Next Article in Journal
A Study on Setting Disaster-Prevention Rainfall by Rainfall Duration in Urban Areas Considering Natural Disaster Damage: Focusing on South Korea
Next Article in Special Issue
The Effect of Submergence and Eutrophication on the Trait’s Performance of Wedelia Trilobata over Its Congener Native Wedelia Chinensis
Previous Article in Journal
Experimental Study of Local Scour around Caissons under Unidirectional and Tidal Currents
Previous Article in Special Issue
Optimization and Analysis of Zeolite Augmented Electrocoagulation Process in the Reduction of High-Strength Ammonia in Saline Landfill Leachate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metallic Iron for Environmental Remediation: Starting an Overdue Progress in Knowledge

1
School of Earth Science and Engineering, Hohai University, Fo Cheng Xi Road 8, Nanjing 211100, China
2
Angewandte Geologie, Universität Göttingen, Goldschmidtstraße 3, D-37077 Göttingen, Germany
3
Hung Vuong University, Nguyen Tat Thanh Str., Viet Tri, Phu Tho 29000, Vietnam
4
Department of Water and Environmental Science and Engineering, Nelson Mandela African Institution of Science and Technology, Arusha P.O. Box 447, Tanzania
5
Faculty of Health Sciences, Campus of Banekane, Université des Montagnes, Bangangté P.O. Box 208, Cameroon
6
Department of Chemistry, Faculty of Sciences, University of Douala, Douala B.P. 24157, Cameroon
7
Faculty of Science, Department of Chemistry, University of Maroua, Maroua BP 46, Cameroon
8
Biosystems and Environmental Engineering Research Group, Department of Soil Science and Agricultural Engineering, University of Zimbabwe, Mt. Pleasant, Harare P.O. Box MP167, Zimbabwe
*
Authors to whom correspondence should be addressed.
Water 2020, 12(3), 641; https://doi.org/10.3390/w12030641
Submission received: 22 January 2020 / Revised: 17 February 2020 / Accepted: 21 February 2020 / Published: 27 February 2020
(This article belongs to the Special Issue Water Quality Engineering and Wastewater Treatment)

Abstract

:
A critical survey of the abundant literature on environmental remediation and water treatment using metallic iron (Fe0) as reactive agent raises two major concerns: (i) the peculiar properties of the used materials are not properly considered and characterized, and, (ii) the literature review in individual publications is very selective, thereby excluding some fundamental principles. Fe0 specimens for water treatment are typically small in size. Before the advent of this technology and its application for environmental remediation, such small Fe0 particles have never been allowed to freely corrode for the long-term spanning several years. As concerning the selective literature review, the root cause is that Fe0 was considered as a (strong) reducing agent under environmental conditions. Subsequent interpretation of research results was mainly directed at supporting this mistaken view. The net result is that, within three decades, the Fe0 research community has developed itself to a sort of modern knowledge system. This communication is a further attempt to bring Fe0 research back to the highway of mainstream corrosion science, where the fundamentals of Fe0 technology are rooted. The inherent errors of selected approaches, currently considered as countermeasures to address the inherent limitations of the Fe0 technology are demonstrated. The misuse of the terms “reactivity”, and “efficiency”, and adsorption kinetics and isotherm models for Fe0 systems is also elucidated. The immense importance of Fe0/H2O systems in solving the long-lasting issue of universal safe drinking water provision and wastewater treatment calls for a science-based system design.

1. Introduction

Metal corrosion is one of the most important problems in industry, transport and agriculture [1,2,3,4]. Understanding metal corrosion comprises its detection (e.g., analytical, visual), its monitoring (e.g., mass loss, H2 evolution) and its long-term characterization under various field conditions [4,5]. The corrosion research aims at determining the durability of metallic structures under operational conditions (e.g., oil and gas pipelines, tanks) and revealing the mechanisms of corrosion process [5,6]. This mechanism can be chemical, electrochemical or mixed [1]. Various tools have been used to characterize the corrosion resistance of different metals under various application conditions [4]. The overall result is the availability of integrated approaches to assess and predict the corrosion processes and thus the longevity of metallic structures (e.g., buried pipes) [3,4]. However, the frequency and sudden nature of metallic pipe failures worldwide indicate the inadequacy of current knowledge related to the longevity of buried metallic pipes.
Metallic iron (Fe0) used as a reactive material in subsurface permeable reactive barriers is comparable to iron pipes with three major differences: (i) corrosion is welcome because it is a rather useful process [7,8,9], (ii) a reactive wall is ideally permanently water saturated, and (iii) the length of used particles (<5 cm) is tiny compared to pipes which are up to 12 m in length. On the one hand, Fe0 specimens used in water treatment comprise steel wool with thickness varying between 25 and 90 μm [10,11]. On the other hand, the length of these particles is comparable to the wall thickness of iron pipes (2–4 mm). There has been no real system analysis for remediation Fe0 materials with the aim to outline the differences making their peculiar characteristics. In addition, traceably deriving the longevity of remediation Fe0 specimens from Fe0 pipes is impossible because of the differences highlighted.
The remediation Fe0 was termed as zero-valent iron (ZVI). This acronym is perhaps the first problem of this still innovative technology. A literature research with “zero-valent iron” as keyword would never reveal the ancient literature on Fe0 for water treatment [12,13,14,15,16]. Indeed, no research group until 2017 has referenced a single article from the ancient use of Fe0 in water treatment [17,18]. In chemistry, metallic elements are characterized by their oxidation state, the one of Fe0 is zero (0). Upon oxidation, Fe0 is transformed to FeII, FeIII or FeIV species. Under environmental conditions, only FeII and FeIII species are stable. The oxidation of Fe0 to FeII is a redox process characterized by an electrode potential whose value is ™0.44 V [1]. According to the first principle of chemical thermodynamics, Fe0 can be oxidized by oxidizing agents from each redox couple having a higher electrode potential (E0 > ™0.44 V). Water (H2O or H+) is a relevant oxidizing agent for Fe0 under environmental conditions. The electrode potential for the redox couple H+/H2 is 0.00 V. Fe0 immersed in (contaminated or polluted) water is corroded to form H/H2 and FeII (and mixed FeII/FeIII) species which are stand-alone reducing agents [19,20,21,22,23,24,25,26]. Clearly, it is not surprising that selected species undergo reductive transformations in an Fe0/H2O system [19,26,27,28,29]. The application of metallic iron in water treatment dates back to the 1890s (Table 1). Suspended matter then left to settle
Table 1 presents a summary of the historical applications of metallic iron in water treatment. The history of Fe0 application in environmental remediation has been the subject of several papers by our research group. Comprehensive research on Fe0 for water treatment revealed that the generation of iron hydroxides and oxides (iron corrosion products or FeCPs) is the root cause of contaminant removal in Fe0/H2O systems [22,23,25,26,53,54]. For example, arsenic [55], carbon tetrachloride [29], chromium [56], fluoride [57,58], hexachloroethane [59] , methylene blue [60], methyl orange [61], Orange II [62], phosphate [63], selenium [64] and zinc [65] are all removed in Fe0/H2O systems despite their differences in charge, redox-reactivity and size.
A critical research review article from 2008 [54] established that adsorption, co-precipitation and size-exclusion were the fundamental mechanisms of contaminant removal in Fe0/H2O systems. Yet to date, 13 years later, researchers are still trying to establish the mechanisms by which aqueous contaminants are removed in the presence of Fe0 [59,63,64,66,67]. Unfortunately, none of them has proven the alternative concept [53,54,68] wrong, and most of them are considering Fe0 as the electron donor for observed transformations (electrochemical reaction) [69,70,71]. Moreover, “contaminant reduction” and “contaminant removal” are mostly randomly interchanged. Therefore, a better argumentation is still needed to convincingly explain the in-depth knowledge on the mechanisms causing contaminant removal in Fe0/H2O systems. Another important application of Fe0 materials entails using Fe0 to induce a pH shift and/or enhance microbial processes involved in methane production in anaerobic digesters [72,73,74,75]. This aspect is not considered herein as it is not focused on contaminant removal as discussed in Section 2.
This paper presents a profound analysis of the Fe0/H2O system and derives the leading causes and factors influencing its efficiency for water treatment. Relevant factors include: (i) the Fe0 specimen including its form and size (intrinsic reactivity), (ii) the water chemistry including the nature of contaminants, the presence of dissolved O2 and the pH value, (iii) the contact time (flow velocity or mixing intensity), and (iv) the Fe0 amount and its proportion in the reactive mixture (thickness of the reactive layer or number of columns). Contaminants are explicitly considered within water chemistry, one of the four main groups of factors influencing iron corrosion. The approaches conventionally adopted to investigate these factors, the major findings, the limitations and the knowledge gaps are presented. It is argued that no progress in knowledge is possible before the research community agrees on the key issue that Fe0 is not a reducing agent (under field conditions).

2. The Fe0/H2O System

2.1. Overview of Fundamental Aspects

There is a transfer of electrons from the Fe0 body (solid state) to the Fe0/H2O interface whenever a piece of a reactive Fe0 specimen is immersed in an aqueous solution (Fe0/H2O system) [1,5,6,76]. This occurs because Fe0 is not stable under environmental conditions or because the redox couple H+/H2 (E0 = 0.00) is higher than that of FeII/Fe0 (E0 = ™0.44) in the electrochemical series [1,22,23,24,25,26]. Equation 1 reveals that the oxidative dissolution of Fe0 by protons (H+) and Equation 1a considers that protons are from water (H2O ⇔ H+ + HO). Fe(OH)2 from Equation 1b tends to polymerize and precipitate but can also be oxidized to even lower soluble Fe(OH)3 by dissolved O2 for example (Equation 2a). Fe(OH)2 and Fe(OH)3 are polymerized and further transformed to various hydroxides and oxides (FeCPs) (Equation 3) [1,21,77]. The different iron corrosion products (FeCPs) depict different adsorptive affinities for dissolved species [77,78,79]. Equation 4 summarizes the process of aqueous iron corrosion.
Fe0 + 2 H+ ⇒ Fe2+ + H2
Fe0 + 2 H2O ⇒ Fe(OH)2 + H2
4 Fe2+ + O2 + 4 H+ ⇒ 4 Fe3+ + 2 H2O
4 Fe(OH)2 + O2 + 2 H2O ⇒ 4 Fe(OH)3
Fe(OH)2, Fe(OH)3 ⇒ FeO, Fe3O4, Fe2O3, FeOOH
Fe0 + H2O + (O2) ⇒H2 + iron hydroxides and oxides
In summary, Equation 4 recalls that immersing a reactive Fe0 in water can be universally used to generate H2, Fe2+ and various FeII, FeIII and FeII/FeIII hydroxides and oxides. Equation 1 demonstrates that Fe0 is a scavenger of humidity (H2O), while Equation 2 demonstrates the O2 scavenging nature of Fe0. These two scavenging characteristics have been exploited in several industrial applications [80,81,82]. For example, Fe0 is used as desiccant in food packaging [81]. In the Fe0 remediation literature, the best illustration for the nature of the Fe0/H2O system as generator of FeCPs is perhaps the excellent research article by Furukawa et al. [83]. These authors used several analytical tools to demonstrate the presence of ferrihydrite, green rust, magnetite and lepidocrocite in an Fe0/H2O system. More importantly, they conclude that an Fe0/H2O system is a temporally and spatially heterogeneous geochemical environment. Concerning the spatial heterogeneity, Furukawa et al. [83] specified that magnetite (Fe3O4) is generated in the vicinity of Fe0, whereas ferrihydrite (Fe(OH)3) precipitates away from the Fe0 surface. This conclusion corroborates ancient findings [84] and recalls that even under oxic external conditions, there is a progressive O2 depletion culminating into anoxic conditions in the vicinity of Fe0. In this context, Stratmann and Müller [85] clearly demonstrated that oxygen is reduced by FeII species within the oxide scale (chemical reaction), while Fe0 is oxidized by water (electrochemical reaction).
This section has recalled that, before Fe0 complete depletion, an Fe0/H2O system is a dynamic and heterogeneous system containing Fe0 and all its corrosion products (Equation 4). Accordingly, even the most accurate measurements and the most precise observations are just a static snap-shots of dynamic processes within the system. The situation is exacerbated by the evidence that, the processes occur over an enormous range of timescales ranging from some few minutes or hours in laboratory investigations to months and years in field applications [86,87,88]. This communication insists on the fact that the frequency of discrepant reports in the scientific literature is rooted on an insufficient system analysis. In this regard, an accurate system analysis constitutes the theory of the system. The theory of the Fe0/H2O system, in turn, is like a guide to constrain the choice of the model. The theory that some contaminants are reduced by electrons from Fe0 (electrochemical mechanism) is flawed as water, even present as humidity or moisture do corrode iron [1]. In fact, even deionized water corrodes iron [89]. Figure 1 depicts the interactions of contaminants with solid phases in a pure adsorbent system and the Fe0-based system.

2.2. Oxide Scale on Fe0 and the Decontamination Process

The universal oxide scale on iron metal (at pH > 4.5) is still regarded by the majority of active researchers on remediation Fe0 as a disturbing factor compromising the electron transfer from the metal body (reactivity loss) [90,91,92]. This view contradicts the evidence that a lag time between the start of experiments and reductive transformation is reported in the literature [93,94]. This time corresponds to the quantitative generation of FeCPs, which act as contaminant scavengers. This means that relevant reducing agents are generated in situ. Another problem of the Fe0 literature is that “contaminant removal” and “contaminant reduction” are randomly interchanged, while no real mass balance of the contaminants has been presented [53,54,95]. On the contrary, contaminants that are not recovered are assumed to be chemically reduced [68,69,70].
A look at the mechanism of oxide scale formation reveals that it cannot be electronically conductive. In fact, the initial scale is very porous and cannot transfer electrons because air and water are not electronically conductive (an aqueous solution can be ionic conductive—electrolyte). In subsequent stages, available pores are filled with nascent FeCPs, but they are never uniform and the oxide scale is a mixture of iron hydroxides and oxides [5]. An oxide scale made up of Fe3O4 alone would have been electronically conductive. However, such an Fe3O4 scale cannot exist under natural conditions (immersed Fe0). All other FeCPs are at best semi-conductors and cannot relay electrons from Fe0 under natural conditions. Clearly, reports justifying the reductive efficiency of Fe0/H2O systems using the semi-conductive nature of FeCPs are mistaken [96]. This assertion encompasses the Fe0/pyrite/H2O system, whose efficiency is mainly justified by the semi-conductive nature of FeS species [97,98,99].
The oxide scale on Fe0 is definitively a diffusion barrier for all dissolved species, including the pollutants. It is also the contaminant scavenger such that electrochemical corrosion of immersed Fe0 induces the generation of contaminant scavengers and other reducing agents. Thus the generation of solid FeCPs is a necessary process which has the (perceived negative) side effect of being expansive. Thus, designing an efficient and sustainable system requires answering the question: how long can FeCPs be generated to satisfactorily treat water while keeping a reasonable hydraulic conductivity (permeability)?

2.3. Chemical Aspects

As discussed earlier (Section 2.1), Fe2+ from Equation 1 is transformed to ferrous hydroxide (Fe(OH)2) and ferric hydroxide (Fe(OH)3) which have a strong tendency to form colloids of particles that normally carry a positive charge [62,100,101]. These minerals are further transformed to other FeII/FeIII minerals (e.g., Fe2O3, FeOOH, green rust) exhibiting different affinities to dissolved species. The nature of oxides in each individual system depends on the intrinsic reactivity of the used Fe0 material and the environmental conditions [5,79]. For example, two different Fe0 specimens corroding under the same environment will not necessarily produce the same iron oxides, because the composition of the oxide scale depends on the relative kinetics of Fe0 dissolution and Fe hydroxide precipitation, which in turn depends on the solution chemistry, including the pH value, dissolved ions and the salinity [5,6,102]. On the other hand, in-situ generated free Fe2+ are adsorbed to the surface of available minerals to form the so-called structural FeII with a reducing power far larger than that of the free Fe2+ (E0 < 0.77 V) and sometimes stronger than Fe0 (E0 < –0.44 V) [103]. The availability of several reducing agents in the Fe0/H2O system, and especially from structural FeII, partly stronger than Fe0 implies that the electrochemical series of metal alone cannot predict the chemistry of the system.

2.4. Physical Aspects

The volumetric expansive nature of iron corrosion is the most important physical phenomenon occurring in Fe0/H2O systems [104]. There is expansion because the parent metal (Fe0) produces in-situ both: (i) H2, occupying a volume about 3100 times larger [26,105], and (ii) each solid oxide and hydroxide is at least twice larger in volume than Fe0 (Voxide > Viron) [106,107]. For example, the specific density of magnetite (Fe3O4) is about one half that of iron (Fe0). It implies that after corrosion a space twice larger than the initial space is occupied [108,109,110]. While it can be assumed that H2 escapes from each open system, no free expansion of oxides in porous systems (e.g., water filters, reactive walls) can be assumed [26]. External or internal free expansion occurs in metallic pipes [111,112] and on the walls of steel canister for radioactive waste repositories [110,113]. On the contrary, free expansion cannot be expected in steel-reinforced concrete structures [108,109]. Table 2 presents a comparison of iron corrosion parameters in Fe0 remediation to that of water pipes and reinforced concrete. Accordingly, considering expansive iron corrosion, which culminates into permeability loss is an essential design parameter for porous Fe0/H2O systems (Fe0 filters). For each Fe0 filter, the temporal production of both H2 and oxide is decisive for the long-term efficiency and the permeability of the system [105,107,114,115].
Another key feature in investigating the remediation Fe0/H2O system is that, at pH > 4.5, the Fe0 surface is permanently covered by an oxide scale. The oxide scale acts both as: (i) conduction barrier for electrons from the metal body, and (ii) physical barrier for dissolved species, including O2 and pollutants of concern [22,23,24,25,26,53,54]. The net result is that Fe0 is oxidized by water (electrochemical reaction—Equation 1), and O2 and dissolved contaminants are reduced by reducing species present in the oxide scale (FeII and FeII/FeIII species, H2) (chemical reaction). Again, any reasoning based on the electrochemical series of elements is a misuse of chemical thermodynamic, as the physics of the system, specifically the expansive nature of iron corrosion and it electronically non-conductive nature are simply ignored.

2.5. Kinetic Aspects

Aqueous corrosion of Fe0 materials under environmental conditions (pH > 4.5) is an electrochemical process involving iron dissolution at the anode and H2 evolution at the cathode (Equation 1). This electrochemical reaction is accompanied by the formation of an oxide scale on Fe0 which is not protective as a rule [6,116,117,118]. In general, oxide scale growth and its protectiveness depend primarily on the precipitation rate of iron hydroxides [5,102]. As the Fe0 surface corrodes under the initial scale, corrosion continuously undermines the scale. Voids are created and are progressively filled up by the ongoing hydroxide precipitation. The relative rate of (i) Fe0 oxidative dissolution and (ii) hydroxide precipitation in the Fe0 vicinity determine the protectiveness of the oxide scale. According to Nesic [5], when the rate of hydroxide precipitation exceeds the rate of Fe0 dissolution, a dense protective oxide scale is formed. Conversely, when Fe0 dissolution undermines the new oxide scale faster than hydroxide precipitation can fill in the voids, a porous and non-protective scale forms.
There are several factors influencing the iron corrosion rate, including the intrinsic reactivity of a material and the solution chemistry. The solution chemistry includes the presence of dissolved O2, contaminants and other (mostly) ubiquitous species. In particular, two different Fe0 specimens may exhibit different degree of protectiveness under the same operational conditions. The most important feature to consider is that the corrosion rate is never linear. Thus, a linear extrapolation of the initial corrosion rates of Fe0 specimens in engineered systems can give an inaccurate estimation of its service life. However, for remediation systems, the non-linear nature of the corrosion kinetics implies a decrease in efficiency. Again, results from pipe corrosion or wall corrosion cannot be transferred to remediation Fe0/H2O systems. However, it is certain that the semi-permeable nature of the oxide scale allows significant and continuous corrosion over long time periods (“rust never rest”). In this context, Roh et al. [119] reported of buried iron pieces from World War I still corroding in the subsurface. Thus, the objective of the Fe0 remediation is to couple this long-term corrosion with efficient contaminant removal [7,88,118,120,121].

2.6. Investigating the Fe0/H2O System

The extent and kinetics of Fe0 oxidative dissolution are the result of the following: (i) the nature of metal (intrinsic reactivity), and (ii) the interactions between Fe0 and the environment in which it is placed [76,122,123,124,125]. Therefore, (i) a change of material, and/or (ii) a change in environment results in changes in the rate and the extent of corrosion. In this study, the influence of the environment is considered at micro-scale, specifically what is happening on the Fe0 surface, in its vicinity or over short distances (within the oxide scale). The oxide/water interface is also considered, but the volume of the solution is not. The aqueous phase is regarded as a reservoir of pollutants and co-solutes, while being the principal Fe0 oxidizing agent [126,127].
Laboratory and pilot-scale investigations are usually conducted in order to obtain reliable information on the interactions of metallic devices with particular operational environments [128,129]. Laboratory tests are designed to simulate some relevant field situations. Laboratory studies are mostly aimed at obtaining data in a more convenient way and in a shorter time [123,124,130]. They also provide mechanistic information for field applications. However, short-term laboratory experiments are always a simplification and this should be borne in mind when interpreting achieved results [76,129,131,132]. Given the diversity of operational parameters that have been proven to influence iron corrosion from individual studies, one can be overwhelmed by their number and the fact that each material is unique in its corrosion behaviour [11,124,125,133]. Therefore, a first attempt toward a systematic investigation of relevant influencing factors goes through the consideration of the electrochemical nature of aqueous iron corrosion.
Fe0 is a good conductor of electricity, and its electrochemical aqueous corrosion depends on the conductive nature of the solution in which it is immersed [1,76]. Understanding the corrosion processes helps in selecting Fe0 materials necessary for designing sustainable systems. Therefore, the best way to design a sustainable system is to consider the fundamentals of iron corrosion at the design stage. This section highlights the knowledge of corrosion principles and the importance of the environment and materials for field design. The four compartments necessary for continuous Fe0 corrosion are: (i) an anodic region on Fe0 where the metal is oxidized and releases Fe2+ (leaving two electrons behind), (ii) an electrolyte to transport released Fe2+ away from the anode, (iii) a cathode where the simultaneous reductive transformation coupled to iron oxidation occurs, and (iv) the Fe0 body transporting electron from the anode to the cathode. In an Fe0/H2O system, the anode and the cathode are different sites on the same Fe0 specimen. In the conventional remediation technology, the size of Fe0 particles is generally small (<5 mm). The four compartments must be electronically connected for the electrochemical process to proceed. This means that if Fe0 is covered by an oxide scale, the scale must be conductive to warrant the transfer of electrons from the metal body to the oxidizing agent within the oxide scale or at the oxide/H2O interface.
A Fe0/H2O system is made up of two interfaces: (i) Fe0/oxide and (ii) oxide/H2O. Because the oxide scale is never electronically conductive and is a diffusion barrier to many species, only water can quantitatively reach the Fe0 surface. The net result is that Fe0 oxidative dissolution is an electrochemical reaction (water is reduced) but all other observed/reported chemical reduction occur within the oxide film or at the oxide/H2O interface [53,54,127]. For the Fe0/H2O remediation system, it means that contaminants are not reduced by electrons from Fe0. This has important implications on the operating principles of Fe0/H2O systems: first, using the electrochemical series to predict the reductive transformation of any species has been a mistake, and second, using the stoichiometry of any electrochemical reaction involving Fe0 is also a mistake. Accordingly, contaminants are ‘just’ dissolved species, capable of modifying: (i) the conductive properties of the electrolyte (H2O), (ii) the ion conductivity of the oxide scale, and (iii) formation and the transformation of the oxide scale.

3. An Overview of the Mistakes of Past Efforts

3.1. Contaminant Removal Mechanisms

The major mistake of the Fe0 remediation literature has been to consider that relevant contaminants are reduced by an electrochemical reaction (electrons from the metal) [134,135,136]. This section summarizes the extent of the confusion using a paper by Kamolpornwijit and Liang [137]. This paper is selected because it considered past efforts in understanding the service life of Fe0 filters and performed long-term experiments (>400 days). Kamolpornwijit and Liang [137] considered porosity loss during nitrate (NO3) removal in an Fe0 barrier. The following reactions were considered:
Fe0 + 2 H2O ⇒ Fe2+ + H2 + OH
4 Fe0 + NO3 + 7 H2O ⇒ 4 Fe2+ + NH4+ + 10 OH
5 Fe0 + 2 NO3 + 6 H2O ⇒ 5 Fe2+ + N2 + 12 OH
10 Fe2+ + 2 NO3 + 6 H2O ⇒10 Fe3+ + N2 + 12 OH
Equation 5 corresponds to Equation 1 and represents aqueous iron corrosion, an electrochemical reaction. Direct electron transfer from Fe0 to NO3 yielding NH4+ (Equation 6) or N2 (Equation 7) are only possible if the oxide scale on Fe0 is electronically conductive. Because an electronically conductive oxide scale does not exist under immersed conditions, Equations 6 and 7 are wrong. In discussing their results, Kamolpornwijit and Liang [137] considered that the H2 volume they measured corresponds to the stoichiometry of Equation 5. Comparing this H2 volume to the N2 volume after Equation 7, they concluded that the contribution of water to the corrosion of iron is minimal. In chemical terms, this is erroneous as Equation 5 shows that 5 moles of Fe0 release only 1 mole of N2. Whether N2 escapes from the system or not, the 5 moles of Fe2+ are further oxidized and or precipitated within the Fe0 filter and filling the initial porosity. In reality, it is 10 moles of Fe0 that are needed to release one mole of N2 (Equation 8).
One major problem of the Fe0 literature has been the discussion of the reaction mechanism without mass balance considerations [31,95,138]. Complete mass balance analysis comprises the one of iron, which is admittedly a difficult task as the system is highly dynamic. However, the analysis made herein for nitrate reduction clearly demonstrates the wrongness of the view that Fe0 is corroded by NO3. On the other hand, the importance of FeCPs in inducing porosity loss is demonstrated. If the discussion of the porosity loss starts with the assumption that one mole of N2 corresponds to 10 moles of corroded and expanded Fe0, a better evaluation of changes in the porosity will be achieved. Kamolpornwijit and Liang [137] is also an excellent illustration on how in-situ generated FeCPs are not properly considered while exotic species like CaCO3 or the flow regime (convection versus laminar) are given key role in pore filling and permeability loss. In essence, if a contaminated water is rich in carbonates a pre-treatment unit for their removal can be used such that decontamination units are HCO3 free. Thus, given the expansive nature of iron corrosion, loss of permeability and porosity even occurs in aqueous systems without CaCO3.
Another important feature from Kamolpornwijit and Liang [137] is that, in case nitrate is reduced to NH4+, regardless of the real mechanisms, NH4+ must be removed from the aqueous phase. This removal occurs by adsorption, co-precipitation and size-exclusion [53,54]. These three mechanisms represent the fundamental mechanisms of contaminant removal in Fe0/H2O systems [26,67]. It is essential to recall that chemical reduction and even chemical precipitation are not relevant contaminant removal mechanisms in the concentration ranges relevant to natural waters [111,139]. In particular, for safe drinking water provision, physical methods (e.g., adsorption, filtration, ion exchange) are always mandatory to cope with the stringent regulations. As an example, water defluoridation by chemical precipitation yields an equilibrium fluoride concentration of about 8 mg L−1 according to the solubility of CaF2 [57,58]. This value is far larger than the maximum permissible contamination level of 1.5 mg L−1 for drinking water. Coming back to the Fe0/H2O system, available results show limited removal extent for fluoride while various removal extents for several contaminants depicting no redox reactivity in the systems has been documented. For this reason, a more rational approach is to consider that all species can be removed regardless of their redox potential, and identify exceptions on a case by case basis. In this regard, recent results by Hildebrant et al. [140] have confirmed that fluoride removal is low, but less reactive materials (EDTA test) comparatively remove more fluoride. This last observation corroborates the view that there is no single Fe0 material for all situations and calls for more systematic investigations to identify appropriate Fe0 materials for specific remediation applications.
The lack of systematic investigations has also affected the selection of Fe0 materials used for pathogen removal [141,142,143,144,145,146]. Using various Fe0 materials and very different experimental conditions, discrepant results have been reported [141,146]. Lu et al. [142] can be regarded as a perfect reflection of the state-of-the-art knowledge. The same authors investigated the mechanism of nitrate reduction and iron cycling by an iron-reducing bacteria strain (strain CC76) and metallic iron. They reported that the strain CC76 was able to utilize Fe2+ (from iron corrosion) as electron donor for the nitrate removal. More importantly, they observed that Fe0 inhibited the growth of strain CC76 in the early stage of the operation. This observation corresponds to CC76 removal by Fe0. However, the authors also reported that after the initial stage, strain CC76 was able “to tolerate” the presence of Fe0, meaning that no or less removal was achieved. This phase corresponds to a decrease in the kinetics of iron corrosion as discussed above and has been described in microbially-influenced corrosion [2]. If we recall the work of Kamolpornwijit and Liang [137] primarily investigating nitrate removal in an abiotic Fe0/H2O system, it then becomes apparently clear that the major source of discrepancy in the literature is the insufficient system analysis. Like with chemical contaminants, all pathogens will be removed in a well-designed system. For a filtration system, the key questions to address are: (i) which Fe0 material, in what amount, and for which contaminant and water? and (ii) with which filtration depth, for which flow velocity and for which operational duration? Another key aspect to consider is the nature of the aggregates (e.g., gravel, MnO2, pumice, sand) to be mixed with Fe0 and the Fe0 proportion in the mixture [107,114]. These aspects are yet to be addressed in the Fe0 remediation literature, but are critical for the design and field application of the technology. Table 3 presents a synthesis of current knowledge including the mistakes, and proposed refinements for future studies on Fe0 remediation.

3.2. Reactivity and Efficiency

In the Fe0 remediation literature the terms “reactivity” and “efficiency” are mostly randomly interchanged. This has introduced confusion in the evaluation of independent results. In an effort to resolve this confusion the notion of “electron efficiency” has been introduced [147,148]. The electron efficiency (in %) is defined as the proportion of electrons from Fe0 oxidation that is used for the target reduction reaction, for example for nitrate reduction by Kamolpornwijit and Liang [137] (Section 3.1). In this approach, electrons used to reduce water (Eq. 1) or dissolved O2 are considered as “excess electrons” or avoidable electron wastage [148]. In other words, the main reaction is devalued to a side effect. Contrary to this still prevailing approach [92], the present work and related ones [149,150,151,152,153] recall that aqueous iron corrosion (Equation 1) is not an unwanted reaction leading to an extra consumption of Fe0, but Equation 1 produces FeII species and H2, which are responsible for the documented chemical reduction and FeCPs which are responsible for contaminant removal and permeability loss [105,106,107].
Each Fe0 material is characterized by its intrinsic reactivity while each system is characterized by its efficiency for water treatment [154,155,156]. According to Miyajima and Noubactep [156], the efficiency is the expression of reactivity in a given system. This means that changing the Fe0 material modifies the efficiency. Conversely, an Fe0 material proven efficient in a system could be useless in another system. In the quest of more efficient systems for water treatment, several Fe0 materials and groups of materials (e.g., bimetallics, iron nails, nano-Fe0, scrap iron, sponge iron, steel wool) have been tested and used for water treatment (Table 1). These efforts have been rendered difficult by a mistaken system analysis [157,158] and the evidence that most comparative works are based on testing materials for the removal of individual contaminants [159,160,161,162]. Lufingo et al. [11] recently discussed approaches that can be considered as universal as they are based on the stoichiometry of Equation 1: (i) H2 evolution [163,164] and (ii) Fe2+ production [11,159]. They concluded that iron dissolution in a dilute solution of 1,10-Phenanthroline (2 mM) (Phen test) is the best available tool to characterize the intrinsic reactivity of Fe0 materials (Section 4).

3.3. Misuse of Adsorption Isotherms and Kinetic Models

Several studies investigating contaminant removal by Fe0 materials have applied isotherm and kinetic models initially developed for materials where contaminant removal occurs via adsorption [131,132,165,166]. A detailed discussion of the various adsorption isotherm and kinetic models, including assumptions and equations are presented in earlier reviews [132]. A comprehensive review of the limitations associated with the applications of such models to Fe0 systems is the subject of another paper, hence is beyond the scope of the current study. Briefly, the extension of such models to Fe0 materials is problematic for the following reasons: first, unlike adsorbents such as activated carbon whose surface area is readily available for adsorption at the beginning of the experiment (time t0 = 0; Figure 1), Fe0 systems are highly dynamic and equilibrium is rarely achieved in such systems (i.e., rust never rests), while the total reactive sites contributing to contaminant removal is not exactly known (Figure 1). Thus, most assumptions for such models are not valid for Fe0 materials. Second, as discussed earlier, contaminant removal in Fe0 systems does not occur solely via adsorption. In view of this, the application of adsorption models to Fe0 systems constitute another fundamental error in Fe0 literature. Yet despite a number of reviews highlight the mistakes, misuse and inconsistencies in the use of adsorption kinetic and isotherm models [53,54,67,68,69,70,71,131,132], the problem persists. Thus there is need to develop models for contaminant removal in Fe0 systems that takes into account the iron corrosion phenomena and the fundamental mechanisms responsible for contaminant removal as highlighted in the current review.
Reference [157] severely questioned the validity of specific rate constants (kSA) in Fe0/H2O systems. These rate constants are based on the stoichiometry of contaminant reduction by electrons from the metal body (Fe0) and are not considering the interactions of contaminants within the oxide scale [167]. As discussed in Section 2.2, the Fe0 surface is not accessible to all contaminants. Even kobs values (which are normalized to the surface area to obtain kSA) currently used in the literature are erroneous as they are rooted on the wrong reaction. However, the most severe mistake has been to use the adsorption capacity for Fe0 which is considered a reducing agent. Moreover, an adsorption capacity is determined in a system where the reactive agent is not completely depleted.

4. Selection and Characterization of Fe0 Materials

As discussed above, the scientific reason for using all Fe0 materials (e.g., granular iron and bimetallics, iron filings, iron nails, iron wire, nano-Fe0, scrap iron, steel wool) in water treatment is the electrode potential of the redox electrode FeII/Fe0: E0 = −0.44 V. Clearly, all reactive Fe0 materials have the same redox potential. The intrinsic reactivity of individual Fe0 specimens depends on a myriad of factors from which some are not readily accessible to the researcher. Relevant influencing factors include: alloying elements, Fe0 form, manufacturing processes, metallography, Fe0 grain size, surface area, and surface oxidation state. Accordingly, each Fe0 material has its own intrinsic reactivity which should be characterized in order to better understand how it is influenced by operational conditions to induce the intended remediation goal.
The long history of using Fe0 for technical chemical applications, including water treatment reveals that there has always been efforts to select appropriate materials for individual applications. For example, the porous “spongy iron” (sponge iron or direct reduced iron) was proven more suitable in filtration systems than dense materials [12,13,168]. Similarly, multi-metallic systems and nanoscale materials were recently developed to address (recalcitrant) contaminants that were less sensitive to treatment with granular materials [169]. For completeness, it should be stated that there is no Fe0 material for all situations such that one should rationally select appropriate materials for each specific application. For example, while treating water in fluidized systems (Anderson Process), dense materials were better than sponge iron [13,168]. Similarly, Hildebrant et al. [140] reported that Fe0 materials of low reactivity according to the EDTA test exhibited a better efficiency for fluoride removal. The question arises how to select the right Fe0 specimen for a given application? This calls for the development of standardized protocols for selection and characterization of Fe0 materials for various applications.
It is unfortunate that material selection has not received the due attention given its central role for the technology. Lufingo et al. [11] recently gave an overview of the available tools for the characterization of the intrinsic reactivity of Fe0 specimens. These authors based their work on an excellent review article by Li et al. [162] and insisted on the quantification of Fe2+ from Equation 1. According to Lufingo et al. [11], quantifying Fe2+ or H2 evolution from Equation 1 are the best tools to characterize the intrinsic reactivity. However, quantifying both Fe2+ and H2 in natural systems suffers from the high reactivity of those primary corrosion products within the system, including their adsorption onto FeCPs and their action as own reducing agents. Clearly, measured amounts of Fe2+ and H2 represent an excess quantity and cannot be strictly used to quantify iron corrosion (at pH values >4.5). These considerations clearly show that the H2 evolution method [163,164] is an approximation, while the EDTA method [170] is disturbed by dissolved O2. All other methods are contaminant-specific, and thus of low value [133,160,161]. Lufingo et al. [11] then proposed iron dissolution in a dilute (2 mM) 1,10-Phenanthroline solution (Phen test) as a facile method free from the inherent shortcomings of all available methods.
In terms of affordability and applicability, the Phen test is currently the best available method to characterize the intrinsic reactivity of Fe0 specimens. The test lasts for less than 36 hours and characterizes the initial kinetics of Fe0 dissolution. It is suggested as a candidate for a standard method and is immediately useful for material selection (screening) and quality control [11]. Each Fe0 specimen is characterized by its kPhen value which reflects its initial dissolution at the pH value of natural waters but without the interaction of the oxide scale. The kPhen values for nine steel wool specimens (Fe0 SW) presented by Lufingo et al. [11] were such that 0.07 ≤kPhen (µg h−1) ≤1.30. This shows a ratio of reactivity of 18.5 for Fe0 SW specimens which are often tested as a uniform class of materials compared to granular materials for example. Hildebrandt et al. [10,140] presented the kEDTA values for 13 reactive Fe0 SW and one granular Fe0 as 3.7 ≤ kEDTA (µg h−1) ≤130.8. The lowest kEDTA value corresponds to granular Fe0, suggesting a reactivity ratio of up to 36 between Fe0 SW and granular Fe0. No data comparing kPhen values for Fe0 SW and granular Fe0 are yet available. However, it is certain that the Phen test provides a confidence Fe0 screening tool and is non-contaminant-specific. Therefore, if a significant body of kPhen data and data for characterized Fe0 specimens for the removal of selected representative contaminants (or contaminant groups) are made available, then site-specific treatability studies would then be required only to fine-tune design criteria for the optimal performance of remediation Fe0/H2O systems.
Our research group has long recognized the need for systematic characterization of Fe0 materials in terms of their intrinsic reactivity and efficiency. The idea followed by this research group as summarized herein is to characterize the Fe0 reactivity and the efficiency of Fe0/H2O systems in a pollutant-independent-manner. The alternative is to agree on probing pollutants while using standard protocols. A reference Fe0 material would also be necessary. The use of EDTA and Phen tests to characterize the intrinsic reactivity has already been discussed. The efficiency of the Fe0/H2O system is characterized using the methylene blue (MB) discoloration method (MB method) [60]. Herein, the low affinity of MB for FeCPs is used to trace the abundance of in-situ generated iron oxides in comparatively long-term experiments [62,101,159,171]. The Phen test provides a reliable guidance in selecting Fe0 from the large catalog of available materials and to control the quality of newly manufactured ones. On the other hand, the MB method provides a reliable guidance for the characterization of the efficiency of Fe0/H2O systems. Selected probing agents enable the discussion of the results, and orange II, methyl orange and reactive red 120 were positively tested in this regard [61,62]. Recent results by Hildebrandt et al. [140] suggest that fluoride could be the next candidate for affordable probing reagent. In fact, it was found that fluoride removal is more efficient by low reactive Fe0 specimens.

5. Future Perspectives and Potential Applications

5.1. Investigating the Fe0/H2O System

The findings on which future research should be rooted must be based on the evidence that the oxide scale on iron is a diffusion barrier and shall never been disturbed in ways that are not reproduced under field situations [22,23,53,54,56,124,126,127,171,172,173,174,175]. This has been the motivation for adopting quiescent bath experiments as a more suitable approach to investigate processes occurring in Fe0 permeable reactive barriers some two decades ago [20]. Then and now, most research groups consider the shaking intensity as a relevant operational parameter to be investigated almost in all instances without a quiescent system as a reference. However, there is no given convincing reason for the chosen mixing intensities tested in these experiments [124,125].
In 2005, Devlin and Allin [173] designed a glass-encased magnet batch reactor (GEM reactor) to investigate the impacts of selected anions on the efficiency of granular Fe0 in removing aqueous contaminants using 4-chloronitrobenzene as a probe molecule. This design aimed to achieve a better comparability between results of different experiments by fixing the stirring method and the stirring rates [169,173]. Using this approach, all experiments should be conducted in the GEM reactor to ensure that the granular iron remains stationary while the solution is stirred. Noubactep et al. [124,127] later demonstrated that, while using a rotary shaker, the shaking intensity should never be larger than 100 rpm. Given that a stirring device (including the GEM reactor) can be difficult to acquire by low-income laboratories, quiescent experiments can be adopted as a rule [171].
Quiescent batch experiments have also been adopted for the determination of the initial corrosion rate of Fe0 in EDTA (kEDTA) and Phen (kPhen). Herein, the experiments are stopped before solution saturation, meaning that Fe0 specimens are characterized under conditions where no oxide scale is available. Contrary to the prevalent approach [152,153,154,155,156,157,158,159,160,161,162,163,164], Fe0 specimens are not characterized for any contaminant removal efficiency [176,177]. Solution saturation corresponds to [Fe] values equivalent to the stoichiometry of iron complexation (e.g.,[Fe] = 112 mg L−1 for 2 mM EDTA). Lufingo et al. [11,120,121] recently compared the EDTA and the Phen test and established the superiority of the Phen test, which is additionally more affordable.
Another key feature of the remediation Fe0/H2O system is related to the oxide scale. The omnipresent oxide scale on Fe0 is positively charged under natural conditions (pH > 5.0) [62,100,101,178]. Thus, the Fe0/H2O system is an ion-selective one and preferentially removes negatively charged species like bacteria [142,146]. One original idea has been to characterize the discoloration of methylene blue (MB) by Fe0/sand/H2O systems (MB method). The MB method is grounded on an historical work by Mitchell et al. [178], who observed that sand adsorbs less MB when it is coated with iron oxide. Thus, mixing the same mass of different Fe0 specimens with the given mass of a sand specimen, and allowing them to equilibrate for the long-time in a MB solution enable the differentiation of the Fe0 reactivity of various materials [62,101,176,177]. As a rule, the most reactive material produces the largest amount of iron hydroxides, which, in turn, in-situ coat sand and thus discolors MB the least. MB is thus not a model contaminant, but an operational tracer [171,177]. To ease the interpretation of results, methylene orange (MO) [61,176] or Orange II [62,101] can be used as anionic dyes since they have molecular sizes similar to that of MB. Phukan et al. [62,101,176] additionally used reactive red 120 which is also an anionic dye, but is much larger in size that MO and Orange II.
One key advantage of the MB method is that its enables a visual observation of preferential flow within a reasonable time scale (some weeks) [62]. In designing an Fe0-based filtration systems, a decrease in hydraulic conductivity (permeability loss) is expected, and an early contaminant breakthrough will be observed due to preferential flow [179,180,181,182,183]. Mineral precipitation is particularly intense in the entrance zone of the filter [181], but the created preferential pathways are extended throughout the whole water column [179,181]. The investigation of the process of creation and extension of preferential pathways has been analytically very challenging [181,182,183]. On the basis of observations using the MB method [176,177], changes generated in the entrance zone can be better followed and considered in modelling efforts. This last aspect is crucial for the development of the technology as it has been convincingly demonstrated that models currently predicting the service life of Fe0 filters are rooted on a wrong premise [114,115].

5.2. Designing the Next Generation Filter

The determination of the amount of a given Fe0 which long-term corrosion kinetics would allow a designed Fe0/aggregate system to satisfactorily treat a given water for a certain time frame (e.g., 12 months) can be regarded as a routine work. This routine engineering application is however complicated by the complexity of the iron corrosion process and mistakes in past research as outlined herein. One major thinking mistake has been to consider that admixing Fe0 with sand (the most used aggregate) would alter the decontamination kinetics [90,183,184]. Thus, admixing with sand has been mainly considered as an economic tool to safe Fe0 costs and the negative effects on the resulting system discussed [185]. It has been recently demonstrated that only hybrid Fe0 systems are sustainable [105,106,107].
Several hybrid systems have been successfully tested for water treatment including Fe0/activated carbon [184], Fe0/Fe3O4 [185], Fe0/MnO2 [186], Fe0/pyrite [99] and Fe0/sand [187]. While inert sand alone has clearly improved the efficiency of Fe0 systems [188], the efficiency of other tested aggregates was attributed to the specific materials. The Fe0/sand is regarded as in-situ sand coating [141]. It is certain that other materials will be similarly covered by FeCPs. Thus, it is questionable whether the systems really operate as described.
Rephrasing Notter [189], the success of an Fe0 filter depends on four main factors: (i) the quantity and quality of the water to be produced (e.g., daily), (ii) the intrinsic reactivity of used Fe0, (iii) the nature of the contaminant(s), and (iv) the availability of Fe0 material (to renew exhausted systems). This key principle remains unchanged, one century later. All is needed are systematic investigations take each Fe0 and each water source as a stand-alone design parameter.
The laboratory procedures used to characterize and evaluate Fe0-based systems in batch and column experiments vary considerably among studies. For example, some studies use agitated batch experiments to investigate contaminant removal by Fe0 [126,127], while others used batch experiments operated in quiescent mode, which is closer to field conditions [61]. This makes direct comparison of results among studies problematic, and could lead to misleading conclusions about the performance of Fe0 materials. Therefore, there is need to develop standardized protocols for the evaluation of Fe0 material in both batch and column experiments. In the case of batch experiments, such protocols should include specifying the particle/grain size, tests for determination of material reactivity (e.g., EDTA, Phen tests), liquid/solid ratio, sampling frequency and duration of experiment. For column experiments, such protocols should specify filter depth, duration of experiment, grain size of filter material and chemical properties of test solution to be used, while accounting for potential interference among solutes. Besides developing dedicated pristine Fe0 materials for the water treatment industry, scope also exists to use Fe0 material generated as wastes from other industries. Similarly, iron oxide-rich spent sludge from Fe0-based water treatment systems can be used as raw materials in other industries such as the production of pigments (e.g., iron oxide red) [190], and even as filter material in the construction industry. This cyclic flow of iron materials between the water treatment industry and other industrial processes could form part of a circular economy. Filter wastes may also be regenerated and recycled to new Fe0 material, and then used in other industry when they contain toxic contaminants such as As or U. The recycling of filter wastes in other industries may require detailed environmental risk assessments, including the evaluation of contaminant leaching and potential ecotoxicological effects.

5.3. Field Applications of Fe0-Based Systems

Fe0-based systems present unprecedented opportunities for wastewater treatment and safe drinking water provision especially in low-income countries, including those in Africa (Table 4) [144,157,191,192,193,194,195,196,197,198,199,200,201]. In fact, the use of Fe0-based systems (e.g., the Bishof Process) for clean water provision has a long history dating back to the 19th century [13,14,202]. The history of Fe0-based drinking water treatment systems is discussed in detail in earlier review papers [17,18,122]. Recently, our research group has proposed the integration of Fe0-based systems in rainwater harvesting systems as a low-cost technology for decentralized drinking water provision, in what is known as the Kilimanjaro Concept [192,193,194].
Fe0-based systems also have potential applications in domestic and industrial wastewater treatment systems (Table 4). For example, the Harza Process [27,203] has been used to remove Se from agricultural drainage water. Rahman et al. [200] and Fronczyk et al. [201] proposed the amendment of treatment media for runoff infiltration trenches/pits with granular Fe0. In addition, Fe0-based systems have been used to treat both domestic [204] and industrial wastewaters [205,206], including acid mine drainage [24,207]. Available studies suggest that Fe0 can be added as filter media in constructed wetlands designed to treat urban stormwater and industrial wastewaters [199]. However, a lot remains to be done to further develop and disseminate Fe0-based technologies for wastewater treatment and decentralized safe water provision in developing countries, which such low-cost technologies are most needed. Considering that the bulk of studies are limited to laboratory scale applications, there is need to optimize the Fe0-based systems and evaluate them under field conditions.

6. Summary and Conclusions

The corrosion of iron in remediation Fe0/H2O systems is an electrochemical process, coupling Fe0 oxidative dissolution to the reduction of water (protons) and to no other available oxidizing agent, including dissolved O2. This is because the universal oxide scale on Fe0 acts as diffusion barrier to dissolved species and a conduction barrier to electrons from the metal body. In other words, water is the sole chemical which can remove electrons from the Fe0 surface. Fe0 oxidation and water reduction must not necessarily occur at the same locality. The spatial separation of oxidative (anodic) and reductive (cathodic) reactions is possible as the metal body allows the free flow of electrons from anodic to cathodic sites. The tendency of Fe0 to give off electrons (Equation 1) is the same for all Fe0-based materials (E0 = –0.44 V). This makes material selection and characterization critical in designing sustainable Fe0/H2O systems.
The need to characterize Fe0 materials in terms of intrinsic reactivity and efficiency is critical for the design and operation Fe0/H2O systems, a subject that has been addressed by our research group. Specifically, EDTA and Phen tests were used to characterize the intrinsic reactivity of Fe0 materials. In this regard, the Phen test is considered an affordable and appropriate method that provides a reliable guidance in selecting Fe0 from the large catalog of available Fe0 materials and to control the quality of newly manufactured ones. The efficiency of the Fe0/H2O system is characterized using the methylene blue (MB) discoloration method, while other probing agents investigated include orange II, methyl orange and reactive red 120.
The most characteristic issue of remediation Fe0 is the small size (<5 mm) of used materials. Assuming uniform corrosion, the corrosion rates for progressive Fe0 oxidation should be normalized to the individual particles. In other words, expression like mmol year −1 should be expressed as mmol year−1 particle−1 or mmol year−1 grain−1. The next important issue will be to consider the non-linear kinetics of the corrosion rate such that the service life of a designed system can be deduced knowing the size of used particles and the long-term corrosion rate. Once this is known, considering the expansive nature of iron corrosion would help to design sustainable systems. The choice of the admixing aggregates (e.g., gravel, MnO2, pumice, sand) and the mixing ratios are to be investigated on a case-by-case basis.
A better understanding of the long-term corrosion of relevant Fe0 materials under site-specific conditions is envisioned to ultimately aid in the design of affordable, applicable and efficient remediation Fe0/H2O systems. Applications of Fe0-based systems include; (i) a large variety of water treatment systems, (ii) household and small-scale water treatment plants, including rainwater harvesting systems for drinking water supply, (iii) decentralized domestic wastewater treatment, (iv) urban stormwater, agricultural and industrial wastewater treatment, and (v) as filter media in constructed wetlands. Addressing the key knowledge gaps highlighted here, and extending Fe0-based systems to other application domains such as wastewater treatment for agriculture are focal research areas in our group, which brings together collaborators from various countries.

Author Contributions

R.H., H.Y., R.T., X.C., M.X., B.K.A., V.C., M.L., N.P.S.-S., A.I.N.-T., N.G.-B., V.R.S.-T., W.G. and C.N. contributed equally to manuscript compilation and revisions. All authors have read and agreed to the published version of the manuscript.

Funding

The work is supported by the Ministry of Education of the People’s Republic of China through the Program “Research on Mechanism of Groundwater Exploitation and Seawater Intrusion in Coastal Areas” (Project Code 20165037412).

Acknowledgments

We acknowledge support by the German Research Foundation and the Open Access Publication Funds of the Göttingen University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Landolt, D. Corrosion and Surface Chemistry of Metals, 1st ed.; EPFL Press: Lausanne, Switzerland, 2007; p. 615. [Google Scholar]
  2. Enning, D.; Garrelfs, J. Corrosion of iron by sulfate-reducing bacteria: New views of an old problem. Appl. Environ. Microbiol. 2014, 80, 1226–1236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wasim, M.; Shoaib, S.; Mubarak, M.N.; Inamuddin Asiri, A.M. Factors influencing corrosion of metal pipes in soils. Environ. Chem. Lett. 2018, 16, 861–879. [Google Scholar] [CrossRef]
  4. Enikeev, M.R.; Potemkin, D.I.; Enikeeva, L.V.; Enikeev, A.R.; Gubaydullin, I.M. Quantification of non-uniform distribution and growth of corrosion products at steel-concrete interface Analysis of corrosion processes kinetics on the surface of metals. Chem. Eng. 2020, 383, 123–131. [Google Scholar]
  5. Nesic, S. Key issues related to modelling of internal corrosion of oil and gas pipelines—A review. Corros. Sci. 2007, 49, 4308–4338. [Google Scholar] [CrossRef]
  6. Lazzari, L. General Aspects of Corrosion, Chapter 9.1; Encyclopedia of Hydrocarbons, Istituto Enciclopedia Italian: Rome, Italy, 2008; Volume V. [Google Scholar]
  7. Wilson, E.K. Zero-Valent metals provide possible solution to groundwater problems. Chem. Eng. News 1995, 73, 19–22. [Google Scholar] [CrossRef]
  8. Fairweather, V. When toxics meet metal. Civ. Eng. 1996, 66, 44–48. [Google Scholar]
  9. Tratneyk, P.G. Putting corrosion to use: Remediating contaminated groundwater with zero-Valent metals. Chem. Ind. 1996, 13, 499–503. [Google Scholar]
  10. Hildebrant, B. Characterizing the reactivity of commercial steel wool for water treatment. Freib. Online Geosci. 2018, 53, 1–80. [Google Scholar] [CrossRef]
  11. Lufingo, M.; Ndé-Tchoupé, A.I.; Hu, R.; Njau, K.N.; Noubactep, C. A novel and facile method to characterize the suitability of metallic iron for water treatment. Water 2019, 11, 2465. [Google Scholar] [CrossRef] [Green Version]
  12. Bischof, G. The Purification of Water: Embracing the Action of Spongy Iron on Impure Water; Bell and Bain: Glasgow, Scotland, 1873; p. 19. [Google Scholar]
  13. Devonshire, E. The purification of water by means of metallic iron. J. Frankl. Inst. 1890, 129, 449–461. [Google Scholar] [CrossRef]
  14. Tucker, W.G. The purification of water by chemical treatment. Science 1892, 20, 34–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Baker, M.N. Sketch of the history of water treatment. J. Am. Water Works Assoc. 1934, 26, 902–938. [Google Scholar] [CrossRef]
  16. van Craenenbroeck, W. Easton & Anderson and the water supply of Antwerp (Belgium). Ind. Archaeol. Rev. 1998, 20, 105–116. [Google Scholar]
  17. Mwakabona, H.T.; Ndé-Tchoupé, A.I.; Njau, K.N.; Noubactep, C.; Wydra, K.D. Metallic iron for safe drinking water provision: Considering a lost knowledge. Water Res. 2017, 117, 127–142. [Google Scholar] [CrossRef]
  18. Noubactep, C. Metallic iron for water treatment: Lost science in the West. Bioenergetics 2017, 6, 149. [Google Scholar] [CrossRef] [Green Version]
  19. Gould, J.P. The kinetics of hexavalent chromium reduction by metallic iron. Water Res. 1982, 16, 871–877. [Google Scholar] [CrossRef]
  20. Noubactep, C. Untersuchungen zur passiven in-situ-Immobilisierung von U(VI) aus Wasser; Akad. Buchh.; TU Bergakademie Freiberg: Freiberg, Germany, 2003; (ISSN: 1433-1284). [Google Scholar]
  21. Chaves, L.H.G. The role of green rust in the environment: A review. Rev. Bras. Eng. Agríc. Ambient. 2005, 9, 284–288. [Google Scholar] [CrossRef] [Green Version]
  22. Gheju, M. Hexavalent chromium reduction with zero-valent iron (ZVI) in aquatic systems. Water Air Soil Pollut. 2011, 222, 103–148. [Google Scholar] [CrossRef]
  23. Ghauch, A. Iron-based metallic systems: An excellent choice for sustainable water treatment. Freib. Online Geosci. 2015, 38, 1–80. [Google Scholar]
  24. Hu, R.; Cui, X.; Gwenzi, W.; Wu, S.; Noubactep, C. Fe0/H2O systems for environmental remediation: The scientific history and future research directions. Water 2018, 10, 1739. [Google Scholar] [CrossRef] [Green Version]
  25. Hu, R.; Noubactep, C. Redirecting research on Fe0 for environmental remediation: The search for synergy. Int. J. Environ. Res. Public Health 2019, 16, 4465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Hu, R.; Gwenzi, G.; Sipowo, R.; Noubactep, C. Water treatment using metallic iron: A tutorial review. Processes 2019, 7, 622. [Google Scholar] [CrossRef] [Green Version]
  27. Anderson, M.A. Fundamental Aspects of Selenium Removal by Harza Process; Rep San Joaquin Valley Drainage Program; US Department of the Interior: Sacramento, CA, USA, 1989. [Google Scholar]
  28. Myneni, S.C.B.; Tokunaga, T.K.; Brown, G.E., Jr. Abiotic selenium redox transformations in the presence of Fe(II,III) oxides. Science 1997, 278, 1106–1109. [Google Scholar] [CrossRef] [Green Version]
  29. Jiao, Y.; Qiu, C.; Huang, L.; Wu, K.; Ma, H.; Chen, S.; Ma, L.; Wu, L. Reductive dechlorination of carbon tetrachloride by zero-valent iron and related iron corrosion. Appl. Catal. B Environ. 2009, 91, 434–440. [Google Scholar] [CrossRef]
  30. Swete, H. The purification of river water by agitation and metallic iron, as conducted with the water of the river severn at Worcester. J. R. Soc. Promot. Health 1892, 13, 309–315. [Google Scholar] [CrossRef]
  31. Guinochet, E. Les Eaux D’alimentation: Épuration, Filtration, Sterilization; J.B. Baillière et Fils: Paris, France, 1894; p. 380. [Google Scholar]
  32. Bischof, G. On putrescent organic matter in potable water. Proc. R. Soc. Lond. 1877, 26, 258–261. [Google Scholar]
  33. Bischof, G. On putrescent organic matter in potable water II. Proc. R. Soc. Lond. 1878, 27, 152–156. [Google Scholar]
  34. Burton, C.E. The Water Supply of Towns and the Construction of Waterworks: A Practical Treatise for the Use of Engineers and Students of Engineering; Crosby Lockwood: London, UK, 1898; p. 486. [Google Scholar] [CrossRef]
  35. Anderson, W. The Antwerp Waterworks. Minutes Proc. Inst. Civ. Eng. 1883, 72, 24–44. [Google Scholar]
  36. Anderson, W. The revolving purifier: An apparatus which effects purification of river water and sewage effluent. In the 1885 Anderson Patent; Revolving Purifier Co: London, UK, 1889. [Google Scholar]
  37. Lauderdale, R.A.; Emmons, A.H. A method for decontaminating small volumes of radioactive water. J. Am. Water Works Assoc. 1951, 43, 327–331. [Google Scholar] [CrossRef]
  38. Harza Engineering Co. Selenium Removal Study; Report to Panoche Drainage District; Harza Engineering Co.: Firebaugh, CA, USA, 1986. [Google Scholar]
  39. Khudenko, B.M. Feasibility evaluation of a novel method for destruction of organics. Water Sci. Technol. 1991, 23, 1873–1881. [Google Scholar] [CrossRef]
  40. James, B.R.; Rabenhorst, M.C.; Frigon, G.A. Phosphorus sorption by peat and sand amended with iron oxides or steel wool. Water Environ. Res. 1992, 64, 699–705. [Google Scholar] [CrossRef]
  41. Erickson, A.J.; Gulliver, J.S.; Weiss, P.T. Enhanced sand filtration for storm water phosphorus removal. J. Environ. Eng. ASCE 2007, 133, 485–497. [Google Scholar] [CrossRef] [Green Version]
  42. Wakatsuki, T.; Esumi, H.; Omura, S. High performance and N, P removable on-site domestic wastewater treatment system by multi-soil-layering method. Water Sci. Technol. 1993, 27, 31–40. [Google Scholar] [CrossRef]
  43. Bojic, A.; Purenovic, M.; Kocic, B.; Perovic, J.; Ursic-Jankovic, J.; Bojic, D. The Inactivation of Escherichia Coli by Microalloyed Aluminium Based Composite; Facta Universitatis; University of Niš: Niš, Serbia, 2001; Volume 2, pp. 115–124. [Google Scholar]
  44. You, Y.; Han, J.; Chiu, P.C.; Jin, Y. Removal and inactivation of waterborne viruses using zerovalent iron. Environ. Sci. Technol. 2005, 39, 9263–9269. [Google Scholar] [CrossRef]
  45. Ngai, T.K.K.; Murcott, S.; Shrestha, R.R.; Dangol, B.; Maharjan, M. Development and dissemination of Kanchan™ Arsenic Filter in rural Nepal. Water Sci. Technol. Water Supply 2006, 6, 137–146. [Google Scholar] [CrossRef]
  46. Ngai, T.K.K.; Shrestha, R.R.; Dangol, B.; Maharjan, M.; Murcott, S.E. Design for sustainable development—Household drinking water filter for arsenic and pathogen treatment in Nepal. J. Environ. Sci. Health A 2007, 42, 1879–1888. [Google Scholar] [CrossRef]
  47. Hussam, A.; Munir, A.K.M. A Simple and Effective Arsenic Filter Based on Composite Iron Matrix: Development and Deployment Studies for Groundwater of Bangladesh. J. Environ. Sci. Health 2007, 42, 1869–1878. [Google Scholar] [CrossRef]
  48. Fu, F.; Dionysiou, D.D.; Liu, H. The use of zero-valent iron for groundwater remediation and wastewater treatment: A review. J. Hazard. Mater. 2014, 267, 194–205. [Google Scholar] [CrossRef]
  49. Colombo, A.; Dragonetti, C.; Magni, M.; Roberto, D. Degradation of toxic halogenated organic compounds by iron-containing mono-, bi-and tri-metallic particles in water. Inorg. Chim. Acta 2015, 431, 48–60. [Google Scholar] [CrossRef]
  50. Noubactep, C.; Schöner, A.; Woafo, P. Metallic iron filters for universal access to safe drinking water. Clean Soil Air Water 2009, 37, 930–937. [Google Scholar] [CrossRef] [Green Version]
  51. Ingram, D.T.; Callahan, M.T.; Ferguson, S.; Hoover, D.G.; Shelton, D.R.; Millner, P.D.; Camp, M.J.; Patel, J.R.; Kniel, K.E.; Sharma, M. Use of zero-valent iron biosand filters to reduce E. coli O157:H12 in irrigation water applied to spinach plants in a field setting. J. Appl. Microbiol. 2012, 112, 551–560. [Google Scholar] [CrossRef] [PubMed]
  52. Marik, C.M.; Anderson-Coughlin, B.; Gartley, S.; Craighead, S.; Kniel, K.E. The efficacy of zero valent iron-sand filtration on the reduction of Escherichia coli and Listeria monocytogenes in surface water for use in irrigation. Environ. Res. 2019, 173, 33–39. [Google Scholar] [CrossRef] [PubMed]
  53. Noubactep, C. Processes of contaminant removal in “Fe0–H2O” systems revisited. The importance of co-precipitation. Open Environ. Sci. 2007, 1, 9–13. [Google Scholar] [CrossRef] [Green Version]
  54. Noubactep, C. A critical review on the mechanism of contaminant removal in Fe0–H2O systems. Environ. Technol. 2008, 29, 909–920. [Google Scholar] [CrossRef] [Green Version]
  55. Neumann, A.; Kaegi, R.; Voegelin, A.; Hussam, A.; Munir, A.K.M.; Hug, S.J. Arsenic removal with composite iron matrix filters in Bangladesh: A field and laboratory study. Environ. Sci. Technol. 2013, 47, 4544–4554. [Google Scholar] [CrossRef]
  56. Gheju, M.; Balcu, I. Sustaining the efficiency of the Fe(0)/H2O system for Cr(VI) removal by MnO2 amendment. Chemosphere 2019, 214, 389–398. [Google Scholar] [CrossRef]
  57. Heimann, S. Testing granular iron for fluoride removal. Freib. Online Geosci. 2018, 52, 1–80. [Google Scholar]
  58. Heimann, S.; Ndé-Tchoupé, A.I.; Hu, R.; Licha, T.; Noubactep, C. Investigating the suitability of Fe0 packed-beds for water defluoridation. Chemosphere 2018, 209, 578–587. [Google Scholar] [CrossRef]
  59. Rodrigues, R.; Betelu, S.; Colombano, S.; Masselot, G.; Tzedakis, T.; Ignatiadis, I. Elucidating the dechlorination mechanism of hexachloroethane by Pd-doped zerovalent iron microparticles in dissolved lactic acid polymers using chromatography and indirect monitoring of iron corrosion. Environ. Sci. Pollut. Res. 2019, 26, 7177–7194. [Google Scholar] [CrossRef]
  60. Miyajima, K. Optimizing the design of metallic iron filters for water treatment. Freib. Online Geosci. 2012, 32, 1–60. [Google Scholar]
  61. Gatcha-Bandjun, N.; Noubactep, C.; Loura, B.B. Mitigation of contamination in effluents by metallic iron: The role of iron corrosion products. Environ. Technol. Innov. 2017, 8, 71–83. [Google Scholar] [CrossRef]
  62. Phukan, M. Characterizing the Fe0/sand system by the extent of dye discoloration. Freib. Online Geosci. 2015, 40, 1–70. [Google Scholar]
  63. Sleiman, N.; Deluchat, V.; Wazne, M.; Mallet, M.; Courtin-Nomade, A.; Kazpard, V.; Baudu, M. Phosphate removal from aqueous solution using ZVI/sand bedreactor: Behavior and mechanism. Water Res. 2016, 99, 56–65. [Google Scholar] [CrossRef] [PubMed]
  64. Liang, L.; Yang, W.; Guan, X.; Li, J.; Xu, Z.; Wu, J.; Huang, Y.; Zhang, X. Kinetics and mechanisms of pH-dependent selenite removal by zero valent iron. Water Res. 2013, 47, 5846–5855. [Google Scholar] [CrossRef] [PubMed]
  65. Kishimoto, N.; Iwano, S.; Narazaki, Y. Mechanistic consideration of zinc ion removal by zero-valent iron. Water Air Soil Pollut. 2011, 221, 183–189. [Google Scholar] [CrossRef]
  66. Yu, J.; Hou, X.; Hu, X.; Yuan, H.; Wang, J.; Chen, C. Efficient degradation of chloramphenicol by zero-valent iron microspheres and new insights in mechanisms. Appl. Catal. B Environ. 2019, 256, 117876. [Google Scholar] [CrossRef]
  67. Noubactep, C. The fundamental mechanism of aqueous contaminant removal by metallic iron. Water SA 2010, 36, 663–670. [Google Scholar] [CrossRef] [Green Version]
  68. Noubactep, C. Flaws in the design of Fe(0)-based filtration systems? Chemosphere 2014, 117, 104–107. [Google Scholar] [CrossRef]
  69. Noubactep, C. Research on metallic iron for environmental remediation: Stopping growing sloppy science. Chemosphere 2016, 153, 528–530. [Google Scholar] [CrossRef]
  70. Noubactep, C.; Makota, S.; Bandyopadhyay, A. Rescuing Fe0 remediation research from its systemic flaws. Res. Rev. Insights 2017. [Google Scholar] [CrossRef] [Green Version]
  71. Noubactep, C. The operating mode of Fe0/H2O systems: Hidden truth or repeated nonsense? Fresenius Environ. Bull. 2019, 28, 8328–8330. [Google Scholar]
  72. Oleszkiewicz, J.A.; Sharma, V.K. Stimulation and inhibition of anaerobic processes by heavy metals—A review. Biol. Wastes 1990, 31, 45–67. [Google Scholar] [CrossRef]
  73. Shen, C.F.; Kosaric, N.; Blaszczyk, R. The effect of selected heavy metals (Ni, Co and Fe) on anaerobic granules and their extracellular polymeric substance (EPS). Water Res. 1993, 27, 25–33. [Google Scholar] [CrossRef]
  74. Zhang, Y.; An, X.; Quan, X. Enhancement of sludge granulation in a zero valence iron packed anaerobic reactor with a hydraulic circulation. Process Biochem. 2011, 46, 471–476. [Google Scholar] [CrossRef]
  75. Liu, Y.; Wang, Q.; Zhang, Y. Zero valent iron significantly enhances methane production from waste activated sludge by improving biochemical methane potential rather than hydrolysis rate. Sci. Rep. 2015, 5, 8263. [Google Scholar] [CrossRef] [Green Version]
  76. Hammonds, P. An Introduction to Corrosion and its Prevention (Chapter 4). Compr. Chem. Kinet. 1989, 28, 233–279. [Google Scholar]
  77. Sikora, E.; Macdonald, D.D. The passivity of iron in the presence of ethylenediaminetetraacetic acid I. General electrochemical behavior. Electrochem. Soc. 2000, 147, 4087–4092. [Google Scholar] [CrossRef]
  78. Antia, D.D.J. Desalination of water using ZVI, Fe0. Water 2015, 7, 3671–3831. [Google Scholar] [CrossRef]
  79. van Genuchten, C.M.; Behrends, T.; Stipp, S.L.S.; Dideriksen, K. Achieving arsenic concentrations of <1 μg/L by Fe(0) electrolysis: The exceptional performance of magnetite. Water Res. 2020, 168, 115–170. [Google Scholar]
  80. Tewari, G.; Jayas, D.S.; Jeremiah, L.E.; Holley, R.A. Absorption kinetics of oxygen scavengers. Int. Food Sci. Technol. 2002, 37, 209–217. [Google Scholar] [CrossRef]
  81. Polyakov, V.A.; Miltz, J. Modeling of the humidity effects on the oxygen absorption by iron-based scavengers. Food Sci. 2010, 75, 91–99. [Google Scholar] [CrossRef] [PubMed]
  82. Xu, B.; Jia, M.; Men, J. Preparation, characterization and deoxygenation performance of modified sponge iron. Arab. J. Sci. Eng. 2014, 39, 31–36. [Google Scholar] [CrossRef]
  83. Furukawa, Y.; Kim, J.-W.; Watkins, J.; Wilkin, R.T. Formation of ferrihydrite and associated iron corrosion products in permeable reactive barriers of zero-valent iron. Environ. Sci. Technol. 2002, 36, 5469–5475. [Google Scholar] [CrossRef] [PubMed]
  84. Wranglen, G. Review article on the influence of sulfide inclusions on the corradability of iron and steel. Corros. Sci. 1969, 9, 585–602. [Google Scholar] [CrossRef]
  85. Stratmann, M.; Müller, J. The mechanism of the oxygen reduction on rust-covered metal substrates. Corros. Sci. 1994, 36, 327–359. [Google Scholar] [CrossRef]
  86. Phillips, D.H.; Van Nooten, T.; Bastiaens, L.; Russell, M.I.; Dickson, K.; Plant, S.; Ahad, J.M.E.; Newton, T.; Elliot, T.; Kalin, R.M. Ten year performance evaluation of a field-scale zero-valent iron permeable reactive barrier installed to remediate trichloroethene contaminated groundwater. Environ. Sci. Technol. 2010, 44, 3861–3869. [Google Scholar] [CrossRef]
  87. Wilkin, R.T.; Acree, S.D.; Ross, R.R.; Puls, R.W.; Lee, T.R.; Woods, L.L. Fifteen-year assessment of a permeable reactive barrier for treatment of chromate and trichloroethylene in groundwater. Sci. Total Environ. 2014, 468–469, 186–194. [Google Scholar] [CrossRef]
  88. Wilkin, R.T.; Lee, T.R.; Sexton, M.R.; Acree, S.D.; Puls, R.W.; Blowes, D.W.; Kalinowski, C.; Tilton, J.M.; Woods, L.L. Geochemical and isotope study of trichloroethene degradation in a zero-valent iron permeable reactive barrier: A twenty-two-year performance evaluation. Environ. Sci. Technol. 2019, 53, 296–306. [Google Scholar] [CrossRef]
  89. Luo, P.; Bailey, E.H.; Mooney, S.J. Quantification of changes in zero valent iron morphology using X-ray computed tomography. J. Environ. Sci. 2013, 25, 2344–2351. [Google Scholar] [CrossRef]
  90. Henderson, A.D.; Demond, A.H. Impact of solids formation and gas production on the permeability of ZVI PRBs. J. Environ. Eng. 2011, 137, 689–696. [Google Scholar] [CrossRef]
  91. Henderson, A.D.; Demond, A.H. Long-term performance of zero-valent iron permeable reactive barriers: A critical review. Environ. Eng. Sci. 2007, 24, 401–423. [Google Scholar] [CrossRef] [Green Version]
  92. Guan, X.; Sun, Y.; Qin, H.; Li, J.; Lo, I.M.C.; He, D.; Dong, H. The limitations of applying zero-valent iron technology in contaminants sequestration and the corresponding countermeasures: The development in zero-valent iron technology in the last two decades (1994–2014). Water Res. 2015, 75, 224–248. [Google Scholar] [CrossRef] [PubMed]
  93. Schreier, C.G.; Reinhard, M. Transformation of chlorinated organic compounds by iron and manganese powders in buffered water and in landfill leachate. Chemosphere 1994, 29, 1743–1753. [Google Scholar] [CrossRef]
  94. Hao, Z.-W.; Xu, X.-H.; Wand, D.-H. Reductive denitrification of nitrate by scrap iron filings. J. Zhejiang Univ. Sci. 2005, 6, 182–186. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, G.; Rho, S.; Jahng, D. Design considerations for groundwater remediation using reduced metals. Chem. Eng. 2004, 21, 621–628. [Google Scholar] [CrossRef]
  96. Yoshino, H.; Kawase, Y. Kinetic modeling and simulation of zero-valent iron wastewater treatment process: Simultaneous reduction of nitrate, hydrogen peroxide, and phosphate in semiconductor acidic wastewater. Ind. Eng. Chem. Res. 2013, 52, 17829–17840. [Google Scholar] [CrossRef]
  97. Du, M.; Zhang, Y.; Hussain, I.; Du, X.; Huang, S.; Wen, W. Effect of pyrite on enhancement of zero-valent iron corrosion for arsenic removal in water: A mechanistic study. Chemosphere 2019, 233, 744–753. [Google Scholar] [CrossRef]
  98. Lü, Y.; Li, J.; Li, Y.; Liang, L.; Dong, H.; Chen, K.; Yao, C.; Li, Z.; Li, J.; Guan, X. The roles of pyrite for enhancing reductive removal of nitrobenzene by zero-valent iron. Appl. Catal. 2019, 242, 9–18. [Google Scholar] [CrossRef]
  99. Chen, K.; Han, L.; Li, J.; Lü, Y.; Yao, C.; Dong, H.; Wang, L.; Li, Y. Pyrite enhanced the reactivity of zero valent iron for reductive removal of dyes. Chem. Technol. Biotechnol. 2020. [Google Scholar] [CrossRef]
  100. Sato, N. Surface oxides affecting metallic corrosion. Corros. Rev. 2001, 19, 253–272. [Google Scholar] [CrossRef]
  101. Phukan, M.; Noubactep, C.; Licha, T. Characterizing the ion-selective nature of Fe0-based filters using three azo dyes in batch systems. J. Environ. Chem. Eng. 2016, 4, 65–72. [Google Scholar] [CrossRef]
  102. Whitman, W.G. Corrosion of iron. Chem. Rev. 1926, 2, 419–435. [Google Scholar] [CrossRef]
  103. White, A.F.; Peterson, M.L. Reduction of aqueous transition metal species on the surfaces of Fe(II)-containing oxides. Geochim. Cosmochim. Acta. 1996, 60, 3799–3814. [Google Scholar] [CrossRef]
  104. Pilling, N.B.; Bedworth, R.E. The oxidation of metals at high temperatures. J. Inst. Metals 1923, 29, 529–591. [Google Scholar]
  105. Noubactep, C. Metallic iron for environmental remediation: Prospects and limitations. In A Handbook of Environmental Toxicology: Human Disorders and Ecotoxicology; D’Mello, J.P.F., Ed.; CAB International: Wallingford, UK, 2019; Chapter 36; pp. 531–544. [Google Scholar]
  106. Caré, S.; Crane, R.; Calabrò, P.S.; Ghauch, A.; Temgoua, E.; Noubactep, C. Modeling the permeability loss of metallic iron water filtration systems. CLEAN—Soil Air Water 2013, 41, 275–282. [Google Scholar] [CrossRef] [Green Version]
  107. Domga, R.; Togue-Kamga, F.; Noubactep, C.; Tchatchueng, J.B. Discussing porosity loss of Fe0 packed water filters at ground level. Chem. Eng. J. 2015, 263, 127–134. [Google Scholar] [CrossRef]
  108. Caré, S.; Nguyen, Q.T.; L’Hostis, V.; Berthaud, Y. Mechanical properties of the rust layer induced by impressed current method in reinforced mortar. Cem. Concr. Res. 2008, 38, 1079–1091. [Google Scholar] [CrossRef]
  109. Zhao, Y.; Ren, H.; Dai, H.; Jin, W. Composition and expansion coefficient of rust based on X-ray diffraction and thermal analysis. Corros. Sci. 2011, 53, 1646–1658. [Google Scholar] [CrossRef]
  110. Pusch, R.; Kasbohm, J.; Knutsson, S.; Yang, T.; Nguyen-Thanh, L. The role of smectite clay barriers for isolating high-level radioactive waste (HLW) in shallow and deep repositories. Procedia Earth Planet. Sci. 2015, 15, 680–687. [Google Scholar] [CrossRef] [Green Version]
  111. Snoeyink, V.L.; Jenkins, D. Water Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 1980; p. 480. [Google Scholar]
  112. Sarin, P.; Snoeyink, V.L.; Bebee, J.; Jim, K.K.; Beckett, M.A.; Kriven, W.M.; Clement, J.A. Iron release from corroded iron pipes in drinking water distribution systems: Effect of dissolved oxygen. Water Res. 2004, 38, 1259–1269. [Google Scholar] [CrossRef]
  113. Grauer, R. Ueber Moegliche Wechselwirkungen Zwischen Verfuellmaterial und Stahlbehälter in Endlager C. Interner Bericht 86-01, NAGRA; Baden, Switzerland, 1986; Available online: https://www.nagra.ch/de (accessed on 27 February 2020).
  114. Moraci, N.; Lelo, D.; Bilardi, S.; Calabrò, P.S. Modelling long-term hydraulic conductivity behaviour of zero valent iron column tests for permeable reactive barrier design. Can. Geotech. J. 2016, 53, 946–961. [Google Scholar] [CrossRef]
  115. Noubactep, C. Predicting the hydraulic conductivity of metallic iron filters: Modeling gone astray. Water 2016, 8, 162. [Google Scholar] [CrossRef] [Green Version]
  116. Dickerson, R.E.; Gray, H.B.; Haight, G.P., Jr. Chemical Principles, 3rd ed.; Benjamin/Cummings Inc.: London, UK, 1979; p. 944. [Google Scholar]
  117. Shi, C.; Wei, J.; Jin, Y.; Kniel, K.E.; Chiu, P.C. Removal of viruses and bacteriophages from drinking water using zero-valent iron. Sep. Purif. Technol. 2012, 84, 72–78. [Google Scholar] [CrossRef]
  118. Noubactep, C. The suitability of metallic iron for environmental remediation. Environ. Prog. Sustain. Energy 2010, 29, 286–291. [Google Scholar] [CrossRef] [Green Version]
  119. Roh, Y.; Lee, S.Y.; Elless, M.P. Characterization of corrosion products in the permeable reactive barriers. Environ. Geol. 2000, 40, 184–194. [Google Scholar] [CrossRef]
  120. Hu, R.; Cui, X.; Xiao, M.; Qiu, P.; Lufingo, M.; Gwenzi, W.; Noubactep, C. Characterizing the suitability of granular Fe0 for the water treatment industry. Processes 2019, 7, 652. [Google Scholar] [CrossRef] [Green Version]
  121. Hu, R.; Ndé-Tchoupé, A.I.; Lufingo, M.; Xiao, M.; Nassi, A.; Noubactep, C.; Njau, K.N. The impact of selected pre-treatment procedures on iron dissolution from metallic iron specimens used in water treatment. Sustainability 2019, 11, 671. [Google Scholar] [CrossRef] [Green Version]
  122. Nanseu-Njiki, C.P.; Gwenzi, W.; Pengou, M.; Rahman, M.A.; Noubactep, C. Fe0/H2O filtration systems for decentralized safe drinking water: Where to from here? Water 2019, 11, 429. [Google Scholar] [CrossRef] [Green Version]
  123. Noubactep, C.; Meinrath, G.; Merkel, J.B. Investigating the mechanism of uranium removal by zerovalent iron materials. Environ. Chem. 2005, 2, 235–242. [Google Scholar] [CrossRef] [Green Version]
  124. Noubactep, C.; Schöner, A.; Meinrath, G. Mechanism of uranium (VI) fixation by elemental iron. J. Hazard. Mater. 2006, 132, 202–212. [Google Scholar] [CrossRef] [Green Version]
  125. Noubactep, C.; Kurth, A.M.; Sauter, M. Evaluation of the effects of shaking intensity on the process of methylene blue discoloration by metallic iron. J. Hazard. Mater. 2009, 169, 1005–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Noubactep, C. Investigating the processes of contaminant removal in Fe0/H2O systems. Korean J. Chem. Eng. 2012, 29, 1050–1056. [Google Scholar] [CrossRef] [Green Version]
  127. Noubactep, C.; Schöner, A.; Sauter, M. Significance of oxide-film in discussing the mechanism of contaminant removal by elemental iron materials. In Photo-Electrochemistry & Photo-Biology for the Sustainability; Union Press: Osaka, Japan, 2012; pp. 97–122. [Google Scholar]
  128. Gillham, R.W.; O’Hannesin, S.F. Enhanced degradation of halogenated aliphatics by zero-valent iron. Groundwater 1994, 32, 958–967. [Google Scholar] [CrossRef]
  129. Richardson, J.P.; Nicklow, J.W. In situ permeable reactive barriers for groundwater. Soil Sedim. Contam. 2002, 11, 241–268. [Google Scholar] [CrossRef]
  130. Wang, T.-H.; Li, M.-H.; Teng, S.-P. Bridging the gap between batch and column experiments: A case study of Cs adsorption on granite. Hazard. Mater. 2009, 161, 409–415. [Google Scholar] [CrossRef] [PubMed]
  131. Tien, C. Remarks on adsorption manuscripts revised and declined: An editorial. Purif. Technol. 2007, 54, 277–278. [Google Scholar] [CrossRef]
  132. Tran, H.N.; You, S.-J.; Hosseini-Bandegharaei, A.; Chao, H.-P. Mistakes and inconsistencies regarding adsorption of contaminants from aqueous solutions: A critical review. Water Res. 2017, 120, 88–116. [Google Scholar] [CrossRef]
  133. Miehr, R.; Tratnyek, P.G.; Bandstra, Z.J.; Scherer, M.M.; Alowitz, J.M.; Bylaska, J.E. Diversity of contaminant reduction reactions by zerovalent iron: Role of the reductate. Environ. Sci. Technol. 2004, 38, 139–147. [Google Scholar] [CrossRef]
  134. Matheson, L.J.; Tratnyek, P.G. Reductive dehalogenation of chlorinated methanes by iron metal. Environ. Sci. Technol. 1994, 28, 2045–2053. [Google Scholar] [CrossRef]
  135. Weber, E.J. Iron-mediated reductive transformations: Investigation of reaction mechanism. Environ. Sci. Technol. 1996, 30, 716–719. [Google Scholar] [CrossRef]
  136. O´Hannesin, S.F.; Gillham, R.W. Long-term performance of an in situ “iron wall” for remediation of VOCs. Groundwater 1998, 36, 164–170. [Google Scholar] [CrossRef]
  137. Kamolpornwijit, W.; Liang, L. Investigation of gas production and entrapment in granular iron medium. Contam. Hydrol. 2006, 82, 338–356. [Google Scholar] [CrossRef] [PubMed]
  138. Lavine, B.K.; Auslander, G.; Ritter, J. Polarographic studies of zero valent iron as a reductant for remediation of nitroaromatics in the environment. Microchem. J. 2001, 70, 69–83. [Google Scholar] [CrossRef]
  139. Howe, K.J.; Hand, D.W.; Crittenden, J.C.; Trussell, R.R.; Tchobanoglous, G. Principles of Water Treatment; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012; p. 647. [Google Scholar]
  140. Hildebrant, B.; Ndé-Tchoupé, A.I.; Lufingo, M.; Licha, T.; Noubactep, C. Steel wool for water treatment: Intrinsic reactivity and defluoridation efficiency. Processes 2020, 8, 265. [Google Scholar] [CrossRef] [Green Version]
  141. Bradley, I.; Straub, A.; Maraccini, P.; Markazi, S.; Nguyen, T.H. Iron oxide amended biosand filters for virus removal. Water Res. 2011, 45, 4501–4510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Lu, J.; Lian, T.; Su, J. Effect of zero-valent iron on biological denitrification in the autotrophic denitrification system. Res. Chem. Intermed. 2018, 44, 6011–6022. [Google Scholar] [CrossRef]
  143. Bradshaw, R.N. Assessment of Zero-Valent Iron Capabilities to Reduce Food-Borne Pathogens Via Filtration and Residual Activities in Irrigation Water; Master of Public Health, University of Maryland: College Park, MD, USA, 2017. [Google Scholar]
  144. Chopyk, J.; Kulkarni, P.; Nasko, D.J.; Bradshaw, R.; Kniel, K.E.; Chiu, P.; Sharma, M.; Sapkota, A.R. Zero-valent iron sand filtration reduces concentrations of virus-like particles and modifies virome community composition in reclaimed water used for agricultural irrigation. BMC Res. Notes 2019, 12, 223. [Google Scholar] [CrossRef]
  145. Kulkarni, P.; Raspanti, G.A.; Bui, A.Q.; Bradshaw, R.N.; Sapkota, A.R. Zerovalent iron-sand filtration can reduce the concentration of multiple antimicrobials in conventionally treated reclaimed water. Environ. Res. 2019, 172, 301–309. [Google Scholar] [CrossRef]
  146. Tepong-Tsindé, R.; Ndé-Tchoupé, A.I.; Noubactep, C.; Nassi, A.; Ruppert, H. Characterizing a newly designed steel-wool-based household filter for safe drinking water provision: Hydraulic conductivity and efficiency for pathogen removal. Processes 2019, 7, 966. [Google Scholar] [CrossRef] [Green Version]
  147. Liu, H.; Wang, Q.; Wang, C.; Li, X. Electron efficiency of zero-valent iron for groundwater remediation and wastewater treatment. Chem. Eng. J. 2013, 215–216, 90–95. [Google Scholar] [CrossRef]
  148. Wu, C.-Q.; Fu, D.-T.; Chen, X. Nitrobenzene degradation by micro-sized iron and electron efficiency evaluation. Chem. Pap. 2014, 68, 1350–1357. [Google Scholar] [CrossRef]
  149. Noubactep, C. Relevant reducing agents in remediation Fe0/H2O systems. Clean—Soil Air Water 2013, 41, 493–502. [Google Scholar] [CrossRef] [Green Version]
  150. Noubactep, C.; Temgoua, E.; Rahman, M.A. Designing iron-amended biosand filters for decentralized safe drinking water provision. CLEAN—Soil Air Water 2012, 40, 798–807. [Google Scholar] [CrossRef] [Green Version]
  151. Makota, S.; Nde-Tchoupe, A.I.; Mwakabona, H.T.; Tepong-Tsindé, R.; Noubactep, C.; Nassi, A.; Njau, K.N. Metallic iron for water treatment: Leaving the valley of confusion. Appl. Water Sci. 2017, 7, 4177–4196. [Google Scholar] [CrossRef] [Green Version]
  152. Naseri, E.; Ndé-Tchoupé, A.I.; Mwakabona, H.T.; Nanseu-Njiki, C.P.; Noubactep, C.; Njau, K.N.; Wydra, K.D. Making Fe0-based filters a universal solution for safe drinking water provision. Sustainability 2017, 9, 1224. [Google Scholar] [CrossRef] [Green Version]
  153. Ebelle, T.C.; Makota, S.; Tepong-Tsindé, R.; Nassi, A.; Noubactep, C. Metallic iron and the dialogue of the deaf. Fresenius Environ. Bull. 2019, 28, 8331–8340. [Google Scholar]
  154. Lacy, W.J. Removal of radioactive material from water byslurrying with powdered metal. J. Am. Water Works Assoc. 1952, 44, 824–828. [Google Scholar] [CrossRef]
  155. Banerji, T.; Chaudhari, S. A cost-effective technology for arsenic removal: Case study of zerovalent iron-based iit bombay arsenic filter in west bengal. In Water and Sanitation in the New Millennium; Nath, K., Sharma, V., Eds.; Springer: New Delhi, India, 2017. [Google Scholar]
  156. Miyajima, K.; Noubactep, C. Characterizing the impact of sand addition on the efficiency of granular iron for water treatment. Chem. Eng. 2015, 262, 891–896. [Google Scholar] [CrossRef]
  157. Noubactep, C. On the validity of specific rate constants (kSA) in Fe0/H2O systems. J. Hazard. Mater. 2009, 164, 835–837. [Google Scholar] [CrossRef] [Green Version]
  158. Noubactep, C. An analysis of the evolution of reactive species in Fe0/H2O systems. J. Hazard. Mater. 2009, 168, 1626–1631. [Google Scholar] [CrossRef] [Green Version]
  159. Btatkeu-K, B.D.; Miyajima, K.; Noubactep, C.; Caré, S. Testing the suitability of metallic iron for environmental remediation: Discoloration of methylene blue in column studies. Chem. Eng. J. 2013, 215–216, 959–968. [Google Scholar]
  160. Kim, H.; Yang, H.; Kim, J. Standardization of the reducing power of zero-valent iron using iodine. J. Environ. Sci. Health A 2014, 49, 514–523. [Google Scholar] [CrossRef] [PubMed]
  161. Li, S.; Ding, Y.; Wang, W.; Lei, H. A facile method for determining the Fe(0) content and reactivity of zero valent iron. Anal. Methods 2016, 8, 1239–1248. [Google Scholar] [CrossRef]
  162. Li, J.; Dou, X.; Qin, H.; Sun, Y.; Yin, D.; Guan, X. Characterization methods of zerovalent iron for water treatment and remediation. Water Res. 2019, 148, 70–85. [Google Scholar] [CrossRef] [PubMed]
  163. Reardon, J.E. Anaerobic corrosion of granular iron: Measurement and interpretation of hydrogen evolution rates. Environ. Sci. Technol. 1995, 29, 2936–2945. [Google Scholar] [CrossRef] [PubMed]
  164. Velimirovic, M.; Carniatoc, L.; Simons, Q.; Schoups, G.; Seuntjens, P.; Bastiaens, L. Corrosion rate estimations of microscale zerovalent iron particles via direct hydrogen production measurements. J. Hazard. Mater. 2014, 270, 18–26. [Google Scholar] [CrossRef]
  165. Allen-King, R.M.; Halket, R.M.; Burris, D.R. Reductive transformation and sorption of cis-and trans-1, 2-dichloroethene in a metallic iron–water system. Environ. Toxicol. Chem. 1997, 16, 424–429. [Google Scholar] [CrossRef]
  166. Mamindy-Pajany, Y.; Hurel, C.; Marmier, N.; Roméo, M. Arsenic (V) adsorption from aqueous solution onto goethite, hematite, magnetite and zero-valent iron: Effects of pH, concentration and reversibility. Desalination 2011, 281, 93–99. [Google Scholar] [CrossRef]
  167. Johnson, T.L.; Scherer, M.M.; Tratnyek, P.G. Kinetics of halogenated organic compound degradation by iron metal. Environ. Sci. Technol. 1996, 30, 2634–2640. [Google Scholar] [CrossRef]
  168. Anderson, W. On the purification of water by agitation with iron and by sand filtration. J. Soc. Arts 1886, 35, 29–38. [Google Scholar] [CrossRef]
  169. Ghauch, A.; Abou Assi, H.; Baydoun, H.; Tuqan, A.M.; Bejjani, A. Fe0-based trimetallic systems for the removal of aqueous diclofenac: Mechanism and kinetics. Chem. Eng. J. 2011, 172, 1033–1044. [Google Scholar] [CrossRef]
  170. Noubactep, C.; Meinrath, G.; Dietrich, P.; Sauter, M.; Merkel, B. Testing the suitability of zerovalent iron materials for reactive Walls. Environ. Chem. 2005, 2, 71–76. [Google Scholar] [CrossRef]
  171. Btatkeu-K, B.D.; Tchatchueng, J.B.; Noubactep, C.; Caré, S. Designing metallic iron based water filters: Light from methylene blue discoloration. J. Environ. Manag. 2016, 166, 567–573. [Google Scholar] [CrossRef] [PubMed]
  172. Gheju, M.; Balcu, I. Removal of chromium from Cr(VI) polluted wastewaters by reduction with scrap iron and subsequent precipitation of resulted cations. J. Hazard. Mater. 2011, 196, 131–138. [Google Scholar] [CrossRef] [PubMed]
  173. Devlin, J.F.; Allin, K.O. Major anion effects on the kinetics and reactivity of granular iron in glass-encased magnet batch reactor experiments. Environ. Sci. Technol. 2005, 39, 1868–1874. [Google Scholar] [CrossRef]
  174. Pierce, E.M.; Wellman, D.M.; Lodge, A.M.; Rodriguez, E.A. Experimental determination of the dissolution kinetics of zero-valent iron in the presence of organic complexants. Environ. Chem. 2007, 4, 260–270. [Google Scholar] [CrossRef]
  175. Noubactep, C.; Licha, T.; Scott, T.B.; Fall, M.; Sauter, M. Exploring the influence of operational parameters on the reactivity of elemental iron materials. J. Hazard. Mater. 2009, 172, 943–951. [Google Scholar] [CrossRef] [Green Version]
  176. Phukan, M.; Noubactep, C.; Licha, T. Characterizing the ion-selective nature of Fe0-based filters using azo dyes. Chem. Eng. J. 2015, 259, 481–491. [Google Scholar] [CrossRef]
  177. Tepong-Tsindé, R.; Phukan, M.; Nassi, A.; Noubactep, C.; Ruppert, H. Validating the efficiency of the MB discoloration method for the characterization of Fe0/H2O systems using accelerated corrosion by chloride ions. Chem. Eng. J. 2015, 279, 353–362. [Google Scholar] [CrossRef]
  178. Mitchell, G.; Poole, P.; Segrove, H.D. Adsorption of methylene blue by high-silica sands. Nature 1955, 176, 1025–1026. [Google Scholar] [CrossRef]
  179. Morrison, S.J.; Mushovic, P.S.; Niesen, P.L. Early Breakthrough of Molybdenum and Uranium in a Permeable Reactive Barrier. Environ. Sci. Technol. 2006, 40, 2018–2024. [Google Scholar] [CrossRef] [PubMed]
  180. Li, L.; Benson, C.H.; Lawson, E.M. Impact of mineral fouling on hydraulic behavior of permeable reactive barriers. Groundwater 2005, 43, 582–596. [Google Scholar] [CrossRef] [PubMed]
  181. Lai, K.C.K.; Lo, I.M.C.; Birkelund, V.; Kjeldsen, P. Field monitoring of a permeable reactive barrier for removal of chlorinated organics. J. Environ. Eng. 2006, 132, 199–210. [Google Scholar] [CrossRef]
  182. McGeough, K.L.; Kalin, R.M.; Myles, P. Carbon disulfide removal by zero valent iron. Environ. Sci. Technol. 2007, 41, 4607–4612. [Google Scholar] [CrossRef]
  183. Bi, E.; Devlin, J.F.; Huang, B. Effects of mixing granular iron with sand on the kinetics of trichloroethylene reduction. Groundw. Monit. Remediat. 2009, 29, 56–62. [Google Scholar] [CrossRef]
  184. Tang, H.; Zhu, D.; Li, T.; Kong, H.; Chen, W. Reductive dechlorination of activated carbon-adsorbed trichloroethylene by zero-valent iron: Carbon as electron shuttle. Environ. Qual. 2011, 40, 1878–1885. [Google Scholar] [CrossRef]
  185. Huang, Y.H.; Tang, C.L.; Zeng, H. Removing molybdate from water using a hybridized zero-valent iron/magnetite/Fe(II) treatment system. Chem. Eng. J. 2012, 200, 205–263. [Google Scholar] [CrossRef]
  186. Btatkeu-K, B.D.; Olvera-Vargas, H.; Tchatchueng, J.B.; Noubactep, C.; Caré, S. Characterizing the impact of MnO2 on the efficiency of Fe0-based filtration systems. Chem. Eng. J. 2014, 250, 416–422. [Google Scholar] [CrossRef]
  187. Westerhoff, P.; James, J. Nitrate removal in zero-valent iron packed columns. Water Res. 2003, 37, 1818–1830. [Google Scholar] [CrossRef]
  188. Song, D.-I.; Kim, Y.H.; Shin, W.S. A simple mathematical analysis on the effect of sand in Cr(VI) reduction using zero valent iron. Korean J. Chem. Eng. 2005, 22, 67–69. [Google Scholar] [CrossRef]
  189. Notter, J.L. The purification of water by filtration. Br. Med. J. 1878, 12, 556–557. [Google Scholar] [CrossRef] [Green Version]
  190. Chen, Z.; Wang, X.; Ge, Q.; Guo, G. Iron oxide red wastewater treatment and recycling of iron-containing sludge. J. Clean. Prod. 2015, 87, 558–566. [Google Scholar] [CrossRef]
  191. Chaudhari, S.; Banerji, T.; Kumar, P.R. Domestic- and community-scale arsenic removal technologies suitable for developing countries. Water Reclam. Sustain. 2014, 2014, 155–182. [Google Scholar]
  192. Marwa, J.; Lufingo, M.; Noubactep, C.; Machunda, R. Defeating fluorosis in the East African Rift Valley: Transforming the Kilimanjaro into a rainwater harvesting park. Sustainability 2018, 10, 4194. [Google Scholar] [CrossRef] [Green Version]
  193. Ndé-Tchoupé, A.I.; Tepong-Tsindé, R.; Lufingo, M.; Pembe-Ali, Z.; Lugodisha, I.; Mureth, R.I.; Nkinda, M.; Marwa, J.; Gwenzi, W.; Mwamila, T.B.; et al. White teeth and healthy skeletons for all: The path to universal fluoride-free drinking water in Tanzania. Water 2019, 11, 131. [Google Scholar] [CrossRef] [Green Version]
  194. Qi, Q.; Marwa, J.; Mwamila, T.B.; Gwenzi, W.; Noubactep, C. Making Rainwater Harvesting a Key Solution for Water Supply: The Universality of the Kilimanjaro Concept. Sustainability 2019, 11, 5606. [Google Scholar] [CrossRef] [Green Version]
  195. Latrach, L.; Ouazzani, N.; Hejjaj, A.; Mahi, M.; Masunaga, T.; Mandi, L. Two-stage vertical flow multi-soil-layering (MSL) technology for efficient removal of coliforms and human pathogens from domestic wastewater in rural areas under arid climate. Int. J. Hyg. Environ. Health 2018, 22, 64–80. [Google Scholar] [CrossRef]
  196. Allred, B.J.; Racharaks, R. Laboratory comparison of four iron-based filter materials for drainage water phosphate treatment. Water Environ. Res. 2014, 86, 852–862. [Google Scholar] [CrossRef]
  197. Murphy, A.P. Removal of selenate from water by chemical reduction. Ind. Eng. Chem. Res. 1988, 27, 181–191. [Google Scholar] [CrossRef]
  198. Oldright, G.L.; Keyes, H.E.; Miller, V.; Sloan, W.A. Precipitation of Lead and Copper from Solution on Sponge Iron; BuMines B 281; Washington, DC, USA, 1928; 131 p, Available online: https://searchworks.stanford.edu/view/7623313 (accessed on 27 February 2020).
  199. Wu, S.; Vymazal, J.; Brix, H. Critical Review: Biogeochemical Networking of Iron in Constructed Wetlands for Wastewater Treatment. Environ. Sci. Technol. 2019, 53, 7930–7944. [Google Scholar] [CrossRef]
  200. Rahman, M.A.; Wiegand, B.A.; Badruzzaman, A.B.M.; Ptak, T. Hydrogeological analysis of the upper Dupi Tila Aquifer, towards the implementation of a managed aquifer-recharge project bei Dhaka City, Bangladesh. Hydrogeology 2013, 21, 1071–1089. [Google Scholar] [CrossRef]
  201. Fronczyk, J.; Markowska-Lech, K.; Bilgin, A. Treatment assessment of road runoff water in zones filled with ZVI, activated carbon and mineral materials. Sustainability 2020, 12, 873. [Google Scholar] [CrossRef] [Green Version]
  202. Davis, F. An Elementary Handbook on Potable Water; Silver, Burdett & Co: New York, NY, USA; Boston, MA, USA; Chicago, IL, USA, 1891; 118 p. [Google Scholar]
  203. Lee, E.W. Drainage Water Treatment and Disposal. In Management of Water Use in Agriculture. Advanced Series in Agricultural Sciences; Tanji, K.K., Yaron, B., Eds.; Springer: Berlin/Heidelberg, Germany, 1994; Volume 22. [Google Scholar]
  204. Hussam, A. Contending with a development disaster: Sono filters remove arsenic from well water in Bangladesh. Innovations 2009, 4, 89–102. [Google Scholar] [CrossRef]
  205. Bartzas, G.; Komnitsas, K.; Paspaliaris, I. Laboratory evaluation of Fe0 barriers to treat acidic leachates. Miner. Eng. 2006, 19, 505–514. [Google Scholar] [CrossRef]
  206. Tansel, B. New technologies for water and wastewater treatment: A survey of recent patents. Recent Pat. Chem. Eng. 2008, 1, 17–26. [Google Scholar] [CrossRef]
  207. Obiri-Nyarko, F.; Grajales-Mesa, S.J.; Malina, G. An overview of permeable reactive barriers for in situ sustainable groundwater remediation. Chemosphere 2014, 111, 243–259. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram comparing the interactions of contaminants (black points) with solid phases in a pure adsorbent system (left) and the Fe0-based system (right). The red points represent iron corrosion products (FeCPs) which are either coated on solids or suspended in the pore solution.
Figure 1. Schematic diagram comparing the interactions of contaminants (black points) with solid phases in a pure adsorbent system (left) and the Fe0-based system (right). The red points represent iron corrosion products (FeCPs) which are either coated on solids or suspended in the pore solution.
Water 12 00641 g001aWater 12 00641 g001b
Table 1. Historical application of metallic iron in water treatment (modified from Mwakabona et al. [17]).
Table 1. Historical application of metallic iron in water treatment (modified from Mwakabona et al. [17]).
Age/TimeMaterialUsageTarget ImpuritiesApplicationReference
Pre–1850Old iron nailsShaken with impure waterSuspended matter then left to settleUsed in West England[30,31]
1850–1900Iron plates/wires
(1857)
Suspended in flowing impureSuspended matter waterTested but not widely applied[13,31,32]
Magnetic carbide/
Iron oxide
Filter mediaSuspended matter/microbesWidely applied household scale filters[12,13,31]
PolariteFilter mediaSuspended matter/colour/microbesHousehold filters[32,33]
Spongy ironFilter mediaSuspended matter/colour/microbes/chemical contaminantsHousehold and large scale
filters
[30,31,34,35]
Iron fillingsAgitated in the evolving purifierSuspended matter/colour/microbesLarge scale in treatment plants[31,34,35]
1900–1950Iron fillingsAgitated in the revolving purifierSuspended matter/colour/microbes/chemical contaminantsLarge scale in treatment plants[16,31,35,36]
1950–1990Steel woolFilter mediaRadionuclidesWidely tested household scale filters[37]
Iron fillingsFilter mediaSelenium from agricultural drainage waterField scale[38]
Iron fillingsStatic or dynamicHalogenated hydrocarbonsConcept[39]
Post–1990sIron fillings, steel woolFilter mediaPhosphate from agricultural drainage waterField scale [40,41]
Iron fillingsFilter mediaDomestic wastewaterField scale [42]
Al0/Fe0 compositeFilter media or batch systemsPathogen and chemical removalLaboratory demonstration [43]
Iron fillingsFilter mediaPathogen removalLaboratory demonstration[44]
Iron nailsFilter mediaKanchan Arsenic filterField scale[45,46]
Composite iron matrixFilter mediaSONO Arsenic filterField scale[47]
Metallic ironReactive media in PRBsChemical contaminantsField scale[22,23,48,49]
Metallic ironFilter media or batch systemsAll classes of contaminantsConcept[50]
Metallic ironFilter mediaE. coli and Listeria monocytogenes removal from surface water for irrigation Field scale[51,52]
Table 2. Comparison of iron corrosion parameters in Fe0 remediation to that of water pipes and reinforced concrete.
Table 2. Comparison of iron corrosion parameters in Fe0 remediation to that of water pipes and reinforced concrete.
ParameterWater PipingReinforced ConcreteFe0 Remediation
Iron corrosionDestructiveDestructiveConstructive
EnvironmentUnknown soil Known concreteKnown mixture
Fe sizeseveral meterssome cm<5 mm
some cm thicksome mm<2 mm
Rust expansionMinor or no issuemajor issuemajor issue
Corrosion effectPipe damagecrackingpermeability loss
Linear corrosionConservativenon applicableAbsurd
Knowledge status Quite goodQuite goodRather poor
Service life>50 years>50 yearsUp to 30 years
ReferenceEnning et al. [2]Caré et al. [108]Hu et al. [26]
Table 3. Summary of synthesis of current knowledge and proposed refinements for future studies on Fe0 remediation.
Table 3. Summary of synthesis of current knowledge and proposed refinements for future studies on Fe0 remediation.
Previous StudySynthesisRefinement for Future Studies
Electrochemical reductionChemical reduction
Protons as concurrent Fe0 oxidizerProtons as sole oxidizing agents for Fe0
Decontamination mechanismAdsorption as possible mechanismAdsorption as fundamental mechanism
Co-precipitation as possible mechanismCo-precipitation as fundamental mechanism
Size-exclusion rarely consideredSize-exclusion as fundamental mechanism
Driving forceE0 value = -0.44 VAffinity to iron oxides and hydroxides
SelectivityE0 value of the contaminantSurface charge and size
H2 evolutionFe0 dissolution in 1,10-Phenanthroline
Material selectionRemoval efficiency for selected speciesFe0 dissolution in EDTA
Physical characterization toolsNot really useful
Fe0 dissolution in complexing agentsPhen test as candidate for standard method
Batch experimentsShaken or stirredQuiescent or shaken at < 100 rpm
Lasting for some hours to daysLasting for some weeks to months
Mostly lasting for some few monthsLasting for at least six months
Many contain only Fe0Never contain more than 50% (vol/vol) Fe0
Column/field experimentsSome accelerated column experimentsNever artificially accelerated iron corrosion
System modellingBased on the stoichiometry of an electrochemical reduction reactionConsider protons as the sole oxidizers of Fe0Consider porosity loss related to each Fe0 atom
Table 4. Potential field applications of Fe0-based systems in drinking water and wastewater treatment. The given reference refers to the oldest known application.
Table 4. Potential field applications of Fe0-based systems in drinking water and wastewater treatment. The given reference refers to the oldest known application.
Field of applicationRemarksReferences
A: Safe drinking water provision:
1.Centralized safe drinking water systemsBoth filter beds and fluidized beds are used[168]
2. Decentralized water treatment for small communitiesSteel wool is used against radionuclides[37,154]
3. Household filters against arsenicTraditional filters are amended with iron nails [45,46]
4. Household filters against pathogensBiosand filters are amended with Fe0 materials[141]
5. Community-scale Fe0-based systems against arsenicNatural water equilibrates with iron nails and flocs are filtered on gravel [191]
6. Decentralized rainwater harvesting systems for drinking provision Fe0 filters are efficient to remediate expected contaminants[192,193,194]
B: Wastewater treatment systems:
1. Decentralized domestic wastewater treatmentA Fe0 unit in implemented mostly to remove PO43-[42,195]
2. Wastewater for agriculture and aquacultureIron filings are used in filter beds to remove Se[196,197]
3. Industrial wastewaters/effluentsSponge iron is used to precipitate Cu and Pb from industrial wastes [198]
4. Constructed Wetlands for Wastewater TreatmentGranular iron is added to reactive materials[79,199]
5. Urban stormwater treatmentFe0 is amended to other media to optimize the treatment of runoff water[200,201]

Share and Cite

MDPI and ACS Style

Hu, R.; Yang, H.; Tao, R.; Cui, X.; Xiao, M.; Amoah, B.K.; Cao, V.; Lufingo, M.; Soppa-Sangue, N.P.; Ndé-Tchoupé, A.I.; et al. Metallic Iron for Environmental Remediation: Starting an Overdue Progress in Knowledge. Water 2020, 12, 641. https://doi.org/10.3390/w12030641

AMA Style

Hu R, Yang H, Tao R, Cui X, Xiao M, Amoah BK, Cao V, Lufingo M, Soppa-Sangue NP, Ndé-Tchoupé AI, et al. Metallic Iron for Environmental Remediation: Starting an Overdue Progress in Knowledge. Water. 2020; 12(3):641. https://doi.org/10.3390/w12030641

Chicago/Turabian Style

Hu, Rui, Huichen Yang, Ran Tao, Xuesong Cui, Minhui Xiao, Bernard Konadu Amoah, Viet Cao, Mesia Lufingo, Naomi Paloma Soppa-Sangue, Arnaud Igor Ndé-Tchoupé, and et al. 2020. "Metallic Iron for Environmental Remediation: Starting an Overdue Progress in Knowledge" Water 12, no. 3: 641. https://doi.org/10.3390/w12030641

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop