SB (V)
SB (V)
SB (V)
Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol
A R T I C L E I N F O A B S T R A C T
Keywords: Antimonate is the dominant form of antimony (Sb) in Sb mine water. The treatment of high-Sb mine water
Oxyanions essentially reduces the discharge of antimonate oxyanions ([Sb(OH)6]− ) in it. Biochar obtained from
Antimonate removal phosphogypsum-modified anaerobic digested distillers’ grain (PADC) can effectively adsorb antimonate from
Modified biochar
water. In this work, using batch adsorption experiments, mathematical models, and characterization methods,
Adsorption mechanisms
the mechanism of Sb(V) adsorption by PADC was studied. Compared with pristine biochar, PADC biochar
showed abundant lamellar and vesicular structures with significant calcium ion loading on the surface. The
kinetics data of the adsorption of Sb(V) on the PADC biochar followed the Elovich equation (R2 = 0.992),
indicating that heterogeneous adsorption had occurred. The results also showed that intraparticle diffusion
played an important role in controlling Sb(V) adsorption by PADC biochar. The Redlich–Peterson model best fit
the Sb(V) adsorption isotherm (R2 = 0.997), indicating that the adsorption was a combination of the Langmuir
and Freundlich models. The maximum adsorption capacity of PADC biochar for Sb(V) is 8123 mg/kg, which is
more than twice that of the pristine biochar (3487 mg/kg) and is sufficient for Sb(V) treatment in most mine
water. Fourier transform infrared (FTIR) spectra, X-ray photoelectron spectroscopy (XPS), X-ray diffractometry
(XRD), and Transmission electron microscopy with energy dispersive X-ray spectroscopy (TEM-EDS) analyses
revealed that the dominant mechanism of Sb(V) removal by PADC biochar was the formation of Ca–O–Sb
complexes or amorphous surface precipitation as well as electrostatic adsorption. This work demonstrated the
potential of PADC biochar in the treatment of Sb-contaminated mine water.
1. Introduction typically less than 1 μg/L (Filella et al., 2002), while Sb-contaminated
water in Sb mine areas can have Sb contents tens to thousands of
Antimony (Sb), as an emerging global environmental pollutant, has times as high (He et al., 2019). In abandoned Sb mines, the Sb con
had a negative impact on human life due to its increasing accumulation centration in water can be as high as 13,350 μg/L (Li et al., 2019). The
in the environment and has gradually attracted considerable concern in direct discharge of Sb-containing wastewater can pose a great threat to
recent years (Nishad and Bhaskarapillai, 2021; Rinklebe et al., 2020; human health through drinking water and the food chain. Thus,
Herath et al., 2017). As the country with the largest Sb storage, by 2018, Sb-contaminated wastewater treatment is critically needed (Li et al.,
China was still the leading producer of Sb, accounting for more than 2018a).
70% of the world’s production (USGS, 2019). Due to the importance of High-Sb mine water isthe key target of Sb pollution treatment in
Sb in flame retardants and alloys, the demand for Sb is increasing, and mine areas. Mine water Sb removal is essentially the removal of [Sb
Sb mining activities are extensive and frequent. As a result, Sb pollution (OH)6]− . Because of this, [Sb(OH)6]− is the most stable form in most
has become a common environmental issue in China (Filella and He, oxygen-containing water bodies with a wide pH range. In general,
2020; He et al., 2012). In non-polluted natural water, Sb content is adsorption, coagulation/flocculation, ion exchange, membrane
☆
This paper has been recommended for acceptance by Jörg Rinklebe.
* Corresponding author.
E-mail address: pwu@gzu.edu.cn (P. Wu).
https://doi.org/10.1016/j.envpol.2022.119032
Received 6 October 2021; Received in revised form 11 February 2022; Accepted 17 February 2022
Available online 22 February 2022
0269-7491/© 2022 Elsevier Ltd. All rights reserved.
L. Li et al. Environmental Pollution 301 (2022) 119032
separation, electrochemistry, and extraction are the primary methods PADC600 biochar as an adsorbent to carry out the adsorption study of Sb
used in Sb removal (Long et al., 2020; Li et al., 2018a). Among these, (V). ADC600 biochar was taken as its comparative adsorbent sample to
adsorption is the most widely used due to its simple operation and high better reveal the adsorption mechanism.
removal yield. Biochar, as an adsorbent for wastewater treatment, is The specific surface area and pore size of biochar were calculated
attracting growing attention (Amen et al., 2020, Wang et al., 2017; based on the Brunauer–Emmett–Teller (BET) gas adsorption method
Oliveira et al., 2017). Compared with traditional biochar, the (GB/T 19,587–2004) using the 3H–2000PS2 system from Beishide In
metal-biochar composites have a higher capacity to remove oxyanions strument (Beijing, China). The experiments were conducted at a pre-set
such as antimonate (Jia et al., 2020; Wang et al., 2019, 2018), but the temperature of 150 ◦ C for 8 h to measure the absorption of high-purity
production cost is high and the metal-biochar composites are not suit liquid nitrogen. The zeta potential was measured with a Malvern Zeta
able for large-scale promotion or application in actual production (Li meter (Malvern Panalytical Instruments Ltd., United Kingdom). To
et al., 2018b). By using low-cost biochar produced from anaerobically measure the pH, biochar and DI-water were mixed completely in a mass
digested residues, the oxyanions can be removed efficiently, with ex ratio of 1:20, and the mixture was allowed to stand for 10 min. The pH
amples having been reported for phosphorous (P) removal (Inyang et al., was then measured by a PHS-3C pH reader using a Shanghai Leici
2010; Yao et al., 2011a, 2011b; Zhang et al., 2012). Moreover, the Instrument.
anaerobic digestion of biochar after phosphogypsum (industrial waste)
modification has a more significant adsorption capacity for oxyanions. 2.3. Batch adsorption of antimonate
Our recent studies showed that the maximum phosphate and chromate
adsorption capacity of modified biochar obtained via anaerobic pyrol All the adsorption tests in this study were based on the optimal
ysis at different temperatures by mixing anaerobic digested distillers’ conditions (Figs. S1, S2, and S3), which are described below. Adsorption
grain and phosphogypsum are respectively 5 times and 2.5 times of that kinetics: First, 0.1 g of biochar (ADC600 or PADC600) was added to a
before modification (Wang et al., 2020; Lian et al., 2019). Antimonate is 50-mL centrifuge tube and mixed with 24.36 mg/L of an Sb(V) solution
an analogue of phosphate and chromate (Shi et al., 2021; Marouane to a total of 40 mL. The pH of the solution was adjusted to 7.5. The
et al., 2021). Thus, this study proposed that based on the characteristics solution was then shaken at room temperature at speed of 200 r/min.
of Sb-contaminated mine water, the use of this Samples were extracted after 0.1, 0.25, 0.5, 1, 2, 4, 8, 16, 24, 36, and 48
phosphogypsum-modified anaerobic digested biochar (PADC) to h. These samples were then centrifuged and filtered with a 0.22-μm filter
remove Sb(V) could be an economically feasible treatment method. membrane. The filtrates were reserved for Sb(V) concentration
In this work, tests and analyses on the feasibility of Sb(V) removal measurement.
using this PADC biochar were performed. To further explore the removal Adsorption isotherms: First, 0.1 g of biochar (ADC600 or PADC600)
mechanism and to estimate its potential applications for Sb mine water was placed in a 50-mL centrifuge tube, and they were mixed with a
treatment based on the pollution characteristics, Sb(V) adsorption with series of Sb(V) solutions with different concentrations (0–181.5 mg/L)
biochar was studied using the batch adsorption method. Typical char to obtain a total volume of 40 mL. The pH of the solution was adjusted to
acterization methods were used to examine the following: (1) the 7.5. The solution was then shaken at room temperature at a speed of 200
characteristics of biochar before and after modification and after Sb(V) r/min for 24 h. These samples were then centrifuged and filtered with a
adsorption, (2) the kinetic process and equilibrium isotherm of PADC 0.22-μm filter membrane. The filtrates were saved for Sb(V) concen
biochar after Sb(V) adsorption to determine its adsorption behavior, and tration measurements.
(3) the mechanism of Sb(V) removal by PADC biochar. In this study, the
effectiveness of such an approach in the use of PADC biochar to treat Sb 2.4. Characterization before and after modification and adsorption
mine water was examined.
Scanning electron microscopy (SEM) and energy dispersive X-ray
2. Materials and methods spectroscopy (EDS) (JSM-6460LV (EDAX-Genesis) were utilized to
qualitatively analyze the surface morphologies and chemical composi
2.1. Materials and reagents tions of the biochar before and after modification. Transmission electron
microscopy (TEM) and energy dispersive X-ray spectroscopy (EDS)
Anaerobic digested distillers’ grain and phosphogypsum were (Tecnai G2 F20 S-TWIN) were utilized to analyze the surface morphol
respectively collected at a distillery and a phosphate fertilizer plant in ogies and chemical compositions of the biochar before and after Sb(V)
Guizhou. Potassium pyroantimonate (K2H2Sb2O7⋅4H2O) was purchased adsorption. The main mineralogical phase composition was determined
from Xi’an Tianmao Chemical. All the reagents were of analytical grade by X-ray diffractometry (XRD, D8 ADVANCE, BRUKER). Fourier trans
or higher. DI water was used in all of the experiments (18.2 MΩ cm− 1, form infrared (FTIR) spectra (FTIR-850, GangDong) was used to analyze
Milli-Q). the composition and changes of chemical bonds and functional groups in
the biochar samples before and after adsorption of Sb(V). The scanning
2.2. Preparation of PADC biochar and determination of its properties wavelength of the IR spectrometer was 400–4000 cm− 1, and the reso
lution was 4 cm− 1. X-ray photoelectron spectroscopy (XPS, PHI5000
In this study, distillers’ grain was used as the biomass material, and VersaProbe) was performed on PADC before and after Sb(V) adsorption,
phosphogypsum was used as the modifier. PADC biochar can be pre and charge calibration was based on the alkyl carbon or outer carbon
pared by mixing these two materials in a 2:1 wt ratio in a tube furnace peak in the C 1s spectrum (284.8 eV). Representative biochar samples of
under a nitrogen atmosphere. Biochar was obtained by pyrolysis at each stage were ground with an agate mortar (Jiang et al., 2022; Wei
temperatures of 300, 400, 500, and 600 ◦ C. The resulting PADC biochar et al., 2020). Except for the powder (<300 mesh) used for XRD analysis,
samples were named PADC300, PADC400, PADC500, and PADC600, powder (<100 mesh) biochar samples were used for other character
respectively. For comparison, the pristine biochar samples from the ization analyses.
anaerobic digestion of distiller’s grain (ADC) without phosphogypsum
modification at the same pyrolysis temperatures were named ADC300, 3. Results and discussion
ADC400, ADC500, and ADC600, respectively. The detailed procedure
can be found in our previous publication (Wang et al., 2020). Based on 3.1. Physical and chemical properties of biochar
the results of the adsorption test (Fig. S1), it was found that biochar with
a pyrolysis temperature of 600 ◦ C had the best adsorption effect. The basic properties of the two biochars used in this work have been
Therefore, the batch adsorption experiment of this study will mainly use reported previously (Wang et al., 2020). The PADC600 and ADC600
2
L. Li et al. Environmental Pollution 301 (2022) 119032
biochars are both alkaline. After modification, the specific surface area increased from 3.06 to 7.74 (Lian et al., 2019). This indicated that the
increased from 0.92 to 13.67 m2/g and the pore volume increased from PADC600 biochar was positively charged at neutral pH. Since Sb(V) was
0.0005 to 0.0096 m3/g, while the average pore size decreased from present in the form of the oxyanion [Sb(OH)6]− over a wide pH range of
252.97 to 19.97 nm. This indicated that most pores in the PADC600 2.0–14.0 (He et al., 2019),the positively charged adsorbent had a
biochar could provid more adsorption sites for the adsorption of Sb(V). stronger adsorption capability than the negatively charged adsorbent.
The surface charge of a particle at a given pH can be determined by Therefore,the surface zeta potential of PADC600 gradually became
the zeta potential (Erdemoglu and Sarikaya, 2006). In general, particles positive, which may be favorable for oxyanion electrostatic adsorption.
with zeta potentials between − 10 and + 10 mV are neutral, and particles Previous research showed that the zeta potential on the biochar
with zeta potentials higher than +30 mV and lower than − 30 mV are surface can be increased by the penetration of calcium (Ca) ions into its
strong cations and strong anions, respectively (Clogston and Patri, pores. As the Ca ion concentration gradually increases, the zeta potential
2011). In this study, zeta potential of the ADC600 biochar was − 36.61 can become positive (Viallis-Terrisse et al., 2001). The Ca content of the
mV, while that of the PADC600 biochar was +0.95 mV (Wang et al., biochar modified by phosphogypsum increased from 1.03% to 20.11%.
2020). After the biochar was modified by phosphogypsum, its pHpzc This could explain how the biochar after modification had a positive zeta
Fig. 1. The SEM images (a, b, c & d) and corresponding EDS spectra (e & f) of the two biochars: (a) ADC, 300X; (b) PADC, 300X; (c) ADC, 1500X; (d) PADC, 1500X;
(e) ADC; and (f) PADC.
3
L. Li et al. Environmental Pollution 301 (2022) 119032
Fig. 2. The XRD patterns: (a) ADC and PADC; (b) Sb-laden PADC and PADC.
4
L. Li et al. Environmental Pollution 301 (2022) 119032
fit to the adsorption kinetics data, with R2 values for the ADC and PADC
1
biochars as high as 0.986 and 0.992, respectively. This indicated that the RL = (1)
1 + KL C0
Sb adsorption by the biochars was a heterogeneous surface adsorption
process. where KL (L/mg) is the Langmuir adsorption coefficient and C0 is the
Intra-particle surface diffusion has a strong impact on the adsorption initial maximum Sb concentration. Adsorption is favorable when 0 < RL
process (Jang et al., 2018; Axe and Trivedi, 2002). The intra-particle < 1; otherwise, it is not. When RL = 0, adsorption is irreversible,; while
model was used to understand whether Sb adsorption on the biochar RL = 1 corresponds to linear adsorption. In this study, the RL values for
was limited by diffusion (Yao et al., 2011a). Fig. 3b shows the correla the PADC600 and ADC600 biochars were 0.187 and 0.147, respectively
tions of the amount of surface Sb adsorption and the square root of the (Table S2), indicating that Sb(V) adsorption with both biochars was
adsorption time for the ADC (R2 = 0.964) and PADC (R2 = 0.945) bio favorable. The Langmuir maximum adsorption capacity of PADC600
chars, which followed linear relationships. This indicated that was 8123 mg/kg, which was more than twice that of ADC600 (3487 mg/
intra-particle surface diffusion significantly impacted Sb(V) adsorption kg).
by biochar.
3.4. Adsorption isotherm 3.5. Mechanism of Sb(V) removal by the PADC biochar
An adsorption isotherm describes the interactions between the To understand the adsorption mechanism, TEM-EDS, XRD, FTIR, and
adsorbate and adsorbent until equilibrium is reached. To explore the XPS analyses were performed on PADC600 before and after Sb(V)
features of the antimonate adsorption isotherms induced by the biochar, adsorption. Sb was successfully loaded into the PADC biochar (Fig. S5).
three equilibria isotherm models were used to fit the experimental data: FTIR spectra of PADC before and after Sb(V) adsorption are shown in
Langmuir, Freundlich, and Redlich–Peterson (Eqs. (S6)–(S8)) (see Sup Fig. 5. Bands between 3600 and 3000 cm− 1 correspond to the OH
plementary material). (Rinklebe et al., 2020; Nakamoto, 2009) and NH groups (Cardenas et al.,
The Langmuir model assumes that a single molecular adsorption 2004). A peak at 1621 cm− 1 is also the stretching mode of the OH band
process occurs on a homogeneous surface, which is dominated by ion for free water (Dong et al., 2015). Other bands between 2965 and 2880
exchange. Once the adsorbate occupies a specific location on the cm− 1 were attributed to the stretching vibrations of CH, CH2, and CH3
adsorbent surface, there will be no further adsorption at this location. (Cardenas et al., 2004). The peaks at 1155 and 599 cm− 1 are the typical
The Freundlich model assumes that a multi-layer adsorption process wavelengths of CaSO4 (Song et al., 2007), which also confirmed the
occurs on a heterogeneous surface, which is dominated by complexa successful doping of CaSO4 into the PADC biochar. Peaks at 1121 cm− 1
tion. The Redlich–Peterson model is a three-parameter adsorption and the shoulder at 1046 cm− 1 were assigned to the symmetric and
isotherm model and is a combination of the Langmuir and Freundlich asymmetric stretches of SO42− (Boily et al., 2010), since phosphogyp
models. At higher concentrations, the model is closer to the Freundlich sum (CaSO4⋅2H2O) was used as a modifier and PADC contained sulfates.
model, while at lower concentrations, it is closer to the Langmuir model. After Sb(V) adsorption, the intensity of the OH peak at 1621 cm− 1
Sb(V) adsorption data of the biochar before and after modification decreased, but a shoulder corresponding to Sb(OH)6− appeared at 1648
was fit well by all three models (Fig. 4), with R2 values greater than 0.95 cm− 1 (McComb et al., 2007). At the same time, the peak intensities of
(Table S2). The Redlich–Peterson model had the best fit (R2 > 0.995), CaSO4 at both 1155 and 599 cm− 1 also decreased noticeably, indicating
indicating that Sb(V) adsorption was a heterogeneous process controlled that after PADC adsorbed Sb, a new chemical bond formed between
by both the Langmuir and Freundlich models. This result is consistent CaSO4 and Sb(OH)6− . The SO42− spectrum did not change significantly
with the results of the kinetics study, that is, the adsorption of Sb(V) on after adsorption. Although the peak at 1121 cm− 1 had weakened
the surface of the PADC600 biochar is a mononuclear and polynuclear slightly, the shoulder at 1046 cm− 1 was slightly increased, suggesting
adsorption process on a heterogeneous surface. that the S–O bond was not involved in the formation of new compounds
To further confirm the effectiveness of the Langmuir isotherm, so after Sb(V) adsorption. Therefore, the FTIR results revealed that for
that the maximum amount of Sb(V) adsorption could be better esti mation after Sb(V) adsorption complexes of Ca ions and Sb(OH)6−
mated, a unitless factor RL was defined as follows: formed, which can further form a mineral phase-stabilized calcium
Fig. 4. Adsorption isotherms of Sb(V) onto ADC and PADC. (Symbols are
adsorption data and lines are modeling results). Fig. 5. FTIR spectra of PADC before and after Sb(V) adsorption.
5
L. Li et al. Environmental Pollution 301 (2022) 119032
antimonate precipitate on the surface of PADC (Rinklebe et al., 2020; areas are affected by carbonate from the surrounding rock (She et al.,
Cornelis et al., 2008). 2021), the water in an Sb mine area always shows high pH and Sb
Wide-scan XPS spectra (Fig. 6a) showed that the chemical compo contents (Li et al., 2019; Guo et al., 2018; Li et al., 2018c; Wang et al.,
sition of PADC was C, Ca, S, and O, which agreed with the EDS and XRD 2011; Majzlan et al., 2011). The optimal pH range of the biochar in this
results. Regardless of whether the initial Sb(V) concentration was 1 or study was 7–9 (Fig. S3), i.e., neutral to weakly alkaline, indicating that
1.5 mM, the intensity of the Ca peak of PADC significantly decreased Sb(V) adsorption by this PADC biochar was favorable in most mine
after adsorption. Thus, Sb(V) removal by PADC was indeed related to the water in the Sb mine area. The maximum Sb(V) adsorption amount by
participation of Ca. The binding energy (BE) showed a positive shift by this biochar was 8123 mg/kg, which is sufficiently high for practical Sb
increasing from 530.7 to 531.1 eV after adsorption (Fig. 6b), indicating treatment in mine water. Compared with other metal-biochar complex
that a redox reaction had occurred. However, there was an overlap be adsorbents (Li et al., 2018b), the biochar used in this work is environ
tween the O 1s and Sb 3d5/2 spectra, and the 531.1 eV peak after mentally friendly. It also has a low cost and is easy to use. Thus, it has
adsorption was attributed to these. It can be seen that Sb had been potential application in the removal of Sb in mine water.
chemically bound to PADC, causing the Sb 3d3/2 peak to appear in the
narrow-scan O 1s + Sb 3d spectrum (Fig. 6b). The corresponding (BE) of 4. Conclusion
Sb 3d3/2 was located at 539.6 eV, which is the typical Sb 3d3/2 core for
Sb(V) (Chen et al., 2020; Miao et al., 2014). These confirmed the for In recent years, the Sb concentration in the environment has
mation of Ca–O–Sb complexes on the surface of the PADC biochar. increased rapidly as mining for Sb has increased. This has created a need
Compared with the PADC biochar before adsorption (Fig. S4b), TEM- for the study of different treatment options for the removal of Sb from
EDS analysis showed that certain agglomerate particles with Ca and O as mine water. Antimonate is the target for most Sb-containing wastewater
the main chemical components dominated antimonate adsorption treatment methods. In this study, the adsorption of antimonate using
(Fig. S5). The XRD patterns (Fig. 2b) showed that the characteristic anaerobic digested distillers’ grain modified by phosphogypsum was
absorption peak positions on the surface of the PADC biochar before and examined. In a neutral aqueous environment, the positive charge carried
after adsorption did not change significantly. The characteristic ab on the surface of the modified biochar provided the possibility of the
sorption peaks were all attributed to CaSO4, indicating that the crystal electrostatic adsorption of antimonate. FTIR spectra suggested that a
form of PADC did not change after the adsorption of Sb(V) and that new compound was formed between Ca ions and Sb(OH)6− , which may
adsorbed Sb(V) mainly existed in amorphous form. lead to the formation of a calcium antimonate precipitate. The XRD
patterns further indicated that the precipitate is amorphous. TEM-EDS
and XPS analyses confirmed that the agglomerate particles with Ca
3.6. Implication and O in their primary chemical composition dominated the Sb(V)
adsorption, resulting in the formation of Ca–O–Sb complex and could be
Mining is the main source of Sb introduced by humans into the a mechanism for the removal of Sb(V) from water.
environment. In mine water, Sb mainly comes from the oxidation of Sb- The maximum Langmuir Sb(V) adsorption amount by the modified
containing sulfides (such as stibnite and Sb2S3) (Ashley et al., 2003). The biochar was 8123 mg/kg, which in theory is sufficiently high for Sb(V)
oxidative and aqueous dissolution of Sb-containing sulfides or oxides is treatment in mine water. Meanwhile, this PADC biochar can also be used
the most common pathway for Sb to enter the environment (Yan et al., as a pH buffer agent for acidic mine water. The next step will be to
2020; Hu et al., 2017; Zhou et al., 2017; Nyirenda et al., 2016). When Sb perform antimonate removal in natural mine water using this PADC
enters a water body, sulfate-type acidulation can occur (Biver and biochar to provide substantial scientific evidence for the superior Sb
Shotyk, 2013, 2012), meaning that the pH decreases and the Sb content pollution treatment capabilities of this material in high-Sb mine water.
increases. On one hand, the biochar used in this work was alkaline
(Wang et al., 2020), similar to the pH of other high-temperature Credit author statement
anaerobic biochars (Inyang et al., 2010; Yao et al., 2011a, 2011b).
Due to the high pH of this biochar, it can be used as a buffering agent for Li Ling: Conceptualization, Formal analysis, Data curation, Writing –
acidic Sb-containing mine water. On the other hand, the content and original draft, Writing- Reviewing and Editing, Funding acquisition.
distribution of Sb in water are closely correlated to water–rock in Liao Lu: Investigation, Data curation, Validation. Wang Bing: Resources,
teractions and the type of water (Wu et al., 2017). Since most Sb mine
Fig. 6. The XPS spectra of PADC and Sb-laden PADC: (a) wide scan; (b) narrow scan spectra of O 1s and Sb 3d.
6
L. Li et al. Environmental Pollution 301 (2022) 119032
Methodology. Li Wei: Methodology. Liu Taoze: Methodology. Wu Pan*: Guo, W., Fu, Z., Wang, H., Song, F., Wu, F., Giesy, J.P., 2018. Environmental
geochemical and spatial/temporal behavior of total and speciation of antimony in
Conceptualization, Supervision, Writing- Reviewing and Editing,
typical contaminated aquatic environment from Xikuangshan, China. Microchem. J.
Funding acquisition. Xu Qingya: Visualization. Liu Shirong: 137, 181–189. https://doi.org/10.1016/j.microc.2017.10.010.
Investigation. He, M.C., Wang, N., Long, X., Zhang, C., Ma, C., Zhong, Q., Wang, A., Wang, Y.,
Pervaiz, A., Shan, J., 2019. Antimony speciation in the environment: recent
advances in understanding the biogeochemical processes and ecological effects.
J. Environ. Sci. 75, 14–39. https://doi.org/10.1016/j.jes.2018.05.023.
Declaration of competing interest
He, M.C., Wang, X.Q., Wu, F.C., Fu, Z.Y., 2012. Antimony pollution in China. Sci. Total
Environ. (421–422), 41–50. https://doi.org/10.1016/j.scitotenv.2011.06.009.
The authors declare that they have no known competing financial Herath, I., Vithanage, M., Bundschuh, J., 2017. Antimony as a global dilemma:
geochemistry, mobility, fate and transport. Environ. Pollut. 223, 545–559. https://
interests or personal relationships that could have appeared to influence
doi.org/10.1016/j.envpol.2017.01.057.
the work reported in this paper. Hu, X.Y., He, M.C., Li, S., Guo, X., 2017. The leaching characteristics and changes in the
leached layer of antimony-bearing ores from China. J. Geochem. Explor. 176, 76–84.
https://doi.org/10.1016/j.gexplo.2016.01.009.
Acknowledgement Inyang, M., Gao, B., Pullammanappallil, P., Ding, W.C., Zimmerman, A.R., 2010. Biochar
from anaerobically digested sugarcane bagasse. Bioresour. Technol. 101 (22),
The authors sincerely appreciate Zhao Wang, a PhD student in the 8868–8872. https://doi.org/10.1016/j.biortech.2010.06.088.
Jang, H.M., Yoo, S., Choi, Y.-K., Park, S., Kan, E., 2018. Adsorption isotherm, kinetic
School of Earth Sciences and Engineering, Nanjing University, for his
modeling and mechanism of tetracycline on Pinus taeda-derived activated biochar.
help in the XPS spectrum analysis of this study. The authors also deeply Bioresour. Technol. 259, 24–31. https://doi.org/10.1016/j.biortech.2018.03.013.
appreciate the valuable comments from the editor and reviewers. This Jia, X.C., Zhou, J.W., Liu, J., Liu, P., Yu, L., Wen, B., Feng, Y., 2020. The antimony
sorption and transport mechanisms in removal experiment by Mn-coated biochar.
work was funded by the National Key Research and Development
Sci. Total Environ. 724, 138158. https://doi.org/10.1016/j.scitotenv.2020.138158.
Project (2020YFC1807701), the National Natural Science Foundation of Jiang, Y.J., Wei, X.D., He, H.P., She, J.Y., Liu, J., Fang, F., Zhang, W.H., Liu, Y.Y.,
China (No. U1612442) and the Project of Science and Technology Wang, J., Xiao, T.F., Tsang, D.C.W., 2022. Transformation and fate of thallium and
Department of Guizhou Province (RENCAI[2016]5664; ZHICHENG accompanying metal(loid)s in paddy soils and rice: a case study from a large-scale
industrial area in China. J. Hazard Mater. 423, 126997. https://doi.org/10.1016/j.
[2019]2838; JICHU[2020]1Z040). jhazmat.2021.126997.
Karim, A.A., Kumar, M., Ray, A., Hariprasad, D., Dhal, N.K., 2021. Biomass mediated
conversion of acidic phosphogypsum into alkaline material through thermal
Appendix A. Supplementary data
treatments. J. Sci. Ind. Res. 80, 924–928.
Li, J.Y., Zheng, B.H., He, Y.Z., Zhou, Y.Y., Chen, X., Ruan, S., Yang, Y., Dai, C.H., Tang, L.,
Supplementary data to this article can be found online at https://doi. 2018a. Antimony contamination, consequences and removal techniques: a review.
Ecotoxicol. Environ. Saf. 156, 125–134. https://doi.org/10.1016/j.
org/10.1016/j.envpol.2022.119032.
ecoenv.2018.03.024.
Li, L., Li, H.X., Liu, H., 2018c. Distribution and migration of antimony and other trace
References elements in a Karstic river system, Southwest China. Environ. Sci. Pollut. Res. 25
(28), 28061–28074. https://doi.org/10.1007/s11356-018-2837-x.
Li, L., Tu, H., Zhang, S., Wu, L., Wu, M., Tang, Y., Wu, P., 2019. Geochemical behaviors
Amen, R., Bashir, H., Bibi, I., Shaheen, S.M., Niazi, N.K., Shahid, M., Hussain, M.M.,
and environmental impact of antimony in mining-effected water environment
Antoniadis, V., Shakoor, M.B., Al-Solaimani, S.G., Wang, H.L., Bundschuh, J.,
(Southwest China). Environ. Geochem. Hlth. 41 (6), 2397–2411. https://doi.org/
Rinklebe, J., 2020. A critical review on arsenic removal from water using biochar-
10.1007/s10653-019-00285-8.
based sorbents: the significance of modification and redox reactions. Chem. Eng. J.
Li, R.H., Wang, J.J., Gaston, L.A., Zhou, B., Li, M., Xiao, R., Wang, Q., Zhang, Z.,
396 (15) https://doi.org/10.1016/j.cej.2020.125195, 125195.
Huang, H., Liang, W., Huang, H., Zhang, X., 2018b. An overview of carbothermal
Ashley, P., Craw, D., Graham, B., Chappell, D., 2003. Environmental mobility of
synthesis of metalebiochar composites for the removal of oxyanion contaminants
antimony around mesothermal stibnite deposits, New South Wales, Australia and
from aqueous solution. Carbon 129, 674–687. https://doi.org/10.1016/j.
southern New Zealand. J. Geochem. Explor. 77, 1–14. https://doi.org/10.1016/
carbon.2017.12.070.
S0375-6742(02)00251-0.
Li, X.Q., Elliott, D.W., Zhang, W.X., 2006. Zero-valent iron nanoparticles for abatement
Axe, L., Trivedi, P., 2002. Intraparticle surface diffusion of metal contaminants and their
of environmental pollutants: materials and engineering aspects. CRIT REV SOLID
attenuation in microporous amorphous Al, Fe, and Mn oxides. J. Colloid Interface
STATE 31 (4), 111–122. https://doi.org/10.1080/10408430601057611.
Sci. 247, 259–265. https://doi.org/10.1006/jcis.2001.8125.
Lian, G., Wang, B., Lee, X., Li, L., Liu, T., Lyu, W., 2019. Enhanced removal of hexavalent
Biver, M., Shotyk, W., 2012. Stibnite (Sb2S3) oxidative dissolution kinetics from pH 1 to
chromium by engineered biochar composite fabricated from phosphogypsum and
11. Geochem. Cosmochim. Acta 79, 127–139. https://doi.org/10.1016/j.
distillers grains. Sci. Total Environ. 697, 134119. https://doi.org/10.1016/j.
gca.2011.11.033.
scitotenv.2019.134119.
Biver, M., Shotyk, W., 2013. Stibiconite (Sb3O6OH), senarmontite (Sb2O3) and
Long, X.J., Wang, X., Guo, X.J., He, M.C., 2020. A review of removal technology for
valentinite (Sb2O3): dissolution rates at pH 2-11 and isoelectric points. Geochem.
antimony in aqueous solution. J. Environ. Sci. 90, 189–204. https://doi.org/
Cosmochim. Acta 109, 268–279. https://doi.org/10.1016/j.gca.2013.01.033.
10.1016/j.jes.2019.12.008.
Boily, J.-F., Gassman, P.L., Peretyazhko, T., Szanyi, J., Zachara, J.M., 2010. FTIR spectral
Majzlan, J., Lalinská, B., Chovan, M., Bläß, U., Brecht, B., Göttlicher, J., Steininger, R.,
components of schwertmannite. Environ. Sci. Technol. 44 (4), 1185–1190, 10.1021/
Hug, K., Ziegler, S., Gescher, J., 2011. A mineralogical, geochemical and micro-
es902803u.
biological assessment of the antimony- and arsenic-rich neutral mine drainage
Cardenas, G., Cabrera, G., Taboada, E., Miranda, S.P., 2004. Chitin characterization by
tailing near Pezinok, Slovakia. Am. Mineral. 96, 1–13. https://doi.org/10.2138/
SEM, FTIR, XRD, and C-13 cross polarization/mass angle spinning NMR. J. Appl.
am.2011.3556.
Polym. Sci. 93, 1876–1885. https://doi.org/10.1002/app.20647.
Marouane, B., Klug, M., As, K.S., Engel, J., Reichel, S., Janneck, E., Peiffer, S., 2021. The
Chen, L.C., Cui, W., Li, J.Y., Wang, H., Dong, X., Chen, P., Zhou, Y., Dong, F., et al., 2020.
potential of granulated schwertmannite adsorbents to remove oxyanions (SeO32-,
The high selectivity for benzoic acid formation on Ca2Sb2O7 enables efficient and
SeO42-, MoO42-, PO43-, Sb(OH)6-) from contaminated water. J. Geochem. Explor.
stable toluene mineralization. Appl. Catal. B: Environ. 271, 118948. https://doi.org/
223, 106708. https://doi.org/10.1016/j.gexplo.2020.106708.
10.1016/j.apcatb.2020.118948.
McComb, K.A., Craw, D., McQuillan, A.J., 2007. ATR-IR spectroscopic study of
Clogston, J.D., Patri, A.K., 2011. Zeta potential measurement. Methods Mol. Biol. 697,
antimonate adsorption to iron oxide. Langmuir 23 (24), 12125–12130. https://doi.
63–70. https://doi.org/10.1007/978-1-60327-198-1_6.
org/10.1021/la7012667.
Cornelis, G., Johnson, C.A., Gerven, T.V., Vandecasteele, C., 2008. Leaching mechanisms
Miao, Y.Y., Han, F.C., Pan, B.C., Niu, Y.J., Nie, G.Z., Lv, L., et al., 2014. Antimony(V)
of oxyanionic metalloid and metal species in alkaline solid wastes: a review. Appl.
removal from water by hydrated ferric oxides supported by calcite sand and
Geochem. 23, 955–976 https://doi:10.1016/j.apgeochem.2008.02.001.
polymeric anion exchanger. J. Environ. Sci. 26, 307–314. https://doi.org/10.1016/
Dong, S.X., Dou, X.M., Mohan, D., Pittman Jr., C.U., Luo, J.M., 2015. Synthesis of
S1001-0742(13)60418-0.
graphene oxide/schwertmannite nanocomposites and their application in Sb(V)
Nakamoto, K., 2009. Infrared and Raman Spectra of Inorganic and Coordination
adsorption from water. Chem. Eng. J. 270, 205–214. https://doi.org/10.1016/j.
Compounds, sixth ed. John Wiley, New York, p. 58.
cej.2015.01.071.
Nishad, P.A., Bhaskarapillai, A., 2021. Antimony, a pollutant of emerging concern: a
Erdemoglu, M., Sarikaya, M., 2006. Effects of heavy metals and oxalate on the zeta
review on industrial sources and remediation technologies. Chemosphere 277,
potential of magnetite. J. Colloid Interface Sci. 300, 795–804. https://doi.org/
130252. https://doi.org/10.1016/j.chemosphere.2021.130252.
10.1016/j.jcis.2006.04.004.
Nyirenda, T., Zhou, J.W., Mapoma, H., Xie, L., Li, Y., 2016. Hydrogeochemical
Filella, M., Belzileb, N., Chen, Y., 2002. Antimony in the environment: a review focused
characteristics of groundwater at the xikuangshan antimony mine in south China.
on natural waters II. Relevant solution chemistry. Earth Sci. Rev. 59, 265–285.
Mine Water Environ. 35, 86–93. https://doi.org/10.1007/s10230-015-0341-9.
https://doi.org/10.1016/S0012-8252(02)00089-2.
Oliveira, F.R., Patela, A.K., Jaisib, D.P., Adhikaric, S., Lu, H., Khanal, S.K., 2017.
Filella, M., He, M.C., 2020. Foreword to the special issue on ‘antimony in the
Environmental application of biochar: current status and perspectives. Bioresour.
environment: a Chinese perspective. Environ. Chem. 17, 303. https://doi.org/
Technol. 246, 110–122. https://doi.org/10.1016/j.biortech.2017.08.122.
10.1071/ENv17n4_FO.
7
L. Li et al. Environmental Pollution 301 (2022) 119032
Rinklebe, J., Shaheen, S.M., El-Naggar, A., Wang, H.L., Laing, G.D., Alessi, D.S., Ok, Y.S., Wang, L., Wang, J.Y., Wang, Z.X., He, C., Lyu, W., Yan, W., Yang, L., 2018. Enhanced
2020. Redox-induced mobilization of Ag, Sb, Sn, and Tl in the dissolved, colloidal antimonate (Sb(V)) removal from aqueous solution by La-doped magnetic biochars.
and solid phase of a biochar-treated and un-treated mining soil. Environ. Int. 140, Chem. Eng. J. 354, 623–632. https://doi.org/10.1016/j.cej.2018.08.074.
105754. https://doi.org/10.1016/j.envint.2020.105754. Wang, X.Q., He, M.C., Xi, J.H., Lu, X.F., 2011. Antimony distribution and mobility in
She, J.Y., Wang, J., Wei, X.D., Zhang, Q., Xie, Z.Y., Beiyuan, J.Z., Xiao, E.Z., Yang, X., rivers around the world’s largest antimony mine of Xikuangshan, Hunan Province,
Liu, J., Zhou, Y.T., Xiao, T.F., Wang, Y.X., Chen, N., Tsang, D.C.W., 2021. Survival China. Microchem. J. 97 (1), 4–11. https://doi.org/10.1016/j.microc.2010.05.011.
strategies and dominant phylotypes of maize-rhizosphere microorganisms under Wei, X.D., Zhou, Y.T., Tsang, D.C.W., Song, L., Zhang, C.S., Yin, M.L., Liu, J., Xiao, T.F.,
metal(loid)s contamination. Sci. Total Environ. 774, 145143. https://doi.org/ Zhang, G.S., Wang, J., 2020. Hyperaccumulation and transport mechanism of
10.1016/j.scitotenv.2021.145143. thallium and arsenic in brake ferns (Pteris vittata L.): a case study from mining area.
Shi, L.D., Wang, Z., Liu, T., Wu, M.X., Lai, C.Y., Rittmann, B.E., Guo, J.H., Zhao, H.P., J. Hazard Mater. 388, 121756. https://doi.org/10.1016/j.jhazmat.2019.121756.
2021. Making good use of methane to remove oxidized contaminants from Wu, D.B., Purnomo, B.J., Sun, S.P., 2017. As and Sb speciation in relation with physico-
wastewater. Water Res. 197, 117082. https://doi.org/10.1016/j. chemical characteristics of hydrothermal waters in Java and Bali. J. Geochem.
watres.2021.117082. Explor. 173, 85–91. https://doi.org/10.1016/j.gexplo.2016.12.003.
Song, Z.L., Zhang, M.Y., Ma, C.Y., 2007. Study on the oxidation of calcium sulfide using Yan, L., Chan, T., Jing, C., 2020. Mechanistic study for stibnite oxidative dissolution and
TGA and FTIR. Fuel Process. Technol. 88 (6), 569–575. https://doi.org/10.1016/j. sequestration on pyrite. Environ. Pollut. 262, 114309. https://doi.org/10.1016/j.
fuproc.2007.01.014. envpol.2020.114309.
USGS, 2019. U.S. Geological Survey, Mineral Commodity Summaries. https://prd-wret. Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X.D., Pullammanappallil, P.,
s3-us-west-2.amazonaws.com/assets/palladium/production/s3fs-public/atoms/files Yang, L.Y., 2011a. Removal of phosphate from aqueous solution by biochar derived
/mcs-2019-antim.pdf. February 2019. from anaerobically digested sugar beet tailings. J. Hazard Mater. 190 (1–3),
Viallis-Terrisse, H., Nonat, A., Petit, J.-C., 2001. Zeta-potential study of calcium silicate 501–507. https://doi.org/10.1016/j.jhazmat.2011.03.083.
hydrates interacting with alkaline cations. J. Colloid Interface Sci. 244, 58–65. Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X.D., Pullammanappallil, P.,
https://doi.org/10.1006/jcis.2001.7897. Yang, L.Y., 2011b. Biochar derived from anaerobically digested sugar beet tailings:
Wang, B., Gao, B., Fang, J., 2017. Recent advances in engineered biochar productions characterization and phosphate removal potential. Bioresour. Technol. 102 (10),
and applications. Crit. Rev. Environ. Sci. Technol. 47 (22), 2158–2207. https://doi. 6273–6278. https://doi.org/10.1016/j.biortech.2011.03.006.
org/10.1080/10643389.2017.1418580. Zhang, M., Gao, B., Yao, Y., Xue, Y., Inyang, M., 2012. Synthesis of porous MgO-biochar
Wang, B., Lian, G., Lee, X., Gao, B., Li, L., Liu, T., Zhang, X., Zheng, Y., 2020. nanocomposites for removal of phosphate and nitrate from aqueous solutions. Chem.
Phosphogypsum as a novel modifier for distillers grains biochar removal of Eng. J. 210, 26–32. https://doi.org/10.1016/j.cej.2012.08.052.
phosphate from water. Chemosphere 238, 124684. https://doi.org/10.1016/j. Zhou, J.W., Nyirenda, T., Xie, L., Li, Y., Zhou, B., Zhu, Y., Liu, H., 2017. Mine waste
chemosphere.2019.124684. acidic potential and distribution of antimony and arsenic in waters of the
Wang, L., Wang, J.Y., Wang, Z.X., Feng, J.T., Li, S.S., Yan, W., 2019. Synthesis of Ce- Xikuangshan mine. China. Appl. Geochemistry. 77, 52–61. https://doi.org/10.1016/
doped magnetic biochar for effective Sb(V) removal: performance and mechanism. j.apgeochem.2016.04.010.
Powder Technol. 345, 501–508. https://doi.org/10.1016/j.powtec.2019.01.022.