Sunteți pe pagina 1din 7

Applied Geochemistry 23 (2008) 21232129

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Accurate determination of Fe(II) concentrations in the presence of a very high soluble Fe(III) background
Youri Gendel, Ori Lahav *
Faculty of Civil and Environmental Engineering, Technion Israel Institute of Technology, Haifa, 32000, Israel

a r t i c l e

i n f o

a b s t r a c t
Analytical methods used for determining dissolved Fe(II) often yield inaccurate results in the presence of high Fe(III) concentrations. Accurate analysis of Fe(II) in solution when it is less than 1% of the total dissolved Fe concentration (FeT) is sometimes required in both geochemical and environmental studies. For example, such analysis is imperative for obtaining the ratio Fe(II)/Fe(III) in rocks, soils and sediments, for determining the kinetic constants of Fe(II) oxidation in chemical or biochemical systems operating at low pH, and is also important in environmental engineering projects, e.g. for proper control of the regeneration step (oxidation of Fe(II) into Fe(III)) applied in ferric-based gas desulphurization processes. In this work a method capable of yielding accurate Fe(II) concentrations at Fe(II) to FeT ratios as low as 0.05% is presented. The method is based on a pretreatment procedure designed to separate Fe(II) species from Fe(III) species in solution without changing the original Fe(II) concentration. Once separated, a modied phenanthroline method is used to determine the Fe(II) concentration, in the virtual absence of Fe(III) species. The pretreatment procedure consists of pH elevation to pH 4.24.65 using NaHCO3 under N2(g) environment, followed by ltration of the solid ferric oxides formed, and subsequent acidication of the Fe(II)-containing ltrate. Accuracy of Fe(II) analyses obtained for samples (Fe(II)/FeT ratios between 2% and 0.05%) to which the described pretreatment was applied was >95%. Elevating pH to above 4.65 during pretreatment was shown to result in a higher error in Fe(II) determination, likely resulting from adsorption of Fe(II) species and their removal from solution with the ferric oxide precipitate. 2008 Elsevier Ltd. All rights reserved.

Article history: Received 27 September 2007 Accepted 27 March 2008 Available online 16 May 2008 Editorial handling by R. Fuge

1. Introduction The commonly used analytical methods for determining dissolved Fe(II) concentration (e.g. the phenanthroline method) yield inaccurate results in the presence of a high Fe(III) background (Herrera et al., 1989). The need for a technique that yields an accurate Fe(II) analysis in cases where the Fe(II) concentration is less than 1% of FeT arises particularly in crystallographic and geochemical studies (e.g. for obtaining the ratio between Fe(II) and Fe(III) in rocks, soils and sediments), and in studies concerned with determining the kinetic constants of Fe(II) oxidation in
* Corresponding author. Tel.: +972 4 8292191; fax: +972 4 8228898. E-mail address: agori@techunix.technion.ac.il (O. Lahav). 0883-2927/$ - see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.apgeochem.2008.03.016

chemical or biochemical systems operating at low pH (Tufekci and Sarikaya, 1996; Nemati et al., 1998; Gomez and Cantero, 2003; Molchanov et al., 2007). The same analysis is also needed in environmental engineering, for example in the context of the wide spread ferric-based gas desulphurization processes in which H2S(g) is oxidized to elemental S in an acidic (normally pH < 2) Fe(III) solution, and the generated Fe(II) is concurrently oxidized back to Fe(III) (typically using Acidithiobacillus ferrooxidans bacteria) to allow for steady state continuous operation (Asai et al., 1989; Ebrahimi et al., 2002). In both cases the rate of Fe(II) chemical or biological oxidation is highly dependent on the Fe(III) concentration, which is in many cases 2 or 3 orders of magnitude higher than the Fe(II) concentration (Jones and Kelly, 1983; Nemati and Webb, 1998;

2124

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129

Molchanov et al., 2007). As an example of the difculties encountered with present Fe(II) analysis methods, Molchanov et al. (2007) used batch culture experiments to obtain the extant kinetic constants of Fe(II) oxidation by A. ferrooxidans, that included the effect of the presence of Fe(III) on the Fe(II) oxidation rate. The value of Ks (Monods afnity to substrate coefcient) under the conditions investigated was found to be 31 4 mg Fe(II)/L. Competitive inhibition by Fe(III) was observed and a kinetic equation was developed that included a Fe(III) inhibition term. However, empirical validation of the kinetic model was restricted to Fe(III) concentrations lower than 2.4 g Fe(III)/L (i.e. approximately 2 orders of magnitude higher than the Ks for Fe(II)) due to analytical limitations. At higher Fe(III) concentrations no reliable Fe(II) analysis was available for obtaining accurate kinetic data. To extend kinetic models to higher Fe(III) concentrations, an appropriate technique for the determination of Fe(II) concentrations, which is not compromised by a high background of Fe(III), needs to be developed. In the gas desulphurization process, a similar problem exists: H2S-laden air is bubbled through an acidic aqueous solution of Fe(III). The latter oxidizes H2S to elemental S while Fe(III) is reduced to Fe(II). Typically in such systems the Fe(II) concentration is much lower than the Fe(III) concentration, which is often as high as 910 g/L, because it is designed to allow for optimal scavenging of the H2S(g) present in the biogas stream (Asai et al., 1989). Analysis of Fe(II) using the colorimetric phenanthroline method (www.standardmethods.org) is common although problems are encountered when attempting to attain an accurate Fe(II) measurement in the presence of a high concentration of soluble Fe(III) species. The interference in the method apparently emanates from the formation of color due to a reaction between Fe(III) species and the phenanthroline reagent (Herrera et al., 1989). Another possible interference mechanism was suggested by Muir and Andersen (1977), who postulated that the interference is caused by partial reduction of Fe(III) to Fe(II) by phenanthroline in the presence of sunlight; both sources of error result in overestimation of the Fe(II) concentration. To overcome the problem of over estimation, that was shown, for example, to amount to as much as 38% when the Fe(II) to FeT ratio was $5%, Herrera et al. (1989) suggested the addition of 0.5 M NaF solution as part of the analytical procedure. Fluoride forms a colorless complex with Fe(III) minimizing the Fe(III) concentration available for reaction with the phenanthroline reagent. Another Fe(III) chelating agent that may be used for this purpose is nitrilotriacetic acid (Fadrus and Maly, 1975), however, its application is cumbersome and laborious because color intensity of the analyzed solution must be measured against a sample prepared without the addition of 1,10-phenathroline to compensate for the color of the Fe(III)-NTA complex. Additionally, the mixture of the buffer solution, color reagent (1,10-phenathroline) and Fe(III) masking reagent (NTA solution) must be prepared each time immediately before use and cannot be stored, rendering this technique unsuitable for routine Fe(II) determination. Using the modied uoridephenanthroline technique on synthetic samples Herrera et al. (1989) showed that the error in Fe(II) analysis can be reduced to 2.6% for a Fe(II):FeT ratio of 5%

and to 5.8% when Fe(II) was 0.5% of the total Fe concentration (as compared with a 57% error attained with the standard phenanthroline method for the same sample). After applying the modied method to natural Fe-containing water samples Herrera et al. (1989) concluded that the modied method can be considered accurate (less than 2.4% error, using the authors denition) as long as the Fe(II):FeT ratio is higher than around 5%. Below this ratio the modied phenanthroline method resulted in an overestimated Fe(II) concentration, with an error of around 12% at a Fe(II):FeT ratio of 0.5%. At Fe(II) to FeT ratios lower than 0.5% the error becomes excessive and the method cannot be used (Herrera et al., 1989). A similar uoridemasking approach was proposed by Krishnamurti and Huang (1990) for the spectrophotometric determination of Fe(II) with TPTZ (2,4,6-tri(2-pyridyl)-1,3,5-triazine). These authors showed that when a high uoride concentration is added to the sample Fe(II) analysis can be carried out even when the Fe(II):Fe(III) ratio is as low as 0.1%. However, it was shown that the presence of uoride may increase the rate of Fe(II) oxidation by dissolved O2, especially at pH > 3.45. To prevent this special care should be given to the pH at which the method is carried out (Krishnamurti and Huang, 1990). Inhibition of Fe(II) analysis at high Fe(III) concentrations is not restricted to the phenanthroline and TPTZ methods. The Kolthoff titration method (Herrera et al., 1989), using potassium dichromate as the oxidizing agent and barium diphenylamine-4-sulfonate as indicator, is considered cumbersome and thus less commonly used, despite being somewhat more accurate at low Fe(II) to total dissolved Fe (FeT) ratios. However, Herrera et al. (1989) showed that Kolthoffs titration method also yields accurate results only when the Fe(II) concentration constitutes more than 1% of FeT. Below this ratio the use of this technique is also not recommended. Another well established Fe(II) analysis technique is the ferrozine method (Stookey, 1970), which is based on the reaction of ferrozine (3-(2-pyridyl)-5, 6-diphenyl-1, 2, 4-triazine) with Fe(II) to form a stable magenta-colored complex with maximum absorbance at 562 nm. Ferrozine forms a colored complex with Fe(II) but not with Fe(III). However, when Fe(III) is present in solution (in particular at high concentrations) it tends to oxidize the ferrozine reagent (Hong and Kester, 1986; Viollier et al., 2000; Giokas et al., 2002; Akl et al., 2006). In this reaction Fe(III) is reduced to the Fe(II) form and color intensity increases, resulting in an overestimated Fe(II) concentration. In this paper a method capable of yielding accurate Fe(II) concentrations at Fe(II) to FeT ratios as low as 0.05% is presented. The method is based on a pretreatment procedure designed to separate Fe(II) species from Fe(III) species in solution without changing the original Fe(II) concentration. Once separated, the modied phenanthroline method (Herrera et al., 1989) is used to determine the Fe(II) concentration, practically in the absence of Fe(III). 2. Proposed pretreatment procedure For samples suspected of having a Fe(II):Fe(III) ratio lower than $2%, the following pretreatment is suggested prior to applying the modied phenanthroline method.

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129

2125

The purpose of this step is to reduce the Fe(III) concentration in the sample to a minimum value with as little change as possible in the dissolved Fe(II) concentration. The procedure is based on the fact that ferric species precipitate at a much lower pH than Fe(II) species. Following a pH increase to >pH 4 and precipitation of the Fe(III) species, ltration is used to separate the precipitates from the aqueous phase. The pH range over which the ferric species are precipitated and separated from solution should be wide enough and simple to obtain, but during the procedure ferrous species should neither oxidize nor adsorb to the surface of the separated ferric species. The suggested pretreatment procedure comprises the following steps: pH is rst increased to 4.0 < pH < 4.65 by the addition of 80 g/L NaHCO3 solution. At this pH range Fe(III) species precipitate almost completely as either amorphous Fe(OH)3(s) (Ksp = 1037.1) or as a variety of ferric oxides/oxyhydroxides, depending on the specic solution characteristics (see Section 5). The addition of NaHCO3 solution (a weak base) was chosen for pH adjustment over the use of a strong base (such as NaOH) to avoid the possible formation of localized high pH conditions during base addition. Formation of such conditions may result in either partial oxidation of Fe(II) by dissolved O2 or precipitation of Fe(OH)2, possibly affecting the accuracy of the analysis. In order to further minimize spontaneous oxidation of Fe(II) during NaHCO3 addition the sample was continuously purged by N2(g) to remove dissolved O2 during pH rise. Once pH had stabilized and Fe(III) species precipitated, the solids were immediately separated from solution using a 0.22 lm syringe lter. The ltrate containing Fe(II) was immediately acidied to pH < 2 to avoid possible Fe(II) oxidation and to allow storage before analysis using the phenanthroline method. The experiments described here focus on establishing the optimal pH range at which Fe(III) species should be precipitated without affecting the soluble Fe(II) concentration. In this respect two different phenomena have to be taken into account: (1) spontaneous Fe(II) oxidation by dissolved O2, which is a strongly pH-dependent reaction, should be minimized; and (2) the chosen pH range should be such that possible adsorption of Fe(II) to the surface of the precipitated Fe(III) species (Coughlin and Stone, 1995; Jeon et al., 2001) or co-precipitation of Fe(II) with Fe(III) (Ruby et al., 2003) and its subsequent removal from the sample with the ltered solids, would be negligible. The purpose of the results presented here is to establish the optimal pH range for the pretreatment procedure and to demonstrate that accurate analysis of Fe(II) in samples where Fe(II) constitutes less than 2% of FeT, and even as low as 0.05% of FeT, can be achieved by applying the proposed procedure.

using a mechanical titrator (Metrohm 718 STAT Titrino). Iron(II) was analyzed using the modied phenanthroline method proposed by Herrera et al. (1989). Iron(III) concentration was determined using the sulphosalicilic acid (SSA) method (Zoltov, 2001). 3.2. Reagents Ferric sulfate stock solution was prepared by dissolving 33.22 g of Fe2(SO4)3 reagent (Honeywell Riedel-de Han, purity 97%) in 1 L of distilled water. The Fe(III) concentration of the stock solution was 9.0 0.004 g/L Fe at pH 1.79 (no pH adjustment was performed). Two Fe(II) stock solutions were prepared using ferrous ammonium sulfate hexahydrate reagent (FAS) (Merck, GR, 99101.5% purity): Solution A (200 mg/L Fe) and solution B (4000 mg/L Fe), the concentrations were veried by the phenanthroline method. The pH of both stock solutions was adjusted to 1.77 by addition of H2SO4. The concentration of the NaHCO3 stock solution used for pH elevation was 80 g/L. 3.3. Pretreatment procedure Twenty milliliter aliquots of Fe2(SO4)3 solution were transferred to 50 mL beakers and appropriate volumes of Fe(II) stock solution were added to make up the required Fe(II) concentrations. Nitrogen purging was commenced (a glass diffuser was used) and after a purging time of about 1 min, NaHCO3 solution was added until the desired pH was attained and stabilized. Once Fe(III) solids precipitated about 10 mL of solution were passed through a 0.22 lm syringe lter. 5 mL of the ltrate were immediately acidied by the addition of 0.1 mL of 1.8 N H2SO4 (to attain a pH in the range 1.01.2). 3.4. Experimental 3.4.1. Preliminary experiment To determine the approximate pH range where most of the Fe(III) precipitates (for the conditions tested, i.e. Fe2(SO4)3 solution containing 9 g/L Fe), a preliminary experiment was performed. Ten 20 mL samples of Fe2(SO4)3 solution containing 9 g/L Fe were prepared and varying volumes of NaHCO3 (80 g/L) solution were added to each of them so that the samples had a nal pH of 3.95, 4.11, 4.15, 4.22, 4.35, 4.51, 4.64, 4.88, 4.96 and 5.30, respectively (Table 1). Following pH stabilization and solids removal by ltration, the samples were acidied, and the residual Fe(III) concentration in the ltrates was analyzed using the SSA method. 3.4.2. Suggested procedure verication experiments To investigate the effect of the Fe(II) to FeT ratio on the accuracy of the results obtained with the modied method, six sets of experiments were performed. The Fe(III) background concentration in all the experiments was 9 g/L Fe, while the Fe(II) concentration was varied to generate Fe(II):FeT ratios of between 2% and 0.05%. To determine the effect of pH, different volumes of NaHCO3 stock solution were added to the samples within each set (i.e. with a similar Fe(II):FeT ratio). The conditions of all of the tests

3. Materials and methods 3.1. Analyses The pH was measured with a Metrohm 827 pH-meter equipped with a Metrohm 6.0239.100 glass electrode. An exact volume of NaHCO3 solution was added to the sample

2126

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129

Table 1 Residual Fe(III) concentrations in samples subjected to the phenanthroline method (i.e. following pretreatment) as a function of the pH in the pretreatment procedure (ND = Fe(III) concentration not detectable by the SSA method) Final pH of solution before acidication 3.95 4.11 4.15 4.22 4.35 4.51 4.64 4.88 4.96 5.30 Fe(III) concentration (mg/L) 26.6 9.6 2.1 ND ND ND ND ND ND ND

constant at 2.4 0.06 mg/L Fe (Table 2 set #1), the small variation in results being attributed to normal analytical uctuations. At pH values higher than pH 4.86 reduced measured Fe(II) values were observed, probably indicating that some of the Fe(II) originally present was separated from solution with the Fe(III) precipitates. Despite the fact that determining the true value of the Fe(II) impurity was not possible, a value of 2.4 mg/L was used in further calculations to represent the Fe(II) background concentration in the 9 g Fe(III)/L solution (i.e. a Fe(II) impurity of 0.026% in the Fe2(SO4)3 solid). 4.3. Applying the suggested method (pretreatment + analysis by the modied phenanthroline method) to samples with different Fe(II):FeT ratios at various pH values during pretreatment Table 2 also shows the results of the six analytical experiment sets that were carried out. The accuracy of the Fe(II) determination in samples where the Fe(III) was separated from solution in the pH range of 4.204.65 was very high (all errors below 5% and typically much lower) relative to the expected Fe(II) concentration. Such errors are not excessively higher than the minimal accuracy of 3% (under optimal conditions) associated with the standard phenanthroline method (www.standardmethods.org). Fig. 1 shows, similarly to the results obtained in set experiments #1 (determination of the Fe(II) concentration in pure Fe2(SO4)3 samples), that at pH values higher than $4.80 a noticeable drop in the accuracy of Fe(II) determination was observed in all the analyses. More generally speaking, the results show that within the applied pH range a decrease in Fe(II) determination accuracy is observed when the pH is higher, suggesting that elevating the pH to around pH 4.2pH 4.3 would probably yield the best results. However, acknowledging that attaining such a narrow pH range is cumbersome and that an error of <5% is acceptable, it appears that the raising pH to between pH 4.2 and 4.6 would yield satisfactory results. 5. Discussion The fact that there appears to be a narrow pH range at which the Fe(III) species can be separated from solution without altering the Fe(II) concentration may be theoretically explained by three different phenomena. First, Fe(II) species are known to spontaneously oxidize to Fe(III) throughout the pH range, with an oxidation rate that is strongly pH-dependent. Ferrous iron oxidation kinetics can be described (Stumm and Morgan, 1996; Morgan and Lahav, 2007) for the simple case in which no Fe(II) complexes (other than hydroxide complexes) exist:

performed are listed in Table 2. The Fe2(SO4)3 stock solution contained some Fe(II) impurity. Thus, the rst experiment set was used for determining the Fe(II) content of the Fe2(SO4) solution, that was in subsequent experiments considered as the blank concentration of Fe(II). 4. Results 4.1. Determining the minimal pH required for efcient Fe(III) precipitation in the pretreatment procedure The results of the preliminary experiments, aimed at testing the efciency of the method for Fe(III) separation, are listed in Table 1. The results show that at pH values higher than 4.2 practically all Fe(III) was separated from the samples using the precipitation/ltration pretreatment procedure. Based on these results further experiments were conducted at pH > 4.2 to determine the pH range at which the precipitation/ltration procedure does not alter the dissolved Fe(II) concentration. A Fe(III) concentration of 9 g/L was chosen in these experiments because this is the most common Fe(III) composition in growth media used for A. ferrooxidans (Silverman and Lundgren, 1959) and thus the highest Fe(III) concentration typically found in engineered systems. Obviously, the results obtained are specic to the solution composition tested (i.e. 9 g/L FeFe2(SO4)3 solution) and under other conditions (e.g. a different Fe(III) concentration or ionic strength or counter anion composition that would generate a different dissolved Fe(III) complex distribution) the results would be slightly different. However, the difference is expected to be sufciently small to not change the conclusion that increasing pH to above pH4.2 would result in the precipitation of all Fe(III) species. 4.2. Determining the Fe(II) impurity concentration in the Fe2(SO4)3 background solution Table 2 (1st data set) lists the results of the residual Fe(II) concentrations measured in the pure Fe2(SO4)3 samples (no external Fe(II) added), as measured by the phenanthroline method after applying the pretreatment technique at various pH values. Over the pH range of 3.964.86 the measured Fe(II) concentrations were almost

dFe2 ko Fe2 k1 FeOH k2 FeOH0 2aq DO dt 1


where ko, k1, k2 and k3 are oxidation rate constants (time1), and DO is the dissolved O2 concentration (mg/L). Hydrolyzed Fe(II) species are more readily oxidized than non-hydrolyzed species in the following order

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129 Table 2 Conditions and results of the analytical experiments performed Set no. Set #1 Fe2(SO4)3 solution, Fe(II)/FeT 0.026% Test no. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 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 43 44 45 46 47 49 50 51 52 53 54 Volume of NaHCO3 solution added (mL) 6.364 6.382 6.414 6.424 6.556 6.59 6.652 6.752 6.844 7.102 7.400 6.442 6.516 6.750 6.884 6.912 6.874 7.012 7.084 6.458 6.574 6.612 6.702 6.818 6.822 7.192 6.465 6.558 6.610 6.720 6.820 6.832 7.090 6.556 6.602 6.678 6.808 6.838 6.92 7.056 6.596 6.610 6.702 6.738 6.774 6.996 7.07 6.614 6.718 6.748 6.918 7.016 7.156 Final pH 3.96 4.08 4.24 4.34 4.59 4.73 4.86 4.96 5.21 5.44 5.77 4.28 4.54 4.74 4.85 5.03 5.11 5.27 5.53 4.25 4.37 4.48 4.67 4.88 4.97 5.26 4.22 4.33 4.44 4.65 4.78 4.92 5.14 4.24 4.44 4.62 4.73 4.87 4.99 5.18 4.25 4.31 4.54 4.65 4.84 4.95 5.14 4.21 4.45 4.62 4.76 4.98 5.18 Expected Fe(II) conc (mg/L) NA NA NA NA NA NA NA NA NA NA NA 4.35 4.35 4.35 4.35 4.35 4.35 4.35 4.35 7.34 7.34 7.34 7.34 7.34 7.34 7.34 12.39 12.39 12.39 12.39 12.39 12.39 12.39 40.66 40.66 40.66 40.66 40.66 40.66 40.66 97.67 97.67 97.67 97.67 97.67 97.67 97.67 191.17 191.17 191.17 191.17 191.17 191.17 Measured Fe(II) conc.(mg/L) 2.43 2.42 2.36 2.36 2.34 2.33 2.40 2.31 2.26 2.21 1.92 4.32 4.29 4.25 4.27 4.25 4.09 3.91 3.51 7.03 6.99 7.02 7.01 6.87 6.75 6.45 12.25 12.24 12.30 12.18 12.14 11.84 10.95 40.74 40.56 40.25 39.58 39.01 38.56 37.29 96.99 96.40 94.92 94.5 91.18 90.00 88.50 186.07 181.18 181.20 179.75 176 170.68

2127

Error in Fe(II) analysis (%) NA NA NA NA NA NA NA NA NA NA NA 0.69 1.38 2.30 1.84 2.30 5.98 10.11 19.31 4.22 4.77 4.36 4.5 6.4 8.04 12.13 1.13 1.21 0.73 1.69 2.02 4.44 11.62 0.20 0.25 1.01 2.66 4.06 5.16 8.29 0.7 1.3 2.82 3.25 6.64 7.85 9.39 2.67 5.23 5.22 5.97 7.94 10.72

Set #2, Fe(II)/FeT = 0.05%

Set #3, Fe(II)/FeT = 0.08%

Set #4, Fe(II)/FeT = 0.14%

Set #5, Fe(II)/FeT = 0.45%

Set #6, Fe(II)/FeT = 1.06%

Set #7, Fe(II)/FeT = 2.01%

Initial Fe(III) concentration in all the samples was 9.0 g Fe/L; The concentration of the NaHCO3 reagent solution was 80 g/L; Expected and measured Fe(II) concentrations in sets #2 to #7 reect the Fe(II) concentration dosed, i.e. relative to the blank Fe(II) impurity concentrations determined from set experiments #1.

2 FeOH0 (Luther, 1990). Correspond2aq ) FeOH ) Fe ingly, it has been shown in a number of works (Lowson, 1982; Millero, 1985) that the difference between k2 and k1 and between k1 and k0 is approximately ve orders of magnitude. Since the distribution between the hydrolyzed Fe(II) species is pH-dependent, the oxidation rate is also strongly pH-dependent as follows: At pH values below $4, it can be assumed that the species Fe2+ dominates in

Eq. (1) and the oxidation rate is very slow; at $4 > pH > $5, the concentration of the species FeOH+ increases linearly with pH and the potential rate of oxidation also increases according to Eq. (1). At pH > $5 the species FeOH0 2 dominates in Eq. (1) because it is far more readily oxidized than Fe2+ and FeOH+ as reected by the magnitude of its rate constant. It can be concluded thus that as the pH is increased to a higher value during the

2128

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129

Error in Fe(II) determination (%)

18 15 12 9 6 3 0 4

Recommended pH range

4.2

4.4

4.6

4.8 pH

5.2

5.4

5.6

Fe(II)/FeT = 0.05% Fe(II)/FeT = 0.45%

Fe(II)/FeT = 0.08% Fe(II)/FeT = 1.06%

Fe(II)/FeT = 0.14% Fe(II)/FeT = 2.01%

Fig. 1. Relative error in Fe(II) determination as a function of pH at which Fe(III) was separated in the pretreatment procedure.

Fe(III) separation procedure, unwanted oxidation of Fe(II) becomes more likely. However, taking into account that the pretreatment procedure was carried out under a N2 atmosphere, that the overall oxidation rate of Fe(II) is very low even at pH 5, and that the procedure lasted only several minutes before the samples were acidied following the ltration step, it would appear that Fe(II) oxidation cannot explain the relatively large errors obtained in the procedure at pH values higher than around pH 4.65. Note also that the presence of complexing agents such as carbonate species, Cl or SO2 ions has been 4 shown to retard Fe(II) oxidation rates (e.g. King, 1998), thus the above conclusion, i.e. that Fe(II) oxidation is an unlikely cause for the observed analytical errors, can be considered general. More reasonable explanations for the removal of Fe(II) from solution at pH values higher than 4.65 are that Fe(II) co-precipitates with Fe(III) species to form an intermediate III solid (such as green rust II, FeII 4 Fe2 OH12 SO4 , for example) and/or that Fe(II) adsorbs onto the surface area of precipitated Fe(III) solids. In either case the Fe(II) involved will be lost along with the Fe(III) precipitates during the ltration step. It is well known that most Fe oxides and oxide hydroxides (e.g. hematite, magnetite, goethite, lepidocrocite, akaganeite, ferrihydrite and schwertmannite) have the capability to adsorb divalent metals on their surfaces (Cornell and Schwertmann, 2003). According to Ruby et al. (2003) who studied the co-precipitation of Fe(II) and Fe(III) in sulfate-rich aqueous solutions, the species that is likely to precipitate under the conditions of the present experiments (4 < pH < 5) is the hydroxysulphate basic salt Fe(OH)2.75(SO4)0.125 that corresponds to a Schwertmannite-type compound, which rapidly transforms into other Fe(III)oxide hydroxides such as ferrihydrite over the pH range applied in the pretreatment procedure. Upon aging, the ferrihydrite may transform into goethite (a-FeOOH). Ruby et al. (2006) suggested that the surface of FeOOH solid species acts as a hydroxylating ligand towards Fe(II) ions in solution and that at pH $ 5 partially hydroxylated Fe(II) species may adsorb on FeOOH surfaces. Alter-

natively, amorphous fresh Fe(III) precipitates may also transform directly into goethite and hematite (a-Fe2O3) (Ruby et al., 2003). However the formation of hematite under acidic conditions has been shown to occur at temperatures higher than 25 C, thus hematite is less likely to be present in the precipitates. In any event, all the abovementioned Fe phases are known to adsorb metal cations with an adsorption capacity which rises with increasing pH (Cornell and Schwertmann, 2003). More specically, Fe2+ ions have been shown to adsorb to the surface of both goethite and hematite (Coughlin and Stone, 1995; Jeon et al., 2001), the adsorption being more pronounced at higher pH values. Based on the above, and irrespective of the particular combination of Fe oxides or oxide hydroxides that formed under the experimental conditions, it is likely that the inaccuracy observed in Fe(II) measurements at pH values higher than 4.65 was largely the result of adsorption of Fe(II) species onto the surface area of the ferric oxide/oxyhydroxide precipitate, a phenomenon which has been shown to be more pronounced at higher pH. Accordingly, the conclusion, with respect to the method proposed in this paper, is that in order to minimize the error in Fe(II) determination, pH should be increased during the application of the pretreatment procedure to a value not higher than about pH 4.65.

6. Conclusions  A pretreatment procedure was introduced to separate Fe(III) species from acidic solutions to allow for accurate determination of Fe(II) by the phenanthroline method for samples in which Fe(II) constitutes less than 1% of the total dissolved Fe concentration.  Pretreatment consists of pH elevation to between pH 4.2 and pH 4.65 using NaHCO3 under N2(g) environment followed by ltration of the solid Fe(III) oxides formed, and subsequent acidication of the Fe(II)-containing ltrate.

Y. Gendel, O. Lahav / Applied Geochemistry 23 (2008) 21232129

2129

 If pretreatment is properly executed the determination of Fe(II) concentration has been shown to be accurate (error < 5%) for samples with Fe(II):FeT ratios as low as 0.05%.  pH elevation to above pH 4.65 during pretreatment was shown to result in measurable losses of Fe(II). These losses are likely the result of Fe(II) species adsorption to the Fe(III) oxide/oxyhyroxide precipitate.  The accuracy of Fe(II) analysis obtained with the proposed method was found to be unrelated to the Fe(II)/ FeT ratio in the original sample. It therefore appears that the technique can be also used for Fe(II)/FeT ratios lower than 0.05%, the limiting factor being the accuracy of the phenanthroline technique rather than the Fe(II)/FeT ratio.

Acknowledgements This work was supported by Research Grant No. IS-3522-04 from BARD, the United States Israel Binational Agricultural Research and Development Fund. References
Akl, M.A., Mori, Y., Sawada, K., 2006. Solvent sublation and spectrometric determination of iron (II) and total iron using 3-(2-pyridyl)-5,6-bis(4phenylsulfonic acid)-1,2,4-triazine and tetrabutylammonium bromide. Anal. Sci. 22, 11691174. Asai, S., Konishi, Y., Yabu, T., 1989. Kinetics of absorption of hydrogen sulde into aqueous ferric sulfate solution. AIChE J. 35, 12711281. Cornell, R.M., Schwertmann, U., 2003. The iron oxides, second ed. WileyVCH Gmbh & Co. KGaA, Weinheim, Germany. Coughlin, B.R., Stone, A.T., 1995. Nonreversible adsorption of divalent metal ions (MnII, CoII, NiII, CuII, and PbII) onto goethite: effects of acidication, FeII addition, and picolinic acid addition. Environ. Sci. Technol. 29, 24452455. Ebrahimi, S., Leerebezem, R., Loosdrecht, M.C.M., Heijnen, J.J., 2002. Kinetics of the reactive absorption of hydrogen sulde into aqueous ferric sulfate solution. Chem. Eng. Sci. 58, 417427. Fadrus, H., Maly, J., 1975. Suppression of iron (III) interference in the determination of iron (II) in water by the 1,10-phenanthroline method. Analyst 100, 549554. Giokas, D.L., Evangelos, K.P., Karayannis, M.I., 2002. Speciation of Fe(III) and Fe(II) by the modied ferrozine method, FIA-spectrophotometry, and ame AAS after cloud- point extraction. Anal. Bioanal. Chem. 373, 237243. Gomez, J.M., Cantero, D., 2003. Kinetic study of biological ferrous sulphate oxidation by iron-oxidising bacteria in continuous stirred tank and packed bed bioreactors. Process Biochem. 38, 867875. Herrera, L., Ruiz, P., Aguillon, J.C., Fehrmann, A., 1989. A new spectrophotometric method for the determination of ferrous iron in the presence of ferric iron. J. Chem. Tech. Biot. 44, 171181.

Hong, H., Kester, D.R., 1986. Redox state of iron in the offshore waters of Peru. Limnol. Oceanog. 31, 512524. Jeon, B.H., Dempsey, B.A., Burgos, W.D., Royer, R.A., 2001. Reactions of ferrous iron with hematite. Colloid Surf. A 191, 4155. Jones, C.A., Kelly, D.P., 1983. Growth of Thiobacillus ferrooxidans on ferrous iron in chemostat culture: inuence of product and substrate inhibition. J. Chem. Tech. Biot. 33B, 241261. King, W., 1998. The role of carbonate speciation on the oxidation rate of Fe(II) in aquatic systems. Environ. Sci. Technol. 32, 29973003. Krishnamurti, G.S.R., Huang, P.M., 1990. Spectrophotometric determination of Fe(II) with 2,4,6-tri(20 pyridyl)-1,3,5-triazine in the presence of large quantities of Fe(III) and complexing ions. Talanta 37, 745748. Lowson, R.T., 1982. Aqueous oxidation of pyrite by molecular oxygen. Chem. Rev. 82, 461499. Luther, G.W., 1990. The frontier-molecular-orbital theory approach in geochemical processes. In: Stumm, W. (Ed.), Aquatic Chemical Kinetics. Wiley-Interscience, New York, pp. 73199. Millero, F.J., 1985. The effect of ionic interactions on the oxidation of metals in natural waters. Geochim. Cosmochim. Acta 49, 547553. Molchanov, S., Gendel, Y., Ioslvich, I., Lahav, O., 2007. An improved experimental and computational methodology for determining the kinetic equation and extant kinetic constants of Fe(II) oxidation by Acidithiobacillus ferrooxidans. Appl. Environ. Microbiol. 73, 17421752. Morgan, B.E., Lahav, O., 2007. The effect of pH on the kinetics of spontaneous ferrous oxidation by O2 in aqueous solution basic principles and a simple heuristic description. Chemosphere 68, 2080 2084. Muir, M.K., Andersen, T.N., 1977. Determination of ferrous iron in copperprocess metallurgical solutions by ortho-phenanthroline colorimetric method. Metall. Trans. B 8, 517518. Nemati, M., Webb, C., 1998. Inhibition effect of ferric iron on the kinetics of ferrous iron. Biotechnol. Lett. 20, 873877. Nemati, M., Harrison, S.T.L., Hansford, G.S., Webb, C., 1998. Biological oxidation of ferrous sulphate by Thiobacillus ferrooxidans: a review on the kinetics aspects. Biochem. Eng. J. 1, 171190. Ruby, C., Aissa, R., Gehin, A., Cortot, J., Abdelmoula, M., Genin, J.M., 2006. Green rusts synthesis by coprecipitation of FeIIFeIII ions and massbalance diagram. C. R. Geosci. 338, 420432. Ruby, C., Gehin, A., Abdelmoula, M., Genin, J.M.R., Jolivet, J.P., 2003. Coprecipitation of Fe(II) and Fe(III) cations in sulphated aqueous medium and formation of hydroxysulphate green rust. Solid State Sci. 5, 10551062. Silverman, M.P., Lundgren, D.G., 1959. Studies on the chemoautotrophic iron bacterium Ferroobacillus ferrooxidans: I. An improved medium and a harvesting procedure for securing high cell yields. J. Bacteriol. 77, 642647. Stookey, L.L., 1970. Ferrozine: a new spectrophotometer reagent for iron. Anal. Chem. 42, 779781. Stumm, W., Morgan, J.J., 1996. Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. Wiley Interscience, New York. Tufekci, N., Sarikaya, H.Z., 1996. Catalytic effects of high Fe(III) concentrations on Fe(II) oxidation. Water Sci. Technol. 34, 389396. Viollier, E., Inglett, P.W., Hunter, K., Roychoudhury, A.N., Cappellen, V.P., 2000. The ferrozine method revisited: Fe(II)/Fe(III) determination in natural waters. Appl. Geochem. 15, 785790. Zoltov, Y.A., 2001. Fundamentals of Analytical Chemistry Practical Guide. Visshaya Shkola, Moscow.

S-ar putea să vă placă și