Next Article in Journal
Combining Precision Viticulture Technologies and Economic Indices to Sustainable Water Use Management
Next Article in Special Issue
Visible Light Photocatalyst and Antibacterial Activity of BFO (Bismuth Ferrite) Nanoparticles from Honey
Previous Article in Journal
Regional Evapotranspiration Estimation by the Improved MOD16-sm Model and Its Application in Central China
Previous Article in Special Issue
A Review on the Catalytic Remediation of Dyes by Tailored Carbon Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metallic Iron for Water Remediation: Plenty of Room for Collaboration and Convergence to Advance the Science

1
School of Earth Science and Engineering, Hohai University, Fo Cheng Xi Road 8, Nanjing 211100, China
2
Applied Geology, University of Göttingen, Goldschmidtstraße 3, D-37077 Göttingen, Germany
3
Biosystems and Environmental Engineering Research Group, Department of Agricultural and Biosystems Engineering, University of Zimbabwe, Mount Pleasant, Harare P.O. Box MP167, Zimbabwe
4
Centre for Modern Indian Studies (CeMIS), University of Göttingen, Waldweg 26, D-37073 Göttingen, Germany
5
School of Material Energy Water and Environmental Science (MEWES), Department of Water Environmental Science and Engineering (WESE), The Nelson Mandela African Institution of Science and Technology, Arusha P.O. Box 447, Tanzania
6
Faculty of Science and Technology, Campus of Banekane, Université des Montagnes, Bangangté P.O. Box 208, Cameroon
*
Author to whom correspondence should be addressed.
Submission received: 6 April 2022 / Revised: 28 April 2022 / Accepted: 1 May 2022 / Published: 6 May 2022

Abstract

:
Scientific collaboration among various geographically scattered research groups on the broad topic of “metallic iron (Fe0) for water remediation” has evolved greatly over the past three decades. This collaboration has involved different kinds of research partners, including researchers from the same organization and domestic researchers from non-academic organizations as well as international partners. The present analysis of recent publications by some leading scientists shows that after a decade of frank collaboration in search of ways to improve the efficiency of Fe0/H2O systems, the research community has divided itself into two schools of thought since about 2007. Since then, progress in knowledge has stagnated. The first school maintains that Fe0 is a reducing agent for some relevant contaminants. The second school argues that Fe0 in-situ generates flocculants (iron hydroxides) for contaminant scavenging and reducing species (e.g., FeII, H2, and Fe3O4), but reductive transformation is not a relevant contaminant removal mechanism. The problem encountered in assessing the validity of the views of both schools arises from the quantitative dominance of the supporters of the first school, who mostly ignore the second school in their presentations. The net result is that the various derivations of the original Fe0 remediation technology may be collectively flawed by the same mistake. While recognizing that the whole research community strives for the success of a very promising but unestablished technology, annual review articles are suggested as an ingredient for successful collaboration.

1. Introduction

Permeable reactive barriers of metallic iron (Fe0 PRBs) are reportedly a well-established technology for groundwater remediation [1,2,3,4]. Since the first four peer-reviewed articles on testing Fe0 for groundwater remediation [5,6,7,8], numerous studies have investigated various aspects on the design and installation of efficient Fe0 PRBs, including the reaction mechanisms and the long-term hydraulic properties [9,10,11]. Moreover, alternatives to granular Fe0 materials were developed, including nano-scale Fe0 and sulfidized Fe0 [12,13,14,15,16]. The results have been presented in more than 5000 research and review articles [17,18]. The research to date is reported to have followed a conventional collaborative approach [19,20,21,22,23]. In particular, intensive international research collaborations have been documented, as revealed by two recent bibliometric analyses [17,18]. One major aim of research collaboration is boosting the advancement of knowledge while accounting for the increasing professionalization and specialization of science [22]. A closer look behind the publications on the Fe0/H2O system reveals that, despite increased scientific productivity, there have been no frank concerted efforts for the “advancement of knowledge” for the past 15 years.
The first problem of the Fe0 PRB technology and the post-1990 Fe0 remediation literature is that it was introduced with very little critical literature research [24]. In particular, the old textbook knowledge that granular Fe0 generates flocs for water treatment, just like iron salts [25], has not been considered. In one study, it was claimed that a thorough patent research was made, and it revealed some works from the 1970s [26]. This view was directly contradicted in a recent review article by Antia [27] revealing that patents on Fe0 for water treatment dating back to the 1850s exist, more than 140 years before the advent of Fe0 PRBs. The question then arises of whether or not progress in knowledge can be achieved without a common starting point? Unfortunately, all warnings from investigators suggesting that Fe0 might not be a reducing agent under environmental conditions [28,29,30,31,32,33,34] were overlooked by the majority of active researchers. In particular, Lavine et al. [31] used differential pulse polarography to simultaneously monitor the disappearance of nitrobenzene and the appearance of Fe2+ in the Fe0/H2O system. Their studies were very “informative” but could not deliver any information about ‘’crucial mechanistic details on the chemical processes controlling the reduction of organic compounds” in the Fe0/H2O system. Jiao et al. [34] and many others have questioned the view that electrons from the metal body (Fe0) are of any significance for the reductive transformation of soluble species, including contaminants and O2 [24,35].
To the authors’ knowledge, only a few research groups have designed experiments to cope with the science of the Fe0/H2O system [36,37,38,39,40,41,42,43,44,45]. In particular, Li et al. [39] have alloyed Fe0 to lower the corrosion rate in an effort to have less dissolved iron in the effluent. The three most prominent research groups are the following: (i) Ghauch’s group in Lebanon [46], (ii) Gheju’s group in Romania [47], and (iii) our group in Germany [48]. Interestingly, until 2010/2011, all three research groups reported on the electrochemical nature of contaminant removal in Fe0/H2O systems [49,50,51]. In other words, critics of the reductive transformation model are not enemies of the Fe0 technology, but are proponents for the integrity of science [24,50,51,52,53,54]. The aim of the present overview article is to point out the flaws or limitations of the named concept while demonstrating that even leading scientists or proponents contribute to perpetuating the concept. The objective is not to blame any single colleague, but rather to correct past errors and lead early-career researchers of Fe0/H2O systems on the right path. The presentation starts with an explanation of the methodology used (Section 2), followed by a summary of the fundamental flaws (Section 3). Then, an overview on how leading scientists have been ignoring the state-of-the-art knowledge is presented (Section 4). A short conclusion will close the presentation.

2. Methodology

No systematic approach using search terms and combinations thereof was used to identify published scientific literature related to “Fe0 for water remediation”. Thus, this paper is not a systematic review, but rather, a critical evaluation of two related aspects that are currently differently appreciated within the research community: (i) permeability loss, and (ii) reactivity loss. These two aspects were well-known in 2007 [9,48,55,56,57] as the major limitations of the Fe0 PRBs technology. This critical review is rooted on the premise that the current confusion within the Fe0 research community results from the fact that the arguments of Noubactep [48,54] and others [46,58] are still largely overlooked within the research community. For this demonstration, recent publications by researchers from the two research groups that had introduced the technology are evaluated: (i) Paul G. Tratnyek from Oregon (Portland, OR, USA), and (ii) Robert W. Gillham from Waterloo (Ontario, OT, Canada).
While there is agreement on the fact that there is “reactivity loss” and permeability loss in Fe0 PRBs, there is a net discrepancy in the appreciation of the origins of these limitations and, thus, on ways to mitigate them. The discrepancy is, in turn, rooted in the significance of the electrochemical nature of aqueous iron corrosion for the process of contaminant removal in Fe0/H2O systems. The majority of active researchers consider that some contaminants can exchange electrons with Fe0 (direct reaction, electrochemical mechanism), while a few authors [24,46,47] maintain that all reductive transformations within Fe0/H2O systems are mediated by secondary reducing agents such as FeII species, Fe3O4, or H2 (indirect reaction, chemical mechanism). A reader can be irritated by such a statement, but it suffices to read Whitney [59] to realize that the science of aqueous iron corrosion was distorted while introducing the Fe0 PRBs technology [24,60].
For the presentation herein, two papers by Dr. Gillham [26,61] and two papers by Dr. Tratnyek [4,14] are considered. From each article, the applied research approach is presented and the soundness of the conclusions is evaluated.

3. Fundamental Flaws in Investigating the Fe0/H2O System

Batch and column experiments are conventionally used to investigate processes relevant for environmental remediation [33,62,63,64,65,66,67,68,69]. However, little guidance is available on: (i) how to perform them, (ii) how they should be set up, and (iii) even how their performance should be evaluated or monitored. The main guidance for investigating Fe0/H2O systems has been the Interstate Technology and Regulatory Council [18,70]. This section covers the design of laboratory experiments to investigate the efficiency of Fe0/H2O systems for water treatment. The presentation is limited to the two major aspects that have avoided or delayed progress in knowledge: (i) the use of vigorously homogenized batch experiments, and (ii) the used of pure Fe0 column systems (100% Fe0).
Fe0 + 2H+ ⇒ Fe2+ + H2
Fe2+ + 3H2O ⇒ FeOOH.H2O + 3H+ + e
Fe0 + RCl + H+ ⇒ RH + Fe2+ + Cl
2Fe2+ + RCl + H+ ⇒ RH + 2Fe3+ + Cl
The system to be investigated is made up of Fe0 that is electrochemically corroded It is fine like this by water or H+ (Equation (1)). For pH > 4.5, Fe2+ is further oxidized, and the resulting Fe3+ precipitates in the vicinity of Fe0 to form an oxide scale (oxide film). In such a system, no Fe0/H2O interface exists, but rather, there are two interfaces: (i) Fe0/oxides, and (ii) oxides/H2O (Figure 1) [24,35]. Clearly, under environmental conditions, the Fe0 surface is permanently shielded by a non-conductive oxide scale acting as: (i) a diffusion barrier for dissolved species, including contaminants and O2, and (ii) a conduction barrier for electrons from the metal body. In other words, the reductive transformation of a chlorinated hydrocarbon (RCl) by Fe0 (Equation (3)) is not possible. Rather, when a reductive transformation is documented, it results from reactions similar to the one depicted in Equation (4) [24,34]. The proponents of the reductive transformation concept fail to explain or totally ignore how the contaminants and O2 by-pass the oxide scale barrier. The question remains as to why the reductive transformation theory is still prevailing 15 years after it was radically refuted? [48,55]. Some answers are given below while considering the design of batch and column experiments. It should be kept in mind that knowledge of the kinetic behavior of Fe0 particles in soils and Fe0 filters is very limited [71]. Thus, one major problem in Fe0 remediation research is that used materials are not characterized for their (long-term) intrinsic reactivity. In essence, the suitability of all relevant materials (including nano Fe0) is based on the electrode potential of the FeII/Fe0 redox couple: E0 = –0.44 V [72,73,74]. Thus, documented differences are due to kinetic factors.

3.1. Batch Experiments

Data to characterize Fe0/H2O systems in batch experiments are conventionally obtained using the bottle-point technique [7,64]. This experimental procedure was borrowed from adsorption studies in which the bottle-point technique is used to determine the adsorption capacity of inert adsorbents. No unified procedure for obtaining adsorption isotherms has been established [62,63,75]. Consequently, each researcher employs a different experimental procedure for the collection of isotherm data. The procedures differ with respect to: (i) buffer application, (ii) the duration of experiments, (iii) Fe0 particle size, (iv) Fe0 pre-treatment, (v) Fe0 type, (vi) shaking type, (vii) shaking intensity and mode, (viii) the volume of the experimental vessels used, (ix) the volume of the added solution, and (x) temperature. As a result, it is very difficult or even impossible to compare independent works from the literature, even for the same contaminants and the same Fe0 material and particle size [75]. The fact that achieved results are often modeled using the wrong reaction (Equation (3) instead of Equation (4)) explains why no efficient descriptor could be presented for the characterization of Fe0/H2O systems [76,77].
The worst mistake in investigating Fe0/H2O systems in batch experiments has been the use of vigorously shaken or homogenized experiments [33,78]. In fact, a key feature of the remediation Fe0/H2O system is the stratification of Fe0/oxides/H2O [48,55] and the existence of two different interfaces (Fe0/oxides and oxides/H2O) [35,60]. While vigorously shaking or stirring their experimental vessels, researchers have produced possibly reproducible data with no practical significance [24,60,75]. This is because such vigorously shaken conditions do not occur in real environmental systems, such as permeable reactive barriers. Such vigorous shaking may also putatively interfere with and disrupt the oxide scale, a phenomenon which may not occur under real environmental conditions [48,55,74]. Moreover, such experimental designs are not able to address the question of whether electron transfer can be effective at pits or crevices as maintained by Huang et al. [4]. This is because everything is suspended and the generated Fe2+ ions have no opportunity to be further transformed in the vicinity of Fe0 (Equation (2)). To the best of the authors’ knowledge, Devlin et al. [79,80] were the first researchers to point out the necessity to keep Fe0 fixed in batch vessels. Noubactep et al. [49] went a step further to let the whole remediation process be strictly diffusion-controlled in quiescent batch experiments (no shaking). It is important to mention that the ITRC [18] has not mentioned this aspect at all.
In summary, the bulk of batch experiments performed for the past 30 years were not designed to answer the key research questions, in particular the mechanism of contaminant removal in Fe0/H2O systems. The current corresponding author established these mechanisms 12 years ago [81,82], but it has been largely ignored (Section 4). In particular, the authors identified herein as leading scientists have not even dared to mention that any alternative idea exists [4,14,26,61]. A similar trend was noticed in recent overview articles [3,4,83,84]. Yet, the named alternative concept is rooted on the science of the system, and has been confirmed by historical findings as summarized by Baker [25] in his seminal reference textbook “The quest for pure water”, which has been re-edited a multiple times.

3.2. Column Experiments

Column experiments enable the investigation of the attenuation of specific species within a specific Fe0/H2O system (Fe0 filter). The design parameters for Fe0 filters include: (i) Fe0 size and type, (ii) filter size, (iii) water flow velocity or residence time, (iv) water quality, including pH value, and (v) temperature. By adjusting these five operational parameters, an infinite number of different efficient Fe0 filters can be designed. This is equally the reason why published results are difficult to compare to each other. Sarr [85] roots the design of Fe0 filters in the stoichiometry of reactions similar to Equation (3). There is already a mistake in such an approach, as water is the sole relevant corroding agent (Equation (1)) [35,59,60]. However, the literature is full of examples of studies comparing the efficiency of Fe0/H2O systems containing various amounts of Fe0 [86,87]. For example, O et al. [87] used 1364.2, 1700.6, 1846.4, and 2045.3 g of four different Fe0 specimens in an effort to characterize the impact of the initial iron corrosion rate on the long-term performance of Fe0 filters. In another example, Bi et al. [86] fixed the “Fe0 + sand” mass to 80 g and varied the Fe0 mass to investigate the impact of the Fe0: sand ratio on the hydraulic conductivity of Fe0 filters. The discussion of both series of results is not addressed herein. The question arises of how the efficiency of systems can be compared without reference to the mass balance [33].
Designing Fe0 filters starting from the mass balance is possible and appropriate, but the mass balance should be the one of Fe. As Fe0 is progressively transformed to its solid corrosion products (FeCPs), a volumetric expansion occurs that is related to the decrease in the density (Table 1). The doxide/diron values in Table 1 suggest that oxides occupy volumes up to 2.3 times larger than the parent metal (Fe0). This means that permeability loss is inherent to Fe0 filters [88,89]. While rooting the design of Fe0 PRBs in the stoichiometry of electrochemical reactions (Equation (3)), it was considered that admixing inert materials such as sand to Fe0 represents a dilution of the reactive material (i.e., Fe0) with possible negative impacts on the contaminant removal capacity of the resulting Fe0 PRBs [86]. However, the expansive nature of aqueous iron corrosion implies that space for accommodating FeCPs must be foreseen and/or the proportion of expansive materials should be decreased. In other words, admixing Fe0 with non-expansive aggregates is even a prerequisite for sustainable Fe0 filters [88,89]. Clearly, only hybrid Fe0/aggregate systems are sustainable (Figure 2) [90]. This evidence is still not clear to many active researchers, mainly because the Fe0 PRB technology was demonstrated with a hybrid system (22% Fe0 w/w) [91], but the majority of the first-generation Fe0 PRBs were built with 100% Fe0. Even today, hybrid Fe0/aggregate systems are still designed without rationalizing their choice [84]. Such a state of affairs unfortunately implies that the results from most column experiments are not transferable to other systems. but are only valid for the specific experimental setup used [24,68].
Iron oxides have characteristic properties that make them useful for a wide range of applications, including environmental remediation [94,95]. They exist under three different main structures with oxygen ratios varying between 22% for wüstite and 37% for FeOOH (Table 1). The oxides with lower oxygen ratios (FeO, Fe3O4, and Fe2O3) are semi-conductors, while FeOOH and hydroxides are insulators. For example, hematite and magnetite have significantly different conductive characteristics despite having less than a 3% difference in oxygen content. Hematite has a corundum structure, while magnetite has an inverse spinel structure. Both oxides are n-type semi-conductors [95]. Hematite has a resistivity of over 105 Ω cm, while magnetite has a lower resistivity of approximately 10−3 Ω cm. These two examples illustrate how the electrical properties of the oxide scale on Fe0 changes with the oxygen content. In essence, electron transfer from the metal (conductor) to FeO and Fe3O4 (semi-conductors) is still theoretically possible, but the transfer is already difficult with Fe2O3 and is hindered by insulators (FeOOH and hydroxides). This is the reason why: (i) no electron transfer from Fe0 to dissolved contaminants is possible, and (ii) batch experiments pertaining to designing Fe0 PRBs should be exclusively performed under quiescent conditions [48,55].
In summary, column experiments can provide excellent estimates of all relevant design parameters for field situations [9]. In addition, the obtained results are almost always directly transferable to water treatment at a field scale. The challenge for the future is to systematically test a range of hybrid Fe0/aggregate systems to derive a series of applicable designs for water treatment. Such experiments should last for several months, ideally not less than one year [96,97].

4. Disregarding the State-of-the-Art Knowledge

The issue of this communication is that the lack of true collaboration is the major challenge to the Fe0 remediation research community. This section illustrates this by presenting five selected recent publications on the removal of methylene blue (MB) in Fe0/H2O systems [98,99,100,101,102]. These works were published more than three years after the MB method was presented in an overview article in the Journal of Environmental Management—Elsevier [103]. The MB method itself was first published a decade ago [104,105]. While the MB method has roots in the historical work of Mitchell et al. [106] and the evidence that iron oxides are poor adsorbents for MB [35], the five referenced papers [98,99,100,101,102] used MB as a model contaminant to characterize some aspects of the Fe0/H2O system. The following section summarizes the studies, and highlights avoidable mistakes by these authors.
Table 2 and Table 3 summarize the experimental conditions of the five articles. It is seen that the MB adsorptive capacities of nano-Fe0 and their modifications were investigated in batch studies. Table 3 also presents the key findings of individual studies. The MB initial concentrations varied from 10 to 2000 mg/L, and the Fe0 loading varied from 0.5 to 14 g/L. The initial pH value varied from 2.0 to 13.0, and the volume of the working solution varied from 20 to 250 mL. Several mixing speeds and intensities were tested, and the experimental duration, where specified, varied from 5 min to 72 h. The suitability of these operational parameters to meet the intended goal in individual studies has been discussed above (Section 3.1) [78,107]. It suffices to recognize that experimental results obtained under such variable conditions are highly qualitative. Moreover, the five papers have collectively ignored the MB method in their presentation. The fact that quantitative MB removal is reported in all these articles is a proof that co-precipitation is the main removal path in Fe0/H2O systems. The following should be kept in mind: (i) MB adsorbs only very weakly onto iron oxides (FeCPs) [106], and (ii) MB discoloration cannot be achieved by any reductive reaction in the Fe0/H2O system [108].
Table 2 reveals that Wang et al. [101] used the longest experimental duration (72 h or 3 d) and one of the highest mixing speeds (120 rpm) among the five papers. It is certain that Fe0 cannot be completely exhausted within 3 days. In other words, it is a mistake to use “adsorption capacity” in discussing the results. There are at least two reasons for this: (i) MB removal by iron oxides is not an adsorptive process, and (ii) no equilibrium state can be achieved in such a dynamic system before total Fe0 depletion. Clearly, once immersed in water, Fe0 continuously generates FeCPs, which progressively age within the system [104,109,110,111]. The next concern is that the corrosion rate (corrosion kinetics) is not known [97,112,113,114,115].
In summary, the novelty of the five papers is collectively questioned because they failed to consider the state-of-the-art knowledge while designing their experiments and discussing their results. Unfortunately, the five papers, which were selected for having investigated MB removal, are not an exception, but rather, represent the rule within the Fe0 remediation research community. The current most holistic understanding of the Fe0/H2O system has been presented by our research group, but is currently ignored by the large majority of active researchers. In other words, an artificial knowledge gap exists, and within this virtual gap, a circular reasoning has already been established for three decades.

5. Scientists Disregard Progress

Using Fe0 for in-situ aquifer remediation has undergone rapid development during the past three decades [16,17]. Related applications offer the promise of a more thorough treatment of several classes of contaminants [1,12]. However, there is controversy in the scientific literature about the key issues determining the efficiency and sustainability of Fe0 filters (Section 3). A starting researcher, seeking the state-of-the-art knowledge on “remediation using Fe0” may identify research groups with the longest expertise [116]. Two such research groups are those of Dr. Paul G. Tratnyek (Oregon/USA) and Dr. Robert W. Gillham (Ontario/Canada). In this section, one research article and one overview article of each group is assessed with regard to the extent to which they have falsified the current state-of-the-art knowledge, which was achieved 11 years ago [81,82]. Interestingly, none of these four articles referenced a single paper out of several of them from Noubactep.

5.1. The Overview Articles

The overview book chapter single-authored by Dr. Gillham [5] retraced the history of “Fe0 for groundwater remediation” from the fortuitous observation of Glen Reynolds [117] to about 2007, when Dr. Noubactep first presented strong arguments against the “reductive transformation” concept in the peer-reviewed literature [48,55]. The idea was published before in the grey literature [118,119]. The wording of Dr. Gillham is reproduced here for the sake of clarity:
Glen Reynolds “was investigating the potential for sampling bias caused by sorption of chlorinated organic contaminants to well casings and other materials commonly used in groundwater sampling. From the results of laboratory batch tests, it was clear that contaminants were lost from solution through diffusion into polymer materials such as PVC and Teflon™. Unexpectedly however, losses were also observed from solutions in contact with a variety of metals, though the pattern of loss was inconsistent with a diffusion process. No literature was found that was directly relevant, however reductive dechlorination was proposed as the most likely cause [117]. Though initially observed in 1984, the potential significance of these early results with respect to groundwater remediation was not recognized until 1989. The results were shown to be reproducible, leading to the “laboratory testing and verification” stage of development.”
The mistake made by Dr. Gillham was an incomplete literature review for at least two reasons: (i) it was known since 1903 that, under ambient conditions, Fe0 is corroded by water and by water alone [59], and (ii) it is textbook knowledge that Fe0 is used in organic synthesis, for example, for the synthesis of aniline (Bechamp Reduction or Bechamp Process, 1854). It is interesting in this regard to state that, while rationalizing nitrobenzene reduction to aniline using Fe0 and acid (e.g., HCl), the direct transfer of electrons from Fe0 to nitrobenzene has never been convincingly established [120]. On the other hand, while using Fe0 in hydrometallurgy for the reductive dissolution of manganese ore (e.g., MnO2) in acidic solutions, Bafghi et al. [121] also could not established direct electron transfer to MnO2. Even while using Fe0 for wastewater treatment during the 1980s and 1990s, direct electron transfer to contaminants was not established [122,123,124]. In particular, Khudenko [124] used the cementation of Cu2+ by Fe0 as a tool to induce the reductive “destruction of organics”. Clearly, in the Khudenko process, Fe0 is oxidized by Cu2+ to produce Fe2+ and H2, which chemically reduce organics. Whitney [59] and Khudenko [124] were not considered while introducing the reductive degradation concept for organics [7,125,126]. This was the beginning of a circular reasoning that cannot be corrected until electron transfer from Fe0 is taken out of the equation [24].
The overview article co-authored by Dr. Tratnyek [4] just perpetuated the named mistakes. The article is entitled “Fe(II) redox chemistry in the environment” and comprises a section on “Zero-Valent Iron” wherein the “prominent role of Fe0 in the environment” is presented. The following three major mistakes were extracted from Huang et al. [4]:
(i)
Acknowledging the non-stoichiometric nature of decontamination reactions while using pseudo-first-order rate constants (kobs, kM, and kSA) or the corresponding half-lives (t1/2) to describe the systems.
(ii)
Considering that any Fe0/H2O interface exists, including within pits, crevices, or equivalent defects in the oxide scale on Fe0. Direct electron transfer from Fe0 to contaminants (electrochemical reduction) supposedly occurs at such sites. Because the formation of oxide scale is spontaneous, it also occurs at defects such that there is always an Fe0/oxides interface.
(iii)
Improperly considering the kinetics of iron corrosion, which is non-constant and non-linearly varies with time [90,114,115].

5.2. The Research Articles

The context of Cai et al. [14] is the “development of sulfidated Fe0 materials for environmental applications” as recently reviewed by Fan et al. [127]. This application is regarded as the convergence of two developments over the past three decades: (i) the diversification of the use of Fe0/H2O systems for water treatment, and (ii) the optimization of the role of FexSy minerals in contaminant transformation processes [14,127]. It is certain that sulfidation of Fe0 increases the number of galvanic cells (Fe-S) in the metal body or at its surface and, thus, enhances the reactivity of Fe0. However, enhancing electron transfer within the metal body does not change the non-conductive nature of the oxide scale in the vicinity of Fe0. This fact was revealed to the research community as early as 2009 [128], but has been largely overlooked in the majority of published studies. In fact, according to Google Scholar, Noubactep [128] has been referenced only 33 times during the past 12 years. That is less than three times per year. The specific aims of Cai et al. [14] are not addressed herein, as they fall under the shortcoming of batch experiments, as discussed in Section 3. On the other hand, these authors could have investigated parallel systems for MB discoloration (MB method) to improve the discussion of their results (Section 4).
Jeen et al. [61] characterized the degradation of two chlorofluorocarbons (CFC11 and CFC113) by granular Fe0, Fe0/Ni0, and Fe0/Pd0 in flow-through column tests. The results revealed rapid and better degradation efficiencies in bimetallic systems (Fe0/Ni0 and Fe0/Pd0) relative to Fe0 for both compounds. However, for CFC113 the average pseudo-first-order rate constants were only about two to three times greater than for Fe0. These smaller-than-expected differences in degradation rate constants between Fe0 and its bimetallic counterparts was a clear hint that the underlying reaction mechanism may be unstable. Instead of checking this alternative explanation, Jeen et al. [61] suggested very complex explanations to cope with the reductive degradation concept. The approach of Jeen et al. [61] is rather characteristic within the Fe0 remediation research community.

6. Discussion

Fe0 PRBs have been used for groundwater remediation since the early 1990s [26]. It is a tangible fact that Fe0 oxidation by groundwater releases ferrous iron species (FeII) and hydrogen (H2) and typically increases the pH value (i.e., H+ consumption) (Equation (1)) while lowering the oxidation/reduction potential (creating reducing conditions). The generated FeII species are further transformed to various FeII/FeIII oxides and hydroxides, which are excellent contaminant scavengers. These modifications cause the reductive degradation of some parent chemicals (including contaminants), but ideally cause the removal of all species (contaminants and daughter products) by adsorption and co-precipitation. The contaminant scavengers account for the removal of non-reducible contaminants such as methylene blue (MB) and Zn2+. In contrast, proponents of the reductive transformation concept fail to explain how non-reducible contaminants are removed in Fe0/H2O systems or consider that only such contaminants are removed by adsorption and co-precipitation. To date, the evaluation of the performance of Fe0 PRBs at field sites has focused on monitoring the following: (i) compliance with groundwater standards, (ii) permeability alterations, (iii) identifying the nature of mineral precipitates, (iv) the extent of Fe0 exhaustion, and (v) potential biologically mediated reactions [129,130,131,132]. The discussion of Fe0 longevity has been a function of amount and size, but has not yet considered the Fe0 ratio within the reactive zone. Unfortunately, without considering the Fe0 ratio no convincing discussion of the biological activity can be given [97,131,132,133,134,135]. Ideally, the reactivity and size of the admixing aggregates (e.g., FeS2, MnO2, and sand) should also be considered. In other words, existing guidance documents for the design of Fe0 PRBs (e.g., ref. [18]) have to be amended.
Apart from the MB method, no other quantification method to account for the extent of Fe0 “passivation” on the efficiency of Fe0 filters has been presented. On the other hand, the permeability of a Fe0 PRB depends both on Fe0 reactivity and the amount of in-situ generated iron minerals (FeCPs). However, the law of generation of these FeCPs is not known and its extent depends on the (residual) porosity and the pore size distribution [114,115]. Virtually, the porosity increases if the following three conditions are observed: (i) there is no inflow of suspended particles, (ii) there is no significant accumulation of removed contaminants, and (iii) there is no mineral precipitation. This idealized case corresponds to the situation where contaminants are present in trace amounts. However, at pH values larger than 4.5, mineral precipitation is certain and quantitative. Moreover, the density of individual FeCPs (d < 5 g/cm3) is lower compared to that of Fe0 (d = 7.86 g/cm3), meaning that each FeCP occupies a larger volume than Fe0 in the porous media. This implies that porosity decrease is the rule for field Fe0 PRBs, regardless of the material’s intrinsic reactivity. Consequently, to understand the mechanisms underlying porosity and permeability changes occurring in PRBs, the reasoning should start with the extent of Fe0 depletion. The problem with the available models is that they have considered iron corrosion by water as a side reaction, thereby neglecting the main pore filling factor. It is, thus, not surprising that no conclusive discussion was presented [89,97]. In one case, a time-dependent increase in the permeability of Fe0 PRBs was predicted [136].
Another possible reason accounting for the continued propagation of the reductive transformation concept despite its established flaws [81,82,137,138,139] is the bandwagon or ‘Matthew’ effect. The bandwagon or Matthew effect implies that a concept (even a flawed one) that has been widely publicized and cited may attract even more subsequent research attention compared to other competing theories. Overall, this creates bias in research with respect to research focus and the allocation of scarce resources, including research grants [140]. This is particularly important for early career researchers who may be desperate to be published or secure research grants. The Matthew effect, its impacts on research, and possible solutions have been discussed extensively in earth sciences [141,142] and environmental sciences [143,144]. The Fe0/H2O remediation community need to develop active measures to guard against such phenomenon and root the field in the fundamentals of aqueous iron corrosion [24,145,146,147].

7. Questioning the Reputation of Journals

There is a current discussion on the meaning of journal ranking for what can be called the survival of science [140,148]. Reputed journals advertise with their rejection quotes [148]. Some journals publish submitted manuscripts only if the comments of at least 66% of the invited reviews are favorable. From the standpoint of the Fe0 literature, it is not exaggerated to assume that part of the rejected manuscripts, constituting up to 33%, could be scientifically sound. This is because, as in any field it is difficult, if not even impossible, to publish a paper with an alternative theory contrary to what is believed to be correct by the majority [140,148]. This is the case, even when the majority may believe in a flawed concept [149,150,151,152]. In fact, as discussed in Section 3, in the case of Fe0/H2O systems, it has been very difficult to publish articles moving “against the tide” that considers Fe0 as a reducing agent [24]. The fundamental reason why Fe0 is suitable for environmental remediation is two-fold: (i) FeII and FeIII species are stable under environmental conditions, and (ii) the structures of Fe0, FeII, and FeIII precipitates (hydroxides and oxides) are considerably different (Table 1) [72,73,153,154]. In fact, Al0 (E0 = –1.676 V) and Zn0 (E0 = –0.762 V) are by far stronger reducing agents than Fe0 (E0 = –0.440 V) and have been reported to be initially more reactive for contaminant reductive transformations [8,28]. However, as corrosion proceeds, Al0 and Zn0 are progressively covered by impervious layers of hydroxide (Al(OH)3) or oxide (ZnO), respectively. Conversely, Fe0 is just shielded by a porous layer of mixed oxides and hydroxides that are non-protective [153,154]. This knowledge was presented as a short communication in the peer-reviewed literature as early as 2010 [154], but has been largely ignored, with just 76 citations in more than 12 years (Google Scholar).
The Fe0/H2O system is certainly a very complex one whose profound understanding requires knowledge from various branches of sciences, including chemistry, electrochemistry, materials science, mechanical engineering, and microbiology [155]. The first four peer-reviewed journal articles on Fe0/H2O systems [5,6,7,8] appeared in reputed journals: Chemosphere [6,8], Environmental Science & Technology [7], and Ground Water [5]. Over the years, both good articles based on the fundamental science of aqueous iron corrosion [47,156] and poor articles perpetuating a flawed concept considering Fe0 as a reducing agent [12,14] were published in all relevant journals. This state of affairs questions the relevance of journal impact factors in evaluating scientists and grant proposals. In fact, a number of cases have been reported in recent years where the impact factors and citations of reputable journals have been manipulated by some members of the journal editorial teams [157,158,159,160]. It seems the right time to consider that each article has its own impact, and should be assessed based on its merit.

8. Conclusions

An increasing amount of attention is directed at documenting and explaining the association between Fe0 oxidative dissolution (iron corrosion) and contaminant removal in Fe0/H2O systems, including Fe0 PRBs. This collective search for the rationale for efficient Fe0-based systems has revealed the key information about the sustainability of Fe0 PRBs. However, the proper interpretation of the available data is biased by the initial consideration that Fe0 is an environmental reducing agent. The biogeochemical complexity of the Fe0/H2O system is obvious, but the fact that Fe0 reactivity declines with the service life of Fe0-based systems (reactivity loss) is also well-known. Therefore, prominent theoretical considerations should be revisited in an effort to determine the explanatory value of standard hypotheses and explore alternative explanations for tangible experimental and field observations. Furthermore, the theoretical perspectives presented herein deserve greater attention because they have already enabled the explanation of many reported controversies. In particular, the current three-fold uncertainty in designing Fe0 PRBs (mechanisms, permeability, and reactivity) is reduced to just one, since mechanisms and reactivity are known. Thus, only changes in the permeability are uncertain, but they are coupled to the following: (i) Fe0/aggregate ratio, (ii) Fe0 amount, (iii) Fe0 reactivity, (iv) Fe0 size, and (v) initial porosity. Conventional methodological tools to evaluate the performance of Fe0 filtration systems are limited in their efficacy, but at the same time, the limitations present good opportunities to explore new tools for progress in this line of research. Deciphering the sources of permeability loss has the potential to establish Fe0 filtration systems as a reference technology for water treatment at several scales. All that is needed is true collaborative research according to the rules of good science. The minimum requirement is the consideration of the state-of-the-art knowledge while presenting one’s own achievements or contributions.
The highly interdisciplinary nature of research on the Fe0/H2O system and the abundance of related publications suggest that annual review articles in relevant journals (including Water/MDPI) can be a tool to offer early-career scientists the best assistance. Currently, it is very difficult to decide where to start, despite the large availability of review articles. Such an annual review can be written as a scoping communication focusing on the papers published within the calendar year (January/December).

Author Contributions

Conceptualization: M.X., R.H. and C.N.; methodology: M.X. and A.I.N.-T.; writing—original draft: M.X., A.I.N.-T., W.G. and C.N.; writing—review and editing: M.X., W.G., R.H. and C.N.; supervision: R.H. and C.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology of China through the Program “Driving process and mechanism of three dimensional spatial distribution of high risk organic pollutants in multi field coupled sites” (Project Code: 2019YFC1804303).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank the peer reviewers for their valuable suggestions and comments that improved this paper. We acknowledge support by the German Research Foundation and the Open Access Publication Funds of the University of Göttingen.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guan, X.; Sun, Y.; Qin, H.; Li, J.; Lo, I.M.; 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]
  2. Naidu, R.; Birke, V. Permeable Reactive Barrier: Sustainable Groundwater Remediation; Taylor & Francis: Abingdon, UK, 2015; p. 333. ISBN 978-1-4822-2447-4. [Google Scholar]
  3. Metzgen, A.D.; Dahmke, A.; Ebert, M. Anaerobic corrosion of zero-valent iron at elevated temperatures. Environ. Sci. Technol. 2021, 55, 8010–8019. [Google Scholar] [CrossRef] [PubMed]
  4. Huang, J.; Jones, A.; Waite, T.D.; Chen, Y.; Huang, X.; Rosso, K.M.; Kappler, A.; Mansor, M.; Tratnyek, P.G.; Zhang, H. Fe (II) redox chemistry in the environment. Chem. Rev. 2021, 121, 8161–8233. [Google Scholar] [CrossRef] [PubMed]
  5. Gillham, R.W.; O’Hannesin, S.F. Enhanced degradation of halogenated aliphatics by zero-valent iron. Ground Water 1994, 32, 958–967. [Google Scholar] [CrossRef]
  6. Lipczynska-Kochany, E.; Harms, S.; Milburn, R.; Sprah, G.; Nadarajah, N. Degradation of carbon tetrachloride in the presence of iron and sulphur containing compounds. Chemosphere 1994, 29, 1477–1489. [Google Scholar] [CrossRef]
  7. Matheson, L.J.; Tratnyek, P.G. Reductive dehalogenation of chlorinated methanes by iron metal. Environ. Sci. Technol. 1994, 28, 2045–2053. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  10. Bartzas, G.; Komnitsas, K. Solid phase studies and geochemical modelling of low-cost permeable reactive barriers. J. Hazard. Mater. 2010, 183, 301–308. [Google Scholar] [CrossRef]
  11. Li, L.; Benson, C.H. Evaluation of five strategies to limit the impact of fouling in permeable reactive barriers. J. Hazard. Mater. 2010, 181, 170–180. [Google Scholar] [CrossRef]
  12. 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. 2015, 431, 48–60. [Google Scholar] [CrossRef]
  13. Warner, S.D. Two Decades of Application of Permeable Reactive Barriers to Groundwater Remediation. In Permeable Reactive Barrier Sustainable Groundwater Remediation; Naidu, R., Birke, V., Eds.; CRC Press: Boca Raton, FL, USA, 2015; pp. 25–39. ISBN 978-1-4822-2447-4. [Google Scholar]
  14. Cai, S.; Chen, B.; Qiu, X.; Li, J.; Tratnyek, P.G.; He, F. Sulfidation of zero-valent iron by direct reaction with elemental sulfur in water: Efficiencies, mechanism, and dechlorination of trichloroethylene. Environ. Sci. Technol. 2021, 55, 645–654. [Google Scholar] [CrossRef] [PubMed]
  15. Wei, K.; Li, H.; Gu, H.; Liu, X.; Ling, C.; Cao, S.; Li, M.; Liao, M.; Peng, X.; Shi, Y.; et al. Strained zero-valent iron for highly efficient heavy metal removal. Adv. Funct. Mater. 2022, 2200498. [Google Scholar] [CrossRef]
  16. Chen, Q.; Fan, G.; Na, W.; Liu, J.; Cui, J.; Li, H. Past, present, and future of groundwater remediation research: A scientometric analysis. Int. J. Environ. Res. Public Health 2019, 16, 3975. [Google Scholar]
  17. Li, X.; Li, Z.; Du, C.; Tian, Z.; Zhu, Q.; Li, G.; Shen, Q.; Li, C.; Li, J.; Li, W.; et al. Bibliometric analysis of zerovalent iron particles research for environmental remediation from 2000 to 2019. Environ. Sci. Pollut. Res. 2021, 28, 4200–34210. [Google Scholar] [CrossRef]
  18. ITRC (Interstate Technology & Regulatory Council). 2011. Permeable Reactive Barrier: Technology Update. PRB-5. Washington, D.C.: Interstate Technology & Regulatory Council, PRB: Technology Update Team. Available online: www.itrcweb.org (accessed on 9 March 2012).
  19. Igliˇc, H.; Doreian, P.; Kronegger, L.; Ferligoj, A. With whom do researchers collaborate and why? Scientometrics 2017, 112, 153–174. [Google Scholar] [CrossRef] [Green Version]
  20. Kyvik, S.; Reymert, I. Research collaboration in groups and networks: Differences across academic fields. Scientometrics 2017, 113, 951–967. [Google Scholar] [CrossRef]
  21. Gheju, M. Progress in understanding the mechanism of CrVI Removal in Fe0-based filtration systems. Water 2018, 10, 651. [Google Scholar] [CrossRef] [Green Version]
  22. Dusdal, J.; Powell, J.J. Benefits, motivations, and challenges of international collaborative research: A sociology of science case study. Sci. Public Policy 2021, 48, 235–245. [Google Scholar] [CrossRef]
  23. Öcalan-Özel, S.; Llerena, P. Industry collaborations of research teams: Are they penalized or rewarded in the grant evaluation process? Front. Res. Metr. Anal. 2021, 6, 707278. [Google Scholar] [CrossRef]
  24. Hu, R.; Nde-Tchoupe, A.I.; Cao, V.; Gwenzi, W.; Noubactep, C. Metallic iron for environmental remediation: The fallacy of the electron efficiency concept. Front. Environ. Chem. 2021, 2, 677813. [Google Scholar] [CrossRef]
  25. Baker, M.N. The Quest for Pure Water: The History of Water Purification from the Earliest Records to the Twentieth Century; American Water Works Assn.: New York, NY, USA, 1948; p. 527. [Google Scholar]
  26. Gillham, R.W. Development of the granular iron permeable reactive barrier technology (good science or good fortune). In Advances in environmental geotechnics. In Proceedings of the International Symposium on Geoenvironmental Engineering, Hangzhou, China, 8–10 September 2007; Chen, Y., Tang, X., Zhan, L., Eds.; Springer: Berlin, Germany; London, UK, 2008; pp. 5–15. [Google Scholar]
  27. Antia, D.D.J. Water treatment and desalination using the eco-materials n-Fe0 (ZVI), n-Fe3O4, n-FexOyHz[mH2O], and n-Fex[Cation]nOyHz[Anion]m [rH2O]. In Handbook of Nanomaterials and Nanocomposites for Energy and Environmental Applications; Kharissova, O.V., Ed.; Springer Nature: Basel, Switzerland, 2020; p. 66-1. [Google Scholar] [CrossRef]
  28. Warren, K.D.; Arnold, R.G.; Bishop, T.L.; Lindholm, L.C.; Betterton, E.A. Kinetics and mechanism of reductive dehalogenation of carbon tetrachloride using zero-valence metals. J. Hazard. Mater. 1995, 41, 217–227. [Google Scholar] [CrossRef]
  29. Fiedor, J.N.; Bostick, W.D.; Jarabek, R.J.; Farrell, J. Understanding the mechanism, of uranium removal from groundwater by zero-valent iron using X-ray photoelectron spectroscopy. Environ. Sci. Technol. 1998, 32, 1466–1473. [Google Scholar] [CrossRef]
  30. Farrell, J.; Bostick, W.D.; Jarabek, R.J.; Fiedor, J.N. uranium removal froum ground water using zero valent iron media. Ground Water 1999, 34, 618–624. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. Lee, G.; Rho, S.; Jahng, D. Design considerations for groundwater remediation using reduced metals. Korean J. Chem. Eng. 2004, 21, 621–628. [Google Scholar] [CrossRef]
  34. Jiao, Y.; Qiu, C.; Huang, L.; Wu, K.; Ma, H.; Chen, S.; Ma, L.; Wu, D. Reductive dechlorination of carbon tetrachloride by zero-valent iron and related iron corrosion. Appl. Catal. B Environ. 2009, 91, 434–440. [Google Scholar] [CrossRef]
  35. Konadu-Amoah, B.; Nde-Tchoupe, A.I.; Hu, R.; Gwenzi, W.; Noubactep, C. Investigating the Fe0/H2O systems using the methylene blue method: Validity, applications and future directions. Chemosphere 2002, 291, 132913. [Google Scholar] [CrossRef]
  36. Lin, Y.-T.; Weng, C.-H.; Chen, F.-Y. Effective removal of AB24 dye by nano/micro-size zero-valent iron. Sep. Purif. Technol. 2008, 64, 26–30. [Google Scholar] [CrossRef]
  37. Mielczarski, J.A.; Atenas, G.M.; Mielczarski, E. Role of iron surface oxidation layers in decomposition of azo-dye water pollutants in weak acidic solutions. Appl. Catal. B 2005, 56, 289–303. [Google Scholar] [CrossRef]
  38. Hou, M.; Li, F.; Liu, X.; Wang, X.; Wan, H. The effect of substituent groups on the reductive degradation of azo dyes by zerovalent iron. J. Hazard. Mater. 2007, 145, 305–314. [Google Scholar] [CrossRef] [PubMed]
  39. Li, T.; Chen, Y.; Wan, P.; Fan, M.; Yang, X.J. Chemical degradation of drinking water disinfection by products by millimeter-sized particles of iron-silicon and magnesium-aluminum alloys. J. Am. Chem. Soc. 2010, 132, 2500–2501. [Google Scholar] [CrossRef] [PubMed]
  40. Xu, F.; Deng, S.; Xu, J.; Zhang, W.; Wu, W.; Wang, W.; Huang, J.; Yu, G. Highly Active and Stable Ni–Fe Bimetal Prepared by Ball Milling for Catalytic Hydrodechlorination of 4-Chlorophenol. Environ. Sci. Technol. 2012, 46, 4576–4582. [Google Scholar] [CrossRef] [PubMed]
  41. Touomo-Wouafo, M.; Donkeng-Dazie, J.; Btatkeu-K, B.D.; Tchatchueng, J.B.; Noubactep, C.; Ludvík, J. Role of pre-corrosion of Fe0 on its efficiency in remediation systems: An electrochemical study. Chemosphere 2018, 209, 617–622. [Google Scholar] [CrossRef]
  42. Vollprecht, D.; Krois, L.-M.; Sedlazeck, K.P.; Müller, P.; Mischitz, R.; Olbrich, T.; Pomberger, R. Removal of critical metals from waste water by zero-valent iron. J. Clean. Prod. 2018, 208, 1409–1420. [Google Scholar] [CrossRef]
  43. Touomo-Wouafo, M.; Donkeng-Dazie, J.; Jirka, I.; Btatkeu-K, B.D.; Tchatchueng, J.B.; Noubactep, C.; Ludvík, J. Electrochemical monitoring of heavy metals removal from aqueous solutions by aged metallic iron. Competitive effects of cations Zn2+, Pb2+ and Cd2+. Monatsh. Chem. 2020, 151, 1511–1523. [Google Scholar] [CrossRef]
  44. Vollprecht, D.; Plessl, K.; Neuhold, S.; Kittinger, F.; Öfner, W.; Müller, P.; Mischitz, R.; Sedlazeck, K.P. Recovery of Molybdenum, Chromium, Tungsten, Copper, Silver, and Zinc from Industrial Waste Waters Using Zero-Valent Iron and Tailored Beneficiation Processes. Processes 2020, 8, 279. [Google Scholar] [CrossRef] [Green Version]
  45. Li, T.; Gao, C.; Wang, W.; Teng, Y.; Li, X.; Wang, H. Strong influence of degree of substitution on carboxymethyl cellulose stabilized sulfidated nanoscale zero-valent iron. J. Hazard. Mater. 2022, 425, 128057. [Google Scholar] [CrossRef]
  46. Ghauch, A. Iron-based metallic systems: An excellent choice for sustainable water treatment. Freiberg Online Geosci. 2005, 32, 1–80. [Google Scholar]
  47. 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] [PubMed]
  48. Noubactep, C. A critical review on the mechanism of contaminant removal in Fe0–H2O systems. Environ. Technol. 2008, 29, 909–920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Noubactep, C.; Meinrath, G.; Dietrich, P.; Merkel, B. Mitigating uranium in groundwater: Prospects and limitations. Environ. Sci. Technol. 2003, 37, 4304–4308. [Google Scholar] [CrossRef] [Green Version]
  50. Gheju, M.; Iovi, A. Kinetics of hexavalent chromium reduction by scrap iron. J. Hazard. Mater. 2006, 135, 66–73. [Google Scholar] [CrossRef] [PubMed]
  51. Ghauch, A.; Tuqan, A. Catalytic degradation of chlorothalonil in water using bimetallic iron-based systems. Chemosphere 2008, 73, 751–759. [Google Scholar] [CrossRef] [PubMed]
  52. Hu, R.; Gwenzi, W.; Sipowo-Tala, V.R.; Noubactep, C. Water treatment using metallic iron: A tutorial review. Processes 2019, 7, 622. [Google Scholar] [CrossRef] [Green Version]
  53. 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] [Green Version]
  54. Hu, R.; Yang, H.; Tao, R.; Cui, X.; Xiao, M.; Amoah, B.K.; Cao, V.; Lufingo, M.; Soppa-Sangue, N.P.; Nde-Tchoupe, A.I.; et al. Metallic iron for environmental remediation: Starting an overdue progress in knowledge. Water 2020, 12, 641. [Google Scholar] [CrossRef] [Green Version]
  55. 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]
  56. Cundy, A.B.; Hopkinson, L.; Whitby, R.L. Use of iron-based technologies in contaminated land and groundwater remediation: A review. Sci. Total Environ. 2008, 400, 42–51. [Google Scholar] [CrossRef]
  57. Thiruvenkatachari, R.; Vigneswaran, S.; Naidu, R. Permeable reactive barrier for groundwater remediation. J. Ind. Eng. Chem. 2008, 14, 145–156. [Google Scholar] [CrossRef]
  58. Gheju, M. Hexavalent chromium reduction with zero-valent iron (ZVI) in aquatic systems. Water Air Soil Pollut. 2011, 222, 103–148. [Google Scholar] [CrossRef]
  59. Whitney, W.R. The corrosion of iron. J. Am. Chem. Soc. 1903, 25, 4, 394–406. [Google Scholar] [CrossRef]
  60. Konadu-Amoah, B.; Hu, R.; Yang, H.; Cao, V.; Nde-Tchoupe, A.; Gwenzi, W.; Ruppert, H.; Noubactep, C. Realizing the potential of metallic iron for environmental remediation: Flee or adapt? Appl. Water Sci. 2022, 12, 641, (Submitted). [Google Scholar]
  61. Jeen, S.W.; Lazar, S.; Gui, L.; Gillham, R.W. Degradation of chlorofluorocarbons using granular iron and bimetallic irons. J. Contam. Hydrol. 2014, 158, 55–64. [Google Scholar] [CrossRef]
  62. Vidic, R.D.; Suidan, M.T.; Brenner, R.C. Adsorption isotherms: Illusive capacity and role of oxygen. War. Res. 1993, 24, 1187–1195. [Google Scholar] [CrossRef]
  63. Vidic, R.D.; Suidan, M.T.; Brenner, R.C. Oxidative coupling of phenols on activated carbon: Impact on adsorption equilibrium. Environ. Sci. Technol 1993, 27, 2079–2085. [Google Scholar] [CrossRef]
  64. Richardson, J.P.; Nicklow, J.W. In situ permeable reactive barriers for groundwater contamination. Soil Sediment Contam. 2002, 11, 241–268. [Google Scholar] [CrossRef]
  65. Dries, J.; Bastiaens, L.; Springael, D.; Agathos, S.N.; Diels, L. Combined removal of chlorinated ethenes and heavy metals by zerovalent iron in batch and continuous flow column systems. Environ. Sci. Technol. 2005, 39, 8460–8465. [Google Scholar] [CrossRef]
  66. Biterna, M.; Antonoglou, L.; Lazou, E.; Voutsa, D. Arsenite removal from waters by zero valent iron: Batch and column tests. Chemosphere 2010, 78, 7–12. [Google Scholar] [CrossRef]
  67. Ali, I. Water treatment by adsorption columns: Evaluation at ground level. Sep. Purif. Rev. 2014, 43, 175–205. [Google Scholar] [CrossRef]
  68. Banzhaf, S.; Hebig, K.H. Use of column experiments to investigate the fate of organic micropollutants—A review. Hydrol. Earth Syst. Sci. 2016, 20, 3719–3737. [Google Scholar] [CrossRef] [Green Version]
  69. Carneiro, M.A.; Pintor, A.; Boaventura, R.A.; Botelho, C. Current trends of arsenic adsorption in continuous mode: Literature review and future perspectives. Sustainability 2021, 13, 1186. [Google Scholar] [CrossRef]
  70. Lanet, P.; Deluchat, V.; Baudu, M. Relevant design parameters for a reactor used in P removal with ZVI-based materials. J. Ind. Eng. Chem. 2021, 104, 8–21. [Google Scholar] [CrossRef]
  71. Lufingo, M.; Nde-Tchoupe, 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]
  72. 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]
  73. Lazzari, L. General aspects of corrosion. In Encyclopedia of Hydrocarbons; Istituto Enciclopedia Italiana: Rome, Italy, 2008; Volume V, pp. 485–505. [Google Scholar]
  74. Noubactep, C.; Care, S.; Crane, R. Nanoscale metallic iron for environmental remediation: Prospects and limitations. Water Air Soil Pollut. 2012, 223, 1363–1382. [Google Scholar] [CrossRef] [Green Version]
  75. Naseri, E.; Nde-Tchoupe, 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]
  76. 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]
  77. 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]
  78. Noubactep, C.; Licha, T.; Scott, T.; 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] [PubMed] [Green Version]
  79. Devlin, J.F.; Klausen, J.; Schwarzenbach, R.P. Kinetics of nitroaromatic reduction on granular iron in recirculating batch experiments. Environ. Sci. Technol. 1998, 32, 1941–1947. [Google Scholar] [CrossRef]
  80. Devlin, J.; Allin, K. 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] [PubMed]
  81. Noubactep, C. The fundamental mechanism of aqueous contaminant removal by metallic iron. Water SA 2010, 36, 663–670. [Google Scholar] [CrossRef] [Green Version]
  82. Noubactep, C. Aqueous contaminant removal by metallic iron: Is the paradigm shifting? Water SA 2011, 37, 419–426. [Google Scholar] [CrossRef] [Green Version]
  83. Thakur, A.K.; Vithanage, M.; Das, D.B.; Kumar, M. A review on design, material selection, mechanism, and modelling of permeable reactive barrier for community-scale groundwater treatment. Environ. Technol. Innov. 2020, 19, 100917. [Google Scholar] [CrossRef]
  84. Njaramba, L.K.; Park, J.B.; Lee, C.S.; Nzioka, A.M.; Kim, Y.J. Permeable reactive barriers with zero-valent iron and pumice for remediation of groundwater contaminated with multiple heavy metals. Environ. Eng. Sci. 2021, 38, 245–255. [Google Scholar] [CrossRef]
  85. Sarr, D. Zero-valent-iron permeable reactive barriers-how long will they last? Remediation 2001, 11, 1–18. [Google Scholar]
  86. Bi, E.; Devlin, J.; Huang, B. Effects of mixing granular iron with sand on the kinetics of trichloroethylene reduction. Ground Water Monit. Remed. 2009, 29, 56–62. [Google Scholar] [CrossRef]
  87. Jeen, S.W.; Gillham, R.W.; Gui, L. Effects of initial iron corrosion rate on long-term performance of iron permeable reactive barriers: Column experiments and numerical simulation. J. Contam. Hydrol. 2009, 103, 145–156. [Google Scholar]
  88. 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]
  89. 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]
  90. Tao, R.; Yang, H.; Cui, X.; Xiao, M.; Gatcha-Bandjun, N.; Kenmogne-Tchidjo, J.F.; Lufingo, M.; Konadu Amoah, B.; Tepong-Tsindé, R.; Ndé-Tchoupé, A.I.; et al. The suitability of hybrid Fe0/aggregate filtration systems for water treatment. Water 2022, 14, 260. [Google Scholar] [CrossRef]
  91. O’Hannesin, S.F.; Gillham, R.W. Long-term performance of an in situ “iron wall” for remediation of VOCs. Ground Water 1998, 36, 164–170. [Google Scholar] [CrossRef]
  92. Balasubramaniam, R.; Kumar, A.R.; Dillmann, P. Characterization of rust on ancient Indian iron. Curr. Sci. 2003, 85, 1546–1555. [Google Scholar]
  93. Caré, S.S.; Nguyen, Q.T.; l’Hostis, V.; Berthaud, Y. Mechanical properties of the rust layer induced by impressed current method in reinforced mortar. Cement Concrete Res. 2008, 38, 1079–1091. [Google Scholar] [CrossRef]
  94. 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]
  95. Ogawa, Y.; Ando, D.; Sutou, Y.; Koike, J. Effects of O2 and N2 flow rate on the electrical properties of Fe-O-N thin films. Mater. Trans. 2014, 55, 1606–1610. [Google Scholar] [CrossRef] [Green Version]
  96. Tepong-Tsindé, R. Designing and Piloting a household filter for the peri-urban population of Douala (Cameroon). Freiberg Online Geosci. 2021, 61, 1–90. [Google Scholar]
  97. Yang, H.; Hu, R.; Ruppert, H.; Noubactep, C. Modeling porosity loss in Fe0-based permeable reactive barriers with Faraday’s law. Sci. Rep. 2021, 11, 16998. [Google Scholar] [CrossRef]
  98. Hamdy, A.; Mostafa, M.; Nasr, M. Regression analysis and artificial intelligence for removal of methylene blue from aqueous solutions using nanoscale zero-valent iron. Int. J. Environ. Sci. Technol. 2019, 16, 357–372. [Google Scholar] [CrossRef]
  99. Sahu, N.; Rawat, S.; Singh, J.; Karri, R.R.; Lee, S.; Choi, J.S.; Koduru, J.R. Process optimization and modeling of methylene blue adsorption using zero-valent iron nanoparticles synthesized from sweet lime pulp. Appl. Sci. 2019, 9, 5112. [Google Scholar] [CrossRef] [Green Version]
  100. Tarekegn, M.M.; Balakrishnan, R.M.; Hiruy, A.M.; Dekebo, A.H. Removal of methylene blue dye using nano zerovalent iron, nanoclay and iron impregnated nanoclay—A comparative study. RSC Adv. 2021, 11, 30109. [Google Scholar] [CrossRef]
  101. Wang, Y.; Chen, T.; Zhang, X.; Mwamulima, T. Removal study of crystal violet and methylene blue from aqueous solution by activated carbon embedded zero-valent iron: Effect of reduction methods. Front. Environ. Sci. 2021, 631. [Google Scholar] [CrossRef]
  102. El-Monaem, A.; Eman, M.; Omer, A.M.; El-Subruiti, G.M.; Mohy-Eldin, M.S.; Eltaweil, A.S. Zero-valent iron supported-lemon derived biochar for ultra-fast adsorption of methylene blue. Biomass Conv. Bioref. 2022, 1–13. [Google Scholar] [CrossRef]
  103. Btatkeu-K, B.; Tchatchueng, J.; Noubactep, C.; Caré, S. Designing metallic iron based water filters: Light from methylene blue discoloration. J. Environ. Manage. 2016, 166, 567–573. [Google Scholar] [CrossRef] [PubMed]
  104. Miyajima, K. Optimizing the design of metallic iron filters for water treatment. Freiberg Online Geosci. 2012, 32, 1–60. [Google Scholar]
  105. Miyajima, K.; Noubactep, C. Effects of mixing granular iron with sand on the efficiency of methylene blue discoloration. Chem. Eng. J. 2012, 200, 433–438. [Google Scholar] [CrossRef] [Green Version]
  106. Mitchell, G.; Poole, P.; Segrove, H. Adsorption of methylene blue by high-silica sands. Nature 1955, 176, 1025–1026. [Google Scholar] [CrossRef]
  107. 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, ISBN-10: 4946428615; ISBN-13: 978-4946428616. [Google Scholar]
  108. Miyajima, K.; Noubactep, C. Characterizing the impact of sand addition on the efficiency of granular iron for water treatment. Chem. Eng. J. 2015, 262, 891–896. [Google Scholar] [CrossRef]
  109. Sikora, E.; Macdonald, D.D. The passivity of iron in the presence of ethylenediaminetetraacetic acid I. General electrochemical behavior. J. Electrochem. Soc. 2000, 147, 4087–4092. [Google Scholar] [CrossRef]
  110. Alcántara, J.; de la Fuente, D.; Chico, B.; Simancas, J.; Díaz, I.; Morcillo, M. Marine atmospheric corrosion of carbon steel: A review. Materials 2017, 10, 406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Ali, N.; Fulazzaky, M.A. The empirical prediction of weight change and corrosion rate of low-carbon steel. Heliyon 2020, 6, e05050. [Google Scholar] [CrossRef] [PubMed]
  112. Romanoff, M. Underground Corrosion; US Government Printing Office: Washingtown, DC, USA, 1957; Volume 579.
  113. Melchers, R.; Petersen, R. A reinterpation of the Romanoff NBS data for corrosion of steels in soils. Corrosion Eng. Sci. Technol. 2018, 53, 131–140. [Google Scholar] [CrossRef]
  114. Stefanoni, M.; Angst, U.; Elsener, B. Electrochemistry and capillary condensation theory reveal the mechanism of corrosion in dense porous media. Sci. Rep. 2018, 8, 7407. [Google Scholar] [CrossRef] [Green Version]
  115. Stefanoni, M.; Zhang, Z.; Angst, U.M.; Elsener, B. The kinetic competition between transport and oxidation of ferrous ions governs precipitation of corrosion products in carbonated concrete. RILEM Tech. Lett. 2018, 3, 8–16. [Google Scholar] [CrossRef]
  116. Painter, B.D. Optimisation of Permeable Reactive Barrier Systems for the Remediation of Contaminated Groundwater. Doctoral Dissertation, Lincoln University, Christchurch, New Zealand, 2005. [Google Scholar]
  117. Reynolds, G.W.; Hoff, J.T.; Gillham, R.W. Sampling bias caused by materials used to monitor halocarbons in groundwater. Environ. Sci. Technol. 1990, 24, 135–142. [Google Scholar] [CrossRef]
  118. Noubactep, C. Contaminant reduction at the surface of elemental iron: The end of a myth. Wiss. Mitt. 2006, 31, 173–179. [Google Scholar]
  119. Noubactep, C. Das Ende eines Mythos: Direkte Schadstoffreduktion durch elementares Eisen (Fe0) widerspricht drei Jahrhunderten Korrosionsforschung. TerraTech 2006, 50, 11–12. [Google Scholar]
  120. Popat, V.; Padhiyar, N. Kinetic study of Bechamp process for p-nitrotoluene reduction to p-toluidine. Int. J. Chem. Eng. Appl. 2013, 4, 401–405. [Google Scholar] [CrossRef] [Green Version]
  121. Bafghi, M.S.; Zakeri, A.; Ghasemi, Z.; Adeli, M. Reductive dissolution of manganese ore in sulfuric acid in the presence of iron metal. Hydrometallurgy 2008, 90, 207–212. [Google Scholar] [CrossRef]
  122. Gould, J.P. The kinetics of hexavalent chromium reduction by metallic iron. Water Res. 1982, 16, 871–877. [Google Scholar] [CrossRef]
  123. Anderson, M.A. Fundamental Aspects of Selenium Removal by Harza Process; Rep San Joaquin Valley Drainage Program; US Dep Interior: Sacramento, CA, USA, 1989.
  124. Khudenko, B.M. Feasibility evaluation of a novel method for destruction of organics. Water Sci. Technol. 1991, 23, 1873–1881. [Google Scholar] [CrossRef]
  125. Roberts, A.L.; Totten, L.A.; Arnold, W.A.; Burris, D.R.; Campbell, T.J. Reductive elimination of chlorinated ethylenes by zero-valent metals. Environ. Sci. Technol. 1996, 30, 2654–2659. [Google Scholar] [CrossRef]
  126. Weber, E.J. Iron-mediated reductive transformations: Investigation of reaction mechanism. Environ. Sci. Technol. 1996, 30, 716–719. [Google Scholar] [CrossRef]
  127. Fan, D.; Lan, Y.; Tratnyek, P.G.; Johnson, R.L.; Filip, J.; O’Carroll, D.M.; Nunez Garcia, A.; Agrawal, A. Sulfidation of iron-based materials: A review of processes and implications for water treatment and remediation. Environ. Sci. Technol. 2017, 51, 13070–13085. [Google Scholar] [CrossRef] [PubMed]
  128. Noubactep, C. On the operating mode of bimetallic systems for environmental remediation. J. Hazard. Mater. 2009, 164, 394–395. [Google Scholar] [CrossRef] [Green Version]
  129. Fan, D.; Lan, Y.; Tratnyek, P.G.; Johnson, R.L.; Filip, J.; O’Carroll, D.M.; Nunez Garcia, A.; Agrawal, A. 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]
  130. 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, 186–194. [Google Scholar] [CrossRef]
  131. Molin, J.; Valkenburg, J.; Seech, A.; Oesterreich, R.; Son, J. Long-term evaluation of an EHC injection permeable reactive barrier in a sulfate-rich, high-flow aquifer. Remediation 2019, 29, 17–29. [Google Scholar] [CrossRef]
  132. 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] [PubMed]
  133. 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]
  134. Noubactep, C. Predicting the Hydraulic Conductivity of Metallic Iron Filters: Modeling Gone Astray. Water 2016, 8, 162. [Google Scholar] [CrossRef] [Green Version]
  135. Yang, H.; Hu, R.; Ndé-Tchoupé, A.I.; Gwenzi, W.; Ruppert, H.; Noubactep, C. Designing the Next Generation of Fe0-Based Filters for Decentralized Safe Drinking Water Treatment: A Conceptual Framework. Processes 2020, 8, 745. [Google Scholar] [CrossRef]
  136. Wu, Q.; Zheng, C.; Zhang, J.; Zhang, F. Nitrate removal by a permeable reactive barrier of Fe0: A model-based evaluation. J. Earth Sci. 2017, 28, 447–456. [Google Scholar] [CrossRef]
  137. Noubactep, C. Flaws in the design of Fe(0)-based filtration systems? Chemosphere 2014, 117, 104–107. [Google Scholar] [CrossRef]
  138. Noubactep, C. Research on metallic iron for environmental remediation: Stopping growing sloppy science. Chemosphere 2016, 153, 528–530. [Google Scholar] [CrossRef]
  139. Noubactep, C.; Makota, S.; Bandyopadhyay, A. Rescuing Fe0 remediation research from its systemic flaws. Res. Rev. Insights 2017, 1, 1–8. [Google Scholar] [CrossRef] [Green Version]
  140. Falkenberg, R.I. Re-invent yourself! How demands for innovativeness reshape epistemic practices. Minerva 2021, 59, 423–444. [Google Scholar] [CrossRef]
  141. Baveye, P.C. Bypass and hyperbole in soil research: Worrisome practices critically reviewed through examples. Eur. J. Soil Sci. 2021, 72, 1–20. [Google Scholar] [CrossRef]
  142. Baveye, P.C. Bypass and hyperbole in soil research: A personal view on plausible causes and possible remedies. Eur. J. Soil Sci. 2021, 72, 21–28. [Google Scholar] [CrossRef]
  143. Daughton, C.G. The Matthew Effect and widely prescribed pharmaceuticals lacking environmental monitoring: Case study of an exposure-assessment vulnerability. Sci. Total Environ. 2014, 466, 315–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Gwenzi, W. The ‘thanato-resistome’-The funeral industry as a potential reservoir of antibiotic resistance: Early insights and perspectives. Sci. Total Environ. 2020, 749, 141120. [Google Scholar] [CrossRef] [PubMed]
  145. Noubactep, C. Metallic iron for environmental remediation: The long walk to evidence. Corros. Rev. 2013, 31, 51–59. [Google Scholar] [CrossRef] [Green Version]
  146. Noubactep, C. Metallic iron for water treatment: A critical review. Clean-Soil Air Water 2013, 41, 702–710. [Google Scholar] [CrossRef] [Green Version]
  147. Noubactep, C. Relevant reducing agents in remediation Fe0/H2O systems. Clean–Soil Air Water 2013, 41, 493–502. [Google Scholar] [CrossRef] [Green Version]
  148. Burgess-Jackson, K. Why I publish in “predatory” journals–and why you should, too. Philos. Int. J. 2020, 3, 1–11. [Google Scholar]
  149. Noubactep, C.; Schöner, A. Fe0-based alloys for environmental remediation: Thinking outside the box. J. Hazard. Mater. 2009, 165, 1210–1214. [Google Scholar] [CrossRef]
  150. Noubactep, C. No scientific debate in the zero-valent iron literature. Clean–Soil Air Water 2016, 44, 330–332. [Google Scholar] [CrossRef]
  151. Noubactep, C. Metallic iron for environmental remediation: How experts maintain a comfortable status quo. Fresenius Environ. Bull. 2018, 27, 1379–1393. [Google Scholar]
  152. Noubactep, C. The operating mode of Fe0/H2O systems: Hidden truth or repeated nonsense? Fresenius Environ. Bull. 2019, 28, 8328–8330. [Google Scholar]
  153. Dickerson, R.E.; Gray, H.B.; Haight, G.P., Jr. Chemical Principles, 3rd ed.; Benjamin/Cummings Inc.: London, UK; Amsterdam, The Netherlands, 1979; p. 944. [Google Scholar]
  154. Noubactep, C. The suitability of metallic iron for environmental remediation. Environ. Prog. Sustain. 2010, 29, 286–291. [Google Scholar] [CrossRef] [Green Version]
  155. Groysman, A. Corrosion for Everybody; Springer: Dordrecht, The Netherlands; Heidelberg, Germany; London, UK; New York, NY, USA, 2010; p. 368. [Google Scholar]
  156. 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]
  157. Rasmussen, M.; Savenije, H.; Thybo, H.; Bamber, J. EGU & Copernicus Report about Citation Stacking in the EGU Journals SE and SOIL. 2017. Available online: https://static2.egu.eu/media/filer_public/07/79/07798eae-e4e4-48f2-a9d0-6b8ce0110302/egu-copernicus-report-about-citation-stacking.pdf (accessed on 22 April 2022).
  158. Retraction Watch, 2017. An Editor in Chief was Caught Manipulating Citations. Now He’s Been Asked to Resign. Available online: http://retractionwatch.com/2017/03/21/editor-chief-caught-manipulating-citations-now-hes-asked-resign/ (accessed on 22 April 2022).
  159. Herteliu, C.; Ausloos, M.; Ileanu, B.V.; Rotundo, G.; Andrei, T. Quantitative and qualitative analysis of editor behavior through potentially coercive citations. Publications 2017, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  160. Chakraborty, J.; Pradhan, D.K.; Nandi, S. On the identification and analysis of citation pattern irregularities among journals. Expert Syst. 2021, 38, e12561. [Google Scholar] [CrossRef]
Figure 1. Schematic of the core/shell structure of the Fe0/H2O system. The following features are labeled: (i) (left) Fe0/FeCPs interface, (ii) FeCPs/H2O interface, (right) reduction of water (H+) by Fe0, and (iv) reduction of RCl by Fe2+ and H2.
Figure 1. Schematic of the core/shell structure of the Fe0/H2O system. The following features are labeled: (i) (left) Fe0/FeCPs interface, (ii) FeCPs/H2O interface, (right) reduction of water (H+) by Fe0, and (iv) reduction of RCl by Fe2+ and H2.
Water 14 01492 g001
Figure 2. A schematic of design to investigate the Fe0/H2O systems. A pure sand system can serve as reference, and no pure Fe0 layer should be considered. Only hybrid and mixed Fe0/aggregates (e.g., Fe0/sand) systems are worth investigating both in laboratory and column studies.
Figure 2. A schematic of design to investigate the Fe0/H2O systems. A pure sand system can serve as reference, and no pure Fe0 layer should be considered. Only hybrid and mixed Fe0/aggregates (e.g., Fe0/sand) systems are worth investigating both in laboratory and column studies.
Water 14 01492 g002
Table 1. Structure, formula, and some characteristics of metallic iron (Fe0) and its main corrosion products. Data are compiled from Balasubramaniam et al. [92] and Caré et al. [93]. bcc stands for body-centered cubic, and fcc stands for face-centered cubic. diron/doxide gives an indication of the extent of volumetric expansion, while the percent Fe gives an idea of the decrease in the electronic conductivity of oxides relative to Fe0.
Table 1. Structure, formula, and some characteristics of metallic iron (Fe0) and its main corrosion products. Data are compiled from Balasubramaniam et al. [92] and Caré et al. [93]. bcc stands for body-centered cubic, and fcc stands for face-centered cubic. diron/doxide gives an indication of the extent of volumetric expansion, while the percent Fe gives an idea of the decrease in the electronic conductivity of oxides relative to Fe0.
SubstanceFormulaStructureDensity (d)dmetal/doxideMWFeO
(-)(g/cm3)(-)(g/mol)(%)(%)
Iron metalFe0bcc7.861.00056100.00.0
WüstiteFeOfcc5.671.3867277.822.2
MagnetiteFe3O4Cubic5.181.51923272.427.6
Hematiteα-Fe2O3Trigonal5.261.49416070.030.0
Maghemiteγ-Fe2O3Cubic4.691.67616070.030.0
Goethiteα-FeOOHOrthorhombic4.281.8368962.937.1
Akaganeiteβ-FeOOHTetragonal3.552.2148962.937.1
Lepidocrociteγ-FeOOHOrthorhombic4.091.9228962.937.1
δ-FeOOHHexagonaln.a.n.a.8962.937.1
FeOOH (Gel)FeOOHCubicn.a.n.a.8962.937.1
FeII hydroxideFe(OH)2Trigonal3.402.3129062.237.8
FeIII hydroxideFe(OH)3Trigonal3.832.05210752.347.7
Table 2. Summary of some experimental conditions used for the experiments in the five articles. Experiments were conducted in batch using nano-size Fe0 (nZVI) modified in different ways (Table 3). “n.s.” stands for “not specified”, MB stands for methylene blue, and [MB] stands for its concentration, while “Nr.” is the number referencing individual articles in Table 3.
Table 2. Summary of some experimental conditions used for the experiments in the five articles. Experiments were conducted in batch using nano-size Fe0 (nZVI) modified in different ways (Table 3). “n.s.” stands for “not specified”, MB stands for methylene blue, and [MB] stands for its concentration, while “Nr.” is the number referencing individual articles in Table 3.
Nr.[MB]
(mg/L)
Fe0 TypeFe0 Dosage (g/L)pH ValueDurationVolume (mL)MixingSpeed (rpm)Ref.
1500–2000nZVI-LBC0.53–115 min20Quiescent0[102]
220–120nZVI and iron impregnated nanoclay.0.4–2.83–1320–140 min50Agitated80[100]
350–1000PAC-mZVI62–1072 h100Stirred120[101]
410–40nZVISLP4–122–10n.s.100Stirred100[99]
510–70nZVI2–142–1210–120 min250Stirred100–250[98]
Table 3. Summary of the use of MB and the key findings for each of the five selected articles used to illustrate the evidence that the lack of true collaboration has been the major obstacle to Fe0 remediation research progress.
Table 3. Summary of the use of MB and the key findings for each of the five selected articles used to illustrate the evidence that the lack of true collaboration has been the major obstacle to Fe0 remediation research progress.
Nr.MB UseKey Findings
1Model contaminant for the investigation of the adsorptive removal capacity of nano zero-valent supported-lemon-derived biochar (nZVI-LBC)NZVI-LBC composite has an excellent adsorptive affinity for MB, making it a potential candidate for the aqueous removal of cationic contaminants.
2Model contaminant for the investigation of the adsorptive removal capacity of nano zero-valent iron (nZVI) and iron impregnated nanoclayIron impregnated nanoclay comparatively has a better adsorptive removal capacity, making it a more suitable adsorbent for industrial applications of cationic dye removal.
3Model contaminant to investigate the adsorption mechanism of palm-activated carbon metallic iron (PAC-mZVI) particlesThe properties of PAC-mZVI and the structure of MB both have significant effects on the removal of MB.
4Model contaminant for the investigation of the adsorptive removal characteristics of nano zero-valent iron synthesized from sweet lime pulp waste (nZVISLP)nZVISLP has a great potential for MB adsorptive removal.
5Model contaminant to investigate the adsorptive removal capacity of nano zero-valent iron (nZVI)nZVI is an excellent absorbent for MB aqueous removal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiao, M.; Hu, R.; Ndé-Tchoupé, A.I.; Gwenzi, W.; Noubactep, C. Metallic Iron for Water Remediation: Plenty of Room for Collaboration and Convergence to Advance the Science. Water 2022, 14, 1492. https://0-doi-org.brum.beds.ac.uk/10.3390/w14091492

AMA Style

Xiao M, Hu R, Ndé-Tchoupé AI, Gwenzi W, Noubactep C. Metallic Iron for Water Remediation: Plenty of Room for Collaboration and Convergence to Advance the Science. Water. 2022; 14(9):1492. https://0-doi-org.brum.beds.ac.uk/10.3390/w14091492

Chicago/Turabian Style

Xiao, Minhui, Rui Hu, Arnaud Igor Ndé-Tchoupé, Willis Gwenzi, and Chicgoua Noubactep. 2022. "Metallic Iron for Water Remediation: Plenty of Room for Collaboration and Convergence to Advance the Science" Water 14, no. 9: 1492. https://0-doi-org.brum.beds.ac.uk/10.3390/w14091492

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