011). Removal of tetracycline from
aqueous solutions using polyvinylpyrrolidone(PVP-K30) modied nanoscale zero valent
iron. J. Hazard. Mater.
192, 4453.
Chen, J., Xiu, Z., Lowry, G. V., and Alvarez, P. J. J. (2011). Effect of natural organic matter on
toxicity and reactivity of nano-scale zero-valent iron. Water Research 45, 19952001.
Crane, R. A. and Scott, T. B. (2012). Nanoscale zero-valent iron: future prospects for an emerg-
ing water treatment technology. J. Hazard. Mater. 211212, 112125.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 459
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Deng, B., and Hu, S. (2001). Reductive dechlorination of chlorinated solvents on zerovalent iron
surfaces. In: Smith, J. A., Burns, S. E., editors. Physicochemical groundwater remediation.
New York: Kluwer Academic:139159.
Dickinson, M., and Scott, T. B. (2010). The application of zero-valent iron nanoparticles for the
remediation of a uranium-contaminated waste efuent. J. Hazard. Mater. 178, 171179.
Dong, H., and Lo, I. M. C. (2013). Inuence of humic acid on the colloidal stability of surface-
modied nano zero-valent iron. Water Res. 47, 419427.
Dukhin, A. S., and Goetz, P. J. (2002). Ultrasound for characterizing colloids: particle sizing,
zeta potential, rheology. New York: Elsevier Science.
Elliott, D. W., and Zhang, W. (2001). Field assessment of nanoscale bimetallic particles for
groundwater treatment. Environ. Sci. Technol. 35, 49224926.
Esfahani, A. R., Firouzi, A. F., Sayyad, G., Kiasat, A., Alidokht, L., and Khataee, A. R. (2014). Pb
(II) removal from aqueous solution by polyacrylic acid stabilized zero-valent iron nanopar-
ticles: process optimization using response surface methodology. Res. Chem. Intermed. 40,
431445.
ESTCP Cost and Performance Report ER 200431. (2010). Emulsied Zero-Valent Nano-Scale
Iron Treatment of Chlorinated Solvent DNAPL Source Areas. US Department of Defence.
https://clu-in.org/ download/ contaminantfocus/ dnapl/ Treatment_Technologies/ DNAPL-
ZVI-ER-200431-C&P.pdf (Accessed 02/12/2015).
Fan,J.,Guo,Y.,Wang,J.,andFan,M.(2009).Rapiddecolorizationofazodyemethyl
orange in aqueo us solution by nanoscale zerovalent iron particles. J. Hazard. Mater.
166, 904910.
Fang, Z., Chen, J., Qiu, X., Cheng, W., and Zhu, L. (2011). Effective removal of antibiotic metro-
nidazole from water by nanoscale zero-valent iron particles. Desalination 268, 6067.
Fang, Z., Qiu, X., Chen, J., and Qiu, X. (2010). Degradation of metronidazole by nanoscale zero-valent
metal prepared from steel pickling waste liquor. Appl.Catal.B:Environ.100, 221228.
Fu, F., Dionysiou, D. D., and Liu, H. (2014). The use of zero-valent iron for groundwater reme-
diation and wastewater treatment: A review. J. Hazard. Mater. 267, 194205.
Geng, B., Jin, Z., Li, T., and Qi, X. (2009). Kinetics of hexavalent chromium removal from water
by chitosan-Fe
0
nanoparticles. Chemosphere 75, 825 830.
Ghauch, A., Tuqan, A., and Assi, H. A. (2009). Antibiotic removal from water: Elimination of
amoxicillin and ampicillin by microscale and nanoscale iron particles. Environ. Pollution
157, 16261635.
Giasuddin, A. B., Kanel, S. R., and Choi, H. (2007). Adsorption of humic acid onto nanoscale
zerovalent iron and its effect on arsenic removal. Environ. Sci. Technol. 41, 20222027.
Gillham, R. W., and OHannesin, S. F. (1994). Enhanced degradation of halogenated aliphatics
by zero-valent iron in ground water. Groundwater 32, 958967.
He, F., and Zhao, D. (2005). Preparation and characterization of a new class of starch- stabilized
bimetallic nanoparticles for degradation of chlorinated hydrocarbons in water. Environ. Sci.
Technol. 39, 33143320.
He, F., Zhao, D., Liu, J., and Roberts, C. B. (2007). Stabilization of Fe¡Pd nanoparticles with
sodium carboxymethyl cellulose for enhanced transport and dechlorination of trichloroethy-
lene in soil and groundwater. Ind. Eng. Chem. Res. 46, 2934.
Hojeong, K., Hye-Jin, H., Juri, J., Seong-Hye, K., and Ji-and Won, Y. (2010). Degradation of tri-
chloroethylene (TCE) by nanoscale zero-valent iron (nZVI) immobilized in alginate bead. J.
Hazard. Mater. 176, 10381043.
Hsu, J., Liao, C., and Wei, Y. (2011). Nitrate removal by synthetic nanoscale zero-valent iron in
aqueous recirculated reactor. Sustain. Environ. Res. 21, 353359.
Hu, J., Lo, I. M., and Chen, G. (2004). Removal of Cr(VI) by magnetite nanoparticle. Water Sci.
Technol. 50, 139146.
460 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Hwang, Y.-H., and Shin, H.-S. (2013). Effects on nano zero-valent iron reactivity of interactions
between hardness, alkalinity, and natural organic matter in reverse osmosis concentrate. J.
Environ. Sci. 25, 21772184.
Hwang, Y.-H., Kim, D.-G., and Shin, H.-S. (2011). Effects of synthesis conditions on the charac-
teristics and reactivity of nano scale zero valent iron. Appl. Catal. B: Environ. 105, 144150.
Hydutsky, B. W., Mack, E. J., Beckerman, B. B., Skluzacek, J. M., and Mallouk, T. E. (2007).
Optimization of nano- and microiron transport through sand columns using polyelectrolyte
mixtures. Environ. Sci. Technol. 41, 64186424.
Interstate Technology and Regulatory Council. (2005). Permeable reactive barriers: lessons
learned/new directions. Report # PRB-4. Washington DC: ITRC.
Iurascu, B., Siminiceanu, J., Vione, D., Vicente, M. A., and Gil, A. (2009). Phenol degradation in
water through a heterogeneous photo-Fenton process catalysed by Fe-treated laponite.
Water Res. 43, 13131322.
Jiang, Z., Lv, L., Zhang, W., Du, Q., Pan, B., Yang, L., and Zhang, Q. (2011). Nitrate reduction
using nanosized zero-valent iron supported by polystyrene resins: Role of surface functional
groups. Water Res. 45, 21912198.
Johnson, R. L., Johnson, G. O., Nurmi, J. T., and Tratnyek, P. G. (2009). Natural organic matter
enhanced mobility of nano zerovalent iron. Environ. Sci. Technol. 43, 545560.
Jortner, J., and Rao, C. N. (2002). Nanostructured advanced materials. Perspectives and direc-
tions. Pure Appl. Chem. 74, 14911506.
Kanel, S. R., and Choi, H. (2007). Transport characteristics of surface-modied nanoscale zero-
valent iron in porous media. Water Sci. Technol. 55, 157162.
Kanel, S. R., Greneche, J. M., and Choi, H. (2006). Arsenic(V) removal from groundwater using
nano scale zero-valent iron as a colloidal reactive barrier material. Environ. Sci. Technol. 40,
20452050.
Kanel, S. R., Manning, B., Charlet, L., and Choi, H. (2005). Removal of arsenic(III) from
groundwater by nanoscale zero-valent iron. Environ. Sci. Technol. 39, 12911298.
Kanel, S. R., Nepal, D., Manning, B., and Choi, H. (2007). Transport of surface-modied iron
nanoparticle in porous media and application to arsenic (III) remediation. J. Nanopart. Res.
9, 725735.
Kassaee, M. Z., Motamedi, E., Mikhak, A., and Rahnemaie, R. (2011). Nitrate removal from
water using iron nanoparticles produced by arc discharge vs. reduction. Chem. Eng. J. 166,
490495.
Keller, A. A., Garner, K., Miller, R. J., and Lenihan, H. S. (2012). Toxicity of nano-zero valent
iron to freshwater and marine organisms. PLoS One 7(8), e43983.
Kim, H. J. (2009). Transport, reactivity and fate of polyelectrolyte modied zero valent iron
nanoparticles used for groundwater remediation in heterogeneous porous media. Doctoral
thesis, Carnegie Mellon University, Pittsburgh, PA, USA
Kim, H. J., Phenrat, T., Tilton, R. D., and Lowry, G. V. (2009). Fe
0
nanoparticles remain mobile
in porous media after aging due to slow desorption of polymeric surface modiers. Environ.
Sci. Technol. 43, 38243830.
Kim, J. S., Shea, P. J., Yang, J. E., and Kim, J. E. (2007). Halide salts accelerate degradation of
high explosives by zerovalent iron. Environ. Pollut. 147, 63441.
Kim, S. A., Kamala-Kannan, S., Lee, K., Park, Y., Shea, P. J., Lee, W., Kim, H., and Oh, B. (2013).
Removal of Pb(II) from aqueous solution by a zeolitenanoscale zero-valent iron composite.
Chem. Eng. J. 217, 5460.
Klimkova, S., Cernik, M., Lacinova, L., Filip, J., Jancik, D., and Zboril, R. (2011). Zero-valent
iron nanoparticles in treatment of acid mine water from in situ uranium leaching. Chemo-
sphere 82, 11781184.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 461
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Konstantina, T., Elpida, P., and Nikolaos, P. N. (2007). Modeling of arsenic immobilization by
zero valent iron. Eur. J. Soil Biol. 43, 356367.
Krajangpan, S., Jarabek, L., Jepperson, J., Chisholm, B., and Bezbaruah, A. (2008). Polymer
modied iron nanoparticles for environmental remediation. Polym. Preprints 49, 921922.
Lee, C., Kim, J. Y., Lee, W. I., Nelson, K. L., Yoon, J., and Sedlak, D. L. (2008). Bactericidal effect
of zerovalent iron nanoparticles on Escherichia coli. Environ. Sci. Technol. 42, 49274933.
Li, H., De Boer, C. V., Buchau, A., Klaas, N., Rucker, W. M., and Hermes, H. (2012). Develop-
ment of an inductive concentration measurement sensor of nano sized zero valent iron.
Paper presented at the 9th International Multi-conference on Systems, Signals and Devices,
Chemnitz, Germany.
Li, H., Zhou, Q., Wu, Y., F J., Wang, T., and Jiang, G. (2009). Effects of waterborne nano-iron
onmedaka (Oryzias latipes): Antioxidant enzymatic activity, lipid peroxidation and histopa-
thology. Ecotoxicol. Environ. Saf. 72, 684692.
Li, X., Elliott, D. W., and Zhang, W. (2006). Zero valent iron nanoparticles for abatement of
environmental pollutants: materials and engineering aspects. Crit. Rev. Solid State Mater.
Sci. 31, 111122.
Li, X., and Zhang, W. (2006). Iron nanoparticles: the core-shell structure and unique properties
for Ni(II) sequestration. Langmuir 22, 46384642.
Li, Z., Greden, K., Alvarez, P. J. J., Gregory, K. B., and Lowry, G. V. (2010). Adsorbed polymer
and NOM limits adhesion and toxicity of nano scale zerovalent iron to E. coli. Environ. Sci.
Technol. 44, 34623467.
Liang, L., Gu, B., and Yin, X. (1996). Removal of technetium-99 from contaminated groundwa-
ter with sorbents and reductive materials. Sep. Technol. 6, 111122.
Liao, C. J., Chung, T. L., Chen, W., and Kuo, S. L. (2007). Treatment of pentachlorophenol-con-
taminated soil using nano-scale zero-valent iron with hydrogen peroxide. J. Mol. Catal. A:
Chem. 265, 189194.
Lien, H. L. and Zhang, W. X. (2001). Nanoscale iron particles for complete reduction of chlori-
nated ethenes. Colloids Surf. A: Physicochem. Eng. Aspects 191, 97105.
Lien,L.,Elliott,D.W.,Sun,Y.P.,andZhang,W.X.(2006).Recentprogressinzero-
valent iron nanoparticles for groundwater remediation. J. Environ. Eng. Man age. 16(6),
371380.
Lin, Y., Weng, C., and Chen, F. (2008). Effective removal of AB24 dye by nano/micro-size zero-
valent iron. Separ. Purif. Technol. 64, 2630.
Lin, Y. H., Tseng, H.-H., Wey, M.-Y., and Lin, M.-D. (2010). Characteristics of two types of sta-
bilized nano zero-valent iron and transport in porous media. Sci. Total Environ. 408, 2260
2267.
Liou, Y. H., Lo, S., Kuan, W. H., Lin, C., and Weng, S. C. (2006). Effect of precursor concentra-
tion on the characteristics of nanoscale zerovalent iron and its reactivity of nitrate. Water
Res. 40, 24852492.
Liu, H. B., Chen, T. H., Chang, D. Y., Chen, D., Liu, Y., He, H. P., Yuan, P., and Frost, R. (2012).
Nitrate reduction over nanoscale zero-valent iron prepared by hydrogen reduction of goe-
thite. Mater. Chem. Phys. 133, 205
211.
Liu, Y. Q., Majetich, S. A., Tilton, R. D., Sholl, D. S., and Lowry, G. V. (2005). TCE dechlorina-
tion rates, pathways, and efciency of nanoscale iron particles with different properties.
Environ. Sci. Technol 39, 13381345.
Liu, Y., Phenrat, T., and 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, 78817887.
Liu, Y., Shen, J. M., Chen, Z. L., and Liu, Y. (2011). Degradation of p-chloronitrobenzene in
drinking water by manganese silicate catalyzed ozonation. Desalination 279, 219224.
462 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Lowry, G. V., and Johnson, K. M. (2004). Congener-specic dechlorination of dissolved PCBs
by microscale and nanoscale zerovalent iron in a water/methanol solution. Environ. Sci.
Technol. 38, 52085216.
Ludwig, R. D., Su, C., Lee, T. R., Wilkin, R. T., Acree, S. D., Ross, R. R., and Keeley, A. (2007). In
situ chemical reduction of Cr(VI) in groundwater using a combination of ferrous sulfate and
sodium dithionite: a eld investigation. Environ. Sci. Technol. 41, 52995305.
Lv, X., Xu, J., Jiang, G., Tang, J., and Cut, X. (2012). Highly active nanoscale zero-valent iron
(nZVI)Fe
3
O
4
nanocomposites for the removal of chromium(VI) from aqueous solutions. J.
Colloid Interface Sci. 369, 460469.
Madhavi, V., Reddy, A. V. B., Reddy, K. G., and Madhavi, G. (2012). A simple method for the
determination of efciency of stabilized Fe
0
nanoparticles for detoxication of chromium
(VI) in water. J. Chem. Pharm. Res. 4, 15391545.
Matheson, L. J., and Tratnyek, P. G. (1994). Reductive dehalogenation of chlorinated methanes
by iron metal. Environ. Sci. Technol. 28, 20452053.
Morgada, M. E., Levy, I. K., Salomone, V., Farias, S. S., Lopez, G., and Litter, I. M. (2009). Arse-
nic (V) removal with nanoparticulate zerovalent iron: Effect of UV light and humic acids.
Catal. Today 143, 261268.
Morrison, I. D., and Ross, S. (2002). Colloidal dispersions: suspensions, emulsions, and foams.
New York: Wiley.
Nazlı, E. (2008). Characterization of the adsorption behavior of aqueous Cd(II) and Ni(II) ions
on nanoparticles of zero-valent iron. Masters thesis, School of Engineering and Science of
_
Izmir Institute of Technology, Turkey.
Nhung, N. T., and Thuong, N. T. K. (2008). Research on the removal of hexavalent chromium
from aqueous solution by iron nanoparticles. VNU J. Sci. (Nat. Sci. Technol.) 233237.
OHara, S., Krug, T., Quinn, J., Clausen, C., and Geiger, C. (2006). Field and laboratory evalua-
tion of the treatment of DNAPL source zones using emulsied zero-valent iron. Remediation
16, 3556.
OCarroll, D., Sleep, B., Krol, M., Boparai, H., and Kocur, C. (2013). Nanoscale zero valent iron
and bimetallic particles for contaminated site remediation. Adv. Water Resour. 51, 104122.
OHannesin, S. F. and Gillham, R. W. (1998). Long-term performance of an in situ ironwall for
remediation of VOCs. Groundwater 36, 164170.
Olegario, J. T., Yee, N., Miller, M., Sczepaniak, J., and Manning, B (2010). Reduction of Se (VI)
to Se(-II) by zerovalent iron nanoparticle suspensions. J. Nanopart. Res. 12, 20572068.
Orth, W. S., and Gillham, R. W. (1996). Dechlorination of trichloroethene in aqueous solution
using Fe(0). Environ. Sci. Technol. 30, 6671.
Phenrat, T., Long, T. C., Lowry, G. V., and Veronesi, B. (2009). Partial oxidation (aging) and
surface modication decrease the toxicity of nanosized zerovalent iron. Environ. Sci. Technol.
43, 195200.
Phenrat, T., Saleh, N., Sirk, K., Tilton, R. D., and Lowry, G. V. (2007). Aggregation and sedimen-
tation of aqueous nanoscale zerovalent iron dispersions. Environ. Sci. Technol. 41, 284 290.
Ponder, S. M., Darab, J. G., and Mallouk, T. E. (2000). Remediation of Cr (VI) and Pb (II) aque-
ous solutions using supported, nano-scale zero-valent iron. Environ. Sci. Technol. 34, 2564
2569.
Ponder, S. M., Darab, J. G., Bucher, J., Caulder, D., Craig, I., Davis, L., Edelstein, N., Lukens, W.,
Nitsche, H., Rao, L. F., Shuh, D. K., and Mallouk, T. E. (2001). Surface chemistry and electro-
chemistry of supported zerovalent iron nanoparticles in the remediation of aqueous metal
contaminants. Chem. Mater. 13, 479486.
Poursaberi, T., Konoz, E., Sarra, A. H. M., Hassanisadi, M., and Hajifathli, F. (2012). Applica-
tion of nanoscale zero-valent iron in the remediation of DDT from contaminated water.
Chem. Sci. Trans. 1, 658668.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 463
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Pradhan, A. A., and Gogate, P. R. (2010). Degradation of p-nitro phenol using acoustic cavita-
tion and Fenton chemistry. J. Hazard. Mater. 173, 517522.
Puls, R. W., Paul, C. J., and Powell, R. M. (1999). The application of in situ permeable reactive
(zero-valent iron) barrier technology for the remediation of chromate-contaminated ground-
water: a eld test. Appl. Geochem. 14, 9891000.
Quinn, J., Geiger, C., Clausen, C., Brooks, K., Coon, C., OHara, S., Krug, T., Major, D., Yoon,
W. S., Gavaskar, A., and Holdsworth, T. (2005). Field demonstration of DNAPL dehalogena-
tion using emulsied zero-valent iron. Environ. Sci. Technol. 39, 13091318.
Reijnders, L. (2006). Cleaner nanotechnology and hazard reduction of manufactured nanopar-
ticles. J. Clean. Prod. 14, 124133.
Ryu, A., Jeong, S., Jang, A., and Choi, H. (2011). Reduction of highly concentrated nitrate using
nanoscale zero-valent iron: Effects of aggregation and catalyst on reactivity. Appl Catal B:
Environ. 105, 128135.
Saad, R., Thiboutot, S., Ampleman, G., Dashan, W., and Hawari, J. (2010). Degradation of trini-
troglycerin (TNG) using zero-valent iron nanoparticles/nanosilica SBA-15 composite
(ZVINs/SBA-15). Chemosphere 81, 853858.
Saleh, N., Kim, H. J., Phenrat, T., Matyjaszewski, K., Tilton, R. D., and Lowry, G. V. (2008).
Ionic strength and composition affect the mobility of surface-modied Fe0 nanoparticles in
water saturated sand columns. Environ. Sci. Technol. 42, 33493355.
Saleh, N., Phenrat, T., Sirk, K., Dufour, B., Ok, J., Sarbu, T., Matyjaszewski, K., Tilton, R. D., and
Lowry, G. V. (2005). Adsorbed triblock copolymers deliver reactive iron nanoparticles to the
oil/water interface. Nano Lett. 5, 24892494.
Saleh, N., Sirk, K., Liu, Y., Phenrat, T., Dufour, B., Matyjaszewski, K., Tilton, R. D., and Lowry,
G. V. (2007). Surface modications enhance nanoiron transport and NAPL targeting in satu-
rated porous media. Environ. Eng. Sci. 24, 4557.
Satapanajaru, T., Chompuchan, C., Suntornchot, P., and Pengthamkeerati, P. (2011). Enhancing
decolorization of reactive black 5 and reactive red 198 during nano zerovalent iron treat-
ment. Desalination 266, 218230.
Selvarani, M., and Prema, P. (2012). Removal of toxic metal hexavalent chromium [cr (vi)] from
aqueous solution using starch stabilized nanoscale zerovalent iron as adsorbent: Equilib-
rium and kinetics. Int. J. Environ. Sci. 2, 19621975.
Shen, J. M., Chen, Z. L., Xu, Z. Z., Li, X. Y., Xu, B. B., and Qi, F. (2008). Kinetics and mechanism
of degradation of p-chloronitrobenzene in water by ozonation. J. Hazard. Mater. 152, 1325
1331
Shi, L., Lin, Y., Zhang, X., and Chen, Z. (2011). Synthesis, characterization and kinetics of ben-
tonite supported nZVI for the removal of Cr (VI) from aqueous solution. Chem. Eng. J. 171,
612617.
Shih, Y. H., Hsu, C. Y., and Su, Y. F. (2011). Reduction of hexachlorobenzene by nanoscale zero-
valent iron: kinetics, pH effect, and degradation mechanism. Separ. Purif. Technol. 76, 268
274.
Shin, K., and Cha, D. K. (2008). Microbial reduction of nitrate in the presence of nanoscale
zero-valent iron. Chemosphere 72, 257262.
Shu,H.,Chang,M.,Chen,C.,andChen,P.(2010).Usingresinsupportednanozero-val-
ent iron parti cles f or decoloration of Acid Blue 113 azo dye solution. J. Hazard. Mater.
184, 499505.
Singh, R., Misra, V., and Singh, R. P. (2011). Synthesis, characterization and role of zero-valent
iron nanoparticle in removal of hexavalent chromium from chromium-spiked soil. J. Nano-
part. Res. 13, 40634073.
Sinha, A., and Bose, P. (2006). Dehalogenation of 2-chloronapthalene by cast iron. Water Air
Soil Poll. 172, 375390.
464 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Sirk, K. M., Saleh, N. B., Phenrat, T., Kim, H.-J., Dufour, B., Ok, J., Golas, P. L., Matyjaszewski,
K., Lowry, G. V., and Tilton, R. D. (2009). Effect of adsorbed polyelectrolytes on nanoscale
zero valent iron particle attachment to soil surface models. Environ. Sci. Technol. 43, 3803
3808.
Stumm, W., and Morgan, J. J. (1996). Aquatic chemistry (3rd ed.). New York: Wiley.
Sun, Y. P., Li, X. Q., Cao, J., Zhang, W. X., and Wang, H. P. (2006). Characterization of zero-val-
ent iron nanoparticles. Adv. Colloid Interface Sci. 120, 4756.
Sun, Y. P., Li, X. Q., Zhang, W. X., and Wang, H. P. (2007). A method for the preparation of sta-
ble dispersion of zero-valent iron nanoparticles. Colloids Surf. A: Physicochem. Eng. Aspects
308, 6066.
Tanboonchuy, V., Hsu, J., Grisdanurak, N., and Liao, C. (2011). Gas-bubbled nano zero-valent
iron process for high concentration arsenate removal. J. Hazard. Mater. 186, 21232128.
Tiraferri, A., Chen, K.-L., Sethi, R., and Elimelech, M. (2008). Reduced aggregation and sedi-
mentation of zero-valent iron nanoparticles in the presence of guar gum. J. Colloid Interface
Sci. 324, 7179.
Tiraferri, A., and Sethi, R. (2009). Enhanced transport of zerovalent iron nanoparticles in satu-
rated porous media by guar gum. J. Nanopart. Res. 11, 635645.
Tratnyek, P. G., and Johnson, R. L. (2006). Nanotechnologies for environmental clean up. Nano
Today 1, 4448.
U.S. Environmental Protection Agency. (2008a). Brownelds and land revitalization programs:
changing American land and lives. Report number: EPA 560-F-08-241. Washington, DC:
Ofce of Solid Waste and Emergency Response.
U.S. Environmental Protection Agency. (2008b). Nanotechnology for site remediation fact
sheet. Report number: EPA 542-F-08-009. Washington, DC: Ofce of Solid Waste and
Emergency Response.
U.S. Environmental Protection Agency. (2010). Field scale Remediation Experience using Iron
Nanoparticles and Evolving Risk-Benet Understanding, U.S. EPA Technology Innovation
and Field Services Division and Contaminated Land: Applications in Real Environments
(CL:AIRE) https://clu-in.org/conf/tio/nano-iron_121410/ (Accessed 02/12/2015).
Uzum, C., Shahwan, T., Eroglu, A. E., Lieberwirth, I., Scott, T. B., and Hallam, K. R. (2008).
Application of zero-valent iron nanoparticles for the removal of aqueous Co
2C
ions under
various experimental conditions. Chem. Eng. J. 144, 213220.
Uzum, C., Shahwan, T., Eroglu, A. E., Lieberwirth, I., Scott, T. B., and Hallam, K. R. (2009). Syn-
thesis and characterization of kaolinite-supported zero-valent iron nanoparticles and their
application for the removal of aqueous Cu
2C
and Co
2C
ions. Appl. Clay Sci. 43, 172181.
Vogel, T. M., Criddle, C. S., and McCarty, P. L. (1987). ES critical reviews: Transformations of
halogenated aliphatic compounds. Environ. Sci. Technol. 21, 722736.
Wang, C. B., and Zhang, W. X. (1997). Synthesizing nanoscale iron particles for rapid and com-
plete dechlorination of TCE and PCBs. Environ. Sci. Technol. 31, 21542156.
Wang, C. M., Baer, D. R., Amonette, J. E., Engelhard, M. H., Antony, J., and Qiang, Y. (2009).
Morphology and electronic structure of the oxide shell on the surface of iron nanoparticles.
J. Am. Chem. Soc. 131, 88248832.
Wen, Z., Zhang, Y., and Dai, C. (2014). Removal of phosphate from aqueous solution using
nanoscale zerovalent iron (nZVI). Colloids Surf. A: Physicochem. Eng. Aspects 457, 433440.
Wiesner, M. R., Lowry, G. V., Alvarez, P., Dionysiou, D., and Biswas, P. (2006). Assessing the
risks of manufactured nanomaterials. Environ. Sci. Technol. 40, 43364345.
Wu, D., Shen, Y., Ding, A., Qiu, M., Yang, Q., and Zheng, S. (2013). Phosphate removal from
aqueous solutions by nanoscale zero-valent iron. Environ. Technol. 34, 26632669.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 465
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
Wu, L., Shamsuzzoha, M., and Ritchie, S. M. C. (2005). Preparation of cellulose acetate sup-
ported zero-valent iron nanoparticles for the dechlorination of trichloroethylene in water. J.
Nanopart. Res. 7, 469476.
Xi, Y., Megharaj, M., and Naidu, R. (2011). Dispersion of zerovalent iron nanoparticles onto
bentonites and use of these catalysts for orange II decolourization. Appl. Clay Sci. 53, 716
722.
Xie, L., and Shang, C. (2005). Role of humic acid and quinine model compounds in bromated
reduction by zerovalent iron. Environ. Sci. Technol. 39, 10921100.
Yan, W., Herzing, A. A., Kiely, C. J., and Zhang, W. (2010). Nanoscale zero-valent iron (nZVI):
Aspects of the core-shell structure and reactions with inorganic species in water. J. Contam.
Hydrol. 118, 96104.
Yang, G. C. C., Tu, H., and Hung, C. (2007). Stability of nanoiron slurries and their transport in
the subsurface environment. Separ. Purif. Technol. 58, 166 172.
Yildirim, I. Z., and Prezzi, M. (2011). Chemical, mineralogical and morphological properties of
steel slag. Adv. Civil Eng. 2011, 463638.
You, Y., Han, J., Chiu, P. C., and Jin, Y. (2005). Removal and inactivation of waterborne viruses
using zerovalent iron. Environ. Sci. Technol. 39, 9263 9269.
Zhang, M., He, F., Zhao, D., and Hao, X. (2011). Degradation of soil-sorbed trichloroethylene by
stabilized zero valent iron nanoparticles: Effects of sorption, surfactants, and natural organic
matter. Water Res. 45, 24012414.
Zhang, W. (2003). Nanoscale iron particles for environmental remediation: An overview. J.
Nanopart. Res. 5, 323332.
Zhang, X., Lin, S., Lu, X., and Chen, Z. (2010). Removal of Pb (II) from water using synthesized
kaolin supported nanoscale zero-valent iron. Chem. Eng. J. 163, 243248.
Zhang, X. W., and Elliott, D. W. (2006). Applications of iron nanoparticles for groundwater
remediation. Remediation 16, 721.
Zhao, Z. S., Liu, J. F., Tai, C., Zhou, Q. F., Hu, J. T., and Jiang, G. B. (2008). Rapid decolorization
of water soluble azo-dyes by nanosized zero-valent iron immobilized on the exchange resin.
Sci. China Series B: Chem. 51, 186192.
Zhu, B.-W., Lim, T.-T., and Feng, J. (2006). Reductive dechlorination of 1 2, 4-trichlorobenzene
with palladized nanoscale Fe0 particles supported on chitosan and silica. Chemosphere 65,
11371143.
Zhu, B.-W., Lim, T.-T., and Feng, J. (2008). Inuences of amphiphiles on dechlorination of a tri-
chlorobenzene by nanoscale Pd/Fe: adsorption, reaction kinetics, and interfacial interactions.
Environ. Sci. Technol. 42, 45134519.
Zhu, H., Jia, Y., Wu, X., and Wang, H. (2009). Removal of arsenic from water by supported
nano zero-valent iron on activated carbon. J. Hazard. Mater. 172, 15911596.
Ziajahromi, S., Zand, A. D., and Khanizadeh, M. (2012). Nitrate removal from water using syn-
thesis nanoscale zero-valent iron (nZVI). International Conference on Applied Life Sciences
(ICALS2012), Turkey, 1012 September 2012, 5963.
466 R. MUKHERJEE ET AL.
Downloaded by [University of Auckland Library] at 20:06 23 March 2016
... In recent years, nanoscale zerovalent iron (nZVI) particles have demonstrated their efficacy in treating various environmental contaminants like pesticides, pharmaceuticals, and chlorinated organic compounds [20]. Abundant surface sites on nZVI particles allow for extensive physical and chemical interactions, including oxidation-reduction reactions, making them highly promising adsorbents for nutrient pollutants like phosphates as well [21,22]. ...
... However, studies on the synthesis of nZVI functionalized alginate aerogels and their ability for phosphate sequestration have not yet been investigated. It is worth mentioning that a major concern with the use of nZVI particles independently in water matrices is their tendency to oxidize and eventually lose their magnetic properties [20]. This makes their separation from water matrices post treatment challenging. ...
Article
Full-text available
Polymeric aerogels, with their versatile physicochemical properties and capacity for functionalization, are innovative materials being increasingly explored for water treatment applications. In this study, novel millimetric sized alginate-based aerogel granules functionalized with nZVI (nanoscale zero-valent iron) were developed and evaluated for their phosphate sequestration performance. Efficient phosphate removal from water is critical as excessive levels of phosphates can lead to eutrophication and negatively impact water quality. nZVI-aerogel granules exhibited significant enhancements in phosphate removal efficiencies (up to 97%) compared to non-functionalized bare-aerogel granules (15%). Average Langmuir removal capacities of 77 mg-PO4³⁻/g were observed consistently for nZVI-aerogel granules across a broad pH range from 3 to 7, which further increased under alkaline conditions reaching up to 180 mg-PO4³⁻/g at pH 11. Kinetic studies were well described by the pseudo first-order kinetic model in the pH 3–7 range, with rates declining from 0.11 h⁻¹ to 0.07 h⁻¹ as pH increased. In contrast, mixed kinetic trends were observed in alkaline pH with rapid phosphate removal followed by a short-term desorption. Solution pH measurements, and analysis of nZVI-aerogel granule surface chemistry and morphology post batch experiments revealed the involvement of multiple sequestration mechanisms including electrostatic adsorption, ion exchange, and surface precipitation. nZVI-aerogel granule morphology remained stable under all tested conditions (except at pH 11) suggesting their strong potential for facilitating efficient post-treatment separation and recovery.
... However, recent research has shown that using a co-catalytic heterogeneous catalyst (CoFe 2 O 4 /MoS 2 ) in the Fenton reaction, which is pH-independent, improves organic pollutant removal (Yan et al. 2021). Iron nanoparticles are another type of catalyst that can be applied in pollutant degradation in wastewater, and their synthesis, applications, and evaluation are well documented (Stefaniuk et al. 2016;Kumar et al. 2016;Sun et al. 2016). Their effectiveness in removing pollutants from wastewater is due to their high surface area and small particle size. ...
Article
Full-text available
In this study, three acid mine drainage (AMD) sources were investigated as potential sources of iron for the synthesis of iron nanoparticles using green tea extract (an environmentally friendly reductant) or sodium borohydride (a chemical reductant). Electrical conductivity (EC), total dissolved solids (TDS), dissolved oxygen (DO), oxidation–reduction potential (ORP), ion chromatography (IC), and inductively coupled plasma-mass spectroscopy (ICP-MS) techniques were used to characterize the AMD, and the most suitable AMD sample was selected based on availability. Additionally, three tea extracts were characterized using ferric-reducing antioxidant power (FRAP) and 2,2-diphenyl-1-picryl-hydrazine-hydrate (DPPH), and the most suitable environmentally friendly reductant was selected based on the highest FRAP (1152 µmol FeII/g) and DPPH (71%) values. The synthesized iron nanoparticles were characterized and compared using XRD, STEM, Image J, EDS, and FTIR analytical techniques. The study shows that the novel iron nanoparticles produced using the selected green tea (57 nm) and AMD were stable under air due to the surface modification by polyphenols contained in green tea extract, whereas the nanoparticles produced using sodium borohydride (67 nm) were unstable under air and produced a toxic supernatant. Both the AMD-based iron nanoparticles can be used as Fenton-like catalysts for the decoloration of methylene blue solution. While 99% decoloration was achieved by the borohydride-synthesized nanoparticles, 81% decoloration was achieved using green tea-synthesized nanoparticles.
... Recently, nano-scale ZVI has developed as an innovative technology that can be utilized to enhance the granular form of ZVI to improve the treatment efficiency by reducing pollutant mass and flux [12]. The nano-zero valent iron (nZVI) is used to remove various inorganic pollutants such as ( ) CU , CO , NI , PO 2 2 2 4 3 + + + and organic contaminants like (azo dyes, pharmaceutical disposal, and halogenated organic components [13]. ...
Article
Full-text available
The textile industry is considered a source of pollution because of the discharge of dye wastewater. The dye wastewater effluent has a significant impact on the aquatic environment. According to the World Bank, textile dyeing, and treatment contribute 17 to 20% of the pollution of water. This paper aims to prepare the bimetallic nano zero-valent iron-copper (Fe⁰-Cu), algae-activated carbon, and their composites (AC-Fe⁰-Cu), which are employed as adsorbents. In this paper, Synthetic adsorbents are prepared and examined for the adsorption and removal of soluble cationic crystal violet (CV) dye. The influence of synthetic adsorbents on the adsorption and removal of soluble cationic crystal violet (CV) dye is investigated using UV-V spectroscopy at different pH (3–10), time intervals (15–180) min, and initial dye concentrations (50–500 ppm). Raw algae exhibit an impressive 96.64% removal efficiency under the following conditions: pH 7, contact time of 180 min, rotational speed of 120 rpm, temperature range of 25 °C–30 °C, concentration of 300 ppm in the CV dye solution, and a dose of 4 g l⁻¹ of raw algae adsorbent. The best removal efficiencies of Raw algae Fe⁰-Cu, and H3PO4 chemical AC-Fe⁰-Cu are 97.61 % and 97.46 %, respectively, at pH = 7, contact time = 150 min, rotational speed = 120 rpm, T = (25–30) °C, concentration = 75 ppm of CV dye solution, and 1.5 g l⁻¹ doses of raw algae F e0-Cu adsorbent and 1 g l⁻¹ dose of H3PO4 chemical AC-Fe⁰-Cu adsorbent. The maximum amounts (q max) of Bi-RA and RA adsorbed for the adsorption process of CV are 85.92 mg g⁻¹ and 1388 mg g⁻¹, respectively. The Bi-H3A-AC model, optimized using PSO, demonstrates superior performance, with the highest adsorption capacity estimated at 83.51 mg g⁻¹. However, the Langmuir model predicts a maximum adsorption capacity (q e ) of 275.6 mg g⁻¹ for the CV adsorption process when utilizing Bi-H3A-AC. Kinetic and isothermal models are used to fit the data of time and concentration experiments. DLS, zeta potential, FT-IR, XRD, and SEM are used to characterize the prepared materials. Response surface methodology (RSM) is used to model the removal efficiency and then turned into a numerical optimization approach to determine the ideal conditions for improving removal efficiency. An artificial neural network (ANN) is also used to model the removal efficiency.
... Nanoscale zero-valent iron (nZVI) has attracted considerable attention due to its strong reduction potential in the remediation of a wide range of contaminants. With a standard potential of E 0 (Fe 2+ /Fe) = 0.447 V (Haynes 2016), nZVI can transform a variety of persistent halogenated organic compounds, nitroaromatic compounds, and reducible inorganic pollutants in wastewater treatment and groundwater remediation for decades (Grieger et al. 2010;Raychoudhury and Scheytt 2013;Fu et al. 2014;Mukherjee et al. 2016). However, nZVI has some technical issues such as easy agglomeration, poor stability, and low durability, which need to be overcome in practical field applications. ...
Article
Full-text available
The widespread occurrence of emerging brominated flame retardant tetrabromobisphenol S (TBBPS) has become a major environmental concern. In this study, a nanoscale zero-valent iron (nZVI) impregnated organic montmorillonite composite (nZVI-OMT) was successfully prepared and utilized to degrade TBBPS in aqueous solution. The results show that the nZVI-OMT composite was very stable and reusable as the nZVI was well dispersed on the organic montmorillonite. Organic montmorillonite clay layers provide a strong support, facilitate well dispersion of the nZVI chains, and accelerate the overall TBBPS transformation with a degradation rate constant 5.5 times higher than that of the original nZVI. Four major intermediates, including tribromobisphenol S (tri-BBPS), dibromobisphenol S (di-BBPS), bromobisphenol S (BBPS), and bisphenol S (BPS), were detected by high-resolution mass spectrometry (HRMS), indicating sequential reductive debromination of TBBPS mediated by nZVI-OMT. The effective elimination of acute ecotoxicity predicted by toxicity analysis also suggests that the debromination process is a safe and viable option for the treatment of TBBPS. Our results have shown for the first time that TBBPS can be rapidly degraded by an nZVI-OMT composite, expanding the potential use of clay-supported nZVI composites as an environmentally friendly material for wastewater treatment and groundwater remediation.
Article
Full-text available
The development of electronics with net zero carbon emissions through more efficient and environmentally friendly materials and processes is still a challenge. Here, alternative chemical synthesis routes of metal conductive nanoparticles, based on biodegradable materials are explored, such as nickel, iron–nickel alloy and iron nanoparticles, to be used, in the long term, as fillers in inks for inject printing. Thus, Ni and FeNi metal nanoparticles of 25–12 nm, forming aggregates of 614–574 nm, respectively, are synthesized in water in the presence of a polyol and a reducing agent and under microwave heating that enables a more uniform and fast heating. Iron nanoparticles of 120 ± 40 nm are synthesized in polyol that limits the aggregation and the oxidation degree. Commercial metal nanoparticles of iron and nickel, are coated with ethylene glycol and used for comparison. The conductivity of nanoparticles when pressed into pellets remains similar for both commercial and synthesized samples. However, when deposited on a strip line and heated, synthesized Ni, FeNi, and Fe nanoparticles show significant conductivity and interesting magnetic properties. It is demonstrated that the nanosize facilitates sintering at reduced temperatures and the capping agents prevent oxidation, resulting in promising conductive fillers for printed electronic applications.
Article
Full-text available
Dyes are one of the most threatening toxins released from industry. Methylene blue (MB) is one of the extremely common dyes in the textile industry. Being complicated in structure and non-biodegradable in nature, removing MB from the aqueous environment is a great challenge. Elimination of dye is usually performed in two ways, degradation, and adsorption. Iron-based nanoparticles, being biocompatible and non-expensive, became a hot field of research in the context of the elimination of toxic dyes. In our review article, we consolidated the data about the synthesis, nature, state, and applications of iron-based nanoparticles to remove MB dye from aqueous solutions specifically via adsorption and degradation. We also reviewed the effect of doping on nanoparticles and their effects on dye removal capacity. Physiological factors such as pH, and temperature play an important role in iron-based nanoparticle synthesis as well as dye degradation and adsorption. A comparative account between adsorption and degradation was tried to depict the elimination of dye in various aspects including efficiency and mechanism. Graphical Abstract
Article
Full-text available
Azo dyes are known to cause environmental pollution and freshwater contamination, posing carcinogenic risks to human health. Consequently, researchers have directed their attention towards studying effective methods for removing these dyes from textile effluent. Amorphous alloys have emerged as a promising candidate for enhancing degradation efficiency on azo dyes, with various metallic alloys undergoing extensive investigation. Notably, the novel catalytic metallic materials exhibit significantly higher degradation performance compared to traditional options. Among the promising candidates are nanometallic alloys based on Fe, Mg, Co, Al, and Mn, which offer improved potential for cleaning diverse types of azo dyes from wastewater. The purpose of this review is to provide an update of the research regarding the use of amorphous alloys in azo dye degradation. The different production methods of amorphous alloys were discussed. The amorphous alloys used in dye degradation studies were subcategorized into three main types: Fe‐based, Mg‐based and some miscellaneous‐based amorphous alloys. Additionally, the study delves into the impact of crucial parameters, such as solution pH and initial dye concentration, providing valuable insights for the efficient treatment of wastewater.
Article
Full-text available
Nanoscale zero-valent iron (NZVI) was synthesized and tested to remediate DDT (Dichlorodiphenyltrichloroethane ) from the contaminated water. The influence of experimental variables such as reaction time, NZVI particle dosage and pH were studied. Mixing an aqueous solution of 3 mg L �1 DDT with 30 mg L -1 Fe 0 resulted in 99.2% loss of DDT within 4 h. Solvent extraction of the Fe 0 revealed that DDT removal was not through adsorption and as Fe 0 treatment of DDT lead to new chromatographic peaks (degradation products) in GC analysis and chloride release, removal occurred through dechlorination. GC/MS analysis confirmed the formation of completely dechlorinated hydrocarbon skeleton of DDT namely, diphenylethane (DPE), as the end product in this treatment; thereby implying the removal of all five chlorine atoms (three alkyl and two aryl) of DDT.
Article
Full-text available
Chromium is an important industrial metal used in various products/processes. Remediation of Cr contaminated sites poses both technological and economic challenges, as conventional methods are often too expensive and difficult to operate. Zero valent iron, an important natural reductant of Cr (VI), is an option in the remediation of contaminated sites, transforming Cr (VI) to essentially nontoxic Cr(III). In the present investigation, an attempt is made to study the efficiency of Fe0 nanoparticles in remediation of Cr contaminated waters. Zero-valent iron (Fe0) nanoparticles were synthesized, characterized, and were tested for removal of Cr (VI) from the water spiked with Cr (VI). Fe0 nanoparticles were synthesized by ferrous sulphate by the reduction of sodium borohydride. The removal efficiency of unstabilised nano Fe0 was compared with Carboxy Methyl Cellulose stabilized Fe0 nano particles. It is observed that the CMC stabilizes the nanoparticles by accelerating the nucleation of atoms during the formation of Fe0 nanoparticles and subsequently forms a bulky and negatively charged layer via sorption of CMC molecules on the Fe0 nanoparticles, thereby preventing the nanoparticles from agglomeration. When a dose of 0.2 g/L of CMC- Fe0 was used for a sample of Cr(VI) (40 mg/L) 100% degradation was observed but the degradation was only 50% when proceeded with unstabilised Fe0 nano particles. The Cr (VI) removal efficiency was decreased significantly with increasing initial pH. Thus the Iron nanoparticles stabilized with CMC are of a good choice for the remediation of heavy metals in groundwater.
Article
Magnetite nanoparticles were synthesized by sol-gel method. The highly crystalline nature of the magnetite structure with diameter of around 10 nm was characterized with transmission electron microscopy and X-ray diffractometry. The surface area was determined to be 198 m2/g. Batch experiments were carried out to determine the adsorption kinetics and mechanism of Cr(VI) by these magnetite nanoparticles. The Cr(VI) uptake was mainly governed by physico-chemical adsorption. The adsorption process was found to be pH and temperature dependent. The adsorption data fit well with the Freundlich isotherm equation. The Freundlich constants were calculated at different temperatures. The adsorption capacity increased with rising temperature. Preliminary results indicate that magnetite nanoparticles may be used as an adsorbent for removal of Cr(VI) from wastewater.
Article
The object of this study was to investigate the capability of nitrate removal by nanoscale zerovalent iron (nZVI) particles in a re-circulated flow reactor of 4.3 L liquid volume. Nitrate removal of 70% was observed with the initial pH of 6.4 with the solution being pre-purged by N gas (anoxic 2 condition) at a rate of 50 mL min-1. However, when the solution was bubbled with CO2 (anoxic condition), the nitrate removal efficiency was less than 5% with the initial pH of 4.0. As the pH was adjusted by H2SO4 and NaOH to the range of 6-8, nitrate removal efficiency of 80-84% could be obtained. Such results demonstrate that carbonic acid generated from CO2 bubbling may play strongly 2 an inhibiting role in nitrate removed by the nZVI. In addition, based on nitrate mass balance, the end product was mostly dominated by ammonium at the initial pH 10. For the initial pH such as 6 and 8, some other nitrogen-containing species have been formed in the reaction, in addition to nitrite and ammonium. © 2011, Chinese Institute of Environmental Engineering. All rights reserved.
Article
Groundwater remediation by nanoparticles has received increasing interests in recent years. The present work was conducted to investigate the feasibility of using new class of stabilized zerovalent iron (ZVI) nanoparticles for in situ reductive immobilization of Cr(VI) in water. The nanoparticles were prepared by chemical reduction method using Ferrous sulfate heptahydrate (FeSO 4.7H 2O) with sodium borohydride (NaBH 4) as a reducer. This sample was subjected to XRD for the determination of crystallinity and average particle size. The size of the particle is 12.4 nm. The morphology of the nanoparticles was observed using SEM. Batch experiments reported that the synthesized starch stabilized zerovalent iron nanoparticles were able to rapidly reduce Cr(VI) in aqueous solution. The extend of Cr(VI) reduction was increased with increasing concentration of iron nanoparticles (0.1 g/L - 0.4 g/L) and inversely with initial Cr(VI) concentration (10 mg/L - 25 mg/L) as well as with decreasing pH (3 - 10). The equilibrium data obtained from the experimental results were then fitted to the Freundlich and Langmuir isotherm models. Kinetic models were examined with pseudo first order rate reaction. The Correlation coefficient between experimental parameters and time shows that there is a strong positive correlation for Cr(VI) reduction. This study indicates that the zerovalent iron nanoparticles, especially those which were starch-stabilized can yield a high removal efficiency in the reduction of Cr(VI) contaminated groundwater.
Article
Nanoscale zerovalent iron (nZVI) was used to remove phosphate from aqueous solution, and the influence of pH, ionic strength and coexisting anions on phosphate removal was investigated. The results agree well with both Langmuir model and Freundlich model, and the calculated maximum adsorption capacity of phosphate was 245.65 mg/g, suggesting significantly higher and excellent uptake of phosphate by nZVI. The removal of phosphate obviously decreased with an increasing pH due to the isoelectric point (IEP) of nZVI, but exhibited no change with ionic strength and coexisting anions (except carbonate anion). The microstructure of fresh and reacted nZVI was characterized by FT-IR, XRD and XPS and these results indicated that no redox reaction occurred to the P(V). These observations shed light on the mechanisms of phosphate removal were mainly adsorption and coprecipitation processes. The higher uptake of phosphate indicates that nZVI was a suitable and effective material for phosphate removal.
Article
Groundwater remediation by nanoparticles has become a major interest in recent years. This report presents the ability of hexavalent chromium removal in aqueous using iron (Fe 0) nanoparticles. Cr(VI) is a major pollution of groundwater and more toxic than Cr(III). Fe 0 reduces Cr(VI), transforming Cr(VI) to nontoxic Cr(III). At a dose of 0.2g/l iron (Fe o) nanoparticles, 100% of Cr(VI) 5mg/l was removed after only 20 minutes. The Cr(VI) removal efficiency increased with decreasing initial pH. Synthesized Fe 0 nanoparticles were compared iron powder in the same conditions. The results show that Fe 0 nanoparticles are more efficient than Fe powder. The final product of the reduction process Cr(VI) was Cr(OH) 3 . It was concluded that iron nanoparticles are a good choice for the removal of heavy metals in water.