Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Chemical Geology 487 (2018) 54–62

Contents lists available at ScienceDirect

Chemical Geology
journal homepage: www.elsevier.com/locate/chemgeo

Dissolution and phase transformation processes of hausmannite in acidic T


aqueous systems under anoxic conditions

Yao Luoa, Wenfeng Tana, Steven L. Suibb, Guohong Qiua, , Fan Liua
a
Key Laboratory of Arable Land Conservation (Middle and Lower Reaches of Yangtse River), Ministry of Agriculture, Hubei Key Laboratory of Soil Environment and
Pollution Remediation, College of Resources and Environment, Huazhong Agricultural University, Wuhan 430070, Hubei Province, China
b
Department of Chemistry, University of Connecticut, 55 North Eagleville Road, Storrs, CT 06269-3060, United States

A R T I C LE I N FO A B S T R A C T

Editor: K. Johannesson Hausmannite is the most widely distributed spinel-structured manganese oxide in soils and sediments. The
Keywords: transformation of this metastable manganese oxide to Mn(IV) oxides with higher adsorption capacity has at-
Hausmannite tracted much research interest, while the transformation mechanisms and influencing factors still remain largely
Disproportionation unknown, especially under acidic condition. In this work, the transformation processes of hausmannite at dif-
Transformation ferent pH values and the influence of cations were studied. Results indicated that hausmannite was transformed
Birnessite into manganite at pH 5.0–9.0. The dissolution of hausmannite was initiated and promoted by protons (≤ 7.0),
Nsutite and the decrease of pH accelerated its conversion to Mn(IV) oxides. The tunnel-structured Mn(IV) oxide was
Cryptomelane
generated via two steps during the dissolution process of hausmannite at pH ≤ 3.0. Hausmannite was dis-
proportionated to δ-MnO2 at first, which was then transformed to nsutite in the presence of Na+ and H+ through
the transfer of electrons from adsorbed Mn(II) to structural Mn(IV). The disproportionation of hausmannite to δ-
MnO2 was not affected by other cations, while the presence of K+ promoted the further transformation of δ-
MnO2 to cryptomelane. The structural rearrangement process of δ-MnO2 was the rate-determining step for the
formation of final products. This work expands the understanding of the formation, transformation and geo-
chemical processes of manganese oxides in supergene environments.

1. Introduction topics of research in soil science, environmental chemistry and miner-


alogy.
Manganese oxides are widely distributed in soils and sediments. Hausmannite, the fifth most common manganese oxide minerals in
Their specific physicochemical properties influence the migration, soil profiles (McKenzie, 1972), is a mixed-valent spinel manganese
transformation, and fate of organic pollutants and toxic metalloids/ oxide with Mn(III) in the distorted octahedral coordination and Mn(II)
metals such as As(III,V), Cr(III,VI), and Se(IV,VI) (Katsoyiannis et al., in the tetrahedral coordination (Giovanoli et al., 1976; Jarosch, 1987).
2004; Saputra et al., 2013; Scott and Morgan, 1996; Weaver and Metastable hausmannite can be generated from the abiotic oxidation of
Hochella, 2003). The interaction processes are responsible for the bio- Mn(II) or reduction of manganese oxides (Murray et al., 1985;
geochemical cycling of corresponding active elements. Crystal structure Lefkowitz et al., 2013). In oxygen-containing natural waters, Mn(II) is
plays an important role in the adsorption and oxidation reactivity of easily oxidized to Mn3O4 and β-MnOOH (Murray et al., 1985). Reaction
manganese oxides (Landrot et al., 2012; Luo et al., 2017; Post, 1999; with Mn(II) causes the reduction of birnessite to β-MnOOH, and the
Qiu et al., 2011; Weaver and Hochella, 2003). Compared with todor- subsequent transformation into hausmannite would occur at pH 8.0–8.5
okite, manganite and cryptomelane, layered birnessite exhibits the (Lefkowitz et al., 2013). Thus, hausmannite is easily generated and is
highest reactivity for sulfide oxidation (Luo et al., 2017; Qiu et al., ubiquitous in supergene environments.
2011). Although δ-MnO2 and acid birnessite both show hexagonal layer Due to the presence of Mn(III) in the structure, hausmannite can act
symmetry, the Cr(III) oxidation capacity of acid birnessite is higher as an oxidizing and reducing agent (Madison et al., 2013). The reaction
than that of δ-MnO2 (Landrot et al., 2012). The reactions between seven of hausmannite with As, Cr, and Pu can affect their mobility and toxi-
manganese oxides and Cr(III) indicated that hausmannite and birnessite city (Cantu et al., 2014; Feng et al., 2006; Garcia et al., 2014;
have the highest Cr(III) oxidation ability (Weaver and Hochella, 2003). Shaughnessy et al., 2003; Weaver and Hochella, 2003). Hausmannite,
Hence, the formation and conversion of manganese oxides are hot which comprises both Mn(II) and Mn(III), has a higher Cr(III) oxidation


Corresponding author.
E-mail address: qiugh@mail.hzau.edu.cn (G. Qiu).

https://doi.org/10.1016/j.chemgeo.2018.04.016
Received 7 December 2017; Received in revised form 5 April 2018; Accepted 11 April 2018
Available online 15 April 2018
0009-2541/ © 2018 Elsevier B.V. All rights reserved.
Y. Luo et al. Chemical Geology 487 (2018) 54–62

reactivity than tunnel-structured manganese oxides including crypto- understand the mineralogy in the entire reaction to identify the possible
melane and pyrolusite (Weaver and Hochella, 2003). The highest Cr(III) intermediates and clarify the growth of tunnel-structured Mn(IV)
and Cr(VI) binding capacities of hausmannite were achieved at pH 4.0 oxides.
and 3.0, respectively (Cantu et al., 2014). Thermodynamic analysis
indicated an adsorption process, but activation energy analysis in- 2. Materials and methods
dicated a chemisorption process for Cr(III,VI) (Cantu et al., 2014),
however, the final products after an hour of reaction was not identified 2.1. Synthesis of manganese oxides
for further understanding the reaction mechanism. When the pH was
controlled between 2.0 and 6.0, the maximum binding capacities of Hausmannite was synthesized following the modified procedure of
hausmannite for As(III) and As(V) were respectively 13.5 and Mckenzie (Feng et al., 2007; McKenzie, 1971). The Mn(OH)2 precipitate
7.5 mg g−1 at pH 3.0 (Garcia et al., 2014). Hausmannite can work as an from MnSO4·H2O and NaOH was oxidized by oxygen in air to obtain the
excellent adsorbent for heavy metals, especially at low pH values, but final product. The XRD pattern and Fourier transform infrared spec-
the adsorption mechanisms are not fully understood. The adsorption troscopy (FTIR) spectrum of synthesized product indicated that pure-
process may be related to the pH-dependent changes in the structure phase hausmannite was generated (Fig. S1). The Brunauer-Emmer-
and transformation compositions of metastable hausmannite, which Teller surface area was measured to be 22 m2 g−1. The preparation of
have been largely ignored in previous studies. references, including cryptomelane (α-MnO2), nsutite (γ-MnO2), feit-
Hausmannite is considered as a vital intermediate in the oxidation knechtite (β-MnOOH), manganite (γ-MnOOH), and δ-MnO2, is de-
process of Mn(II) to Mn(IV) oxides. Mn(III) in the structure of haus- scribed in Supporting Information, and the results of XRD and Mn K-
mannite can be disproportionated to Mn(IV) as indicated by a ther- edge extended X-ray absorption fine structure (EXAFS) are shown in
modynamic model (Hem and Lind, 1983). A two-step cycle for a final Fig. S2.
product of MnO2 includes (Hem, 1978): oxidation of Mn(II) to Mn3O4
(Eq. (1)), and disproportionation of Mn3O4 to MnO2 (Eq. (2)). The total 2.2. Mineral dissolution experiments
reaction may be expressed as Eq. (3). These equations are listed as
follows. Dissolution experiments of hausmannite were performed at a pH
range of 3.0–9.0. The whole experiments were conducted under anoxic
6 Mn(II) + O2 + 6 H2 O → 2 Mn3O4 + 12 H+ (1)
conditions. Distilled deionized water (DDW, 18 MΩ cm resistance) was
boiled for 20 min and cooled to room temperature by bubbling high-
Mn3O4 + 4H+ → MnO2 + 2 Mn(II) + 2 H2 O (2)
purity nitrogen gas (99.999%, Wuhan Iron and Steel (Group) Corp.,
China) before use. About 0.50 g of hausmannite (4.4 mmol L−1) was
2 Mn(II) + O2 + 2 H2 O → 2 MnO2 + 4 H+ (3)
added to 500 mL 0.1 mol L−1 Na2SO4 solution in a five-neck bottle with
The incongruent dissolution processes of hausmannite induced by one neck for pH monitoring, one capped with a rubber stopper
hydrolysis and protonation have been extensively studied under oxic equipped with a port for pH adjustment and sampling, one for bubbling
conditions. The dissolution of hausmannite occurs first, and 7-Å phase of nitrogen and the remaining two were sealed. The reaction solutions
birnessite is then precipitated in high-concentration NaOH and KOH were stirred vigorously with a magnetic stir bar throughout the ex-
solution, while dilute alkaline solution facilitates the transformation of periments to ensure thorough mixing. The pH of the reaction systems
hausmannite to manganite (Cornell and Giovanoli, 1988). In acidic was continuously adjusted manually in the first 12 h using NaOH (0.01
solution, conversion of hausmannite to manganite is also dominant at and 0.1 mol L−1) and H2SO4 solution (0.1 and 1.0 mol L−1), and was
pH ≥ 4.0, and Mn(IV) oxide precipitates are the major secondary mi- then adjusted three times a day. After the prescribed time, 3.0 mL
neral phases at lower pH (Giovanoli et al., 1976; Hem and Lind, 1983). suspension was collected from the port of the bottle by a syringe and
Bricker found that γ-MnO2 is produced in intense acidification and δ- filtered through a 0.22 μm microporous membrane filter. The reaction
MnO2 is produced in mild acidification (Bricker, 1965), but the trans- system was continuously purged with high-purity nitrogen gas in the
formation mechanism from hausmannite to δ- and γ-MnO2 has not been first 12 h, and the nitrogen pressure inside the reactor was maintained
clarified yet. In previous studies, the above reaction processes were to be slightly higher than the atmospheric pressure to prevent air in-
mainly monitored by conventional characterization methods such as gress. After 12 h of reaction, the reactor was transferred to an anaerobic
Mn average oxidation state (Mn AOS), X-ray diffraction (XRD), and glove box to maintain anoxic conditions. Wet samples were identified
transmission electron microscopy (TEM) for the long-term reaction by XRD soon after the reaction. The samples in membrane filter were
products. Besides the limitation of technology, the features of manga- washed with deoxygenated DDW for three times, dried in a vacuum
nese oxides, including fine particles, poor crystallinity, and indis- oven at 40 °C for 12 h, and stored in a vacuum bag for further analyses
tinguishable crystal phases, make it extremely difficult to identify the including FTIR, TEM, X-ray photoelectron spectroscopy (XPS), and X-
reaction intermediates, resulting in a poor understanding of the trans- ray absorption spectra (XAS).
formation mechanisms. Synchrotron radiation XRD (SR-XRD) and Mn To examine the influence of cations on the hausmannite transfor-
K-edge X-ray absorption near edge structure (XANES) spectroscopy mation process under acidic conditions, parallel experiments were
were also used to analyze the dissolution products of hausmannite conducted in 0.1 mol L−1 K2SO4 solution or without background cation
(Peña et al., 2007). Manganite, which was observed by XRD in aged under the same experimental conditions.
hausmannite products for several weeks or a year at pH ≥ 6.0, was
identified by SR-XRD but not by XANES in the 1-week dissolution 2.3. Characterization of intermediates
product at pH 6.0 (Peña et al., 2007). The low content of transformation
products also leads to the poor understanding of reaction process. Wet samples were directly pasted onto the glass slides for XRD
Collectively, hausmannite can be transformed to manganese (hydr) (Bruker D8 ADVANCE) data collection using Cu Kα (λ = 0.15418 nm)
oxides under a wide range of environmental conditions. Although the and an operation voltage of 40 kV and current of 40 mA at a scanning
final products have been well analyzed, the detailed transformation rate of 2° min−1. Fourier transform infrared spectroscopy (FTIR, Bruker
mechanism, especially the short-term intermediates and crystal VERTEX 70) was conducted with a Bruker Equinox 55 model spectro-
growing process, still remains unknown. photometer by making pellets with KBr powder. The Mn AOS of
In the present work, we examined the dissolution and phase trans- hausmannite was measured by the oxalic acid method (Kijima et al.,
formation processes of hausmannite with various pH values and aqu- 2001). About 0.1 g of sample was dissolved into 10 mL H2SO4
eous cations including H+, Na+, and K+. Our objectives are to (0.5 mol L−1) and an excess of 10 mL H2C2O4 (0.5 mol L−1) to reduce

55
Y. Luo et al. Chemical Geology 487 (2018) 54–62

all the Mn(III)/Mn(IV) to Mn(II). The excess H2C2O4 in the solution was
determined by titration with a standard solution of KMnO4
(0.02 mol L−1) at 60 °C. Manganese content in the product was de-
termined by atomic absorption spectroscopy (AAS, Varian AAS240FS).
The concentration of Mn(II) in the filtrate was also determined by AAS.
Transmission electron microscopy (TEM, FEI Talos F200C) were used to
characterize the micromorphology of hausmannite and the transfor-
mation intermediates.
The relative contents of Mn(II), Mn(III), and Mn(IV) in the surface of
samples were analyzed by XPS (VG Multilb2000) with Al Kα at
1486.71 eV. The Mn AOS and mineral compositions of reaction pro-
ducts were further characterized by Mn K-edge XAS at the 1W1B
beamline of the Beijing Synchrotron Radiation Facility (BSRF), China.
Mn K-edge XAS data were collected in transmission mode over the
energy range of 6340–7327 eV. The Mn AOS of hausmannite and
transformation products was also analyzed by XANES spectra using
Combo method in the 6521–6653 eV interval (Manceau et al., 2012),
and the fractional contributions of the mineral compositions in the
transformation intermediates were quantitatively calculated by linear Fig. 1. Concentration of released Mn(II) from the transformation of hausman-
combination fitting of the Mn K-edge EXAFS data. Wet solid inter- nite suspension at pH 3.0–9.0 in 0.1 mol L−1 Na2SO4.
mediates were in situ characterized by attenuated total reflectance
Fourier transform infrared (ATR-FTIR) spectroscopy. The detailed de-
decreasing pH, which agrees well with the Mn(II) release tendency
scriptions for XAS and ATR-FTIR analysis are presented in the Sup-
(Fig. 1). When pH was decreased to 3.0, all the diffraction peaks for the
porting Information.
product were indexed to γ-MnO2 (JCPDS card no. 14–0644), indicating
the formation of pure-phase γ-MnO2. Diffraction peaks at 2θ values of
3. Results
22°, 34.5°, 37.1°, 38.8°, 42.6°, and 56.1° correspond to the (120), (031),
(131), (230), (300), and (160) crystal planes of γ-MnO2 (Fig. 2a).
3.1. Dissolution reaction at various pHs
Fig. 2b shows the FTIR spectra of 15-day transformation products
after washing and drying. The absorption bands at 624, 523, and
The transformation reaction of hausmannite (1.0 g L−1) was con-
426 cm−1 are assigned to the stretching and bending vibrations of
ducted at pH 3.0–9.0 in an anoxic environment. After 15 days of reac-
MneO in hausmannite (Kirillov et al., 2009). The bands at 713, 576,
tion, the suspension retained the color of cinnamon brown just like
and 523 cm−1 are attributed to the MneO stretching vibration of MnO6
pristine hausmannite particles when the pH was between 5.0 and 9.0.
octahedra in γ-MnO2 (Potter and Rossman, 1979). The bands at 1094
With the decrease of pH to 3.0, shortly after the beginning of dissolu-
and 1048 cm−1 correspond to the characteristic adsorption peaks of γ-
tion experiments, visible color changes were observed in the solid
MnOOH and β-MnOOH, respectively (Elzinga, 2011; Kirillov et al.,
phase. The solid turned nearly black, indicating the surface oxidation or
2009). γ-MnOOH and β-MnOOH were formed possibly due to the oxi-
transformation of Mn3O4 to a secondary mineral phase under oxic
dation of the sample by air during the washing and drying process (Qiu
conditions as previously reported (Bricker, 1965). In the anoxic reac-
et al., 2011). The band at 1156 cm−1 is assigned to the OeH bending
tion systems, surface oxidation of Mn3O4 did not occur, suggesting the
vibration of γ-MnOOH (Kirillov et al., 2009). Only γ-MnOOH was ob-
formation of a new phase. These results are consistent with the previous
tained during hausmannite hydrolysis at pH 7.0–9.0 (Giovanoli et al.,
report, which showed apparent dissolution of hausmannite at pH ≤ 4.0
1976). The transformation mechanism exhibited significant differences
(Hem and Lind, 1983). However, under oxic conditions, remarkable
at various pHs as indicated by the transformation products.
color changes were observed in Mn3O4 dissolution experiments which
The pH of the experiments was maintained by the addition of H2SO4
used 50 mmol L−1 formic acid to buffer the pH at 5.0 (Peña et al.,
under acidic conditions, suggesting the consumption of protons. The
2007). The different results in this work indicated that oxygen or buffer
dissolution and disproportionation of hausmannite may proceed as Eq.
solution affects the dissolution process of Mn3O4.
(2) at pH ≤ 3.0 in anoxic environment. γ-MnO2 is usually formed as the
The dissolution rate of hausmmanite was further evaluated by
final product in the intense acidification of hausmannite, and will be
comparing the concentration of dissolved Mn(II) at different pHs. The
further transformed to more stable tunnel-structured β-MnO2
concentration of Mn(II) increased and then reached a plateau after
(Giovanoli et al., 1976). After multicycles of the charge-discharge tests,
1 day of reaction (Fig. 1). With the decrease of pH, the release rate of
hausmannite would be electrochemically oxidized to birnessite (Dubal
Mn(II) from the structure of hausmmanite increased, suggesting that
et al., 2010). Our results revealed that hausmannite was dis-
low pH accelerates the dissolution process of hausmmanite. After 2 h of
proportionated in acidic condition and finally transformed to γ-MnO2
reaction, the release ratio of Mn(II) reached about 1%, 8%, and 53%
after equilibration in the Mn(II)-containing solution in the experimental
when pH was respectively controlled at 7.0, 5.0, and 3.0. At pH 3.0, the
period.
concentration of dissolved Mn(II) showed no significant change after
2 h of reaction. The above results confirmed that protons induce and
promote the dissolution of hausmannite, which is consistent with pre- 3.2. Disproportionation processes
vious reports (Peña et al., 2007).
The mineral composition of wet solid products at different pHs after The fast phase change and disproportionation process were con-
15 days of reaction was characterized by XRD (Fig. 2a). When pH was firmed by the changes in the color and Mn(II) concentration, and the
controlled at 5.0–9.0, the diffraction peaks of hausmannite could al- long-term transformation products were fully characterized. However,
ways be observed without the formation of apparent new peaks. The it is still unknown how the disproportionation proceeds and whether
fact that no other phase was observed may be ascribed to the poor other intermediates are preferentially formed directly from the dis-
crystallinity or low abundance of the transformation products. The proportionation. The mineral compositions for different time intervals
diffraction peaks of hausmannite were attenuated with pH decreasing at pH 3.0 were further characterized by XRD to identify the possible
from 9.0 to 5.0, suggesting an increase of dissolution rate with intermediates (Fig. 3). After 30 min of reaction, the diffraction peaks of

56
Y. Luo et al. Chemical Geology 487 (2018) 54–62

Fig. 2. XRD patterns (a) and FTIR spectra (b) of solid products from the transformation of hausmannite suspension for 15 days at pH 3.0–9.0 in 0.1 mol L−1 Na2SO4.

the disproportionation of Mn(III) in hausmannite. The relative content


of Mn(IV) changed from 8% to 67% after 2 h of reaction, and then
approached to 90% after 7 days. The relative content of Mn(II) de-
creased to below 10% within 1 h, indicating that Mn(II) in the structure
was quickly released from the spinel structure under the induction by
protons. The AOS data further confirmed the formation of Mn(IV) in a
short reaction time, which is in agreement with the color change.
The fractional contributions of the compositions in the dissolution
products at pH 3.0 were further quantitatively calculated by linear
combination fitting of the Mn K-edge EXAFS data. The raw χ spectra
could be well fitted by the linear combinations of the χ spectra of
Mn3O4, β-MnOOH, γ-MnOOH, δ-MnO2, and γ-MnO2. The k3-weighted χ
functions of the transformation products exhibited significant changes
with time (Fig. 5a). In consistence with the XRD results, the fitting
results indicated a progressive increase in the amount of γ-MnO2 gen-
erated from the consumption of hausmannite, and hausmannite almost
completely disappeared after 2 h of reaction (Table S2, Fig. 5b). The
relative contents of β-MnOOH, γ-MnOOH, and δ-MnO2 all increased at
first and then decreased with time. The content of δ-MnO2 accumulated
at the expense of hausmannite and showed a decreasing trend in the
Fig. 3. XRD patterns of wet intermediate products from hausmannite suspen- absence of hausmannite after 2 h of reaction. β-MnOOH and γ-MnOOH
sion in 0.1 mol L−1 Na2SO4 at pH 3.0. The intermediate changes are highlighted were formed as the minor products and exhibited a changing trend si-
by dashed lines.
milar to that of δ-MnO2.
Before the determination by XAS, products were exposed to air and
hausmannite were attenuated, and only the diffraction peaks at 2θ could be potentially oxidized during the washing and tablet preparation
values of ~18 and 36° were observed, indicating a remarkable dis- process, which might lead to changes in the mineral composition. β-
solution process. Two weak peaks were formed at ~37 and 66° after 2 h MnOOH and γ-MnOOH can be easily formed from Mn(II) oxidation by
of reaction. The newly formed peaks correspond to the MneO spacing air (Murray et al., 1985). When interacting with Mn(II), birnessite is
in the octahedral arrangement. The initial disproportionation product reduced to β-MnOOH and γ-MnOOH (Lefkowitz et al., 2013). To further
of hausmannite was MnOx octahedral units with a highly disordered investigate the real-time transformation of hausmannite, the inter-
network. γ-MnO2 was formed as the major product and its diffraction mediate products obtained from the reaction were measured by in-situ
peaks increased over time. The release rate of Mn(II) indicated that the ATR-FTIR (Fig. 6). Among the bands between 700 and 1200 cm−1, a
dissolution of Mn3O4 occurred rapidly, but the formation process of γ- broad vibration band at 1056 cm−1 appeared after 30 min of reaction.
MnO2 was quite slow as indicated by XRD results (Figs. 1 and 3). This band corresponds to β-MnOOH, suggesting that γ-MnOOH is not a
Mn AOS is commonly used as an indicator for the oxidation degree direct transformation product in the dissolution process of hausman-
of manganese oxides. The relative contents of Mn(II,III,IV) and Mn AOS nite. The formation of β-MnOOH may be ascribed to the presence of δ-
in the transformation products at pH 3.0 were quantitatively calculated MnO2, and SR-XRD and in-situ XAS can be used to analyze the forma-
by the Combo method (Fig. 4a). The Mn AOS of pristine hausmannite tion of β-MnOOH in future work.
was determined to be 2.85 by the Combo method (Table S1), which is Fig. 7 shows the TEM images of dissolution products of hausmannite
slightly higher than that obtained by titration (2.69). The precision of at pH 3.0. The as-synthesized hausmannite is uniform flaky particles
the Combo method decreases with increasing content of structural Mn with diameters of about 50–100 nm. Porous products with irregular
(III) due to the strong Jahn-Teller distortion effect caused by Mn(III), particle edges were observed after 1 day of reaction. Small particles
which affects the plot of the XANES spectra (Manceau et al., 2012). The were aggregated to form some spindle-shaped structure as the reaction
Mn AOS value increased to 3.63 after 2 h of reaction, and then reached proceeded. The spindle-shaped particles were observed to increase and
3.90 after 7 days of reaction (Fig. 4b, Table S1). The relative content of were the major products after 15 days of reaction. The lattice fringe
Mn(IV) increased along with the decrease of Mn(II) and Mn(III) in the distance was measured to be 0.40 nm, which corresponds to the (120)
transformation products with the extension of reaction time, indicating planes of γ-MnO2 (Fig. 7d).

57
Y. Luo et al. Chemical Geology 487 (2018) 54–62

Fig. 4. Normalized Mn K-edge XANES spectra of pristine hausmannite and transformation products in 0.1 mol L−1 Na2SO4 at pH 3.0 (Open circles and lines re-
spectively present experimental and best-fit lines, and difference plots are revealed at bottom) (a), and the relative content of Mn(II,III,IV) and Mn AOS determined by
Combo method (b).

3.3. Effect of cations on the disproportionation processes

Cation plays an important role in the crystal structure of manganese


oxides. Manganese oxides with either α-, β-, or γ-structure types or the
mixtures of these structure types are produced from the reaction of
acidic MnSO4 with H+, Li+, Na+, K+, or NH4+ solution (Hill et al.,
2003). The background electrolyte may influence the transformation of
hausmannite. Therefore, K+ or H+ was used instead of Na+ to in-
vestigate the dissolution process of hausmannite.
Fig. 8a shows the XRD pattern of the 15-day reaction product at
pH 3.0 with the addition of K+ under anoxic conditions. The diffraction
peaks of the final products were well indexed to cryptomelane (JCPDS
card no. 72-1982). In the absence of background cation, γ-MnO2 was
formed as the major product (a similar figure to that in the presence of
Na+ is not shown here). Cryptomelane is easily generated during the
preparation of manganese oxides in high K+ solution. Well-crystallized
cryptomelane could also be obtained by holding K-birnessite in a mixed
solution of 0.1 mol L−1 HNO3 and 1 mol L−1 KNO3 at 70 °C (Cornell and
Giovanoli, 1988). How metastable hausmannite is transformed to
cryptomelane in K+ solution still remains elusive. The relative contents Fig. 6. ATR-FTIR spectra of hausmannite dissolution for different time intervals
of compositions in the transformation products after 1 day and 15 days in 0.1 mol L−1 Na2SO4 at pH 3.0.
of reaction were quantitatively calculated by linear combination fitting
of the Mn K-edge EXAFS data (Fig. 8b, Table S3). The fitting results
decreased from 34% to 18%, while that of cryptomelane increased from
indicated that the relative content of δ-MnO2 and γ-MnOOH decreased
42% to 76% when the reaction proceeded for 1 day and 15 days, re-
together with that of hausmannite. The relative content of δ-MnO2
spectively. This changing trend of mineral composition is similar to that

Fig. 5. Mn K-edge k3-weighted EXAFS spectra (The lines and dash dots are the experimental and best-fit curves) (a), and the linear combination fitting results of the
k3-weighted χ spectra (b) of the transformation products in 0.1 mol L−1 Na2SO4 at pH 3.0. The changes of spectra are highlighted by shaded areas.

58
Y. Luo et al. Chemical Geology 487 (2018) 54–62

a 0.25 nm b
0.25 nm

2 nm

100 nm 100 nm
c d
0.40 nm

5 nm

100 nm 100 nm
Fig. 7. TEM images of synthesized hausmannite (a) and the transformation intermediate products of hausmannite suspension in 0.1 mol L−1 Na2SO4 at pH 3.0 for
1 day (b), 5 days (c), and 15 days (d).

in the presence of Na+. That is to say, poorly crystalline δ-MnO2 is the 4. Discussion
initial transformation product of hausmannite, which is further slowly
transformed to cryptomelane. 4.1. Disproportionation pathway
Fig. 9 shows the TEM images of intermediates after 1 day and
15 days of reaction in the presence of K+. The products were a mixture In the reaction system, dissolved Mn(II) was detectable at pH ≤ 7.0
of aggregated amorphous and rod-like structures after 1 day of reaction. and obvious transformation products of hausmannite were only ob-
The lattice fringes at 0.72, 0.54, and 0.32 nm correspond to the (110), served at pH 3.0. These results indicated that the dissolution and dis-
(200), and (310) planes of cryptomelane. As shown in Fig. 9b, the rod- proportionation of hausmannite are initiated by protons. The transfor-
like structures were composed of thin sheets for 1-day reaction product. mation rate and products varied at different pHs, and mild acidification
The characteristic morphology of hausmannite disappeared after only induced the transformation of a small portion of hausmannite.
15 days. Rod-like particles and irregular plates were respectively When the pH was kept at 3.0 with different cations in the reaction
formed as the major and minor products. The lattice spacings of the solution systems, the early formed black precipitate was of poor crys-
irregular plates at 0.25 nm may correspond to the (110) planes of δ- tallinity as indicated by XRD analysis. Broad peaks were only observed
MnO2 (Fig. 9d). at 37° and 66°, which were related to the octahedral structure of

Fig. 8. XRD pattern of transformation product after 15 days of reaction (a) and Mn K-edge k3-weighted EXAFS spectra of intermediate products (b) from the
transformation of hausmannite suspension in 0.1 mol L−1 K2SO4 at pH 3.0. The changes of spectra are highlighted by shaded areas.

59
Y. Luo et al. Chemical Geology 487 (2018) 54–62

a b

0.54 nm
0.72 nm

0.32 nm

100 nm 10 nm
c d

100 nm 1 nm
Fig. 9. HRTEM images of the transformation intermediates of hausmannite suspension for 1 day (a, b) and 15 days (c, d) in 0.1 mol L−1 K2SO4 at pH 3.0.

birnessite, indicating a slight ordering of the product. In fact, only the 4.2. Formation mechanism of tunnel-structured Mn oxides
reflection peak of (00l) planes of birnessite was not observed, which
was due to the disordered stacking of octahedral sheets. The release of Under anoxic conditions, the cinnamon brown color of hausmannite
Mn(II) together with the disproportionation of Mn(III) in the structure rapidly turned to black at pH 3.0. In background electrolyte with H+,
of hausmannite led to the formation of poor crystalline δ-MnO2. Na+, or K+, the black particles subsequently aggregated to form an
MnO4−/Mn2+ was used to synthesize α-MnO2, γ-MnO2, β-MnO2, inhomogeneous suspension. These phenomena suggested that the
and birnessite by controlling the reaction conditions such as pH and physiochemical properties of the δ-MnO2 newly formed from the dis-
cations, and poorly ordered MnO2 was formed as the precursor for the solution process underwent changes during the aging process.
well-crystallized manganese oxides (Portehault et al., 2009). During the Layer structures were compressed and simultaneously closed up to
preparation of manganese oxides from different Mn sources, δ-MnO2 is form tunnel structures by hydrothermal treatment of birnessite in 1 M
responsible for the initial formation of tunnel-structured Mn oxides HNO3, and the formation of MnO2 nanorods/nanowires was attributed
(Wang and Li, 2003). Flower-like birnessite was formed from the ro- to a rolling mechanism (Shen et al., 2006). In this work, the rolling of
tating, orienting and attaching of poorly crystalline δ-MnO2 nano- plate was not observed in the transformation of spinel structure to layer
particles, and these nanopetals grew mainly along the a–b planes via the and tunnel structures. Xiao et al. speculated that nanofibrous crypto-
oriented attachment mechanism (Liang et al., 2017). δ-MnO2 is com- melane is developed from an agglomerated mass of nanoparticles (Xiao
monly formed as the initial product in the growth of manganese oxide, et al., 1998). Orthogonal birnessite probably grows through an oriented
and works as an important precursor for birnessite and tunnel-struc- aggregation mechanism from hexagonal birnessite (including δ-MnO2
tured Mn(IV) oxides. Bricker proposed that birnessite can be formed and acid birnessite), and the smaller size of δ-MnO2 may favor the
from the disproportionation of hausmannite via Eq. (2). δ-MnO2 growth of orthogonal birnessite (Landrot et al., 2012; Zhao et al.,
showed an electron micromorpology of poorly organized crystals with a 2016). In this work, the XRD and EXAFS results demonstrated that a
plat-like appearance (Bricker, 1965). Poorly ordered δ-MnO2 was also large amount of δ-MnO2 disordered along the c axis was generated in
observed as a metastable intermediate in this work. the early stage of dissolution. The micromorphology was porous for the
In the disproportionation process of hausmannite, the early pre- initial intermediate (Figs. 7b and 9a), which might be composed of δ-
cipitate was constituted of MnO6 octahedral layers randomly stacked MnO2. The oriented aggregation of δ-MnO2 facilitated the further
with each other, and was considered as δ-MnO2 that was disordered generation of one-dimensional γ-MnO2 and α-MnO2.
along the c axis. The EXAFS fitting data further confirmed that δ-MnO2 After complete disproportionation of hausmannite, the theoretical
was one of the compositions in the product, which increased at first and Mn(II) concentration was calculated to be 8.7 mmol L−1. The con-
then decreased. The above results indicated that hausmannite was centration of dissolved Mn(II) reached about 7.0 mmol L−1 after 2 h of
broken down to tiny thin plate-like δ-MnO2, which was then trans- reaction at pH 3.0 (Fig. 1). Hence, most of the Mn(II) in the structure
formed to tunnel-structured manganese oxides. Although cations were and disproportionation products of Mn(III) was released to the aqueous
needed to compensate for the charge balance of layer-structured bir- solution. Birnessite has a high affinity for Mn(II), which permits the
nessite, δ-MnO2 was always formed in different cation (H+, Na+, and reaction of birnessite with adsorbed Mn(II). MnOOH and Mn3O4 are
K+) systems, indicating that the presence of other cations is not es- formed through the transfer of electrons from adsorbed Mn(II) to
sential to the formation of δ-MnO2. structural Mn(IV), which causes the rearrangement of the crystal

60
Y. Luo et al. Chemical Geology 487 (2018) 54–62

structure of birnessite (Elzinga, 2011). The transformation from hex- rearrangement with the generation of tunnel-structured γ-MnO2. The
agonal birnessite to orthogonal birnessite follows a similar mechanism total transformation process can be summarized as the following
(Zhao et al., 2016). δ-MnO2 owns a larger specific surface than acid equations.
birnessite (Zhao et al., 2016), which favors the surface adsorption of Mn
(II). The structural rearrangement of δ-MnO2 may also contribute to the Mn(II)Mn(III)2 O4 + 4 H+ → δ‐Mn(IV)O2 + 2 Mn(II) + 2 H2 O (4)
formation of γ-MnO2 and α-MnO2 in this work.
In order to investigate the adsorption process of Mn(II), relative δ‐Mn(IV)O2 (layer) + Mn(II) → γ‐Mn(IV)O2 (tunnel) (5)
contents of manganese with different valence states in the surface of
intermediates were investigated by XPS (Figs. 10 and S3). The Mn AOS In particular, inorganic cation templates play critical roles in the
of pristine hausmannite was calculated to be 2.65 by XPS analysis, formation of different tunnel structures, as the tunnels of different Mn
which is close to the value measured by titration. The major peak at (IV) oxide crystals have different sizes and need different amounts of
~641.4 eV shifted to ~642.2 eV after 10 min of reaction, and then cations to be stabilized (Hill et al., 2003). In the rearrangement of solid
shifted towards the low energy value (Fig. 10b). Mn AOS of the inter- phase, γ-MnO2 was formed in the presence of H+ and Na+; however,
mediates increased at first and then decreased. The relative contents of cryptomelane was formed when K+ was added instead. The ionic radius
surface Mn(II), Mn(III), and Mn(IV) were obtained from XPS Mn 2p3/2 of H+ (0.29 Å) and Na+ (0.95 Å) is far smaller than that of K+ (1.33 Å).
spectra fitting of intermediates at different times, and the relative K+ ions can serve as a template for the formation of the [2 × 2] tunnel
content of Mn(II) increased to 31.9% and 51.6% after 10 min and 12 h structure of α-MnO2. In the early stage of self-assembly process of thin
of reaction, respectively (Fig. S3, Table S4). The relative content of Mn plates, K+ ions play an important role in supporting the tunnel of α-
(II) obtained by XPS had a trend opposite to that obtained by XANES, MnO2.
and the value of the former was apparently lower than that of the latter, This study demonstrates that disproportionation of hausmannite is
because XPS methods only detect a surface depth of ~2 nm in minerals accompanied by the release of Mn(II) in the intense acidification, and
and the XAS results are bulk measurements. The Mn(II) in the crystal hausmannite is be transformed into MnO2 (γ-, α-MnO2) with δ-MnO2 as
structure and disproportionation products of Mn(III) was released to an intermediate in the presence of Mn(II). Solution pH determines the
aqueous solution or adsorbed on the surface of the reaction product. transformation rate and products, significantly affecting the reactivity
The release of Mn(II) was initially rapid and the amount of released Mn of manganese oxides with nutrients and trace toxic elements as well as
(II) was much larger than that of adsorbed Mn(II) in the early stage of the geochemical cycling of related elements in the environment. δ-
reaction (within 10 min), leading to the decrease of Mn(II) on the sur- MnO2 can be formed through biological and abiotic pathways, and is
face of the reaction product. Then, more Mn(II) was adsorbed on the widely distributed in soils, suboxic marine, fresh water sediments, and
surface of the products with the increased content of δ-MnO2 due to the riparian zones. High reactivity of δ-MnO2 influences the migration and
high affinity of δ-MnO2 for Mn(II) (within 2 h). After 2 h of reaction, toxicity of organic compounds and heavy metals such as Se(IV), Cr(III),
there was still a small amount of Mn(II) adsorbed by the products as the and As(III). Hausmannite is easily formed in natural environment from
reaction proceeded. the oxidation of Mn(II) and reduction of Mn(IV) oxides and more easily
Birnessite exhibits a high stability in the absence of Mn(II) at found in alkaline medium. Although ocean and most surficial en-
pH ≤ 3.0, and a certain concentration of Mn(II) induces the transfor- vironments have pH above 7.0, the decomposition of organics, acid
mation of birnessite to γ-MnO2 (Tu et al., 1994). Tu et al. speculated mine drainage, and acid rain may cause very low pH values in local
that Mn(IV) in the crystal structure of birnessite works as the electron areas. In recent years, the pH of acid rain is below 4.5 in large areas of
acceptor and is reduced to Mn(II) by adsorbed Mn(II) in an auto- southern China (Du et al., 2012; Zhang et al., 2007), and can be even as
catalysis process (Tu et al., 1994). Some adsorbed Mn(II) was in- low as 2.7 in Pennsylvania and 1.5 in west Virginia (LaBastille, 1981).
corporated into the birnessite structure, causing changes in the length Mn(II) release and structure alteration of hausmannite may occur in
of MneO bonds and the imbalanced charge. The structure of birnessite these environments. The above transformation processes can com-
became unstable and was rearranged to a new stable phase due to the pletely alter the adsorption and redox reactivity of hausmannite. The
incorporation of a large amount of Mn(II) (Feitknecht et al., 1960; Tu disproportionation and dissolution of hausmannite, which may affect
et al., 1994). Hence, the initial transformation product δ-MnO2 ad- the migration and fates of trace elements and pollutants, should be fully
sorbed Mn(II), and then electrons were transferred from surface Mn(II) considered in these spots.
to structural Mn(IV) in δ-MnO2, followed by the crystal structure

Fig. 10. Mn 2p3/2 spectra of pristine hausmannite (The open circles are experimental curves and the lines are the best-fit linear combination) (a) and Mn 2p3/2
spectra of intermediates after transformation in 0.1 mol L−1 Na2SO4 at pH 3.0 for different times (b).

61
Y. Luo et al. Chemical Geology 487 (2018) 54–62

5. Conclusion Hem, J.D., Lind, C.J., 1983. Nonequilibrium models for predicting forms of precipitated
manganese oxides. Geochim. Cosmochim. Acta 47, 2037–2046.
Hill, L.I., Verbaere, A., Guyomard, D., 2003. Nanofibrous α, β, γ and α·γ-manganese di-
The transformation process and intermediate products of haus- oxides prepared by the hydrothermal-electrochemical technique I. Synthesis and
mannite are influenced by pH and cations. Hausmannite is transformed characterization. J. Electrochem. Soc. 150, D135–D148.
into manganite at pH 5.0–9.0. The decrease of pH promotes the trans- Jarosch, D., 1987. Crystal structure refinement and reflectance measurements of haus-
mannite, Mn3O4. Mineral. Petrol. 37, 15–23.
formation of hausmannite to Mn(IV) oxides, and hausmmanite is Katsoyiannis, I.A., Zouboulis, A.I., Jekel, M., 2004. Kinetics of bacterial As(III) oxidation
completely consumed within 2 h of reaction at pH 3.0. The structural and subsequent As(V) removal by sorption onto biogenic manganese oxides during
Mn(II) and disproportionated Mn(III) are released from hausmannite to groundwater treatment. Ind. Eng. Chem. Res. 43, 486–493.
Kijima, N., Yasuda, H., Sato, T., Yoshimura, Y., 2001. Preparation and characterization of
aqueous solution in the form of Mn(II) ions. When the pH is controlled open tunnel oxide α-MnO2 precipitated by ozone oxidation. J. Solid State Chem. 159,
at 3.0, hausmannite is disproportionated to δ-MnO2 at first, which is 94–102.
then transformed to tunnel-structured Mn(IV) oxide through the Kirillov, S.A., Aleksandrova, V.S., Lisnycha, T.V., Dzanashvili, D.I., Khainakov, S.A.,
García, J.R., Visloguzova, N.M., Pendelyuk, O.I., 2009. Oxidation of synthetic haus-
structural rearrangement of δ-MnO2. Nsutite is formed in the presence
mannite (Mn3O4) to manganite (MnOOH). J. Mol. Struct. 928, 89–94.
of H+ and Na+ in the reaction systems, and the presence of K+ ac- LaBastille, A., 1981. Acid rain how great a menace? Natl. Geogr. 160, 652–679.
celerates the transformation of δ-MnO2 to cryptomelane. The structural Landrot, G., Ginder-Vogel, M., Livi, K., Fitts, J.P., Sparks, D.L., 2012. Chromium(III)
rearrangement process of δ-MnO2 is the rate-determining step for the oxidation by three poorly-crystalline manganese(IV) oxides. 1. Chromium(III)-oxi-
dizing capacity. Environ. Sci. Technol. 46, 11594–11600.
transformation of hausmannite to tunnel-structured Mn(IV) oxide. Lefkowitz, J.P., Rouff, A.A., Elzinga, E.J., 2013. Influence of pH on the reductive trans-
formation of birnessite by aqueous Mn(II). Environ. Sci. Technol. 47, 10364–10371.
Acknowledgements Liang, X.R., Zhao, Z.R., Zhu, M.Q., Liu, F., Wang, L.J., Yin, H., Qiu, G.H., Cao, F.F., Liu,
X.Q., Feng, X.H., 2017. Self-assembly of birnessite nanoflowers by staged three-di-
mensional oriented attachment. Environ. Sci.: Nano 4, 1656–1669.
This work was supported by the National Natural Science Luo, Y., Shen, Y.G., Liu, L.H., Hong, J., Qiu, G.H., Tan, W.F., Liu, F., 2017. In situ de-
Foundation of China (Grant Nos. 41571228 and 41425006), the tection of intermediates from the interaction of dissolved sulfide and manganese
oxides with a platinum electrode in aqueous systems. Environ. Chem. 14, 178–187.
National Key Research and Development Program of China (Grant No. Manceau, A., Marcus, M.A., Grangeon, S., 2012. Determination of Mn valence states in
2017YFD0801000), the Fok Ying-Tong Education Foundation (Grant mixed-valent manganates by XANES spectroscopy. Am. Mineral. 97, 816–827.
No. 141024) and the Fundamental Research Funds for the Central Madison, A.S., Tebo, B.M., Mucci, A., Sundby, B., Luther, G.W., 2013. Abundant pore-
water Mn(III) is a major component of the sedimentary redox system. Science 341,
Universities (Program No. 2662015JQ002). S. L. S. thanks the support 875–878.
of the US Department of Energy, Office of Basic Energy Sciences, McKenzie, R.M., 1971. The synthesis of birnessite, cryptomelane, and some other oxides
Division of Chemical, Biological and Geological Sciences under grant and hydroxides of manganese. Mineral. Mag. 38, 493–502.
McKenzie, R.M., 1972. The sorption of some heavy metals by the lower oxides of man-
DE-FG02-86ER13622.A000. We greatly acknowledge Dr. Lirong Zheng
ganese. Geoderma 8, 29–35.
at beamline 1W1B of Beijing Synchrotron Radiation Facility (BSRF) for Murray, J.W., Dillard, J.G., Giovanoli, R., Moers, H., Stumm, W., 1985. Oxidation of Mn
the technical assistance with data collection and analysis. The authors (II): initial mineralogy, oxidation state and ageing. Geochim. Cosmochim. Acta 49,
also thank Dr. Jianbo Cao at the Public Laboratory of Electron 463–470.
Peña, J., Duckworth, O.W., Bargar, J.R., Sposito, G., 2007. Dissolution of hausmannite
Microscope in Huazhong Agricultural University for the help of TEM (Mn3O4) in the presence of the trihydroxamate siderophore desferrioxamine B.
characterization. Geochim. Cosmochim. Acta 71, 5661–5671.
Portehault, D., Cassaignon, S., Baudrin, E., Jolivet, J.P., 2009. Structural and morpho-
logical control of manganese oxide nanoparticles upon soft aqueous precipitation
Appendix A. Supplementary data through MnO4−/Mn2+ reaction. J. Mater. Chem. 19, 2407–2416.
Post, J.E., 1999. Manganese oxide minerals: crystal structures and economic and en-
Supplementary data to this article can be found online at https:// vironmental significance. Proc. Natl. Acad. Sci. 96, 3447–3454.
Potter, R.M., Rossman, G.R., 1979. The tetravalent manganese oxides: identification,
doi.org/10.1016/j.chemgeo.2018.04.016. hydration, and structural relationships by infrared spectroscopy. Am. Mineral. 64,
1199–1218.
References Qiu, G.H., Li, Q., Yu, Y., Feng, X.H., Tan, W.F., Liu, F., 2011. Oxidation behavior and
kinetics of sulfide by synthesized manganese oxide minerals. J. Soils Sediments 11,
1323–1333.
Bricker, O., 1965. Some stability relations in system Mn-O2-H2O at 25 degrees and 1 Saputra, E., Muhammad, S., Sun, H.Q., Ang, H.M., Tadé, M.O., Wang, S.B., 2013.
atmosphere total pressure. Am. Mineral. 50, 1296–1354. Manganese oxides at different oxidation states for heterogeneous activation of per-
Cantu, Y., Remes, A., Reyna, A., Martinez, D., Villarreal, J., Ramos, H., Trevino, S., oxymonosulfate for phenol degradation in aqueous solutions. Appl. Catal. B Environ.
Tamez, C., Martinez, A., Eubanks, T., Parsons, J.G., 2014. Thermodynamics, kinetics, 142–143, 729–735.
and activation energy studies of the sorption of chromium(III) and chromium(VI) to a Scott, M.J., Morgan, J.J., 1996. Reactions at oxide surfaces. 2. Oxidation of Se(IV) by
Mn3O4 nanomaterial. Chem. Eng. J. 254, 374–383. synthetic birnessite. Environ. Sci. Technol. 30, 1990–1996.
Cornell, R.M., Giovanoli, R., 1988. Transformation of hausmannite into birnessite in al- Shaughnessy, D.A., Nitsche, H., Booth, C.H., Shuh, D.K., Waychunas, G.A., Wilson, R.E.,
kaline media. Clay Clay Miner. 36, 249–257. Gill, H., Cantrell, K.J., Serne, R.J., 2003. Molecular interfacial reactions between Pu
Du, Y.J., Jiang, N.J., Shen, S.L., Jin, F., 2012. Experimental investigation of influence of (VI) and manganese oxide minerals manganite and hausmannite. Environ. Sci.
acid rain on leaching and hydraulic characteristics of cement-based solidified/sta- Technol. 37, 3367–3374.
bilized lead contaminated clay. J. Hazard. Mater. 225, 195–201. Shen, X.F., Ding, Y.S., Hanson, J.C., Aindow, M., Suib, S.L., 2006. In situ synthesis of
Dubal, D.P., Dhawale, D.S., Salunkhe, R.R., Lokhande, C.D., 2010. Conversion of inter- mixed-valent manganese oxide nanocrystals: an in situ synchrotron X-ray diffraction
locked cube-like Mn3O4 into nanoflakes of layered birnessite MnO2 during super- study. J. Am. Chem. Soc. 128, 4570–4571.
capacitive studies. J. Alloys Compd. 496, 370–375. Tu, S., Racz, G.J., Goh, T.B., 1994. Transformations of synthetic birnessite as affected by
Elzinga, E.J., 2011. Reductive transformation of birnessite by aqueous Mn(II). Environ. pH and manganese concentration. Clay Clay Miner. 42, 321–330.
Sci. Technol. 45, 6366–6372. Wang, X., Li, Y.D., 2003. Synthesis and formation mechanism of manganese dioxide
Feitknecht, W., Oswald, H.R., Feitknecht-Steinmann, U., 1960. Über die topochemische nanowires/nanorods. Chem. Eur. J. 9, 300–306.
einphasige reduktion von γ-MnO2. Helv. Chim. Acta 43, 1947–1950. Weaver, R.M., Hochella, M.F., 2003. The reactivity of seven Mn-oxides with Cr3+ aq : a
Feng, X.H., Zu, Y.Q., Tan, W.F., Liu, F., 2006. Arsenite oxidation by three types of comparative analysis of a complex, environmentally important redox reaction. Am.
manganese oxides. J. Environ. Sci. 18, 292–298. Mineral. 88, 2016–2027.
Feng, X.H., Zhai, L.M., Tan, W.F., Liu, F., He, J.Z., 2007. Adsorption and redox reactions Xiao, T.D., Strutt, E.R., Benaissa, M., Chen, H., Kear, B.H., 1998. Synthesis of high active-
of heavy metals on synthesized Mn oxide minerals. Environ. Pollut. 147, 366–373. site density nanofibrous MnO2-base materials with enhanced permeabilities.
Garcia, S., Sardar, S., Maldonado, S., Garcia, V., Tamez, C., Parsons, J.G., 2014. Study of Nanostruct. Mater. 10, 1051–1061.
As(III) and As(V) oxoanion adsorption onto single and mixed ferrite and hausmannite Zhang, J.E., Ouyang, Y., Ling, D.J., 2007. Impacts of simulated acid rain on cation
nanomaterials. Microchem. J. 117, 52–60. leaching from the Latosol in south China. Chemosphere 67, 2131–2137.
Giovanoli, R., Feitknecht, W., Maurer, R., Hani, H., 1976. Reaction of Mn3O4 with acids. Zhao, H.Y., Zhu, M.Q., Li, W., Elzinga, E.J., Villalobos, M., Liu, F., Zhang, J., Feng, X.H.,
Chimia 30, 307–309. Sparks, D.L., 2016. Redox reactions between Mn(II) and hexagonal birnessite change
Hem, J.D., 1978. Redox processes at surfaces of manganese oxide and their effects on its layer symmetry. Environ. Sci. Technol. 50, 1750–1758.
aqueous metal ions. Chem. Geol. 21, 199–218.

62

You might also like