DOI:
10.1039/C6RA16657A
(Review Article)
RSC Adv., 2016,
6, 91025-91044
Carriers for nano zerovalent iron (nZVI): synthesis, application and efficiency
Received
28th June 2016
, Accepted 30th August 2016
First published on 13th September 2016
Abstract
Nanoparticles refer to very small size elements less than 100 nm. These particles have a large surface area to volume ratio with enhanced reactivity. One nanomaterial very interesting for remediation activities is nano zero valent iron (nZVI) estimated to be 10 to 1000 times more reactive than granular ZVI, thereby creating a greater reductive capacity per gram. Thus smaller amounts of iron would be needed to treat a contaminated plume at a faster rate, which would further reduce costs. nZVI also has the potential to treat recalcitrant pollutants such as chlorinated organic compounds, nitrates and hexavalent chromium. But the high reactivity of nZVI alone is not enough to make it an effective remediation agent, the nano iron particles should also be able to form stable dispersions and migrate to contaminated plumes. These expected additional properties are seen to be missing as the nZVI quickly agglomerates and oxidizes once released into the environment. This study looks at the challenges associated with the use of nZVI and the different transporters or carriers that have sought to overcome these shortfalls. Further attention is focused on how these nZVI particles are synthesized on carriers and the results obtained when used to treat pollutants.
 Junias Adusei-Gyamfi | Junias Adusei-Gyamfi is a Master's student of Sustainable Management of Pollution at Catholic University of Lille (ISA), France. He has a bachelor's degree in Chemistry from KNUST, Ghana. Before his master's, he worked as a project officer for an Environmental non-governmental organization, Learnyoung Africa. He considers himself as a very disciplined and focused man with great interest in sustainable environmental development. He desires to help developing countries overcome their myriad environmental issues both through research and field works. |
 Victor Acha | Dr. Victor Acha is Associate Professor of Nanotechnologies and Bioprocess Engineering at LaSalle Beauvais, France. He got his B.Sc. in Chemical Engineering at USFX (Bolivia), M.Sc. in Food Engineering and Ph.D. in Bioengineering at UCL (Belgium). He worked in Colorado State University (USA) for 4 years and UTHSC San Antonio (USA) for 2 years. Cofounder of the Research Unit Hydrise (LaSalle), interested in the treatment of polluted waters by nanoparticles, ozonation, model-based estimations (software sensors and adaptive control), and development of nano(bio)sensors for environmental, biomedical, and food safety applications. Peer reviewer Certificate from Publons. Member of different Research Associations. |
Introduction
Nanotechnology is one of the fastest growing fields of scientific research and promises a potential revolution in approaches to remediation of polluted sites.1 The term nanoparticle is used to classify particles with dimensions less than 100 nanometers (nm) in size. The unique and specific characteristics of different nanomaterials defines their usage and applications in machinery, energy, cosmetics, optics, electronics, drug delivery, and medical diagnostics.2,3 The nano metals commonly used include iron (Fe0), magnesium (Mg0), palladium (Pd0) and silver (Ag0)4 with nano zero valent iron being the most used.5 In 2000 the first field trial of nano zero valent iron (nZVI) for the treatment of trichloroethylene in groundwater at a manufacturing site in Trenton, New Jersey, USA, was documented.6
nZVI is a promising material for groundwater in situ remediation because of its high surface area, high reactivity, fast kinetics, small particle size, and magnetic property that is beneficial for separating and recovering tiny particles from wastewater.7 Though the definition of nano particles sets the threshold size at 100 nm, a review of applications in literature indicates the average particles size is approximately 60 nm (>90%)8 while the critical diameter below which specific nano effect occurs is 30 nm.9 The high specific surface area allows nZVI to be 10 to 1000 times highly reactive than the granular ZVI, and the sorption capacity is also much higher2,10 thereby creating a greater reductive capacity per gram. Thus smaller amounts of nZVI would be needed to treat a contaminated plume that would further reduce costs.11
The principal potential benefits of nZVI are the speed of contaminant degradation due to its high surface area and reactivity that reduces remediation time.12 It also has the potential of treating recalcitrant pollutants such as chlorinated organic compounds, nitrates and hexavalent chromium that can be found in aquatic environments.4 There is almost complete degradation process eliminating the formation of toxic intermediate products and it has the potential for synergistic application with bioremediation techniques.6
Fe0 has standard reduction potential (E0) of −0.440 V for half-reaction between the Fe2+/Fe0 couple (eqn (1)) which confirms that it is an effective electron donor regardless of its particle size. Reduction potential measures the probability of a chemical species being reduced. Thus a higher reduction potential means a higher probability of the species to be reduced. This means that theoretically Fe0 can reduce any pollutant that has a higher reduction potential than −0.440 V.13
In the subsurface environment, the predominant electron receptors are water and to some extent residual dissolved oxygen as described in eqn (2) and (3)
|
Fe0(s) + 2H2O(aq) → Fe2+(aq) + H2(g) + 2OH−(aq)
| (2) |
|
2Fe0(s) + 4H+(aq) + O2(aq) → 2Fe2+(aq) + 2H2O(l)
| (3) |
Since the oxidation of Fe0 consumes H+ and produces OH− ions (eqn (2) and (3)), there is an increase in solution pH and an associated decline in solution potential (Eh). The high surface free energy of nZVI coupled with the dissociative adsorption of water on the iron surface results in the formation of surface-bound hydroxyl species. Thus before reacting with target pollutants, the nZVI rapidly reacts with water and/or oxygen in its surrounding condition, resulting in the formation of passive oxidation layers.7 The structure of the nanoparticle is therefore made up of an inner core which exhibits characteristics of a metallic iron (reductant) and an oxidized shell with characteristics of an iron oxide (sorbent) (Fig. 1).8
 |
| Fig. 1 A core–shell structure for iron nanoparticles in aqueous solution. This figure has been adapted from ref. 8 with permission from Elsevier. | |
The principal mechanism of the core is to act as an electron source for the reduction of pollutants. Iron oxides however can strongly adsorb metals in soils thus immobilizing metallic contaminants in soil but not much is known about the efficiency of the treatment in soils that are not saturated with water. It also serves as the barrier for electron passage from the core.1
The high reactivity of nZVI alone is not enough to make it an effective remediation agent, the nano iron particles should also be able to form stable dispersions and migrate to contaminated plume.14,15
This study looks at the challenges associated with the use of nZVI and the different transporters or carriers that have sought to overcome these challenges. A further attention is focused on how these nZVI particles are synthesized on carriers and the results obtained when used to treat pollutants.
Properties of nZVI
Iron which is a strong lustrous, ductile, malleable metal is seen as one of the most abundant elements in the earth's crust is found in group 8, period 4 and block D of the periodic table. nZVI is a sub-micrometer particles of iron metal. Iron is highly reactive to both air (oxygen) and water, and this reactivity increases significantly at the nanoscale. A summary of the general properties of the bulk Fe is shown in Table 1. The properties of nZVI are usually different from the bulk Fe particle and this difference is mainly caused by factors such as the history of the sample, its handling and processing which would affect the size, structure and even composition of the nanoparticle. Additionally, the close proximity to other nanoparticles can also affect the electronic and magnetic characteristics of individual nanoparticles. The packing or dispersion of nanoparticles for measurement purposes may in some circumstances also change the results for the properties being measured.16
Table 1 General properties of iron
Symbol |
Fe |
Melting point |
1538 °C |
Boiling point |
2861 °C |
Density (g cm−3) |
7.87 |
Relative atomic mass |
55.845 |
Atomic number |
26 |
Key isotopes |
56Fe |
Electron configuration |
[Ar] 3d64s2 |
First ionization energy |
761 kJ mol−1 |
Table 2 Summary of nZVI synthesis methods34
Chemical synthesis methods |
Liquid-phase reduction or borohydride reduction of ferrous salts |
Gas-phase reduction |
Microemulsion |
Controlled chemical co-precipitation |
Chemical vapor condensation |
Pulse electrodeposition |
Liquid flame spray |
Thermal reduction of ferrous iron |
Electrolysis |
Physical synthesis methods |
Polyphenolic plant extract |
Inert gas condensation |
Severe plastic deformation |
High-energy ball milling |
Ultrasound shot peening |
Different characterization tools (in situ and ex situ) such as optical spectroscopy, scanning electron microscopy, transmission electron microscopy, and X-ray diffraction have been developed to study detailed characteristics and properties of the nanoparticles such as size, shape, size distribution, surface area, and chemical state. These tools have helped reveal the core–shell structure of nZVI as shown in Fig. 1. This core-structure dictates the chemical properties of the nanoparticle by exhibiting characteristics of both iron oxides (sorption) and metallic iron (electron source). According to the core–shell model, the mixed valence iron oxide shell is largely insoluble under neutral pH conditions and may protect the ZVI core from rapid oxidation.17
Magnetic properties
nZVI has shown generally to possess some magnetic properties though this properties largely vary on factors such as history (preparation method and storage), size, shape, chemical composition, surface oxidation and dimension of the nanoparticle.18,19 In fact it is this property that results in the bare nanoparticles forming chain like structure and resulting in agglomeration. The strong magnetic properties of nZVI have been exploited in many recovery processes like using magnetic separation of coated nZVI particles to extract algae from solution. The nonlinear hysteresis loops with nonzero remnant magnetization (Mr) and coercivity (Hc) exhibits ferromagnetic properties like iron oxide core–shell structures with saturation magnetization (Ms) value of about 100 emu g−1 (ref. 20 and 21) confirmed this that it is evident that the nanoparticles behave ferromagnetically with Curie temperature much higher than 76.85 °C further adding that both the squareness ratio (Mr/Ms) ratio and coercive field Hc decreases monotonically with increasing temperature, as expected for small particle systems. Actually in monodisperse systems, when magnetic field strength is strong enough to overcome hysteresis of the particles, larger ferromagnetic particles show higher heating performance than smaller superparamagnetic ones.22
Metallic ferromagnetic particles show very large coercivity compared with the value predicted by the simple Stoner–Wohlfarth model under the assumption of bulk anisotropy. nZVI are therefore likely to possess uniaxial anisotropy with their effective anisotropy being much larger than that of bulk Fe particles.23 When the particle size is increased, coercivity and relative remanence increases which corresponds to transition from superparamagnetic to ferrimagnetic behavior. Ferrimagnetic and ferromagnetic share very similar properties but differ in the direction neighboring dipoles are arranged. In ferromagnetic, the dipoles are arranged in the same direction whiles in ferrimagnetic materials, the neighboring dipoles are opposite to each other. Ferromagnetic particles also show higher reversal losses than smaller superparamagnetic particles.22
Generally due to the increase of surface component of anisotropy (Ks), magnetic anisotropy tends to increase with decreasing of particle size. Magnetic anisotropy is generated by the spin–orbit coupling occurred at magnetic cations, and the anisotropy decreases with decreasing spin–orbit coupling, leading to a decrease of coercivity.24
Catalytic properties
Due to factors such as cost, abundance, stability, recyclability and environmental friendliness, transition metals are considered to play an important role in catalysis and as popular substitutes of platinum metal based catalysts. This role is of more important when the metal particle is on the nanoscale due to the modification in its physical structure such as surface area to volume ratio and surface exposure of atoms. These factors tend to contribute positively towards increasing the efficiency of the catalyst. nZVI in the quantum dot range have been reported to be effective catalysts in many reduction reactions such as being used as a catalysts in the hydrogenation reaction of various substituted aromatic ketones to alcohols with NaBH4.25
nZVI is also a very desirable catalyst in Fischer–Tropsch (F–T) synthesis though after the reaction the catalyst/waxy product separation is problematic to regenerate the catalyst. Low product selectivity, catalyst agglomeration and sintering are other concerns of the use of this catalyst. However, iron is still desirable because of its high activation which ensures that the hydrocarbons produced are high in olefinic content.26,27 The synthesis method of the nanoparticles also plays an important role in physical properties and performance of catalysts.28
In the synthesis of both single and multi-walled carbon nanotubes through chemical vapor deposition (CVD) process, nZVI have been used as an effective catalyst. The results indicate that there is an upper limit for the size of the catalyst particles to nucleate the nanotubes and the size of the nanoparticle determines the diameter of the tubes.29 It permits submicron scale patterned nanotube growth that retains a high degree of surface cleanliness30,31 however found carbon nanotubes synthesized with Fe are not very stable. Thus with increasing temperature (>700 °C), diffusion of carbon in iron is favored forming more amorphous carbon.
In Fenton's reaction, the metallic or zero-valent iron, which is a two-electron donor, can directly reduce dissolved molecular oxygen in aqueous solutions to hydrogen peroxide at its solid surface according to the well-studied iron corrosion theory. The hydrogen peroxide produced can then also heterogeneously oxidize zero-valent iron into ferrous iron.32 One main disadvantage however of using Fenton reagent is that the pH has to be lowered to less than 4 to keep the iron in solution.17
Synthesis of nZVI
Nanoparticles can occur naturally in air, water, soil and sediments or be produced intentionally for specialized process. It can also be formed as accidental by-products of industrial processes.1
Although micro particles of iron are inexpensive, reactive nanoparticles can be much more expensive because of the materials and processes needed to make them.33 There exist several physical and chemical methods for the synthesis of nZVI such as grinding, abrasion, lithography, nucleation from homogeneous solutions or gas, annealing at elevated temperatures and reacting with reducing agents. These synthesis methods are grouped under two broad approaches; the bottom-up and top-down approaches.
The bottom-up approach involves assembling individual atoms and molecules to form nano sized structures. It uses a wide range of reductants to convert dissolved iron in solution to nZVI. The top-down approach on the contrary involves the crushing of bulk particles (microscale or granular) of iron to fine nano-sized particles by mechanical or chemical ways.3,6 The choice of synthesis method influences both the size and shape of the nano particles produced. Table 2 summarizes different physical and chemical synthesis methods used by researchers with some of the mostly used techniques detailed in the next section.
Among the chemical synthesis methods, the borohydride reduction of ferrous salts is more popularly used due to its simplicity as no special instruments or materials are needed; in addition, the products have a homogenous structure and are very reactive.12 Another reason for the preference of this technique is its lower associated cost of production at laboratory scale33 estimated to $200 per kg of nZVI.35 Most of other techniques for synthesizing nano materials are not feasible or cost-effective for industrial large scale production.36
Liquid-phase reduction or borohydride reduction
The synthesis of nZVI can be achieved by reacting iron salt or compound with a strong reducing agent such as sodium borohydride (NaBH4). The borohydride solution is slowly added into the iron (ferric or ferrous) salt while stirring vigorously under anaerobic conditions. The resulting black precipitate formed is vacuumed filtered, washed with deionized water or ethanol and subsequently dried.37 The reaction follows the order: |
4Fe3+(aq) + BH4− + 3H2O → 4Fe2+(s) + H2BO3− + 4H+(aq) + 2H2(g)
| (4) |
|
2Fe2+(aq) + BH4− + 3H2O → 2Fe0(s) + H2BO3− + 4H+(aq) + 2H2(g)
| (5) |
|
4Fe3+(aq) + 3BH4− + 9H2O → 4Fe0(s) + 3H2BO3− + 12H+(aq) + 6H2(g)
| (6) |
The chemical reduction method includes four steps:38
• Preparation of supersaturated solution.
• Nucleation of the nZVI cluster.
• Growth of nZVI nuclei.
• Agglomeration of nZVI.
Gas-phase reduction
This process uses gases such as H2, CO2 or CO acting as reducing agents to reduce iron salt or compound to nano form at higher temperature (>500 °C).33 The particles produces have an average particle size ranging from 50 to 300 nm with a specific area of about 7.55 m2 g−1.34 The reaction proceeds according to eqn (7) and (8):33 |
Fe(C2H3O2)2(aq) + C(g) → Fe0(s) + 2CH2CO + CO + H2O
| (7) |
|
Fe3O4(aq) + 2C(g) → 3Fe0(s) + 2CO2
| (8) |
Thermal decomposition
One synthesis method which is well-known commonly for producing higher quality particles with more narrow size distribution is the thermal decomposition of an iron precursor. This method is quite similar to the gas phase reduction previously explained. The iron precursor (iron oxide or an iron salt) are reduced at high temperatures (>500 °C) in the presence of gaseous reducing agents. The precursor/capping agent ratio, rate of heating, the final temperature of the reaction, and the annealing time are all very important to both the size and size distribution of the nanoparticles.39 The reducing agents used for this synthesis include N2, H2, CO, and Ar. If the reducing agent e.g. CO is produced as a result of thermally decomposing carbon-based materials such as biochar or carbon black, the overall synthesis is termed carbothermal decomposition/reduction.40 Hydrothermal syntheses of iron nanoparticles is performed in aqueous media in reactors or autoclaves where the pressure can be higher than 2000 psi and the temperature can be above 200 °C. The longer the reaction time, the larger the size of the nanoparticle synthesized, likewise the higher the water content, the greater the precipitation of nanoparticles. The shape of the nanoparticles also depends on the reflux time.41
Ultrasound assisted method in synthesis of nZVI
This method that was first synthesized by ref. 36 can be used to enhance either the physical or chemical methods. This method like the chemical method uses reducing agents like sodium borohydride and ammonium hydroxide to produce small, uniform and equal-axe grains of iron nanoparticle of average size of 10 nm.12 |
4Fe2+(aq) + BH4−(aq) + 7NH4OH → 4Fe0 + H3BO3(aq) + 7NH4+(aq) + 4H2O
| (9) |
The morphology of the nanoparticles depends on the frequency of the ultrasound that is used to assist the chemical reduction method in defining the shape of the particles.38
Precision milling
Precision milling method consists of applying mechanical force to crush micro iron with steel shot in a high speed rotary chamber for about 8 h without any chemical to achieve highly reactive nanoparticles of diameter 10–50 nm and surface area 39 m2 g−1. Upon contact with the steel shot, the particles are deformed and cracked producing nanoparticles with irregular shapes and with strong tendency to aggregate35 because of its higher surface energy.34
Electrochemical
This is a cheap and fast method in which cathodes are used to attract Fe2+/Fe3+ ions from solution by help of electric current. Cationic surfactants are used to act as a stabilizing agent and ultrasonic waves (20 kHz) as a source of energy necessary for fast removal of iron nanoparticles from the cathode:40 |
Cathode: Fe3+ + 3e− + stabilizer → nFe0
| (10) |
The electrochemical method can be augmented with ultrasound to produce particles of diameter 1–20 nm and specific surface area of 25.4 m2 g−1.42
Green synthesis
Plants. This is an inexpensive and environmentally friendly procedure where highly water soluble, less toxic, and biodegradable plant extracts are used to reduce iron to nanoscale. The extract is heated in water close to the boiling point and then mixed with the iron ion solution causing the irons to be reduced to nZVI in presence of polyphenols.43 The extracts serves two purposes, first as a reducing agent and then as a capping agent for Fe. The capping polyphenols further prevents oxidation as they are potential antioxidants and can also scavenge free radicals.44 The produced nZVI are spherical and range from 5 to 15 nm in size.43 However, one of the major concerns of this method is the destruction of plants and its parts.40
Bacterial. Green synthesis can also be achieved by using iron reducing bacteria such as Acidiphilium spp. Bacteria of the genus Acidiphilium known to be able to reduce Fe3+ to Fe2+ under both aerobic and anaerobic conditions as well as in acidic and iron-rich environments.45 This reduced species of iron including Fe0 have been confirmed by Mössbauer spectroscopy.46Table 3 gives a summary of the different synthesis methods of nZVI that have been described including the pros and cons of each.
Table 3 Comparison of different synthesis methods of nZVI
Method |
Advantages |
Disadvantages |
Liquid-phase reduction or borohydride reduction |
Nanoparticle produced are very reactive |
The reducing agents used can be toxic |
Synthesis procedure and methodology is easy to replicate |
Cost effective |
Ultrasound assisted method |
Simultaneous oxidation reaction occurring during synthesis |
The reducing agents used can be toxic |
Possible for producing very small sized particles |
Extra reagents are required to prevent oxidation of the nanoparticles |
Particles produced are of different morphologies |
Precision milling |
No toxic reagent needed |
Long milling time is required |
Ability to produce easily on large scale |
Nanoparticle size and shape produced are irregular |
Higher tendency of particle aggregating |
Electrochemical |
Short duration required |
Higher tendency of particle agglomerating |
Cost effective method |
Adjusting the current density can control the particle size |
Green synthesis |
Environmentally friendly method |
Possible competition with and destruction of food crops |
Easier to replicate on large scale |
Long duration required for bacterial based synthesis |
Cost effective |
Greater expertise required |
Thermal decomposition |
Possible to control particle size and morphology |
Higher energy requirement |
Increased yield |
Long duration |
The use of nZVI particles have proven to be an effective remediation technique for the reduction of chlorinated hydrocarbons including PCE, TCE, and PCBs using the dehalogenation pathway,2,47 dense non-aqueous phase liquid (DNAPL) contaminants in aquifers,48 metallic and metalloid contaminants, including Pb2+, Ni2+, Co2+, Cu2+, AsO43− and AsO33−.2 However, like any technique there exist some challenges to their application.
Challenges of nZVI
The application of nZVI works best under anaerobic conditions since it is otherwise quickly transformed to iron oxides under aerobic conditions.49 The large surface area of the nZVI is the strength of this technique as the degradation reaction occurs on the surface of the particles. Any modification to the surface such as oxidation would therefore affect the performance of the particles. Some of the major challenges identified with the use of nZVI are the rapid aggregation of the particles, passivation (quick oxidation by non-target compounds), sorption to aquifer materials and rapid sedimentation that consequently limits the mobility of nanoparticles in the aquatic environment.6,37 This also makes long term storage of nano iron particles materials not feasible as they must be used soon after production unless they are stabilized on a support to inhibit oxidation.50
One reason for its rapid sedimentation and high settling velocity would be the density of the iron particle which is 7800 kg m−3 coupled with the extremely short settling distances in aquifers.37 For instance in groundwater, they migrate a few inches to a few feet.51 The surface charge or zeta (ζ) potential of the iron nanoparticles influences their suspension stability and mobility in a particular matrix. Aquifer materials mostly have negative charges on their surface under typical groundwater conditions of neutral pH, thus iron nanoparticles with positive charges at pH lower than 8.3 are attracted to them which reduces their mobility.8,52
Also, once the particle concentration is increased, there is increase in intrinsic magnetic moment leading to strong magnetization which eventually reduces its surface area and reactivity. Aside the magnetic forces, dipole–dipole attraction may occur between the magnetic moments of the particles thus affecting their size and dispersion stability and so settle into aquifer media pores.14,15,53 When suspensions are highly concentrated with low viscosity, the kinetics of agglomeration is very high and Brownian motion is responsible for bringing particles within the small range of the interaction forces.52 The relatively high ionic strength of groundwater makes it favorable for the reduction of electrostatic repulsion between particles and increase of colloidal aggregation in water.6,37,54
Carriers used for nZVI and their synthesis
To overcome the challenges outlined in the section above, there would be the need to alter the surface charge of the iron particles.10 Two solutions proposed are electrostatic repulsion (imparting or increasing the surface charge) and steric stabilization (adsorption of long chain organic molecules).6,37,54 Electrostatic repulsion introduces strong long ranged repulsive forces needed to overcome the magnetic attraction between the iron nanoparticles. Steric stabilization on the other hand pre-coats the iron nanoparticles with hydrophilic polymers whose long loops and tails extend out into solution.55 The coats used can be organic or inorganic molecules but practically it is worthy that the protection shells do not only stabilize the nano particle but also serve as functionalization.56 The stabilization of the metal on a support thus prevents aggregation of the metal and further leaching.47,57 Another solution proposed is the coupling of the nZVI with other metals like palladium or nickel to form a bimetallic system.58
Organic substances
Monomers, surfactants and polymers. Organic compounds may be used as a support or carrier to passivate the surface of the nano iron particles during or after the synthesis procedure to limit their agglomeration. The carriers used are linked to the nano particles by chemical bond and may not only provide sufficient repulsive interactions to prevent agglomeration but also provides reactive functional groups such as aldehyde groups, hydroxyl groups, carboxyl groups, amino groups, etc. The organic substances used can be grouped under three categories; hydrophobic, hydrophilic and amphiphilic. Hydrophobic organic substances such as fatty acids and alkyl phenol have little affinity for water while hydrophilic substances like ammonium salt, polyol and lysine have strong attraction for water. Amphiphilic substances such as sulfuric lysine on the other hand exhibit both hydrophobic and hydrophilic properties.56Iron particles of hydrophobic supported functional groups form long-chains. Those supported on hydrophilic groups however have a better dispersion. The oil-soluble support can be transformed into water-soluble type functionalized form by ligand-exchange reaction.59
Polymers used can either be natural or synthetic. It may be used as a support to compensate or augment functional group deficit provided by monomers. Coating the surface of the iron particle with a polymer increases the repulsive forces to balance the magnetic and the van der Waals attractive forces acting on the nano particles. The functional groups introduced by the polymer would also decrease the saturation magnetization value of iron particle.56 Some examples of monomers, surfactant and polymers that have been used as carriers, their respective method of synthesis and the characterization results obtained are presented below:
Polyvinyl alcohol-co-vinyl acetate-co-itaconic acid (PV3A). PV3A is a nontoxic and biodegradable surfactant with molecular weight within the range of 4300–4400 (Fig. 2).
 |
| Fig. 2 Molecular structure of polyvinyl alcohol-co-vinyl acetate-co-itaconic. This figure has been adapted from ref. 37 with permission from Elsevier. | |
From Transmission Electron Microscopy (TEM) images, the bare nanoparticles are spherical and form a chain-like aggregate with median and mean size of 59.4 nm and 105.7 nm respectively (Fig. 3a). In contrast, PV3A-stabilized iron particles (Fig. 3b) show an even distribution with little aggregation and median and mean size of 7.9 nm and 15.5 nm respectively. Addition of PV3A alters the surface charge on the iron nanoparticle to predominately negative charges over a much wider pH range probably due to the dissociation of the carboxylic acid groups on the PV3A molecules, and as a result increases the surface potential and stability.
 |
| Fig. 3 TEM images of: (a) iron particles prepared with the sodium borohydride method and (b) iron particles stabilized with PV3A. The scale bar for both images is 200 nm. This figure has been adapted from ref. 37 with permission from Elsevier. | |
Polyvinylpyrrolidone (PVP). The synthesis of PVP-nZVI is done by chemical process using NaBH4 aqueous solution. The average diameter of the PVP-nZVI particles is 50–80 nm (Fig. 4). These particles tend to aggregate to form chain structures, as a result of both their magnetic properties and their tendency to remain in the most thermodynamically favorable state.15
 |
| Fig. 4 TEM image of PVP-nZVI. This figure has been adapted from ref. 15 with permission from Springer. | |
Polyacrylic acid (PAA). The PAA-nZVI synthesis is achieved by chemical processes using NaBH4 aqueous solution by adding 8% PAA (v/v) to iron precursors so that iron complexes would form through carboxylate ligand interaction. The particles produced have a diameter of about 9–15 nm with an average of 12 nm.60
Ethyl acrylate & 1,7-octadiene copolymer. A copolymer in the form of beads is prepared by mixing ethyl acrylate and 1,7-octadiene before adding to benzoyl peroxide. The mixture is then suspended in an aqueous phase of gum-Arabic and gelatin under mechanical stirring at a temperature of 80 °C. The subsequent copolymer-nZVI synthesis is done by sonication. The particles sizes of the product ranges from 25–75 nm with some overlaps (Fig. 5).
 |
| Fig. 5 Scanning electron microscope image of synthesized iron nanoparticles (a) at 10 μm and (b) at 0.5 μm. This figure has been adapted from ref. 61, open access (Hindawi Publishing corporation). | |
Guar gum. Guar gum is a nontoxic biodegradable green polymer polysaccharide of high molecular weight. Iron nanoparticles that are stabilized on guar gum are much smaller with no aggregation even at very high ionic strength. Guar gum effectively reduces the hydrodynamic radius of bare nanoparticles from 500 nm to less than 200 nm and prevents aggregation of nanoparticles even at very high salt concentrations (0.5 M NaCl and 3 mM CaCl2).55
Xanthan gum. Xanthan gum is a polymer with the ability to increase the velocity of a liquid and thus attain kinetic stabilization. The xanthan gum-nZVI synthesis is done by ultra-sonication. 6 g L−1 of xanthan gum gels is able to avoid the sedimentation of 30 g L−1 and the aggregation of 15 g L−1 iron nanoparticles over a period of 10 days.62
Polystyrene. Polystyrene is a porous polymer that may be able to remain stable under reaction conditions of higher temperatures. The polystyrene-nZVI synthesis is achieved by thermal reduction process. TEM images show a well dispersed nano Fe particles with average size of 44 nm.26
Polyvinylpyrrolidone (PVP-K30). Polyvinylpyrrolidone (PVP-K30) is water-soluble, low-cost, environmentally friendly and used commonly in food processing. The PVP-K30-nZVI synthesis is done following the chemical reduction process using NaBH4. The composite particles has a surface area of 36.90 m2 g−1 and size range of 10–40 nm in diameter.63
Poly(methylmethacrylate) (PMMA)/anisole co-modified nZVI (PnZVI). The PMMA/anisole support is prepared by dissolving PMMA particles in anisole. The PMMA/anisole-nZVI synthesis is done by chemical reduction process using NaBH4. The particle specific surface area of the composite is 42.32 m2 g−1.7
Polystyrene resins. nZVI was encapsulated within porous polystyrene resins functionalized with –CH2Cl and –H2N+(CH3)3 respectively. The synthesis was done by the chemical reduction process using NaBH4. The particle size was in the range of 5–20 nm (ref. 64 and 65) using the similar procedure synthesized nZVI on PolyFlo resin.
Carboxymethyl cellulose (CMC). CMC is a biodegradable food-grade ingredient which is formed by modifying cellulose by replacing the native CH2OH group in the glucose unit with a carboxymethyl group (Fig. 6). CMC can accelerate nucleation of Fe atoms during the formation of nZVI. The composite products have a bulky negatively charged layer which provides electrosteric stabilization, thus a combination of electrostatic stabilization and steric stabilization.66
 |
| Fig. 6 The molecular structure of CMC. This figure has been adapted from ref. 67, open access (AUTEX RJ). | |
The CMC-nZVI synthesis is done following the chemical processing using NaBH4. The well dispersed composites are spherical in shape and with sizes ranging from 20 to 100 nm (Fig. 7).66
 |
| Fig. 7 TEM images of Fe0 nanoparticles prepared (a) without CMC, and (b) with CMC. This figure has been adapted from ref. 66 with permission from Elsevier. | |
Inorganic substances
Alternatively, the use of inorganic compounds as support or carrier can greatly reduce the oxidation rate of nano iron particles. Some of the most common inorganic substances used include silica, metal, nonmetal, metal oxides, and sulfides. Other advantages of the use of inorganic support are:56
• The coating provides both stability and serves as a barrier to limit inter-particle interactions which can result in agglomeration.
• This composite nano particles possess an improved biocompatibility, hydrophilicity and stability.
• The technology of preparation for size tunable composite nano particles is already mature, and it is quiet easy to control the variation of the shell.
Some examples of inorganic supports include:
Clay. Materials that become plastic when mixed with a small amount of water are often referred to as clay. The two most important and common clay minerals are kaolin and montmorillonite. Their structural configuration, which helps them to hold exchangeable cations and anions on their surfaces, makes them good materials for immobilization of pollutants either by ion exchange mechanism or adsorption. The ions held on their surfaces include Ca2+, Mg2+, H+, K+, SO42−, Cl−, PO43−. Other reaction mechanisms for holding pollutants may include H-bonding, van der Waals interactions or hydrophobic bonding arising from either strong or weak interactions.68
Kaolin. Kaolin is a 1
:
1 structured clay consisting of a tetrahedral sheet of SiO4 and an octahedral sheet with Al3+ as the octahedral cation.68 Kaolin clay is formed as a result of the alteration of alumosilicates (feldspar, feldspathoid, spodumene, sillimanite) and volcanic glasses, sometimes altered by acidic hydrothermal solutions.69 The kaolin-nZVI synthesis is done following the chemical reduction process using NaBH4. The produced nZVI has an average size of 44.3 nm and a mean specific surface area of 26.11 m2 g−1. The mean surface area value obtained is 7.5 times larger than kaolin alone (3.67 m2 g−1) (Fig. 8).70 Using the same support material and synthesis procedure,71 obtained a BET and Langmuir surface areas of 9.6 and 34.8 m2 g−1 respectively and sizes between 10 and 80 nm.
 |
| Fig. 8 SEM scanning images of samples. (a) Kaolin (k); (b) nZVI; (c) and (d) k-nZVI. This figure has been adapted from ref. 70 with permission from Elsevier. | |
Hexadecyl trimethylammonium modified montmorillonite (HDTMA-Mont). The support is prepared by dissolving known grams of montmorillonite in deionized water by ultrasound for 15 min. A solution of HDTMA is subsequently added under magnetic stirring for 3 h at 60 °C. The synthesis of the HDTMA-Mont-nZVI is by chemical process using NaBH4 aqueous solution. The composite product has specific surface area of 38.1 m2 g−1 (Fig. 9).
 |
| Fig. 9 TEM images of: (a) unsupported iron nanoparticles, (b) HDTMA-Mont/iron particles: iron nanoparticles supported by HDTMA-Mont, (c) Mont/iron: iron nanoparticles supported by montmorillonite and (d) Mont: montmorillonite particles. This figure has been adapted from ref. 72 with permission from Elsevier. | |
Bentonite. Bentonite is a traditional low-cost efficient adsorbent, with chemical and mechanical stability, high adsorption capability and unique structural properties. The bentonite-nZVI synthesis is done following the chemical reduction process using NaBH4. The synthesized particles has diameters in the range of 20 and 90 nm and surface area of 39.94 m2 g−1 (Fig. 10).73
 |
| Fig. 10 SEM images of laboratory synthesized iron particles with and without a support material. (a) nZVI (b) B-nZVI before reaction with Cr(VI) solution (c) B-nZVI after reaction with Cr(VI) solution. The scale bar for all images is 500 nm. This figure has been adapted from ref. 73 with permission from Elsevier. | |
Zeolite. Zeolites are microporous aluminosilicate minerals commonly used as adsorbents for removing several pollutants. The zeolite-nZVI synthesis is done following the chemical reduction process using NaBH4. The mean surface area of the composite is 80.37 m2 g−1, compared to 12.25 m2 g−1 for nZVI and 1.03 m2 g−1 for the zeolite alone (Fig. 11).74
 |
| Fig. 11 SEM images of (a) nZVI particles and (b) zeolite. This figure has been adapted from ref. 74 with permission from Elsevier. | |
In a similar synthesis using NaY zeolite as the support,75 produced spherical composite particles with diameters in the range of 50 to 100 nm.
Sepiolite. Sepiolite is a soft white clay mineral which is stable in air; it has a good dispersibility in water, cost-effective and with good adsorption properties. The sepiolite-nZVI synthesis was done following the chemical reduction process using NaBH4. The TEM images of the product show nZVI with particle size range of 20 to 60 nm and with little aggregation (Fig. 12).76
 |
| Fig. 12 TEM images of (a) nZVI, (b) sepiolite-supported nZVI.76 This figure has been adapted from ref. 66, free access (Atlantis press). | |
Marine clay. Marine clay is found in coastal regions around the world. The marine clay-nZVI synthesis is done following the chemical reduction process using NaBH4. The product obtained is composed of homogenous rounded particles with size range from 29 to 57 nm.77
Corals. Coral collected from Persian Gulf coast is prepared by cleaning, crushing and drying at 100 °C before being sieved through a 100 mesh. The synthesis of the nZVI is by chemical process using NaBH4 aqueous solution. The sieved coral is added before NaBH4 during the synthesis procedure. A scanning electron microscope (SEM) image (Fig. 13) shows that most of the sheet structure is changed to irregular small particles with size range of 53–94 nm and a surface area of 16.85 m2 g−1.78
 |
| Fig. 13 SEM of nano iron particles synthesized on coral. The scale bar is 500 nm. This figure has been adapted from ref. 78, open access (CEST). | |
Oyster shells. Oyster shells collected from the coast of the Persian Gulf is prepared by washing in deionized water, crushed and dried in 100 °C oven before sieving with a mesh of 100. The oyster shell-nZVI synthesis is done following the chemical processing using NaBH4.
Biochar (BC). Biochar is derived from a variety of cost effective and readily available biomass materials such as wood, leaves, manure and agricultural wastes through the pyrolysis. In addition, it does not need a treatment for further activation like active carbon.79,80 The preparation of the material is done by mixing biochar (<100 mesh) with 1 M HCl (1/20, v/v) and shaken overnight at room temperature for demineralization. It is subsequently purified using dialysis method until the solution pH is close to neutral, and dried at 80 °C in an oven. The BC-nZVI synthesis is done by chemical process using NaBH4 aqueous solution. A Brunauer, Emmett and Teller (BET) images after synthesis show a surface area of 142.80 m2 g−1. TEM images shows nZVI nanoparticles appeared to be dispersed on the surface and the boundary sites of biochar (Fig. 14)81 used chitosan, a renewable transformed polysaccharide that can be obtained from natural chitin as an ‘organic glue’ to cement nZVI on BC. Chitosan is an abundant biopolymer obtained by the alkaline deacetylation of chitin (a polymer made up of acetylglucosamine units).57 The BC-nZVI synthesis is done by chemical process using 1.2% NaOH aqueous solution.
 |
| Fig. 14 (a) Transmission electron microscope and (b) X-ray diffraction images of nZVI, BC and nZVI/BC. This figure has been adapted from ref. 79, open access (Plos One). | |
Carbon black. Carbon black is a low cost and abundant waste product from the fossil fuel industry. The carbon black-nZVI synthesis is achieved by carbothermal synthesis at 700 °C for 4 h. The produced particles are of surface area 108.67 m2 g−1 and size range of approximate 20–150 nm.82 Composite of carbon functionalized nano iron particles have higher chemical and thermal stability as well as biocompatibility.56
Cationic exchange membrane (CEM). The CEM used consists of hydrophilic polysulfone on polyester support with a pore size of 0.45 μm. The CEM is first prepared by soaking it in 0.2 M Fe(III) solution overnight before washing with distilled water. The subsequent CEM-nZVI synthesis is done by chemical process using NaBH4 aqueous solution. The produced nZVI are spherical in shape with size range of 30–40 nm and surface area of 1.90 m2 g−1 (Fig. 15).
 |
| Fig. 15 SEM images of ZVI-immobilized Fe0-CEM. This figure has been adapted from ref. 58 with permission from Elsevier. | |
Calcium (Ca)–alginate beads. Calcium (Ca)–alginate is one of the most common cost effective materials for immobilizing living cells in food and for drug delivery. The Ca–alginate-nZVI synthesis is done by chemical process using NaBH4 aqueous solution. TEM images indicate nZVI of particle size ranges from 10 to 100 nm with an average size of 35 nm. A higher magnification TEM image shows a 2.5 nm of oxide shell around the nZVI core (Fig. 16).
 |
| Fig. 16 (a) An alginate bead with entrapped nZVI, (b) SEM image of alginate bead. (c) SEM image of alginate bead surface after nZVI entrapment. (d) Higher magnification of SEM image (b), (e) TEM image of Ca–alginate bead section shown in (d), (f) blow-up image of the circled area in image (e). This figure has been adapted from ref. 53 with permission from Elsevier. | |
Silica fume. Silica fume is a cheap, non-toxic and abundant materials produced during silicon metal production from a blast. The silica fume-nZVI synthesis is done following the chemical reduction process using NaBH4. The iron particles produced are spherical with size ranging from 20 to 110 nm.83
Graphene oxide (GO). GO obtained from cheap natural graphite is strongly hydrophilic due to the presence of carboxyl, hydroxyl and epoxide groups on its nano sheets. The GO-nZVI synthesis is done following the chemical reduction process using NaBH4. The produced nano composites have particle size of about 5 nm (Fig. 17).84
 |
| Fig. 17 TEM images of (a and b) GO, (c and d) Fe and (e and f) composite ration 1 : 5. This figure has been adapted from ref. 84 with permission from Elsevier. | |
Multi-walled carbon nanotubes (MWCNTs). Multi-walled carbon nanotubes (MWCNTs) used as a porous-based support material for the nZVI nanoparticles are synthesized by chemical reduction process using NaBH4. The composite particle has a polygon shape with diameter range between 20–80 nm. However, aggregation of nZVI particles is still observed, but much decreased relative to the unsupported nZVI nanoparticles85 (Fig. 18).
 |
| Fig. 18 (a) SEM and TEM images of MWCNTs, (b) SEM and TEM images of nZVI supported on MWCNTs before reaction, (c) SEM images of nZVI supported on MWCNTs after reaction. This figure has been adapted from ref. 85 with permission from Elsevier. | |
Other carriers like poly(styrene sulfonate), carboxymethyl, polyaspartate86 in addition to those already mentioned have proven their efficiency in reducing aggregation and enhancing nZVI mobility. Those that have been used for depollution purposes have also shown to be more effective compared to the bare nZVI.
In as much as it may be advisable to synthesize nZVI on a support in order to reduce aggregation and in effect improve its effectiveness, care must be taken not to add up more pollution and risk to the environment. The reason for a choice of a carrier should be well defined. Table 4 compares several of these carriers and accesses the advantages and disadvantages they carry when used during nZVI synthesis.
Table 4 Comparison of the different nZVI carriers
Carrier |
Method of synthesis |
Advantage |
Disadvantage |
Functional groups |
Reference |
Polyvinyl alcohol-co-vinyl acetate-co-itaconic acid (PV3A) |
Chemical reduction |
Non toxic |
Carrier covers large fraction of reactive surface of the nZVI and can thus affect reactivity especially if there is no evidence of sorption or partitioning between nZVI and the pollutant(s) |
–OH, –CO–, and –COOH |
37 |
Biodegradable surfactant mainly due to the presence of functional groups such as –OH, –CO–, and –COOH |
86% reduction in average size of nZVI particles |
Reduction in the zeta (ζ)-potential at neutral pH |
A shift of the isoelectric point (IEP) |
No observed sedimentation of stabilized nZVI |
Presence of carrier presents both steric and electrostatic repulsions |
Polyvinylpyrrolidone (PVP) |
Chemical reduction |
Cost effective material |
PVP-nZVI has an amorphous structure |
C O |
15 |
80% reduction in average size of nZVI particles |
The presence of excess PVP can serve as a bridge between discrete PVP-nZVI particles and increase the sedimentation time |
Reduced sedimentation time |
Effective even in porous soil media treatment |
Reduced zeta potential than other carriers such as CMC and PAA |
Polyacrylic acid (PAA) |
Chemical reduction |
Aside the partial bidentate bridging of PAA to nZVI, it also forms a gel network through hydrogen bonding to trap colloidal particles to stabilize dispersions. However the presence of this gel can be destroyed with increase in concentration of cations like Ca |
PAA-nZVI forms poorly ordered and amorphous structures |
–OH, –COO–, and C–H |
60 |
Ethyl acrylate & 1,7-octadiene copolymer |
Chemical reduction and sonication |
Acts as a cation exchanger and thus enriches the cationic byproduct of the aqueous media |
The carrier being a weak acid support has a lower reactivity than strong acid supports |
C O and C–H |
61 |
Capable of adsorbing trivalent ions |
Guar gum |
Sonication |
Cost effective material |
Sedimentation occurs after few hours |
O–H |
55 |
Non-hazardous green (of natural origin) material |
Water soluble and biodegradable |
60% reduction in average size of nZVI particles |
Higher molecular weight polymers provide strong long-ranged steric repulsion required for stabilization |
Xanthan gum |
Ultrasonication |
Non-hazardous and biodegradable |
N.A |
C O |
62 |
Possibility of stabilizing for long time (more than 10 days) |
It has wide stable velocity making it simple to dose even on the field |
The polymer is arranged in a network structure making it possible for it to trap colloidal particles and increase stabilization |
Polystyrene |
Thermal decomposition |
Ability to remain stable even at higher temperatures (250–300 °C) |
N.A |
C O and C–H |
26 |
Polyvinylpyrrolidone (PVP-K30) |
Chemical reduction |
Higher adsorption capacity to adsorb both pollutant and degraded by-products |
Adsorption of by-products means that toxicity may be spread from the plume site if the by-products are toxic |
C O, C–N and C–H |
63 |
Cost effective material |
Non-hazardous and environmentally friendly |
Water soluble and biodegradable |
Poly(methylmethacrylate) (PMMA)/anisole co-modified nZVI (PnZVI) |
Chemical reduction |
Reduce significantly surface oxidation of nZVI |
Reduced reactivity of the nZVI due to blocking of reactive site by carrier |
C O, –COO– and C–H |
7 |
Polystyrene resins (–CH2Cl and –H2N+(CH3)3) |
Chemical reduction |
The presence of ammonium group helps in reducing nZVI particle size |
Excess layers of carrier can reduce reactivity |
C–N and C–H |
64 |
Carboxymethyl cellulose (CMC) |
Chemical reduction |
Provides steric stabilisation Environmentally friendly material |
The formation of surface passivation layer on the nZVI reduces the reactivity |
–CO–, –COOH and OH |
66 |
Non-toxic and water soluble |
Kaolin |
Chemical reduction |
Acts as natural scavengers |
Reduces the specific surface area of nZVI |
|
70 and 71 |
Higher mechanical and chemical stability |
Edge sites contain more nZVI particles than the surface sites of the clay mineral |
Higher cation exchange capacity |
Increases the reactivity of nZVI |
Hexadecyl trimethylammonium modified montmorillonite (HDTMA-Mont) |
Chemical reduction |
Abundant, cost effective and environmentally friendly material |
An additional surfactant like hexadecyl trimethylammonium may be needed to increase reactivity |
|
72 |
Has an increased specific surface area and average pore width |
Bentonite |
Chemical reduction |
Traditionally abundant and cost effective absorbent |
Reaction products are deposited on the surface of B-nZVI which consequently decreased the activity of the nZVI |
|
73 |
Higher mechanical and chemical stability |
Unique structural properties |
Zeolite |
Chemical reduction |
Higher sorption capacity for inorganic pollutants |
N.A |
|
74 |
85% increase in mean surface area |
Metallic pollutants removed are incorporated into the internal matrix of the composite making them non exchangeable and non-bioavailable |
Sepiolite |
Chemical reduction |
Cost effective material |
N.A |
|
76 |
Good stability in air |
Good dispersion in water and with equally good adsorption properties |
Corals, oyster shells |
Chemical reduction and wet impregnation |
Cost effective and abundant material |
May not be ideal for treating hydrophilic pollutants |
|
78 and 87 |
Good stability in air |
Biochar (BC) |
Chemical reduction |
Cost effective and readily available material |
Excess biochar loading on the nZVI particles may block reactive sites thereby reducing reactivity |
|
79 |
Higher adsorption capacity and stable structure |
It does not need a treatment for further activation |
85% increase in surface area |
Biochar + chitosan (organic glue) |
Chemical reduction |
Polysaccharide is biodegradable and non-toxic |
Chitosan decomposes at much lower temperature than pyrogenic carbon |
|
81 |
Chitosan has the tendency to dissolve in acid solution and precipitate from basic solution |
Chitosan is nontoxic, hydrophilic, biocompatible and biodegradable polymer |
Carbon black |
Thermal decomposition |
Low cost and abundant material |
Higher temperature of about 700 °C for over 4 h is required to have a higher surface area. This may increase production cost |
|
56 and 82 |
Higher chemical and thermal stability |
Cationic exchange membrane (CEM) |
Chemical reduction |
Release of metallic pollutants from the support is very minimal |
N.A |
|
58 |
Calcium (Ca)–alginate beads |
Chemical reduction |
The porosity of the composite allows solutes to diffuse into the beads and come in contact with the entrapped cells |
Heterogeneity of pore size results in nanoparticle being entrapped more at some places (agglomeration) than others |
|
53 |
Slight reduction in the reactivity of the nZVI |
Non-toxic and biodegradable |
Silica fume |
Chemical reduction |
Non-toxic, abundant and cost effective material |
N.A |
|
83 |
Higher surface area to homogenously disperse nZVI |
The dominant negatively surface is useful for migration through the soil |
Graphene oxide (GO) |
Chemical reduction |
90% reduction in particle size |
N.A |
|
84 |
Good thermal and mechanical properties |
Multi-walled carbon nanotubes (MWCNTs) |
Chemical reduction |
nZVI acts as a micro-anode and MWCNTs as a micro-cathode which prevents the formation of oxide film on the surface of the nZVI |
N.A |
|
85 |
Generally, the choice of a carrier whether organic or inorganic is influence by;
(1) Availability of the product.
(2) Cost of the carrier.
(3) Environmental friendliness of the carrier (biodetgradability).
(4) The intended use or target pollutant to be treated.
Aside the above mention general factors that affect the choice of carriers, the physical and chemical properties of the carrier are also considered. Organic carriers for example are chosen to alter the surface charge of nZVI particles. These surface charges provide both electrostatic stabilization and steric stabilization and thus reduce aggregation. On the other hand, inorganic carriers have naturally good adsorption properties, surface ions, and consequently exhibit good ion exchange mechanisms.
The TEM images show that nZVI are well dispersed on their respective carriers compared to their bare forms. This then confirms that the use of carriers reduces aggregation and preserves the smaller size and large surface areas of nZVI which is the heart of their performance. However the kind of carrier used and the pollutant to be treated would affect the output efficiency during remediation. Table 5 summarizes different carriers, the pollutants they have been used to treat and the results obtained. It wasn't very clear the reasons that informed these studies to choose one particular carrier over the other. Since the carrier would modify the properties of the nZVI including its toxicity in the environment, it is very important to access both advantages and disadvantages of the intended carrier in relation to the pollutant to be treated before selection.
Table 5 Composite products used for pollution remediation
Support material |
Optimum ratio (support–nZVI) |
Pollutant |
Initial concentration of pollutants (mg L−1) |
Duration (min) |
Efficiency (%) |
References |
Coral |
NA |
Humic acid |
2 |
90 |
84.5 |
78 |
Biochar |
5 : 1 |
Methyl orange |
60 |
30 |
98.5 |
79 |
Biochar–chitosan |
1 : 1 : 2 |
Pb |
NA |
NA |
93 |
81 |
Cr |
38 |
As |
95 |
Granular activated carbon (GAC)–Pd |
NA |
2-Chlorobiphenyl (2-ClBP) |
4 |
1152 |
90 |
88 |
Green tea (Camellia sinensis) polyphenols |
NA |
Bromothymol blue |
580 |
60 |
34 |
89 |
Cationic exchange membrane |
NA |
Trichloroethylene (TCE) |
59.9 |
1080 |
80 |
58 |
Calcium (Ca)–alginate beads |
NA |
Nitrate (NO3−) |
100 |
120 |
73 |
53 |
Magnesium-aminoclay (MgAC) |
7.5 : 1 |
Nitrate (NO3−N) |
100 |
30 |
75 |
90 |
Carboxymethyl cellulose (CMC) |
NA |
Perchloroethylene (PCE) |
9.6 |
360 |
80 |
47 |
Trichloroethylene (TCE) |
15 |
180 |
100 |
Kaolin |
5 : 1 |
Pb2+ |
500 |
20 |
96 |
70 |
Kaolin |
5 : 1 |
Pb2+ |
12.75 |
60 |
98.8 |
13 |
Total Cr |
71 |
99.8 |
Cu2+ |
27.25 |
69.5 |
Zn2+ |
216.45 |
28.1 |
Ni2+ |
8 |
23.0 |
Kaolin |
0.2 : 1 |
Cu2+ |
100 |
1440 |
99 |
71 |
Co2+ |
98 |
Oyster shell |
NA |
Humic acid |
5 |
120 |
94.6 |
87 |
Cellulose acetate |
NA |
Trichloroethylene |
80 |
2880 |
53 |
50 |
Chitosan |
NA |
Cr(VI) |
40 |
180 |
74.03 |
91 |
Bentonite |
1 : 1 |
Cr(VI) |
75 |
240 |
100 |
73 |
Pb2+ |
13 |
100 |
Cu2+ |
34 |
92.7 |
Zn2+ |
288 |
59.4 |
Zeolite |
NA |
Pb2+ |
100 |
140 |
96.2 |
74 |
NaY zeolite |
NA |
Potassium acid phthalate (KHP) |
425 |
120 |
79 |
75 |
Sepiolite |
NA |
Decabromodiphenyl ether (BDE-209) |
NA |
23 040 |
48.64 |
76 |
HDTMA-Mont |
NA |
Cr(VI) |
50 |
350 |
82 |
72 |
Silica fume |
NA |
Cr(VI) |
40 |
120 |
88 |
83 |
Graphene oxide |
5 : 1 |
Methyl blue |
0.5 |
15 |
99 |
84 |
Carbon black |
NA |
Uranium |
580 |
10 080 |
98 |
82 |
Multiwalled carbon nanotube (MWCNT) |
2 : 1 |
Cr(VI) |
20 |
120 |
98 |
85 |
Carboxymethyl cellulose (CMC) |
5 : 1 |
Cr(VI) |
8 |
60 |
98 |
66 |
Polystyrene resins |
NA |
Nitrate |
50 |
750 |
97.2 |
64 |
Polyvinylpyrrolidone (PVP-K30) |
NA |
Tetracycline (TC) |
35 |
240 |
93 |
63 |
Polyvinyl alcohol-co-vinyl acetate-co-itaconic acid (PV3A) |
NA |
Trichloroethene (TCE) |
7 |
180 |
99 |
37 |
Polysorbate 20 (Tween® 20) |
NA |
As(III) |
1 |
5760 |
100 |
92 |
Fly ash |
NA |
Phosphate ion |
2 |
50 |
97 |
93 |
Starch |
NA |
Trichloroethene (TCE) |
25 |
60 |
98 |
94 |
PCB |
2.5 |
60 000 |
80 |
Starch |
NA |
Tetracycline (TC) |
500 |
14 400 |
99 |
95 |
Hydrophilic carbon |
NA |
Trichloroethene (TCE) |
13 |
180 |
97 |
96 |
PMMA/anisole |
NA |
Methyl orange |
100 |
60 |
81.4 |
7 |
Sunset yellow |
74 |
Acid fuchsine |
69 |
Bamboo |
NA |
Methylene blue |
140 |
120 |
79.6 |
97 |
Triton X-100 surfactant |
NA |
Reactive Black 5 (RB5) |
500 |
180 |
27 |
98 |
Cetyl trimethyl ammonium bromide (CTAB) surfactant |
35 |
It is however noteworthy that the use of carrier may not automatically signify an improved efficiency. Adsorbed polyelectrolyte coatings like poly(styrene sulfonate) (PSS), carboxymethyl cellulose (CMC), and polyaspartate (PAP) have demonstrated their ability in reducing the reactivity of nZVI to degrade TCE in water.99 Some surfactant coatings were also found to decrease the ZVI reactivity toward Reactive Black 5.98 Additionally PMMA/anisole hybrid coatings demonstrated an inhibition to the discoloration reactivity of nZVI toward organic dyes.7 The combinations of carriers do not also necessarily add any synergic effect and can also eventually inhibit the reactivity of the iron particles.100
Several results from both TEM and SEM showing dispersion and average particle size of nZVI after been synthesized on various carriers have been presented in this study. What is however missing is whether the results are based on the carrier used or on the synthesis method. It would be very interesting to explore this angle using the same carriers but with different synthesis methods. Another part missing in many publications reviewed is the ratio between the carrier used and the nZVI during synthesis. The information is very vital for future replication of similar experiments.
Traceability of nZVI
For a remediation treatment to be effective and fast, the nZVI must be as mobile as possible. The use of carriers to enhance this mobility has proven to work. However in as much as we want to enhance the mobility of nZVI by using carriers, care must be taken not to generate undesirable side-effects specifically on the potential of nZVI to adsorb and transport contaminants away from the primary contamination zone.101 There are strong evidence to buttress this concern as both natural colloids and nZVI have demonstrated the capacity to sorb contaminants.34
It is therefore necessary to track the nZVI used to effectively determine its fate in the environment. This is however a challenge because of the high background of natural iron colloids present in soils and other environmental systems. One suggested solution to this issue would be to label synthesized nZVI to make them easily traceable and to differentiate them from natural sources.34 Very few studies have investigated the fate of synthesized nano particles in the environment though it is estimated that the most released nano particles are titanium dioxide followed by iron and zinc respectively. About 63–91% of the global synthesized nano particles end up into landfills, 8–28% into the soils, 0.4–7% into water bodies and 0.1–1.5% into the atmosphere.102
Conclusion
Several carriers have been used with the aim of overcoming challenges like quick agglomeration and oxidation experience in the application of nZVI some of which has achieved very good results. However, the selection and use of a support should be carefully considered because the kind of support material used can decrease the reactivity of the nZVI especially if the optimum balance ratio between the support and the nano iron is not well done. In such instances, the adsorbed surface excess of the thick layers of coating can directly inhibit the mass transfer of target pollutants, block access to reactive surface sites, reduce indirectly the corrosion reaction of the iron particles with water or H+ and the electron transfer to target pollutants.7,99 With the popularity nZVI is gaining in the field of remediation, it is obvious that its use would keep increasing. However some gray areas about this technique still exist especially in the possibility of the carriers used adsorbing pollutants and thus increasing their migration from the plume as well as the toxicity of the injected nZVI. Future researchers need to focus on these unclear areas and seek to provide some answers to these uncertainties.
Acknowledgements
This study was funded by SFR Condorcet-FR CNRS 3417. The authors would like to express their appreciation to the Institut Polytechnique LaSalle Beauvais, France, and Dr Christophe WATERLOT (Equipe Sols et Environnement-LGCgE) (EA 4515) for their support in various ways.
References
- N. C. Mueller and B. Nowack, Nanoparticles for Remediation: Solving Big Problems with Little Particles, Elements, 2010, 6(6), 395–400 CrossRef CAS.
- X. Li, D. W. Elliott and W. Zhang, Zero-Valent Iron Nanoparticles for Abatement of Environmental Pollutants: Materials and Engineering Aspects, Crit. Rev. Solid State Mater. Sci., 2006, 31(4), 111–122 CrossRef CAS.
- M. R. Wiesner, G. V. Lowry, P. Alvarez, D. Dionysiou and P. Biswas, Assessing the Risks of Manufactured Nanomaterials, Environ. Sci. Technol., 2006, 40(14), 4336–4345 CrossRef CAS PubMed.
- J. P. Vyjayanthi, Immobilization of nanoscale zerovalent metals for the reductive removal of recalcitrant pollutants, Int. J. Environ. Sci., 2012, 2(4), 2173 CAS.
- R. Kober, D. Schafer, M. Ebert and A. Dahmke, Coupled in situ reactors using Fe0 and activated carbon for the remediation of complex contaminant mixtures in groundwater, IAHS Publ., 2002, 435–440 Search PubMed.
- P. Bardos, B. Bone, P. Daly, D. Elliott, S. Jones and G. Lowry, et al., A Risk/Benefit Appraisal for the Application of Nano-Scale Zero Valent Iron (nZVI) for the Remediation of Contaminated Sites, 2014 Search PubMed.
- X. Wang, J. Yang and M. Zhu, Effects of PMMA/anisole hybrid coatings on discoloration performance of nano zerovalent iron toward organic dyes, J. Taiwan Inst. Chem. Eng., 2014, 45(3), 937–946 CrossRef CAS.
- Y.-P. Sun, X. Li, J. Cao, W. Zhang and H. P. Wang, Characterization of zero-valent iron nanoparticles, Adv. Colloid Interface Sci., 2006, 120(1–3), 47–56 CrossRef CAS PubMed.
- M. Auffan, J. Rose, J.-Y. Bottero, G. V. Lowry, J.-P. Jolivet and M. R. Wiesner, Towards a definition of inorganic nanoparticles from an environmental, health and safety perspective, Nat. Nanotechnol., 2009, 4(10), 634–641 CrossRef CAS PubMed.
- K. Mackenzie, A. Schierz, A. Georgi and F.-D. Kopinke, Colloidal Activated Carbon and Carbon-Iron-Novel Materials for in situ Groundwater Treatment, 2008, cited 2016 Mar 11, available from: https://www.researchgate.net/profile/Katrin_Mackenzie2/publication/237555806 Search PubMed.
- S. M. Ponder, J. G. Darab, J. Bucher, D. Caulder, I. Craig and L. Davis, et al., Surface Chemistry and Electrochemistry of Supported Zerovalent Iron Nanoparticles in the Remediation of Aqueous Metal Contaminants, Chem. Mater., 2001, 13(2), 479–486 CrossRef CAS.
- M. R. Jamei, M. R. Khosravi and B. Anvaripour, Investigation of ultrasonic effect on synthesis of nano zero valent iron particles and comparison with conventional method: utlrasonic effect on synthesis of nZVI particles, Asia–Pac. J. Chem. Eng., 2013, 8(5), 767–774 CrossRef CAS.
- X. Zhang, S. Lin, X.-Q. Lu and Z. Chen, Removal of Pb(II) from water using synthesized kaolin supported nanoscale zero-valent iron, Chem. Eng. J., 2010, 163(3), 243–248 CrossRef CAS.
- T. Phenrat, N. Saleh, K. Sirk, R. D. Tilton and G. V. Lowry, Aggregation and Sedimentation of Aqueous Nanoscale Zerovalent Iron Dispersions, Environ. Sci. Technol., 2007, 41(1), 284–290 CrossRef CAS PubMed.
- B. Liang, Y. Xie, Z. Fang and E. P. Tsang, Assessment of the transport of polyvinylpyrrolidone-stabilised zero-valent iron nanoparticles in a silica sand medium, J. Nanopart. Res., 2014, 16(7) DOI:10.1007/s11051-014-2485-0 , available from: http://link.springer.com/10.1007/s11051-014-2485-0.
- D. R. Baer, P. G. Tratnyek, Y. Qiang, J. E. Amonette, J. Linehan and V. Sarathy, et al., Synthesis, Characterization, and Properties of Zero-Valent Iron Nanoparticles, Environmental Applications of Nanomaterials Synthesis, Sorbents and Sensors, 2007 Search PubMed.
- S. M. Cook, Assessing the use and application of zero-valent iron nanoparticle technology for remediation at contaminated sites, Jackson State Univ [Internet], 2009 [cited 2016 Aug 10], available from: http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.570.9438%26rep=rep1%26type=pdf Search PubMed.
- C. Wang, D. R. Baer, J. E. Amonette, M. H. Engelhard, J. Antony and Y. Qiang, Morphology and Electronic Structure of the Oxide Shell on the Surface of Iron Nanoparticles, 2009 Search PubMed.
- M. Krajewski, W. S. Lin, H. M. Lin, K. Brzozka, S. Lewinska and N. Nedelko, et al., Structural and magnetic properties of iron nanowires and iron nanoparticles fabricated through a reduction reaction, Beilstein J. Nanotechnol., 2015, 6, 1652–1660 CrossRef CAS PubMed.
- Y.-C. Lee, K. Lee, Y. Hwang, H. R. Andersen, B. Kim and S. Y. Lee, et al., Aminoclay-templated nanoscale zero-valent iron (nZVI) synthesis for efficient harvesting of oleaginous microalga, Chlorella sp. KR-1, RSC Adv., 2014, 4(8), 4122–4127 RSC.
- X. X. Zhang, G. Wen, S. Huang, L. Dai, R. Gao and Z. L. Wang, Magnetic properties of Fe nanoparticles trapped at the tips of the aligned carbon nanotubes, 2001 Search PubMed.
- V. Patsula, M. Moskvin, S. Dutz and D. Horák, Size-dependent magnetic properties of iron oxide nanoparticles, Phys. Chem. Solids, 2016, 88, 24–30 CrossRef CAS.
- T. Ibusuki, S. Kojima, O. Kitakami and Y. Shimada, Magnetic anisotropy and behaviors of Fe nanoparticles, IEEE Trans. Magn., 2001, 37(4), 2223–2225 CrossRef CAS.
- C. Granata, R. Russo, E. Esposito, A. Vettoliere, M. Russo and A. Musinu, et al., Magnetic properties of iron oxide nanoparticles investigated by nanosquids, Eur. Phys. J. B, 2013, 86(6) DOI:10.1140/epjb/e2013-40051-2 , available from: http://link.springer.com/10.1140/epjb/e2013-40051-2.
- L. Parimala and J. Santhanalakshmi, Studies on the Iron Nanoparticles Catalyzed Reduction of Substituted Aromatic Ketones to Alcohols, J. Nanopart., 2014, 2014, 1–10 CrossRef.
- A. Desai, D. Mahajan and M. Rafailovich, Synthesis and characterization of nano-sized iron particles on a polystyrene support as potential Fischer–Tropsch catalyst, Prepr. Pap. - Am. Chem. Soc., Div. Fuel Chem., 2003, 48(2), 783 CAS.
- H. M. Torres Galvis, J. H. Bitter, C. B. Khare, M. Ruitenbeek, A. I. Dugulan and K. P. de Jong, Supported Iron Nanoparticles as Catalysts for Sustainable Production of Lower Olefins, Science, 2012, 335(6070), 835–838 CrossRef CAS PubMed.
- A. N. Pour, M. R. Housaindokht, S. F. Tayyari and J. Zarkesh, Fischer–Tropsch synthesis by nano-structured iron catalyst, J. Nat. Gas Chem., 2010, 19(3), 284–292 CrossRef CAS.
- Y. Li, J. Liu, Y. Wang and Z. L. Wang, Preparation of Monodispersed Fe–Mo Nanoparticles as the Catalyst for CVD Synthesis of Carbon Nanotubes, Chem. Mater., 2001, 13(3), 1008–1014 CrossRef CAS.
- H. C. Choi, S. Kundaria, D. Wang, A. Javey, Q. Wang and M. Rolandi, et al., Efficient Formation of Iron Nanoparticle Catalysts on Silicon Oxide by Hydroxylamine for Carbon Nanotube Synthesis and Electronics, Nano Lett., 2003, 3(2), 157–161 CrossRef CAS.
- L. M. Hoyos-Palacio, A. G. García, J. F. Pérez-Robles, J. González and H. V. Martínez-Tejada, Catalytic effect of Fe, Ni, Co and Mo on the CNTs production, IOP Conf. Ser.: Mater. Sci. Eng., 2014, 59, 12005 CrossRef.
- H. Wu, J.-J. Yin, W. G. Wamer, M. Zeng and Y. M. Lo, Reactive oxygen species-related activities of nano-iron metal and nano-iron oxides, J. Food Drug Anal., 2014, 22(1), 86–94 CrossRef CAS PubMed.
- L. B. Hoch, E. J. Mack, B. W. Hydutsky, J. M. Hershman, J. M. Skluzacek and T. E. Mallouk, Carbothermal Synthesis of Carbon-supported Nanoscale Zero-valent Iron Particles for the Remediation of Hexavalent Chromium, Environ. Sci. Technol., 2008, 42(7), 2600–2605 CrossRef CAS PubMed.
- L. Chekli, Development of methods for the characterisation of engineered nanoparticles used for soil and groundwater remediation, Dissertation, School of Civil and Environmental Engineering Faculty of Engineering and Information Technology University of Technology, Sydney (UTS), New South Wales, Australia, 2015.
- S. Li, W. Yan and W. Zhang, Solvent-free production of nanoscale zero-valent iron (nZVI) with precision milling, Green Chem., 2009, 11(10), 1618 RSC.
- N. R. Tao, M. L. Sui, J. Lu and K. Lua, Surface nanocrystallization of iron induced by ultrasonic shot peening, Nanostruct. Mater., 1999, 11(4), 433–440 CrossRef CAS.
- Y.-P. Sun, X.-Q. Li, W.-X. Zhang and H. P. Wang, A method for the preparation of stable dispersion of zero-valent iron nanoparticles, Colloids Surf., A, 2007, 308(1–3), 60–66 CrossRef CAS.
- M. R. Jamei, M. R. Khosravi and B. Anvaripour, A novel ultrasound assisted method in synthesis of nZVI particles, Ultrason. Sonochem., 2013, 21(1), 226–233 CrossRef PubMed.
- W. Glasgow, B. Fellows, B. Qi, T. Darroudi, C. Kitchens and L. Ye, et al., Continuous synthesis of iron oxide (Fe3O4) nanoparticles via thermal decomposition, Particuology, 2016, 26, 47–53 CrossRef CAS.
- M. Stefaniuk, P. Oleszczuk and Y. S. Ok, Review on nano zerovalent iron (nZVI): From synthesis to environmental applications, Chem. Eng. J., 2016, 287, 618–632 CrossRef CAS.
- S. Laurent, D. Forge, M. Port, A. Roch, C. Robic and L. Vander Elst, et al., Magnetic Iron Oxide Nanoparticles: Synthesis, Stabilization, Vectorization, Physicochemical Characterizations, and Biological Applications, Chem. Rev., 2010, 110(4), 2574 CrossRef CAS.
- S.-S. Chen, H.-D. Hsu and C.-W. Li, A new method to produce nanoscale iron for nitrate removal, J. Nanopart. Res., 2004, 6(6), 639–647 CrossRef CAS.
- G. E. Hoag, J. B. Collins, J. L. Holcomb, J. R. Hoag, M. N. Nadagouda and R. S. Varma, Degradation of bromothymol blue by “greener” nano-scale zero-valent iron synthesized using tea polyphenols, J. Mater. Chem., 2009, 19(45), 8671 RSC.
- P. Haddad, A. B. Seabra and N. Duran, Biogenic synthesis of nanostructured iron compounds: applications and perspectives, IET Nanobiotechnol., 2013, 7(3), 90–99 CrossRef PubMed.
- N. Durán and A. B. Seabra, Metallic oxide nanoparticles: state of the art in biogenic syntheses and their mechanisms, Appl. Microbiol. Biotechnol., 2012, 95(2), 275–288 CrossRef PubMed.
- D. Kupka, M. Lovás and V. Šepelák, Deferrization of Kaolinic Sand by Iron Oxidizing and Iron Reducing Bacteria, Adv. Mater. Res., 2007, 20–21, 130–133 CrossRef CAS.
- F. He, D. Zhao and C. Paul, Field assessment of carboxymethyl cellulose stabilized iron nanoparticles for in situ destruction of chlorinated solvents in source zones, Water Res., 2010, 44(7), 2360–2370 CrossRef CAS PubMed.
- J. C. Quinn, C. Geiger, K. Clausen Brooks, C. Coon and W.-S. Yoon, et al., Field Demonstration of DNAPL Dehalogenation using Emulsified Zero-Valent Iron [Internet], 2005 [cited 2016 Apr 9], available from: http://www.geosyntec.com/contaminated-sites-publications/item/2976-field-demonstration-of-dnapl-dehalogenation-using-emulsified-zero-valent-iron Search PubMed.
- J. T. Nurmi, P. G. Tratnyek, V. Sarathy, D. R. Baer, J. E. Amonette and K. Pecher, et al., Characterization and Properties of Metallic Iron Nanoparticles: Spectroscopy, Electrochemistry, and Kinetics, Environ. Sci. Technol., 2005, 39(5), 1221–1230 CrossRef CAS PubMed.
- L. Wu, M. Shamsuzzoha and S. M. C. Ritchie, Preparation of Cellulose Acetate Supported Zero-Valent Iron Nanoparticles for the Dechlorination of Trichloroethylene in Water, J. Nanopart. Res., 2005, 7(4–5), 469–476 CrossRef CAS.
- W. Zhang, Nanoscale iron particles for environmental remediation: an overview, J. Nanopart. Res., 2003, 5(3–4), 323–332 CrossRef CAS.
- M. Elimelech, J. Gregory, X. Jia, R. A. Williams, J. Gregory and X. Jia, et al., Particle Deposition and Aggregation: Measurement, Modelling, and Simulation, in Particle Deposition & Aggregation [Internet], Butterworth-Heinemann, Woburn, 1995, [cited 2016 Mar 16]. pp. 157–202, available from: http://www.sciencedirect.com/science/article/pii/B9780750670241500066 Search PubMed.
- A. N. Bezbaruah, S. Krajangpan, B. J. Chisholm, E. Khan and J. J. Elorza Bermudez, Entrapment of iron nanoparticles in calcium alginate beads for groundwater remediation applications, J. Hazard. Mater., 2009, 166(2–3), 1339–1343 CrossRef CAS PubMed.
- J.-E. Otterstedt and D. A. Brandreth, Small Particles Technology [Internet], Springer US, Boston, MA, 1998 [cited 2016 Apr 7], available from, http://link.springer.com/10.1007/978-1-4757-6523-6 Search PubMed.
- A. Tiraferri, K. L. Chen, R. Sethi and M. Elimelech, Reduced aggregation and sedimentation of zero-valent iron nanoparticles in the presence of guar gum, J. Colloid Interface Sci., 2008, 324(1–2), 71–79 CrossRef CAS PubMed.
- W. Wu, Q. He and C. Jiang, Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies, Nanoscale Res. Lett., 2008 Nov, 3(11), 397–415 Search PubMed.
- E. Guibal, Heterogeneous catalysis on chitosan-based materials: a review, Prog. Polym. Sci., 2005, 30(1), 71–109 CrossRef CAS.
- H. Kim, H.-J. Hong, Y.-J. Lee, H.-J. Shin and J.-W. Yang, Degradation of trichloroethylene by zero-valent iron immobilized in cationic exchange membrane, Desalination, 2008, 223(1–3), 212–220 CrossRef CAS.
- Z. P. Chen, Y. Zhang, S. Zhang, J. G. Xia, J. W. Liu and K. Xu, et al., Preparation and characterization of water-soluble monodisperse magnetic iron oxide nanoparticles via surface double-exchange with DMSA, Colloids Surf., A, 2008, 316(1–3), 210–216 CrossRef CAS.
- Y.-H. Lin, H.-H. Tseng, M.-Y. Wey and M.-D. Lin, Characteristics of two types of stabilized nano zero-valent iron and transport in porous media, Sci. Total Environ., 2010, 408(10), 2260–2267 CrossRef CAS PubMed.
- S. W. Ali, M. L. Mirza, T. M. Bhatti, K. Naeem and M. Imran Din, Dispersion of Iron Nanoparticles by Polymer-Based Hybrid Material for Reduction of Hexavalent Chromium, J. Nanomater., 2015, 2015, 1–8 Search PubMed.
- S. Comba and R. Sethi, Stabilization of highly concentrated suspensions of iron nanoparticles using shear-thinning gels of xanthan gum, Water Res., 2009, 43(15), 3717–3726 CrossRef CAS PubMed.
- H. Chen, H. Luo, Y. Lan, T. Dong, B. Hu and Y. Wang, Removal of tetracycline from aqueous solutions using polyvinylpyrrolidone (PVP-K30) modified nanoscale zero valent iron, J. Hazard. Mater., 2011 DOI:10.1016/j.jhazmat.2011.04.089 , available from: http://linkinghub.elsevier.com/retrieve/pii/S0304389411005279.
- Z. Jiang, L. Lv, W. Zhang, Q. Du, B. Pan and L. Yang, et al., Nitrate reduction using nanosized zero-valent iron supported by polystyrene resins: Role of surface functional groups, Water Res., 2011 Mar, 45(6), 2191–2198 Search PubMed.
- S. M. Ponder, J. G. Darab and T. E. Mallouk, Remediation of Cr(VI) and Pb(II) Aqueous Solutions Using Supported, Nanoscale Zero-valent Iron, Environ. Sci. Technol., 2000, 34(12), 2564–2569 CrossRef CAS.
- Q. Wang, H. Qian, Y. Yang, Z. Zhang, C. Naman and X. Xu, Reduction of hexavalent chromium by carboxymethyl cellulose-stabilized zero-valent iron nanoparticles, J. Contam. Hydrol., 2010, 114(1–4), 35–42 CrossRef CAS PubMed.
- K. Boruvkova and J. Wiener, Water absorption in carboxymethyl cellulose, Autex Res. J., 2011, 11(4) Search PubMed , available from: http://www.autexrj.com/cms/zalaczone_pliki/4_0019_11.pdf.
- K. Bhattacharyya and S. Gupta, Adsorption of a few heavy metals on natural and modified
kaolinite and montmorillonite: A review, Adv. Colloid Interface Sci., 2008, 140(2), 114–131 CrossRef CAS PubMed.
- A. Aras, M. Albayrak, M. Arikan and K. Sobolev, Evaluation of Selected Kaolin Clays as a Raw Material for the Turkish Cement and Concrete Industry, 2007 Search PubMed.
- X. Zhang, S. Lin, Z. Chen, M. Megharaj and R. Naidu, Kaolinite-supported nanoscale zero-valent iron for removal of Pb2+ from aqueous solution: Reactivity, characterization and mechanism, Water Res., 2011, 45(11), 3481–3488 CrossRef CAS PubMed.
- C. Uzum, T. Shahwan, A. Eroglu, K. Hallam, T. Scott and I. Lieberwirth, Synthesis and characterization of kaolinite-supported zero-valent iron nanoparticles and their application for the removal of aqueous Cu2+ and Co2+ ions, Appl. Clay Sci., 2009, 43(2), 172–181 CrossRef CAS.
- S. Li, P. Wu, H. Li, N. Zhu, P. Li and J. Wu, et al., Synthesis and characterization of organo-montmorillonite supported iron nanoparticles, Appl. Clay Sci., 2010, 50(3), 330–336 CrossRef CAS.
- L. Shi, X. Zhang and Z. Chen, Removal of Chromium(VI) from wastewater using bentonite-supported nanoscale zero-valent iron, Water Res., 2011, 45(2), 886–892 CrossRef CAS PubMed.
- S. A. Kim, S. Kamala-Kannan, K.-J. Lee, Y.-J. Park, P. J. Shea and W.-H. Lee, et al., Removal of Pb(II) from aqueous solution by a zeolite–nanoscale zero-valent iron composite, Chem. Eng. J., 2013, 217, 54–60 CrossRef CAS.
- W. Wang, M. Zhou, Q. Mao, J. Yue and X. Wang, Novel NaY zeolite-supported nanoscale zero-valent iron as an efficient heterogeneous Fenton catalyst, Catal. Commun., 2010, 11(11), 937–941 CrossRef CAS.
- N. Mu, D. Bi, R. Fu, X. Guo and Z. Xu, Sepiolite-supported nanoscale zerovalent iron to remediate decabromodiphenyl ether contaminated soil, 2015 [cited 2016 Mar 23], available from: http://www.atlantis-press.com/php/download_paper.php?id=22859 Search PubMed.
- H. K. How and W. Z. W. Yaacob, Synthesis and Characterization of Marine Clay-Supported Nano Zero Valent Iron, Am. J. Environ. Sci., 2015, 11(2), 115–124 CrossRef.
- V. Alipour, L. Rezaei, S. Nasseri, R. Nabizadeh and A. Mahvi. Application of coral supported Iron nano zero valent scale for removal of Natural Organic matter from aqueous solutions, in Proceedings of the 14th International Conference on Environmental Science and Technology Rhodes, CEST2015_01394, Greece, September 2015, pp. 3–5 Search PubMed.
- L. Han, S. Xue, S. Zhao, J. Yan, L. Qian and M. Chen, Biochar Supported Nanoscale Iron Particles for the Efficient Removal of Methyl Orange Dye in Aqueous Solutions, PLoS One, 2015, 10(7), e0132067 Search PubMed.
- M. Ahmad, S. S. Lee, X. Dou, D. Mohan, J.-K. Sung and J. E. Yang, et al., Effects of pyrolysis temperature on soybean stover- and peanut shell-derived biochar properties and TCE adsorption in water, Bioresour. Technol., 2012, 118, 536–544 CrossRef CAS PubMed.
- Y. Zhou, B. Gao, A. R. Zimmerman, H. Chen, M. Zhang and X. Cao, Biochar-supported zerovalent iron for removal of various contaminants from aqueous solutions, Bioresour. Technol., 2014, 152, 538–542 CrossRef CAS PubMed.
- R. A. Crane and T. Scott, The removal of uranium onto carbon-supported nanoscale zero-valent iron particles, J. Nanopart. Res., 2014, 23(7), 1211–1218 Search PubMed , available from: http://link.springer.com/10.1007/s11051-014-2813-4.
- Y. Li, T. Li and Z. Jin, Stabilization of Fe0 nanoparticles with silica fume for enhanced transport and remediation of hexavalent chromium in water and soil, J. Environ. Sci., 2011, 23(7), 1211–1218 CrossRef CAS.
- J. Guo, R. Wang, W. W. Tjiu, J. Pan and T. Liu, Synthesis of Fe nanoparticles@graphene composites for environmental applications, J. Hazard. Mater., 2012, 225–226, 63–73 CrossRef CAS PubMed.
- X. Lv, J. Xu, G. Jiang and X. Xu, Removal of chromium(VI) from wastewater by nanoscale zero-valent iron particles supported on multiwalled carbon nanotubes, Chemosphere, 2011, 85(7), 1204–1209 CrossRef CAS PubMed.
- T. Phenrat, N. Saleh, K. Sirk, H.-J. Kim, R. D. Tilton and G. V. Lowry, Stabilization of aqueous nanoscale zerovalent iron dispersions by anionic polyelectrolytes: adsorbed anionic polyelectrolyte layer properties and their effect on aggregation and sedimentation, J. Nanopart. Res., 2008, 10(5), 795–814 CrossRef CAS.
- V. Alipour, S. Nasseri, R. Nabizadeh Nodehi, A. H. Mahvi and A. Rashidi, Preparation and application of oyster shell supported zero valent nano scale iron for removal of natural organic matter from aqueous solutions, J. Environ. Health Sci. Eng., 2014, 12(1) DOI:10.1186/s40201-014-0146-y , available from: http://www.ijehse.com/content/12/1/146.
- H. Choi, S. R. Al-Abed, S. Agarwal and D. D. Dionysiou, Synthesis of Reactive Nano-Fe/Pd Bimetallic System-Impregnated Activated Carbon for the Simultaneous Adsorption and Dechlorination of PCBs, Chem. Mater., 2008, 20(11), 3649–3655 CrossRef CAS.
- G. E. Hoag, J. B. Collins, J. L. Holcomb, J. R. Hoag, M. N. Nadagouda and R. S. Varma, Degradation of bromothymol blue by “greener” nano-scale zero-valent iron synthesized using tea polyphenols, J. Mater. Chem., 2009, 19(45), 8671 RSC.
- Y. Hwang, Y.-C. Lee, P. D. Mines, Y. S. Huh and H. R. Andersen, Nanoscale zero-valent iron (nZVI) synthesis in a Mg-aminoclay solution exhibits increased stability and reactivity for reductive decontamination, Appl. Catal., B, 2014, 147, 748–755 CrossRef CAS.
- B. Geng, Z. Jin, T. Li and X. Qi, Preparation of chitosan-stabilized Fe0 nanoparticles for removal of hexavalent chromium in water, Sci. Total Environ., 2009, 407(18), 4994–5000 CrossRef CAS PubMed.
- S. R. Kanel, D. Nepal, B. Manning and H. Choi, Transport of surface-modified iron nanoparticle in porous media and application to arsenic(III) remediation, J. Nanopart. Res., 2007, 9(5), 725–735 CrossRef CAS.
- S. Yao, J. Li and Z. Shi, Phosphate ion removal from aqueous solution using an iron oxide-coated fly ash adsorbent, Adsorpt. Sci. Technol., 2009, 27(6), 603–614 CrossRef CAS.
- F. He and D. Zhao, Preparation and Characterization of a New Class of Starch-Stabilized Bimetallic Nanoparticles for Degradation of Chlorinated Hydrocarbons in Water, Environ. Sci. Technol., 2005, 39(9), 3314–3320 CrossRef CAS PubMed.
- Y. Fu, L. Peng, Q. Zeng, Y. Yang, H. Song and J. Shao, et al., High efficient removal of tetracycline from solution by degradation and flocculation with nanoscale zerovalent iron, Chem. Eng. J., 2015, 270, 631–640 CrossRef CAS.
- B. Schrick, B. W. Hydutsky, J. L. Blough and T. E. Mallouk, Delivery Vehicles for Zerovalent Metal Nanoparticles in Soil and Groundwater, Chem. Mater., 2004, 16(11), 2187–2193 CrossRef CAS.
- S. Shaibu, F. Adekola, H. Adegoke and O. Ayanda, A Comparative Study of the Adsorption of Methylene Blue onto Synthesized Nanoscale Zero-Valent Iron-Bamboo and Manganese-Bamboo Composites, Materials, 2014, 7(6), 4493–4507 CrossRef.
- S. Chatterjee, S.-R. Lim and S. H. Woo, Removal of Reactive Black 5 by zero-valent iron modified with various surfactants, Chem. Eng. J., 2010, 160(1), 27–32 CrossRef CAS.
- T. Phenrat, Y. Liu, R. D. Tilton and G. V. Lowry, Adsorbed Polyelectrolyte Coatings Decrease Fe0 Nanoparticle Reactivity with TCE in Water: Conceptual Model and Mechanisms, Environ. Sci. Technol., 2009, 43(5), 1507–1514 CrossRef CAS PubMed.
- R. Kober, D. Schafer, M. Ebert and A. Dahmke, Coupled in situ reactors using Fe0 and activated carbon for the remediation of complex contaminant mixtures in groundwater, IAHS Publ., 2002, 435–440 Search PubMed.
- N. C. Müeller and B. Nowack, Report on nanotechnology in the technology sector: Environment, European Commission, ObservatoryNano, 2009 [cited 2016 May 12], available from: http://library.eawag-empa.ch/empa_publications_2009_open_access/EMPA20090447.pdf.
- A. A. Keller, S. McFerran, A. Lazareva and S. Suh, Global life cycle releases of engineered nanomaterials, J. Nanopart. Res., 2013, 15(6) DOI:10.1007/s11051-013-1692-4 , available from: http://link.springer.com.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.