Next Article in Journal
Effects of Temperature on the Characteristics of Nitrogen Removal and Microbial Community in Post Solid-Phase Denitrification Biofilter Process
Previous Article in Journal
Microbiological Safety of Street-Vended Beverages in Chow Kit, Kuala Lumpur
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Redirecting Research on Fe0 for Environmental Remediation: The Search for Synergy

1
School of Earth Science and Engineering, Hohai University, Fo Cheng Xi Road 8, Nanjing 211100, China
2
Department of Applied Geology, Universität Göttingen, Goldschmidtstraße 3, D-37077 Göttingen, Germany
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2019, 16(22), 4465; https://doi.org/10.3390/ijerph16224465
Submission received: 22 October 2019 / Revised: 5 November 2019 / Accepted: 11 November 2019 / Published: 13 November 2019
(This article belongs to the Section Environmental Science and Engineering)

Abstract

:
A survey of the literature on using metallic iron (Fe0) for environmental remediation suggests that the time is ripe to center research on the basic relationship between iron corrosion and contaminant removal. This communication identifies the main problem, which is based on the consideration that contaminant reductive transformation is the cathodic reaction of iron oxidative dissolution. Properly considering the inherent complexities of the Fe0/H2O system will favor an appropriate research design that will enable more efficient and sustainable remediation systems. Successful applications of Fe0/H2O systems require the collective consideration of progress achieved in understanding these systems. More efforts should be made to decipher the long-term kinetics of iron corrosion, so as to provide better approaches to accurately predict the performance of the next generation Fe0-based water treatment systems.

1. Introduction

The broad research community working on “metallic iron (Fe0) for environmental remediation” largely considers Fe0 to be an environmental reducing agent. This idea has been repeated in the introduction and discussion sections of hundreds of research and review articles [1,2,3]. The fundamental flaws identified have been hindering the progress of knowledge on designing efficient and sustainable Fe0-based water treatment systems [2,4].
The idea that Fe0 reduces selected contaminants under operational conditions was introduced in the peer-reviewed literature in 1994 [5] and was a controversial discussion topic for the following five years [6,7,8,9,10,11]. For example, Warren et al. [7] used several elemental metals (Al0, Fe0, Sn0, and Zn0) to treat carbon tetrachloride (CT) and reported “unresolved mechanistic complexities”. CT was one of the organic compounds used in the seminal 1994 work [5]. On the other hand, Weber [9] used 4-aminoazobenzene as a probe molecule to “confirm” that “reductive transformation by Fe0 is a surface-mediated process”. According to Weber [9], electron transfer at the Fe0/H2O interface can be facilitated by appropriate “electron mediators”. Similar controversies have been reported regarding the removal mechanism of several inorganic contaminants [6,10,11]. However, in 1998, a “broad consensus” on reductive transformations as main contaminant removal mechanisms in Fe0/H2O systems was reached [12]. Despite having been proven wrong several times, the reductive transformation concept is still prevalent [13,14,15,16,17,18,19].
Reports disproving the reductive transformation concept are numerous and scattered in the scientific literature. For example, Noubactep et al. [14] demonstrated that there is no uranium removal without “free” precipitating iron corrosion products (FePCs). They used pyrite (a pH-shifting agent) and MnO2 (a Fe2+-consuming agent) to delay the “free” precipitation of FeCPs in the vicinity of Fe0. Ghauch et al. [16] and Gheju and Balcu [19] similarly used MnO2 to delay the removal of diclofenac and chromium, respectively. On the other hand, Jiao et al. [15] clearly demonstrated that observed CT-reductive transformation in Fe0/H2O systems cannot be mediated by the Fe0 surface. Finally, Ebelle et al. [18] systematically demonstrated the non-suitability of the experimental design of Matheson and Tratnyek [5] to achieve the often questioned but still largely accepted conclusions.
The named arguments [14,15,16,17] and the demonstration by Ebelle et al. [18] suggest that the popular view that Fe0 is an environmental reducing agent should be urgently revisited. Such a profound correction will offer new research perspectives towards the realization of the full potential of Fe0-based treatment systems [20]. The perspectives include: (i) a science-based assessment of the limits of Fe0-based remediation systems (advantages, disadvantages, and sustainability), (ii) the design of reliable, efficient, and sustainable Fe0-based remediation systems, and (iii) the design of programs to educate designers and operators of such systems. Proper capacity building of the treatment processes, the expected performance, and the required operational conditions would enable well-designed, efficient, and sustainable systems. This paper aimed to pave the way to a bright future of Fe0-based remediation technology by presenting some key aspects that have been either overseen or not particularly well considered in the past. The discussion is limited to the chemistry of aqueous iron corrosion.

2. Method

The approach used herein considers Fe0 as a generator of FeCPs, which are (i) contaminant scavengers, (ii) stand-alone reducing agents, and (iii) fouling agents in filtration systems. Readers that are not familiar with this decade-old-concept are referred to the corresponding literature [21,22,23,24,25,26,27]. This approach implies that the Fe0 reactivity yielding contaminant-reductive transformation is not an intrinsic property of Fe0 (electrons from Fe0). Contaminant reduction is associated with primary (FeII, H/H2) and secondary FeCPs (e.g., Fe3O4, green rust) as iron is corroded by water (the solvent) [4,28,29,30]. The existing literature was evaluated as to the contribution it has made to disregarding the content of the concept and the concept itself (from 2007 on). The next sections discuss how the confusion has been kept alive within the Fe0 research community in the past three decades.

2.1. A Questionable Literature Review

Fe0 remediation technology is more than 170 years old [31]. Ideally, the further development of this technology should be limited to assembling performance data on operating Fe0/H2O systems in order to improve the design of new systems. However, a critical consideration of the history of the Fe0 remediation technology reveals that no systematic approach has ever been followed in designing systems [30,32,33]. In the past three decades, an exceedingly pragmatic approach has been followed [3,34,35].
The three steps toward establishment of any technology are (i) a science-based feasibility analysis, (ii) laboratory scale treatability studies of derived designs, and (iii) pilot plant testing. This systematic approach is considered to have virtually been followed while establishing the Fe0 technology [36,37,38]. However, the very first step cannot be considered to have occurred for at least three reasons: (i) the literature review was very limited, as several historical aspects were overlooked [2,31,33]; (ii) the nature of Fe0 as a long-term source of iron oxides and hydroxides (contaminant scavengers) [39,40] was not really realized, and instead its “electron-producing” property was favored [3,5,9]; and (iii) the volumetric, expansive nature of iron corrosion was overlooked [41,42]. This initial approximate problem analysis is the major problem of the Fe0 technology and must be urgently fixed, especially since Fe0/H2O systems have the potential to secure universal safe drinking water provision in the developing world within one decade [19,20]. Properly considering that Fe0 is not a reducing agent under environmental conditions would clarify all discrepancies reported in the literature [4,33], and thus, aid to shape research for the design of better systems.

2.2. The Chemistry of the Fe0/H2O System

Once immersed in water, a reactive Fe0 specimen starts its oxidative dissolution to form FeII species and H2 (Equation (1)). Both primary iron corrosion products (FeII and H2) are stand-alone reducing agents. In the presence of oxygen and/or other oxidizing agents (e.g., MnO2), FeII is further oxidized to FeIII (Equation (2)) [14,16,19]. At pH values > 4.5, FeII and FeIII hydroxides have low solubility; they polymerize and precipitate to various types of oxides and hydroxides (Equation (3)).
Fe0 + 2 H+ ⇒ Fe2+ + H2
4 Fe2+ + O2 + 2 H+ ⇒ 4 Fe3+ + 2 OH
Fe3+ + Fe2+ + n H2O 4 Fe2+ + O2 + 2 H+ ⇒ 4 Fe3+ + 2 OH FeII, FeIII, and mixed oxyhydroxides
Fe0 + RX + H+ ⇒ Fe2+ + RH + X
where X represents a halogen ion (e.g., Cl).
In other words, a Fe0/H2O system is a dynamic system containing potential reducing agents (Fe0, FeII, and H2) and adsorbents (oxyhydroxides). It is thus not surprising that adsorption and chemical reduction occur in such systems [25,28,29]. It is rather surprising that the chemical reduction observed is considered to be the cathodic process coupled with the Fe0 oxidative dissolution (Equation (1)) and described by reactions equivalent to Equation (4) [4,32,33,43].
Equation (4) implies that Fe0 is oxidized by RX in an electrochemical process. This corresponds to the view that Fe0 is a reducing agent, and constitutes the heart of discrepancy [4]. The major prerequisite for this mechanism is that the oxide scale on Fe0 is electronically conductive [43,44]. Such a conductive layer has not been reported under environmental conditions. Accordingly, quantitative reduction by electrons from the metal is not likely to occur [4,32,33,43]. The well-documented reduction is thus mediated by other available reducing agents. Consequently, “reduction in the presence of Fe0” should not be considered to be “reduction by Fe0” [4]. An electrochemical mechanism for O2 was already out ruled in the 1980s, as summarized by Stratmann and Müller [45]. Accordingly, (i) Fe0 is oxidized by water (Equation (1)) (electrochemical reaction), and (ii) contaminants and O2 are reduced by FeII, H2, and/or other species like green rust (chemical reaction).
During the precipitation of oxyhydroxides (Equation (3)), dissolved species are mechanically entrapped in their matrix (co-precipitation). The volume of each corrosion product (oxide) is larger than that of the parent metallic iron (Voxide > Viron) [41,42,46]. This evidence implies that in a filter bed containing Fe0, iron corrosion pursued for contaminant removal is the first cause of porosity reduction and thus permeability loss [4,30,33,35,42]. Properly considering the point outlined here will soon unify the research community on the evidence that a new orientation of research on Fe0 is long overdue. It should be explicitly stated that the notion of electron efficiency, defined as the ratio of electrons transferred to the target contaminants, is not acceptable, as it considers electrochemical reduction to be a relevant reduction path [4,47]. The presentation herein has insisted that reducing electrons are from H2 (two electrons per mole of Fe0) and/or from FeII species (one electron per mole of Fe0). Accordingly, a good starting position is three electrons per mole of Fe0 (for pure abiotic reduction). H2 escaping from the system is a fraction of the non-utilized amount because H2 evolution is not quantitative at pH > 4.5 [48].

2.3. The Root Origin of Confusion

A profound analysis of the Fe0/H2O system (Section 2.2) revealed that the root origin of confusion within the Fe0 remediation research community is nothing but the understanding of the corrosion reaction which can be generally given by Equation (4) [49,50]:
metal + oxidizing agent ⇒ oxidized metal + reducing agent
For iron, Equation (5) reads:
Fe0 + oxidizing agent ⇒ FeII/FeIII species + reducing agent
The questions arises, which relevant oxidizing agents are available for Fe0 corrosion under natural conditions? The prevailing approach regards Fe0 as a reducing agent with the electrode potential −0.44 V. Accordingly, Fe0 should reduce all dissolved species with a higher electrode potential, including O2 and contaminants, (E0 > −0.44 V). The research community has been trying to address this issue on a case-by-case basis for three decades already, with very limited success [1,3,19,20,36,38,50]. However, this concept does not consider the evidence that in aqueous solutions at pH > 4.5, the Fe0 surface is constantly covered with an electrically non-conductive oxide scale which hinders any quantitative reaction after Equation (4) [30,50]. A clear experimental proof against this concept has been the repeated observation of a time lag between the start of the experiment and the start of the reducing reaction [33,35]. If the electrochemical reduction (electrons from Fe0) was a relevant reaction mechanism, it would have been quantitative when the Fe0 surface is still free from FeCPs [30,50]. The lag time indicates that even acting reducing agents are generated in situ.
In contrast, the approach presented herein considers water (H2O, the solvent) to be the sole relevant oxidizing agent for Fe0 (Equation (1)). The reason is that the natural oxide scale acts as a diffusion barrier to all dissolved species. It follows that any reasoning using Equation (4) (E0 > −0.44 V) has not properly considered the physicochemistry of the Fe0/H2O system. Thus, trying to consider the stoichiometry of equations similar to Equation (4) in models is simply false. The most obvious damage of such an approach has been the discussion on which species fill the initial porosity of Fe0-based filtration systems: oxide scale (FeCPs), H2 accumulation, or foreign precipitates (e.g., CaCO3, FeCO3). This controversy was related best by Henderson and Demond [37,51,52]. These two authors wrote one of the best review articles on “remediation with Fe0” [37], confirming permeability loss as one of the major problems of Fe0 filters. Their efforts to solve the named problem went through the stoichiometry of reactions similar to Equation (4) [51] and resulted in their recommending FeS-based systems as less-clogging alternatives [52]. However, this approach cannot properly account for H2 evolution, as each mole of oxidized Fe0 generates one mole of H2 occupying a volume more than 31,000 times larger than the volume of the parent Fe0. For illustration, Equation (1) has been re-written to show the changes in the number of moles of Fe0 and FeCPs between the start of the operation (t0 = 0) and Fe0 complete depletion (t):
Fe 0 + 2   H 2 O Fe ( OH ) 2 + H 2 t 0 = 0 n 0 x 0 0 t > t 0 n 0   ( 1 α ) x α   n 0 α   n 0 t 0 x n 0 n 0 Volume 1 - 2.1   to   6.4 > 3100
The values of x, x’, and x” are not relevant for the discussion, as water is the solvent.
The factor 31,000 arises from the evidence that one mole of Fe0iron = 7.84 g cm3) occupying a volume of 7.14 cm3 produces one mole of H2 occupying a volume of 22,400 cm3 (assuming ideal gas behavior). Accordingly, properly reasoning with the changes in volume accompanying aqueous iron corrosion implies that H2 must escape or be used for other processes for the system to be sustainable. The mean feature of the reasoning based on Equation (6) is that modeling porosity loss should have started by considering pore filling by expansive FeCPs [42]. Pore filling by the process of iron corrosion has only recently been demonstrated by X-ray computed tomography of systems containing Fe0 and deionized water [48,53]. Again, with a proper system analysis, this evidence would have been considered in the early phase of the technology development. Instead, it is still under discussion whether mixing Fe0 with non-expansive aggregates (e.g., gravel, pumice, sand) is suitable for efficient systems [48,53].

2.4. Is Multidisciplinarity the Problem?

The subtlety and scientific diversity of the physical and chemical processes involved in water treatment using Fe0-based systems is demonstrated herein. The application of Fe0 materials in water treatment and environmental remediation (“putting corrosion to use” [54]) is, in essence, an engineering task. However, the present work has demonstrated that, despite three decades of intensive research, scientists have not yet established the science of the Fe0/H2O system. This common effort should be derived from the available basics of corrosion, electrochemistry, and surface chemistry of metals [30,43,49]. While the value of multidisciplinarity had been widely recognized in establishing the science of aqueous iron corrosion [43,44,49], the behavior of active scientists on remediation with Fe0 seems to ultimately preclude results.
The current literature on remediation with Fe0 can be regarded as a collection of independent research and review articles, with the term “zero-valent iron” in common. It is difficult for a critical reader to identify signs of common effort even from some critical review articles. As an example, while the intrinsic reactivity of Fe0 materials has been recognized as a crucial design parameter in the early phase of technology development [55,56], only a few contaminant-independent tools have been presented and none of them have really been adopted [57,58,59,60]. This means that researchers have not even tried to characterize granular materials (mm in size) before manufacturing granular bimetallics and nano-scale materials [30,50]. On the contrary, reactivity testing is continued with species as toxic as 99TcVII [61].
The same pragmatic trend can be observed in all other aspects of efforts to enhance the efficiency of Fe0-based systems, of which two are discussed here in some detail: (i) admixing Fe0 with pyrite (FeS2) or other FeS minerals, and (ii) adding complexing agents to the system. Adding FeS2 to a Fe0/H2O system corresponds to a pH shift to lower values. In such a system, the pH value is controlled by two antagonistic processes: FeS2 dissolution, lowering the pH, and Fe0 corrosion, enhancing the pH. Because FeS2 dissolution is kinetically more favorable, the pH of the system first decreases and then subsequently progressively increases to a final value [14]. If FeS2 dissolution initially shifted the pH to values lower that 4.5, quantitative contaminant removal will be observed only after the system’s pH increases to values above 4.5 [14,62]. This observation has been documented the best in quiescent batch experiments. However, most available works have varied the initial pH values and the FeS2 mass loading and reported controversial observations, as summarized by Du et al. [63]. Moreover, authors have generally not monitored the final pH values and the experimental vessels were vigorously shaken or stirred to enable suspension of Fe0 and FeS2, while FeS2 aims to support electron transfer from Fe0 thanks to its semi-conductive properties.
As concerns complexing agents, they replace water molecules in the hydration sphere of a dissolved metal ion (Fe2+ and/or Fe3+) and thus increase Fe solubility [49]. Since the seminal work of Matheson and Tratnyek [5], various complexing agents have been commonly used in investigating the mechanism of contaminant removal. Tested ligands include ethylenediaminetetraacetic acid (EDTA) and 1,10-Phenanthroline (Phen). These ligands are used to avoid or delay the passivation of the Fe0 surface. However, it has been well-documented that individual ligands modify the redox properties of the two involved redox systems (FeII/Fe0 and FeIII/FeII) differently [64]. For example, Phen stabilizes FeII while EDTA stabilizes FeIII. It follows that the addition of EDTA favors reduction by FeII, a reaction that could never quantitatively occur in the absence of EDTA [49,64,65,66]. In other words, in trying to keep the Fe0 surface free, bias is introduced. Again, investigated systems have been shaken or stirred, but the extent to which the complexing agents are consumed have not considered in discussions of the results.
The enumerated mistakes suggest that multidisciplinary study has been misinterpreted in the Fe0 remediation literature for the last 30 years. Exploiting aqueous iron corrosion for water treatment was known before the natures of many pollutants were established [2,31,39]. Thus, successfully using Fe0 to remove contaminants from water does not transform any scientist into an “iron corrosion expert”. Corrosion science is a stand-alone branch of science and at least some extra effort is required to become familiar with the basics of the highly complex process of iron corrosion [49]. Having neglected this key evidence, researchers have been hopelessly searching for common underlying mechanisms for reactions in Fe0/H2O systems “that provide a confidence for design that is non-site-specific” [67]. However, the treatability experiments have been ill-designed (shaken or stirred) and interpreted with incorrect descriptors such as the kSA concept [22,23,24,25].
The authors of the present communication suggest that the past three decades have presented a significant body of data confirming the suitability of Fe0-based systems for the remediation for a large number of pollutants. If relevant data on the iron corrosion rate were made available, efficient remediation systems could be designed using a concept presented in 1991 by Boris M. Khudenko [43] and long-term column experiments for filtration systems [68,69]. Khudenko’s concept [43] has been largely ignored in the scientific literature while contra-intuitive views have been massively published. Khudenko’s concept was recently discovered and proven in line with a concept Noubactep and colleagues have presented more than a decade ago [21,22]. Noubactep’s concept has been similarly ignored by active researchers [50,70,71]. The fact that many researchers work with a falsified view should be of concern to the whole scientific community. This is a problem which cannot be resolved by a few individuals [28,29,71]. This communication was written to make the problem better known to the research community.

3. Conclusions

Using Fe0/H2O systems is a suitable alternative to other methods of water treatment and environmental remediation. Testing their suitability has been unfortunately hampered by several flaws, including (i) incomplete literature reviews; (ii) inappropriate system analysis, in particular overlooking the expansive nature of aqueous iron corrosion; (iii) inappropriate testing conditions (shaking and stirring the vessels); and, of course, (iv) incorrect interpretation of good experimental observations. Because contaminant removal and contaminant reduction have generally been randomly interchanged, even species like aromatic amines, which have been claimed to require microbiological treatment after treatment in Fe0/H2O systems, can be quantitatively removed upon proper design, as reduction is not a relevant removal mechanism. The literature has presented many examples of non-reducible species (including methylene blue) that have been quantitatively removed in Fe0/H2O systems [72,73,74].
One constant myth has been an argument based on the idea of the Fe0 surface as a site accessible or not to contaminants. At the pH range of relevance, quantitative contaminant diffusive transport to the Fe0 surface is impossible, especially since there is no driving force. H2 and Fe2+ extraction from the Fe0 surface (kinetic energy) acts against the transport of any species to the Fe0 surface. For a better understanding of the operating mode of Fe0/H2O systems, long-term experiments under near-nature conditions should be performed to characterize the time-dependent kinetics of iron corrosion and their impact of the efficiency of the system. Hence, both the induction period (lag time prior to contaminant reduction) and the decrease in the corrosion rate will be better understood and used to design more efficient and sustainable Fe0-based systems. All that is needed is a truly collaborative research effort, respecting the basic scientific principles of research and publication.

Author Contributions

R.H. and C.N. contributed equally to manuscript compilation and revisions.

Funding

This work is supported by the Ministry of Science and Technology of China through the Program “Research on Mechanism of Groundwater Exploitation and Seawater Intrusion in Coastal Areas” (Project Code 20165037412) and by the Ministry of Education of China through “the Fundamental Research Funds for the Central Universities” (Project Code: 2015B29314).

Acknowledgments

The manuscript was improved thanks to the insightful comments of anonymous reviewers from International Journal of Environmental Research and Public Health. 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. 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]
  2. 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] [PubMed]
  3. Xin, J.; Tang, F.; Yan, J.; La, C.; Zheng, X.; Liu, W. Investigating the efficiency of microscale zero valent iron-based in situ reactive zone (mZVI-IRZ) for TCE removal in fresh and saline groundwater. Sci. Total Environ. 2018, 626, 638–649. [Google Scholar] [CrossRef] [PubMed]
  4. Noubactep, C. Metallic iron for environmental remediation: A review of reviews. Water Res. 2015, 85, 114–123. [Google Scholar] [CrossRef]
  5. Matheson, L.J.; Tratnyek, P.G. Reductive dehalogenation of chlorinated methanes by iron metal. Environ. Sci. Technol. 1994, 28, 2045–2053. [Google Scholar] [CrossRef]
  6. Cantrell, K.J.; Kaplan, D.I.; Wietsma, T.W. Zero-valent iron for the in situ remediation of selected metals in groundwater. J. Hazard. Mater. 1995, 42, 201–212. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. Weber, E.J. Iron-mediated reductive transformations: Investigation of reaction mechanism. Environ. Sci. Technol. 1996, 30, 716–719. [Google Scholar] [CrossRef]
  10. Fiedor, J.N.; Bostick, W.D.; Jarabek, R.J.; Farrel, 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]
  11. Gu, B.; Liang, L.; Dickey, M.J.; Yin, X.; Dai, S. Reductive precipitation of uranium (VI) by zero-valent iron. Environ. Sci. Technol. 1998, 32, 3366–3373. [Google Scholar] [CrossRef]
  12. 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]
  13. 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]
  14. 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]
  15. 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]
  16. 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]
  17. 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]
  18. 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]
  19. 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]
  20. 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]
  21. 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]
  22. 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]
  23. Noubactep, C. On the operating mode of bimetallic systems for environmental remediation. J. Hazard. Mater. 2009, 164, 394–395. [Google Scholar] [CrossRef] [PubMed]
  24. Noubactep, C. On the validity of specific rate constants (kSA) in Fe0/H2O systems. J. Hazard. Mater. 2009, 164, 835–837. [Google Scholar] [CrossRef] [PubMed]
  25. Noubactep, C. An analysis of the evolution of reactive species in Fe0/H2O systems. J. Hazard. Mater. 2009, 168, 1626–1631. [Google Scholar] [CrossRef] [PubMed]
  26. Noubactep, C. Elemental metals for environmental remediation: Learning from cementation process. J. Hazard. Mater. 2010, 181, 1170–1174. [Google Scholar] [CrossRef]
  27. Noubactep, C.; Caré, S. On nanoscale metallic iron for groundwater remediation. J. Hazard. Mater. 2010, 182, 923–927. [Google Scholar] [CrossRef]
  28. Ghauch, A. Iron-based metallic systems: An excellent choice for sustainable water treatment. Freiberg Online Geosci. 2015, 32, 1–80. [Google Scholar]
  29. Gheju, M. Progress in understanding the mechanism of CrVI Removal in Fe0-based filtration systems. Water 2018, 10, 651. [Google Scholar] [CrossRef]
  30. Hu, R.; Gwenzi, G.; Sipowo, R.; Noubactep, C. Water treatment using metallic iron: A tutorial review. Processes 2019, 7, 622. [Google Scholar] [CrossRef]
  31. Devonshire, E. The purification of water by means of metallic iron. J. Frankl. Inst. 1890, 129, 449–461. [Google Scholar] [CrossRef]
  32. Noubactep, C. Research on metallic iron for environmental remediation: Stopping growing sloppy science. Chemosphere 2016, 153, 528–530. [Google Scholar] [CrossRef] [PubMed]
  33. 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]
  34. Warner, S.D.; Sorel, D. Ten years of permeable reactive barriers: Lessons learned and future expectations. In Chlorinated Solvent and DNAPL Remediation: Innovative Strategies for Subsurface Cleanup; Henry, S.M., Warner, S.D., Eds.; Ser. 837; American Chemical Society: Washington, DC, USA, 2003; pp. 36–50, Symp. [Google Scholar]
  35. 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]
  36. Richardson, J.P.; Nicklow, J.W. In situ permeable reactive barriers for groundwater. Soil Sediment Contam. 2002, 11, 241–268. [Google Scholar] [CrossRef]
  37. 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]
  38. Gillham, R.W. Development of the granular iron permeable reactive barrier technology (good science or good fortune). In Advances in Environmental Geotechnics, Proceedings of the International Symposium on Geoenvironmental Engineering, Hangzhou, China, 8–10 September 2007; Chen, Y., Tang, X., Zhan, L., Eds.; Springer: Berlin/Heidelberg, Germany; London, UK, 2008; pp. 5–15. [Google Scholar]
  39. Lauderdale, R.A.; Emmons, A.H. A method for decontaminating small volumes of radioactive water. J. Am. Water Work. Assoc. 1951, 43, 327–331. [Google Scholar] [CrossRef]
  40. 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.
  41. Pilling, N.B.; Bedworth, R.E. The oxidation of metals at high temperatures. J. Inst. Met. 1923, 29, 529–591. [Google Scholar]
  42. 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]
  43. Khudenko, B.M. Feasibility evaluation of a novel method for destruction of organics. Water Sci. Technol. 1991, 23, 1873–1881. [Google Scholar] [CrossRef]
  44. Inzelt, G. Crossing the bridge between thermodynamics and electrochemistry. From the potential of the cell reaction to the electrode potential. Chemtexts 2014, 1, 2. [Google Scholar] [CrossRef] [Green Version]
  45. Stratmann, M.; Müller, J. The mechanism of the oxygen reduction on rust-covered metal substrates. Corros. Sci. 1994, 36, 327–359. [Google Scholar] [CrossRef]
  46. Zhao, Y.; Yu, J.; Jin, W. Damage analysis and cracking model of reinforced concrete structures with rebar corrosion. Corros. Sci. 2011, 53, 3388–3397. [Google Scholar] [CrossRef]
  47. 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]
  48. 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]
  49. Landolt, D. Corrosion and Surface Chemistry of Metals, 1st ed.; EPFL Press: Lausanne, Switzerland, 2007; p. 615. [Google Scholar]
  50. 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]
  51. 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]
  52. Hendeson, A.D.; Demond, A.H. Permeability of iron sulfide (FeS)-based materials for groundwater remediation. Water Res. 2013, 47, 1267–1276. [Google Scholar] [CrossRef]
  53. Santisukkasaem, U.; Das, D.B. A non-dimensional analysis of permeability loss in zero-valent iron permeable reactive barrier (PRB). Transp. Porous Media 2019, 126, 139–159. [Google Scholar] [CrossRef] [Green Version]
  54. Tratneyk, P.G. Putting corrosion to use: Remediating contaminated groundwater with zero-valent metals. Chem. Ind. 1996, 499–503. [Google Scholar]
  55. 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]
  56. Miehr, R.; Tratnyek, G.P.; 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] [PubMed]
  57. 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]
  58. Kim, H.; Yang, H.; Kim, J. Standardization of the reducing power of zerovalent iron using iodine. Environ. Lett. 2014, 49, 514–523. [Google Scholar]
  59. 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]
  60. 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]
  61. Boglaienko, D.; Emerson, H.P.; Katsenovich, Y.P.; Levitskaia, T.G. Comparative analysis of ZVI materials for reductive separation of 99Tc(VII) from aqueous waste streams. J. Hazard. Mater. 2019, 380, 120836. [Google Scholar] [CrossRef]
  62. 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]
  63. 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]
  64. Rizvi, M.A.; Syed, R.M.; Khan, B. Complexation effect on redox potential of iron(III)−iron(II) couple: A simple potentiometric experiment. J. Chem. Educ. 2011, 88, 220–222. [Google Scholar] [CrossRef]
  65. Naka, D.; Kim, D.; Strathmann, T.J. Abiotic reduction of nitroaromatic compounds by aqueous iron(II)-catechol complexes. Environ. Sci. Technol. 2006, 40, 3006–3012. [Google Scholar] [CrossRef]
  66. Rizvi, M.A. Complexation modulated redox behavior of transition metal systems. Russ. J. Gen. Chem. 2015, 85, 959–973. [Google Scholar] [CrossRef]
  67. 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] [PubMed]
  68. 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]
  69. Noubactep, C. Predicting the hydraulic conductivity of metallic iron filters: Modeling gone astray. Water 2016, 8, 162. [Google Scholar] [CrossRef] [Green Version]
  70. Noubactep, C. No scientific debate in the zero-valent iron literature. CLEAN Soil Air Water 2016, 44, 330–332. [Google Scholar] [CrossRef]
  71. Noubactep, C.; Makota, S.; Bandyopadhyay, A. Rescuing Fe0 remediation research from its systemic flaws. Res. Rev. Insights 2017. [Google Scholar] [CrossRef] [Green Version]
  72. Lackovic, J.A.; Nikolaidis, N.P.; Dobbs, G.M. Inorganic arsenic removal by zero-valent iron. Environ. Eng. Sci. 2000, 17, 29–39. [Google Scholar] [CrossRef]
  73. Jia, Y.; Aagaard, P.; Breedveld, G.D. Sorption of triazoles to soil and iron minerals. Chemosphere 2007, 67, 250–258. [Google Scholar] [CrossRef]
  74. Frost, R.L.; Xi, Y.; He, H. Synthesis, characterization of palygorskite supported zero-valent iron and its application for methylene blue adsorption. J. Colloid Interface Sci. 2010, 341, 153–161. [Google Scholar] [CrossRef] [Green Version]

Share and Cite

MDPI and ACS Style

Hu, R.; Noubactep, C. Redirecting Research on Fe0 for Environmental Remediation: The Search for Synergy. Int. J. Environ. Res. Public Health 2019, 16, 4465. https://doi.org/10.3390/ijerph16224465

AMA Style

Hu R, Noubactep C. Redirecting Research on Fe0 for Environmental Remediation: The Search for Synergy. International Journal of Environmental Research and Public Health. 2019; 16(22):4465. https://doi.org/10.3390/ijerph16224465

Chicago/Turabian Style

Hu, Rui, and Chicgoua Noubactep. 2019. "Redirecting Research on Fe0 for Environmental Remediation: The Search for Synergy" International Journal of Environmental Research and Public Health 16, no. 22: 4465. https://doi.org/10.3390/ijerph16224465

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