1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Effect of pumice and sand on the sustainability of granular iron beds for the aqueous removal of CuII, NiII, and ZnII Stefania Bilardia, Paolo S. Calabròa, Sabine Caréb,*, Nicola Moracia, Chicgoua Noubactepc,d aUniversità degli Studi Mediterranea di Reggio Calabria, MECMAT, Mechanics and Materials Department, Faculty of Engineering, Via Graziella, loc. Feo di Vito, 89122 Reggio Calabria, Italy. bUniversité Paris-Est, Laboratoire Navier (UMR 8205), CNRS, ENPC, IFSTTAR, F-77455 Marne-la-Vallée, France. cAngewandte Geologie, Universität Göttingen, Goldschmidtstraße 3, D-37077, Göttingen, Germany. dKultur und Nachhaltige Entwicklung CDD e.V., Postfach 1502, D-37005 Göttingen, Germany * corresponding author: e-mail: sabine.care@ifsttar.fr; Tel.: + 33 140435466; Fax: + 33 140435450. Running Title: Admixing pumice and sand to reactive Fe0 sustains long-term metal removal. Acronym List PRB Permeable Reactive Barrier Keywords: Hydraulic conductivity, Reactive barriers, Pumice, Sand, Zerovalent iron. 2 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 Abstract Current knowledge of the basic principles underlying the design of Fe0 beds is weak. The volumetric expansive nature of iron corrosion was identified as the major factor determining the sustainability of Fe0 beds. This work attempts to systematically verify developed concepts. Pumice and sand were admixed to 200 g of Fe0 in column studies (50:50 volumetric proportion). Reference systems containing 100 % of each material have been also investigated. The mean grain size of the used materials (in mm) were 0.28 (sand), 0.30 (pumice) and 0.50 (Fe0). The five studied systems were characterized (i) by the time dependent evolution of their hydraulic conductivity (permeability) and (ii) for their efficiency for aqueous removal of CuII, NiII, and ZnII (about 0.30 M of each). Results showed unequivocally that (i) quantitative contaminant removal was coupled to the presence of Fe0, (ii) additive admixture lengthened the service life of Fe0 beds, and (iii) pumice was the best admixing agent for sustaining permeability while the Fe0/sand column was the most efficient for contaminant removal. The evolution of the permeability was well-fitted by the approach that the inflowing solution contained dissolved O2. The achieved results are regarded as starting point for a systematic research to optimise/support Fe0 filter design. 1 Introduction Permeable reactive barriers (PRBs) containing metallic iron (Fe0) as reactive medium have been developed during the past two decades to an established technology for groundwater remediation [1-11]. The original PRB technology containing granular Fe0 has been expanded to the injection of nano-scale Fe0 for source remediation [8-11]. To date, more than 180 Fe0 PRBs have been installed worldwide [8,11]. Successful accomplishment of remedial goals has been typically reported. At some few sites, system failures were recorded [8,12]. Reported failures were attributed to design shortcomings due to poor site characterization (reason 1), poor design selection (reason 2) or installation at sites where the technology is not an appropriate choice (reason 3) [8,12,13]. However, there is clear evidence that the physico-chemistry of the Fe0/H2O system was not properly 3 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 considered [14]. Accordingly, design shortcomings may have reasons different from or additional to reasons 1, 2 and 3. A major concern of Fe0 PRBs is related to the reduction of the hydraulic conductivity (permeability loss) with time [15-17]. Laboratory and field data have also demonstrated diminished Fe0 reactivity with time [9, 18]. Consequently, the sustainability of Fe0 PRBs in terms of both Fe0 reactivity and system permeability has been extensively discussed during the past 15 years [1,18-24]. Reported results are confusing and even conflicting as demonstrated below for trichloroethene (TCE). O’Hannesin and Gillham [1] reported on successful TCE (268 mg/L) reductive degradation by a Fe0/sand mixture containing 22 % Fe0 by weight (laboratory and field test). Bi et al. [21] tested several weight Fe0/sand mixtures (25/75, 50/50, 75/25, 85/15 and 100/0) for TCE (up to 60 mg/L) treatment and reported that the system with less than 50 % Fe0 was not efficient (laboratory test). Ruhl et al. [18] evaluated four dual mixtures (Fe0/anthracite, Fe0/gravel, Fe0/pumice and Fe0/sand) for TCE (about 10 mg/L) treatment (laboratory test). The used masses of additives varied from 24.4 g for pumice to 104.3 g for gravel. The used mass of Fe0 was 100 g resulting in Fe0 weight ratios varying from 49 % for gravel to 80 % for pumice. Ruhl et al. [18] concluded that tested dual systems are not applicable for TCE treatment but “might be applicable for the removal of heavy metals”. The three examples reveal that researchers use varying experimental procedures to characterize processes in Fe0/H2O systems (see Tab. 1) [1,21,25-33]. These procedures differ for instance in Fe0 intrinsic reactivity, Fe0 pre-treatment, Fe0 mass, Fe0 particle size and shape, used admixing additives and their proportions, duration of the experiments, nature and concentration of the contaminant, buffer application, solution flow velocity and water chemistry. As a result, many different reports for the same compound are available in the literature (even for the same Fe0). Water and dissolved inorganic constituents (Ca2+, HCO3-, Mg2+, O2, PO43-, SO42-) react with iron species (Fe0, FeII and FeIII) to form precipitates that progressively fill the inter-particular porosity 4 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 within a Fe0 filter. The potential of these in situ generated precipitates to limit the permeability and the efficiency of Fe0 PRBs filters has been clearly documented [1,4,6-8]. However, the role of Fe0 oxidation by water has not always been properly considered and the role of gas (H2) formation in porosity/permeability loss has been sometimes overestimated [15,16]. Recent theoretical works demonstrated that a Fe0-based filter should be considered as a system in which iron is corroded mostly by water and the micro-pollutants are sequestrated in the matrix of precipitation corrosion products [14,34-36]. This view corroborates concordant reports regarding Fe0 filters as a long-term sink for C, S, Ca, Si, Mg, and N [12,37-39]. The present work is an attempt to improve the design of Fe0 filtration systems based on recent theoretical studies [14]. In the present work, the efficiency of five different systems (A to E) for aqueous contaminant removal is tested in column studies. The volumetric composition of individual systems is given as: (A) 100 % sand, (B) 100 % pumice, (C) 100 % Fe0, (D) 50:50 Fe0:pumice, and (E) 50:50 Fe0:sand. The model solution contained about 0.30 M of CuII, NiII, and ZnII. The evolution of the systems is characterized by determining the (i) extend of contaminant removal, and (ii) evolution of hydraulic conductivity. 2 Materials and methods 2.1 Chemicals and solutions Copper(II) nitrate hydrate (purity 99.999), nickel(II) nitrate hexahydrate (purity 99.999) and zinc(II) nitrate hexahydrate (purity 99.000) were obtained from Sigma-Aldrich. All chemicals used for experiments and analysis were of analytical grade. The used solutions were obtained by dissolving copper nitrate, nickel nitrate and zinc nitrate in distilled water. The molar concentration of the resulting solution was as follows: 0.27 M Cu, 0.29 M Ni and 0.37 M Zn. The corresponding mass concentrations are 17 mg/L Cu, 17 mg/L Ni, and 23 mg/L Zn. 2.2 Solid materials 2.2.1 Porosity of binary granular media 5 93 94 The total porosity Φ of a binary granular medium composed of two kinds of particles P1 and P2 (here P1 corresponds to Fe0 particles) is given by : V M V MM1 V V V VV1f rz 2sa2 2 rz 2sa21sa1 rz 2a 2 rz 2a1a pppper ρϕ+ρ+ρ−=Φ ϕ++−=ϕ+Φ=Φ /.]//[ .][.int (1) 95 96 Where: (i) Φinter is inter-particular porosity (ii) Vai, Mai, ρai are respectively the apparent volume of the particles i, the mass and the apparent specific weight, and ϕ2 is the intra-particular porosity of the particles 2 with 97 02 =ϕ for non porous particles (sand) and 02 ≠ϕ for porous particles (pumice) and (iii) V 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 rz is the total packing volume of the granular medium. 2.2.2 Metallic iron (Fe0) The used Fe0 is of the type FERBLAST RI 850/3.5, distributed by Pometon S.p.A., Mestre - Italy. The powdered material contents mainly iron (> 99.74 %). Identified impurities included mainly Mn (0.26 %), O, S and C. The material is characterized by uniform grain size distribution. The coefficient of uniformity U (ratio between the diameters corresponding to 60 and 10 % finer in the grain size distribution) is 2. The mean grain size (d50) is about 0.5 mm and the initial porosity of used Fe0 medium has been estimated to be Φ0 = 49.6 % (see Tab. 2, Eq. 1). 2.2.3 Pumice The used pumice originates from Lipari (Aeolian Islands, Sicily – Italy). Its mineralogical composition was determined as follows: SiO2: 71.75 %; Al2O3: 12.33 %; K2O: 4.47 %; Na2O: 3.59 %; Fe2O3: 1.98 %; moreover it contains about 4 % of bound water entrapped in the pumice structure during the sudden cooling of magma and traces of other compounds (e.g. CaO, SO3, MgO, TiO2, FeO, MnO, P2O5). The material is characterized by uniform grain size distribution. The coefficient of uniformity U is 1.4. The mean grain size (d50) is about 0.3 mm. This type of pumice has been chosen since it was the available fraction closest to Fe0 in dimension. The initial porosity of the 6 115 pumice granular medium has been estimated to be Φ0 = 73.3 % and the inner porosity of the pumice (intra particular porosity ϕpp ) to be 41 % (Tab. 2, Eq. 1) through Mercury Intrusion Porosity (MIP) measurements under the hypothesis that the relative density (packing) of granular mixtures in the columns and during MIP experiments were the same. 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 2.2.4 Sand The used quartz sand was obtained from a river quarry. The sand was carefully washed with distilled water and sieved before use. The material is characterized by uniform grain size distribution. The coefficient of uniformity U is 2.2. The mean grain size (d50) is about 0.28 mm. The material was used without any further characterization. The initial porosity of the sand medium has been estimated to be Φ0 = 45.0 % (Tab. 2, Eq. 1). 2.3 Column operation Laboratory scale polymethyl methacrylate (Plexiglas) columns were operated in up-flow mode. The influent solution was pumped upwards from a single PE bottle using a precision peristaltic pump (Ismatec, ISM930). The flow rate was maintained constant at a value of 0.5 mL/min. Tygon tubes were used to connect inlet reservoir, pump, columns and outlet. Five plexiglas columns (50 cm long, 5.0 cm inner diameter) were used in the experiments (Fig. 1). Five different systems were investigated (Systems A through E) (Tab. 3). System A was the operational reference system containing only sand (0 % Fe0), System B was the second operational reference containing only pumice (0 % Fe0) and system C was a pure iron bed (100 % Fe0). The volumetric proportion of Fe0 in the 2 other systems (D, E) was 50 %. In systems C to E, the mass of iron was fixed to 200 g. This mass represented either 100 % of the reactive zone (rz) or the relevant volumetric proportion of rz (Fig. 1, Tab. 3). In system B the pumice volume was set to be the same occupied by Fe0 in system C while the pumice mass was obviously the same as in system E. The total porosity of the all systems varies between 45 % and 73 % (Tab. 2, Eq. 1). 7 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 The hydraulic conductivity [40] was determined during the column tests, by either constant-head (k > 10-6 m/s) or variable-head (k < 10-6 m/s) permeability methods, at selected dates to assess the permeability of the systems. The experiments were performed at room temperature (21 ± 4 °C). Samples for analysis were collected at periodic intervals and the experiments where prolonged until contaminant breakthrough (systems A and B) or a significant loss of the hydraulic conductivity (systems C to E) was observed. 2.4 Analytical methods Samples from the columns were centrifuged at 3000 rpm (ALC, PK121 Multispeed Centrifuge). The supernatant was vacuum filtered through a 0.45 μm glass filters. The aqueous concentrations of Cu, Fe, Ni and Zn were then measured by Atomic Absorption Spectrophotometry (AAS - Shimadzu AA – 6701F; wavelengths: Cu 324.75 nm, Ni 232.00 nm, Zn 213.86 nm, Fe 248.33 nm) using air- acetylene flame and according to conventional Standard Methods [41]. The used AAS device was calibrated using three operational standard solutions covering the expected concentration range of the samples (after dilution if applicable). Each operational standard solution was prepared by an appropriate dilution of a 1000 ppm (Cu(NO3)2, Fe(NO3)3, Ni(NO3)2, Zn(NO3)2) certified atomic absorption stock solution from Merck (Germany). The minimum correlation coefficient of calibration curves was of 0.997. The pH value was measured by combination glass electrodes (WTW GmbH, inolab pH/Cond 720). 2.4.1 MIP measurements MIP measurements have been carried out using a Micromeritics instrument apparatus type (AutoPore IV 9500). The instrument is capable of a minimum intruding pressure of 3.4 kPa and a maximum pressure of 227 MPa, so that the pore radius ranges from 2.7 nm to 180 μm. For pumice particles the measured pore data allow determining the inter-particular and intra- particular porosities of the pumice particles, the apparent specific weight ρas (defined as the ratio of the mass and the apparent volume of the pumice particles) and the specific weight ρs (defined as the 8 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 ratio of the mass and the volume of the solid phase of the pumice particles). 2.5 Expression of the experimental results In order to characterize the magnitude of tested systems for contaminant removal, the removal efficiency (E) and the specific removal (Es) were calculated using Eq. 2 and Eq. 3 [33]. E = mrem/min*100 (2) Es = mrem/mFe*100 (3) where min is the mass of contaminant flowed into the column, mrem is the mass of removed contaminant, and mFe the mass of Fe0 present in the column. 2.6 Evaluation of the residual porosity When iron corrodes, porous oxide layers are formed at the Fe0/H2O interface. The volume of the corrosion product (Voxide) is higher than that of the original metal (VFe). The ratio (η) between the volume of expansive corrosion product and the volume of iron consumed in the corrosion process is called ‘‘coefficient of volumetric expansion’’ [42,43]. Generally, Voxide is 2.08 to 6.40 times larger than V0 (2.08 ≤ η ≤ 6.40 for free expansion). At any time (t > 0), Voxide can be calculated using Eq. 4: Voxide = η*(V0 –Vt) (4) Where η is the coefficient of volumetric expansion, (V0 –Vt) is the consumed Fe0 volume with V0 the initial volume of Fe0 and Vt its residual Fe0 at time t. The effective volumetric expansion ΔV (Eq. 5) corresponding to the volume of pores that is occupied by iron corrosion products is a measure of the extent of porosity loss. ΔV = (η - 1)* (V0 –Vt) (5) The residual porosity of the system at time t (Φ(t)) may be estimated by (Eq. 6): V VV1t rz t0 0 )).(()( −−η−Φ=Φ (6) 186 187 Where is the initial porosity of the reactive zone given in Tab. 3, and VΦ0 rz is the volume of the 9 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 reactive zone. 3 Results and discussion 3.1 pH variation and Fe breakthrough Metal ions are known to be removed from the aqueous phase in packed Fe0 beds by adsorption, co- precipitation and adsorptive size-exclusion when the pH > 4.5 [44-48]. Figure 2a clearly demonstrates that the pH value of the initial solution (t = 0) and that of the effluent from all columns was larger that 5.5. This suggests that contaminant removal could be quantitative (see Tab. 4) if the residence time is sufficient to enable the formation of enough iron corrosion products for contaminant retention in the column. A hint that quantitative contaminant removal was likely is given by the evolution of the iron concentration (Fig. 2b). Fig. 2b clearly shows that the effluent iron concentration was less than 0.2 mg/L and reached values close to up to 1 mg/L only shortly before clogging for the system with 50 % pumice (system E). This observation could be attributed to accelerated transport through preferential flow paths [26]. 3.2 Metal breakthrough Fig. 3 and Tab. 4 summarize the results of contaminant removal in systems containing Fe0. It is clear from Fig. 3a that no Cu breakthrough occurs. Ni breakthrough occurs first (Fig. 3b). In fact Ni breakthrough occurs before day 8 in the system with 100 % Fe0. Zn is the next less retained metal with a breakthrough occurring at day 10 in the system with 100 % Fe0 (Fig. 3c). The observed order of removal efficiency corresponds to the selectivity sequence for iron oxides and soils: Ni < Zn < Cu [49-51]. For example, Fontes and Gomes [50] found that in competitive adsorption CuII maintains its strong affinity with the surface, while NiII and ZnII are displaced from the surface. This observation corroborates the view that species with higher affinity to iron oxides are better treated by Fe0 filters [23,24,52,53]. Another important result from Fig. 3 is that no contaminant breakthrough was observed in the system with 50 % sand (system D). This system is less porous than the system with 50 % pumice 10 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 (Tab. 3). The differential behaviour of the systems with pumice and sand (D and E) illustrates the dilemma of sustaining efficiency (maximum Fe0 ratio) while maintaining permeability, for instance by using porous pumice in place of sand [33]. As discussed in details elsewhere [14,36] this dilemma could be solved by using an appropriate thickness of the Fe0-based layer for each relevant additive (e.g. activated carbon, anthracite, gravel, pumice, sand) to achieve water treatment under site specific conditions. Relevant site specific parameters include the nature of contaminant, the water chemistry and the water flow velocity. In other words, a proper design (reason 2, § 1) should be extended to the width of the Fe0 PRB, the nature of the admixing agent (e.g. type and grain size distribution) and the proportion of Fe0 therein. The last important issue on contaminant breakthrough concerns the suitability of specific removal (Eq. 3) for a dynamic system in which reactive species are progressively generated. Es values from Tab. 4 show that the lowest specific removal (1.74 mg Ni/g Fe0) was obtained in system C (100 % Fe0). While this result seems contradictory, it corroborates the view that iron corrosion is self- inhibitory and that decreasing the proportion of Fe0 is a powerful tool to increase sustainability ([54] and ref. cited therein). Tab. 5 shows that correcting Es by considering the extent of Fe0 depletion at tlimit (Es,eff) restores the intuitive trend that “the greater the adsorbent amount, the larger the Es value“. Accordingly, the highest Es,eff values were obtained in system C (absolute value) which clogged at first. This result corroborates previous findings that filtration systems containing a 100 % Fe0 layer are efficient but not sustainable [55,56]. Moreover, the fact that the effective specific removals for the three systems are similar (4.0 ≤ Es,eff ≤ 6.8) is a hint that the calculation of the consumed iron is right. Note that, Es,eff values are derived from Es values on the basis of the extent of Fe0 depletion at tlimit (Es, tlimit), not at the depletion at the end of the experiment. This result means that no significant breakthrough was observed before tlimit. 3.3 Hydraulic conductivity 11 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 The results presented in Fig. 4 clearly demonstrate that the hydraulic conductivity decreases with time for the systems containing Fe0 particles (systems C, D, E) then remains constant at time tlimit (Tab. 5). The initial hydraulic conductivity K0 for all systems is about 5.10-4 m/s. The hydraulic conductivity tends to about the limit value Klimit = 5.10-9 m/s at time tlimit. The results show that the decrease of the permeability is down to about five orders of magnitude K0/Klimit = 1.10-5 (-). The time tlimit seems to depend on the investigated system (Tab. 5). The data in Tab. 5 clearly indicate that Fe0 admixture with sand and pumice resulted in extended service life. The longest service life was observed for the system with pumice particles and is consistent with the fact that intra-particle porosity has contributed to increased permeability [57,58]. Among the proposed models in the literature, the Kozeny-Carman equation is often considered to evaluate the evolution of the hydraulic conductivity [59]. This equation was developed after considering a porous material as an assembly of capillary tubes and yielded the hydraulic conductivity K as function of the porosity Φ, the specific surface S (m2/kg of solids) and a factor C to take into account the shape and tortuosity of channels. The first approximation is to accept the Kozeny-Carman equation [60-64]: )(.)(.)( Φ− Φ− Φ Φ= 1 01 0 KtK 23 0 (6) 253 254 255 256 257 258 where K0 is the initial hydraulic conductivity and Φ0 the initial porosity. For the evaluation of the residual porosity Φ as a function of the time t, uniform corrosion for spherical particles with initial radius R0 (here R0 = 500 μm) is assumed. Individual particles corrode independently with the same kinetics until material depletion. Under these assumptions, it is considered that the actual radius R(t) of Fe0 particles varies linearly with time t according to: t tRRtR depletion 0 0 , .)( ∞ −= (7) 259 where t depletion,∞ is the time at Fe0 depletion. 260 261 262 From Eq. 5, 6 and 7, it is possible to simulate the decrease of the hydraulic conductivity as a function of time (Figure 4b-d). The modelling has been applied for the coefficient of volumetric expansion η = 6.40 in coherence with high O2 levels and for two times at Fe0 depletion ( t depletion,∞ = 50 days and 62.5 days) [43]. It can be noticed that the maximum volume of Fe 263 264 265 0 which may corrode is the one which leads to clogging (VFe,clogging) and is expressed by: 1 V=V rz0gingcFe −η Φ . log, . (8) 266 267 At time tlimit it is assumed that the volume of consumed Fe0 tends to the one leading to the clogging of the column ( ).(. log,VV0011Vt gingcFe−= in calculations) and remains constant for t > tlimit. Under these assumptions, the permeability K 268 269 270 271 272 limit corresponding to the time tlimit is reached at the clogging of the columns (Φ ≈ 0, see Tab. 5). The obtained results show that there is a good agreement between experiment and modelling concerning the kinetic of the decrease of the permeability with time and the Klimit value. It can be noticed that the linear law for corrosion process (Eq. 7) with the two considered times ( t depletion,∞ ) as a first approximation of corrosion kinetic, allows to well reproduce the decrease of the permeability at the beginning of the filtration process. However, the evolution of the permeability around t 273 274 275 276 277 278 279 280 281 282 283 limit can not be accurately predicted. The proposed modelling is a first attempt to predict the time-dependent decrease of the hydraulic conductivity (permeability loss) on the basis of the volumetric expansion of corroding iron. This work shows that the evolution of the hydraulic conductivity may be predicted without considering the evolution of the tortuosity or the specific surface in the Kozeny-Carman equation and is the consequence of the filling of the porosity by expansive iron corrosion products. 3.4 Discussion The achieved experimental results and the proposed modelling show that there is a significant effect of the inner porosity of the pumice (system E). This effect is a clogging delay compared to the 12 13 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 Fe0/sand system (Fig. 4). This result is explained by the internal porosity of pumice particles which may store iron corrosion products, delaying the filling of the inter-granular porosity. Although theoretically sound [58,65,66], this hypothesis has to be confirmed in future works, for instance, using X-ray micro-tomography to probe inner porosity of the pumice specimen and considering various pumice material with differential pore connectivity. For a better understanding of the evolution of the initial porosity as iron corrosion proceeds, it is imperative to couple imaging (visualization) and mathematical modelling. The first attempt to visualize the deposition of iron particles (nano-scale) in the context of remediation with Fe0 was recently published [67]. It is expected that the use of X-ray microtomography visualization (and other appropriate techniques) will enable a better understanding of the effects of corrosion products on the bed clogging and to interpret the evolution of the residual porosity. An increase of the sustainability of the Fe0 bed is noticed (Tab. 5). More iron could be consumed at the time tlimit. The extent of Fe0 depletion is increased by using admixtures. This result corroborates the view that admixing Fe0 with non-expansive material is a tool to induce sustainability [14,24]. Accordingly, the repeatedly reported cost reduction (Fe0 costs) ([21] and ref. cited therein) should be regarded as a positive side-effect. In other words, while using admixtures, material wastage [12] is avoided and service life is increased. It seems that the Fe0 proportion in efficient real systems should be lower than 50 % (1:1, v/v) used here [23,24]. In fact, the efficiently permeable reactive barrier at Borden (Ontario, Canada) contained only 22 % Fe0 (w/w) [1]. On the other hand, while testing Fe0 for viruses and bacteriophages removal from drinking, Shi et al. [68] found out that a sand filter containing only 15 % Fe0 (w/w) was very efficient for microbe removal. The design of Shi et al. [68] consisted in a column packed with sand (sand filter) containing a reactive Fe0/sand layer (50:50, v/v). While the Fe0/sand ratio was the same as the one discussed here, this work and related studies propose that parameters such as the characteristics of the column, the mass of Fe0, the chemical reactivity of Fe0, the thickness of the Fe0/sand layer, the proportion of Fe0, the relative 14 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 size of used particles (δ values), the porosity of the additives are routinely given to enable the comparison of results. 4 Conclusions The concept that dual Fe0/inert additive systems are more sustainable than pure Fe0 systems for water treatment is validated using pumice and sand as additives and CuII, NiII and ZnII as model contaminants. As expected the sand system was more efficient for contaminant removal and the pumice system more permeable. The order of contaminant removal efficiency corresponds to the selectivity sequence for iron oxides. This observation corroborates the view that chemical reduction (if applicable) is of secondary important for the process of contaminant removal. The presented work highlights the volumetric expansive nature of iron corrosion as the most important clogging factor in water treatment. Filter clogging is demonstrated to be related to pore blocking in the inlet zone. Therefore, reliable strategies have to be developed to design sustainable Fe0 filters under environmental conditions (water works). Systematic research at laboratory scale is needed to understand the impact of various factors on the clogging process. These factors include: the temperature, the nature of Fe0 (chemical reactivity) and used additives (reactivity, porosity), the shape and dimension of Fe0 and additives, the relative dimensions of Fe0 and additives (δ values) and the quality of the inflowing aqueous solution (pH, O2 level, HCO3-, humic substances, contaminants). The possibility to use various Fe0 materials of different reactivity in the treatment chains should be carefully considered. For example, a readily reactive Fe0 material (e.g. powdered) can be used in the first column(s) as O2 scavenger and substituted when clogging occurs. In such a configuration, less reactive materials (e.g. granular) are used in subsequent columns for effective water treatment. The net output of such a systematic research will be the minimization of uncertainties on the long term efficiency (sustainability) of Fe0-based filtration systems, including nano-scale Fe0. The lack of systematic approach has already led to difficulties in finding research funding [69] after more 15 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 than 15 years intensive research. Moreover, there is currently a hindrance in the spreading of this efficient technology in Europe [70]. At the term, a small number of avoidable failures could result in dismissal of a proven efficient technology. Acknowledgements The authors are grateful to Dott Giuseppe Panzera for its essential assistance during this research activity. This paper is the results of the joint work of the authors. References [1] S.F. O´Hannesin, R.W. Gillham, Long-term performance of an in situ "iron wall" for remediation of VOCs. Ground Water 1998, 36, 164–170. [2] K. Komnitsas, G. Bartzas, I. Paspaliaris, Inorganic contaminant fate assessment in zero-valent iron treatment walls. Environ. Forensics 2006, 7, 207–217. [3] K. Komnitsas, G. Bartzas, I. Paspaliaris, Modeling of reaction front progress in fly ash permeable reactive barriers. Environ. Forensics 2006, 7, 219–231. [4] A.D. Henderson, A.H. Demond, Long-term performance of zero-valent iron permeable reactive barriers: a critical review. Environ. Eng. Sci. 2007, 24, 401–423. [5] R. Thiruvenkatachari, S. Vigneswaran, R. Naidu, Permeable reactive barrier for groundwater remediation. J. Ind. Eng. Chem. 2008, 14, 145–156. [6] G. Bartzas, K. Komnitsas, Solid phase studies and geochemical modelling of low-cost permeable reactive barriers. J. Hazard. Mater. 2010, 183, 301–308. [7] L. Li, C.H. Benson, Evaluation of five strategies to limit the impact of fouling in permeable reactive barriers. J. Hazard. Mater. 2010, 181, 170–180. [8] R.W. Gillham, 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, China, 16 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 September 8-10, 2009"; Y. Chen, X. Tang, L. Zhan (Eds); Springer Berlin/London (2010) 5–15. [9] S. Comba, A. Di Molfetta, R. Sethi, A Comparison between field applications of nano-, micro-, and millimetric zero-valent iron for the remediation of contaminated aquifers. Water Air Soil Pollut. 2011, 215, 595–607. [10] M. Gheju, Hexavalent chromium reduction with zero-valent iron (ZVI) in aquatic systems. Water Air Soil Pollut. 2011, 222, 103–148. [11] ITRC, Permeable reactive barrier: Technology update. PRB-5. Washington, D.C.: Interstate Technology & Regulatory Council, PRB: Technology Update Team (2011). www.itrcweb.org (access: 09.03.2012). [12] R.W. Puls, Long-term performance of permeable reactive barriers: lessons learned on design, contaminant treatment, longevity, performance monitoring and cost – an overview. In: I. Twardowska et al. (eds.), Soil and Water Pollution Monitoring, Protection and Remediation, 3–23 (2006) 221–229. [13] S.J. Morrison, P.S. Mushovic, P.L. Niesen, Early breakthrough of molybdenum and uranium in a permeable reactive barrier. Environ. Sci. Technol. 2006, 40, 2018–2024. [14] S. Caré, R. Crane, P.S. Calabro, A. Ghauch, E. Temgoua, C. Noubactep, Modelling the permeability loss of metallic iron water filtration systems. Clean – Soil, Air, Water 2012, doi: 10.1002/clen.201200167. [15] Y. Zhang, R.W. Gillham, Effects of gas generation and precipitates on performance of Fe° PRBs. Ground Water 2005, 43, 113–121. [16] A.D. Henderson, A.H. Demond, Impact of solids formation and gas production on the permeability of ZVI PRBs. J. Environ. Eng. 2011, 137, 689–696. 17 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 [17] L. Carniato, G. Schoups, P. Seuntjens, T. Van Nooten, Q. Simons, L. Bastiaens, Predicting longevity of iron permeable reactive barriers using multiple iron deactivation models. J. Contam. Hydrol. 2012, doi:10.1016/j.jconhyd.2012.08.012. [18] A.S. Ruhl, N. Ünal, M. Jekel, Evaluation of two-component Fe(0) fixed bed filters with porous materials for reductive dechlorination. Chem. Eng. J. 2012, 209, 401–406. [19] P.D. Mackenzie, D.P. Horney, T.M. Sivavec, Mineral precipitation and porosity losses in granular iron columns. J. Hazard. Mater. 1999, 68, 1–19. [20] D. Sarr, Zero-valent-iron permeable reactive barriers - how long will they last? Remediation 2001, 11, 1–18. [21] E. Bi, J.F. Devlin, B. Huang, Effects of mixing granular iron with sand on the kinetics of trichloroethylene reduction. Ground Water Monit. Remed. 2009, 29, 56–62. [22] S. Ulsamer, A model to characterize the kinetics of dechlorination of tetrachloroethylene and trichloroethylene by a zero valent iron permeable reactive barrier. Master thesis, Worcester Polytechnic Institute, 2011, 73 pp. [23] K. Miyajima, Optimizing the design of metallic iron filters for water treatment. Master thesis, University of Göttingen, 2012, 70 pp. [24] K. Miyajima, C. Noubactep, Effects of mixing granular iron with sand on the efficiency of methylene blue discoloration. Chem. Eng. J. 2012, 200–202, 433–438. [25] C. Su, R.W. Puls, In situ remediation of arsenic in simulated groundwater using zerovalent iron: laboratory column tests on combined effects of phosphate and silicate. Environ. Sci. Technol. 2003, 37, 2582–2587. [26] P. Westerhoff, J. James, Nitrate removal in zero-valent iron packed columns. Water Res. 2003, 37, 1818–1830. [27] D.I. Kaplan, T.J. Gilmore, Zero-valent iron removal rates of aqueous Cr(VI) measured under flow conditions. Water Air Soil Pollut. 2004, 155, 21–33. 18 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 [28] S. Bang, G.P. Korfiatis, X. Meng, Removal of arsenic from water by zero-valent iron. J. Hazard. Mater. 2005, 121, 61–67. [29] O.X. Leupin, S.J. Hug, A.B.M. Badruzzaman, Arsenic removal from Bangladesh tube well water with filter columns containing zerovalent iron filings and sand. Environ. Sci. Technol. 2005, 39, 8032–8037. [30] K. Komnitsas, G. Bartzas, K. Fytas, I. Paspaliaris, Long-term efficiency and kinetic evaluation of ZVI barriers during clean-up of copper containing solutions. Miner. Eng. 2007, 20, 1200– 1209. [31] A.M. Gottinger, Chemical-free arsenic removal from potable water with a ZVI-amended biofilter, Master thesis, University of Regina (Saskatchewan, Canada) (2010) 90 pp. [32] J. Li, Y. Li, Q. Meng, Removal of nitrate by zero-valent iron and pillared bentonite. J. Hazard. Mater. 2010, 174, 188–193. [33] N. Moraci, P.S. Calabrò, Heavy metals removal and hydraulic performance in zero-valent iron/pumice permeable reactive barriers. J. Environ. Manage. 2010, 91, 2336–2341. [34] C. Noubactep, An analysis of the evolution of reactive species in Fe0/H2O systems. J. Hazard. Mater. 2009, 168, 1626–1631. [35] C. Noubactep, S. Caré, Designing laboratory metallic iron columns for better result comparability. J. Hazard. Mater. 2011, 189, 809–813. [36] C. Noubactep, E. Temgoua, M.A. Rahman, Designing iron-amended biosand filters for decentralized safe drinking water provision. Clean: Soil, Air, Water 2012, 40, 798–807. [37] R.L. Johnson, R.B. Thoms, R.O'B. Johnson, T. Krug, Field evidence for flow reduction through a zero-valent iron permeable reactive barrier. Ground Water Monit. Remed. 2008, 28, 47–55. 19 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 [38] R.L. Johnson, R.B. Thoms, R.O'B. Johnson, J.T. Nurmi, P.G. Tratnyek, Mineral precipitation upgradient from a zero-valent iron permeable reactive barrier. Ground Water Monit. Remed. 2008, 28, 56–64. [39] S. Tuladhar, L.S. Smith, SONO filter: An excellent technology for save water in Nepal. SOPHEN 2009, 7, 18–24. [40] K.H. Head, G.P. Keeton, Permeability, shear strength & compressibility tests, In: Manual of Soil Laboratory Testing, vol. 2. Whittles Publishing: United Kingdom, 2008. [41] APHA, AWWA, WEF, Standard Methods for the examination of water and wastewater, 21st ed. American Public Health Association, Washington D.C. (USA), 2005. [42] S. Caré, Q.T. Nguyen, V. L'Hostis, Y. Berthaud, Mechanical properties of the rust layer induced by impressed current method in reinforced mortar. Cement Concrete Res. 2008, 38, 1079–1091. [43] Y. Zhao, H. Ren, H. Dai, W. Jin, Composition and expansion coefficient of rust based on X- ray diffraction and thermal analysis. Corros. Sci. 2011, 53, 1646–1658. [44] C. Noubactep, Processes of contaminant removal in “Fe0–H2O” systems revisited. The importance of co-precipitation. Open Environ. J. 2007, 1, 9–13. [45] C. Noubactep, A critical review on the mechanism of contaminant removal in Fe0–H2O systems. Environ. Technol. 2008, 29, 909–920. [46] C. Noubactep, The fundamental mechanism of aqueous contaminant removal by metallic iron. Water SA 2010, 36, 663–670. [47] C. Noubactep, The suitability of metallic iron for environmental remediation. Environ Prog. Sustain Energy 2010, 29, 286–291. [48] C. Noubactep, Aqueous contaminant removal by metallic iron: Is the paradigm shifting? Water SA 2011, 37, 419–426. 20 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 [49] P.C. Gomes, M.P.F. Fontes, A.G. Silva, E.S. Mendonca, A.R. Netto, Selectivity sequence and competitive adsorption of heavy metals by Brazilian soils. Soil Sci. Soc. Am. J. 2001, 65, 1115–1121. [50] M.P.F. Fontes, P.C. Gomes, Simultaneous competitive adsorption of heavy metals by the mineral matrix of tropical soils. Appl. Geochem. 2003, 18, 795–804. [51] Y.N. Vodyanitskii, The role of iron in the fixation of heavy metals and metalloids in soils: a review of publications. Eurasian Soil Sci. 2010, 43, 519–532. [52] B. Saha, S. Das, J. Saikia, G. Das, Preferential and enhanced adsorption of different dyes on iron oxide nanoparticles: A comparative study. J. Phys. Chem. C 2011, 115, 8024 –8033. [53] T.B. Scott, I.C. Popescu, R.A. Crane, C. Noubactep, Nano-scale metallic iron for the treatment of solutions containing multiple inorganic contaminants. J. Hazard. Mater. 2011, 186, 280– 287. [54] C. Noubactep, S. Caré, K.B.D. Btatkeu, C.P. Nanseu-Njiki, Enhancing the sustainability of household Fe0/sand filters by using bimetallics and MnO2. Clean – Soil, Air, Water 2012, 40, 100–109. [55] A. Hussam, Contending with a Development Disaster: SONO Filters Remove Arsenic from Well Water in Bangladesh. Innovations, 2009, 4, 89–102. [56] D. Pokhrel, B.S. Bhandari, T. Viraraghavan, Arsenic contamination of groundwater in the Terai region of Nepal: An overview of health concerns and treatment options. Environ. Int. 2009, 35, 157–161. [57] P.S. Calabrò, N. Moraci, P. Suraci, Estimate of the optimum weight ratio in zero-valent iron/pumice granular mixtures used in permeable reactive barriers for the remediation of nickel contaminated groundwater. J. Hazard. Mater. 2012, 207–208, 111–116. [58] C. Noubactep, S. Caré, Enhancing sustainability of household water filters by mixing metallic iron with porous materials. Chem. Eng. J. 2010, 162, 635–642. 21 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 [59] B. Courcelles, A. Modaressi-Farahmand-Razavi, D. Gouvenot, A. Esnault-Filet, Influence of precipitates on hydraulic performance of permeable reactive barrier filters. Int. J. Geomech. 2011, 11, 142–151. [60] P.C. Carman, Fluid flow through a granular bed. Trans. Inst.Chem. Eng. 1937, 15, 150–166. [61] W.D. Carrier III, Goodbye, Hazen; Hello, Kozeny-Carman. J. Geotechn. Geoenviron. Eng. 2003, 129, 1054–1056. [62] R.P. Chapuis, Predicting the saturated hydraulic conductivity of sand and gravel using effective diameter and void ratio. Canadian Geotech. J. 2004, 41, 787–795. [63] P.W.J. Glover, E. Walker, Grain-size to effective pore-size transformation derived from electrokinetic theory. Geophysics 2009, 74, E17–E29. [64] Y. Chen, R. Hu, W. Lu, D. Li, Ch. Zhou, Modelling coupled processes of non steady seepage flow and non linear deformation for a concrete-faced rockfill dam. Comput. Struct. 2011, 89, 1333–1351. [65] C. Noubactep, S. Caré, Dimensioning metallic iron beds for efficient contaminant removal. Chem. Eng. J. 2010, 163, 454–460. [66] C. Noubactep, S. Caré, F. Togue-Kamga, A. Schöner, P. Woafo, Extending service life of household water filters by mixing metallic iron with sand. Clean – Soil, Air, Water 2010, 38 951–959. [67] Q. Wang, J.-H. Lee, S.-W. Jeong, A. Jang, S. Lee, H. Choi, Mobilization and deposition of iron nano and sub-micrometer particles in porous media: A glass micromodel study. J. Hazard. Mater. 2011, 192, 1466–1475. [68] C. Shi, J. Wei, Y. Jin, K.E. Kniel, P.C. Chiu, Removal of viruses and bacteriophages from drinking water using zero-valent iron. Separ. Purif. Technol. 2012, 84, 72–78. 22 502 503 504 505 506 507 508 509 [69] N.C. Mueller, J. Braun, J. Bruns, M. Cerník, P. Rissing, D. Rickerby, B. Nowack, Application of nanoscale zero valent iron (NZVI) for groundwater remediation in Europe. Environ. Sci. Pollut. Res. 2011, 19, 550–558. [70] A.S. Ruhl, A. Weber, M. Jekel, Influence of dissolved inorganic carbon and calcium on gas formation and accumulation in iron permeable reactive barriers. J. Contam. Hydrol. 2012, 142–143, 22–32. 23 509 510 511 512 Table 1: Variability of the operational conditions for column experiments as illustrated by the dimension of the column (H, D), the nature of contaminants (X), the nature of additives, the Fe0 mass, and the relative dimension of particles (δ). X H D Additive Fe 0 dFe(∗) dadditive(∗) δ (∗∗) Ref. (cm) (cm) (g) (mm) (mm) (-) TCE. PCE 50 3.8 Sand n.a. 0.25 1.315 0.19 [1] As 31 2.5 Sand 75 0.42 0.275 0.65 [25] NO3- 30 5 Sand 1636 0.3 n.a. (-) [26] Cr 10 2.5 Sand n.a. 1.45 0.638 0.44 [27] As 17.8 5.1 Sand 400 0.15 0.4 0.38 [28] As 4 0.1 Sand 1.5 n.a. 0.5 (-) [29] Cu 45 5 Sand 525 0.7 0.8 0.88 [30] TCE 40 1.59 Sand 80 1.355 0.118 0.09 [21] As n.a. n.a. Sand n.a. 0.5 0.5 1.0 [31] NO3- 20 2 Sand 90 0.1 0.55 0.19 [32] Cu. Ni 50 5 Pumice 240 0.5 0.30 0.6 [33] Cu. Ni, Zn 50 5 Sand 200 0.5 0.28 0.6 present Cu. Ni, Zn 50 5 Pumice 200 0.5 0.30 0.6 present 513 514 515 516 517 (∗) dFe and dadditive are the the mean grain size of Fe0 and the additive respectively. (∗∗) δ is the diameter ratio of the smaller particle size to the larger one (Fe0 or additive). 24 517 518 Table 2: Characteristics of used materials tested by Mercury Intrusion Porosity (MIP). Fe0 Pumice Sand Specific weight ρs (g/cm3) 7.78 1.92 2.60 Apparent specific weight ρas (g/cm3) 7.78 1.14 2.60 Compactness C(-) 0.51 0.45 0.55 Inter particular porosity Φinter (%) 49.6 54.8 45.0 Intra particular porosity ϕpp (%) - 41.0 - Porosity Φ0 (%) 49.6 73.3 45.0 519 520 25 520 521 Table 3: Main characteristics of the studied columns. “rzeff” is the measured reactive zone. The estimated porosity is also given. System Media Fe0 Fe0 Additive Duration rzeff. Porosity* (vol %) (g) (g) (days) (cm) (%) A sand 0.0 0.0 1060 28 40.0 45.05 B pumice 0.0 0.0 27 28 2.6 72.6 C Fe0 100.0 200 0.0 17** 2.6 49.6 D Fe0 + sand 50.0 200 76.4 15** 5.2 46.05 E Fe0 + pumice 50.0 200 27.0 28** 5.0 59.5 * in this values the internal porosity of the pumice is also included. 522 523 524 525 526 527 **-marked systems were stopped because of excessive permeability loss. 26 527 Table 4: Magnitude of contaminant removal in investigated systems. System min (mg) E (%) Es (mg/g) Ni Cu Zn Ni Cu Zn Ni Cu Zn A 685 685 927 57.3 99.8 65.7 n.a. B 685 685 927 57.2 51.5 52.1 n.a. C 367 367 497 94.7 99.9 99.8 1.74 1.83 2.48 D 367 367 497 99.3 99.8 99.9 1.82 1.83 2.48 E 612 612 828 93.3 99.9 99.6 2.86 3.06 4.12 528 529 530 27 530 Table 5: Modelled characteristics of Fe0-containing columns. The extent of Fe0 depletion (%) is given by: V VV100consumedFe 0 t0 −= .% where (V0 –Vt) is the consumed Fe0 volume with V0 the initial volume of Fe 531 532 533 0 and Vt its residual Fe0 at time t. System tlimit (*) Fe0 depletion Fe 0 residual Φ/Φ0 (**) Es.eff (mg/g) (***) (days) (%) (g) (%) Ni Cu Zn Fe0 7.5 17.1 168.8 4.6 5.0 5.0 6.8 Fe0:sand 10.0 33.2 133.6 2.0 4.0 4.1 5.5 Fe0:pumice 16.0 41.4 117.2 1.4 4.3 4.4 6.0 534 535 536 537 538 539 540 (*) “tlimit” is the time which corresponds to a constant hydraulic conductivity. (**) Φ/Φ0 is the residual porosity. Φ0 is the initial porosity of the column. (***) Es,eff is the specific removal. 28 540 541 542 543 544 545 546 547 548 549 550 551 552 Figure captions Figure 1: Schematic diagram of the experimental design. Used materials were (i) Fe0 (0 or 200 g), (ii) pumice (0 or 27 g), and (iii) quartz gravel (0 or 1060 g). Figure 2: Time-dependant evolution of the pH value (a) and the iron concentration (b) of column effluent. The lines are not fitting functions, they simply connect points to facilitate visualization. Figure 3: Magnitude of Cu (a), Ni (b), and Zn (c) breakthrough from the columns containing Fe0. The lines are not fitting functions, they simply connect points to facilitate visualization. Figure 4: Time-dependant evolution of the hydraulic conductivity in the three systems containing Fe0 and the reference system pumice: (a) experimental K/K0 values; (b, c and d) relative permeability K/K0, experimental and modelled values. η = 6.4, model 1: t depletion,∞ = 50 days, and model 2: 553 t depletion,∞ = 62.5 days. For the three Fe0-systems, the value at time tlimit has been adjusted so that the residual content of Fe 554 555 556 557 0 is slightly superior to the Fe0 volume which is not consumed at clogging.