You are on page 1of 8

Article

pubs.acs.org/est

Early Stage Formation of Iron Oxyhydroxides during Neutralization


of Simulated Acid Mine Drainage Solutions
Mengqiang Zhu,†,* Benjamin Legg,‡ Hengzhong Zhang,‡ Benjamin Gilbert,† Yang Ren,§
Jillian F. Banfield,†,‡ and Glenn A. Waychunas†

Earth Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States

Department of Earth and Planetary Science, University of California-Berkeley, Berkeley, California, 94720, United States
§
Advanced Photon Source, Argonne National Laboratory, Argonne, Illinois 60439, United States
*
S Supporting Information

ABSTRACT: The phases and stability of ferric iron products


formed early during neutralization of acid mine drainage
waters remain largely unknown. In this work, we used in situ
and time-resolved quick-scanning X-ray absorption spectros-
copy and X-ray diffraction to study products formed between 4
min and 1 h after ferric iron sulfate solutions were partially
neutralized by addition of NaHCO3 ([HCO3−]/[Fe3+] < 3).
When [HCO3−]/[Fe3+] = 0.5 and 0.6 (initial pH ∼ 2.1 and
2.2, respectively), the only large species formed were sulfate-
complexed ferrihydrite-like molecular clusters that were stable
throughout the duration of the experiment. When [HCO3−]/[Fe3+] = 1 (initial pH ∼ 2.5), ferrihydrite-like molecular clusters
formed initially, but most later converted to schwertmannite. In contrast, when [HCO3−]/[Fe3+] = 2 (initial pH ∼ 2.7),
schwertmannite and larger ferrihydrite particles formed immediately upon neutralization. However, the ferrihydrite particles
subsequently converted to schwertmannite. The schwertmannite particles formed under both conditions aggregated extensively
with increasing time. This work provides new insight into the formation, stability and reactivity of some early products that may
form during the neutralization of natural acid mine drainage.

■ INTRODUCTION
When exposed to air and water, metal sulfide deposits and coal
Fe oxyhydroxide phase stability and their behavior in the
environment.
It is generally believed that solvated Fe3+ undergoes
mines produce iron-rich sulfuric acid solutions that can contain
hydrolysis with subsequent formation of low-molecular-weight
a broad range of toxic elements.1,2 These solutions, referred to
Fe3+ species (hydrated and hydrolyzed monomers, dimers,
as acid mine drainage (AMD), degrade environmental quality.
trimers, etc.).6 Fe oxyhydroxide particles may nucleate from the
Neutralization of AMD occurs when solutions mix with natural
polynuclear species6 and grow via atom-by-atom addition or
waters, such as natural rivers and streams, during wetland oriented aggregation.7 In addition to hydrolysis, olation and
treatment,3 through reaction with rock surfaces, or following oxolation processes are responsible for these condensation
lime addition in remediation processes. As a result, ferric iron processes.8 The resultant Fe oxyhydroxide phases depend on
(Fe3+) hydrolyzes and precipitates as Fe oxyhydroxide anion type and pH. The impact of anions in phase organization
nanoparticles such as ferrihydrite and schwertmannite. These originates from their Fe3+ binding ability, and this is strongly
Fe oxyhydroxide nanoparticles can transform to more affected by pH. In general, weak Fe3+ binding anions, such as
thermodynamically stable phases goethite and hematite with NO3− and ClO4−, favor ferrihydrite formation. Stronger Fe3+-
time. binding ligands, such as Cl− and SO42‑, lead to formation of
During the formation of Fe oxyhydroxide nanoparticles, toxic akaganeite and schwertmannite, respectively, under acidic
elements in solution can be incorporated into the nanoparticle conditions. Under neutral and alkaline pH conditions,
structure. During or after formation, contaminants can also be ferrihydrite tends to be the resultant phase because OH−
adsorbed onto nanoparticle surfaces or trapped in interfacial outcompetes Cl− and SO42‑ to bind Fe3+. In the presence of
regions between aggregated nanoparticles. Such processes very strong Fe3+-binding ligands, such as phosphate and silicate,
influence contaminant fate and transport.4 Nanoparticles also
can interact with microorganisms and their extracellular Received: March 30, 2012
polymers,5 and hence extracellular substances, and other Revised: June 25, 2012
solution species may promote or retard particle growth, Accepted: July 5, 2012
transformation and aggregation. These processes can control Published: July 5, 2012

© 2012 American Chemical Society 8140 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

Fe oxyhydroxide particle formation can be substantially and two-line ferrihydrite reference materials were synthesized
impaired.9,10 using the procedures described by Schwertmann and Cornell.14
Aqueous speciation in pure Fe3+ solutions is relatively well UV-Vis Spectroscopy and pH Measurement. Each of
understood. In recent years, much has been learned about the mixed solutions was immediately transferred to a quartz
nanoparticle structure, growth, and phase transformation cuvette with 1 mm path length for time-resolved UV-vis spectra
behavior.11 However, a major gap in knowledge concerns the collection using an Agilent UV-vis spectrophotometer (model
nucleation pathway whereby aqueous monomeric Fe3+ species 8453). UV-vis spectra were continuously recorded with 5 s of
are converted into molecular clusters and nanoparticles. Prior integration time. Time-dependent pH values of the above
work attempted to study early stage Fe oxyhydroxide formation solutions were automatically recorded every 20 s. The pH
kinetics and mechanisms in partially neutralized Fe3+ solutions meter was calibrated using pH 1, 1.68, 2, and 4 buffer solutions.
using various methods, including titration, UV-vis spectroscopy, Synchrotron X-ray Diffraction (SXRD). Time-resolved
wide-angle X-ray scattering diffraction (WAXS), small-angle X- SXRD 2-D images were collected at beamline 11-ID-B at the
ray scattering (SAXS), extended X-ray absorption fine structure Advanced Photon Source (APS). The high flux of the X-rays of
(EXAFS) spectroscopy, electron microscopy, and analytical this beamline is the key for collecting high signal-to-noise ratio
centrifugation.6,12,13 Effects of anion type, ratio of neutralization XRD data from particles in dilute particle suspensions. A
base to Fe3+, Fe3+ concentration, aging period, and temperature, detailed description for XRD collection and processing is
were examined. However limited knowledge was obtained provided in the Supporting Information (SI-1).
about the precipitate phase, average local atomic structure, QEXAFS Spectroscopy. Each of the mixed solutions was
particle size and morphology, and aggregate structure. In transferred to a 2 mm thick, 15 mm long, and 10 mm wide cell
particular, due to the limited time-resolution of the made from acrylic plastic plate. Both sides of the cell were
experimental methods, these investigations were confined to sealed with Kapton film to hold the solutions. QEXAFS spectra
late reaction stages (>30 min) when particle aggregation and were collected at one scan per 2 s in transmission mode. Fifteen
phase transformation comprise the major events. Although the or 14 continuous scans of QEXAFS spectra were averaged to
final precipitate phases could be identified in these studies, the improve counting statistics, resulting in a time-resolution of 30
forms and structures of the early formed products in the s. The measurements were conducted at beamline X18A at the
neutralization process remained largely unknown, and hence National Synchrotron Light Source (NSLS), using a specially
their formation kinetics and mechanism are yet to be explored. equipped monochromator with a Si (111) crystal. The
In the current study, we were able to investigate the formation monochromator was detuned by 35% with respect to I0 to
kinetics and phase identities of early formed products in minimize higher order harmonic X-rays. The detailed beamline
partially neutralized Fe3+-sulfate solutions using time-resolved and monochromator setup for the QEXAFS measurement can
UV-vis spectroscopy, quick-scanning EXAFS (QEXAFS) spec- be found elsewhere.15
troscopy, and synchrotron XRD, by combining these rapid Powder samples, including the Fe oxyhydroxide references
methods with a slowed neutralization process. and the solid precipitates in the mixed solutions at h = 1 and 2,

■ EXPERIMENTAL SECTION
Sample Preparation. All chemicals used were of analytical
were ground finely and uniformly spread on Scotch tape. The
tapes were folded to four or eight layers for EXAFS data
collection in transmission mode at beamline 4-1 at the Stanford
grade. A 0.8 M Fe3+ stock solution ([SO42‑]/[Fe3+] = 1.5) was Synchrotron Radiation Lightsource (SSRL), using a Si (220)
prepared by adding 4.16 g of ferric iron sulfate powder monochromator crystal.
(containing 21−23% Fe by weight) to 16 mL deionized (DI) After background removal and normalization the EXAFS
water. The Fe solution was used within 48 h after preparation, spectra were converted into the k-weighted function, k3χ(k),
although there were no visible changes in color and turbidity and Fourier transforms |χ(R)| were calculated over the k range
over 1 month. Reactions were initiated by slowly pumping 4 of 3−14 Å−1 using the Bessel-Kaiser window function. Linear
mL DI water, 0.2 M, 0.24 M, 0.4 M, or 0.8 M NaHCO3 combination fitting (LCF) analysis of QEXAFS spectra was
solution at 5 mL·min−1 using a syringe pump, into a 30 mL performed over 3−12 Å−1 in k space. In the fits, the component
plastic bottle containing 4 mL of 0.4 M Fe3+ sulfate solution (2 weight sum was forced to equal 1.0. The program Athena16 was
mL 0.8 M Fe stock solution plus 2 mL DI water) to achieve used for all of the above data reduction and processing. Further
[HCO3−]/[Fe3+] molar ratios (represented by h) of 0, 0.5, 0.6, EXAFS shell-by-shell fitting was performed using the SixPack
1, and 2. An acidified Fe solution was prepared by adding 4 mL program suite17 to obtain the detailed local atomic environment
of concentrated HNO3 to 4 mL of 0.4 M Fe3+ sulfate solution, of iron atoms for representative samples.
referred to as FeAcid. This solution is supposed to contain Ferron Assay. Ferron (8-hydroxy-7-iodo-5-quinoline sul-
mainly Fe(H2O)63+. The final concentration of total Fe3+ in all fonic acid) has been used for semiquantification of Fe3+
solutions was 0.2 M. Vigorous stirring was enforced during the monomers, polymers and precipitates based on distinct
base addition to minimize local oversaturation with respect to ferron−Fe complex formation rates. Fe3+ monomers are
Fe oxyhydroxide phases. Each base addition step took 48 s. complexed by ferron rapidly, but the breakdown of larger
Upon completion, one additional minute of stirring was used to Fe3+ species to form ferron−Fe complexes occurs much more
remove CO2 gas resulting from NaHCO3 decomposition slowly.18 It has been assumed that reactions of monomeric Fe3+
during the neutralization reaction. Then, the solutions were species with ferron are complete within 1 min (denoted as
immediately transferred for the following time-resolved Fe_monomer).18 Species that form ferron−Fe complexes over
analyses. Measurements of reaction time started from the first the next 12 h were interpreted as Fe polymers, and any Fe
drop of base addition. When the first data sets were recorded, species that remained unreactive after 12 h were considered as
about 2−5 minutes had elapsed, depending on time-resolved stable precipitates.18 Following this method, we used the ferron
analytical approaches (see below). All experiments were assay to estimate Fe3+ precipitation reaction progress during
conducted at room temperature of 24−25 °C. Schwertmannite neutralization of ferric salts by NaHCO3, based on Fe_mo-
8141 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

nomer fractions measured at different times as the neutraliza- addition, ferric Fe species with an extended structure, such as
tion reaction proceeded. The specific experimental procedures Fe oxyhydroxide particles, have a characteristic absorption
were described in the Supporting Information (SI-2). The feature called electron pair transition (EPT) band.22 The EPT
accuracy of ferron assay-based inferences was re-evaluated in band is significant for particles but very weak for molecular
light of other results of this study (see below). clusters, and thus can be used to differentiate the two types of

■ RESULTS
Color and Turbidity. The visible appearance of the
Fe species.
Figure 2a−d shows UV-vis spectra of the solutions at
different h. For clarity, the spectra were presented in two
solutions immediately following bicarbonate addition was different wavelength ranges. The spectra of h = 1 and h = 2
dependent on the base to ferric iron molar ratio (h = solutions evolved with time whereas others remained the same.
[HCO3−]/[Fe3+]). All h ≥ 0.5 solutions displayed red wine The acidified solution (FeAcid) only had a very weak band at
color, and the color deepened as h increased. Moreover, the ∼806 nm (Figure 2b). The h = 0 solution had two broad
cloudiness (or turbidity) of the h = 1 and h = 2 solutions absorption bands centered at approximately 425 and 830 nm.
increased over time, whereas other solutions remained clear For the h = 0.5 and 0.6 solutions, three bands at about 430,
during the experimental period. To monitor the turbidity 448, and 850 nm can be discerned (Figure 2b and d), and the h
changes with time, the measured absorbance at 600 nm was = 0.6 solution bands are more intense. The barely changed
used as a turbidity indicator, as no samples had UV-vis spectra of the above solutions indicate that Fe species in these
absorption bands at this wavelength (see below). Assuming that solutions were stable within the experimental time period. The
Rayleigh scattering applies, increased absorbance was due to first spectrum of the h = 1 solution also showed the same three
increase of size and/or numbers of scattering objects.19 As bands as the h = 0.5 solution (Figure 2a and b). However, the
shown in Figure 1, absorbance at 600 nm increased with first two bands gradually disappeared with reaction time, and
the 850 nm band decreased its intensity and shifted its position
to 862 nm. Another band at 485 nm, not obvious in the first
spectrum of the h = 1 solution, emerged and became
increasingly pronounced (Figure 2b). This band did not exist
in the spectra of the h = 0, 0.5, and 0.6 solutions.
The observed bands at 425, 430, and 448 nm can be assigned
to 6A1 g → (4A1, 4Eg) transitions and the bands at 830, 850, and
862 nm to 6A1 g→ 4T1 g transitions.11,21 The substantial
existence of all these bands indicates formation of Fe clusters
or particles.21,23 The 485 nm band is assigned to EPT.22 The
absence of this band in the h = 0, 0.5, 0.6 and the initial h = 1
solutions suggests that the Fe species were molecular clusters.
The disappearance of the three bands of the h = 1 solution
shows that the Fe cluster was decreasing in abundance, whereas
the increasing EPT absorbance indicates that a Fe oxyhydroxide
phase was forming and increasing in abundance. The UV-is
spectrum of the filtered h = 1 solution after 2 h of reaction time
Figure 1. UV-vis absorbance at 600 nm versus reaction time for Fe still showed the three bands, although they were much less
sulfate solutions at various h values (h = [NaHCO3]/[Fe3+], i.e., intense (Figure 2b and d), indicating that the Fe clusters had
neutralization ratio). Since no significant UV-vis absorption bands exist not completely disappeared.
at this wavelength, the measured absorbance was due to particle
scattering, indicating turbidity of the solutions.
Regarding the h = 2 solution, the EPT band at 485 nm is
pronounced in the first collected spectrum, indicating that Fe
oxyhydroxide particles formed from the beginning. A band at
increasing h when reaction time < 27 min, indicating particle 865 nm (assigned to 6A1 g→ 4T1 g) decreased in intensity with
formation. For the h = 1 solution, the 600 nm absorbance time without significant position change. In addition, the
increased in the first ∼10 min (see inset) and dramatically after absorbance between 525 and 600 nm decreased with time,
∼26 min. For the h = 2 solution, an initially subtle decrease was suggesting disappearance of a metastable phase. The filtered h =
followed by a dramatic increase in scattering after ∼28 min 2 solution after 2 h had a weak peak at 850 nm besides the 485
(Figure 1). The dramatic increases correlated with the nm band (Figure 2b and d), indicating the solution also had a
transitions from clear red wine-like solutions to cloudy yellow small amount of Fe clusters similar to those present in the h = 1
suspensions. The cloudiness was due to the massive formation solutions. The lack of obvious 430 and 448 nm bands in this
of large particle aggregates.19 The aggregates then settled to case could be due to the overlap with the intense absorbance
form precipitates that are referred to as Ppt.h1 and Ppt.h2 for h tail (300−600 nm) from the larger particles.
= 1 and h = 2, respectively. Phase Analysis by Synchrotron XRD (SXRD). To
Fe3+ Molecular Clusters and Particles Determined by identify the phases of particles being formed, in situ SXRD
UV-vis Spectroscopy. Fe3+ monomers, molecular clusters and data were collected from the neutralized solutions. As shown in
particles have different optical excitation features that can be Supporting Information Figure SI-1, the first XRD pattern of
used for identification. Ligand-field or d → d transitions are the h = 1 solution cannot be firmly attributed to known phases
spin forbidden and thus are extremely weak in Fe 3+ due to the low particle abundance and/or small particle size.
monomers.20,21 By contrast, the ligand-field bands of Fe3+ The first pattern of the h = 2 solution presented broad peaks,
clusters and particles are intense mainly due to antiferromag- indicating formation of very fine crystalline particles. With
netic coupling between next neighboring Fe atoms.11,21 In increasing reaction time, the initial broad peaks of both h = 1
8142 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

Figure 2. UV-vis spectra of Fe sulfate solutions at various h values. The spectra of the h = 1 and 2 solutions were time dependent and the arrows
indicate the reaction directions with increased reaction time (2 min < t < 27 min). The “filtered” spectra were collected after filtration of the cloudy
solutions to remove solids. For closer inspection, the UV-vis spectra were presented in two different wavelength regimes (note that the absorbance
scales differ). The spectral amplitude of the h = 1 and 2 solutions increased over the whole wavelength range due to particle scattering, correlating
with the transition from clear to cloudy solutions due to onset of precipitation. The spectra collected after the solutions became cloudy (top black
curves in panels (a) and (c)) are not shown in (b) and (d).

Figure 3. Fe K-edge EXAFS spectra of HNO3 acidified Fe sulfate solution (FeAcid) (brown), the neutralized Fe sulfate solutions at different reaction
times and different h values (pink, green, cyan, blue), the final precipitates at h = 1 and 2 (red), with 2-line ferrihydrite (Fhy2L) and schwertmannite
(Schw) reference spectra. (a) k-weighted χ(k) and (b) the magnitude (|χ(R)|) of their Fourier transforms.

and h = 2 solutions sharpened (2.4 Å−1) and split (4 Å−1 and combination fitting (LCF), were used to further identify the
5.5 Å−1), increasingly resembling those of the schwertmannite particle phases and quantify their amounts versus reaction time.
reference phase. The SXRD results showed that the precipitates Local Atomic Environment of Fe. Fe K-edge EXAFS
(Ppt.h1 and Ppt.h2) formed in the two solutions were primarily spectroscopy provides information about the local atomic
environment around central Fe (within 6 Å). Figure 3 displays
schwertmannite (Supporting Information Figure SI-1). How-
the results of in situ QEXAFS spectroscopy acquired from
ever, XRD phase characterization for particles in solutions samples under the same conditions as the XRD data of
based on poorly defined peaks has large uncertainties due to Supporting Information Figure SI-1. The peaks at 1.5−1.6 Å (R
solution interference. In the following two sections, QEXAFS +ΔR) in |χ(R)| are assigned to the nearest O shell at a Fe−O
spectroscopic analyses, including EXAFS curve fitting and linear distance of ∼2 Å, corresponding to Fe−O bond lengths. As the
8143 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

Figure 4. Molar fractions of ferrihydrite (red circles), schwertmanite (blue triangles), and h = 0.5 solutions (green triangles) were obtained from
linear combination fitting for (a) h = 1 and (b) h = 2. Values measured via the ferron assay are indicated by black boxes. Comparison of the inferred
molar fraction of ferrihydrite plus the h = 0.5 solution (open boxes) to the ferron assay values indicates that the ferron assay actually measures both
of these components at h = 1.

neutralization degree increased from the acidified solution to introduces further diversity.25,26 As it is impossible to identify
the h = 2 solutions, the Fe−O bond length shortened, the O and prepare LCF standards for all possible Fe species in the
peaks broadened, and the amplitudes decreased (Figure 3). The neutralized solutions, we assume that the h = 0.6, 1, and 2
bond length shortening can be due to the formation of particles solutions can be represented by a combination of solutions at
and clusters that usually have shorter Fe−O bonds in Fe−OH lower neutralization ratios (representing monomeric Fe3+ and
and Fe−OFe moieties compared to those in Fe−OH2 of small Fe3+ clusters) and Fe oxyhydroxide solid phases
monomeric Fe species. These changes also widen the overall approximating particles in aqueous solutions. FeAcid, the h =
Fe−O bond length distribution and cause destructive 0, 0.2, and 0.5 solutions, schwertmannite, Fhy2L, goethite,
interference in the χ(k) functions and accordingly the reduced lepidocrocite, hematite, Ppt.h1, Ppt.h2, and their combinations,
O peak amplitudes in their Fourier transforms.24 The wider thus were tested with LCF analysis to determine if well-defined
Fe−O bond distribution is consistent with the increasing general fits could be obtained with a minimum set of
Debye−Waller factors (σ2) obtained from fitting using a single components. It was found that a combination of the h = 0.5
O shell (Supporting Information SI-3). solution, Ppt.h2 and Fhy2L gave the best fits with the least
The peaks at ∼2.6 and ∼3.05 Å (R+ΔR)| are assigned to the number of components (fits in Supporting Information SI-6
edge- and corner-sharing Fe shells at the Fe−Fe distances of and results in Figure 4). Use of Ppt.h2 and h = 0.5 solution only
∼3.0 and ∼3.35 Å, respectively. Sulfate can complex with in the fits decreased the goodness of fit (in terms of reduced χ2)
Fe25,26 and sulfur scattering may slightly modify the peaks up to 33% (the first spectrum of the h = 1 solution), showing
corresponding to approximately these distances. As the necessity of the Fhy2L component.
neutralization degree increases, the two Fe peaks become The LCF analysis indicates that Ppt.h1 contained Fhy2L at a
more and more pronounced, indicating Fe polymerization. phase fraction 17% and Ppt.h2 at 83%. The h = 0.6 solution
FeAcid did not exhibit the edge-sharing Fe peak (Figure 3), contained Fhy2L at 4% and the h = 0.5 solution at 96%, but this
showing that the high concentration of acid suppressed was subject to large uncertainties due to the low abundance of
formation of Fe polymers. The pronounced peaks between Fhy2L. Although the EXAFS spectrum of Ppt.h2 was almost
2.8 and 3.75 Å (R+ΔR) may result from both multiple- identical to that of schwertmannite reference phase (Supporting
scattering paths within the first O shell of Fe octahedra and Information SI-5), replacing Ppt.h2 with the schwertmannite
single-scattering paths from the O atoms in the second reference slightly decreased the fitting accuracies but gave
hydration shell.27 The peak at 3.1 Å (R+ΔR) overlaps with similar component fraction values. Henceforth, for clarity, we
the corner-sharing Fe peak, making it hard to discern the Fe use “schwertmannite” to represent “Ppt.h2” in further
shell when particle abundance is low, such as the spectra of the discussion of the LCF results that were summarized in Figure 4.
h < 1 solutions and the initial spectrum of the h = 1 solution. The Fhy2L component accounted for ∼20% of Fe species
For the h = 1 and 2 solutions, with increasing reaction time, initially in the h = 1 solution whereas the schwertmannite
the O shells did not change significantly, but the two Fe shells quantity was negligible (Figure 4a). With time, the Fhy2L
underwent substantial increases in peak amplitude and shifts in component fraction gradually decreased to ∼5% whereas
peak positions (representative spectra in Figure 3 and the schwertmannite developed to 15−20% (Figure 4a). About
complete set in Supporting Information SI-4). As revealed by 80% of the Fe could be represented by the h = 0.5 solution and
further analysis described below, these changes are mainly due this quantity stayed almost constant with time, indicating
to particle phase transformation. formation of schwertmannite from conversion of Fhy2L. For
Time-Dependent Evolution of Fe3+ Species. EXAFS the h = 2 solution, both Fhy2L and schwertmannite accounted
linear combination fitting (LCF) analysis can be used to for ∼30%, respectively, initially (Figure 4b). With time, the
identify and quantify each species in a system containing mixed Fhy2L component gradually disappeared and the schwert-
Fe phases.28 We anticipate that the neutralized solutions mannite fraction increased to ∼70%. The h = 0.5 solution
contain many different Fe species, and complexation by sulfate component decreased from 40% to 30% with time.
8144 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

Figure 5. (a) The fraction of Fe species (Fe_monomer) that reacts with ferron within 1 min, and (b) the change in pH over time for h = 0 to h = 2
solutions. The empty symbols in (a) are replicates measured after filtration to remove precipitates. The solid symbols indicate the fraction of Fe
species measured by ferron assay for unfiltered solutions.

As the two Fe peaks of Fhy2L have lower amplitude and schwertmannite over time. Combining the above two analyses
longer corner-sharing Fe−Fe distance than those of schwert- indicates that the initially formed Fhy2L component, detected
mannite (Figure 3), the relative changes of the two Fe phases as molecular clusters in the h = 1 solution and as particles in the
with time resulted in increasing amplitude of the two Fe peaks h = 2 solution, transformed to schwertmannite over time. UV-
and shortening of the corner-sharing Fe distance (Supporting vis and EXAFS results are also consistent with the SXRD data
Information SI-4). For the same reason, the Fe peaks of the h = that schwertmannite particles were present initially in the h = 2
1 solution were not significant initially, though Fhy2L solution and not in the h = 1 solution, but the SXRD analysis
accounted for 20%. failed to identify the ferrihydrite particles. The XRD peak
Reaction Progress Measured by Ferron Assay and pH sharpening and splitting seen in both solutions were due to the
Changes. In addition to the above EXAFS LCF analysis, the transformation of the Fhy2L particles and clusters, showing
ferron assay was also used to determine reaction progress of poorly defined XRD peaks, to schwertmannite particles
particle formation. Figure 5a shows the time-dependent change showing better defined peaks. The clusters in the h = 1
in the proportion of Fe3+ species that react with the ferron solution were too small to contribute much to the turbidity
probe within one minute, assumed to represent monomeric measured at 600 nm (I ∼ d6/λ4). Therefore, the observed
ferric iron. Initially, almost all Fe was in the Fe_monomer gradual turbidity increase of the h = 1 solution was mainly due
fraction for the h = 1 solution, whereas less than half of the Fe to the formation of schwertmannite particles.
was in this fraction for the h = 2 solution (Figure 5b). In Transformation of Ferrihydrite to Schwertmannite.
addition, the ferron-Fe formation rates of h = 1 and 2 solutions The transformation of the ferrihydrite clusters and particles to
gradually decreased with the increasing neutralization reaction schwertmannite indicates that schwertmannite was the
time (Supporting Information Figure SI-2). In contrast, for the thermodynamically more stable phase in the h = 1 and h = 2
h = 0, 0.5, and 0.6 solutions, ferron−Fe complex formation solutions. This is anticipated, given the presence of sulfate
reaction completed almost instantaneously (data not shown) anions. However, the transformation in the h = 1 solution was
and the rates did not change with the neutralization reaction not complete (5% remained). In addition, ferrihydrite clusters
time, indicating that the solution was dominated by the existed in the h = 0.5 and 0.6 solutions without transforming to
Fe_monomer fraction. It is noteworthy that for the h = 2 schwertmannite within the studied time period. These
solution, a second abrupt decrease in Fe_monomer occurred at observations indicate that ferrihydrite clusters were more stable
∼30 min when the solution became cloudy, possibly due to a than schwertmannite in the presence of sulfate under more
second nucleation event. acidic conditions, although their abundances were low. Given
Fe3+ oxyhydroxide particle formation is a hydrolysis and the large structural difference of the two phases, the
condensation process, releasing H+. Therefore, the pH change transformation likely occurred via dissolution and reprecipita-
over reaction time reflects particle formation progress. As tion.
Figure 5b shows, the pH of the h = 1 and 2 solutions slowly
The EXAFS LCF result shows that 70% of Fe existed as
decreased with the reaction time, suggesting particle formation.


schwertmannite eventually in the h = 2 solution, greater than
the sum (60%) of the initial ferrihydrite and schwertmannite
DISCUSSION fractions. A second nucleation event occurred at 30 min in the
According to the UV-vis results, Fe molecular clusters formed h = 2 solution, as suggested by the sudden drop of the dissolved
initially in the h = 1 solution but transformed to Fe Fe concentration in the ferron assay (Figure 5a). Therefore, the
oxyhydroxide phases over time; in contrast, the initially formed additional 10% schwertmannite resulted from the formation of
Fe species in the h = 2 solution were not molecular clusters but new nuclei rather than growth of the existing particles. This
Fe oxyhydroxide particles, some of which decreased and others indicates that schwertmannite can directly form from dissolved
increased in abundance with time. The QEXAFS LCF analysis Fe species, and that ferrihydrite is not an indispensible
shows that the Fhy2L component was the only phase formed precursor for its formation.
initially in the h = 1 solution, whereas for the h = 2 solution, According to a previous study using an electric field jump
both Fhy2L and schwertmannite phases were present initially; relaxation kinetic technique,29 hydrolysis reactions of mono-
the Fhy2L components in both solutions transformed to meric Fe3+ are extremely fast and should be complete within
8145 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

seconds. The slow pH decreases of the h = 1 and 2 solutions Using this value the cluster concentrations can be estimated as
must be due to other slower chemical reactions. In this case, the 2.4 ± 0.2, 3.1 ± 0.2, and 5.3 ± 0.4 mM in h = 0.5, 0.6 and the
pH decrease can be attributed to the phase transformation, that initial h = 1 solutions, respectively, according to their absolute
is schwertmannite formation from ferrihydrite clusters and absorbance at 850 nm.
particles leads to a net release of protons. Phase transformation- Environmental Implications. For the first time, the
induced pH decreases have been observed previously in identities and abundances of iron molecular clusters and
ferrihydrite transformation to goethite, hematite, and lepidoc- particles formed early in ferric iron solution neutralization have
rocite.30 been determined. The result provides new insights into the
Invalidation of Ferron Assay for Fe Monomers. The precipitation process and demonstrates the necessity of using in
LCF-determined soluble Fe fraction (represented by the h = situ fast diffraction and spectroscopic measurements for analysis
0.5 solution) in the h = 2 solution is basically consistent with of rapid ferric iron oxyhydroxide particle formation processes.
the ferron assay result (Figure 4b). The large discrepancy The concentrations of ferric iron and sulfate used in this study
between the two measurements for the h = 1 solution (Figure are comparable to those in AMD solutions, and the particle
4a) is removed if the ferron assay actually measures both Fhy2L phases identified are commonly found in AMD-impacted
(clusters) and the Fe species approximated by the h = 0.5 environments. In addition, the various neutralization ratios are
solution. The finding that the Fhy2L-like clusters reacted with applicable for natural and engineered AMD neutralization
ferron within one minute invalidates the ferron assay for processes. Therefore, the formation and phase transformation
quantification of Fe monomers. There is less evidence for processes revealed in this study are likely relevant in a variety of
reaction of ferron with the Fhy2L particles in the h = 2 solution AMD environments, though are most directly related to those
within one minute, consistent with the conclusion that the main with minimal ferrous iron and low concentration of heavy
component was Fhy2L particles. The decreased ferron−Fe metal(loid)s and dissolved organic matter. Given that AMD can
complex formation rates with the neutralization reaction time contain a wide variety of toxic elements that are attenuated by
showed that the developing schwertmannite particles were less adsorption onto particle surfaces or through coprecipitation
reactive than ferrihydrite clusters and particles regarding during neutralization, knowledge of early stage particle
ferron−Fe formation. formation and transformation processes contribute to an
The Ferrihydrite Components. The Fhy2L-like cluster in improved understanding of attenuation mechanisms.
the h = 1 solution might resemble the Fe13 motif in the
ferrihydrite structure, that is, [FeO4Fe12(OH)24(H2O)12]7+.31
This molecular cluster was proposed by Bradley and Kydd32

*
ASSOCIATED CONTENT
S Supporting Information
decades ago based on infrared spectroscopy, although few
SXRD, the ferron assay procedures, all time-series QEXAFS
studies followed this suggestion. These authors found this
spectra, QEXAFS shell-by-shell fitting results and comparisons
cluster was unstable, but could be stabilized by sulfate binding.
of QEXAFS spectra with their linear combination fits. This
The structure of the aluminum Keggin cluster (Al13),
material is available free of charge via the Internet at http://
[AlO4Al12(OH)24(H2O)12]7+, was determined after crystalliza-
pubs.acs.org.
tion of the cluster from a solution as a sulfate salt.33 It is not
surprising that sulfate binding could stabilize similar Fe13
molecular clusters in solution, possibly by preventing their
growth into larger ferrihydrite particles and their dissolution by
■ AUTHOR INFORMATION
Corresponding Author
acid as well. The sulfate binding becomes weaker at higher *Phone: (510) 643-9120; fax: (510) 486-5686; e-mail: mzhu@
pH.34 Therefore, when adding more base for the h = 2 case, lbl.gov.
OH− could outcompete sulfate to bind with Fe3+ facilitating the
Notes
growth of ferrihydrite clusters to large particles. This is
The authors declare no competing financial interest.


consistent with the fact that ferrihydrite can be synthesized
by rapidly increasing Fe3+ sulfate solution pH to pH 7. On the
other hand, the weak sulfate binding at h = 2 was not able to ACKNOWLEDGMENTS
protect ferrihydrite from reaction, and accordingly ferrihydrite This work was supported by the Director, Office of Science,
converted completely as the pH was still very acidic (pH < Office of Basic Energy Sciences, Chemical Sciences, Geo-
2.65). sciences, and Biosciences Division, of the U.S. Department of
The Fhy2L-like cluster was also present in the h = 0.5 and 0.6 Energy under Contract No. DE-AC02-05CH11231. The
solutions, according to UV-vis spectroscopy. Since the h = 0.5 authors are grateful to beamline scientists Dr. Syed Khalid at
solution was one of the three components used in the LCF beamline X18B at NSLS, Brookhaven National Laboratory
analyses of the h = 0.6, 1, and 2 solutions, the abundances of (BNL), Dr. Karena Chapman, Dr. Olaf Borkiewicz and Dr.
the Fhy2L-like clusters in these solutions were underestimated. Peter Chupas at beamline 11-ID-B at APS, Argonne National
This means that the clusters were also present in the h = 2 Laboratory (ANL), and Dr. John Bargar at beamline 4-1 at
solution, consistent with the UV-vis spectrum of the filtered Stanford Synchrotron Radiation Lightsource (SSRL) for
solution, although the majority of the ferrihydrite units were providing technical help on data collection. Use of the National
particles. Synchrotron Light Source, Brookhaven National Laboratory
Abundance of the Clusters. Assuming the cluster was supported by the U.S. DOE Office of Science, Office of
contains 13 Fe atoms, based on the LCF-determined fractions Basic Energy Sciences, under Contract No. DE-AC02-
and UV-vis absorbance differences at 850 nm, we estimated its 98CH10886. Use of APS was supported by the U.S. DOE
absorption coefficient (μ) as 44.5 mol−1·cm−1 using the h = 1 Office of Science under Contract No. DE-AC02-06CH11357.
and h = 0.5 results and 52.0 mol−1·cm−1 using the h = 0.6 and h Portions of this research were carried out at the Stanford
= 0.5 results, with an average value of 48.3 ± 3.7 mol−1·cm−1. Synchrotron Radiation Laboratory, a national user facility
8146 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147
Environmental Science & Technology Article

operated by Stanford University on behalf of the U.S. QEXAFS and UV/Vis simultaneous monitoring of the TiO2-
Department of Energy, Office of Basic Energy Sciences. nanoparticles formation by hydrolytic sol-gel route. J. Phys. Chem. C


2010, 114 (14), 6228−6236.
REFERENCES (20) Lopes, L.; de Laat, J.; Legube, B. Charge transfer of iron(III)
monomeric and oligomeric aqua hydroxo complexes: Semiempirical
(1) Evangelou, V. P.; Zhang, Y. L. A reviewPyrite oxidation investigation into photoactivity. Inorg. Chem. 2002, 41 (9), 2505−
mechanisms and acid-mine drainage prevention. Crit. Rev. Environ. Sci. 2517.
Technol. 1995, 25 (2), 141−199. (21) Rossman, G. R. Spectroscopic and magnetic studies of ferric
(2) Burgos, W. D.; Borch, T.; Troyer, L. D.; Luan, F.; Larson, L. N.; iron hydroxy sulfatesIntensification of color in ferric iron clusters
Brown, J. F.; Lambson, J.; Shimizu, M. Schwertmannite and Fe oxides bridged by a single hydroxide ion. Am. Mineral. 1975, 60 (7−8), 698−
formed by biological low-pH Fe(II) oxidation versus abiotic 704.
neutralization: Impact on trace metal sequestration. Geochim. (22) Smolakova, L.; Grygar, T.; Capek, L.; Schneeweiss, O.; Zboril,
Cosmochim. Acta 2011, 76, 29−44. R. Speciation of Fe in Fe-modified zeolite catalysts. J. Electroanal.
(3) Johnson, D. B.; Hallberg, K. B. Acid mine drainage remediation Chem. 2010, 647 (1), 8−19.
options: A review. Sci. Total Environ. 2005, 338 (1−2), 3−14. (23) Scheinost, A. C.; Chavernas, A.; Barron, V.; Torrent, J. Use and
(4) Waychunas, G. A.; Kim, C. S.; Banfield, J. F. Nanoparticulate iron limitations of second-derivative diffuse reflectance spectroscopy in the
oxide minerals in soils and sediments: Unique properties and visible to near-infrared range to identify and quantify Fe oxide minerals
contaminant scavenging mechanisms. J. Nanopart. Res. 2005, 7, in soils. Clay Clay Miner. 1998, 46 (5), 528−536.
409−433. (24) Webb, S. M.; Tebo, B. M.; Bargar, J. R. Structural
(5) Chan, C. S.; De Stasio, G.; Welch, S. A.; Girasole, M.; Frazer, B. characterization of biogenic Mn oxides produced in seawater by the
H.; Nesterova, M. V.; Fakra, S.; Banfield, J. F. Microbial marine Bacillus sp. strain SG-1. Am. Mineral. 2005, 90, 1342−1357.
polysaccharides template assembly of nanocrystal fibers. Science (25) Majzlan, J.; Myneni, S. C. B. Speciation of iron and sulfate in
2004, 303 (5664), 1656−1658. acid waters: Aqueous clusters to mineral precipitates. Environ. Sci.
(6) Flynn, C. M. Hydrolysis of inorganic iron(III) salts. Chem. Rev. Technol. 2004, 39 (1), 188−194.
1984, 84 (1), 31−41. (26) Dousma, J.; den Ottelander, D.; de Bruyn, P. L. The influence of
(7) Banfield, J. F.; Welch, S. A.; Zhang, H.; Ebert, T. T.; Penn, R. L. sulfate ions on the formation of iron(III) oxides. J. Inorg. Nucl. Chem.
Aggregation-based crystal growth and microstructure development in 1979, 41 (11), 1565−1568.
natural iron oxyhydroxide biomineralization products. Science 2000, (27) Sakane, H.; Munoz-Pαez, A.; Diaz-Moreno, S.; Martinez, J. M.;
289 (5480), 751−754. Pappalardo, R. R.; Sanchez Marcos, E. Second hydration shell single
(8) Jolivet, J.-P.; Henry, M.; Livage, J., Metal Oxide Chemistry and scattering versus first hydration shell multiple scattering in M(H2O)63+
Synthesis: From Solution to Solid State; John Wiley & Sons Inc: New EXAFS spectra. J. Am. Chem. Soc. 1998, 120 (40), 10397−10401.
York, 2000. (28) Hansel, C. M.; Benner, S. G.; Fendorf, S. Competing Fe(II)-
(9) Rose, J.; Manceau, A.; Bottero, J.-Y.; Masion, A.; Garcia, F. induced mineralization pathways of ferrihydrite. Environ. Sci. Technol.
Nucleation and growth mechanisms of Fe oxyhydroxide in the 2005, 39 (18), 7147−7153.
presence of PO4 ions. 1. Fe K-Edge EXAFS study. Langmuir 1996, 12 (29) Hemmes, P.; Rich, L. D.; Cole, D. L.; Eyring, E. M. Kinetics of
(26), 6701−6707. hydrolysis of ferric ion in dilute aqueous solution. J. Phys. Chem. 1971,
(10) Masion, A.; Doelsch, E.; Rose, J.; Moustier, S.; Bottero, J. Y.; 75 (7), 929−932.
Bertsch, P. M. Speciation and crystal Chemistry of iron(III) chloride (30) Liu, H.; Li, P.; Zhu, M.; Wei, Y.; Sun, Y. Fe(II)-induced
hydrolyzed in the presence of SiO4 ligands. 3. Semilocal scale structure transformation from ferrihydrite to lepidocrocite and goethite. J. Solid
of the aggregates. Langmuir 2001, 17 (16), 4753−4757. State Chem. 2007, 180 (7), 2121−2128.
(11) Cornell, R. M.; Schwertmann, U., The Iron Oxides: Structure, (31) Michel, F. M.; Ehm, L.; Antao, S. M.; Lee, P. L.; Chupas, P. J.;
Properties, Reactions, Occurrences and Uses, 2nd ed.; Wiley-VCH: New Liu, G.; Strongin, D. R.; Schoonen, M. A. A.; Phillips, B. L.; Parise, J. B.
York, 2003; p 703. The structure of ferrihydrite, a nanocrystalline material. Science 2007,
(12) Combes, J. M.; Manceau, A.; Calas, G.; Bottero, J. Y. Formation 316, 1726−1728.
of ferric oxides from aqueous solutions: A polyhedral approach by X- (32) Bradley, S. M.; Kydd, R. A. Comparison of the species formed
ray absorption spectroscdpy: I. Hydrolysis and formation of ferric gels. upon base hydrolyzes of gallium(III) and iron(III) aqueous solutions -
Geochim. Cosmochim. Acta 1989, 53 (3), 583−594. the possibility of existence of an [FeO4Fe12(OH)24(H2O)12]7+
(13) Bottero, J. Y.; Tchoubar, D.; Arnaud, M.; Quienne, P. Partial polyoxocation. J. Chem. Soc., Dalton 1993, 15, 2407−2413.
hydrolysis of ferric nitrate salt. Structural investigation by dynamic (33) Johansson, G.; Lundgren, G.; Sillén, L. G.; Söderquist, R. On the
light scattering and small-angle x-ray scattering. Langmuir 1991, 7 (7), crystal structure of a basic aluminium sulfate and the corresponding
1365−1369. selenate. Acta Chem. Scand. 1960, 14, 771.
(14) Schwertmann, U.; Cornell, R. M., Iron Oxides in the Laboratory: (34) Fukushi, K.; Sverjensky, D. A. A surface complexation model for
Preparation and Characterization, 2nd ed.; Wiley-VCH: New York, sulfate and selenate on iron oxides consistent with spectroscopic and
2000. theoretical molecular evidence. Geochim. Cosmochim. Acta 2007, 71
(15) Khalid, S.; Caliebe, W.; Siddons, P.; So, I.; Clay, B.; Lenhard, T.; (1), 1−24.
Hanson, J.; Wang, Q.; Frenkel, A. I.; Marinkovic, N.; Hould, N.;
Ginder-Vogel, M.; Landrot, G. L.; Sparks, D. L.; Ganjoo, A., Quick
extended x-ray absorption fine structure instrument with millisecond
time scale, optimized for in situ applications. Rev. Sci. Instrum. 2010,
81, (1).
(16) Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS:
Data analysis for X-ray absorption spectroscopy using IFEFFIT. J
Synchrotron Radiat. 2005, 12, 537−541.
(17) Webb, S. SIXPack: A graphical user interface for XAS analysis
using IFEFFIT. Phys. Scr. 2005, T115, 1011−1014.
(18) Jiang, J. Q.; Graham, N. J. D. Observations of the comparative
hydrolysis/precipitation behaviour of polyferric sulphate and ferric
sulphate. Water Res. 1998, 32 (3), 930−935.
(19) Stotzel, J.; Lutzenkirchen-Hecht, D.; Frahm, R.; Santilli, C. V.;
Pulcinelli, S. H.; Kaminski, R.; Fonda, E.; Villain, F.; Briois, V.

8147 dx.doi.org/10.1021/es301268g | Environ. Sci. Technol. 2012, 46, 8140−8147

You might also like