Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Groundwater for Sustainable Development 15 (2021) 100694

Contents lists available at ScienceDirect

Groundwater for Sustainable Development


journal homepage: www.elsevier.com/locate/gsd

Review article

Groundwater remediation using zero-valent iron


nanoparticles (nZVI)
Abdul Mannan Zafar a, 1, Muhammad Asad Javed a, 1, Ashraf Aly Hassan a, b,
Mohamad Mostafa Mohamed a, c, *
a
Department of Civil and Environmental Engineering, United Arab Emirates University, Al Ain, P.O. Box 15551, AlAin, UAE, United Arab Emirates
b
Department of Civil and Environmental Engineering, University of Nebraska-Lincoln, Lincoln, NE, USA
c
National Water and Energy Center, United Arab Emirates University, Al Ain, P.O. Box 15551, AlAin, UAE, United Arab Emirates

A R T I C L E I N F O A B S T R A C T

Keywords: The underground water environment and ecology are considerably threatened by groundwater
Nanoparticles contamination. Several new remediation strategies and techniques have been implemented dur­
Nano zero-Valent iron ing the past two decades to eliminate contamination. These strategies and techniques have been
Groundwater continuously improved through various methods including the augmentation of different nano­
Pollutants particles and modified composites. These nanoparticles are useful for the remediation of
Remediation
groundwater contaminated by both organic or inorganic pollutants. This paper focuses on the
Removal efficiency
contaminant removal efficiency of a particular type of iron nanoparticle, i.e., nano zero-valent
iron (nZVI) particles. Further, this paper details the usage of simple nZVI and modified nZVI in
the context of contamination removal based on their dosages. It also describes the effects of
coatings of different metallic and metallic oxide compounds used in manufacturing modified
forms of nZVI. Results show that nZVI particles have almost 80% removal effeciency of both
organic and inorganic pollutants. Further, the usage of modified nZVI particles, such as carbox­
ymethyl cellulose (CMC)–Carbo-Iron® colloids (CICs), nano-Fe, and sodium alginate nZVI par­
ticles, resulted in contaminant removal of almost 100%. Compared to nano-Fe, CMC–CIC yielded
more significant results in groundwater remediation at low dosages. Thus, this paper proves that
even a few nanoparticles (NPs) chosen appropriately are sufficient to remediate groundwater. The
results of this study can help improve the synthesis of modified nZVI.

1. Introductions
Several methods have been adopted for groundwater remediation. Innovative methods for groundwater restoration include
improved steam extraction, thermal conduction, use of surfactants, enhanced bioremediation reduction, permanganate application,
and use of permeable reactive (sorption) barriers (Phenrat et al., 2019, Mohamed and Hatfied 2005; Mohamed et al., 2006, Mohamed
et al., 2007, Mohamed et al., 2010a,b,c; Mohamed and Hatfied 2011). Nanoparticle (NP)-based processes have been modified and
applied in various scientific fields to achieve different purposes. More than 59 NP-based technologies are available for groundwater

* Corresponding author. Department of Civil and Environmental Engineering, United Arab Emirates University, Al Ain, P.O. Box 15551, AlAin, UAE, United Arab
Emirates.
E-mail addresses: 201990200@uaeu.ac.ae (A.M. Zafar), 201990206@uaeu.ac.ae (M.A. Javed), alyhassan@uaeu.ac.ae, ahassan@unl.edu (A.A. Hassan), m.
mohamed@uaeu.ac.ae (M.M. Mohamed).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.gsd.2021.100694
Received 6 March 2021; Received in revised form 20 September 2021; Accepted 23 October 2021
Available online 27 October 2021
2352-801X/© 2021 Elsevier B.V. All rights reserved.
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

remediation. However, the most viable and robust technology is the use of nano zero-valent iron (nZVI) particles (Teefy, 1997).
Because of the small pore size, ZVI is adopted as a passive strategy for treating pollutants at depth in groundwater and facilitates
groundwater remediation. Additional experiments have to be conducted, and unique mechanisms have to be developed for nZVI
particles to improve the efficiency of metal NPs and eliminate multiple (e.g., organic and inorganic) contaminants.
The synthesis of iron nanoparticles via the Fe(II) and Fe(III) borohydride reduction was first demonstrated more than 50 years ago
(Smith and Rodrigues, 2015). In 1990, ZVI was added to a permeable reactive barrier to micrometer-sized iron particles (Gillham and
O’Hannesin, 1994; Reynolds et al., 1990). The application of nZVI to remove environmental contaminants was first suggested by Wang
and Zhang (1997). nZVI is an exceptional electron donor that can transubstantiate the reduction potential (E0H ) of contaminants greater
than − 0.447 V (Oppegard et al., 1961). The process involves the removal of organic compounds to achieve dechlorination using
metallic NPs to degrade compounds such as tetrachloroethylene (PCE), vinyl chloride (VC), and trichloroethylene (TCE). A metal E0H of
less than − 0.41 V can be mainly attributed to the sorption/surface complexation resulting from detoxification by nZVI (Li and Zhang,
2007).

Fe+2 + 2e− → Fe0 EH0 = − 0.447V

As the nZVI particles are extremely small, the surface atoms form quite a significant fraction of the NPs, and the surface energy is
excessive, and the surface ratio volumes are high (Wang and Zhang, 1997). Therefore, contaminants decay or are immobilized by mass
remediation agents rather than by bulk items. Facilities for in situ injections of the particles for field remediation are also feasible
because of the size of the nZVI (Elliott and Zhang, 2001; Schrick et al., 2004). Elliot and Zhang (2001) used bimetallic nZVI particles for
the first field trial because of their ease of synthesis and reactiveness. In this analysis, removal efficiency of 96% was achieved in four
weeks for TCE, which is the highest value observed for nZVI. This field test used uncoated bimetallic nZVI, and nZVI dispersion was
decreased to 1.5 g/L to minimize nZVI clogging of the surface pores via aggregation. Soon after the initial field analysis, scientists
realized that nZVI was not sufficiently effective as an in-situ remediation agent with high reactivity and limited size. nZVI often must be
readily dispersed in water to allow scattering by a water-saturated pore medium to the contaminated region and can be added to
contaminants, such as dense nonaqueous phase liquids (DNAPLs), at a higher NP concentration (3.5–20 g/L) for efficient sourced
groundwater remediation.
The mobility of the uncoated nZVI was determined to be limited in saturated pore media (i.e., the reasonable transport distances for
uncoated particles were just a few centimeters, i.e., approx. 0.05–0.5 cm). Researchers assumed that either deposition or accumulation
or maybe both processes decreased the mobility of nZVI. Schrick et al. (2004) first suggested a carbon-based NP synthesis in 2004. In
2005, the Lowry and Tilton Group suggested using triblock copolymers and homopolymers as a surface modifier for pre-synthesized
nZVI (Saleh et al., 2007). Polymeric surface adjustment not only allows for electrostatic refraction but also limits the number of
functional groups targeting nonaqueous phase liquids (NAPLs) by clustering and accumulation (Phenrat et al., 2009b, 2010; Phenrat
and Lowry, 2009).
In 2005, a research group suggested the single-pot synthesis of nZVI-modified carboxymethyl cellulose (CMC) to increase the
mobility of nZVI in the saturated zone (He and Zhao, 2005). In this synthesis, nZVI was supplied to the NAPL source area to produce
emulsified zero-valent iron (EZVI). The EZVI emulsion contained oil and surfactants, which increased biodegradation, enhancing
abiotic degradation (Quinn et al., 2005). The first synthesized nZVI (modified) was prepared by John’s research group, where the nZVI
was encapsulated in porous silica particles (Zhan et al., 2008; Zheng et al., 2008). Phenrat and Lowry (2009) demonstrated the
magnetic characteristics of nZVI, considering its ability for rapid agglomeration and reduced mobility in pore media. Subsequent to
these studies, nZVI accumulation and porous media transportation gained importance in research and development on nZVI. In
addition, polymeric surface improvement, including making changes to the CMCs, became a standard laboratory study preparation
method.
In 2006, the possibility that nanotechnology may cause harmful effects became a concern (Wiesner et al., 2006). Several research
groups studied the fate and transport of NPs, focusing on the impact of NPs on the consumer end and their toxicity (Saleh et al., 2007).
The mobility of polymeric nZVI particles at low concentrations was essential for assessing their environmental consequences after their
deployment for remediation. The first report on the quantification of nZVI mobility (with low particle concentration) by the clean bed
filtration theory was published in 2007 (Saleh et al., 2007). The electrostatic stability of low-particle (30 mg/L) polymer-modified nZVI
particles resists nZVI particles adhesion to aquifer materials, contributing to the potential mobility of modified nZVI of over tens to
hundreds of meters across unconsolidated sandy aquifers. The nZVI polymer agglomeration is minimal at low particle concentrations,
and the conventional filtration theory can effectively model the transportation behavior of the nZVI. In 2010, the environmental
tradeoffs of in situ nZVI remediations were first assessed (Grieger et al., 2010); this study reported no evidence of a potential threat to
the environment using nZVI particles. However, they also found that the onsite experiment’s effect on human health was mostly
unknown because they lacked adequate data for environmental concerns (i.e., the ability to remain active/nonactive, bio­
accumulation, and toxicity). Advances have been achieved in terms of both the reactivity and location of nZVI applications. The first
studies on the use of nZVI for treating nZVI source areas and the combination of nZVI and bioremediation were published in 2010
(Taghavy et al., 2010; Xiu et al., 2010). Taghavy et al. (2010) observed 30% conversion of PCE to ethane and 5% DNAPL removal when
60 g/L of nZVI was provided along with sand. Within and below the DNAPL source region, the positioning of nZVI improved the
removal efficiency of PCE, which reduced the PCE concentration in the effluent. The removal efficiency of the DNAPL area was a
critical characteristic for estimating the flow rate and location of the field site.
Around 2010, some studies were performed on the potential effect of bioremediation by nZVI. In these studies, the possible
combination of nZVI with bioremediation or natural attenuation was examined. Xiu et al. (2010) suggested that nZVI could serve as a

2
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

biostimulating reagent related to H2 production and an inhibitor of cell surface contact during anaerobic corrosion. Although nZVI
significantly stimulated methanogenic species, nZVI initially inhibited TCE-dechlorinating bacterium but subsequently recovered after
partial oxidation. Coating nZVI with sulfur (FeS) was also a promising solution to nZVI, a vital obstacle to practical in situ remedial
nZVI for nonselective oxidation. In 2013, for example, sulfurization of nZVI was regarded as a viable approach to increase its reactive
durability because sulfurization would inhibit the Fe0–H2O response but preserve substantial reactivity to contaminants such as
trichloroethylene (Fan et al., 2013, 2016). Semerád et al. (2020) investigated that sulfidated nZVI particles were more stable than
unmodified nZVI in terms of Fe0 content and diameter size. Their results found that α-Fe0 content was higher in sulfidated nZVI
(78.5%) wrapped with only 5% of sulfur than unmodified nZVI (35.3%), enhancing nanoparticle aging for treatment.
Researchers proposed that nZVI should be used for treating vadose zones in addition to the saturated area. Ding et al. (2013)
showed that the first nZVI foam-based vadose rehabilitation synthesis solves the challenge of dispersing nZVI-based water into the
vadose region, leading to the unexpected suspension and relocation of chlorinated volatile organic compounds (CVOCs) in aquifers.
Although the vadose zone has inadequate water or suitable CVC in the pore water for substantial reductive nZVI dechlorination, LF
EMF foam-based nZVI can be applied for soil vapor extraction (SVE), and thermally enhanced SVE for restoring the vadose zone may
serve as a collective remedy. Foam-based nZVI (41 g/L) in the liquid phase resulted in heat generation to reach temperatures up to
77 ◦ C within 15 min under LF EMF conditions with a frequency of 150 kHz and a current of 13 A. In contrast to reactors in which LF
EMF is not used, a forty-fold increase in volatilization was achieved for the TCE–DNAPL in the vadose zone. While these latest nZVI
applications are still under development, they have the potential for field applications. nZVI has been used in pilot and field appli­
cations since 2001. In 2009, 44 carcinogenic VOC-polluted sites were recovered using nZVI (Karn Barbara et al., 2009). In 2013, 59
nZVI treated sites were accomplished (Bardos et al., 2014). A survey conducted in 2017 reported that nZVI remediation sites
contaminated with CVOC reached 77 except for one arsenic-contaminated site, and these sites were located in the USA, Canada,
Taiwan, and Europe (Karn Barbara et al., 2009).
In this review, the efficiency of the removal of organic and inorganic pollutants was estimated from the data collected in the last
decade, i.e., 2010–2020. The primary focus was to determine the effect of the nZVI and modified nZVI particles on different organic
and inorganic pollutants. Although, the dosage of NPs applied to the groundwater contaminants affects the removal efficiency.
Therefore, the dosage effect in the application of NPs to decontaminate the polluted sites was estimated. With regard to data collection,
few studies were onsite injection-based studies on nZVI particles in field experiments, whereas some were laboratory-based studies.
Moreover, contact time is an essential factor because of the significant gaps between the times for different studies; however, the
contact time was neglected in the previous works. Literature was collected from Google Scholar, the Web of Science, and Science Direct
by applying keywords such as “the effect of nZVI in groundwater-contaminated sites” and “groundwater remediation by nanoparticles
and nZVI nanoparticles. This review is mainly covered by comparing the nZVI and modified nZVI in field applications and lab-scale
studies. In this review, a detailed quantitative analysis has been performed on the data collected from the previous studies conducted
on nZVI and modified nZVI particles to estimate the removal efficiencies. This review article also discusses the environmental risk
factors once nZVI particles are injected into the groundwater.

2. Types of nZVI particles


Inorganic nanomaterials with organic contaminants (e.g., metal/metallic oxide/metallic sulfide NPs and carbon-based nano­
materials) can be classified as organic nanomaterials, and they are commonly used for groundwater remediation. Organic nano­
materials (e.g., organic metallic structures, nanomaterials with nanomembranes/films, and polymeric organic compounds) are usually
based on molecules. Oxidation, degradation, and adsorption are the most common methods for the remediation of organic pollutants in
water (Lu and Astruc, 2020).
For many years, the focus was on remediating the sites contaminated by pesticides and halogenated organic pollutants using
metallic NPs. Considering the synergy effects of some transition metals, monoelementary transition metal NPs, such as Fe, Cu, Au, Ag,
and Co, and elementary metallic NPs, such as Fe/Ni and Fe/Cu, are being used for groundwater remediation (He et al., 2007; Zhu and
Lim, 2007). With regard to catalytic degradation, bimetallic NPs have shown remarkably better removal performance than the
monometallic NPs.
The behavior, particle size, and stability of monometallic/bimetallic NPs play critical roles in determining the efficiency of the
organic contaminant remediation of water. The palladium (Pd) NPs developed biologically in recent years are used to remediate large-
scale recalcitrant and toxic organochlorine (c-hexachlorocyclohexane) contaminants in groundwater (Hennebel et al., 2011). To avoid
accumulation and passivation of nZVI NPs, several dispensing agents and supporters, such as activated carbon, zeolites, porous silicon
oxides, and organic molecules, were tested to stabilize the transitional metal NPs (Lu and Astruc, 2020).

2.1. CMC-modified nZVI


CMC-stabilized Fe–Pd NPs (containing 0.1% by weight of NPs and 0.1%–0.6% by weight of CMC) was used to remove groundwater
contaminants such as VOCs: TCE (1.6–23.8 mg/L), PCE (1.2–12 mg/L), VC (1.1–2.2 mg/L), cis-DCE (8.5–20 mg/L), and PCB (6.9–97.4
μg/L) (He et al., 2010). In another study, TCE and PCE were treated using CMC-modified nZVI (Bennett et al., 2010; Kocur et al., 2014,
2015). These NPs were also used for assessing the injection and dispersion in groundwater in a controlled test-scale study (Kocur et al.,
2016).

2.2. CMC-modified Carbo-Iron® colloids (CICs)


CMC-modified nZVI was used to target aromatic hydrocarbons such as benzene (approximately 20 mg/L), methyl tert-butyl ether

3
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

(MTBE, approximately 3.9 mg/L), and ammonium (approximately 55 mg/L) (Busch et al., 2015). Another study reported that one ton
of CVOCs and mostly PCE was treated by CMC–CIC-modified nZVI (Mackenzie et al., 2016).

2.3. Pd-modified nZVI


Pd-modified nZVI forms a composite known as Iron-Osorb®. It is swellable organosilica Osorb and is organophilic with 2%–10% by
weight of nZVI mixed with 0.02%–0.1% by weight of palladium acetate. This composite was used in three injection phases to treat TCE
at concentrations of 0.5–1 mg/L (Edmiston et al., 2011). Similarly, another study reported that Pd-coated nZVI was applied via two
injection rounds to treat VC, 1,2-DCA (1,2-dichloroethane), 1,1-DCA (1,1-dichloroethane), 1,1-DCE, cis-1,2-DCE, and TCE (Wei et al.,
2012).

2.4. NANOFER
The NANOFER treatment of PCE, TCE, and DCE with concentrations of up to 70 mg/L were performed in 82 pumping wells under a
pressure of 0.8 MPa (Mueller et al., 2012). Another study reported the injection of nano-Fe–nZVI in a liquid state with food-grade
inorganic material to remediate soil containing PCE (2.7 g/kg) and shallow groundwater containing PCE (41 mg/L) (Jordan et al.,
2013). Similarly, the remediation of CVOCs (VC, DCE, PCE, and TCE) at a maximum contamination concentration of 79.9 mg/L by
NANOFER STAR (87% by weight of FeO) has been reported (Lacina et al., 2015). Yang et al.(2021) studied the effect of NANOFER
STAR nZVI particle to decontaminate the TCE, which were chemically synthesized with sodium nitrate in a lab scale experiments. The
study reported upto 90% removal of TCE with the passivation corrosion of NaNO3 within 14 days of treatment.

2.5. nZVI
nZVI particles are usually composed of iron ores (particle size: 70 nm) stabilized by polycarboxylic acid. The use of nZVI (30 g/L)
and 2.5-μm iron (60 g/L) to remediate PCE (1–2 tons) has been reported (Mueller et al., 2012). Flake-like nZVI (85% FeO) particles
coated by polyethylene glycol were studied to treat PCE (20–50 mg/L; 10–14 m below ground surface (bgs) were used) (Köber et al.,
2014). Similarly, a variant of nZVI in monoethylene glycol (FerMEG12) was used to remove CVOCs, including PCE, TCE, and hexa­
chloroethane, at concentrations of up to 20 g/kg (Stejskal et al., 2017). Another study used the nZVI particle (12 g/L) on field scale
application (Gangwon-do road, Wonju, Korea) to remediate the site contaminated from TCE (Ahn et al., 2021).

2.6. nZVI-biochar
Another modified form of nZVI applied in the studies is a use of biochar to encapsulate the nZVI particle to increase the stability of
Fe0 particle for the transfer of electrons. Qian et al. (2020) used nZVI particle in their first injection on-site and in the second injection
they used a mixture of nZVI mixed with biochar. 200 kg of nZVI was applied in first injection well to the total volume of 100 m3 in

Fig. 1. a) Process of nZVI particle injection through a well. nZVI particle reaction to the pollutant (e.g., nonaqueous phase liquids (NAPLs)) in the groundwater-
contaminated sites. b) Generalized illustration of the adsorption, precipitation, coprecipitation, oxidation, and reduction process of nZVI. c, d) Example of nZVI
and modified nZVI particle for dechlorination mechanism of trichloroethylene (TCE [C2HCl3]) into acetylene and hydrogenation of ethene or ethane.

4
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

suspension form to reach the final concentration of 10 g/L. In the second injection, 100 kg of nZVI and 200 kg of biochar were mixed
together to form a suspension of 50 m3 of volume. Semerád et al. (2021) g 1 g of hematite, 5 g of sawdust and 20 mL of water on a
homogenizer for 5 min. Sawdust was further dried at 60–80 ◦ C. And then the mixture was pyrolyzed in a furnace at the temperature of
700 ◦ C to form the nZVI-biochar NPs. The study evaluated the nZVI-biochar NP on the groundwater contaminated from TCE having a
concentration of 47.5 mg/L 0.7 g/L of nZVI-biochar concentration was applied in the lab-scale study to evaluate the effectiveness of
nZVI-biochar NPs to decontaminate TCE.

3. A brief mechanism of nZVI and modified nZVI particles treatment in the groundwater
The field-scale application and definition of the field mobility of the nZVI particles are depicted in a. nZVI particles were injected
into the ground through wells. NPs have shown different treatment levels for the contaminant plume: the NPs were grouped to form
strains or clusters with solid particles (i.e., sand). In some cases, aggregation similar to that in microbial colonies was observed. Such
aggregates may dissociate to provide single (Fe0) particles. Attachment and detachment continuously occur in the groundwater
regime, while NPs are present in the saturated zone. Generally, nZVI particles are also charged that they tend to accumulate with
contaminants, such as NAPL, as shown in Fig. 1a and b illustrates the generalized mechanism of reduction, precipitation, co-
precipitation, oxidation, and adsorption that could happen in the groundwater regime after nZVI particles are introduced to reme­
diation sites. Fig. 1c and d demonstrate the dechlorination mechanism pertaining reduction efficiency of an iron particle (Fe0). Fig. 1c
shows the nZVI (Fe0 particle) distribution of electrons of around 92% to reduce into H2 gas from hydronium ions (H+). However, only
8% of electrons are transferred to decontaminate trichloroethylene (TCE [C2HCl3]) and reduced to acetylene (C2H2) using unmodified
nZVI. While, on the other hand, in Fig. 1d, modified nZVI (Fe0 particle coated with sulfur) shows 87% electrons transfer to dechlorinate
the TCE to acetylene. At the same time, the remaining 13% of electrons are transferred to produce hydrogen gas. The production of
acetylene reaction is a slow process, and due to enough hydrogen available in the water, a hydrogenation reaction takes place to
produce ethene (C2H4) or ethane (C2H6), as shown in Fig. 1d. The process is described as an example to compare the procedure of nZVI
particle and modified nZVI particle reduction potential to remediate the contaminated sites from NAPL.

4. Performance of nZVI and modified nZVI for different pollutants


As mentioned earlier, nZVI and modified nZVI NPs have been used to treat sites contaminated by organic or inorganic particles. In
this study, the collected data, which include data on different types of pollutants, i.e., organic (e.g., TCE, PCE, and VOC’s), inorganic
(heavy metals such as Cr(VI) and As), and nitrates, were analyzed. The NPs performed well in terms of onsite decontamination or
laboratory experiments, achieving maximum output for the nZVI and modified nZVI NPs. The NP performance for different persistent
organic and inorganic pollutants is shown in Fig. 2 to understand the overall nZVI NP performance.

4.1. nZVI
Iron is the most abundant element in the environment, and it plays a vital role in several contaminants. The standard two states of
iron valence are the highly water-soluble Fe-II (ferrous iron) and the highly water-insoluble Fe-III (ferric iron). Ferric iron is stable in a
neutral to alkaline pH environment and oxygen-rich environments. ZVI or Fe0 is also present under rigorous conditions such as those
related to metamorphous rocks and meteorites under varying environmental and geological conditions. However, owing to its high
reactivity, ZVI is seldom found on Earth’s surface. Iron is also one of the most critical tracing components for biological processes
owing to its variable position—it can coordinate with O2, N2, and S atoms and bind with small molecules. It also plays a crucial role in
specific interactions within living organisms. As an electron donor, iron plays an essential role in pollutant mobility, sorption, and

Fig. 2. Overall performance of nZVI particles in removing inorganic and organic persistent pollutants in different studies collected.

5
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

degradation in the atmosphere during the oxidation of Fe2+ to Fe3+.


Several nZVI systems have been explored until now for groundwater remediation. At a study site in Bornheim, Germany,
byproducts (e.g., TCE and DCE) were reduced by approximately 90% of CVOCs. After two years of injection, no rebound has been
observed, and the downward trend of pollutants still continues (Mueller et al., 2012). For in situ control, five new nZVI systems were
used for perchloroethylene. DP liner soil sampling was used for one year of surveillance to analyze the nZVI radius of influence (ROI).
PCE concentrations decreased from 20 to 30 mg/L to 0.1–2 mg/L in the well source. Mechanical mobilization from the close by
PCE–NAPL sources may explain the initial increase in PCE concentration after nZVI injection (Köber et al., 2014).
Further, the contaminant sampling concentrations within the injection zone decreased by 49%–89% for PCE, by 81%–97% for TCE,
and by 97%–98% for hexachloroethane at the sampling level within three months of injection. After one year of injection, there was no
detectable minor rebound, and a high ethane concentration was observed because of dechlorination. In this regard, the fact that a high-
volume injection of FerMEG12 suspension likely led to source dissolution and enhanced contaminant dissolution into groundwater
may have played a significant role (Stejskal et al., 2017).

4.2. Modified nZVI


Fe amorphous oxyhydroxides (HFOs) are mainly precipitated and permeate due to their strong reactivity attributed to their high
specific surface area. HFOs are mostly considered potent adsorbents of various pollutants (Cundy et al., 2008; Cundy and Hopkinson,
2005). The disposal of contaminants from waste streams is a method that can be used with the precipitation of (hydrous) ferric oxides
for a range of industrial processes; for example, high-density sludge systems are used in mining and treatment of textile dye effluents
for arsenic control. However, HFO has been developed over the years to realize more orderly transformations (crystalline), such as
transformation to goethite or hematite (Phenrat et al., 2009a) with relatively smaller surface areas, and they are usually less reactive
and efficient than sorbent substrates. For radionuclides, either surface adsorption, coprecipitation, or absorption into the Fe-oxide
structure may be applicable. For a host of contaminants, non-oxide Fe phases (e.g., sulfides, carbonates, and phosphates) are often
used as essential sorbents for (co) precipitants. Nevertheless, the long-term preservation of iron oxyhydroxide and oxides of radio­
nuclides and pollutant metals depends on the pH and oxidation-reduction potential (Eh) because chemical and microbial reduction or
increased acidity reactions may result in the disintegration solubilization of iron. The reduction of the associated Fe-oxide emissions
from organic sediments in Bengal, India, has led to substantial groundwater depletion and significant community health
complications.
nZVI sulfide particles are mainly introduced in groundwater for the treatment of chlorinated pollutants. The small size (approx­
imately 5 μm) of ZVI particles provides a larger surface area. Contaminant removal efficiency is high because of the excellent reactivity
and distribution combined with good chemical-reducing agents. The thin layer of iron sulfide imparts several advantages for treatment
with each type of ZVI particle. The reactions of chlorinated ethenes are strengthened by sulfidation, minimizing the passivity by water
reactions (TCE-kinetics = 30 × barley ZVI) and increasing the persistence of the treatment. The amount of daughter products is
reduced, and more than 90% of the TCE is transformed into ethene/ethane, as shown in Fig. 1.
Therefore, the contaminant sorption for iron sulfides, carbonates, and phosphates is highly dependent on and may even be
reversible under local Eh/pH conditions. The release of metals during pyrite oxidation and marcasite-rich mining waste and exposure
to air and water during mining procedures is a notable example in the case of iron sulfides. This may contribute to the persistent,
widespread drainage of acid mines and associated traces of metal enrichment, leading to substantial environmental degradation
(Blodau, 2006; Kalin et al., 2006) through iron oxidation (and the oxidation of other metal sulfides) in combination with ferrous iron
oxidation and hydrolysis. Many of the reactions of iron in soils, sediments, and soil waters are microbial, and Fe2+ is the predominant
microbial respiration electron transmitter in many surface habitats. Restricted and discharge contaminants, for example, were shown
to be microbe-mediated processes involving Fe. Several high-priority contaminant metals, such as U, Cr, and Tc, can be reduced by the
enzyme dealing with Fe3+ activity reduction in microorganisms.
Recent rapid advancements in nanotechnology have led to substantial improvements, such as improved reactivity, in the use of
iron-based NPs for soil and groundwater treatment. Because of the high surface area and enhanced mobility in iron NPs (1 nm = 10− 9
m), faster and more economical contaminant removal compared to the standard iron-based technologies (U.S. EPA Nanotechnology
White Paper, 2007) can be achieved. Laboratory experiments have demonstrated the utility of strengthening NPs predominantly with
ZVI as a bulk reducer of various organic compounds such as chlorinated methane, ethane, benzene, and biphenyls. Quinn et al. (2005)
also remarked that ZVI NPs could handle and remediate dissolved solvents such as DNAPLs.
The used particles are typically pumped directly into the slurry. A subsurface region can be suspended in the hydrophobic fluid to
remove particle agglomerations and increase reactiveness. The mobility helps decontaminate groundwater plumes and the main
contaminant source areas by injecting an emulsion without any digging. Typically, NPs are composed of ZVI coated with catalytic
metals such as Pd and Pt. In some methods, the magnetic properties of iron minerals, such as magnetite (Fe3O4), are employed to create
a form of insulation after treatment. Significant reduction in soil levels (>80%) for the DNAPL trichloroethylene (TCE) contaminant
and significant decreases of TCE in groundwater (60%–100% decrease) was observed during field experiments in the United States
(Hara et al., 2006; Quinn et al., 2005). However, some positive results, specifically, the decrease in TCE for some test well injections,
were generally less efficient than those achieved using zero-valent NPs; this result may be attributed to the premature passivation of
the NPs used or the use of an inadequate scale such as a large iron-to-soil ratio (Gavaskar et al., 2005).
PCE and TCE degradations were the highest in the case of CMC-stabilized Fe–Pd NPs. Kocur et al. (2016) studied the influence of
CMC-stabilized Fe–Pd NPs by rapid ethene degradation. Preinjection levels recovered to chlorinated ethene concentrations approxi­
mately two weeks after the injection, although the nZVI was drained after two weeks, and long-term CVOC biodegradation was
stimulated (He et al., 2010). Chlorinated ethene was less affected because it was extracted from the injected solution (Bennett et al.,

6
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

2010). Another research showed that 71% of the overall Fe injection breakthrough occurred at 1 m from the injection well. Increased
dechlorination of chlorinated ethenes to nontoxic ethene was also observed over the two years after nZVI injection, indicating
increased bioremediation. This effect has been verified by an increase in the order of magnitude of the region affected by the
CMC-modified nZVI (Kocur et al., 2014, 2015, 2016).
Fig. 3 illustrates the removal efficiency achieved by nZVI and modified nZVI. In some experiments, the performance of CMC-CIC
was visually observed. In a study of an on-site injection well, 5 m away from the injection site, roughly 12% of the removal efficiency
was achieved using the CMC-CIC modified nZVI particles (Busch et al., 2015). These results indicate that Carbo-Iron® is more versatile
than the uncoated nZVI (Busch et al., 2015). Another study found a substantial decrease in PCE in a test well 4 m away from the
injection site over the first 2.5 months of the first injection test. Initially, byproducts, such as VC disappeared entirely, and low
concentrations of DCEs were detected. Over time, PCE concentrations started increasing from the much lower initial value. However,
high ethane and ethene concentrations were observed than expected based on CIC reactivity, indicating biodegradation. After 200
days, the PCE concentration reduced from 19 to 1.5 mg/L (~92.5% removal efficiency) in the second injection test (Mackenzie et al.,
2016).
TCE decreased drastically (>50%) when Pd-modified nZVI was added for remediation. The cleanup criteria (TCE 5 μg/L) were
fulfilled for groundwater after 60 days. Only one component of the chlorinated portion, i.e., cis-DCE, had a concentration of 1–3 μg/L.
The water had low toxin levels for 120 days (Edmiston et al., 2011). A substantial decrease in ORP (approximately 400 mV) was also
cataloged at the injection point, indicating optimal reaction conditions. The decrease in 1,2-DCA reveals the beginning of bio­
stimulation (Wei et al., 2012). The literature-based on nZVI experiments shows that nZVI prepared synthetically in the laboratory is
more reactive than commercial/industrial products of nZVI.
NANOFER is the only reported commercial nZVI product that has been proven more effective than many of the industrial nZVI.
Research showed that the initial contaminant at the Horice site in the Czech Republic decreased by 60%–75% using NANOFER
(Mueller et al., 2012). Another study reported using NANOFER by eliminating PCE within one month, resulting in substantial amounts
of byproducts cis-DCE (concentration 65 mg/L over 12 months) and ethane. The effect was observed from a distance of 2.44 from the
locations where pH and ORP were affected; physical nZVI measurements were not performed (Jordan et al., 2013).

4.3. Effect of NP dosage


Another essential factor in NP application is the number of NPs supplied to the contaminated site. Hence, the numbers of different
types of nZVI and modified nZVI particles supplied into groundwater in the field or laboratory experiments were examined in this
paper. As shown in Fig. 4, unmodified nZVI particles, such as nano-Fe and NANOFER STAR, were used in larger quantities, e.g.,
approximately 21,500 and 2500 mg/L, respectively. In comparison, modified nZVI particles were used in less amount compared to
uncoated particles. For instance, CMC-CIC modified nZVI particles have shown approximately 100% removal efficiency with only
~1200 mg/L of dosage.
Similarly, other modified nZVI NPs, such as graphene oxide-sulfidated nZVI NPs, sodium alginate nZVI, and sulfidated nZVI, were
used in minimal quantities, indicating that modified nZVI was more stable and reactive for groundwater remediation. The stability of

Fig. 3. Analysis of the different modified forms of nZVI particles to remove contaminants (mainly TCE and PCE).

7
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Fig. 4. Effect of nanoparticle dosage on the contaminant removal of organic and inorganic nanoparticles.

modified nZVI particles has more extended aging than unmodified nZVI particles. Therefore the modified nZVI persists for longer
durations in the groundwater regime and can remediate contaminated sites more efficiently. The size, weight, dosage, contact time,
and removal efficiency of the NPs are listed in Table 1. One explanation for the wide range of values is that these studies include field
experiments and laboratory-scale studies. However, a considerably greater supply of NPs is required to remediate the groundwater in
the field because of the substantial groundwater volume involved in onsite remediation. Another reason may be that modified nZVI
NPs have a larger surface area than unmodified nZVI particles; the more significant the surface, the higher is the efficiency. Therefore,
there is much potential in developing, synthesizing, and on-site production of modified nZVI particles.

5. Case studies of nZVI and modified nZVI particles used on-site groundwater remediation
5.1. A case study on investigation of DNAPL contaminant using (modified) sulfidated nZVI particle
Sulfidation of nZVI particles is relatively a new practice in the field of groundwater remediation, which gives more stability and
enhanced electron transfer of nZVI to decontaminate chlorinated contaminants such as DNAPLs (Nunez Garcia et al., 2020). In Ontario,
Canada, Nunez Garcia et al. (2020) studied a contaminated site to dechlorinate the DNAPL. The depth of DNAPL at the site was
measured around 4–5 m bgs. Previously the same site was also trialed by injecting into four boreholes with the 620 L volume of 1 g/L
CMC nZVI particle (Kocur et al., 2014, 2015, 2016). The injection wells were installed with piezometers made up of Teflon tubes of
different lengths and color-coded to monitor well as an adapted methodology for on-site remediation technique. CMC-nZVI particle
was engineered and coated with a sulfur layer by synthesizing CMC-nZVI with sodium dithionite to form CMC-S-nZVI for injecting
wells. The final concentration of suspension was kept at 1 g/L, and a total volume of 620 L was prepared for the on-site remediation.
The particle size was measured to be ~90 nm, and iron was the principal constituent of the nZVI particle. For CVOCs analysis, the
electron capture detector (ECD) and flame ionization detector was equipped with gas chromatography (GC). The samples were
collected from the on-site constructed bore logs at a depth of 2.9, 3.2, 3.51, 3.81, 4.12, and 4.42 m on days 0, 1, 3, and 17.
Fig. 5 illustrates the iron, boron, chloride, and CVOCs compound’s depth profiles detected on days 0, 3, and 17. The trend of total
iron in Fig. 5a showed an exciting trend as day 0 and day 17 were identical; however, day 3 showed an increase in the total iron
concentration (750 μM) at a depth of 4.42 m. The same pattern was followed in Fig. 5b, where approx. 5500 μM concentration was
observed for total boron concentration on day three at 4.42 m of depth, which is higher than day 0 and day 17 at the same depth. The
exciting trend was observed using modified nZVI particle where PCE concentration was significantly dropped on day 17. At bgs 4.42 m,
the PCE reduction was more on day 17 than day 0, and around 82.5% of removal efficiency was achieved. However, at the 4.42 m bgs,
for TCE, 72.2% removal efficiency was achieved. There is a clear indication that the performance of CMC-S-nZVI particles was more as
the depth increased below the ground surface.

8
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Table 1
Performance of nanoparticles (NPs) in the removal of various groundwater pollutants and their removal efficiencies w.r.t dosages.

Sr NPs NP NP NP Target pollutant Pollutant Process Contact References


weight volume doses reduction time

1 Zero-valent Sn – – 10 mg Lead II as the toxic – Microwave 20 s Mahmoud and


NPS heavy metal from adsorptive method Abdelwahab
water (Chemisorption (2020)
reaction)
2 CMC-modified 114 g 0.57 m3 0.2 g/ – – Injection well 2 weeks He et al. (2010)
nZVI L
CMC-modified 567 g 0.587 1 g/L VOCs, TCE, PCE, 40% Ch. ethane, Injection well 2 weeks He et al. (2010)
Nzvi m3 cis & trans-VCE, 87% PCB, 29%
Vinyl chloride VC
3 CMC-modified 0.69 g 0.12 m3 0.8 wt TCE and PCE 20% of initial Push-Pull well 7 days Bennett et al.
nZVI % conc. (2010)
CMC-modified 0.21 g 0.11 m3 0.4 wt – 20% of initial – – Bennett et al.
nZVI % conc. (2010)
CMC-modified 0.34 g 0.33 m3 0.27 – 20% of initial – – Bennett et al.
nZVI wt% conc. (2010)
4 CMC-modified – – 5.7 g/ Estimation and – Injection well – Johnson et al.
nZVI L determination (2013)
5 CMC-modified – 760 L 1 g/L TCE and PCE 71% Constant-head 2 years Kocur et al.
nZVI breakthrough gravity injection (2014)
6 CMC-modified 120 kg 850 L – Benzene, MTBE, 100% efficiency Injection and – Busch et al.
carbo-iron and ammonium extraction well (2015)
colloids (CIC)
7 CMC-modified 130 kg 0.85 m3 – VOCs, PCE, and 97% efficiency – 200 days Mackenzie et al.
carbo-iron DCE (2016)
colloids (CIC)
8 Pd-modified 52 kg 0.57 m3 – TCE, cis-DCE >50% Multiple pressure 6 Edmiston et al.
nZVI (iron- injection points months (2011)
Osorb)
9 Pd-modified 30 kg – 18.25 VC, DCA, DCE, 90% Gravity-fed injection – Wei et al. (2012)
nZVI (iron- g/L TCE through an injection
Osorb) well
10 NANOFER 300 kg – – PCE, TCE, and 67.50% – – Mueller et al.
DCE (2012)
11 Nano-Fe 726 kg – 21.5 PCE 100% Injection and 1 month Jordan et al.
g/L monitoring well (2013)
12 NANOFER STAR 100 kg – 2 g/L VOCs (PCE, TCE, 45% – 6 Lacina et al.
DCE, and VC) months (2015)
13 nZVI – – – PCE approx. 90% – 2 years Mueller et al.
reduction (2012)
14 nZVI coated by 280 kg – 10 g/L PCE approx. 95% Pressure activated 1 year Köber et al.
polyethylene injection (2014)
glycol
15 Sulfidated nZVI – – 3 g/L TCE 100% @600min Laboratory study 600 min Chen et al.
(2020)
16 Sulfidated nZVI – – 3 g/L TCE 80% @600min Laboratory study 600 min Chen et al.
(2020)
17 Nano-Fe/Pd – – 5 g/L Cr(VI) 70% Laboratory study 250 min He et al. (2020)
bimetal loaded
zeolite
18 Nano-Fe/Pd – – 5 g/L Nitrate 72.50% Laboratory study 250 min He et al. (2020)
bimetal loaded
zeolite
19 Sulfidated nZVI – – 1 g/L TCE 100% Laboratory study 60 min Brumovsky et al.
(2020)
20 Sodium alginate – – 15 Cr(VI) 99% Laboratory study 20 min Li et al. (2019)
nZVI mg/L
21 CMC-modified – – 15 Cr(VI) 70% Laboratory study 20 min Li et al. (2019)
nZVI mg/L
22 Cu2O@CALB - – 1.5 g/ TCE 95% Laboratory study 10 min Losada-Garcia
MeNPs L et al. (2020)
23 PLA-ZVIa – – 10 g/L TCE and methyl 95% Laboratory study 40 min Pandey and Saha
orange (2020)
24 GO-nZVI-Sb – – 0.6 g/ Cr(VI) 92% Laboratory study 350 min Wu et al. (2020)
L
3
25 nZVI 13.5 kg 2.8 m 12 g/L TCE 95.7% Field scale 0.52 Ahn et al. (2021)
days
26 0.2 g – 5 g/L TCE >90% Laboratory study 14 daysc Yang et al.(2021)
(continued on next page)

9
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Table 1 (continued )

Sr NPs NP NP NP Target pollutant Pollutant Process Contact References


weight volume doses reduction time

nZVI (NANOFER
STAR)
27 nZVI-biochar – 30 mL 0.7 g/ Chlorinated 99% Laboratory study 30 days Semerád et al.
L ethenes (CEs/TCE) (2021)
a
Polylactic acid zero-valent iron.
b
Graphene oxide-sulfidated iron.
c
With passivation corrosion of Na + -NO-3

Fig. 5. Depth profile of a) total iron, b) total boron, c) chloride along with a color-coding indication of sample collected at different depths, d-j) CVOCs and
dechlorination byproducts, and k) ethene at NB-1 (Nunez Garcia et al., 2020). Access granted.

5.2. A case study on comparing nZVI and bio-char modified nZVI particles performance to decontaminate chlorinated solvents
Qian et al. (2020) investigated a study by injecting the nZVI and modified nZVI particles by direct push method into the
groundwater to decontaminate the chlorinated solvents. The groundwater contaminants studied in this study were trichloroethylene
(TCE) and trichloromethane (TCM). nZVI used in this study was NANOFER STAR which was purchased commercially. Modified nZVI
was encapsulated with biochar to increase the stability of nZVI in the groundwater. The study used two injections in each trial; the first
injection was used with the help of nZVI particles (NANOFER STAR), while the second injection was incorporated with the help of
modified biochar-nZVI particles mixture. A comparison is shown in Fig. 6 for the brief discussion on two different injections.
The first injection is based on 200 kg of nZVI particles with a final concentration of 10 g/L in a 100 m3 suspension was injected with
the help of Geoprobe for the direct push. The injection depth interval was varied with the distance below the ground surface (bgs) of
3.5 m, 4.5 m, and 5.5 m. The suspension of the nZVI particle was continuously stirred while injecting to keep the concentration of nZVI
constant. The pressure range was set between 1 and 12 bars. For the second injection, 100 kg of nZVI particles were mixed with 200 kg
of biochar in a mixing tank with a volume of 50 m3 to make a suspension. The injection depth interval was kept the same for the second
injection. This setup was injected with a pressure range between 1 and 5 bars. The TCE and TCM were detected using hexane as a
reference in the liquid-liquid separation technique equipped with a gas chromatograph electron capture detector (GC-ECD).
Fig. 6a–d illustrates TCE concentration detected in micropump wells at different nZVI particle injection locations, i.e., MPW1,
MPW2, MPW3, and MPW4. On day 0, the first injection of nZVI particles was placed in all these mentioned wells at different bgs
injection depths. MPW1 site was more contaminated and showed the initial TCE concentration approx.—3500 μg/L. After the first
injection, the TCE concentration in Fig. 6a reduced from 3500 to 1000 μg/L, and around 71.48% removal efficiency was achieved at
the bgs level of 5.5 m. However, the TCE concentration started increasing after 1–2 days of nZVI injection. On day 14, when the TCE
concentration at bgs 5.5 m reached 3000 μg/L, a second injection was conducted with a biochar-nZVI particles suspension. This in­
jection showed that the modified nZVI particles stayed longer in the groundwater and reduced the TCE concentration to 0 μg/L giving
the 100% removal efficiency on day 28, which remained at 0 until the treatment cycle ended on the 56th day. The exact process
occurred at almost all the micropump wells MPW2, MPW3, and MPW4. The modified nZVI increased the surface area and increased the
strength of iron particles inside, which keep transferring the electron flow from the inner core of the nZVI particle to the outer surface
encapsulated by the surface modifiers.

10
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Fig. 6. TCE concentration in micropump wells (MPW 1–4) (a–d) by injecting the nZVI (NANOFER STAR) at first injection and biochar-nZVI particle at the second
injection (Qian et al., 2020). Access granted.

6. Discussion on NPs in groundwater contamination removal


This paper presents a comprehensive review of the effect of NPs on different types of pollutants. However, the application of NPs
may help optimize the cost efficiency of onsite preparation and treatment. To determine the sustainability of NPs, aspects, such as
toxicity, have to be considered before implementation at the site. Possible sustainable options for NP-based remediation involves the
destruction of persistent organic and inorganic pollutants. Other technologies, such as in situ chemical oxidation, in situ heating, and
the use of surfactants, can reduce soil fertility. Therefore, nZVI remains a potential candidate for recovering organic and inorganic
pollutants from the soil without disturbing soil characteristics. Kirschling et al. (2010) determined that the nZVI particles did not
damage soil bacteria; in an onsite experiment, H2 was produced through nZVI stimulation in the soil, and the sulfates and methanogens
in the groundwater regime were reduced. Still, the impact of the large dosage on groundwater remediation is unknown because iron
can corrode and form insoluble byproducts that can be precipitated in the aquifer.
Similarly, these sediments can settle down in the voids of soil, thereby reducing the permeability of the aquifer. However, to the
best of the knowledge of the authors, the use of nZVI for groundwater contamination remediation is considered more sustainable than
any known technology as yet. Nevertheless, the engineered approach of using nZVI particle as a permeable reactive barrier can
enhance the efficiency and the ease of extracting NPs from the treated groundwater sources.
However, the use of NPs to treat contaminated groundwater may be reduced, considering possible risks. The Royal Society/Royal
Academy of Engineering (RS/RAE) has banned the usage of NPs for groundwater remediation in the United Kingdom. The primary
reason behind this restriction is the lack of field data on the use of NPs for contaminated-site remediation. The second reason is the lack
of research in this emerging field of study, as the environmental consequences of site remediation after the nZVI treatment are un­
known. Similarly, research on the observed risks of nZVI is lacking (Boxall et al., 2007; Nowack and Bucheli, 2007; Wiesner et al.,
2006). Evidence from laboratory-based studies shows that NPs/microparticles can enter the human body and be deposited into the
tissues (Renn, 2008).
Karn et al. (2009) considered the NP technology more beneficial than harmful owing to the lack of risk data. Another study
supported the idea of using iron oxide-based NPs because these particles and their ores are already present in the groundwater in the
form of rusty water (Watlington, 2005). Iron NPs were not completely understood at that time, and no literature reported the risk
analysis and toxicity potential of site remediation experiments using nZVI particles. The scientific community is concerned about the
production process of nZVI particles, which may have a persistent effect on humans and the environment (Karn Barbara et al., 2009).

11
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

7. Risk factors affecting the environment upon the release of nZVI into groundwater
Considering the potential danger of using nZVI, many countries have introduced preliminary regulatory policies. Some countries
have imposed a voluntary ban on manufacturing engineered NPs owing to general environmental concerns regarding NPs. The usage of
nZVI for in situ remediations may cause risks for humans and may affect ecological well-being. In some laboratory-based experiments
on aquifers, the artificially synthesized nZVI particles transformed into much similar naturally occurring iron II and III oxides and
hydroxides. Transport of nZVI will be constrained by the accumulation and sorption processes depending on their densities. Because of
passivation, transportation of easily reactive nZVI will decrease rapidly. In general, reactivity and mobility are interrelated; the more
reactive particles agglomerate more easily, thereby reducing the transport distances. Generally, nZVI is modified using different
coatings or by forming composites to enhance mobility. Such modification can potentially reduce the reactivity, implying that the
particles with the more significant reactivity hazard are also the most readily contained.
Little information is available about the direct health consequences of nZVI. The toxicity of nZVI can be generally separated into
questions related to the impact of the toxicity of iron and the size of the particles on the nanometer scale. With regard to the former,
only NPs smaller than 30 nm have been observed to be toxic, while bigger particles can be considered bulk substances. Usually, when
the particles are synthesized in the 10–100 nm range, the nZVI tends to get agglomerated to form larger particles. The key receptors of
interest are groundwater and surface water near the injection field and the ecological conditions. Risks associated with the use of nZVI
can probably be managed. The data available, typically based on laboratory data and modeling, show that nZVI has a short envi­
ronmental persistence (<1 year, usually days or weeks) and incredibly reduced mobility (<1 m–100 m). Further, modified nZVI has
been hypothesized to travel small distances (<100 m) from an injection site and even shorter distances if the injection site limits the
mobility conditions.
There are the following environmental factors related to nZVI treatments for groundwater remediations:

7.1. Toxicity
The remedial applications of nZVI tends to contribute to the toxicity or the possibility of harmful effects on environment and human
health. Surprisingly, the natural iron NPs existing in the atmosphere are more toxic than the nZVI (Keane, 2009). Nevertheless, several
papers have indicated that NPs smaller than 30 nm in size may have a different shape and cytotoxicity (Handy et al., 2008; Wiesner
et al., 2006). Further, NPs smaller than 30 nm may be crystalline depending on the material size, and this may impart such NPs
radically different properties compared to the bulk material (Auffan et al., 2009). Auffan and others also recommended that the NPs be
classified into those smaller than 30 nm and those larger than 30 nm.
In vitro bacterial cells are toxically affected by nZVI (Auffan et al., 2006; Li et al., 2010; Macé et al., 2006). nZVI has also been found
to surpass the traditional in vitro iron toxicity in terms of bactericidal effects on Escherichia coli (E. coli). However, under deaerating
conditions, the bactericidal properties were substantially higher, suggesting toxicity due to lack of oxygen (Lee et al., 2008). Fe (II) and
Fe (III) had a more significant inactivation effect on both gram-negative and gram-positive bacteria strains with regard to the toxicity
mechanisms of nZVI (Chen et al., 2020). Their results suggest that nZVI toxicity may be due to the ionic Fe present on the particle
surface of nZVI.
Other studies showed the mixed impacts of nZVI on soil and groundwater microorganisms. For example, three types of bacteria,
including E. coli, experienced no significant growth effects under the influence of nZVI (Wang et al., 2012). However, a proteomic
study revealed signs of oxidation stress responses and a lack of bactericidal impact of nZVI with regard to Klebsiella oxytoca (Saccà
et al., 2013). Overexpression of oxidative stress-related proteins in Bacillus cereus was observed in a proteomic study, indicating that
nZVI may have harmful effects on bacteria, causing rapid changes in soil microbes (Fajardo et al., 2012). nZVI at a concentration below
10 mg/kg did not have any toxic effect on gram-negative Klebsiella planticola, whereas it adversely affected gram-positive Batillus
nealsoni under comparable concentrations (Fajardo et al., 2012).

7.2. Bioaccumulation
NPs can be associated with suspended solids or soil where small organisms may ingest them, leading to their bioaccumulation and
allowing NPs to penetrate the food chain. Finally, the NPs are bioamplified. The particle characteristics and environmental system
depend on these mechanisms (Boxall et al., 2007). However, little is known about the survival and aggregation of nZVI in organisms or
about their bioamplification in food chains. Nanoscale oxide products of nZVI persist in biological environments owing to their low
solubility. The persistent impacts of these products, such as mutagenicity, are little understood (Auffan et al., 2006).
Thus, research indicates that so long as NPs are clustered and absorbed, they are typically less mobile but can be consumed by filter
feeders and other sedimentary species (Karn Barbara et al., 2009). However, there is no evidence of the possibility of biomagnification
of NPs (Biswas and Wu, 2005). Another study has shown that Ag NPs have not been bioaccumulated in Eisenia fetida earthworm,
indicating that bioaccumulation is impossible in some NPs (Coutris et al., 2012). However, Co NPs are found to bioaccumulate in
E. fetida, though with a lower rate than ionic Co (Coutris et al., 2012). Co bioaccumulation occurred because it is a blood variable.
Hence, bioaccumulation of nZVI too is probable because Fe is a constituent of blood. Another study showed that nZVI bioaccumulates
in embryonic medaka fish faster than nonoxidized nZVI (Chen et al., 2013).

7.3. Other NP inclusions


Components, such as doped metals (for e.g., Pd) or coatings, may also be inserted into iron NP products to improve remediation
efficiency. Such inserted components may have two considerable effects on the toxicity of nZVI:

12
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

• Coatings applied to the nZVI have been proven to significantly reduce toxicity (Chen et al., 2013), while other studies exhibited
enhancement in nZVI performance (Wiesner et al., 2006). Therefore, the coated NPs showed less toxicity in the environment.
• A coating on nZVI itself may cause toxicity (because of doping by metals). However, the layers, including nanoorganic matter, can
also decrease toxicity (Phenrat and Lowry, 2009).
Any risk estimation of such doped metals must be proportionate, provided that the percentages of these catalysts in the subsurface
are minimal, e.g., usually less than 1%.

8. Prospects and recommendations


For future applications of nZVI particles, few recommendations must be adopted for management and engineering in situ site
experiments. The technology implementation must be based on improved NPs. The nZVI particles must be studied first with regard to
their stability, mobility, reactivity, selectivity, and reduction of toxicity before field experiments. NP manufacturing may require more
investigations into product synthesis, life cycle analysis, manufacturing methods (characterization), behavior, and effect of storage on
nZVI particles. Similarly, other technical improvements may require optimizing field applications and developing reliable and robust
means for delivering NPs into groundwater. In addition to the technical specifications, potential health impacts and environmental
effects of the NPs themselves must be studied instead of merely studying the impacts of nZVI products. These studies require inves­
tigation of bare iron, bimetal composites, and coated nZVI particles. Further, nZVI application may affect the microbial (bacterial)
population in soil with time; hence, the microbial population must be reviewed. Fate and transport studies should detect the nZVI
particles in environmental media.
Furthermore, concentration reduction of NPs in the field experiments, assessment of the time frame of application of NPs, and
electron donor evaluation capacity in the field and the key factors that influence them may be investigated. The valence state of iron
and essential risk assessment may require the development of technological solutions to remediate groundwater. The transport dis­
tance of the nZVI particles in groundwater and the diffusion and dispersion of nZVI particles may provide actual knowledge about
these particles. Monitoring the transformed sites and controlling the reaction products may be good scientific strategies to examine the
future impacts of nZVI on groundwater. In addition, determination of the transport distances for colloids, which will reflect the
mobility of the nZVI particles in the real-world scenario, may be necessary.
Another important aspect of using nZVI particles is that they can be utilized in the vadose zone and groundwater regime to remove
persistent organic and inorganic contaminants. Mostly, the nZVI particles have shown good removal ability for NAPLs, and the
mechanism involved was demonstrated in this study. However, many factors are yet to be considered before the application of nZVI
particles for groundwater remediation. The effect of nZVI particles and dosage optimization of different modified NPs have also been
discussed. The authors believe that nZVI particles have potential for groundwater contamination treatment in terms of research,
synthesis, and deployment. Many alternatives to iron hydroxides, sulfidated iron, and nZVI particles exist, and this research area has
much research potential; nevertheless, it must be well investigated, and the effect of deployment of NPs into the groundwater must be
examined thoroughly to avoid any risk to humans and the environment.

Funding
This work was supported by the National Water and Energy Center and United Arab Emirates University (UAEU) [Grants Numbers
G00003034 (31R191), G00003297, and G00003296].

Author contributions
Conceptualization: Abdul Mannan Zafar, Muhammad Asad Javed; Mohamad Mostafa Mohamed; Formal Analysis: Abdul
Mannan Zafar, Muhammad Asad Javed; Data curation: Abdul Mannan Zafar, Muhammad Asad Javed; Funding Acquisition:
Mohamad Mostafa Mohamed, Ashraf Aly Hassan; Investigation: Abdul Mannan Zafar, Muhammad Asad Javed; Methodology: Abdul
Mannan Zafar, Muhammad Asad Javed; Mohamad Mostafa Mohamed; Project Administration: Mohamad Mostafa Mohamed, Ashraf
Aly Hassan; Supervision: Mohamad Mostafa Mohamed, Ashraf Aly Hassan; Visualization: Abdul Mannan Zafar, Muhammad Asad
Javed; Writing – original draft: Abdul Mannan Zafar, Muhammad Asad Javed; Writing – review & editing: Mohamad Mostafa
Mohamed, Ashraf Aly Hassan.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Appendix A. Supplementary data


Supplementary data to this article can be found online at https://doi.org/10.1016/j.gsd.2021.100694.

13
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

References
Ahn, Jun-Young, Kim, Cheolyong, Jun, Seong-Chun, Hwang, Inseong, 2021. Field-scale investigation of nanoscale zero-valent iron (NZVI) injectionparameters for
enhanced delivery of NZVI particles to groundwater. Water Res. 202, 117402 https://doi.org/10.1016/j.watres.2021.117402. In press.
Auffan, M., Decome, L., Rose, J., Orsiere, T., De Meo, M., Briois, V., Chaneac, C., Olivi, L., Berge-lefranc, J., Botta, A., Wiesner, M.R., Bottero, J., 2006. In vitro
interactions between DMSA-coated maghemite nanoparticles and human fibroblasts:A physicochemical and cyto -GenotoxicalStudy. Environ. Sci. Technol. 40,
4367–4373. https://doi.org/10.1021/es060691k.
Auffan, M., Rose, J., Bottero, J.-Y., Lowry, G.V., Jolivet, J.-P., Wiesner, M.R., 2009. Towards a definition of inorganic nanoparticles from an environmental, health and
safety perspective. Nat. Nanotechnol. 4, 634–641. https://doi.org/10.1038/nnano.2009.242.
Barbara, Karn, Todd, Kuiken, Otto, Martha, 2009. Nanotechnology and in situ remediation: a review of the benefits and potential risks. Environ. Health Perspect. 117,
1813–1831. https://doi.org/10.1289/ehp.0900793.
Bardos, P., Bone, B., Daly, P., Elliott, D., Jones, S., Lowry, G., Merly, C., 2014. A risk/benefit appraisal for the application of nano-scale zero valent iron (nZVI) for the
remediation of contaminated sites. WP9 NanoRem.
Bennett, P., He, F., Zhao, D., Aiken, B., Feldman, L., 2010. In situ testing of metallic iron nanoparticle mobility and reactivity in a shallow granular aquifer. J. Contam.
Hydrol. 116, 35–46. https://doi.org/10.1016/j.jconhyd.2010.05.006.
Biswas, P., Wu, C.-Y., 2005. Nanoparticles and the environment. J. Air Waste Manag. Assoc. 55, 708–746. https://doi.org/10.1080/10473289.2005.10464656.
Blodau, C., 2006. A review of acidity generation and consumption in acidic coal mine lakes and their watersheds. Sci. Total Environ. 369, 307–332. https://doi.org/
10.1016/j.scitotenv.2006.05.004.
Boxall, A.B.A., Tiede, K., Chaudhry, Q., 2007. Engineered nanomaterials in soils and water: how do they behave and could they pose a risk to human health?
Nanomedicine 2, 919–927. https://doi.org/10.2217/17435889.2.6.919.
Brumovský, M., Filip, J., Malina, O., Oborná, J., Sracek, O., Reichenauer, T.G., Andrýsková, P., Zbořil, R., 2020. Core–shell Fe/FeS nanoparticles with controlled shell
thickness for enhanced trichloroethylene removal. ACS Appl. Mater. Interfaces 12, 35424–35434. https://doi.org/10.1021/acsami.0c08626.
Busch, J., Meißner, T., Potthoff, A., Bleyl, S., Georgi, A., Mackenzie, K., Trabitzsch, R., Werban, U., Oswald, S.E., 2015. A field investigation on transport of carbon-
supported nanoscale zero-valent iron (nZVI) in groundwater. J. Contaminant Hydrol. Fate Trans. Biocolloids Nano. Soil Groundwater Syst. 181, 59–68. https://
doi.org/10.1016/j.jconhyd.2015.03.009.
Chen, Q., Li, J., Wu, Y., Shen, F., Yao, M., 2013. Biological responses of Gram-positive and Gram-negative bacteria to nZVI (Fe0), Fe2+ and Fe3+. RSC Adv. 3,
13835–13842. https://doi.org/10.1039/C3RA40570B.
Chen, J., Dong, H., Tian, R., Li, R., Xie, Q., 2020. Remediation of trichloroethylene-contaminated groundwater by sulfide-modified nanoscale zero-valent iron
supported on biochar: investigation of critical factors. Water Air Soil Pollut. 231, 432. https://doi.org/10.1007/s11270-020-04797-3.
Coutris, C., Hertel-Aas, T., Lapied, E., Joner, E.J., Oughton, D.H., 2012. Bioavailability of cobalt and silver nanoparticles to the earthworm Eisenia fetida.
Nanotoxicology 6, 186–195. https://doi.org/10.3109/17435390.2011.569094.
Cundy, A.B., Hopkinson, L., 2005. Electrokinetic iron pan generation in unconsolidated sediments: implications for contaminated land remediation and soil
engineering. Appl. Geochem. 20, 841–848. https://doi.org/10.1016/j.apgeochem.2004.11.014.
Cundy, A.B., Hopkinson, L., Whitby, R.L.D., 2008. Use of iron-based technologies in contaminated land and groundwater remediation: a review. Sci. Total Environ.
400, 42–51. https://doi.org/10.1016/j.scitotenv.2008.07.002.
Ding, Y., Liu, B., Shen, X., Zhong, L., Li, X., 2013. Foam-assisted delivery of nanoscale zero valent iron in porous media. J. Environ. Eng. 139, 1206–1212. https://doi.
org/10.1061/(ASCE)EE.1943-7870.0000727.
Edmiston, P.L., Osborne, C., Reinbold, K.P., Pickett, D.C., Underwood, L.A., 2011. Pilot scale testing composite swellable organosilica nanoscale zero-valent
iron—iron-Osorb®—for in situ remediation of trichloroethylene. Remed. J. 22, 105–123. https://doi.org/10.1002/rem.21302.
Elliott, D.W., Zhang, W., 2001. Field assessment of nanoscale bimetallic particles for groundwater treatment. Environ. Sci. Technol. 35, 4922–4926. https://doi.org/
10.1021/es0108584.
Fajardo, C., Ortíz, L.T., Rodríguez-Membibre, M.L., Nande, M., Lobo, M.C., Martin, M., 2012. Assessing the impact of zero-valent iron (ZVI) nanotechnology on soil
microbial structure and functionality: a molecular approach. Chemosphere 86, 802–808. https://doi.org/10.1016/j.chemosphere.2011.11.041.
Fan, D., Anitori, R.P., Tebo, B.M., Tratnyek, P.G., Lezama Pacheco, J.S., Kukkadapu, R.K., Engelhard, M.H., Bowden, M.E., Kovarik, L., Arey, B.W., 2013. Reductive
sequestration of pertechnetate (99TcO4–) by nano zerovalent iron (nZVI) transformed by abiotic sulfide. Environ. Sci. Technol. 47, 5302–5310. https://doi.org/
10.1021/es304829z.
Fan, D., O’Brien Johnson, G., Tratnyek, P.G., Johnson, R.L., 2016. Sulfidation of nano zerovalent iron (nZVI) for improved selectivity during in-situ chemical
reduction (ISCR). Environ. Sci. Technol. 50, 9558–9565. https://doi.org/10.1021/acs.est.6b02170.
Gavaskar, A., Tatar, L., Condit, W., 2005. Cost and Performance Report Nanoscale Zero-Valent Iron Technologies for Source Remediation. NAVAL FACILITIES
ENGINEERING SERVICE CENTER PORT HUENEME CA.
Gillham, R.W., O’Hannesin, S.F., 1994. Enhanced degradation of halogenated aliphatics by zero-valent iron. Groundwater 32, 958–967. https://doi.org/10.1111/
j.1745-6584.1994.tb00935.x.
Grieger, K.D., Fjordbøge, A., Hartmann, N.B., Eriksson, E., Bjerg, P.L., Baun, A., 2010. Environmental benefits and risks of zero-valent iron nanoparticles (nZVI) for in
situ remediation: risk mitigation or trade-off? J. Contaminant Hydrol. Manuf. nano. Subsurf. Syst. 118, 165–183. https://doi.org/10.1016/j.
jconhyd.2010.07.011.
Handy, R.D., von der Kammer, F., Lead, J.R., Hassellöv, M., Owen, R., Crane, M., 2008. The ecotoxicology and chemistry of manufactured nanoparticles.
Ecotoxicology 17, 287–314. https://doi.org/10.1007/s10646-008-0199-8.
Hara, S.O., Krug, T., Quinn, J., Clausen, C., Geiger, C., 2006. Field and laboratory evaluation of the treatment of DNAPL source zones using emulsified zero-valent
iron. Remed. J. 16, 35–56. https://doi.org/10.1002/rem.20080.
He, F., 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, 3314–3320. https://doi.org/10.1021/es048743y.
He, F., Zhao, D., Liu, J., Roberts, C.B., 2007. Stabilization of Fe− Pd nanoparticles with sodium carboxymethyl cellulose for enhanced transport and dechlorination of
trichloroethylene in soil and groundwater. Ind. Eng. Chem. Res. 46, 29–34. https://doi.org/10.1021/ie0610896.
He, F., Zhao, D., Paul, C., 2010. Field assessment of carboxymethyl cellulose stabilized iron nanoparticles for in situ destruction of chlorinated solvents in source
zones. Water Res. 44, 2360–2370. https://doi.org/10.1016/j.watres.2009.12.041.
He, Y., Lin, H., Luo, M., Liu, J., Dong, Y., Li, B., 2020. Highly efficient remediation of groundwater co-contaminated with Cr(VI) and nitrate by using nano-Fe/Pd
bimetal-loaded zeolite: process product and interaction mechanism. Environ. Pollut. 263, 114479. https://doi.org/10.1016/j.envpol.2020.114479.
Hennebel, T., Simoen, H., Verhagen, P., De Windt, W., Dick, J., Weise, C., Pietschner, F., Boon, N., Verstraete, W., 2011. Biocatalytic dechlorination of
hexachlorocyclohexane by immobilized bio-Pd in a pilot scale fluidized bed reactor. Environ. Chem. Lett. 9, 417–422. https://doi.org/10.1007/s10311-010-
0295-x.
Johnson, R.L., Nurmi, J.T., O’Brien Johnson, G.S., Fan, D., O’Brien Johnson, R.L., Shi, Z., Salter-Blanc, A.J., Tratnyek, P.G., Lowry, G.V., 2013. Field-scale transport
and transformation of carboxymethylcellulose-stabilized nano zero-valent iron. Environ. Sci. Technol. 47, 1573–1580. https://doi.org/10.1021/es304564q.
Jordan, M., Shetty, N., Zenker, M.J., Brownfield, C., 2013. Remediation of a former dry cleaner using nanoscale zero valent iron. Remed. J. 24, 31–48. https://doi.
org/10.1002/rem.21376.
Kalin, M., Fyson, A., Wheeler, W.N., 2006. The chemistry of conventional and alternative treatment systems for the neutralization of acid mine drainage. Sci. Total
Environ. 366, 395–408. https://doi.org/10.1016/j.scitotenv.2005.11.015.
Keane, E., 2009. Fate, Transport and Toxicity of Nanoscale Zero-Valent Iron (nZVI) Used during Superfund Remediation.
Kirschling, T.L., Gregory, K.B., Minkley, E.G., Lowry, G.V., Tilton, R.D., 2010. Impact of nanoscale zero valent iron on geochemistry and microbial populations in
trichloroethylene contaminated aquifer materials. Environ. Sci. Technol. 44, 3474–3480. https://doi.org/10.1021/es903744f.

14
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Köber, R., Hollert, H., Hornbruch, G., Jekel, M., Kamptner, A., Klaas, N., Maes, H., Mangold, K.-M., Martac, E., Matheis, A., Paar, H., Schäffer, A., Schell, H.,
Schiwy, A., Schmidt, K.R., Strutz, T.J., Thümmler, S., Tiehm, A., Braun, J., 2014. Nanoscale zero-valent iron flakes for groundwater treatment. Environ. Earth Sci.
72, 1. https://doi.org/10.1007/s12665-014-3239-0.
Kocur, C.M., Chowdhury, A.I., Sakulchaicharoen, N., Boparai, H.K., Weber, K.P., Sharma, P., Krol, M.M., Austrins, L., Peace, C., Sleep, B.E., O’Carroll, D.M., 2014.
Characterization of nZVI mobility in a field scale test. Environ. Sci. Technol. 48, 2862–2869. https://doi.org/10.1021/es4044209.
Kocur, C.M.D., Lomheim, L., Boparai, H.K., Chowdhury, A.I.A., Weber, K.P., Austrins, L.M., Edwards, E.A., Sleep, B.E., O’Carroll, D.M., 2015. Contributions of abiotic
and biotic dechlorination following carboxymethyl cellulose stabilized nanoscale zero valent iron injection. Environ. Sci. Technol. 49, 8648–8656. https://doi.
org/10.1021/acs.est.5b00719.
Kocur, C.M.D., Lomheim, L., Molenda, O., Weber, K.P., Austrins, L.M., Sleep, B.E., Boparai, H.K., Edwards, E.A., O’Carroll, D.M., 2016. Long-term field study of
microbial community and dechlorinating activity following carboxymethyl cellulose-stabilized nanoscale zero-valent iron injection. Environ. Sci. Technol. 50,
7658–7670. https://doi.org/10.1021/acs.est.6b01745.
Lacina, P., Dvorak, V., Vodickova, E., Barson, P., Kalivoda, J., Goold, S., 2015. The application of nano-sized zero-valent iron for in situ remediation of chlorinated
ethylenes in groundwater: a field case study. Water Environ. Res. 87, 326–333. https://doi.org/10.2175/106143015X14212658613596.
Lee, C., Kim, J.Y., Lee, W.I., Nelson, K.L., Yoon, J., Sedlak, D.L., 2008. Bactericidal effect of zero-valent iron nanoparticles on Escherichia coli. Environ. Sci. Technol.
42, 4927–4933. https://doi.org/10.1021/es800408u.
Li, X., Zhang, W., 2007. Sequestration of metal cations with zerovalent iron NanoparticlesA study with high resolution X-ray photoelectron spectroscopy (HR-XPS).
J. Phys. Chem. C 111, 6939–6946. https://doi.org/10.1021/jp0702189.
Li, Z., Greden, K., Alvarez, P.J.J., Gregory, K.B., Lowry, G.V., 2010. Adsorbed polymer and NOM limits adhesion and toxicity of nano scale zerovalent iron to E. coli.
Environ. Sci. Technol. 44, 3462–3467. https://doi.org/10.1021/es9031198.
Li, Z., Xu, S., Xiao, G., Qian, L., Song, Y., 2019. Removal of hexavalent chromium from groundwater using sodium alginate dispersed nano zero-valent iron. J. Environ.
Manag. 244, 33–39. https://doi.org/10.1016/j.jenvman.2019.04.130.
Losada-Garcia, N., Rodriguez-Otero, A., Palomo, J.M., 2020. High degradation of trichloroethylene in water by nanostructured MeNPs@CALB biohybrid catalysts.
Catalysts 10, 753. https://doi.org/10.3390/catal10070753.
Lu, F., Astruc, D., 2020. Nanocatalysts and other nanomaterials for water remediation from organic pollutants. Coord. Chem. Rev. 408, 213180. https://doi.org/
10.1016/j.ccr.2020.213180.
Macé, C., Desrocher, S., Gheorghiu, F., Kane, A., Pupeza, M., Cernik, M., Kvapil, P., Venkatakrishnan, R., Zhang, W., 2006. Nanotechnology and groundwater
remediation: a step forward in technology understanding. Remed. J. 16, 23–33. https://doi.org/10.1002/rem.20079.
Mackenzie, K., Bleyl, S., Kopinke, F.-D., Doose, H., Bruns, J., 2016. Carbo-Iron as improvement of the nanoiron technology: from laboratory design to the field test. Sci.
Total Environ. 563–564, 641–648. https://doi.org/10.1016/j.scitotenv.2015.07.107.
Mahmoud, M.E., Abdelwahab, M.S., 2020. One-step synthesis of zero-valent Sn nanoparticles and potential microwave remediation of lead from water. Mater. Res.
Bull. 111090 https://doi.org/10.1016/j.materresbull.2020.111090.
Mohamed, M.M., Hatfield, K., 2005. Modeling microbial-mediated reduction in batch reactors. Chemosphere 59 (8), 1207–1217. https://doi.org/10.1016/j.
chemosphere.2004.12.013.
Mohamed, M., Hatfield, K., 2011. Dimensionless Monod parameters to summarize the influence of microbial inhibition on the attenuation of groundwater
contaminants. Biodegradation 22, 877–896.
Mohamed, M.M.A., Hatfield, K., Hassan, A.E., 2006. Monte Carlo evaluation of microbial-mediated contaminant reactions in heterogeneous aquifers. Adv. Water
Resour. 29 (8), 1123–1139. https://doi.org/10.1016/j.advwatres.2005.09.008.
Mohamed, M., Hatfield, K., Perminova, I.V., 2007. Evaluation of Monod kinetic parameters in the subsurface using moment analysis: theory and numerical testing.
Adv. Water Resour. 30 (9), 2034–2050. https://doi.org/10.1016/j.advwatres.2007.04.006.
Mohamed, M., Hatfield, K., Hassan, A.E., Klammler, H., 2010a. Stochastic evaluation of subsurface contaminant discharges under physical, chemical, and biological
heterogeneities. Adv. Water Resour. 33 (Issue 7), 801–812.
Mohamed, M.M., Saleh, N.E., Sherif, M.M., 2010b. Sensitivity of benzene natural attenuation to variations in kinetic and transport parameters in Liwa Aquifer. UAE
Bulletin Environ. Contamination Toxicol. 84 (4), 443–449. https://doi.org/10.1007/s00128-010-9957-4.
Mohamed, M., Saleh, N., Sherif, M., 2010c. Modeling in-situ benzene bioremediation in the contaminated liwa aquifer (UAE) using the slow-release oxygen source
technique. Environ. Earth Sci. 61 (7), 1385–1399.
Mueller, N.C., Braun, J., Bruns, J., Černík, M., Rissing, P., Rickerby, D., Nowack, B., 2012. Application of nanoscale zero valent iron (NZVI) for groundwater
remediation in Europe. Environ. Sci. Pollut. Res. 19, 550–558. https://doi.org/10.1007/s11356-011-0576-3.
Nowack, B., Bucheli, T.D., 2007. Occurrence, behavior and effects of nanoparticles in the environment. Environ. Pollut. 150, 5–22. https://doi.org/10.1016/j.
envpol.2007.06.006.
Nunez Garcia, A., Boparai, H.K., Chowdhury, A.I.A., de Boer, C.V., Kocur, C.M.D., Passeport, E., Sherwood Lollar, B., Austrins, L.M., Herrera, J., O’Carroll, D.M., 2020.
Sulfidated nano zerovalent iron (S-nZVI) for in situ treatment of chlorinated solvents: a field study. Water Res. 174, 115594. https://doi.org/10.1016/j.
watres.2020.115594.
Oppegard, A.L., Darnell, F.J., Miller, H.C., 1961. Magnetic properties of single-domain iron and iron-cobalt particles prepared by borohydride reduction. J. Appl. Phys.
32, S184–S185. https://doi.org/10.1063/1.2000393.
Pandey, K., Saha, S., 2020. Microencapsulated Zero Valent Iron NanoParticles in Polylactic acid matrix for in situ remediation of contaminated water. J. Environ.
Chem. Eng. 8, 103909. https://doi.org/10.1016/j.jece.2020.103909.
Phenrat, T., Lowry, G.V., 2009. Chapter 18 - physicochemistry of polyelectrolyte coatings that increase stability, mobility, and contaminant specificity of reactive
nanoparticles used for groundwater remediation. In: Savage, N., Diallo, M., Duncan, J., Street, A., Sustich, R. (Eds.), Nanotechnology Applications for Clean
Water, Micro and Nano Technologies. William Andrew Publishing, Boston, pp. 249–267. https://doi.org/10.1016/B978-0-8155-1578-4.50027-5.
Phenrat, T., Kim, H.-J., Fagerlund, F., Illangasekare, T., Tilton, R.D., Lowry, G.V., 2009a. Particle size distribution, concentration, and magnetic attraction affect
transport of polymer-modified Fe0 nanoparticles in sand columns. Environ. Sci. Technol. 43, 5079–5085. https://doi.org/10.1021/es900171v.
Phenrat, T., Long, T.C., Lowry, G.V., Veronesi, B., 2009b. Partial oxidation (“Aging”) and surface modification decrease the toxicity of nanosized zerovalent iron.
Environ. Sci. Technol. 43, 195–200. https://doi.org/10.1021/es801955n.
Phenrat, T., Kim, H.-J., Fagerlund, F., Illangasekare, T., Lowry, G.V., 2010. Empirical correlations to estimate agglomerate size and deposition during injection of a
polyelectrolyte-modified Fe0 nanoparticle at high particle concentration in saturated sand. J. Contaminant Hydrol. Manuf. nano. Subsurf. Syst. 118, 152–164.
https://doi.org/10.1016/j.jconhyd.2010.09.002.
Phenrat, T., Lowry, G.V., Babakhani, P., 2019. Nanoscale zerovalent iron (NZVI) for environmental decontamination: a brief history of 20 Years of research and field-
scale Application. In: Phenrat, T., Lowry, G.V. (Eds.), Nanoscale Zerovalent Iron Particles for Environmental Restoration: from Fundamental Science to Field Scale
Engineering Applications. Springer International Publishing, Cham, pp. 1–43. https://doi.org/10.1007/978-3-319-95340-3_1.
Qian, L., Chen, Y., Ouyang, D., Zhang, W., Han, L., Yan, J., Kvapil, P., Chen, M., 2020. Field demonstration of enhanced removal of chlorinated solvents in
groundwater using biochar-supported nanoscale zero-valent iron. Sci. Total Environ. 698, 134215. https://doi.org/10.1016/j.scitotenv.2019.134215.
Quinn, J., Geiger, C., Clausen, C., Brooks, K., Coon, C., O’Hara, S., Krug, T., Major, D., Yoon, W.-S., Gavaskar, A., Holdsworth, T., 2005. Field demonstration of DNAPL
dehalogenation using emulsified zero-valent iron. Environ. Sci. Technol. 39, 1309–1318. https://doi.org/10.1021/es0490018.
Renn, O., 2008. White paper on risk governance: toward an integrative framework. In: Renn, O., Walker, K.D. (Eds.), Global Risk Governance: Concept and Practice
Using the IRGC Framework, International Risk Governance Council Bookseries. Springer Netherlands, Dordrecht, pp. 3–73. https://doi.org/10.1007/978-1-4020-
6799-0_1.
Reynolds, G.W., Hoff, J.T., Gillham, R.W., 1990. Sampling bias caused by materials used to monitor halocarbons in groundwater. Environ. Sci. Technol. 24, 135–142.
https://doi.org/10.1021/es00071a017.

15
A.M. Zafar et al. Groundwater for Sustainable Development 15 (2021) 100694

Saccà, M.L., Fajardo, C., Nande, M., Martín, M., 2013. Effects of nano zero-valent iron on Klebsiella oxytoca and stress response. Microb. Ecol. 66, 806–812. https://
doi.org/10.1007/s00248-013-0269-1.
Saleh, N., Sirk, K., Liu, Y., Phenrat, T., Dufour, B., Matyjaszewski, K., Tilton, R.D., Lowry, G.V., 2007. Surface modifications enhance nanoiron transport and NAPL
targeting in saturated porous media. Environ. Eng. Sci. 24, 45–57.
Schrick, B., Hydutsky, B.W., Blough, J.L., Mallouk, T.E., 2004. Delivery vehicles for zerovalent metal nanoparticles in soil and groundwater. Chem. Mater. 16,
2187–2193. https://doi.org/10.1021/cm0218108.
Semerád, J., Filip, J., Ševců, A., Brumovský, M., Nguyen, N.H.A., Mikšíček, J., Lederer, T., Filipová, A., Boháčková, J., Cajthaml, T., 2020. Environmental fate of
sulfidated nZVI particles: the interplay of nanoparticle corrosion and toxicity during aging. Environ. Sci.: Nano 7, 1794–1806. https://doi.org/10.1039/
D0EN00075B.
Smith, S.C., Rodrigues, D.F., 2015. Carbon-based nanomaterials for removal of chemical and biological contaminants from water: a review of mechanisms and
applications. Carbon 91, 122–143. https://doi.org/10.1016/j.carbon.2015.04.043.
Stejskal, V., Lederer, T., Kvapil, P., Slunsky, J., Skácelová, P., 2017. NanoRem pilot site–spolchemie I, Czech republic: nanoscale zero-valent iron remediation of
chlorinated hydrocarbons. NanoRem bulletin 1–8.
Taghavy, A., Costanza, J., Pennell, K.D., Abriola, L.M., 2010. Effectiveness of nanoscale zero-valent iron for treatment of a PCE–DNAPL source zone. J. Contaminant
Hydrol. Manuf. nano. Subsurf. Syst. 118, 128–142. https://doi.org/10.1016/j.jconhyd.2010.09.001.
Teefy, D.A., 1997. Remediation technologies screening matrix and reference guide: version III. Remed. J. 8, 115–121. https://doi.org/10.1002/rem.3440080111.
USEPA, 2007–. (Accessed 1 February 2007).
Wang, C.-B., Zhang, W., 1997. Synthesizing nanoscale iron particles for rapid and complete dechlorination of TCE and PCBs. Environ. Sci. Technol. 31, 2154–2156.
https://doi.org/10.1021/es970039c.
Wang, Q., Jeong, S.-W., Choi, H., 2012. Removal of trichloroethylene DNAPL trapped in porous media using nanoscale zerovalent iron and bimetallic nanoparticles:
direct observation and quantification. J. Hazard Mater. 213–214, 299–310. https://doi.org/10.1016/j.jhazmat.2012.02.002.
Watlington, K., 2005. Emerging Nanotechnologies for Site Remediation and Wastewater Treatment. Environmental Protection Agency USA.
Wei, Y.-T., Wu, S., Yang, S., Che, C.-H., Lien, H.-L., Huang, D.-H., 2012. Biodegradable surfactant stabilized nanoscale zero-valent iron for in situ treatment of vinyl
chloride and 1,2-dichloroethane. J. Hazard Mater. 211–212, 373–380. https://doi.org/10.1016/j.jhazmat.2011.11.018.
Wiesner, M.R., Lowry, G.V., Alvarez, P., Dionysiou, D., Biswas, P., 2006. Assessing the risks of manufactured nanomaterials. Environ. Sci. Technol. 40, 4336–4345.
https://doi.org/10.1021/es062726m.
Wu, H., Li, L., Chang, K., Du, K., Shen, C., Zhou, S., Sheng, G., Linghu, W., Hayat, T., Guo, X., 2020. Graphene oxide decorated nanoscale iron sulfide for highly
efficient scavenging of hexavalent chromium from aqueous solutions. J. Environ. Chem. Eng. 8, 103882. https://doi.org/10.1016/j.jece.2020.103882.
Xiu, Z., Jin, Z., Li, T., Mahendra, S., Lowry, G.V., Alvarez, P.J.J., 2010. Effects of nano-scale zero-valent iron particles on a mixed culture dechlorinating
trichloroethylene. Bioresour. Technol. 101, 1141–1146. https://doi.org/10.1016/j.biortech.2009.09.057.
Yang, Lei, Chen, Yun, Ouyang, Da, Yan, Jingchun, Qian, Linbo, Han, Lu, Chen, Mengfang, Li, Jing, Gu, Mingyue, 2021. Mechanistic insights into adsorptive and
oxidative removal of monochlorobenzene in biochar-supported nanoscale zero-valent iron/ persulfate system. Chem. Eng. J. 400, 125811 https://doi.org/
10.1016/j.cej.2020.125811. In press.
Zhan, J., Zheng, T., Piringer, G., Day, C., McPherson, G.L., Lu, Y., Papadopoulos, K., John, V.T., 2008. Transport characteristics of nanoscale functional zerovalent
iron/silica composites for in situ remediation of trichloroethylene. Environ. Sci. Technol. 42, 8871–8876. https://doi.org/10.1021/es800387p.
Zheng, T., Zhan, J., He, J., Day, C., Lu, Y., McPherson, G.L., Piringer, G., John, V.T., 2008. Reactivity characteristics of nanoscale zerovalent Iron− Silica composites
for trichloroethylene remediation. Environ. Sci. Technol. 42, 4494–4499. https://doi.org/10.1021/es702214x.
Zhu, B.-W., Lim, T.-T., 2007. Catalytic reduction of chlorobenzenes with Pd/Fe nanoparticles:reactive sites ,catalyst stability ,particle aging ,and regeneration.
Environ. Sci. Technol. 41, 7523–7529. https://doi.org/10.1021/es0712625.

16

You might also like