Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Environmental Technology & Innovation 35 (2024) 103689

Contents lists available at ScienceDirect

Environmental Technology & Innovation


journal homepage: www.elsevier.com/locate/eti

Electrochemical study characterizations of pyrite weathering in


simulated acidic soil: Iron transformation, sulfur conversion and
environmental implications
Xiaonan Feng a, b, 1, Zhijie Chen c, 1, Shuai Wang a, b, Bing-Jie Ni c, Ling Cen a, b,
Qingyou Liu d, *
a
Key Laboratory of High-temperature and High-pressure Study of the Earth’s Interior, Institute of Geochemistry, Chinese Academy of Sciences,
Guiyang 550081, China
b
University of Chinese Academy of Sciences, Beijing 100039, China
c
Water Research Centre, School of Civil and Environmental Engineering, The University New South Wales, Sydney, NSW 2052, Australia
d
School of Materials and Environmental Engineering, Shenzhen Polytechnic University, Shenzhen 518055, China

A R T I C L E I N F O A B S T R A C T

Keywords: Pyrite in natural environments is susceptible to weathering, which results in acid mine drainage
Pyrite weathering (AMD) and heavy metal pollution-related environmental problems. The increase in mining ac­
Acidic soil tivities has led to a large amount of pyrite entering surroundings, which may aggravate soil
Electrochemical techniques
acidification and heavy metal pollution. In situ electrochemical techniques were used to inves­
Surface analysis
Environmental implications
tigate the weathering behavior of pyrite in simulated acidic soil solutions, which is conducive to
understanding the weathering mechanism and environmental impact. In this study, the polari­
zation curves and electrochemical impedance spectroscopy reveal that a higher soil solution
concentration, temperature and/or acidity accelerate pyrite weathering rate, with an activation
energy of 21.21 kJ⋅mol-1. Further investigations reveal that the pyrite is initially oxidized to
initially oxidized to S0 and released Fe(II), the Fe(II) is easily transformed into Fe(III), and part of
the Fe species ultimately transforms into goethite γ-FeOOH and hematite α-Fe2O3. The S0 ulti­
mately transform into sulfate, accompanied by the release of hydrogen ions (H+), resulting in
serious soil acidification. These results specifically reveal the mechanism of pyrite weathering,
especially the process of iron transformation and sulfur conversion, and they also quantitatively
identify the pyrite weathering rate and the quantities of heavy metal ions released, providing an
experimental basis for the risk assessment of pyrite in acid soils.

1. Introduction

Pyrite (FeS2) is the most widespread metal sulfide material in the crust and is often associated with chalcopyrite, galena, sphalerite,
and precious metals like Au (Gao et al., 2015). Pyrite readily oxidizes and hence the mining process can potentially disrupt ecological
and environmental balances, such as heavy metal contamination, acid mine drainage (AMD), and tailing impoundment among others
(Fan et al., 2017; Ogbughalu et al., 2020). Meanwhile, pyrite oxidation is utilized for desulfurizing coal, mineral flotation, new

* Corresponding author.
E-mail address: liuqingyou@szpu.edu.cn (Q. Liu).
1
Xiaonan Feng and Zhijie Chen contribute equally to this work.

https://doi.org/10.1016/j.eti.2024.103689
Received 28 January 2024; Received in revised form 11 May 2024; Accepted 21 May 2024
Available online 22 May 2024
2352-1864/© 2024 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

materials, soil science, and the biogeochemical cycling of S and Fe (Wu et al., 2020). Thus, it is very important to study the weathering
mechanism of pyrite due to the double-edged sword nature of pyrite.
Acidic soils occupy approximately 30 % or 3950 m ha of the world’s ice-free land area (von Uexküll and Mutert, 1995), and they
contain Na+, Fe3+, Cl-, SO2- -
4 , and HCO3. The presence of Fe
3+
has a strong oxidizing characteristic, promoting pyrite weathering. Cl- has
been proposed as an aggressive anion that destroys passive films via adsorption or diffusion control and promotes pyrite weathering
(Zheng et al., 2019). In addition, HCO-3 is suggested to benefit pyrite oxidation due to its high ionic strength and smaller electric
double-layer resistance (Caldeira et al., 2003).
A large amount of pyrite migrated into acidic soils with mining activities, exacerbating soil acidification and heavy metal pollution.
The above-mentioned ions in acidic soil affect pyrite weathering, and in turn, pyrite weathering may cause further acidification of
acidic soil and heavy metal ions pollution. Up to now, there is still a lack of data on the thermodynamics, kinetics and intermediate
mechanism of pyrite weathering in acid soils. Thus, In situ-electrochemical techniques (open circuit potential (OCP), cyclic voltam­
metry (CV), polarization curves, and electrochemical impedance spectroscopy (EIS)) and surface techniques (scanning electron mi­
croscopy with energy dispersive x-ray spectroscopy (SEM-EDS), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS))
were used to study the weathering behavior of pyrite in simulated acidic soil solution in this work.

2. Materials and methods

2.1. Preparation of pyrite samples and acidic soil solutions

The single pyrite crystals and powdered samples used in this experiment were obtained from a Hg mine in Dongchuan Pb-Zn mine,
Yunnan Province, China. X-ray diffraction measurements revealed that the single samples used were pure and uniform pyrite; electron
microprobe analysis confirmed they contained 52.96 % S and 46.89 % Fe (wt%) (Table S1). First, the single pyrite samples were cut
into 0.5 cm × 0.5 cm × 1.0 cm cuboids. The upper surface was attached to a copper conductor using silver paint, and the rest was sealed
with epoxy resin so that only the bottom surface was exposed as a working electrode. Then, other block pyrite samples (Lanmuchang
mine, Guizhou Province, China, impurity pyriteis shown in Table S2) were crushed, and the pieces were microscopically selected,
ground, and screened to obtain 200 mesh (~74 μm) FeS2 powders.
The simulated acidic soil solutions were prepared according to the soil composition of Yingtan, Jiangxi Province, which is one of
the most representative and typical acidic soil areas in southeastern China. The chemical components of the simulated acidic soil
solutions are shown in Table 1 (Liu et al., 2009). Four chemical concentrations were used to simulate different acidic soil conditions.
Sulfuric acid (H2SO4) and sodium hydroxide (NaOH) solutions were used to adjust the pH to 3.0, 4.5, and 6.0.

2.2. Electrochemical experiment

The electrochemical characterizations were obtained using a traditional three-electrode system, the FeS2 electrode was used as the
working electrode, a platinum electrode was used as the auxiliary electrode, and a saturated calomel electrode (SCE) was used as the
reference electrode. The electrolyte was approximately 25 ml, and a water bath system was used to adjust the experimental tem­
perature. All of the potentials in this study are in reference to the SCE.
Before all electrochemical measurements, OCP was conducted until its value changed by less than 2.0 mV over 5 min. The CV tests
provide information about the various redox reactions that may occur on the interface. First, the potential was scanned from the OCP to
700 mV (vs. SCE) in the positive direction. Then, it was scanned from 700 mV to − 1000 mV in the negative direction. Finally, it was
scanned back to the OCP at a scan rate of 20 mV⋅s-1. The polarization curve can provide information about the weathering rate and the
related kinetic and thermodynamic parameters. It was conducted at 10 mV⋅s-1 from − 250 mV to +250 mV (vs. OCP) in this study. EIS
can be used to obtain information about the electrochemical process at the electrode/electrolyte interface. The EIS was performed at
the OCP with an amplitude of 10 mV in the frequency range of 10-3 to 105 Hz. The ZSimpWin 3.20 and PowerSuit softwares were used
to simulate the impedance and polarization data after the tests, respectively. All of the electrochemical experiments were conducted at
least in triplicate to ensure credibility and reproducibility.

2.3. Pyrite weathering batch soaking experiments and characterization

Eleven polished pyrite block specimens were analyzed with SEM-EDX and Raman spectra. One was a pristine pyrite sample. The
other 9 samples were divided into three groups and were soaked in 10 different solutions for 10 weeks: (1) 25 ◦ C, pH 6.0 with different

Table 1
Chemical composition of simulated acidic soil solutions (g⋅L-1).
Solution Chemical composition (g⋅L-1)

MgSO4⋅7 H2O NaCl Na2SO4 CaCl2 KNO3 NaHCO3

CS, 1 - - - - - -
CS, 2 0.020 0.047 0.014 0.011 0.029 0.015
CS, 3 0.100 0.235 0.070 0.055 0.145 0.075
CS, 4 0.400 0.940 0.280 0.220 0.580 0.300

2
X. Feng et al.
3

Environmental Technology & Innovation 35 (2024) 103689


Fig. 1. OCP for the pyrite electrode in simulated acidic soil solutions at different concentrations (a), different temperatures (b), and different acidities (c); Polarization curves for the pyrite electrode in
simulated acidic soil solutions at different concentrations (d), different temperatures (e), and different acidities (f).
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

concentrations as Cs,1, Cs,2, Cs,3, Cs,4 (Table 1); (2) Cs,3, pH 6.0 at different temperatures of 5, 15, 25, and 35 ◦ C; and (3) 25 ◦ C, Cs,3 with
different pH values of 3.0, 4.5, and 6.0. After weathering for 10 weeks, a JSM-6460LV SEM coupled with an EDX apparatus was
employed to observe the changes in the surface morphologies and relevant elemental contents. Raman analysis (British Renishaw
instrument) was used to obtain Raman data for determining the type of chemical bonds.
0.5 g of pyrite powder and 50 ml of the 10 different solutions mentioned above were added to a 100-ml Erlenmeyer flask. They were
subjected to XPS testing after 10 weeks of soaking using the same method used for the block samples. They were analyzed using XPS
after two weeks of soaking, and the original power samples were analyzed using an ESCALAB 250XI instrument to obtain the XPS
spectra for characterizing elemental chemical states.

3. Results and discussion

3.1. Electrochemical characterizations

3.1.1. OCP study


The OCP mainly reflects the stability of the entire electrochemical system. From Fig. 1, all of the potentials change slightly with
time and reach a quasi-steady state at approximately 1500 s, indicating that a spontaneous passivation film was formed on the surface
of the pyrite. Table 2 lists the final stable OCP values under different conditions. The OCP values gradually increase with higher
concentration, decrease with rising temperature, and also decrease with higher pH values.

3.1.2. Polarization curve study


Fig. 1 presents the polarization curves of pyrite in simulated acidic soil solutions at (d) different concentrations, (e) different
temperatures, and (f) different pHs. The similar shapes of the curves suggest the same weathering mechanism. When the pyrite erodes
under higher concentration or acidity conditions, the curves move toward the upper right of the coordinate axis, implying that the
corrosion current (icorr) increased and the corrosion potential (Ecorr) became more positive. When the temperature was increased, the
curves moved toward the lower right, indicating a larger icorr and a more negative Ecorr. Specifically, the values of Ecorr and icorr can be
obtained based on Tafel extrapolation, and the weathering rate (k) of the pyrite can be obtained according to Faraday’s equation,
which is presented in Table 3. From a thermodynamics point of view, a more positive corrosion potential Ecorr means more stability, i.
e., the pyrite has a smaller weathering tendency. From a kinetic standpoint, a larger corrosion rate k implies a faster weathering rate.
Micorr
k= . (1)
nF

where k (mg⋅cm-2•h-1) and icorr (µA⋅cm-2) are weathering rate and corrosion current density, respectively; n (dimensionless) represents
the element’s valence state; M (g⋅mol-1) is the atomic weight; and F (C⋅mol-1) is the Faraday constant.
When pyrite samples were weathered in soil solutions with different concentrations, the results show that pyrite has a more positive
corrosion potential Ecorr and a larger corrosion rate k when the soil solution has a higher concentration. According to the chemical
composition of the soil solution, a larger concentration means a thick passive film will be formed during the spontaneous oxidation
process, indicating that the pyrite is difficult to be weathered further. However, a larger concentration results in a larger Cl- con­
centration, leading to increased penetration ability that can more easily breach the passive film.
When pyrite is weathered in an acidic soil solution at different temperatures, the results show that a higher temperature causes a
more negative Ecorr and a larger k meaning that the pyrite is more easily weathered under higher temperatures. This is because the
higher temperature results in more internal energy being transformed into electrochemical energy. When pyrite undergoes weathering
in an acidic soil solution at varying acidities, the findings indicate that a higher acidity (lower pH) leads to a more positive Ecorr,
indicating a reduced weathering tendency. In addition, the increased acidity leads to a higher k value, indicating that a stronger acidity
accelerates the weathering process of pyrite.

Table 2
Open circuit potential of pyrite in simulate acidic soil solutions at different conditions.
Experimental condition pH Temperature (◦ C) Cs (g⋅L-1) OCP (mV)

Different soil solution concentrations 6.0 25 CS, 1 79.8


6.0 25 CS, 2 101.4
6.0 25 CS, 3 113.6
6.0 25 CS, 4 121.9
Different temperatures 6.0 5 CS, 3 156.3
6.0 15 CS, 3 130.1
6.0 25 CS, 3 113.6
6.0 35 CS, 3 97.1
Different pHs 3.0 25 CS, 3 290.9
4.5 25 CS, 3 215.6
6.0 25 CS, 3 113.6

4
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

Table 3
Electrochemical parameters of pyrite electrodes in simulate acidic soil solutions at different conditions.
condition pH Temperature Cs Ecorr icorr k(Fe(II)) kHg kAs kTi kSb
(◦ C) (g⋅L-1) (mV) (µA⋅cm-2) (g⋅m-2⋅y-1) (mg⋅m-2⋅y-1) (mg⋅m-2⋅y-1) (mg⋅m-2⋅y-1) (mg⋅m-2⋅y-1)

concentrations 6.0 25 CS, 1 –38.6 1.03 94.00 302.92 270.61 145.40 8.48
6.0 25 CS, 2 –27.2 1.48 135.07 435.27 388.84 208.93 12.19
6.0 25 CS, 3 –15.1 1.74 158.79 511.71 457.13 245.62 14.33
6.0 25 CS, 4 –10.6 2.32 211.72 682.28 609.50 327.49 19.10
temperatures 6.0 5 CS, 3 29.9 0.78 71.18 229.38 204.91 110.10 6.42
6.0 15 CS, 3 5.8 1.27 115.90 373.49 333.65 179.28 10.46
6.0 25 CS, 3 –15.1 1.74 158.79 511.71 457.13 245.62 14.33
6.0 35 CS, 3 –30.5 2.11 192.56 620.53 554.34 297.86 17.37
pHs 3.0 25 CS, 3 192.4 2.29 208.99 673.48 601.64 323.27 18.86
4.5 25 CS, 3 93.8 2.04 186.17 599.94 535.95 287.97 16.80
6.0 25 CS, 3 –15.1 1.74 158.79 511.71 457.13 245.62 14.33

Ecorr: corrosion potential; icorr: corrosion current density; k corrosion rate.

3.1.3. EIS study


The EIS analysis reveals the electrochemical process at the electrode/electrolyte interface during the oxidation of pyrite, which
provides an effective explanation for the weathering mechanism of pyrite. Fig. 2 shows Nyquist and Bode plots of the pyrite electrode.
All of the Bode diagrams consist of two time constants, reflecting that Nyquist plots have two capacitive loops. The high-frequency loop
results from the resistance of the passive film (Rf) caused by the pseudo-capacitance impedance, while the low-frequency loop is caused
by the double layer charge transfer resistance (Rt) at the pyrite/electrolyte interface. The deviation of the capacitor loops from the ideal
semicircle shape may be attributed to the frequency variation resulting from the passive film’s unevenness (Córdoba-Torres et al.,
2001). The equivalent electrochemical circuit shown in Fig. S1 was used to simulate the pyrite/electrolyte interface. Rs represents the
solution resistance, and CPEf and CPEdl are constant phase elements replacing the passive film capacitance and the charge transfer
capacitance of the double layer, respectively.
Table 4 lists the fitting results of the equivalent circuit model. When pyrite is weathered in an acidic soil solution with different
concentrations, the results show that both the transfer resistance Rt and the film resistance Rf decrease and the value of transfer
capacitance CPEdl and film capacitance CPEf increase. The smaller resistance and larger capacitance reveal that a higher concentration
is beneficial to the charges migrating at the double layer and the passive film. Thus, a higher concentration promotes pyrite weath­
ering. When pyrite undergoes weathering at varying temperatures, higher temperatures facilitate the migration of charges at the
double layer and the passive film, ultimately promoting the weathering of pyrite. When pyrite is weathered in an acidic soil solution at
different acidities, a higher acidity results in a smaller Rt and a larger CPEdl at the double layer. Therefore, a higher acidity promotes
charge migration at the double layer, and it is advantageous to pyrite oxidation. However, a higher acidity causes a larger film
resistance Rf and a smaller film capacitance CPEf, suggesting that it may inhibits pyrite oxidation. This is due to the anode reaction
(Reaction (3)) and cathode reaction (Reaction (2)). On the one hand, a higher acidity (H+) promotes the cathode reaction (Reaction
(2)) and further accelerates the anode reaction (Reaction 3) and enhances the pyrite oxidization. On the other hand, the more severe
reaction (Reaction 3) simultaneously produces more passivation and S0, inhibiting further oxidation of pyrite. In comparison, the EIS
results show that Rt decreases much more than Rf, suggesting the charge transfer is dominant. This is why a higher acidity (H+)
promotes pyrite weathering.

O2+ 4H++ 4e-→H2O (2)

3.1.4. CV study
CV curves can provide information about various redox reactions that may occur on the interface, which is useful to determine the
weathering mechanism of pyrite. From Fig. 3, all of the CV curves are similar in terms of their oxidation-reduction peaks, meaning that
pyrite has the same weathering mechanisms under these different conditions, but the current changes as the conditions change,
indicating that the pyrite weathering rate is related to the concentration, temperature, and pH.
The CV curves all have two anode peaks and three cathode peaks. The first anodic peak, A1, at approximately 400 mV, is likely
caused by pyrite oxidation to Fe2+ and S0 (Reaction (3)) (Antonijevic et al., 2005). Then, the oxidized products (Fe2+ and S0) are
further oxidized at nearly 650 mV (peak A2) (Reactions (4) and (5)).

A1: FeS2 → Fe2+ + 2S0 + 2e- (3)

A2: Fe2+ → Fe3+ + e- (4)


0
S + O2 + 2H2O → 4H + +
SO2-
4 + 2e-
(5)
3+ 2+
During the negative scan, the first cathodic peak (C1) appears at ~ 400 mV, corresponding to the reduction of Fe to Fe as in
Reaction (6) (Tu et al., 2017). The second cathodic peak (C2) appears at ~ − 400 mV, due to the reduction of sulfur (Reaction (7))
(Almeida and Giannetti, 2003). The third peak (C3) appears at ~ − 800 mV, and it is ascribed to the reduction of pyrite (Reaction (8)).

5
X. Feng et al.
6

Environmental Technology & Innovation 35 (2024) 103689


Fig. 2. Nyquist (a, b, c) and Bode plots (d, e, f) for arsenopyrite in simulated acidic soil solutions at different conditions, where ○, □ and × represent the experimental values and ─ represents
simulated values.
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

Table 4
Equivalent circuit model parameters for pyrite in simulate acidic soil solutions at different conditions.
Experimental condition pH Temperature Cs CPEf, Y0 n Rf CPEdl, Y0 n Rt χ2
(◦ C) (g⋅L-1) (S⋅cm-2⋅s-n) (Ω⋅cm2) (S⋅cm-2⋅s-n) (Ω⋅cm2)

Different soil solution concentrations 6.0 25 CS, 1 1.59E-4 0.7485 7.67E4 1.67E-5 0.9736 2.48E5 6.01E-3
6.0 25 CS, 2 1.72E-4 0.8293 4.83E4 6.50E-5 0.8982 1.60E5 1.90E-3
6.0 25 CS, 3 2.49E-4 0.6664 471.1 1.25E-4 0.8957 1.01E5 7.25E-4
6.0 25 CS, 4 2.94E-4 0.7456 156.0 2.28E-4 0.8763 4.99E4 2.33E-3
Different temperatures 6.0 5 CS, 3 1.49E-4 0.6195 5787 3.88E-5 0.6864 4.81E5 4.41E-4
6.0 15 CS, 3 1.80E-4 0.7014 1445 8.45E-5 0.7016 2.49E5 2.46E-4
6.0 25 CS, 3 2.49E-4 0.6664 471.1 1.25E-4 0.8957 1.01E5 7.25E-4
6.0 35 CS, 3 2.83E-4 0.5795 68.88 4.18E-4 0.8597 5.81E4 1.28E-3
Different pHs 3.0 25 CS, 3 8.42E-5 0.7402 1118.0 1.46E-4 0.6294 6.04E4 8.93E-4
4.5 25 CS, 3 2.24E-4 0.7962 993.7 1.58E-4 0.6136 8.41E4 2.86E-4
6.0 25 CS, 3 2.49E-4 0.6664 471.1 1.25E-4 0.8957 1.01E5 7.25E-4

Rf: passive film resistance; Rt: charge transfer resistance; n: dimensionless number;
CPEdl: constant phase element of double layer; CPEf: constant phase element of passive film.

Fig. 3. CV for the pyrite electrode in simulated acidic soil solutions at different conditions (scan rate 20 mV⋅s-1).

It should be noted that the FeS produced in Reaction (8) is easily decomposed in an acidic environment (Reaction (9)) (Tao et al., 2003)

C1: Fe3+ + e- → Fe2+ (6)


0 -
C2: S + 2H + 2e → H2S
+
(7)

C3: FeS2 + 2H+ + 2e-→ FeS + H2S (8)

FeS + 2H+ = Fe2+ + H2S (9)

During the positive return sweep, the last anode peak, at ~ − 200 mV (A3), caused by the reduction of H2S (Giannetti et al., 2001).

A3: H2S → S0 + 2H++ 2e- (10)

3.2. Physicochemical characterizations of weathered pyrite

3.2.1. Pyrite surface morphology analysis


SEM and EDX were employed to identify the morphologies and compositions of the pristine and weathered pyrite. Fig. 4 shows SEM
images of the pristine pyrite and the pyrite weathered for 10 weeks. The surface of the pristine pyrite is smooth and homogeneous.
After 10 weeks of weathering, corrosion products have accumulated on the surface of the pyrite, resulting in uneven, rough surfaces
that even contain corrosion holes. Noticeable cracks appear on the surface, indicating a heterogeneous weathering process. As Mikhlin
et al. (2006) and McGuire et al. (2001) reported, the surface passive film existed as a patch instead of a “blanket-cover layer”, allowing
ions to diffuse through the surface film. In addition, the number of corrosion products and the size of the corrosion holes on the eroded
surface of the Cs,4 sample were greater than those on the surface of the Cs,1 sample at 25 ◦ C and a pH of 6.0. The same phenomenon was
observed for the different acidities and temperatures, implying that a higher concentration, temperature, and acidity promotes pyrite
weathering, which is consistent with the electrochemical results.
Table S3 lists the EDX result of the pristine and weathered pyrite samples. The pristine pyrite specimen contained 68.2 % S and
31.8 % Fe (at%), corresponding with the atomic stoichiometric ratio of pyrite. After 10 weeks of weathering, all of the weathered

7
X. Feng et al.
8

Fig. 4. SEM image of the pristine and pyrite after weathered for 10 weeks. (a) pristine; (b) 25 ◦ C, CS, 1, pH 6.0; (c) 25 ◦ C, CS, 2, pH 6.0; (d) 25 ◦ C, CS, 3, pH 6.0; (e) 25 ◦ C, CS, 4, pH 6.0; (f) 5 ◦ C, CS, 3, pH

Environmental Technology & Innovation 35 (2024) 103689


6.0; (g) 15 ◦ C, CS, 3, pH 6.0; (h) 35 ◦ C, CS, 3, pH 6.0; (i) 25 ◦ C, CS, 3, pH 3.0; (j) 25 ◦ C, CS, 3, pH 4.0.
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

pyrite specimens contain O, Fe, and S elements. Compared with the original pyrite specimen, a larger concentration, higher tem­
perature, or higher acidity results in a higher oxygen content and lower Fe and S contents. These results confirm that a larger con­
centration, higher temperature, or higher acidity promotes pyrite weathering. In addition, the corrosion products generated on the
surface may be iron or sulfur oxides, which will be further discussed in the next section.

3.2.2. Raman spectral analysis


Raman spectroscopy is an important technique for determining chemical and structural information. Fig. 5 shows the Raman
spectra of the pristine and weathered pyrite specimens. For the pristine pyrite, two peaks appeared at ~343 and 379 cm-1, which are
the pyrite characteristic peaks (Lara et al., 2015). For the pyrite specimens eroded in simulated acidic soil solutions for 10 weeks, all of
the samples exhibited three new peaks at 220, 280, and 612 cm-1. The peak at 220 cm-1 is associated with the vibration of the S-S bond
in S0 (Xia et al., 2010; Zhu et al., 2014). The peak at 280 cm-1 is ascribed to goethite (γ-FeOOH), and the peak at 612 cm-1 is ascribed to
hematite (α-Fe2O3). Table S4 summarizes the Raman peak positions of FeS2, S0, γ-FeOOH, and α-Fe2O3.

3.2.3. Surface elemental analysis


To further uncover the electronic valence states and chemical bonding of the S and Fe of the pristine and weathered pyrite
specimens, XPS analysis was performed. The peak parameters for the Fe and S spectra and the contents of the species in the pristine and
eroded pyrite are listed in Table S5 and Table S6, respectively.
In the Fe 2p spectra of the pristine pyrite (Fig. 6), five peaks appeared at about 707.3, 708.3, 709.4, 710.6, and 711.8 eV, which are
ascribed to Fe(II)-S (707.3 eV, 70.15 %), Fe(II)-O (708.3 and 709.4 eV, 15.03 %), and Fe(III)-O (710.6 and 711.8 eV, 11.46 %)
respectively. The Fe and S mainly exist as Fe(II)-S (Mycroft et al., 1990). In the eroded pyrite, an extra peak appears at approximately
713.4 eV, which is in good agreement with Fe(III)-SO (Cai et al., 2009; Tu et al., 2017). As the concentration increased from Cs,1 to Cs,4
(pH of 6.0 and 25 ◦ C), the content of Fe(III)-SO increased from 3.57 % to 10.43 %, while the content of Fe(II)-S decreased from 67.17 %
to 43.13 %. As the temperature increase from 5 ◦ C to 35 ◦ C (pH of 6.0 and Cs,3), the Fe(III)-SO increase from 3.02 % to 8.43 %, while
the Fe(II)-S decrease from 66.84 % to 61.98 %. As the pH increases from 3.0 to 4.5, the contents of Fe(III)-SO and Fe(II)-S increase from
9.61 % to 44.77 % and from 7.24 % to 55.43 %, respectively. All of these results indicate that a larger concentration, higher tem­
perature, or higher acidity facilitates the conversion of Fe(II) to Fe(III), which accelerates the oxidation of pyrite.
For S 2p spectra (Fig. 7), the pristine pyrite exhibits three double peaks, namely, S2- (161.7 eV), S2- 0
2 (162.7 eV), and S (163.9 eV),
2-
and the main component (S2 ) accounted for 80.49 % (Nesbitt et al., 1995; Lara et al., 2015). The eroded pyrite contains two additional
doublets at 166.0 and 168.6 eV, which are attributed to the SO2- 2-
3 and SO4 produced by the pyrite oxidation, repsectively (Cai et al.,
2009). In addition, the contents of the oxidized components (S0, SO2- 2- 2- 2-
3 , and SO4 ) increase and the contents of S and S2 decrease with
increasing salt concentration, temperature, or acidity. These results suggest that a larger concentration, acidity, or temperature
promotes pyrite weathering, which is consistent with the previously described electrochemical and SEM-EDS results.

3.3. Summary of the weathering mechanism of pyrite

Based on the combined results of the electrochemistry and surface measurements, Fig. 8 presents the weathering mechanism of
pyrite. Pyrite weathering is an electrochemical process. At the cathode, the dissolved oxygen is reduced in the acidic solution (Reaction
2). At the anode, first, S in FeS2 is oxidized to a S0 passive film covering the mineral surface. The peak of A1 and A3 in CV curves
detected the presence of S0, and EDX, Raman and XPS results also confirmed the formation of elemental sulfur. Then, the S0 is gradually
transformed into high-valence soluble sulfite (SO2- 2-
3 ) and sulfate (SO4 ), accompanied by the release of hydrogen (H ) ions, resulting in
+

serious soil acidification, it can be seen from Reaction (10) that sulfate and hydrogen ions are formed during the reaction and
tetravalent sulfur and hexavalent sulfur are confirmed by XPS.
The Fe is initially transformed to Fe2+ and subsequently converted into Fe3+. A2 and C1 are the redox peaks of iron in CV curves,
and XPS results also detected the presence of ferrous iron and ferrous iron. In addition, due to the presence of water, Fe3+ is further

Fig. 5. Raman spectra of the pristine and weathered pyrite.

9
X. Feng et al.
10

Environmental Technology & Innovation 35 (2024) 103689


Fig. 6. XPS spectra for Fe 2p (a-j) in the pristine and eroded surfaces of pyrite.
X. Feng et al.
11

Environmental Technology & Innovation 35 (2024) 103689


Fig. 7. XPS spectra for S 2p (a-j) in the pristine and eroded surfaces of pyrite.
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

Fig. 8. Scheme of the weathering mechanism of pyrite in acidic soil solutions.

evolved into iron oxide-hydroxide γ-FeOOH, and it is ultimately transformed into α-Fe2O3. From Raman spectra, the peak at 280 cm-1
is ascribed to goethite (γ-FeOOH), while the peak at 612 cm-1 is to hematite (α-Fe2O3). The Fe3+ comes in contact with the water vapor
in the air and is converted into iron oxide-hydroxide (FeOOH) (Reaction (11)) (Shim and Duffy, 2001; Tu et al., 2017). Iron oxy­
hydroxide (γ-FeOOH) is unstable and is further transformed into α-Fe2O3 (Reaction (12)) (Nesbitt et al., 1995).

2Fe3+ + 4H2O → 2γ-FeOOH + 6H+ (11)

2γ-FeOOH → 2α-FeOOH → α-Fe2O3 + H2O (12)

4. Environmental implications

4.1. Influences of experimental parameters on the weathering tendency of pyrite

Heavy metal ion and H+ levels are important to evaluate surrounding pollution. Based on the open circuit potential tests, the
thermodynamic parameters of pyrite weathering were obtained, and the value of the OCP reflects the pyrite weathering trend. The
OCP value is more positive, indicating less corrosion or more difficulty in weathering, and vice versa. The results presented in Table 2
show that a higher concentration causes a more positive potential, meaning the pyrite has a smaller weathering trend and a thicker
spontaneous passivation film forms. A higher temperature corresponds to a stronger weathering trend because the higher temperature
destroys the passivation film formed on the electrode surface. A higher pH results in a smaller weathering tendency because it reduces
the pyrite’s electrochemical activity (Moslemi et al., 2011).

4.2. Influences of experimental parameters on the weathering rate of pyrite

According to the above polarization curve study, the kinetics parameters of pyrite weathering were obtained. We further quan­
titatively determined the process of pyrite weathering in acid soil solutions and amounts of H+ and heavy metal ions released in a
certain period based on the Faraday equation. From Table 3, it can be seen that higher concentration, temperature and acidity are
consistent with a higher weathering rate. Using a pyrite ore sample from the Lanmuchang mine (Guizhou Province, China) as an
example (Table S2), the polarization results indicate that the Fe(II) release rate is 94.00 g⋅m-2⋅y-1 at a pH of 6.0 and a temperature of 25

C. During the weathering process in a year, 302.92 g of Hg, 270.61 g of As, 145.40 g of Ti, and 8.48 g of Sb will also enter the soil,
resulting in potential pollution problems. With elevated temperature, concentration, and acidity levels, there is a corresponding in­
crease in the release of heavy metal ions.

4.3. The activation energy of pyrite weathering in simulated acid soil

Generally speaking, the rate of a chemical reaction is closely related to its activation energy with lower activation energies cor­
( )
responding to faster reaction rates (Laidler, 1996). The Arrhenius equation defines the activation energy (Ea) as ln kk21 = − ERa T12 − T11 ,
where, k (mol⋅L-1⋅s-1) represents the reaction rate constant, A (mol⋅L-1⋅s-1) is pre-exponential factor, e (dimensionless) means Napierian
base, Ea (J⋅mol-1) represents activation energy, R (J⋅mol-1⋅K-1) is the molar gas constant(8.314 J⋅mol-1⋅K-1), and T (K) means tem­
perature. The relationship is lnk = -2844/T+3.1718 (R2 = 0.9723), as shown in Fig. S2.
The activation energy of pyrite weathering in a simulated acidic soil solution is 21.21 kJ⋅mol-1, which is less than 40 kJ⋅mol-1,
indicating that pyrite weathering readily occurs in acidic soil (Badawy, 2000). Lasaga (1998) pointed out that when the activation
energy is > 20 kJ⋅mol-1, the reaction is controlled by the surface reaction; and when it is < 20 kJ⋅mol-1, it is controlled by diffusion. In
this study, the value of Ea is > 20 kJ⋅mol-1, suggesting that pyrite weathering is related to the surface reaction.
The weathering of pyrite in an acidic soil solution will lead to further soil acidification and result in heavy metal ion pollution. The
soil solution’s concentration, pH, and temperature affect the trend of the weathering and the dissolution rate. Thus, the surrounding
environmental conditions, such as the soil geological background, acid rain, latitude, and season will all affect the degree of soil
acidification and the amount of heavy metal ion pollution.

12
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

5. Conclusions

The weathering mechanism of pyrite in acidic soil was investigated by electrochemistry coupled with surface analysis techniques.
The OCP results show that pyrite is easily weathered in larger temperature, concentration or acidity soil solutions. The polarization
curves show that a larger temperature, concentration or acidity promotes the pyrite weathering rate. Furthermore, the EIS mea­
surements confirm that the causes are the greater capacitance and smaller resistance. The oxidative dissolution of pyrite is controlled
by surface interactions with an activation energy of 21.21 kJ⋅mol-1. The CV curves and surface analysis data show that the pyrite is
initially oxidized to Fe2+ and S0, which are then transformed into Fe3+ and SO2- 4 , and H ions are released and increase soil acidifi­
+

cation. Some of the Fe3+ may be further transformed into γ-FeOOH and α-Fe2O3. These results provide a demonstration of a rapid and
quantitative calculation of the metal ions released, reveal the geochemical cycle of Fe and S in mining acidic soil containing pyrite, and
demonstrate the significance of assessing the risks of pyrite in acidic soil.

CRediT authorship contribution statement

Qingyou Liu: Writing – review & editing, Writing – original draft, Project administration, Methodology, Investigation, Funding
acquisition, Data curation, Conceptualization. Xiaonan Feng: Writing – original draft, Resources, Methodology, Formal analysis, Data
curation. Bing-Jie Ni: Writing – original draft, Methodology. Ling Cen: Writing – original draft, Methodology, Data curation. Zhijie
Chen: Writing – original draft, Methodology, Data curation. Shuai Wang: Writing – original draft, Methodology, Data curation.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data Availability

The data that has been used is confidential.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (41873074). We thank LetPub (www.
letpub.com) for its linguistic assistance during the preparation of this manuscript.

Appendix A. Supporting information

Supplementary data associated with this article can be found in the online version at doi:10.1016/j.eti.2024.103689.

References

Almeida, C.M., Giannetti, B.F., 2003. Electrochemical study of arsenopyrite weathering. Phys. Chem. Chem. Phys. 5 (3), 604–610.
Antonijevic, M.M., Dimitrijevic, M.D., Serbul, S.M., Dimitrijevi, V.L.J., Bogdanovic, G.D., Milic, S.M., 2005. Influence of inorganic anions on electrochemical
behaviour of pyrite. Electrochim. Acta 50, 4160–4167.
Badawy, W.A., Al-Kharafi, F.M., Al-Ajmi, J.R., 2000. Electrochemical behaviour of cobalt in aqueous solutions of different pH. J. Appl. Electrochem. 30 (6), 693–704.
Cai, Y., Pan, Y., Xue, J., Xue, J., Sun, Q., Su, G., Li, X., 2009. Comparative XPS study between experimentally and naturally weathered pyrites. Appl. Surf. Sci. 255
(21), 8750–8760.
Caldeira, C.L., Ciminelli, V.S.T., Osseo-Asare, K., 2003. Pyrite oxidation in alkaline solutions: nature of the product layer. Int. J. Miner. Process. 72, 373–386.
Córdoba-Torres, P., Nogueira, R.P., de Miranda, L., Brenig, L., Wallenborn, J., Fairén, V., 2001. Cellular automaton simulation of a simple corrosion mechanism:
mesoscopic heterogeneity versus macroscopic homogeneity. Electrochim. Acta 46 (19), 2975–2989.
Fan, R., Short, M., Zeng, S.-J., Qian, G., Li, J., Schumann, R., et al., 2017. The formation of silicate-stabilised passivating layers on pyrite for reduced acid rock
drainage. Environ. Sci. Technol. 51 (19), 11317–11325.
Gao, T.Y., Shi, Y., Liu, F., Zhang, Y.S., Feng, X.H., Tan, W.F., Qiu, G.H., 2015. Oxidation process of dissolvable sulfide by synthesized todorokite in aqueous systems.
J. Hazard. Mater. 290, 106–116.
Giannetti, B.F., Bonilla, S.H., Zinola, C.F., Raboczkay, T., 2001. A study of the main oxidation products of natural pyrite by voltammetric and photoelectrochemical
responses. Hydrometallurgy 60, 41–53.
Laidler, K.J., 1996. A glossary of terms used in chemical kinetics, including reaction dynamics. Pure Appl. Chem. 68, 149–192.
Lara, R.H., Monroy, M.G., Mallet, M., 2015. An experimental study of iron sulfides weathering under simulated calcareous soil conditions. Environ. Earth Sci. 73 (4),
1849–1869.
Lasaga, A., 1998. Kinetic theory in the earth sciences. Princeton series in geochemistry. Princeton university press, princeton, New Jersey.
Liu, Z.Y., Li, X.G., Du, C.W., Lu, L., Zhang, Y.R., Cheng, Y.F., 2009. Effect of inclusions on initiation of stress corrosion cracks in X70 pipeline steel in an acidic soil
environment. Corros. Sci. 51, 895–900.
McGuire, M.M., Jallad, K.N., Ben-Amotz, D., Hamers, R.J., 2001. Chemical mapping of elemental sulfur on pyrite and arsenopyrite surfaces using near-infrared Raman
imaging microscopy. Appl. Surf. Sci. 178, 105–115.
Mikhlin, Y.L., Romanchenko, A.S., Asanov, I.P., 2006. Oxidation of arsenopyrite and deposition of gold on the oxidised surfaces: a scanning probe microscopy,
tunneling spectroscopy and XPS study. Geochim. Cosmochim. Acta 70, 4874–4888.

13
X. Feng et al. Environmental Technology & Innovation 35 (2024) 103689

Moslemi, H., Shamsi, P., Habashi, F., 2011. Pyrite and pyrrhotite open circuit potentials study: effects on flotation. Miner. Eng. 24 (10), 1038–1045.
Mycroft, J.R., Bancroft, G.M., McIntyre, N.S., Lorimer, J.W., Hill, I.R., 1990. Detection of sulphur and polysulphides on electrochemically oxidized pyrite surfaces by
X-ray photoelectron spectroscopy and Raman spectroscopy. J. Electroanal. Chem. 292 (1), 139–152.
Nesbitt, H.W., Muir, I.J., Pratt, A.R., 1995. Oxidation of arsenopyrite by air and airsaturated, distilled water and implications for mechanisms of oxidation. Geochim.
Cosmochim. Acta 59, 1773–1786.
Ogbughalua, O.T., Vasileiadis, S., Schumann, R.C., Gerson, A.R., Li, J., Smart, R.S.C., Short, M.D., 2020. Role of microbial diversity for sustainable pyrite oxidation
control in acid and metalliferous drainage prevention. J. Hazard. Mater. 393, 122338.
Shim, S.H., Duffy, T.S., 2001. Raman spectroscopy of Fe2O3 to 62 GPa. Am. Min. 87, 318–326.
Tao, D.P., Richardson, P.E., Luttrell, G.H., Yoon, R.H., 2003. Electrochemical studies of pyrite oxidation and reduction using freshly-fractured electrodes and rotating
ring-disc electrodes. Electrochim. Acta 48, 3615–3623.
Tu, Z., Wan, J., Guo, C., Fan, C., Zhang, T., Lu, G., Reinfelder, J.R., Dang, Z., 2017. Electrochemical oxidation of pyrite in pH 2 electrolyte. Electrochim. Acta 239,
25–35.
von Uexküll, H.R., Mutert, E., 1995. Global extent, development and economic impact of acid soils. Plant Soil 171, 1–15.
Wu, W., Qu, S., Nel, W., Jia, J., 2020. The impact of natural weathering and mining on heavy metal accumulation in the karst areas of the Pearl River Basin, China. Sci.
Total Environ. 734, 139480.
Xia, L., Yin, C., Dai, S., Qiu, G., Chen, X., Liu, J., 2010. Bioleaching of chalcopyrite concentrate using leptospirillum ferriphilum, acidithiobacillus ferrooxidans and
acidithiobacillus thiooxidans in a continuous bubble column reactor. J. Ind. Microbiol. Biotechnol. 37, 289–295.
Zheng, K., Liu, Q., Li, H., 2019. Assessing the influence of humic acids on the weathering of pyrite: electrochemical mechanism and environmental implications.
Environ. Pollut. 51, 738–745.
Zhu, T., Lu, X., Liu, H., Li, J., Zhu, X., Lu, J., Wang, R., 2014. Quantitative X-ray photoelectron spectroscopy-based depth profiling of bioleached arsenopyrite surface
by Acidithiobacillus ferrooxidans. Geochim. Cosmochim. Acta 127, 120–139.

14

You might also like