Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Hydrometallurgy 97 (2009) 221227

Contents lists available at ScienceDirect

Hydrometallurgy
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / h yd r o m e t

The dissolution of scorodite in gypsum-saturated waters: Evidence of CaFeAsO4 mineral formation and its impact on arsenic retention
Marie-Claude Bluteau, Levente Becze, George P. Demopoulos
Department of Mining and Materials Engineering, McGill University, Montreal, Canada H3A 2B2

a r t i c l e

i n f o

a b s t r a c t
This paper reports the results of an investigation on scorodite behavior in a gypsum-saturated aqueous environment in terms of arsenic release and solid phase transformation. The experiments were conducted at xed pH (59) and temperature (22 C and 75 C). The apparent equilibrium arsenic concentrations obtained in a gypsum-saturated system were lower than in a pure (gypsum-free) system. At pH 7 and 22 C, the apparent equilibrium arsenic concentration was 3.6 mg/L in the presence of gypsum compared to 5.9 mg/L in a pure system. At higher temperatures and pH values, this effect was even more signicant. One possible explanation for this reduced arsenic concentration level in the presence of gypsum may be the formation of calciumiron(III)arsenate compounds. Experimental evidence for this was obtained at 75 C and pH 7 and 9 where scorodite was found to partly transform to yukonite (Ca2Fe3(AsO4)4(OH)12H2O). Crown Copyright 2009 Published by Elsevier B.V. All rights reserved.

Article history: Received 14 December 2008 Received in revised form 16 March 2009 Accepted 16 March 2009 Available online 26 March 2009 Keywords: Arsenic Scorodite Yukonite Gypsum Dissolution Solubility Stability Tailings Porewaters

1. Introduction Arsenic is a major contaminant in a number of non-ferrous ores. Upon processing of these ores, it reports to aqueous efuents or smelter gases from which it is removed and disposed as a waste. The metallurgical industry has been challenged to nd arsenic-carriers having low solubility and high long-term stability (Riveros et al., 2001). Scorodite (FeAsO42H2O), a crystalline ferric arsenate found in nature, has been proposed as potentially good carrier for the xation of arsenic in the case of arsenic-rich and iron-decient process solutions (Demopoulos, 2005). Scorodite has a high arsenic content, a stoichiometric (Fe/As molar ratio = 1) iron demand and excellent dewatering characteristics. It can be either produced by hydrothermal precipitation at temperatures higher than 150 C (Swash and Monhemius, 1994) or by controlled precipitation at 95 C under atmospheric pressure (Filippou and Demopoulos, 1997; Demopoulos et al., 2003; Singhania et al., 2005; Singhania et al., 2006). In addition, scorodite or its amorphous/poorly crystalline precursors may be formed in the case of mine tailings produced by neutralization of hydrometallurgical process liquors rich in arsenic (N600 mg/L) (Langmuir et al., 1999, 2006). Hence, the behavior and long-term stability of scorodite in lime-neutralized tailings is of great environmental interest.

The long-term stability (over 1 year equilibrium) of scorodite in a pure (gypsum-free) aqueous system in terms of solubility and kinetics of dissolution was discussed recently by Bluteau and Demopoulos (2007). Some scorodite dissolution rate data was also reported by Harvey et al. (Harvey et al., 2006) but in this study very short leaching time (4 h only) was employed and no characterization of phases was attempted. According to the former study (Bluteau and Demopoulos, 2007), the incongruent dissolution of scorodite in neutral or slightly alkaline waters at 22 C is extremely slow reaching apparent equilibrium after several months with a colloidal 2-line ferrihydrite phase (FeOOH0.5H2O) (Majzlan et al., 2004) as per following reaction: FeAsO4 2H2 O + OH

= FeOOH 0:5H2 O + H2 AsO4 + 0:5 H2 O: 1

Corresponding author. Tel.: +1 514 398 2046; fax: +1 514 398 4492. E-mail address: george.demopoulos@mcgill.ca (G.P. Demopoulos).

At neutral pH and 22 C the apparent equilibrium arsenic concentration was found to be 5.9 mg/L. Because of the widespread use of lime or limestone for neutralisation of the typically sulphate-containing acidic efuents (Langmuir et al., 1999; Moldovan et al., 2003; Moldovan and Hendry, 2005; Jia and Demopoulos, 2005, 2008) the porewaters of mining and metallurgical tailings ponds are gypsum-saturated. This raises the possibility of scorodite interaction with gypsum (or its constituents Ca2+ or SO2). Therefore, the stability of scorodite, in such environ4 ment needs to be investigated. No similar study has been undertaken before. There exists, however, evidence in literature that suggests calcium ions introduced in the tailings generation process either as

0304-386X/$ see front matter. Crown Copyright 2009 Published by Elsevier B.V. All rights reserved. doi:10.1016/j.hydromet.2009.03.009

222

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227

Table 1 Summary of all dissolution experiments and equilibrium As concentration data. Temperature (C) Nominal pH Equilibrium pH Gypsum (g/L) Name Duration (weeks) As concentration (mg/L) 22 5 4.99 30 G225 56 0.28 7 6.98 30 G227 57 3.55 9 7.93 30 G229 57 115 5 5.01 0 225 65 0.35 7 6.99 0 227 66 5.89 9 8.98 0 229 66.4 386 75 5 4.99 20 G755 20.7 1.18 7 6.80 20 G757 20.7 272 9 20 G759 20.7 2580 5 4.99 0 755 9 3.94 7 6.97 0 757 32.4 384 9 9.04 0 759 4.6 6816

No apparent equilibrium reached, All scorodite dissolved.

lime or gypsum contribute to lowering the aqueous concentration of arsenic when the latter is removed from solution by co-precipitation or adsorption with ferric iron. Thus Harris and Monette (Harris and Monette, 1987) and Krause and Ettel (Krause and Ettel, 1989) found Ca (OH)2 to be superior to NaOH when used as a base in Fe(III)As(V) (Fe/As = 4) co-precipitation from sulphate solutions at pH N 5 as it was found to yield lower residual arsenic concentrations. No explanation for this behaviour was offered, although reference to possible formation of unidentied calcium arsenate phases was made (Krause and Ettel, 1989). Later, Emett and Khoe (Emett and Khoe, 1994), who studied the co-precipitation of Fe(III), As(V) and Ca(II) by neutralisation with NaOH, found similarly Ca2+ to lower signicantly the dissolved arsenate concentration in neutral-to-alkaline pH range conditions. The co-adsorption of Ca2+ with arsenate on ferrihydrite was one of the mechanisms advanced by the authors explaining the enhanced retention of arsenic by ferrihydrite. A similar mechanism was postulated by Jia and Demopoulos (Jia and Demopoulos, 2005) who observed soluble calcium sulphate to enhance the uptake of arsenate by ferrihydrite previously prepared from sulphate solutions. Emett and Khoe (Emett and Khoe, 1994), however, did not rule out the possibility of Ca2+ ions to have contributed to enhanced arsenic removal via surface transformation of the arsenate-bearing ferrihydrite to a calcium-iron arsenate phase, such as yukonite (Ca2Fe3 (AsO4)4(OH)12H2O), lazarenkoite (CaFeAs3O7H2O), arseniosiderite (Ca 2 Fe3(AsO 4) 3O 23H2O) or kolfanite (Ca2 Fe3 O2 (AsO 4 )3 2H2 O). Unfortunately, the solids were not characterized so it was not possible to validate this mechanism. Recently, Paktunc et al. (Paktunc et al., 2003, 2004), using a variety of characterization techniques, identied the presence of CaFe arsenate phases in gold ores and cyanidation tailings along with scorodite and arsenic-bearing ferric oxyhydroxide. The cyanidation experiments, aiming to simulate industrial practice, used lime (CaO) to maintain the pH. Interestingly enough, in several ore samples, the authors found a replacement of scorodite by CaFe arsenate phases. Arseniosiderite and yukonite were among the mineral phases identied. It is the object of this work to assess the stability of scorodite in gypsum-saturated waters and present for the rst time evidence of the formation of a CaFeAsO4 phase contributing to lowering the aqueous arsenic concentration level. To this end, extended (up to 57 weeks) scorodite dissolution experiments were performed in gypsum-saturated environments in the pH 59 region at ambient (22 C) and elevated (75 C) temperature and compared to recently published dissolution data in gypsum-free waters (Bluteau and Demopoulos, 2007). Elevated temperature was included with the purpose of accelerating the equilibration process and facilitating the production of easily recognizable crystalline phases. 2. Materials and methods The scorodite used in this work was produced following the hydrothermal method described by Dutrizac and Jambor (Dutrizac and Jambor, 1988). One liter of 0.3 M As(V) and 0.3 M Fe(III) solution was prepared from single-distilled water, hydrated arsenic pentoxide and hydrated ferric nitrate. The solution was heated to 160 C in autoclave while continuously stirred for 24 h. After cooling, the

scorodite was ltered out of the solution on a 0.2 m pore size membrane using a pressure ltration unit, washed 4 times using 1 L of boiling distilled water, dried at 110 C for 2 h and at 50 C for 2472 h. Following scorodite synthesis a product cleaning method was used prior to dissolution experiments in order to remove any soluble arsenate material co-precipitated with scorodite. This method involved the equilibration of scorodite in pH 59 adjusted water for several 24 h periods until a constant arsenic concentration was obtained (Bluteau and Demopoulos, 2007). The dissolution experiments were carried out using deionized water (resistivity N18 Mcm 1) in 250 mL Erlenmeyer asks sealed with rubber stoppers. The slurries were agitated using 38 mm length magnetic bars. For the 75 C experiments, the asks were placed in a water bath maintained at 75 0.2 C by an immersion circulator. The rst stage of the experiments consisted of a CaSO4 equilibration period. During this period 200 mL of water at the designated pH value and temperature was saturated with CaSO4 by the addition of 4 6 g gypsum (reagent grade). The slurry was allowed to equilibrate for 1 week. During this period, the pH was frequently adjusted; samples were taken every day and ltered using 0.1 m pore size syringe lters. After 1 week, the pre-cleaned scorodite was added (5 g on dry basis). The pH was adjusted many times a day. pH adjustment typically involved addition of 0.55 N NaOH solutions as the dissolution reaction (refer to Eq. (1)) involved OH consumption. Four weeks before the end of the experiments, the pH was let to adjust by itself. Slurry samples of 3 mL were retrieved at predetermined times using a syringe and ltered with 0.1 m or 0.025 m pore size syringe membrane lters. At the end of the experiments, the slurries were ltered as well using 0.1 or 0.025 m pore size membrane and the nal solid material was split into three fractions. Part of it was not washed, part of it was water-washed and part of it was acetonewashed. The solids were dried for at least seven days at room temperature. The experiments in a pure system (Bluteau and Demopoulos, 2007) were conducted using the same procedure but this time no CaSO4 was used.

Fig. 1. As Concentrations as a Function of Time for G22-5, G22-7, 22-5 and 22-7 Experiments (closed symbols correspond to 0.1 m while open symbols to 0.02 or 0.025 m lter pore size) (for Test No identication refer to Table 1).

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227

223

Fig. 2. As Concentration as a Function of Time for G22-9 and 22-9 Experiments and Solution Ca/S Molar Ratio for G22-9 Experiment (for Test No identication refer to Table 1).

Fig. 3. XRD Pattern of the Final Solids of the Experiment G22-9 (Not Washed).

All solutions were analysed for As, Fe and Na concentrations by ICP-AES using a Thermo Jarrel Ash Trace Scan machine. Total inorganic carbon concentrations were measured for some solution samples using a TOC Analyzer Rosemount Dohrmann DC-80 equipped with a non-dispersive infrared detector. 2-mL samples were taken, ltered on 0.025-m pore size membranes and quickly analyzed. The chemical composition of the nal solids was determined by HCl digestion followed by ICP-AES analysis. All solids were analyzed by XRay diffraction using a Rigaku Rotaex D-Max diffractometer with CuK1 radiation at 40 kV and 150 mA. The particle size distribution of the starting scorodite material was determined using a HORIBA LA920 Particle Size Analyzer. Some selected samples were analyzed by a TGA Perkin-Elmer thermogravimetric analyzer model TGA-7. 3. Results and discussion Table 1 gives a summary of all dissolution experiments. It is noted that the rst letter (G) of the test name indicates gypsum-saturated aqueous media followed by the temperature and nominal pH of the solution. As an example, G22-5 corresponds to gypsum-saturated solution, 22 C and pH 5. Experiments which were performed in gypsum-free solutions have no G letter. Experiments were run up to 66 weeks. Most of the experiments reached apparent equilibrium in terms of As concentration except the experiments run at pH 9. 3.1. Characterization of the starting material The material produced for use in this study had the characteristic light green colour of scorodite and was composed of 33.0 wt.% As and 24.5 wt.% Fe corresponding to a Fe/As molar ratio of 1.0. During TGA analysis, 15.6% weight was lost between 170 C and 225 C, which corresponds exactly to 2 mol of crystallization water. The mean particle size by volume was 1.98 0.81 m. The comparison of the
Table 2 Chemical composition of the nal solids. Experiment G22-5 G22-7 G22-9 G75-5 G75-7 G75-9 Washing W A W A W A W W W As (wt.%) 23.6 18 24.3 17.5 22.7 17.7 22.3 20.5 20.4 Fe (wt.%) 16.7 12.3 17 12.8 16 13.4 16.5 15.6 19.2 Na (wt.%) 0.3 0.2 0 0.3 0.2 0.7 0.4 0.1 0.6

XRD pattern of the solid with the standard 26-0778 from JCPDS conrmed that the material was highly crystalline scorodite. The scorodite particles consisted of agglomerates of small crystallites of 0.1 to 0.3 m in size. The XRD pattern along with SEM pictures of the hydrothermal scorodite material used in this work may be found elsewhere (Bluteau and Demopoulos, 2007). 3.2. The scoroditegypsumwater system at 22 C Table 1 gives the nal arsenic concentration as a function of pH. All the experiments performed in gypsum-saturated media exhibited lower arsenic solubilities than those obtained in the pure system. The aqueous iron concentration was extremely low in all samples of all experiments. It was usually lower than 0.3 mg/L and never higher than 5 mg/L. This indicates that it reprecipitated as soon as it dissolved. In the gypsum-free system (Bluteau and Demopoulos, 2007), iron was found to reprecipitate as nano-sized ferrihydrite particles. Colloidal ferrihydrite was apparently stabilized in that case by the presence of arsenate in solution due to the adsorption of the latter. Figs. 1 and 2 give the As concentration as a function of time for the experiments run at pH 5 and 7 (the rst Figure) and pH 9 (the second Figure). They show that the As concentration increased with pH and the dissolution rate decreased with time. When compared to the gypsum-free experiments, the As concentration was slightly lower in the gypsum-saturated system. At the end of the experiments, the solution Ca/S molar ratio was 0.91 at pH 5 and 0.83 at pH 7. Similar but more accelerated behavior was observed in the results obtained at pH 9 (refer to Fig. 2). In this case, the Ca/S molar ratio in solution decreased with time down to ~0.1. This suggests the precipitation of calcium either with arsenic (as calcium arsenate) or arsenic and iron (as calcium-iron-arsenate) or CO2 (as CaCO3). The nal solids of the experiments were subjected to washing either with water or acetone. The latter was used for the purpose of

Ca (wt.%) 4.3 12.3 5.2 11.5 11.9 12.5 3.2 8.9 9.4

S (wt.%) 4.4 10.1 6.4 9.9 0.3 2.8 0.1 0.1 0.2

Fe/As mol 0.95 0.91 0.94 0.98 0.95 1.02 1 1.02 1.26

Ca /S mol 0.79 0.97 0.65 0.93 31.15 3.57 25.6 63.81 39.8

As/Ca mol 2.93 0.78 2.47 0.81 1.02 0.76 3.72 1.23 1.16

W: Water washed, A: Acetone washed.

224 Table 3 Composition of the solutions used in chemical modelling. Experiment G22-5 Sample 1 2 3 1 2 3 1 2 3 Time (weeks) 52 54 56 53 55 57 53 55 57 pH 4.99 4.99 4.99 7.11 7.00 6.98 9.02 8.18 7.93

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227

pE 9.78 10.56 9.52 6.99 7.21 8.03 7.09 6.98 8.08

Concentrations (mM) As 0.0034 0.0041 0.0037 0.0489 0.0538 0.0474 1.2453 1.6217 1.5383 Fe 0.001214 0.001432 0.000732 0.000032 0.000025 0.000247 0.000014 0.001808 0.000412 Ca 15.44 15.66 16.69 15.82 15.39 15.93 10.76 10.62 10.82 S 17.18 16.98 18.28 18.61 17.57 19.09 90.75 85.08 89.75 Na 2.002 2.140 1.975 6.988 7.782 7.782 128.08 151.20 135.39 C from inorganic CO2 3

0.027

G22-7

0.070

G22-9

2.528

Calculated from the ORP measurements.

preserving water-soluble phases such as gypsum. Table 2 shows the chemical composition of the nal solids. As evidenced by the lowering of the Ca/S ratio, water washing caused partial dissolution of the CaSO4 phase at all pH values at 22 C. This led to an enrichment of the solid material in As and Fe. The low sulphur content and high calcium content of the nal solid of G22-9 conrmed the hypothesis that much of the gypsum had dissolved and that the calcium had re-precipitated under a different form such as a Ca-arsenate compound. XRD patterns of the nal solids of experiments G22-5 and G22-7 indicated that the solids were crystalline and revealed only the presence of scorodite and gypsum. The XRD pattern of the nal unwashed solid material of experiment G22-9 (see Fig. 3) revealed in addition to the presence of scorodite and gypsum (peaks slightly moved to the left) some weak peaks indicative of one or more minor unknown phases. No CaAsO4 or CaCO3 phase could be identied. The high background is indicative of one or more amorphous or poorly crystalline components such as ferrihydrite (Bluteau and Demopoulos, 2007) or unknown CaAsO4 containing phases. 3.3. Chemical modelling To verify the attainment of equilibrium or not in the above described dissolution experiments, pH was let to adjust by itself over a period of 4 weeks before termination of the tests. Solution samples were taken 4 and 2 weeks before the end and at the end of the experiments. Solution analyses including inorganic carbon concentration are summarized in Table 3. These solutions were modelled in terms of attainment of equilibrium for a number of phases making use of the geochemical code PHREEQC (Parkhust and Appelo, 1999) and a modied thermodynamic database for the arsenic species recently published by Langmuir et al. (2006). It is noted that pE is a measure of electron activity calculated from the ORP values (pE = Eh/0.059). For all solution compositions, Fe3+ and AsO3 activities were rst 4 calculated as a function of pH followed by calculation of log (IAP (Ion Activity Product)) values for scorodite and ferrihydrite taken here as ferric hydroxide (i.e. log (aFe3+ (aOH)3)). The Saturation Index (SI) of scorodite was calculated using the most recent value of KSP provided by Langmuir et al. (2006). The Saturation Index was not

calculated for ferric hydroxide, as it is not a crystalline compound with a constant composition. Table 4 shows the calculated average values and standard deviations for the three G22-5 samples and the individual values for G22-7 and G22-9 samples. The Saturation Indexes (SI) for scorodite, gypsum, calcite (CaCO3), and johnbaumite (Ca5(AsO4)3OH) are shown as well. It can be seen from the table that all solutions were in equilibrium with respect to gypsum as all SI values were close to 0. The solution log (IAP) for scorodite at pH 5 is close to the log (KSP) reported by Langmuir et al. (2006) which is 25.83. For the experiments G22-7 (pH 7) and G22-9 (pH 9 8) the log (IAP) for ferric hydroxide (ferrihydrite) increased with the pH value from 37.9 at pH 5 to 36.5 at pH 9. Similar increase for the apparent log KSP of Fe(OH)3 was determined by Langmuir et al. (2006), namely 39.49 at pH 2.18 to 33.5 at pH 7.37. The Saturation Index for calcite was calculated. The solutions at pH 5 and 7 (experiments G22-5 and G22-7) were undersaturated with respect to calcite whereas experiment G22-9 (pH 9 8) was oversaturated with respect to it. Similarly the solution at pH 9 8 of experiment G22-9 was found to be initially supersaturated and at the end to be in equilibrium with jonhbaumite (Ca5 (AsO4)3OH), a calcium hydroxy-arsenate. The modelling results imply that at least in part at pH N 8 formation of aragonite and/or jonhbaumite could have taken place during the long-term equilibration of scorodite and gypsum. However, as it was noted earlier XRD analysis failed to detect calcium carbonate or arsenate in the G22-9 equilibrium mixture. Similarly the limited dissolution of scorodite at ambient temperature makes the identication of any minor calcium arsenate precipitated phases practically impossible. Under these circumstances, it was decided to perform experiments at elevated temperature (75 C) in order to promote scorodite dissolution and simultaneously favor the production of crystalline phases hence facilitating the characterization of the transformation process. 3.4. Characterization of the transformation process at 75 C Fig. 4 shows the As concentration as a function of time for the 75 C tests at pH 5 and 7 as well as the aqueous Ca/S molar ratio for G75-5 (pH 5) and G75-7 (pH 7). In general, slightly lower terminal arsenic

Table 4 Calculated log (IAP) values for scorodite and ferric hydroxide and SI for scorodite, gypsum, calcite, and johnbaumite. G22-5 Average Sampling time (weeks) pH Log (aFe3+ aAsO3) 4 SI (Scorodite) 3+ Log (aFe (aOH)3) SI (Gypsum) SI (Calcite) SI (Johnbaumite) 5256 4.99 25.1 0.78 37.9 0.04 5.17 21.5 Std Dev 0.17 0.17 0.20 0.02 0.01 0.20 G22-7 1 53 7.11 26.5 0.69 37.7 0.04 1.42 5.39 2 55 7.00 26.6 0.76 38.0 0.02 1.54 5.75 3 57 6.98 25.6 0.23 37.6 0.05 1.57 6.02 G22-9 1 53 9.02 29.1 3.22 37.4 0.03 1.42 2.58 2 55 8.18 24.9 0.95 36.5 0.03 0.69 0.92 3 57 7.93 25.4 0.46 36.5 0.06 0.44 0.00

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227

225

Fig. 4. As Concentration as a Function of Time for G75-5, G75-7, 75-5 and 75-7 Experiments and Solution Ca/S Molar Ratio for G75-5 and G75-7 Experiments (for Test No identication refer to Table 1).

Fig. 6. XRD Pattern of the Final Solid of the Experiment G75-7 (Water Washed).

concentrations were reached in the gypsum-saturated system vis-vis the pure system. The specic As concentration data can be evaluated in Table 1. Referring to Fig. 4 it is interesting to note the variation of the Ca/S molar ratio with time. The Ca/S molar ratio at pH 5 was stable at around 0.95 for the 8 rst weeks and then decreased abruptly down to 0.65. As this variation did not impact on As concentration, it is unlikely that a CaAs phase precipitated. At pH 7, the As concentration increased initially more slowly in the gypsumsaturated system than in the pure system. The Ca/S ratio decreased abruptly from the beginning of the experiment and stabilized at a value of 0.1. When the ratio stabilized, the As concentration stabilized as well. It is likely that a CaAs or CaFeAs phase precipitated in that experiment. Fig. 5 shows the As concentration as a function of time at pH 9 as well as the Ca/S molar ratio for G75-9. In the gypsum-saturated system, the arsenic concentration increased less rapidly and stabilized at a value 4 times lower than in the pure system. The difference could have been even more signicant if the original scorodite concentration had been higher, as apparently no saturation was reached in the pure system (all the scorodite had dissolved). The aqueous Ca/S molar ratio decreased very rapidly in the gypsum saturated experiment indicating the precipitation of a CaAs or CaAsFe compound. After less than 5 weeks, all the gypsum had dissolved. The composition of the nal solids at 75 C as a function of pH is reported in Table 2. The solids had been washed with water prior to their analysis. It can be seen from these data that, the higher the pH, the higher the Ca content of the washed solids. There was no sulphur

remaining in the solid, indicating that all gypsum had been dissolved. At all pH values, calcium was present in the solids phases, suggesting the precipitation of calcium-arsenate-containing phases. XRD analysis of the washed nal solids from experiments G75-7 (pH 7) and G75-9 (pH 9) revealed the presence of yukonite a calciumiron(III) arsenate phase in addition to scorodite. The relevant XRD patterns proving the presence of yukonite (albeit somewhat poorly crystalline) are shown in Figs. 6 and 7. In other words, the XRD data of this work (Figs. 6 and 7) suggest the transformation of scorodite to yukonite as evidenced by the relative size of the corresponding peaks of scorodite and yukonite. SEM examination revealed ne particle morphology indicative of a product obtained via a dissolution-reprecipitation mechanism (Blesa and Matijevic, 1989). Yukonite is indeed a poorly crystalline CaFeAsO4 phase of composition approaching Ca2Fe3(AsO4)4(OH)12H2O. It has been found in Takish Lake in Yukon, at the Sterling Hill mine, Ogdensburg in New Jersey (Dunn, 1982) and in the Ketza River Mine ore, partly as replacement product of arsenopyrite and scorodite (Paktunc et al., 2003). The latter mineralogical evidence seems to conrm the transformation process identied in this work. 3.5. Implications It has been established previously (Krause and Ettel, 1988; Bluteau and Demopoulos, 2007; Langmuir et al., 2006) that the dissolution of

Fig. 5. As Concentration as a Function of Time for G75-9 and 75-9 Experiments and Solution Ca/S Molar Ratio for G75-9 Experiment (for Test No identication refer to Table 1).

Fig. 7. XRD Pattern of the Final Solid of the Experiment G75-9 (Water Washed).

226

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227

scorodite is incongruent at pH N 3 resulting in the formation of a colloidal (nanosized) ferrihydrite phase (Eq. (1)) stabilized by the adsorption of arsenate ions on it (Eq. (2)) The latter most likely occurs via bidentate-binuclear complexation (Waychunas et al., 1993). A minor fraction of Ca2+ ions may report to the surface of ferrihydrite via co-adsorption (Jia and Demopoulos, 2005). HAsO4
2

+ uFeOH + H

YuFeHAsO4 ads + H2 O:

was found to partly transform to yukonite (Ca2Fe3(AsO4)4(OH) 12H2O) as evidenced clearly by XRD analysis. Although, the experiment was done at elevated temperature that may or may not reect closely reactions taking place at ambient conditions, it can be tentatively concluded that the presence of calcium/gypsum in tailings pore waters signicantly affects the stability of scorodite at pH N 7 causing its partial transformation to yukonite. As a result of this transformation the amount of arsenic that may potentially be released from the scorodite-containing tailings solids is signicantly reduced. Acknowledgment The support of this work by Natural Sciences and Engineering Research Council of Canada (NSERC) via a strategic project grant is gratefully acknowledged, as is the sponsorship of Areva Resources Inc., Barrick Gold Corporation, Cameco, Hatch and Teck Cominco. References
Blesa, M.E., Matijevic, E., 1989. Phase transformations of iron oxides, oxohydroxides and hydrous oxides in aqueous media. Adv. Colloid Interface Sci. 29, 173231. Bluteau, M.C., Demopoulos, G.P., 2007. The incongruent dissolution of scorodite solubility, kinetics and mechanism. Hydrometallurgy 87, 163177. Bothe Jr., J.V., Brown, P.W., 1999. Arsenic immobilization by calcium arsenate formation. Environ. Sci. Technol. 33, 38063811. Demopoulos, G.P., 2005. On the preparation and stability of scorodite. In: Reddy, R.G., Ramachandran, V. (Eds.), Arsenic Metallurgy. TMS, Warrendale, PA, pp. 2550. Demopoulos, G.P., Lagno, F., Wang, Q., Singhania, S., 2003. The atmospheric scorodite process. In: Riveros, P.A., Dixon, D., Dreisinger, D.B., Menacho, J. (Eds.), Copper 2003- Hydrometallurgy of Copper, vol. IV. CIM, Montreal, QC, pp. 597616. Dunn, P.J., 1982. New data for pitticite and a second occurrence of Yukonite at Sterling Hill, New Jersey. Mineral. Mag. 46, 261264. Dutrizac, J.E., Jambor, J.L., 1988. The synthesis of crystalline scorodite, FeAsO4.2H2O. Hydrometallurgy 19, 377384. Emett, M.T., Khoe, G.H., 1994. In: Warren, G. (Ed.), Environmental stability of arsenic bearing hydrous iron oxide. Warrendale, PA, TMS, pp. 153166. Filippou, D., Demopoulos, G.P., 1997. Arsenic immobilization by controlled scorodite precipitation. JOM 49, 5255. Harris, G.B., Monette, S., 1987. The stability of arsenic-bearing residues. In: Reddy, R.G., Hendrix, J.L., Queneau, P.B. (Eds.), Arsenic Metallurgy: Fundamentals and Applications. TMS, Warrendale, PA, pp. 469488. Harvey, M.C., Schreiber, M.E., Rimstidt, J.D., Grifth, M.M., 2006. Scorodite dissolution kinetics: implications for arsenic release. Environ. Sci. Technol. 40, 67096714. Jia, Y.F., Demopoulos, G.P., 2005. Adsorption of arsenate onto ferrihydrite from aqueous solution: inuence of media (sulfate vs. nitrate), added gypsum, and pH alteration. Environ. Sci. Technol. 39, 95239527. Jia, Y.F., Demopoulos, G.P., 2008. Coprecipiation of arsenate with iron (III) in aqueous sulphate media: effect of time, lime as base and co-ions on arsenic retention. Water Res. 42 (3), 661668. Krause, E., Ettel, V.A., 1988. Solubility and stability of scorodite, FeAsO4.2H2O: new data and further discussion. Am. Mineral. 73, 850854. Krause, E., Ettel, V.A., 1989. Solubilities and stabilities of ferric arsenate compounds. Hydrometallurgy 22, 311337. Langmuir, D., Mahoney, J., MacDonald, A., Rowson, J., 1999. Predicting arsenic concentrations in the porewaters of buried uranium mill tailings. Geochim. Cosmochim. Acta 63, 33793394. Langmuir, D., Mahoney, J., Rowson, J., 2006. Solubility products of amorphous ferric arsenate and crystalline Scorodite (FeAsO42H2O) and their application to arsenic behavior in buried mine tailings. Geochim. Cosmochim. Acta 70, 29422956. Majzlan, J., Navrotsky, A., Schwertmann, U., 2004. Thermodynamics of iron oxides: part III. Enthalpies of formation and stability of ferrihydrite (Fe(OH)3), schwermannite ( FeO(OH) 3/4(SO4) 1/8), and -Fe 2O 3 . Geochim. Cosmochim. Acta 68 (5), 10491059. Moldovan, B.J., Hendry, M.J., 2005. Characterizing and quantifying controls on arsenic solubility over a pH range of 111 in a uranium mill-scale experiment. Environ. Sci. Technol. 39, 49134920. Moldovan, B.J., Jiang, D.T., Hendry, M.J., 2003. Mineralogical characterization of arsenic in uranium mine tailings precipitated from iron-rich hydrometallurgical solutions. Environ. Sci. Technol. 37, 873879. Paktunc, D., Foster, A., Laamme, G., 2003. Speciation and characterization of arsenic in Ketza river mine tailings using X-ray absorption spectroscopy. Environ. Sci. Technol. 37, 20672074. Paktunc, D., Foster, A., Heald, S., Laamme, G., 2004. Speciation and characterization of arsenic in gold ores and cyanidation tailings using X-ray absorption spectroscopy. Geochim. Cosmochim. Acta 68 (5), 969983. Parkhust, D.L., Appelo, C.A.J., 1999. User's guide to PHREEQC (Version 2), a computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations. U.S. Geol. Surv. Water-Resour. Invest. Rep. 994259. Riveros, P.A., Dutrizac, J.E., Spencer, P., 2001. Arsenic disposal practices in the metallurgical industry. Can. Metall. Q. 40, 395420.

Preliminary calculations published elsewhere (Demopoulos, 2005) have suggested the stability of scorodite in the incongruent dissolution zone to be dependant on the crystallinity of the ferrihydrite phase. Analysis of the experimental data in the present study (refer to Table 4) showed indeed the log (IAP) for ferrihydrite (Fe(OH)3) to increase with pH, i.e. ferrihydrite to become less stable or less crystalline. This implies enhanced stability for scorodite as Langmuir et al. have discussed in detail (Langmuir et al., 2006). It is now reported for the rst time that in addition to the above stabilizing factors the presence of gypsum may inuence the release or retention of arsenic from scorodite solids. Thus evidence obtained (albeit at elevated temperature) in the present work suggests that in the presence of gypsum and at pH 7 partial transformation of scorodite is likely to occur leading to the formation of an apparently less soluble CaFeAsO4 phase such as yukonite. The reactions representing the latter are given below (for simplication the protonation reactions of arsenate ions are omitted): CaSO4 2H2 O = Ca
2+

+ SO4

+ 2H2 O + 2H2 O

3 4 5

FeAsO4 2H2 O = Fe 2 Ca
2+

3+

+ AsO4
3+

+ 4 AsO4

+ 3Fe

+ OH

+ 12 H2 O

= Ca2 Fe3 AsO4 4 OH 12H2 O

Eqs. (2) and (5) are assumed to contribute to the reduction of arsenic concentration in calcium-containing tailings pore waters. Although the occurrence of other reactions, such as the formation of johnbaumite (Ca5(AsO4)3OH), may well contribute to arsenic immobilization (Bothe and Brown, 1999) it is the rst time that arsenic retention via the formation of yukonite-like phases is proposed. 4. Conclusions In this study the dissolution behavior of scorodite in gypsumsaturated waters over the pH range 59 was investigated and compared with its behaviour in the absence of gypsum. During the various long-term (up to 57 weeks) stability tests the concentration of arsenic, calcium and sulphur (being in the form of sulphate) was monitored and the remaining solid phases were characterized. It was found that the release of arsenic from the scorodite solids, while increasing with pH and time as reported previously, was partially mitigated in the presence of gypsum. The benecial effect of gypsum on arsenic retention was more pronounced at pH N 7. Chemical modelling with the aid of PHREEQC suggested that the formation of johnbaumite ((Ca5(AsO4)3OH) may be responsible for the reduction of arsenic concentration at pH 9. However, X-ray diffraction analysis of the equilibrated solids at 22 C failed to nd evidence of this or other calcium arsenate phases. In an effort to isolate and characterize the product of the reaction between scorodite and gypsum that was potentially responsible for the observed reduction in arsenic release, long-term equilibration tests were run at 75 C. At this temperature and pH 7 and 9, scorodite

M.-C. Bluteau et al. / Hydrometallurgy 97 (2009) 221227 Singhania, S., Wang, Q., Filippou, D., Demopoulos, G.P., 2005. Temperature and seeding effects on the precipitation of scorodite from sulphate solutions under atmospheric-pressure conditions. Metall. Mater. Trans., B 36B, 327333. Singhania, S., Wang, Q., Filippou, D., Demopoulos, G.P., 2006. Acidity, valency and thirdion effects on the precipitation of scorodite from mixed sulfate solutions under atmospheric-pressure conditions. Metall. Mater. Trans., B 37 B, 189197.

227

Swash, P.M., Monhemius, A.J., 1994. Hydrothermal precipitation from aqueous solutions containing iron (III), arsenate and sulphate. Hydrometallurgy'94. Chapman & Hall, Cambridge, U.K., pp. 177190. Waychunas, G.A., Rea, C.B.A., Fuller, C., Davis, J.A., 1993. Surface chemistry of ferrihydrite: part 1. EXAFS studies of the geometry of coprecipitated and adsorbed arsenate. Geochim. Cosmochim. Acta 57, 22512269.

You might also like