Metallic iron for water treatment: A knowledge system challenges mainstream science 1 2 3 4 5 6 7 8 Chicgoua Noubactep Angewandte Geologie, Universität Göttingen, Goldschmidtstraße 3, D - 37077 Göttingen, Germany. Kultur und Nachhaltige Entwicklung CDD e.V., Postfach 1502, D - 37005 Göttingen, Germany. 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 Tel.: +49 551 39 3191, Fax.: +49 551 399379, e-mail: cnoubac@gwdg.de Abstract A knowledge system (KS) is a knowledge that is unique to a given group of persons. This form of knowledge may have a local or natural origin and is linked to the community that has produced it. On the contrary, the core of mainstream science (MS) is the desire to profoundly understand processes, through sequential studies such as hypothesis formulation, experiment and prediction. Thus, KS is communitarian and MS is universal. KS can be understood and rendered universal through MS. In general, a process discovery (know-how) may be intuitive, accidental, conjectural or inspirational but outcomes should be predictable and repeatable as soon as the know-why is achieved by MS. This paper argues that the technology of using metallic iron for water treatment has all the characteristics of a KS and that promoters of this technology have deliberately rejected scientific arguments leading to the know-why of the fortuitous discovery. Consequently, the technology has developed into an impasse where controversial discoveries are reported on all relevant aspects. It is concluded that the integrity of science in endangered by this communitarian behaviour. Keywords: False premise, Knowledge system, Mainstream science, Peer review, Zerovalent iron. 1 1 Introduction 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 The use of elemental iron (Fe0) for water treatment has attracted much attention thanks to its great potential for removing several classes of substances from the aqueous phase [1-9]. Fe0 has been proven the most efficient material for subsurface reactive permeable barriers (reactive walls) [6, 10, 11]. Actually, there are about 180 Fe0 reactive walls installed worldwide [6]. The large majority of them been meeting their design goals. The use of Fe0 in reactive walls was discovered around 1990 as summarized in ref. [6]. Gillham and his colleagues [12] fortuitously found that trichloroethylene was reduced in a steel canister. This result was reproducible. Subsequent “laboratory testing and verification" experiments were performed [13-15]. To understand the importance of this discovery, one should consider that chlorinated organic compounds are more toxic and less biodegradable than their non-chlorinated reduction products. Accordingly, reducing chlorinated organic compounds is an ideal way to make them less toxic and more biodegradable. The discovery of Gillham et al. [12] coincided with the active search of appropriate materials for reactive walls after the concept for groundwater remediation presented by McMurty and Elton [16]. From 1990 on, Fe0 has been tested at several scales and is now an established technology for water treatment (groundwater remediation, wastewater treatment, safe drinking water production) [4, 5, 7-9, 11]. However, despite observed efficiency, the Fe0 reactive wall technology as a whole is based on a false premise which should be clarified before the technology develops properly. The objective of this communication is to demonstrate that leading scientists or research groups working on “water treatment with Fe0” are too confident with the false premise that iron is a reducing agent for contaminant reductive transformations. Accordingly, each alternative concept is systematically disregarded, endangering the integrity of science and questioning the efficiency of the peer review system. For the sake of clarity, the popular state- 2 of-the-art knowledge on the mechanism of contaminant removal in Fe0/H2O systems will first be presented. 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 2 Mechanism of contaminant removal Since the introduction of Fe0 bed filtration technology for groundwater remediation, several contaminants and contaminant groups have been reported to be removed by reductive transformations [17-19]. Clearly, contaminants were considered to be removed because of their chemical transformations possibly making them less harmful (reductive degradation) or less mobile (reductive precipitation). Accordingly, the case for which reduction products are toxic is still actively discussed [20]. Moreover, the formation of the universal oxide film on Fe0 (reactivity loss) and the pore filling by iron corrosion products (permeability loss) have been regarded as the major inhibitive factors for the process of contaminant removal [3, 21- 23]. The formation of the universal oxide film at the surface of Fe0 is a characteristic of iron corrosion at pH > 4.5 [24, 25]. The film results from the precipitation of iron (hyrdr)oxides at the surface or in the vicinity of Fe0. In separating Fe0 from water, the formed porous film necessarily lowers the kinetic of iron oxidation. On the other hand, the permeability loss results from the expansive nature of iron corrosion. In fact, upon corrosion, the volume of formed iron oxide is 2.08 to 6.40 times larger than the volume of the parent metal (V0) in the lattice (Fe0) [26,27]. The volumetric expansion (ΔV - Eq. 1) corresponds to the extent of porosity loss. ΔV = (η - 1)*V0 (1) Where 2.08 ≤ η ≤ 6.40. The hitherto presentation shows that three major opened questions of the Fe0 technology are: (i) how can harmful reaction products be removed? (Question 1) (ii) how can reactivity loss be prevented? (Question 2) and how can permeability loss be properly considered? (Question 3) Answering these three questions will accelerate technology development and could open 3 new perspectives. However, it is not likely that the popular state-of-the-art knowledge presented above will enable the adequate search of answers because it has neglected several key aspects. It should be emphasized that Question 2 and Question 3 have been recently theoretically addressed. It was shown that mixing Fe 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 0 (< 52 vol-%) with inert or reactive but non expansive additive will sustain long term corrosion and prolong Fe0 bed’s lifespan [28, 29] 3 Fundamental flaws of the current accepted mechanism The premise that aqueous contaminants are reduced by Fe0 is founded on the thermodynamic instability of Fe0 in aqueous solutions (immersed iron corrosion). Accordingly, all oxidized contaminants which electrode potential is larger than that of Fe0 (E0Fe2+/Fe0 = -0.44 V) should be reduced by Fe0. However, water (H2O or H+ - E0H+/H2 = 0.0 V) as solvent is also corrosive for Fe0 and is present in stoichiometric abundance. Accordingly, irrespective from the presence of any reducible species (including dissolved oxygen and contaminants), water will oxidize Fe0. This reaction can not be considered a side reaction as has been mistakenly done [17] (Mistake 1). Moreover, it is the oxidation of Fe0 by water that produces the oxide scale responsible for the reported reactivity loss. Accordingly, reactivity loss is an inherent property of Fe0 and should properly considered if Fe0 is to be used in aqueous solutions (at pH > 4.5). In this respect, enhancing the reactivity of Fe0 by reducing its particle size (μm and nm) for example will not avoid the formation of the universal oxide scale. The oxide scale is necessarily a diffusion barrier for all dissolved species as its shields the Fe0 surface [24, 25]. Note that the initial oxide scale is porous alloying retarded transport of water and dissolved species including FeII, H/H2 and eventually molecular O2. The initial porous film may develop to an impervious one which stops corrosion. The process of film formation is rigorously the same under anoxic and oxic conditions. The major difference being the kinetics. According to Cohen [25] the film forms 65 times slower under anoxic conditions. Given that adsorbed FeII and H/H2 are powerful reducing agents, they will more likely contribute to contaminant 4 reduction (if applicable) than the surface of Fe0 [30]. Accordingly, even when contaminant reduction is quantitative, effective reducing agents are likely primary iron corrosion products (indirect reduction by Fe 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 II and H/H2) (Mistake 2). The paramount mistake (Mistake 3) has been to confound “contaminant removal” and “contaminant reduction”. The goal of water treatment is contaminant removal and it is very difficult to say whether the reduction of a species has been achieved by electrons from Fe0, FeII or H/H2 (direct or indirect reduction). It is clear that a reduced contaminant should be removed from water to obtain clean water. The five major contaminant removal mechanisms in Fe0 beds are: adsorption, co-precipitation, precipitation, size-exclusion and volatilization. Biological and chemical transformations could influence the removal capability of a substance but are not removal mechanisms. Accordingly, Question 1 (how can harmful reaction products be removed? – Section 2) should be revisited because contaminants and transformation products should be removed from the aqueous phase. Moreover, contaminant transformation products have been removed from the aqueous phase, unless the technology would have not been efficient. The problem is that “contaminant removal” and “contaminant reduction” were randomly interchanged. The hitherto presentation has unmasked the popular state-of-the-art knowledge on the process of aqueous contaminant removal by Fe0 as non accurate. This false premise is well-accepted and repeated in the introduction of many articles and books. The question arises why the technology is efficient. 4 The real mechanism of contaminant removal in Fe0/H2O systems The real mechanism of aqueous contaminant removal in the presence of Fe0 (e.g. Fe0/H2O systems) can be derived from the pH dependence of the solubility of iron (FeII, FeIII) in aqueous solutions (Fig. 1) [31-33]. In other words, Fe0 can be regarded as generator for Fe species and the behaviour of these species used to discuss the efficiency of contaminant removal. In this case, the major question for Fe0/H2O sustainability is how to maintain the 5 corrosion process? (Question 4). Answering Question 4 is over the scope of this communication. However, is should be noticed that the continuation of the corrosion process depends on the porosity of the initial oxide scale on Fe 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 0 [34]. In other words, answering Question 4 will consist in investigating the effects of operational conditions on the porosity of the oxide scale in long-term experiments. In such experiments, the oxide scale should allow to form and transform on the surface of Fe0 (no or slow agitation, shaking or stirring). Figure 1 shows clearly, that for pH > 5 the solubility of Fe under oxic conditions is smaller than 10–5 M (0.56 mg). Under anoxic conditions, FeII has a relatively high solubility. The saturation concentration of FeII species approaches 0.5 M (28 g/L) at room temperature. However, dissolved Fe concentrations drop below 10-12 M (5.6*10-11 g/L) if FeII is polymerized e.g. [Fe(OH)2]n or oxidized to FeIII species [35]. This reaction results in the precipitation of Fe phases. FeIII species can also form under anoxic conditions, for example, when water contents oxidizing species like NO3-, MnO2 or MnOOH. FeII oxidation and Fe phase formation is often catalyzed by microorganisms. FeIII phases which form from solution begin as small clusters that evolve into larger polymeric units with time, eventually reaching colloidal sizes [36]. Contaminants are necessarily enmeshed in the matrix of corrosion Fe phases and/or adsorbed at their surface [37, 38]. Adsorbed or enmeshed contaminants can be further reduced by diffusing FeII and H/H2. Reduced contaminants will be adsorbed/enmeshed. That is the way contaminants have been successfully removed in Fe0/H2O systems [39, 40]. This statement is supported by results of some few authors (e.g. ref. [41-45], but the validity of the reductive transformation concept has never been systematically questioned. For example, Furukawa et al. [43] stated that “the ubiquitous presence of ferrihydrite suggests that the use of Fe0-PRBs may be extended to applications that require contaminant adsorption rather than, or in addition to, redox-promoted contaminant degradation”. Jia et al. [44] used Fe0 as adsorbent generator of the removal of non-reducible organic species (triazoles). The new concept stipulates that adsorption and co- 6 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 precipitation are the fundamental mechanism of contaminant removal, regardless of the redox properties. It is very important to note that contaminants are not removed by Fe0 or individual Fe phases but by the whole process of aqueous iron corrosion in a porous system (e.g. bed, wall) [39, 40]. Although this plausible explanation is compatible with the open literature on the interaction of iron oxides with contaminants in the hydrosphere as recently demonstrated [7, 40], some leading scientists [46-48] are continuing to defend the false premise with supposedly scientific arguments [49, 50]. This attitude has degraded the research community on Fe0 to a sort of modern communitarian knowledge system. 5 A modern knowledge system Currently the expression knowledge system is used in the scientific literature to describe efficient knowledge that has been used by small communities in the third world and by indigenous peoples in North America and Australia [51-55]. According to Mapara [55], indigenous knowledge systems (IKS) are bodies of knowledge of the indigenous people of particular geographical areas that have survived on for a very long time. They are knowledge forms that have failed to die despite colonial onslaught and scholar arrogance [55]. In social science, it is commonplace to scientifically investigate a IKS in order to extent its applicability [55, 56]. Sometimes, despite proven amelioration of IKS by mainstream science, indigenous people are not ready to modify their original knowledge unless its holistic socio- cultural and spiritual dimensions are conserved. Presently, the same trend is observed in using Fe0 for water treatment as demonstrated above. Whatever modern experimental tools have demonstrated to sustain the false premise, the objective of investigations were falsified and/or good results were misinterpreted. There is no need to multiply examples. Interested readers are referred to two recent articles [57, 58]. The sole example will be that of efforts to establish a mass balance to demonstrate the efficiency of Fe0 to reduce contaminants [59]. In real systems contaminants are removed within the 7 oxide scale on Fe0. It is not likely that water will release enmeshed contaminants because water can not dissolve iron oxides [60]. Accordingly, researchers reporting on mass balance without dissolving corrosion products have necessarily done something wrong like mechanically stirring the solution [59]. 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 The fact that many researchers are confident to this modern knowledge system is related by the huge number of peer review articles, that have been published on this topic during the last 17 years (from 1994 on). Peer review is the process that requires experts in a given field to evaluate an author’s work and ideas in that same field, usually for the purpose of publishing a paper or awarding a grant [61]. Considering this definition, it is fair to say that experts in the field of water treatment with Fe0 have built a “modern knowledge system” which is now challenging mainstream science. It is superfluous to recall that good manuscripts and proposals have been rejected, the curse of their authors being to think outside of the box. 6 Concluding remarks The present communication has recalled the danger of the expert feedback system when peer reviewers are not competent. The situation is exacerbated when a college of experts has been built around a false premise for years (yet two decades) [62]. During this time at least four generations of PhD students have spent their youth working on an idea defying mainstream science. This harsh criticism is not to make colleagues look bad, but rather to question the efficiency of the interdisciplinary approach. Iron corrosion is an old research field, where chemists, electrochemists, physicists have controversially discussed for decades [34, 63-65]. It is not likely that researchers without good chemical background will understand the important features of iron corrosion without guidance. The process of discovery may be intuitive, accidental, conjectural or inspirational but outcomes should be developed to predictable and repeatable issues. Despite 20 years of intensive research, no repeatable or predictable results have been achieved while using metallic iron for water treatment [57, 58]. Accordingly the whole achieved results is highly 8 qualitative. Furthermore, all existing models for the process of contaminant removal in Fe 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 0/H2O systems are faulty as none up to date properly considers the fundamental process of co-precipitation. To exploit the huge potential of Fe0 for environmental remediation and safe drinking water provision, a concerted research scientific effort is needed [39, 67-70]. The first step to this great future will be to abandon false premises and develop the actual communitarian knowledge system to a real scientific community. Acknowledgments The original manuscript was improved by the insightful comments from Daniel Fotio (University of Maroua – Cameroon). The manuscript was further improved by the insightful comments of anonymous reviewers from Fresenius Environmental Bulletin. Cited references [1] Bigg, T. and Judd, S.J. (2000) Zero-valent iron for water treatment. Environ. Technol. 21, 661–670. [2] Scherer, M.M., Richter, S., Valentine, R.L. and Alvarez, P.J.J. (2000) Chemistry and microbiology of permeable reactive barriers for in situ groundwater clean up. Rev. Environ. Sci. Technol. 30, 363–411. [3] Henderson, A.D. and Demond, A.H. (2007) Long-term performance of zero-valent iron permeable reactive barriers: a critical review. Environ. Eng. Sci. 24, 401–423. [4] Cundy, A.B., Hopkinson, L. and Whitby, R.L.D. (2008) Use of iron-based technologies in contaminated land and groundwater remediation: A review. Sci. Tot. Environ. 400, 42–51. [5] Thiruvenkatachari, R., Vigneswaran, S. and Naidu, R. (2008) Permeable reactive barrier for groundwater remediation. J. Ind. Eng. Chem. 14, 145–156. [6] Gillham, R.W. (2010) 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 in Hangzhou, 9 China, September 8-10, 2009"; Y. Chen, X. Tang, L. Zhan (Eds); Springer Berlin/London, 5– 15. 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 [7] Vodyanitskii, Y.N. (2010) The role of iron in the fixation of heavy metals and metalloids in soils: a review of publications. Eurasian Soil Sci. 43, 519 –532. [8] Comba, S., Di Molfetta, A. and Sethi R. (2011) A Comparison between field applications of nano-, micro-, and millimetric zero-valent iron for the remediation of contaminated aquifers. Water Air Soil Pollut. 215, 595–607. [9] Gheju, M. (2011) Hexavalent chromium reduction with zero-valent iron (ZVI) in aquatic systems. Water Air Soil Pollut., doi 10.1007/s11270-011-0812-y. [10] Jambor, J.L., Raudsepp, M. and Mountjoy, K. (2005) Mineralogy of permeable reactive barriers for the attenuation of subsurface contaminants. Can. Miner. 43, 2117–2140. [11] Phillips, D.H., Van Nooten, T., Bastiaens, L., Russell, M.I., Dickson, K., Plant, S., Ahad, J.M.E., Newton, T., Elliot, T. and Kalin, R.M. (2010) Ten year performance evaluation of a field-scale zero-valent iron permeable reactive barrier installed to remediate trichloroethene contaminated groundwater. Environ. Sci. Technol. 44, 3861–3869. [12] Reynolds, G.W., Hoff, J.T. and Gillham, R.W. (1990) Sampling bias caused by materials used to monitor halocarbons in groundwater. Environ. Sci. Technol. 24, 135–142. [13] Gillham, R.W. and O'Hannesin, S.F. (1992) Metal-catalysed abiotic degradation of halogenated organic compounds. IAH Conference: Modern Trends in Hydrogeology, Hamilton, Ontario, Canada, 94–103. [14] O'Hannesin, S.F. (1993) A field demonstration of a permeable reaction wall for the in situ abiotic degradation of halogenated aliphatic organic compounds. M.Sc. Thesis, University of Waterloo, Ontario, Canada. [15] Gillham R.W. and O’Hannesin S.F (1994) Enhanced degradation of halogenated aliphatics by zero-valent iron. Ground Water 32, 958–967. 10 [16] McMurty, D.C. and Elton, R.O. (1985) New approach to in-situ treatment of contaminated groundwaters. Environ. Progr. 4/3, 168 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 –170. [17] Matheson, L.J. and Tratnyek, P.G. (1994) Reductive dehalogenation of chlorinated methanes by iron metal. Environ. Sci. Technol. 28, 2045–2053. [18] Weber, E.J. (1996) Iron-mediated reductive transformations: investigation of reaction mechanism. Environ. Sci. Technol. 30, 716–719. [19] O´Hannesin, S.F. and Gillham, R.W. (1998) Long-term performance of an in situ "iron wall" for remediation of VOCs. Ground Water 36, 164–170. [20] Alvarado, J.S., Rose, C. and LaFreniere, L. (2010) Degradation of carbon tetrachloride in the presence of zero-valent iron. J. Environ. Monit. 12, 1524–1530. [21] Li, L. and Benson, C.H. (2010) Evaluation of five strategies to limit the impact of fouling in permeable reactive barriers. J. Hazard. Mater. 181, 170–180. [22] Bartzas, G. and Komnitsas, K. (2010) Solid phase studies and geochemical modelling of low-cost permeable reactive barriers. J. Hazard. Mater. 183, 301–308. [23] Moraci, N. and Calabrò, P.S. (2010) Heavy metals removal and hydraulic performance in zero-valent iron/pumice permeable reactive barriers. J. Environ. Manage. 91, 2336–2341. [24] Caule, E.J. and Cohen, M. (1955) The formation of thin films of iron oxide. Can. J. Chem. 33, 288–304. [25] Cohen, M. (1959) The formation and properties of passive films on iron. Can. J. Chem. 37, 286–291. [26] Caré, S., Nguyen, Q.T., L'Hostis, V. and Berthaud, Y. (2008) Mechanical properties of the rust layer induced by impressed current method in reinforced mortar. Cement Concrete Res. 38, 1079–1091. [27] Zhao, Y., Ren, H., Dai, H. and Jin, W. (2011) Composition and expansion coefficient of rust based on X-ray diffraction and thermal analysis. Corros. Sci. 53, 1646–1658. 11 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 [28] Noubactep, C., Caré, S., Togue-Kamga, F., Schöner, A. and Woafo, P. (2010) Extending service life of household water filters by mixing metallic iron with sand. Clean – Soil, Air, Water 38, 951–959. [29] Noubactep, C. and Caré, S. (2010) Enhancing sustainability of household water filters by mixing metallic iron with porous materials. Chem. Eng. J. 162, 635–642. [30] Stratmann, M. and Müller, J. (1994) The mechanism of the oxygen reduction on rust- covered metal substrates. Corros. Sci. 36, 327–359. [31] Liu, X. and Millero, F.J. (1999) The solubility of iron hydroxide in sodium chloride solutions. Geochim. Cosmochim. Acta 63, 3487–3497. [32] Liu, X. and Millero, F.J. (2002) The solubility of iron in seawater. Marine Chem. 77, 43– 54. [33] Rickard, D. (2006) The solubility of FeS. Geochim. Cosmochim. Acta 70, 5779–5789. [34] Nesic, S. (2007) Key issues related to modelling of internal corrosion of oil and gas pipelines – A review. Corros. Sci. 49, 4308–4338. [35] Baes, C.F., and Mesmer, R.E. (1976) The hydrolysis of cations, John Wiley & Sons, New York 489 pp. [36] Combes, J.M., Manceau, A., Calas, G. and Bottero, J.Y. (1989) Formation of ferric oxides from aqueous solutions: A polyhedral approach by X-ray absorption spectroscdpy: I. Hydrolysis and formation of ferric gels. Geochim. Cosmochim. Acta 53, 583–594. [37] Eusterhues, K., Rennert, T., Knicker, H., Kgel-Knabner, I., Totsche, K.U. and Schwertmann, U. (2011) Fractionation of organic matter due to reaction with ferrihydrite: Coprecipitation versus adsorption. Environ. Sci. Technol. 45, 527–533. [38] Hanna, K. and Boily, J.-F. (2010) Sorption of two naphthoic acids to goethite surface under flow through conditions. Environ. Sci. Technol. 44, 8863–8869. [39] Noubactep, C. (2010) Metallic iron for safe drinking water worldwide. Chem. Eng. J. 165 740–749. 12 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 [40] Noubactep, C. (2010) The fundamental mechanism of aqueous contaminant removal by metallic iron. Water SA 36, 663–670. [41] Lavine, B.K., Auslande,r G. and Ritter, J. (2001) Polarographic studies of zero valent iron as a reductant for remediation of nitroaromatics in the environment. Microchem. J. 70, 69–83. [42] Mantha, R., Taylor, K.E., Biswas, N. and Bewtra, J.K. (2001) A continuous system for Fe0 reduction of nitrobenzene in synthetic wastewater. Environ. Sci. Technol. 35, 3231–3236. [43] Furukawa, Y., Kim, J.-W., Watkins, J. and Wilkin R.T. (2002) Formation of ferrihydrite and associated iron corrosion products in permeable reactive barriers of zero-valent iron. Environ. Sci. Technol. 36, 5469–5475. [44] Jia, Y., Aagaard, P. and Breedveld G.D. (2007) Sorption of triazoles to soil and iron minerals. Chemosphere 67, 250–258. [45] Jiao, Y., Qiu, C., Huang, L., Wu, K., Ma, H., Chen, S., M,a L. and Wu L. (2009) Reductive dechlorination of carbon tetrachloride by zero-valent iron and related iron corrosion. Appl. Catal. B: Environ. 91, 434–440. [46] Schnoor, J. (2009) Environmental Science & Technology presents the 2009 Excellence in Review Awards. Environ. Sci. Technol. 43, 8003–8003. [47] Schnoor, J. (2010) Environmental Science & Technology presents the 2010 Excellence in Review Awards. Environ. Sci. Technol. 44, 7985–7985. [48] Nagpal, V., Bokare A.D., Chikate, R.C., Rode, C.V. and Paknikar, K.M. (2011) Reply to comment on “Reductive dechlorination of ?-hexachlorocyclohexane using Fe–Pd bimetallic nanoparticles”, by C. Noubactep. J. Hazard. Mater., doi:10.1016/j.jhazmat.2011.04.015. [49] Tratnyek, P.G. and Salter, A.J. (2010) Response to comment on “Degradation of 1,2,3- Trichloropropane (TCP): hydrolysis, elimination, and reduction by iron and zinc”. Environ. Sci. Technol. 44, 3198–3199. 13 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 [50] Xie, Y. and Cwiertny, D.M. (2010) Use of dithionite to extend the reactive lifetime of nanoscale zero-valent iron treatment systems. Environ. Sci. Technol. 44, 8649–8655. [51] Snively, G. and Corsiglia, J. (2001) Discovering indigenous science: Implications for science education. Sci. Ed. 85, 6–34. [52] Quigley, C. (2001) Globalization and science education: The implications for indigenous knowledge systems. Int. Educ. Stud. 2, 76–88. [53] Barnhardt, R. and Kawagley, A.O. (2005) Indigenous knowledge systems and Alaska native ways of knowing. Anthr. Educ. Quart. 36, 8–23. [54] Aikenhead, G.S. and Ogawa, M. (2007) Indigenous knowledge and science revisited. Cult. Stud. Sci. Educ. 2, 539–620. [55] Mapara, J. (2009) Indigenous knowledge systems in zimbabwe: Juxtaposing postcolonial theory. J. Pan African Studies 3, 139–155. [56] Cash, D.W., Clark, W.C., Alcock, F., Dickson, N.M., Eckley, N., Guston, D.H., Jäger, J. and Mitchell, R.B. (2003) Science and technology for sustainable development special feature knowledge systems for sustainable development. Proc. Natl. Acad. Sci. U S A. 100, 8086– 8091. [57] Noubactep, C., Licha, T., Scott, T.B., Fall, M. and Sauter, M. (2009) Exploring the influence of operational parameters on the reactivity of elemental iron materials, J. Hazard. Mater. 172, 943–951. [58] Noubactep, C. (2010) Characterizing the reactivity of metallic iron in Fe0/EDTA/H2O systems with column experiments. Chem. Eng. J. 162, 656–661. [59] Hwang, Y.-H., Kim, D.-G. and Shin, H.-S. (2011) Mechanism study of nitrate reduction by nano zero valent iron, J. Hazard. Mater. 185, 1513–1521. [60] Noubactep, C., Meinrath, G., Dietrich, P. and Merkel, B. (2003) Mitigating uranium in ground water: prospects and limitations. Environ. Sci. Technol. 37, 4304–4308. 14 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 [61] McCormack, N. (2009) Peer review and legal publishing: what law librarians need to know about open, single-blind, and double-blind reviewing. Law Library J. 101, 59–70. [62] Noubactep, C. (2011) Aqueous contaminant removal by metallic iron: Is the paradigm shifting? Water SA 37, 419–426. [63] Whitman, G.W., Russel, R.P. and Altieri, V.J. (1924) Effect of hydrogen-ion concentration on the submerged corrosion of steel. Indust. Eng. Chem. 16, 665–670. [64] Holmes, D.R. and Meadowcroft, D.B. (1977) Physical methods in corrosion technology. Phys. Technol. 8, 2–9. [65] Farelas, F., Galicia, M., Brown, B., Nesic, S. and Castaneda, H. (2010) Evolution of dissolution processes at the interface of carbon steel corroding in a CO2 environment studied by EIS. Corros. Sci. 52, 509–517. [66] Obanijesu, E.O., Pareek, V., Gubner, R. and Tade, M.O. (2010) Corrosion education as a tool for the survival of natural gas industry. Nafta Sci. J. 61, 541–554. [67] Pradeep T. and Anshup (2009) Noble metal nanoparticles for water purification: A critical review. Thin Solid Films 517, 6441–6478. [68] Antia, D.D.J. (2010) Sustainable zero-valent metal (ZVM) water treatment associated with diffusion, infiltration, abstraction and recirculation. Sustainability 2, 2988–3073. [69] Litter, M.I., Morgada, M.E. and Bundschuh J. (2010) Possible treatments for arsenic removal in Latin American waters for human consumption. Environ. Pollut. 158, 1105–1118. [70] Antia, D.D.J. (2011) Modification of aquifer pore-water by static diffusion using nano- zero-valent metals. Water 3, 79–112. 15 Fig. 1: Solubility data of FeII in 0.1 M NaCl (25 °C) and FeIII in 0.01 M NaCl (25 °C) as a function of pH. Data for Fe 380 381 382 II are from ref. [33] and data for FeIII from ref. [31]. 2 4 6 8 10 12 -10 -8 -6 -4 -2 0 [Fe] = 10-5 M FeII solubility FeIII solubility lo g [F e] pH value 383 384 16