You are on page 1of 9

Environ. Sci. Technol.

2003, 37, 2734-2742

nonionic H3AsO3, which does not adsorb as strongly to


Iron-Catalyzed Oxidation of mineral surfaces (1) as As(V) (4). As(III) is thus more mobile
Arsenic(III) by Oxygen and by in groundwaters and is also more difficult to remove in arsenic
removal treatments (5, 6). Typically, preoxidation to As(V) is
Hydrogen Peroxide: pH-Dependent necessary in arsenic removal procedures (7). Although As(V)
is thermodynamically favored under oxic conditions, As(III)
Formation of Oxidants in the Fenton is only slowly oxidized by dissolved O2 (e.g., with a half-life
of around 9 d in air-saturated water with low iron contents
Reaction and pH 7.6-8.5) (8). The As(III) oxidation by H2O2 is also
slow at neutral and acidic pH, as only H2AsO3- and HAsO32-
STEPHAN J. HUG* AND OLIVIER LEUPIN but not H3AsO3 react with H2O2 (9). The half-life of As(III) in
Swiss Federal Institute for Environmental Science and 1 mM H2O2 at pH 7.5 is 2.1 d. The influence of transition
Technology (EAWAG), Überlandstrasse 133, metals on the oxidation of As(III) with H2O2 has recently
CH-8600 Dübendorf, Switzerland been studied (10), and it was found that Fe(II) and Fe(III)
both increased the oxidation rates to a similar extent, but the
effects of Fenton chemistry were not addressed.
The iron-catalyzed oxidation of As(III) with O2 and H2O2
The oxidation kinetics of As(III) with natural and technical is important in both natural and technical systems. The
oxidants is still not well understood, despite its importance connection between iron and arsenic redox reactions is one
in understanding the behavior of arsenic in the environment of the key factors in understanding the fate of arsenic in the
and in arsenic removal procedures. We have studied the environment. For example, arsenic in the iron crusts of recent
sediments from Lake Baikal was found to be enriched by a
oxidation of 6.6 µM As(III) by dissolved oxygen and hydrogen
factor of 58 above background levels (11). In water treatment,
peroxide in the presence of Fe(II,III) at pH 3.5-7.5, on a oxidation followed by coprecipitation with iron (hydr)oxides
time scale of hours. As(III) was not measurably oxidized by is one of the most economic and efficient arsenic removal
O2, 20-100 µM H2O2, dissolved Fe(III), or iron(III) (hydr)- options (12). The oxidation of As(III) by Fe(II) and H2O2
oxides as single oxidants, respectively. In contrast, As(III) (Fenton reaction) at pH 1-3 has been studied as early as
was partially or completely oxidized in parallel to the 1963 (13) and has been tested for the treatment of ground-
oxidation of 20-90 µM Fe(II) by oxygen and by 20 µM water with high arsenic concentrations at pH 2.5-3.0 (14).
H2O2 in aerated solutions. Addition of 2-propanol as an •OH- Treatment of As(III)-spiked groundwater by a combination
radical scavenger quenched the As(III) oxidation at low of Fenton chemistry and filtration through scrap iron at
pH but had little effect at neutral pH. High bicarbonate neutral pH has recently been reported (15), but kinetic and
mechanistic issues were not discussed.
concentrations (100 mM) lead to increased oxidation of As-
In a recent study of iron-catalyzed photochemical oxida-
(III). On the basis of these results, a reaction scheme is
tion and removal of arsenic at neutral pH (16), we found that
proposed in which H2O2 and Fe(II) form •OH radicals at low arsenic was partly oxidized in parallel to the dark oxidation
pH but a different oxidant, possibly an Fe(IV) species, at of Fe(II) by dissolved O2. We have attempted to explain these
higher pH. With bicarbonate present, carbonate radicals observations with the formation of HO2•/•O2-, H2O2, and •OH
might also be produced. The oxidant formed at neutral pH radicals in the generally stated sequence of Fe(II,III) oxida-
oxidizes As(III) and Fe(II) but does not react competitively tion/reduction reactions and Fenton chemistry (17-21).
with 2-propanol. Kinetic modeling of all data simultaneously However, the involvement of free •OH radicals was not
explains the results quantitatively and provides estimates consistent with quenching experiments in which •OH scav-
for reaction rate constants. The observation that As(III) engers such as 2-propanol and 3-butanol had almost no effect
is oxidized in parallel to the oxidation of Fe(II) by O2 and by on the oxidation of As(III) at neutral pH (16). An important
role of •O2- as an oxidant for As(III) was also excluded on the
H2O2 and that the As(III) oxidation is not inhibited by •OH-
basis of the results of the dark and photochemical reactions.
radical scavengers at neutral pH is significant for the
A number of newer studies have suggested that oxidizing
understanding of arsenic redox reactions in the environment species different from free •OH radicals (e.g., Fe(IV) species)
and in arsenic removal processes as well as for the are formed in the Fenton reaction (22-25) and in the
understanding of Fenton reactions in general. oxidation of Fe(II) with O2 (26). Fe(IV) intermediates in iron
porhyrins have been isolated and studied by low-temperature
spectroscopy (27). Spectroscopic evidence for a ferryl species
Introduction and an Fe(III) dimer formed from ferryl and Fe(II) was also
With arsenic-contaminated groundwater affecting 30-50 found in the reaction of uncomplexed Fe(II) with O3 at pH
million people in Bangladesh (1) and possibly up to10 million below 2 (28) and in a pulse radiolysis study (29). Rate
in Vietnam (2) and with the debate around lower limits in constants for the reaction of ferryl with various compounds
the United States (3), arsenic has been recognized as one of at low pH were reported (30). Furthermore, Fe(IV) species
the most widespread and problematic water contaminants. were suggested in the reaction of Fe(II)-NTA and Fe(II)-
The behavior of arsenic in the environment and in water EDDA (31) and of Fe(II)-DTPA (32) with H2O2 at neutral pH.
treatment processes is mainly determined by its redox The relative contribution of reaction pathways leading to
chemistry. As a water contaminant, As(III) is more prob- Fe(IV) and •OH, respectively, seem to depend on various
lematic than As(V). Unlike the arsenate anions H2AsO4- and parameters, such as pH and the concentrations of organic
HAsO42-, the dominant As(III) species up to pH 8 is the and inorganic ligands (33-35).
In the degradation of pollutants, it is generally observed
* Corresponding author phone: +41-1-823-5454; fax: +41-1-823- that the efficiency of dark and photo-Fenton systems
5210; e-mail: hug@eawag.ch. decreases with increasing pH (36). For example, atrazine
2734 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 37, NO. 12, 2003 10.1021/es026208x CCC: $25.00  2003 American Chemical Society
Published on Web 05/17/2003
As(III) Analysis. As(III) was measured with an atomic
TABLE 1. Experimental Parameters (Initial Concentrations fluorescence spectrometer (AFS)sMillenium Excalibur (PS
after Mixing)a Analytical Ltd, Kent, U.K.). The instrument works with hydride
species oxidation oxidation with oxidation with generation by continuous-flow mixing of the acidified
(concn) with O2 H2O2 at >pH 6 H2O2 at <pH 6 (normally 1 M HCl) sample with 0.7% NaBH4 in a Teflon
As(III) (µM) 6.6 6.6 6.6 valve and a liquid-gas separator. To selectively detect As-
Fe(II) (µM) 90 0.1-20 20 (III), we used a pH 5 citrate buffer (0.5 M disodium hydrogen
H2O2 (µM) 0 20 20 citrate) instead of HCl. Only As(III) (and no As(V)) is converted
initial pH 7.0 and 7.5 7.0 and 7.5 pH 3.5, 4.4, and 5.0 to AsH3 under these conditions (41). Aliquots (100 or 200 µL)
HCO3- (mM) 8.30 0,b 2.1, 100c 0 from the reaction mixtures were diluted 1:50 into 0.5 M citrate
Ca, mg/L (mM) 2.50 0.63 0.63 buffer. The dilution and lowering of pH stopped the reaction
Mg, mg/L (mM) 1.65 0.41 0.41
SO42- (µM) 90 0.1-20 1110 between Fe(II), oxygen, and H2O2. Both dissolved and
2-propanol (mM) 0, 14 0, 14 0, 14 adsorbed As(III) were measured as the pH 5 citrate buffer led
a The temperature was 22-25 °C. b With H PO -/HPO 2- (pH 7.4) )
to the desorption of As(III) and to the dissolution of the
2 4 4
2 mM. c In two experiments with 100 mM NaHCO3 (with and without
amorphous iron hydroxide colloids. We verified that the
2-propanol and formate). reactions did not proceed after the sampling and dilution by
performing the AFS analysis at several times 5-120 min after
mixing of the sample aliquot with the buffer.
degradation in dark and-photo-Fenton systems (37, 38) Reproducibility. Each experiment was repeated 2-3
decreased sharply from pH 4-7, and it was suggested that times. The analysis of As(III) and As(V) showed standard
•OH is formed only at low pH. In contrast, the oxidation of
deviations of (7%. The standard deviations of the Fe(II)
As(III) in Fenton and photo-Fenton reactions appears to be measurements (phenanthroline method) was (5%. The
possible at neutral pH. To better understand the reaction reproducibility between different experiments was mostly
pathways and kinetics, we have investigated the dark reaction limited by the pH increases due to loss of CO2 from the
of Fe(II) with O2 and with H2O2 as a function of pH and solutions during the sampling and due to lack of buffering
bicarbonate concentrations. below pH 6. The deviations between repeated experiments
were approximately (10%.
Experimental Section FTIR Spectra of Colloids. Colloids were collected on 0.25-
µm Teflon filters and measured on a Bio-Rad FTS 575C FT-
Chemicals. All chemicals were reagent grade from Aldrich IR instrument equipped with a Harrick Meridian Diamond
or Fluka and were used as received. split pea ATR unit.
Kinetic Experiments. Synthetic groundwater was pre- Formation of Acetone from 2-Propanol. The formation
pared by dissolution of CaCO3 and MgCO3 with an excess of of acetone during the reaction of 20 µM H2O2, 20 µM Fe(II),
CO2 in high-purity 18 MΩ water (Millipore). The initial pH and 0.5 mM 2-propanol as a function of pH was measured
for the experiments at pH 7.0-7.5 was reached by outgassing in two experiments with direct-injection GC-MS (GC8000
of CO2 by vigorous stirring in open beakers. H2SO4 was used and MD800, Fisons, Manchester) (42).
to adjust pH values below 6. The initial pH, measured with Kinetic Modeling. Kinetic modeling was performed with
a combined glass electrode, was (0.1 unit within the indicated ACUCHEM (43) and with improved global fitting routines
value and typically increased by 0.1-0.2 pH unit during the written in Matlab (The MathWorks Inc.) but otherwise similar
kinetic runs due to loss of CO2. as in previous studies of illuminated Fe(III) systems (38, 44-
The 90 mM (5.0 g/L) Fe(II) stock solutions were always 46). The list of reactions in Table 2 was used as the input for
freshly prepared by dissolving FeSO4‚7H2O in 1 mM HCl. ACUCHEM. Equilibrium reactions were written as separate
The 100 mM H2O2 stock solutions were prepared from 30% forward and back reactions, with fast, non-rate-determining
H2O2 (determined by titration with permanganate). For the forward/back rate constants, and the pH was kept constant
oxidation experiments with dissolved O2, acidic Fe(II) stock for each experiment by formulating all equilibrium reactions
solution was quickly mixed into 100-200 mL of the pH- as H+-preserving reactions. The concentration of dissolved
adjusted air-saturated synthetic groundwater solutions that O2 was kept constant by equilibrium with a large reservoir
already contained the As(III). Aliquots were sampled at of air. The input file for ACUCHEM is provided as Supporting
specific time intervals. In the experiments with H2O2, two Information. The adjustable rate constants were changed by
equal volumes with the separated reactants at twice the the Matlab routine until the total sum of normalized squared
targeted initial concentrations were prepared separately. For differences between data and model for all experiments was
example, a pH-adjusted solution containing H2O2, As(III), minimized with the Nelder-Mead simplex routine.
and the •OH-radical scavenger (e.g., 2-propanol) was quickly
mixed with the same volume of Fe(II) solution. This approach Results and Discussion
ensured that the concentrations during the mixing never Oxidation of Fe(II) and As(III) in Aerated Solutions with
exceeded the intended initial concentrations. The parameters O2. The concentration change of As(V) and Fe(II) after
for all kinetic experiments are listed in Table 1. addition of 90 µM Fe(II) to air-saturated 6.6 µM As(III)
As(V) Analysis. In the experiments where iron was solutions are shown in Figure 1. The solid lines in these and
oxidized by dissolved O2, the sum of dissolved and adsorbed in all following figures represent the output from the model
As(V) was measured spectrophotometrically with the mo- calculation (Tables 2 and 3, model 2b). In the following
lybdenum blue method (39). Under the acidic and reducing discussion, we refer to the reactions in Table 2, but the
conditions of the reagents, colloidal iron(III) (hydr)oxides reaction model will be discussed in more detail in the
were dissolved. To check for interference with traces of following section.
phosphate impurities, absorbance spectra were recorded It can be seen that, parallel to the disappearance of Fe(II)
from 500 to 900 nm and fitted with the distinct spectra of due to oxidation by dissolved O2, As(III) is partly oxidized to
molybdenum blue formed from phosphate and arsenate (40). As(V). After the oxidation of Fe(II) is complete, As(III) is not
The As(V)-specific molybdenum blue method was more exact oxidized further with the fraction of oxidized As(III) remaining
than forming differences from As(III)-specific measurements between 25 and 30%. The higher oxidation rates at pH 7.5
(see below) in experiments where only small fractions of as compared to pH 7.0 are explained by the pH-dependent
As(III) were oxidized. conditional rate constants for Fe(II) oxidation taken from

VOL. 37, NO. 12, 2003 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 2735
TABLE 2. Kinetic Modela
reactions rate constants (s-1) or (M-1 s-1 ref.
Arsenic Oxidations
A1 As(III) + •OH f As(IV) 8 × 109 49
A2 As(III) + CO3•- f As(IV) + HCO3- 1.1 × 108 49
A3 As(IV) + O2 f As(V) + O2•- 1.1 × 109 49
FeII Oxidations
B1 Fe2+ + O2 f O2•- + Fe3+ k9, k10 47
B2 Fe2+ + O2•- + 2H+ f Fe3+ + H2O2 1 × 107 19
B3 Fe2+ + HO2• + H+ f Fe3+ + H2O2 1.2 × 106 19
B4 Fe2+ + •OH f Fe3+ + OH- 3.2 × 108 59
B5 Fe2+ + CO3• + H+ f Fe3+ + HCO3- 3.6 × 108 for Fe(CN)44- (60)
Fenton Reactions
f1 Fe2+ + H2O2 f Fe3+ + •OH + OH-
f2 FeIIOH+ + H2O2 f FeIIIOH2+ + •OH + OH-
f3 FeIICO3 + H2O2 f FeIIICO3+ + •OH + OH-
Modified Fenton Reactions
F1 Fe2+ + H2O2 f INT k11 61
F2 FeIIOH+ + H2O2 f INT-OH k12 61
F3 FeIICO3 + H2O2 f INT-CO3 k1
I1 INT-OH +H+ a INT K2
I2 INT f FeIII(OH)2+ + •OH 1 × 106*
I3 INT-OH f Fe(IV) 1 × 106*
I4 INT-CO3 f FeIIIOH2+ + CO3•- 1 × 106*
I5 INT-CO3 f Fe(IV) + HCO3- k3
Fe(VI) Reactions
J1 Fe(IV) + Fe2+ f Fe3+ + Fe3+ k4
J2 Fe(IV) + FeIIOH+ f Fe3+ + Fe3+ k5
J3 Fe(IV) + FeIICO3 f Fe3+ + FeIIICO3+ k6
J4 As(III) + Fe(IV) f As(IV) + Fe3+ k7
J5 FeIIICO3+ + H+ f Fe3+ + HCO3- 1 × 109*
Fe(III) Reductions
R1 Fe3+ + O2•- f Fe2+ + O2 1.5 × 108 19
R2 FeIIIOH2+ + O2•- f Fe2+ + O2 + OH- 1.5 × 108 19
R3 FeIII(OH)2+ + O2•- f Fe2+ + O2 + 2OH- 1.5 × 108 19
Reactive Oxygen Radical Reactions
O1 HO2• + O2•- + H+ f H2O2 + O2 9.7 × 107 19
O2 HO2• + HO2• f H2O2+ O2 8.3 × 105 19
O3 H2O2 + •OH f HO2• + H2O 3 × 107 62
O4 HCO3- + •OH f CO3•- + H2O 8.5 × 106 62
O5 H2O2 + CO3•- f HO2• + HCO3- 4.3 × 105 63
Q1 •OH + 2-prop f O •- + acetone 1.9 × 109 62
2
Q2 CO3•- + 2-prop f HCO3- + products 5 × 104 64
Fe(III) Equilibria
E1 Fe3+ a FeIIIOH2+ + H+ 4.57 × 109/1 × 1012 (pK1 2.3) 65
E2 FeIIIOH2+ a FeIII(OH)2+ + H+ 2.51 × 108/1 × 1012 (pK2 3.6) 65
E3 FeIII(OH)2+ a FeIII(OH)3 + H+ 1.17 × 104/1 × 1012 (pK3 7.9) 65
Fe(III) Precipitation
P1 FeIII(OH)3 + FeIII(OH)3 f precip k8 see Table 3
P2 FeIII(OH)3 + precip f precip k8 see Table 3
Fe(II) Equilibria
E5 Fe2+ a FeIIOH+ + H+ 3.09 × 102/1 × 1012 (pK1 ) 9.51) 65
E6 Fe2+ + CO32- a FeIICO3 4.89 × 1012/1 × 107 (log K ) 5.69) 65
Superoxide Equilibria
E7 O2•- + H+ a HO2• 1 × 1012/1.58 × 107 (pKa 4.8) 19
E8 air a O2 1 × 106/1 × 1012 constant [O2]
(250 M air reservoir)
Carbonate Equilibria
E9 HCO3- a CO32- + H+ 4.67 × 101/1 × 1012 (pKa ) 10.33) 65
E10 CO2 a HCO3- + H+ 4.46 × 105/1 × 1012 (pKa ) 6.35) 65
a List of reactions used for modeling. Species in italics are written to keep equations balanced but were not included in the model. Rate constants

in bold were optimized by simultaneous fitting. Those marked with an asterisk (*) are nonrate determining and were estimated; all others were
taken from the literature. The conditional, pH-dependent k9 and k10 were calculated from Millero et al. (47): 14 M-1 s-1 at pH 7.42, 3.0 M-1 s-1 at
pH 7.10, and 1.3 × 10-4 M-1 s-1 at pH 5.0 (I ) 0.01 M). In some optimizations, k9-k12 were kept at fixed values and left adjustable in others (see
Table 3).

the literature (47). The oxidation of As(III) must occur by shows that •OH cannot be the oxidant for As(III). At the added
reactive oxygen species (O2•-, H2O2, •OH, or other oxidizing concentration of 14 mM, 2-propanol would completely inhibit
intermediates) that are formed as intermediates during the the oxidation of As(III) by quantitatively scavenging the •OH
reduction of O2 with Fe(II). The small effect of 2-propanol radicals. Also, H2O2 did not oxidize As(III) on a time scale of

2736 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 37, NO. 12, 2003
and 20 µM H2O2 in aerated solutions, without and with 14
mM 2-propanol, at different pH values. At pH 7, the oxidation
of As(III) is rapid but incomplete. The fraction of As(III)
oxidized decreased only slightly in the presence of 2-propanol.
At pH 5.0, As(III) was completely oxidized but slightly slower
in the presence of 2-propanol. At pH 4.4, the effect of
2-propanol was clearly evident, both in the smaller oxidation
rate and in the reduced fraction of oxidized As(III). At pH 3.5,
2-propanol almost completely inhibited the oxidation of As-
(III). These experiments clearly demonstrate that the reaction
mechanism undergoes significant changes between pH 3.5
and pH 7.5. The most important conclusion is that •OH might
be the oxidant for As(III) at low pH, where the oxidation of
As(III) is almost completely quenched by the reaction of •OH
with 2-propanol. However, with increasing pH, an additional
oxidant that does not react competitively with 2-propanol
must be formed. HO2• and O2•- as significant oxidants can
be discounted for the following reasons: At low pH, •OH
reacts with 2-propanol mainly to acetone and HO2• (48). Since
the As(III) oxidation is almost completely quenched by
2-propanol at pH 3.5, HO2• cannot be a significant oxidant
for As(III) in this system. At higher pH, O2•- is formed. Since
O2•- is a weaker oxidant than HO2• and H3AsO3 remains the
dominant As(III) species over the entire pH range used, O2•-
can also not be a significant oxidant for As(III). This is
consistent with our former photochemical study (16). In the
pulse-radiolysis study of Klaning et al. (49), oxidation of As-
(III) by O2•- was also not reported. Thus, an additional oxidant
must be formed in the Fenton reaction at higher pH. To
account for these observations, we added the reactions (F1-
F3) in which Fe(II) and H2O2 form an intermediate that leads
to formation of Fe(IV) at neutral pH and to •OH at low pH
(I1-I5). Fe(IV) oxidizes Fe(II) (J1-J3) and As(III) (J4) but not
2-propanol. The mechanistic reasoning for this model will
FIGURE 1. Oxidation of 6.6 µM As(III) after the addition of 90 µM be discussed in more detail in a separate paragraph below.
Fe(II) in aerated solutions. (A) The formation of As(V) and (B) the The relative reaction rates of the proposed oxidant with
disappearance of Fe(II) in the same experiments is indicated by
the competitive reductants Fe(II) and As(III) at close to neutral
symbols. Error bars ((10%) in these and the following figures are
pH were studied by varying the amount of added Fe(II) to
not shown for clarity. The lines represent the output of the model
2b (Table 2), which simultaneously fits all the data in these and the the otherwise constant concentrations of 20 µM H2O2 and
following figures. Adding 14 mM 2-propanol as an OH-radical 6.6 µM As(III) at pH 7.2-7.5. The results are shown in Figure
scavenger did not significantly lower the oxidation of As(III). 3. At low concentrations of added Fe(II), several times more
As(III) is oxidized than Fe(II) added. RAs/Fe, the ratio of As(III)
hours, in agreement with other studies (9). The role of O2•- oxidized over Fe(II) added, is about 3 for 0.5 µM added Fe-
as an oxidant was discounted in our previous work (16). To (II). At higher concentrations of added Fe(II), RAs/Fe drops to
better understand this system, we have conducted experi- below 1 (e.g., below 0.33 for 20 µM Fe(II)). Overall, the model
ments in which As(III) was oxidized with H2O2 and with Fe(II) accounts well for the observed trends. The addition of very
at various pH values. low concentrations of Fe(II) starts a chain reaction. The first
Oxidation of Fe(II) and As(III) with Added H2O2. Figure step is the reaction of Fe(II) with H2O2 and the formation of
2A-D shows the oxidation of 6.6 µM As(III) with 20 µM Fe(II) Fe(IV). The reaction of Fe(IV) with As(III) forms Fe(III) and

TABLE 3. Effect of Model Variationsa


rate constant log(ki) model 1 model 2a model 2b model 2c reported
(M-1 s-1) (P1) (P1 + P2) (P1 + P2) (P1 + P2) log(k)
k1, FeIICO3 + H2O2 6.11 5.99 5.94 5.95 4.34d, 7.52d
pK2, INT a INT-OH + H+ 5.16 5.14 5.24 5.34
k3, CO3 partitioning 6.66 6.61 6.63 6.59
k4, Fe2+ + Fe(IV) 3.92 3.09 3.28 3.59
k5, FeIIOH+ + Fe(IV) 8.51 7.93 8.13 8.10
k6, FeIICO3 + Fe(IV) 7.22 6.63 6.56 6.55
k7, As(III) + Fe(IV) 6.28 5.65 5.62 5.58
k8, Fe(OH)3 Precip. 9.52 8.30 8.29 8.36
k9, Fe(II) +O2 (pH 7.5) 1.15 1.15 1.17 1.15a
k10, Fe(II) +O2 (pH 7.2) 0.48 0.48 0.47 0.48a
k11, Fe2+ + H2O2 1.60 1.60 1.43 2.11a 1.80b, 1.92c
k12, FeIIOH+ + H2O2 7.21 7.21 7.12 7.21a 6.77b, 5.58c
SSD 1.46 1.32 1.26 1.66
a Fitted rate constants and sum of squared differences (SSD) for different model variations. Numbers in italics indicate that rate constants were

kept at fixed values while the rate constants in bold were optimized by simultaneous fitting. The last column lists log(k) values reported in other
studies under similar conditions: aMillero et al. (47, 61); bSedlak and Hoigne (44); cKing and Farlow (56); dlog(k) for the reaction of FeIICO3 and
FeII(CO3)22- with H2O2, respectively (56). (All iron(II)-carbonate species in our model were represented by FeIICO3.)

VOL. 37, NO. 12, 2003 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 2737
FIGURE 3. Oxidation of 6.6 µM As(III) in 20 µM H2O2 and 0.1-20
µM Fe(II) at pH 7.2-7.5. The ratio of As(III) oxidized over Fe(II)
added decreases with increasing concentrations of added Fe(II).

As(IV) (J4). As(IV) reacts almost diffusion controlled mainly


with dissolved O2 (A3) (which is an order of magnitude more
concentrated than H2O2) to As(V) and O2•-. The chain is
propagated by the reduction of Fe(III) with O2•- to re-form
Fe(II) (R1-R3). With increasing concentrations of Fe(II), Fe-
(IV) reacts with Fe(II) to two Fe(III) in a chain-terminating
reaction (J1-J3). With increasing pH, the increasing fraction
of FeIIOH+ (E5) reacts more quickly with Fe(IV). This explains
the incomplete oxidation of As(III) at pH 7 (Figure 2A) as
compared to the complete oxidation at pH 5 (Figure 2B) in
the presence of H2O2 as well as the only 20-30% oxidation
at the higher Fe(II) concentrations with O2 (Figure 1A).
Influence of Bicarbonate. The possible role of iron(II)
carbonate species and of carbonate radicals was investigated
in experiments with 20 µm H2O2, 6 µM Fe(II), and different
bicarbonate concentrations. The results in Figure 4A show
that about twice as much As(III) is oxidized in the presence
of 100 mM HCO3- than with only 2 mM HCO3-. In solutions
buffered with phosphate, even less As(III) is oxidized. The
same experiments with 14 mM added formate are shown in
Figure 4B (another experiment with 14 mM 2-propanol is
not shown). Within the scatter of the data, there is no clear
difference between the results with and without formate or
2-propanol. Bicarbonate increases the fraction of oxidized
As(III), possibly by formation of carbonate radicals or by
changing the speciation of Fe(II). The small or absent
influence of formate and 2-propanol again confirms that, at
pH 7.5, •OH radicals are not the main oxidant of As(III).
Formation of Acetone from 2-Propanol as a Function
of pH. The final concentration of acetone formed 2 h after
mixing of equal volumes of Fe(II) and H2O2 solutions to initial
concentrations of 20 µM H2O2, 20 µM Fe(II), and 0.5 mM
2-propanol are shown in Figure 5. The large error bars in the
pH values represent the pH changes during the reactions in
the unbuffered solutions. Despite the detection limit of 2 µM
for acetone and the uncertainty in the pH values, it is clearly
apparent that the formation of acetone decreased steeply
with increasing pH. Assuming a 65% yield of acetone in the
reaction of •OH with 2-propanol and subsequent oxidation
of the organic radical by O2 (summarized in reaction Q1 in
Table 2), the model correctly predicts the acetone formation
as a function of pH. (The reported maximum yield of acetone
formation from 2-propanol and •OH in aerated solutions is
87% (48)). These measurements confirm that, at higher pH,
2-propanol is not oxidized and thus does not quench the
FIGURE 2. Effect of pH on the oxidation of 6.6 µM As(III) with 20 oxidation of As(III) at neutral pH.
µM Fe(II) and 20 µM H2O2. The addition of 14 mM 2-propanol had Characterization of Colloids. The colloids formed from
only a small effect above pH 5 (A and B) but lowered the rate and the oxidation of 90 µM Fe(II) by dissolved O2 in aerated
the fraction of As(III) oxidation at pH 4.4 (C) and almost completely solutions were investigated by ATR-FTIR spectroscopy (Figure
inhibited the As(III) oxidation at pH 3.5 (D). 6) and compared to spectra of known Fe(III) phases (50).

2738 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 37, NO. 12, 2003
FIGURE 6. ATR-FTIR spectra of the colloids formed from the
oxidation of 90 µM Fe(II) by dissolved O2 at pH 7.0-7.3. Spectrum
1, in synthetic groundwater with As(III). Spectrum 2, same as 1 but
with the addition of 9 µM crystalline lepidocrocite. Spectrum 3,
without As(III) at pH 7.0-7.2. Spectrum 4, same as 3 but with As(III).
Spectrum 5, same as 3 but with As(V). Spectral assignments:
preformed crystalline lepidocrocite, 1024 and 751 cm-1. Lepidocrocite
formed from Fe(II) oxidation, 1017 and 743 cm-1. Adsorbed As(V),
804 and 841 cm-1. Adsorbed (bi)carbonate, 1375 and 1492 cm-1.

akaganeite, and δ-FeOOH can be excluded as none of their


spectral characteristics are observed (50). The lepidocrocite
fraction is estimated at 5-10% by comparison with spectrum
2, where 10% of crystalline lepidocrocite was added. Spectra
3-5 are of precipitates formed at pH 7.0-7.2, without and
with As(III) and with As(V) (6.6 µM), respectively. In the
presence of As(III) and As(V), less lepidocrocite and more
FIGURE 4. Oxidation of 6.6 µM As(III) in 20 µM H2O2 and 6 µM Fe(II) adsorbed carbonate is observed. This indicates a lower
at pH 7.2-7.5 with different concentrations of bicarbonate (A). Panel crystallinity, consistent with the inhibition of crystal growth
B shows the same experiments with 14 mM formate added. Addition by strongly adsorbing anions such as arsenate. Adsorbed
of 2-propanol (not shown) had a similarly small effect as formate. arsenate is visible between 800 and 900 cm-1 (spectrum 5)
(51) and also in the precipitates formed in the presence of
only As(III) (spectrum 4).
Hypothetical Reaction Model and Possible Reaction
Mechanism. The model in Table 2 was constructed by
collecting all the pertinent reactions for which rate constants
have been reported in the literature and in the on-line
database of the Notre Dame Radiation Laboratory (52). Since
this list was insufficient to explain the observed pH depen-
dence, we replaced the Fenton reactions as usually formu-
lated (f1-f3) by the more detailed equations (F1-F3, I1-I5,
J1-J4, and P1-P2). The resulting reaction model is also shown
as a pictorial reaction scheme in Figure 7. Rectangular boxes
mark the added reactions and intermediates.
For simplicity, we have treated all experiments as ho-
mogeneous systems and considered only the reactions
between dissolved species. Other than as a sink for dissolved
Fe(III), we have ignored the role of solids and of surface
reactions. This approximation can be justified by the following
considerations: (i) In all experiments with H2O2, only 20 µM
FIGURE 5. Formation of acetone after mixing of 40 µM H2O2 and iron(III) hydroxides (ca. 2 mg of Fe(OH)3/L) are formed at
1 mM 2-propanol with an equal volume of 40 µM Fe(II) at various the end of the reaction, and most of the Fe(II) and H2O2 is
pH values. Measured data points are indicated with error bars. The consumed before all colloids have formed. (ii) Even at the
lines are the predicted acetone concentrations by the model (Table larger Fe(III) concentration (90 µM) used in the oxidation
2), assuming a yield for acetone of 86% (upper line) and 65% (lower experiments with O2, only small fractions of the reactants
line). (O2, As(III), and Fe(II)) and the intermediates (H2O2) are
adsorbed on the iron (hydr)oxide colloids at circumneutral
Spectrum 1 is of the colloids formed from oxidation of Fe(II) and lower pH values. In the course of the formation of colloids,
in synthetic groundwater at pH 7.0-7.3. The only clear As(V) adsorption does become important, but the fate of
spectral features are the two rather small peaks characteristic As(V) as a final product does not influence the reaction
of lepidocrocite at 743 and 1017 cm-1 and two peaks of kinetics. However, the role of forming colloids (P1 and P2)
adsorbed carbonate at 1375 and 1492 cm-1. The rather as a sink for dissolved Fe(III) is important and was included
uncharacteristic increasing absorbance toward lower wave- in the model in order not to overestimate the reduction of
numbers indicates amorphous ferric hydroxides. Geothite, dissolved Fe(III) by •O2- (R1-R3).

VOL. 37, NO. 12, 2003 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 2739
The key reaction is the Fenton reaction between Fe(II)
and H2O2. As discussed above, several oxidants have been
suggested to be formed, depending on the reaction condi-
tions. The direct formation of free •OH radicals and Fe(III)
in the common formulation of the Fenton reaction (f1-f3)
would imply an outer-sphere electron-transfer reaction
between H2O2 and Fe(II). However, it is widely accepted that
the reaction between Fe(II) and H2O2 occurs via an inner-
sphere mechanism (23, 34, 53), where in a first step one
water ligand is exchanged by H2O2:

[FeII(H2O)6]2+ + H2O2 f
[(H2O)5FeII-O2H2]2+ + H2O (1a)

[(H2O)5FeII(OH)]+ + H2O2 f
[(H2O)4(HO)FeII-O2H2]+ + H2O (1b)

Since the ligand exchange is faster than the following FIGURE 7. Pictorial representation of the model in Table 2. The
reactions but slower than the protonation equilibrium of the different species of Fe(II), As(III), As(V), and the reactive oxygen
water ligands with the surrounding solution, we would expect radicals are not shown in this diagram.
that [(H2O)5FeII-O2H2]2+ and [(H2O)4(HO)FeII-O2H2]+ are at
equilibrium before the next steps occur. could then be formed, more likely by hydrogen abstraction
from a surrounding water molecule than by ligand exchange.
In the rate-determining electron-transfer and internal
In INT-OH, formulated as [(H2O)3FeIII(OH)2(•OH)]+, a second
rearrangement reactions, two intermediates (INT and INT- electron transfer, favored by the electron-donating hydroxyl
OH) are formed: groups, leads to the formation of an Fe(IV) species (e.g.,
[(H2O)3FeIV(OH)3]+).
[(H2O)5FeII-O2H2]2+ f INT (2a) (ii) INT and INT-OH could be both Fe(IV) species. A recent
theoretical study using density functional theory (DFT) (52)
[(H2O)4(HO)FeII-O2H2]+f INT-OH (2b) found that [(H2O)4FeIV(OH)2]2+ is formed by Fe-facilitated
OH-OH bond breaking, loss of an H2O ligand, and hydrogen
abstractions from Fe-bound H2O ligands while formation of
The formulation of the Fe(II) oxidation as parallel reactions a free •OH radical was not favored. In a second DFT study
of differently protonated Fe(II) species retains consistency including surrounding water molecules (55), it was again
with the generally accepted formulation of the Fe(II) oxidation concluded that an Fe(IV) species is favored over the formation
(47, 54). For the rate-determining formation of INT and INT- of a free •OH radical, but it was suggested that the two forms
OH, we have used values close to the reported rate constants [(H2O)4FeIV(OH)2]2+ and [(H2O)3FeIV(OH)3]+ (possibly equiva-
for the oxidation of Fe2+ and FeII(OH)+ by H2O2 (Table 3). lent to INT and INT-OH in our model) are in a protonation
Since the two ligand-exchange reactions 1a and 1b are not equilibrium with the solution. Again, [(H2O)4FeIV(OH)2]2+
rate-determining, we have summarized the reactions 1a and could form a free or bound •OH radical, while [(H2O)3FeIV-
2a as well as 1b and 2b each as a single combined, rate- (OH)3]+ could itself be the final Fe(IV) product with less
determining reaction (F1 and F2) in our model (Table 2). We oxidizing power (expected to be the consequence of an
then assume that once INT and INT-OH are formed, they are additional OH- group) or it could form a ferryl ([OdFeIV-
in a fast, not-rate-determining protonation equilibrium with (H2O)5]2+) species (23, 52, 55). It was suggested that ferryl
the surrounding water: exists in the hydrolitic forms FeIV(OH)22+ and FeIV(OH)3+ above
pH 3 (30), and we would expect [(H2O)3FeIV(OH)3]+ and
possibly [(H2O)3FeIV(OH)4] to dominate at neutral pH. The
INT-OH + H+ a INT (3, I1)
following rate constants for ferryl with other compounds
were measured at pH 1-3 (30): kHCOOH ) 160 M-1 s-1, kHCOO-
The final products are formed in reactions that are faster ) 3 × 105 M-1 s-1, and kethanol ) 2.5 × 103 M-1 s-1. Since these
than 2a and 2b and are thus also not rate determining: rate constants are small in comparison with the expected
rate constants for the reactions of Fe(IV) with As(III) and
INT f FeIII(OH)2+ and •OH (4a, I2) Fe(II) (and are expected to still be lower at higher pH), we
have not included the reaction of ferryl with 2-propanol or
INT-OH f Fe(IV) (4b, I3) with HCOO- in our model.
The suggested partitioning of the reaction pathways does
The protonation equilibrium between INT and INT-OH leads not change the accepted overall kinetics of the reaction of
to a pH-dependent partitioning of the reactions into •OH Fe(II) with H2O2. Both products, •OH and Fe(IV), react with
radicals at low pH and more weakly oxidizing Fe(IV) species Fe(II) in fast and non-rate-determining reactions to Fe(III).
a higher pH. However, the partitioning of the Fenton reaction into •OH-
or •OH-like radicals at low pH and a less reactive Fe(IV) species
The nature of INT and INT-OH and of the final products
at higher pH explains our quenching experiments and is also
cannot be stated with certainty and depends on how the consistent with many other studies and suggestions. Gallard
reactions proceed in detail. We propose two possible et al. studied the degradation of atrazine by the Fenton
sequences: reaction between pH 1 and pH 8 and also suggested an
(i) INT is formed by an one-electron transfer (followed by intermediate that forms •OH at low pH and a different
dissociation of HO-OH and loss of an H2O ligand) and could intermediate, which does not react with atrazine, at higher
be formulated as [(H2O)4FeIII(OH)(•OH)]2+. A free •OH radical pH (37). While the intermediate suggested by Gallard et al.

2740 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 37, NO. 12, 2003
The effect of changes in the reaction scheme on the
optimized rate constants and the SSD is shown in Table 3.
In model 1, Fe(III) precipitation was described simply as a
dimerization (P1) and in model 2a-2c as a polymerization
(P1 and P2). For P1 and P2, the same adjustable rate constants
were used to avoid the introduction of more parameters.
Model 2a led to a lower SSD and to better fits of the data in
Figure 1A as well as to more plausible rate constants for
Fe(III) polymerization than model 1. The rate constant for
dimerization of Fe(OH)3 has been estimated at >1 × 107 M-1
s-1 (58). The description of Fe(III) removal by precipitation
is important, as it affects the reduction of dissolved Fe(III)
species by •O2- (R1-R3) and with it the fits in Figures 3 and
4. Following the suggestion by Sedlak and Hoigne (44), we
have used the same rate constants for the reaction of FeIII-
(OH)2+ and FeIII(OH)2+ with •O2-. Assuming that only FeIII-
(OH)2+ reacts with •O2- still leads to good fits in Figures 1 and
2 but to unsatisfactory fits of the data in Figures 3 and 4. The
FIGURE 8. Sum of squared differences between the model and the role of Fe(III) speciation and precipitation in Fenton systems
data (205 data points in the 27 experiments shown in Figures 1-4) at neutral pH is still difficult to describe and will require
as a function of varying each constant around the optimized value,
further attention.
while keeping the other constants fixed at the optimal value. The
resulting parabolas show that all rate constants except k4 are well The reaction scheme in models 2a-2c is identical, but
constrained. the rate constants for Fe(II) oxidation are different. In model
2b, all rate constants for the Fe(II) oxidation were left
adjustable. This increased flexibility slightly improved the
was not able to oxidize atrazine, it might well be able to
fits, but the output of model 2a (not shown) is almost
oxidize As(III).
indistinguishable from the output of model 2b (shown in the
Role of Carbonate and Carbonate Radicals. In ground- figures). Model 2c, with the rate constants for Fe(II) oxidation
water with high bicarbonate concentrations, additional fixed to the values according to Millero et al., results in a
pathways are possible. King and Farlow (56) have found that somewhat faster oxidation of As(III) than observed in Figure
iron(II) carbonate complexes contribute significantly to the 2B-D, but the fits for all other figures were the same as those
oxidation of Fe(II) by H2O2 in the presence of 2 mM NaHCO3. of model 2b. Various studies have found slightly different
The importance of Fe(II) species decreases in the order: FeII- rate constants for the oxidation of Fe2+ and FeIIOH+ with
CO3 > FeIIOH- > FeII(CO3)22- at pH 5.5-7.3 and FeIICO3 > H2O2 (see last column in Table 3). Given the experimental
FeII(CO3)22- > FeIIOH- at pH >7.3. uncertainties (e.g., 0.1-0.2 pH unit increase during the
In our model, we have included only the FeIICO3, as each reactions and ∼10% relative errors in the As(III) determina-
additional species in the model also requires the addition of tions) and the simplification in the model (e.g., only one
the corresponding reactions with Fe(IV) and with •O2-. The carbonate species), the log(k) values for the oxidation of Fe-
fitted log(k) for the reaction of H2O2 with FeIICO3 was between (II) species are in good agreement with the values reported
the values reported for FeIICO3 and for FeII(CO3)22- (see Table by others. The most important part of the model, the pH-
3). dependent partitioning of the Fenton reaction into •OH and
For the reaction of FeIICO3 with H2O2, we propose the another oxidizing species (suggested to be Fe(IV)) was rather
formation of an intermediate INT-CO3, which can form either invariant to differences in the other rate constants or to the
carbonate radicals, Fe(IV), or both. (As the pKa of the exact description of Fe(III) precipitation.
bicarbonate/carbonate radical is reported as <0 (57), we have Finally, it should be noted that the proposed model does
considered only CO3•-. The partitioning into Fe(IV) and CO3•- not mean that no •OH radicals are formed at neutral pH.
pathways was left open by leaving one of the fast and non- Rather, the model states that the yield of •OH formation drops
rate-determining rate constants (I4 and I5) adjustable, and sharply between pH 5 and pH 6 and 10-fold with each pH
the fitting suggests that 4.5 times more Fe(IV) than CO3•- (I4 unit increase above pH 6. Compounds that only react with
and I5) is formed. The simplified model for carbonate •OH can still be oxidized in the Fenton reaction at pH 7,

reproduces the trends observed in Figure 4 but is not able although with low yields, and the addition of 2-propanol or
to fully account for the experimental results. The role of iron- other •OH scavengers inhibits their oxidation.
(II) carbonate species and of carbonate radicals will need Environmental and Technical Significance. The finding
further attention; however, we do not currently have sufficient that As(III) is partly oxidized in parallel to the oxidation of
data to add more reactions into the proposed model. Fe(II) is significant in all situations where oxygen is introduced
Uncertainties of the Model. Figure 8 shows the variation into circumneutral waters containing As(III) and Fe(II). The
of the adjustable rate constants and the resulting sum of co-oxidation of As(III) with Fe(II) by O2 might be one of the
squared differences (SSD) for model 2b. For the fitting, all most important abiotic pathways for As(III) oxidation and
data were normalized such that the highest concentration might help to explain why both As(III) and As(V) are typically
was 1.0 in each kinetic run to give all experiments the same found in groundwater samples after they have been pumped
weight. The minimum total SSD of 1.26 (205 data points in to the surface.
27 experiments) translates to an average standard deviation The oxidation of As(III) with H2O2 and Fe(II) is important
of 8.3%, which agrees with the estimated 10% error between in natural and technical systems at pH values up to 8, below
the different experiments. The steep parabolic shapes (except which As(III) is not oxidized at appreciable rates by H2O2
for k4) indicate that the fits are well constrained and that the alone. H2O2 can occur naturally at micromolar concentrations
resulting rate constants are reliable to within 0.5 log unit in in rainwater and in soil waters, and it is used at millimolar
the constraint of this model and data set. The value for k4 concentrations in water treatment. The observation that the
cannot be determined better in these experiments, as the As(III) oxidation at circumneutral pH is not quenched by
reaction of Fe(IV) with Fe2+ does not contribute significantly typical •OH scavengers such as 2-propanol or formate (in
under the employed experimental conditions. contrast to the almost complete quenching at low pH) is

VOL. 37, NO. 12, 2003 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 2741
important in natural and technical systems as it can be (27) Nakamoto, K. J. Mol. Struct. 1997, 408/409, 11-16.
expected that naturally present dissolved organic matter (28) Logager, T.; Holcman, J.; Sehested, K.; Pedersen, T. Inorg. Chem.
would also not significantly affect As(III) oxidation by 1992, 31, 3523-3529.
(29) Rush, J. D.; Bielski, B. H. J. J. Am. Chem. Soc. 1986, 108, 523-
scavenging •OH. Since the rate and extent of the As(III) 525.
oxidation strongly depends on pH and As(III), Fe(II), O2, H2O2, (30) Jacobsen, F.; Holcman, J.; Sehested, K. Int. J. Chem. Kinet. 1998,
and HCO3- in concentrations ranges that are common in 30, 215-221.
natural and technical systems, a good quantitative under- (31) Rush, J. D.; Koppenol, W. H. J. Am. Chem. Soc. 1988, 110, 4957-
standing of these reactions is important. The proposed kinetic 4963.
model provides a base for the understanding of the important (32) Rahhal, S.; Richter, H. W. J. Am. Chem. Soc. 1988, 110, 3126-
3133.
factors in these systems and for the optimization of arsenic (33) Luzzatto, E.; Cohen, H.; Stockheim, C.; Wieghardt, K.; Meyerstein,
removal methods. D. Free Radical Res. 1995, 23, 453-463.
(34) Goldstein, S.; Meyerstein, D.; Czapski, G. Free Radical Biol. Med.
Acknowledgments 1993, 15, 435-445.
We thank Thomas Ruettimann (EAWAG) for technical (35) Goldstein, S.; Meyerstein, D. Acc. Chem. Res. 1999, 32, 547-550.
(36) Aplin, R.; Waite, T. D. Water Sci. Technol. 2000, 42, 345-354.
assistance; Bernhard Wehrli, Silvio Canonica, Urs von Gunten, (37) Gallard, H.; de Laat, J.; Legube, B. New J. Chem. 1998, 22, 263-
and Barbara Sulzberger (EAWAG) for valuable discussions; 268.
and the Alliance for Global Sustainability for financial support (38) Balmer, M.; Sulzberger, B. Environ. Sci. Technol. 1999, 33, 2418-
in the framework of the arsenic mitigation project in 2424.
Southeast Asia. (39) Haywood, M. G.; Riley, J. P. Anal. Chim. Acta 1976, 85, 219-230.
(40) Johnson, D. L.; Pilson, M. E. Q. Anal. Chim. Acta 1972, 58, 289-
299.
Supporting Information Available (41) Yamamoto, M.; Urata, K.; Murashige, K.; Yamamoto, Y. Spec-
The input file for ACUCHEM with the complete reaction trochim. Acta Part B 1981, 36, 671-677.
scheme of model 2b. This material is available free of charge (42) Zwank, L.; Schmidt, T. C.; Haderlein, S. B.; Berg, M. Environ.
via the Internet at http://pubs.acs.org. Sci. Technol. 2002, 36, 2054-2059.
(43) Braun, W.; Herron, J. T.; Kahaner, D. ACUCHEM, Computer
Program for Modeling Complex Reaction Systems; National
Literature Cited Bureau of Standards: Gaithersburg, MD.
(1) Kinniburgh, D. G.; Smedley, P. L. Arsenic contamination of (44) Sedlak, D. L.; Hoigné, J. Environ. Sci. Technol. 1994, 28, 1898-
groundwater in Bangladesh; Final Report Summary; Bangladesh 1906.
Department for Public Health Engineering, British Geological (45) Voelker, B. M.; Morel, F. M. M.; Sulzberger, B. Environ. Sci.
Survey: 2000. Technol. 1997, 31, 1004-1011.
(2) Berg, M.; Tran, H. C.; Nguyen, T. C.; Pham, H. V.; Schertenleib, (46) Hug, S. J.; Laubscher, H.-U.; James, B. R. Environ. Sci. Technol.
R.; Giger, W. Environ. Sci. Technol. 2001, 35, 2621-2626. 1997, 31, 160-170.
(3) National Research Council. Arsenic in drinking water; National (47) Millero, F. J.; Sotolongo, S.; Izaguirre, M. Geochim. Cosmochim.
Academy Press: Washington, DC, 1999. Acta 1987, 51, 793-801.
(4) Manning, B. A.; Goldberg, S. Soil Sci. 1997, 162 (12), 886-895. (48) Asmus, K. D.; Mockel, H.; Henglein, A. J. Phys. Chem. 1973, 77,
(5) Hering, J. G.; Chen, P. Y.; Wilkie, J. A.; Elimelech, M.; Liang, S. 1218-1221.
J. Am. Water Work Assoc. 1996, 88, 155-167. (49) Klaning, U. K.; Bielski, B. H. J.; Sehested, K. Inorg. Chem. 1989,
(6) Meng, X. G.; Bang, S.; Korfiatis, G. P. Water Res. 2000, 34 (4), 28, 2717-2724.
1255-1261. (50) Cornell, R. M.; Schwertmann, U. The Iron Oxides; VCH Ver-
(7) Borho, M.; Wilderer, P. Water Sci. Technol. 1996, 34, 25-31. lagsgesellschaft: Weinheim, Germany, 1996.
(8) Kim, M. J.; Nriagu, J. Sci. Total Environ. 2000, 247, 71-79. (51) Voegelin, A.; Hug, S. J. Environ. Sci. Technol. 2003, 37, 972-978.
(9) Pettine, M.; Campanella, L.; Millero, F. J. Geochim. Cosmochim. (52) The Radiation Chemistry Data Center of the Notre Dame
Acta 1999, 63, 2727-2735. Radiation Laboratory, 2002; http://allen.rad.nd.edu/.
(10) Pettine, M.; Millero, F. J. Mar. Chem. 2000, 70, 223-234. (53) Buda, F.; Ensing, B.; Gribnau, M. C. M.; Baerends, E. J. Chem.-
(11) Muller, B.; Granina, L.; Schaller, T.; Ulrich, A.; Wehrli, B. Environ. Eur. J. 2001, 7, 2775-2783.
Sci. Technol. 2002, 36, 411-420. (54) Stumm, W.; Lee, G. F. Ind. Eng. Chem. 1961, 53, 143-146.
(12) United States Environmental Protection Agency, Office of Water (55) Ensing, B.; Buda, F.; Blochl, P.; Baerends, E. J. Angew. Chem. Int.
(4606). EPA 815-R-00-028, 2000; Vol. 2000. http://www.epa.gov/ Ed. 2001, 40, 2893-2895.
safewater/ars/treatments_and_costs.pdf. (56) King, D. W.; Farlow, R. Mar. Chem. 2000, 70, 201-209.
(13) Woods, R.; Kolthoff, I. M.; Meehan, E. J. J. Am. Chem. Soc. 1964, (57) Czapski, G.; Lymar, S. V.; Schwarz, H. A. J. Phys. Chem. A 1999,
86, 1698. 103, 3447-3450.
(14) Banerjee, K.; Helwick, R. P.; Gupta, S. Environ. Prog. 1999, 18, (58) Schneider, W. Chimia 1988, 42, 9-20.
280-284. (59) Stuglik, Z.; Zagorski, Z. P. Radiat. Phys. Chem. 1981, 17, 229-
(15) Krishna, B. M. V.; Chandrasekaran, K.; Karunasagar, D.; 233.
Arunachalam, J. J. Hazard. Mater. 2001, B84, 229-240.
(60) Huie, R. E.; Shoute, L. C. T.; Neta, P. Int. J. Chem. Kinet. 1991,
(16) Hug, S. J.; Canonica, L.; Wegelin, M.; Gechter, D.; von Gunten,
23 (6), 541-552.
U. Environ. Sci. Technol. 2001, 35, 2114-2121.
(61) Millero, F. J.; Sotolongo, S. Geochim. Cosmochim. Acta 1989, 53,
(17) Haber, F.; Weiss, J. Proc. R. Soc. London, Ser. A 1934, 332-351.
1867-1873.
(18) Barb, W. G.; Boxendale, J. H.; George, P.; Hargrove, K. R. Trans.
(62) Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. J.
Faraday Soc. 1951, 47, 591-616.
Phys. Chem. Ref. Data 1988, 17, 513-886.
(19) Rush, J. D.; Bielski, B. H. J. J. Phys. Chem. 1985, 89, 5062-5066.
(63) Draganic, Z. D.; Negron-Mendoza, A. S., K.; Vujosevic, S. I.;
(20) Koppenol, W. H. Redox Rep. 2001, 6, 229-234.
Navarro-Gonzales, R.; Albarran-Sanchez, M. G.; Draganic, I. G.
(21) Liochev, S. I.; Fridovich, I. Redox Rep. 2002, 7, 55-57.
Radiat. Phys. Chem. 1991, 38 (3), 317-321.
(22) Wink, D. A.; Nims, R. W.; Desrosiers, M. F.; Ford, P. C.; Keefer,
(64) Lilie, J.; Hanrahan, R. J.; Henglein, A. Radiat. Phys. Chem. 1978,
L. K. Chem. Res. Toxicol. 1991, 4, 510-512.
11, 225-227.
(23) Bossmann, S. H.; Oliveros, E.; Göb, S.; Siegwart, S.; Dahlen, E.
P.; Payawan, L., Jr.; Straub, M.; Wörner, M.; Braun, A. M. J. Phys. (65) Smith, R. M., Martell, A. E., Eds. Critical Stability Constants;
Chem. A 1998, 102, 5542-5550. Plenum Press: New York, 1976.
(24) Kremer, M. L. Phys. Chem. Chem. Phys. 1999, 1, 3595-3605.
(25) Pignatello, J. J.; Liu, D.; Huston, P. Environ. Sci. Technol. 1999, Received for review October 3, 2002. Revised manuscript
33, 1832-1839. received April 2, 2003. Accepted April 7, 2003.
(26) Reinke, L. A.; Rau, J. M.; McCay, P. B. Free Radical Biol. Med.
1994, 16, 485-492. ES026208X

2742 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 37, NO. 12, 2003

You might also like