You are on page 1of 12

Environmental Pollution 317 (2023) 120723

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Adsorption and immobilization performance of pine-cone pristine and


engineered biochars for antimony in aqueous solution and military
shooting range soil: An integrated novel approach
Basit Ahmed Khan a, c, Mahtab Ahmad a, *, Sajid Iqbal b, Fath Ullah a, Nanthi Bolan c,
Zakaria M. Solaiman c, Munib Ahmed Shafique d, Kadambot H.M. Siddique c
a
Department of Environmental Sciences, Faculty of Biological Sciences, Quaid-i-Azam University, Islamabad, 45320, Pakistan
b
Analytical Chemistry Group, Chemistry Division, Pakistan Institute of Nuclear Science and Technology (PINSTECH), P.O. Nilore, Islamabad, Pakistan
c
UWA School of Agriculture and Environment, and the UWA Institute of Agriculture, The University of Western Australia, Perth, WA, 6009, Australia
d
Central Analytical Facility Division, Pakistan Institute of Nuclear Science and Technology (PINSTECH), P.O. Nilore, Islamabad, Pakistan

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

Keywords: Antimony (Sb–V), a carcinogenic metalloid, is becoming prevalent in water and soil due to anthropogenic ac­
Remediation tivities. Biochar could be an effective remedy for Sb(V)-contaminated water and soil. In this study, we used
Sorption pristine and engineered pinecone-derived biochar as an innovative approach for treating Sb(V)-contaminated
Water treatment
water and shooting range soil. Biochar was produced from pine–cone waste (pristine biochar) and enriched
Army firing range
Stabilization
with Fe and Al salts via saturation (engineered biochar). Adsorption tests in water revealed that iron-modified
Biochar modification biochar showed higher adsorption capacity (8.68 mg g− 1) than that of the pristine biochar (2.49 mg g− 1) and
aluminum-modified biochar (3.40 mg g− 1). Isotherm and kinetic modeling of the adsorption data suggested that
the adsorption process varied from monolayer to multilayer, with chemisorption as the dominant interaction
mechanism between Sb(V) and the biochars. The post-adsorption study of iron-modified biochar by Fourier
Transform Infrared (FTIR) and X-ray diffraction (XRD) further supported the chemical bonding and outer-sphere
complexation of Sb(V) with Fe, N–H, O–H, C–O and C– –C components. The pristine and iron-modified biochars
also successfully immobilized Sb(V) in a shooting range soil, more so in the latter. Subsequent sequential ex­
tractions and post-analysis by scanning electron microscopy (SEM), energy dispersive X-ray analysis (EDX), and
elemental dot mapping revealed that Sb in the treated soil transformed to a more stable form. It was concluded
that iron-modified biochar could act as an efficient material for the adsorption and immobilization of Sb(V) in
water and soil, respectively.

1. Introduction Bolivia (Bolan et al., 2022).


Leaching from mine wastes, weathering of sulfide ores, and man-
An increasing number of industrially introduced anionic pollutants made activities (smelting, industry effluents, metallurgy processes,
that pose severe environmental and public health threats have con­ and military shooting discharges) cause contamination of soil and
cerned people worldwide (Lito et al., 2012; Xu et al., 2013). Frequent aquatic ecosystems with Sb (Diquattro et al., 2020). Moreover, Sb is
anionic species including arsenate and arsenite, chromate and chromite, generated primarily from pharmaceuticals (Zhuang et al., 2018), in­
and antimonate and antimonite, are some of the most harmful pollutants dustrial and dumps from mining waste (Dupont et al., 2016), wastewater
in most water bodies and wastewater (Xu et al., 2013). Antimony (Sb), a sewage sludge (Campos et al., 2019), automobile emissions (Fort et al.,
metalloid with a 1.9 million ton worldwide reserve, is one of the most 2016), seepage from plastic waste (Chu et al., 2021; Hu et al., 2020),
frequently used metals (Bolan et al., 2022). More than 80% of Sb-related leachate directly draining from construction sites that dump solid waste
materials, including toys, clothing for fire-fighters, vehicle seat covers, or electronic waste (Intrakamhaeng et al., 2020), industrial spills, wood
paint, batteries, and metal alloys, are produced in China, Russia, and preservation, and metal processing (Herath et al., 2017).

* Corresponding author.
E-mail address: mahmad@qau.edu.pk (M. Ahmad).

https://doi.org/10.1016/j.envpol.2022.120723
Received 30 August 2022; Received in revised form 20 November 2022; Accepted 21 November 2022
Available online 24 November 2022
0269-7491/© 2022 Elsevier Ltd. All rights reserved.
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

The species of Sb that is most common in the environment is Sb(V). carbonized at 600 ◦ C under the low oxygen conditions at 5 ◦ C min− 1
The Sb(V) species mostly exists in Sb(OH)-6 state under aerobic condi­ heating rate, to produce pristine biochar.
tions across a wide range of pH, and it is more mobile under the alkaline Engineered biochars were produced using the impregnation modi­
and neutral conditions. Sb(V) is also present in natural waters as oxy­ fication method reported by Khan et al. (2022). The pristine biochar was
anions and Sb(OH)-6, and it is apparently very stable under wide range of then soaked in a 3 M mineral salt solution (FeCl3 or AlCl3) in a 1:10
pH (Zhu et al., 2021). Ashley et al. (2003) pointed out that stibnite (w/v) ratio, agitated for 6 h at room temperature. The suspension was
(Sb2S3) would undergo a series of chemical reactions such as oxidation centrifuged at 3000 rpm, and rinsed thoroughly using deionized water to
and leaching under the joint action of oxygen, water, and microorgan­ remove excess amounts of salts, and dried at 80 ◦ C in an oven for 24 h.
isms that eventually generates soluble SbO−3 in the wastewater. The engineered or modified biochars were named Fe-biochar and
Both forms of Sb(III) or Sb(V)-derived complexes are very dangerous Al-biochar.
oncogenic contaminants; if exposed, animals and humans may experi­
ence adverse health effects via dermal and oral uptake, and inhalation 2.2. Biochar characterization
(Bagherifam et al., 2019; Wu et al., 2011). Sb(V) can enter the human
body through skin contact and tainted food, disrupting the viability, Proximate, chemical, and ultimate analysis of all biochars were un­
synthesis, and release of metallothionein proteins. Humans who dertaken. The chemical analysis for biochars includes pH, EC, and point
consume too much Sb(V) are at risk of respiratory problems, rashes, of zero charges (pHPZC) whereas moisture, mobile/volatile matter, ash
diarrhea, and nausea (Hua et al., 2021). content, and resident/fixed carbon content were measured in the
The oxides of aluminum (Al), manganese (Mn), and iron (Fe) can proximate analysis (Ahmad et al., 2012b; Akhtar et al., 2021). A Nikolet
naturally adsorb Sb discharged into soils; however, the soluble antim­ iS50 FTIR spectrometer (USA) and D/Max-B X-ray diffractometer
onate is found mainly under alkaline and oxidized states as Sb(V) (Ji (Japan) were utilized to identify the functional groups and determine
et al., 2017). The soil’s metalloid mobilization is influenced by several the crystallography of biochars. A CHNSO (carbon, hydrogen, nitrogen,
factors, such as precipitation of minerals, ion exchange, adsorption/­ sulphur, oxygen) elemental analyzer (Flash 2000 CHNSO analyzer, USA)
desorption, complexation, pH, electrical conductivity, leaching, phy­ was used for the ultimate analysis.
tostabilization, and biotransformation (El-Naggar et al., 2018;
Palansooriya et al., 2020). The average Sb concentration in soil is 1 mg 2.3. Soil collection, processing, and characterization
kg− 1 (Johnson et al., 2005). However, increased Sb levels in soil samples
have been reported (up to 4400 mg kg− 1), and pore space water samples A military shooting range site in the Kharian, district Gujrat of
(up to 1700 mg L− 1) from an abandoned mine (Cidu et al., 2014). The Pakistan, was selected to collect soil samples. The top 15–20 cm of the
shooting range soils are often examined to be contaminated with Sb soil was taken from several locations within a 0.5 km radius of the
because it makes up 2–8% of the weight of bullets (He et al., 2019). For shooting range. The collected samples were mixed and homogenized in a
instance, in Switzerland, the Sb concentration in top layer firing range polyethylene bag to produce a composite sample. The soil samples after
soil ranged from 35 to 17,500 mg kg− 1 (Johnson et al., 2005). A military air drying were passed through 2 mm sieve to remove pebble, gravels,
shooting range field in Canada had an Sb concentration of 570 mg kg− 1 and bullet fragments to ensure that only homogeneous soil particles
(Laporte-Saumure et al., 2011), while a shooting range in Norway had were obtained. The acid digestion method (Ahmad et al., 2012a) was
830 mg kg− 1 Sb in the topsoil (Mariussen et al., 2017). used to analyze the total Sb concentration in the soil samples using an
Sb-containing wastewater is often treated via adsorption, electro­ inductively coupled plasma (ICP) spectrophotometer (iCAP 6500 of
chemical, biological, and co-precipitation/coagulation processes (Cao Thermo Fisher Scientific UK). The soil was characterized for basic pa­
et al., 2019). Adsorption technique is appealing because it is efficient, rameters (pH, EC, texture, organic matter, and water holding capacity)
low cost, and easy to use as compared to other techniques (Khan et al., following the methods reported in Mun-Gai & Owido, (2021). The
2022; Murad et al., 2022). Among various adsorbents, carbon-based supplementary material includes the information about soil properties
materials like biochar have recently gained attention due to its low (Table S2). The shooting range soil with a pH of 8.68 and silty clay
cost and easy accessibility. However, the efficiency of pristine biochar, texture contained 28 mg kg− 1 total Sb.
particularly for anionic contaminants, remains controversial, mainly
due to its negatively charged surfaces (Amin et al., 2020). Therefore, 2.4. Adsorption experiments
engineering or modifying the pristine biochar is needed to improve its
adsorption efficiency for anionic contaminants such as As and Sb. In this The adsorption experiments were carried out to study the effects of
context, impregnating pristine biochar with iron and aluminum mineral adsorption variables including, solution pH, adsorbent dose, initial
salts could be advantageous. The cationic Al and Fe elements can coat concentration, and total contact time. To determine pH’s impact on the
the biochar surface converting it from a negatively charged to a posi­ adsorption process, four different solutions of Sb(V) at 10 mg L− 1 con­
tively charged surface, thereby enhancing the electrostatic attraction of centration whose pHs were adjusted from 3 to 9, separately. All exper­
anionic contaminants. Furthermore, a variety of iron oxides can strongly iments were conducted over 24 h and the biochar dose was 1 g L− 1. The
adsorb Sb(III) and Sb(V), significantly impacting their mobility, speci­ initial concentration effect of Sb(V) on biochars’ adsorption potential
ation, and environmental fate (Bolan et al., 2022). was measured in seven concentration ranges, i.e., 0.5, 1.0, 2.0, 5.0, 10,
This study aims to utilize waste pine-cones to produce pristine bio­ 20, and 40 mg L− 1. For each biochar a dosage of 1 g L− 1 was taken and
char modified with Fe and Al salts to produce engineered biochars which pH of the solutions was set at 7 for a contact period of 24 h. In order to
are then tested for their adsorption potential in Sb(V)-contaminated determine the biochars’ adsorption capacity for contact time experi­
water and immobilization efficiency in Sb-contaminated shooting ment, 1 g L− 1 dose of each biochar was treated with a 10 mg L− 1 solution
range soil. Mechanisms of Sb(V) adsorption in water and immobilization of Sb(V) at pH 7. The mixtures were left in contact for 15 min to 24 h and
in soil with pristine and engineered biochars are also explored. analyzed solutions at regular intervals. Each experiment was performed
at 120 rpm in an orbital shaker, and filtration was made by using
2. Materials and methods Whatman 42 filter paper. The concentration of Sb(V) was analyzed by
using ICP spectrophotometer.
2.1. Production of pristine and engineered biochars
2.5. Adsorption modeling
Waste pine-cones were converted thermally to biochar via slow py­
rolysis in a muffle furnace (Khan et al., 2022). The ground feedstock was The Sb(V) amount adsorbed by biochars was calculated by using the

2
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

equation below: elements such as O, N, Si, S, Pb, Sb, and Fe on the soil particles.
V
Adsorbed amount = (Cₒ − Cₑ) × 3. Results and discussion
W

where Co is the initial concentration, Ce represents the final Sb(V) con­ 3.1. Characteristics of prepared biochars
centration (mg L− 1), V is considered as volume of the solution (L), and W
represents biochar mass (g). The properties of pristine and engineered biochars are shown in
The four commonly used equilibrium isotherm models (Langmuir, Table S1. The engineered biochars (Fe- and Al-biochars) had compara­
Freundlich, Temkin, and Dubinin-Radushkevich) were employed to tively higher mobile/volatile matter than that of the pristine biochar due
determine the adsorption process of all biochars. Similarly, four non- to the volatile organic matter’s strong affinity with FeCl3 and AlCl3
linear kinetics models (pseudo-second-order, intraparticle diffusion, (Beesley and Marmiroli, 2011). The engineered biochars had acidic pH
Elovich, and power function) were also used to further investigate the values, i.e. 2.33 and 4.35 for Fe- and Al-biochar, respectively due to the
adsorption mechanisms. The equations of various isotherm and kinetic FeCl3 and AlCl3 impregnation (Shetty & Prakash, 2020). The ultimate
models are given in Table S3 in the supplementary material. elemental analysis revealed that the Fe-biochar and pristine biochar had
higher amounts of total C (81.60% and 80.65%, respectively) than the
Al-biochar (59.03%). The low C content of Al-biochar may be due to
2.6. Soil incubation experiments
enrichment with AlCl3, resulting in reduction of the inorganic C content
after the engineering of pristine biochar (Hafeez et al., 2022). The
Pristine and iron-modified biochars were selected to remediate the
engineered biochars had higher H concentrations than the pristine
Sb-contaminated shooting range soil. A soil incubation experiment
biochar, indicating the supremacy of the H-containing components due
included three applications of each biochar (1%, 2.5%, and 5%; w/w),
to their acidic (<7) pH values. Similarly, the engineered biochars had
with 500 g soil placed in an air-tight polyethylene container (600 mL of
slightly higher total N contents than pristine biochar, possibly because of
capacity) and mixed well with each biochar at the specified application
their relative drop in the mass intensity after impregnation. The
rate. The soil was moistened with 47 mL, or 50% of soil’s water holding
Al-biochar had significantly higher O content than the other biochars,
capacity by using de-ionized water. The containers were tightly covered,
indicating the prevalence of O-containing functional groups on its
and the incubation experiment took place for 60 days in the dark at room
surface.
temperature. The containers were checked periodically for weights, with
any weight loss was compensated by adding the desired amount of water
3.2. SEM, EDX, FTIR and XRD descriptions
to keep the moisture contents constant.

The SEM images revealed the morphological features of the biochars


2.7. Soil extraction for Sb immobilization (Fig. S1). All biochars had distinctive porous and tubular structures. The
presence of Fe (5.72%) and Al (7.30%) was detected in the EDX spectra
After incubation, all the samples of soil from each treatment were air- of the Fe-biochar and Al-biochar, respectively. The few other common
dried, and Sb availability was measured using a variety of single elements (C, O, and Cl) were also observed. Considering pristine bio­
extraction and sequential extraction techniques to analyze the soil sta­ char, primary elements (C, O, N, K, and S) were also present.
bilization effect of the used biochars on Sb immobilization. The water- The FTIR spectra for the surface functioning of pristine biochar, Fe-
soluble fraction of Sb was determined by extracting 2 g of incubated biochar, and Al-biochar are shown in Fig. 1a. All biochars had O–H
soil by using 50 mL deionized water for 2 h. In order to determine the stretching developed in broad band at 3372 cm− 1 due to the vibration
exchangeable amounts of Sb, the soil was extracted by using 50 mL of 1 effect of alcohols, carboxylic acid, and phenols in the cellulose and
M NH4NO3 solution for 2 h. The impact of the biochar treatments on Sb lignin. Moreover, the bending vibration of O–H in pristine biochar was
leachability was assessed using the TCLP (toxicity characteristics attributed to the peak at 1036 cm− 1 (Hongtao et al., 2018; Jia et al.,
leaching procedure) method. Briefly, the soil sample was extracted in a 2020; Rahman et al., 2021b); the same peak shifted in the Fe-biochar
solution of glacial acetic acid (pH 4.93) at a ratio of 1:20 (soil/extrac­ and Al-biochar to 996 and 1026 cm− 1, respectively, due to the
tant) after shaking at 30 rpm for 18 h. A physiological-based extraction impregnation (Fan et al., 2018). For pristine biochar, the peak at 863
test (PBET) was conducted to examine the biochar efficiency for cm− 1 was associated with the P–O–C aliphatic group; however, this peak
decreasing Sb bioavailability in human bodies. Briefly, 0.35 g of soil shifted in the Al- and Fe-biochars to 790 cm− 1 and 873 cm− 1, respec­
sample was shaken at 30 rpm with synthetic gastric solution (35 mL) for tively (Amin et al., 2020), possibly due to the significant oscillation in
1 h at 37 ◦ C. The gastric solution composed of 0.4 M glycine adjusted to inorganic Fe and Al salts. Only the Fe- and Al-biochars had distinct peaks
pH 2.2. The above mentioned single extraction tests followed the at 584 and 515 cm− 1, reflecting Al–O and Fe–O complexes due to the
methods described in Alaboudi et al. (2019), Ahmad et al. (2012a) and impregnation with salts (Karimipour et al., 2019; Din et al., 2021; Le
USEPA (1992). Bozec et al., 2002).
For the sequential extraction method, five component fractions were The surface minerology and crystallinity of biochars were also
measured in a series by using the method of Tessier et al. (1979): (i) examined using XRD spectroscopy (Fig. 1b). All biochars were amor­
non-specifically sorbed considering the readily soluble form, (ii) spe­ phous in nature due their organic feedstock. The spectra of all biochars
cifically sorbed form, (iii) interconnected with amorphous and imper­ peaked at two theta degrees of 22.6, denoting existence of crystalline
fectly crystalline ferrous and aluminum hydrous oxides, (iv) linked with cellulose (Zhu et al., 2016; Poletto et al., 2014). The peak at 26◦ in all
the well-crystallized Fe- and the Al-hydrous oxides, and (v) residual soil biochars indicated an inorganic component mainly comprising quartz.
fraction. Furthermore, the peak at two theta degrees of 40 denoted the presence
of Al2O3 (Du et al., 2009) in the XRD pattern of the Al-biochar. In the
2.8. SEM, EDX, and elemental dot mapping of soil Fe-biochar, magnetite (Fe3O4) peaks were visible at two theta degrees of
35.6, 43.8, and 57.6 (Fan et al., 2018).
Scanning electron microscope (SEM) and the elemental dot mapping
were also performed by using SEM apparatus (FESEM, MAIA-3, 3.3. Adsorption of antimony onto biochars and isotherm modeling
TESEAN, USA) equipped with the energy dispersive X-ray (EDX) spec­
troscopy systems to examine morphological and elemental changes in The pH significantly affected biochar effectiveness in absorbing Sb
the soil. X-ray dot maps were then used to highlight the distribution of (V) and the adsorption process. The solution pH was set to 3, 5, 7, or 9

3
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Fig. 1. (a) FTIR and (b) XRD spectra of pristine biochar, Fe-biochar, Al-biochar, sand post-adsorbed PFe-biochar.

for adsorption studies. The Sb(V) adsorption for engineered biochars contributed to the electrostatic attraction of Sb(V) ions to the biochar
changed considerably with changing solution pH (Fig. 2a). The Fe- surface. The Fe- and Al-biochars had positively charged surfaces at pH 3
biochar had a higher Sb adsorption capacity at pH 7 (2.26 mg g− 1) and negatively charged at pH > 5, with pHPZC values of 4.98 and 4.89,
than the Al-biochar (1.66 mg g− 1) and the pristine biochar (0.39 mg respectively. The adsorbent surface had more negative charges at pH >
g− 1). The Qe of the engineered biochars increased slightly at pH 5 and 7, 5, making them unsuitable for binding Sb(V). The OH− content in the
most likely due to competitive binding sites between Sb(OH)6 and H+ solution increased as pH increased and had a stronger affinity for
(Rahman et al., 2021a). The pHPZC is a significant marker for the surface sorption than Sb(OH)–6, however, competition between OH− and Sb−
charge at various solution pHs. For solution pH ranging from 3 to 9 ions in the aqueous solution decreased the capacity for Sb(V) adsorption
(pHPZC 7.58), the pristine biochar had a positive charge, which may have (Jia et al., 2020; Khan et al., 2022) (see Fig. 3).

Fig. 2. Effect of (a) pH, (b) dose, (c) initial concentration, and (d) contact time on Sb adsorption onto pristine biochar, Fe-biochar, and Al-biochar. Experiments were
performed at an initial concentration of 10 mg L− 1 (except for Fig. (c)), a dosage of 1 g L− 1 (except for Fig. (b)), pH of 7 (except for Fig. (a)), and a contact time of 24 h
(except for Fig. (d)) for a volume of 40 mL.

4
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Fig. 3. Fittings of (a) Langmuir, (b) Freundlich, (c) Temkin, and (d) Dubinin Radushkevich isotherm models to the experimental data of Sb adsorption onto pristine
biochar, Fe-biochar, and Al-biochar. Solid lines indicate the respective models fittings.

Fig. 2b represents the effect of doses on the adsorption capacity of (0.821), and D-R models (0.869); the Fe-biochar data fitted well to the
three biochars. It was observed that adsorption capacity for all biochars Langmuir (0.841), Freundlich (0.826), and D-R (0.819) models, while
decreased as the adsorbent dose increased. The adsorption capacity for the Al-biochar data only fitted well to the D-R isotherm model (0.869).
low dose was higher because of the availability of more active sites to The Langmuir model represents monolayer adsorption on homogeneous
bind the Sb(V) in the solution. However, Fe-biochar showed higher surfaces, while the Freundlich, Temkin and D-R models represents het­
adsorption capacity than pristine biochar and Al-biochar because of erogeneous adsorption (Amin et al., 2021; Liu et al., 2010), indicating
stronger affinity with Sb(V) due to the presence of Fe-functional groups the overall heterogeneity of the biochar surfaces. The Fe-biochar had a
and more active sites on its surface (Zhang et al., 2019). Fig. 2c illus­ greater KF (Freundlich predicted adsorption capacity) value (0.706 mg
trates how the initial concentration affects the ability of various biochars g− 1) than the other two biochars (0.609 and 0.350 mg g− 1). Similarly,
to adsorb Sb(V). Among the three biochars, the Fe-biochar had the the Fe-biochar had a greater QD (D-R predicted adsorption capacity)
highest adsorption capacity (3.95 mg g− 1) at concentration of 40 mg value (3.574 mg g− 1) than the pristine biochar (2.28 mg g− 1) and
L− 1. Because Sb(V) ions cover active sites on the surface of biochar and Al-biochar (2.854 mg g− 1). According to the Langmuir model, the
interact with adsorbed functional groups, and all biochars can adsorb Fe-biochar had a higher adsorption capacity (Qmax = 4.825 mg g− 1) than
more Sb(V) as the initial concentration increases (Liu, 2015). pristine (2.681 mg g− 1) and Al-biochar (4.247 mg g− 1). These results
Increasing the contact time generally increases biochars’ Sb(V) demonstrate that Fe-biochar favors Sb(V) adsorption in water. .
adsorption (Fig. 2d). However, within a relatively short period, the
number of active sites on biochar declines, thereby decreasing Sb(V)
3.4. Adsorption kinetics
adsorption. Sb(V) began to bind quickly to the biochars within 15 min,
resulting in 5.20 mg g− 1 adsorption for Fe-biochar, 2.52 mg g− 1 for Al-
Adsorption kinetics was used to assess the effectiveness of pristine
biochar, and 2.41 mg g− 1 for pristine biochar, and all biochars had
and engineered biochars at adsorbing Sb(V). In the first 5 h, all biochars
reached equilibrium by 5 h. After 24 h the Fe-biochar has the highest
significantly increased their ability to adsorb Sb(V) (Fig. 2d), which was
adsorption capacity (8.68 mg g− 1), followed by Al-biochar (3.40 mg g− 1)
related to adsorption by the high-affinity sites or the precipitation of
and pristine biochar (2.49 mg g− 1). All biochars reacted immediately at
adsorbate on the adsorbent surface. Table 1b presents the constant pa­
the start of the agitation process because the surface of biochars has
rameters for the different kinetic models. The pseudo-second-order R2
more active sites, which initially led to increased adsorption.
and χ2 values for various biochars ranged from 0.755 to 1 and
Four isotherm models (Langmuir, Freundlich, Temkin, and Dubinin-
0.008–0.175. Pristine biochar had a higher rate constant (k) (0.067 mg
Radushkevich (D-R)) were compared to check the fitness of Sb(V)
g− 1 min− 1) than the other two biochars (0.011 and 0.037 mg g− 1
adsorption data for the different biochars. Table 1a shows the equilib­
min− 1), while Fe-biochar had the highest adsorption capacity (8.534 mg
rium isotherm models calculated parameters. The pristine biochar
g− 1) than pristine biochar (5.211 mg g− 1) and Al-biochar (2.381 mg
adsorption data fitted well to the Langmuir (R2 = 0.841), Temkin
g− 1), respectively. In the Elovich model, the pristine biochar, Fe-

5
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

biochar, and Al-biochar had R2 values of 0.993, 0.971, and 0.872,


respectively, and χ2 values of 0.004, 0.185, and 0.091, respectively.

0.065
0.086
0.150
E (kJ mol− 1)
Experimental data for the intraparticle model did not fit well, while the

V
8.839
7.180
5.315
kinetic experimental data fit well to the power function model (R2 >
0.90 and χ2 > 0.01). The Fe-biochar had a higher power function

K (mg g− 1)
adsorption capacity (4.83 mg g− 1) than the pristine biochar (1.68 mg

0.943
1.68
4.83
g− 1) and Al-biochar (0.943 mg g− 1). The fitting of these models clarifies
QD (mg g− 1)
the chemical interaction between the biochars and Sb(V) that could

2.281
3.574
2.854
trigger a physicochemical sorption or adsorption process via valence

0.005
0.233
0.073
Power function
forces by sharing or electron transfer between the sorbent and sorbate

χ2
Dubinin Radushkevich

(chemisorption) (Chen et al., 2019; Wu et al., 2019). Furthermore, all


models predicted that Fe-biochar significantly outperformed the pristine
1.204
3.832
3.103

0.992
0.967
0.900
and Al-biochars in terms of Sb(V) adsorption.
χ2

R2 Interestingly, several adsorption mechanisms can be implicated in


1.520 the removal of Sb(V) from water due to the biochar’s multifunctional
4.670
1.019
0.869
0.819
0.869

properties. Monolayer-to-multilayer, surface complexation, and ion-


R2

exchange adsorption of Sb onto the biochar surface may be the domi­


nant adsorption mechanisms in this study(Fig. 4; Table 1).
kd (mg g− 1)
A (L g− 1)

The Fe-biochar showed highest adsorption capacity (8.68 mg g− 1)


1.917
3.034
1.484

0.042
0.149
0.058

towards Sb(V) removal in this study, which is greater than some other
reported engineered biochars. For example, Jia et al. (2020) used
Intra-particle diffusion

Mn-coated biochar and reported only 0.94 and 0.73 mg g− 1 of Sb(III)


B (J mol–)

0.364
3.760
0.194

and Sb(V) adsorption, respectively. Likewise, Wang et al. (2018) re­


χ2
926.0
848.8
609.9

ported 4.85 mg g− 1 Sb(V) adsorption onto magnetic biochar. However, a


greater adsorption of Sb(V) is observed in some studies using engineered
0.434
0.476
0.730

Fe-biochars. For instance, Wang et al. (2018) further modified the


R2
Parameters calculated by (a) isothermal, and (b) kinetic models for Sb adsorption onto pristine biochar, Fe-biochar, and Al-biochar.

0.004
0.185
0.091

magnetic biochar by doping with La that improved the Sb(V) adsorption


χ2

to 18.92 mg g− 1. Similarly, Rahman et al. (2021a) reported 98.04 mg


min− 1)

g− 1 Sb(V) adsorption using Zr–Fe-biochar. This comparison with other


Temkin

studies reveal that adsorption efficiency of Fe-biochar could be


0.821
0.763
0.688

1
R2

enhanced by doping or coating with additional transition metals.


β (mg g−

6.339

However, this will add to the cost of the adsorbent.


1.46
3.60
KF (mg g− 1)

3.5. Adsorption mechanism(s) of biochar interactions with Sb(V) in water


min− 1)
0.609
0.706
0.350

The adsorption experiments revealed that Fe-biochar is more effec­


3.48 × 1046
2.71 × 1039
1

3.1 × 105

tive than pristine biochar and Al-biochar, which might be driven by


α (mg g−

modification during FeCl3 engineering. For example, the Fe-biochar had


0.391
0.488
0.597

iron-associated and positively charged functional groups on its surface


n

(FTIR; Fig. 1a) that could be involved in increased adsorption. To sup­


0.004
0.185
0.091

port this finding, FTIR and XRD analysis of Fe-biochar after adsorption
0.005
0.262
0.082

χ2

were carried out. The peak in PFe-biochar shifted from 1026 to 1136
χ2

cm− 1 with increasing amounts of iron oxide adsorbed via the inner-
Freundlich

Elovich

0.993
0.971
0.872

sphere complex and the presence of hydroxyl groups from adsorbed


0.686
0.826
0.748

R2

antimonate. This suggests that ion exchange or electrostatic attraction


R2

mechanism play a vital role in Fe-biochar adsorbing Sb(V) (McComb


min− 1)

et al., 2007). All biochars exhibited a band at 2356 cm− 1 indicating the
KL (L g− 1)

C–– C stretching vibrations in alkyne groups and the affinity of Sb(V)



0.191
0.103
0.058

with alkynes. Specifically, after Sb(V) adsorption onto Fe-biochar, the


K (g mg−

C–– C peak intensity increased and its position changed (Amin et al.,
0.067
0.011
0.037


2021; Ayoob et al., 2008; Saka, 2012). After adsorption the N–H peak
shifted from 1577 cm− 1 to 1599 cm− 1, the O–H peak shifted from 1026
Qmax (mg g− 1)

cm− 1 to 1136 cm− 1, and the Fe–O peak shifted from 584 cm− 1 to 587
qe (mg g− 1)

cm− 1 in the PFe-biochar. These spectral alterations demonstrate in­


2.681
4.825
4.247

5.211
8.534
2.381

teractions between the Sb(V) ion and Fe–O–OH, N–H, and O–H surface
functional groups in the Fe-biochar surface. The Fe–O–Sb bands were
Pseudo-second-order

attributed to the spectroscopic peak at 587.8 cm− 1, showing that the


5.70
0.17
0.18

0.008
0.018
0.175

iron oxides on the surface of the Fe-biochar adsorbed Sb(V) (Wang et al.,
χ2

χ2

2020; Zhang et al., 2019). According to McComb et al. (2007), peak


Langmuir

intensities at around 1100 cm− 1 and 3200–3400 cm− 1 change as the


0.841
0.841
0.790

0.988
1.000
0.755

inner-sphere complex formation due to adsorption of Sb(V) onto iron


R2

R2

oxide with the help of hydroxyl groups from adsorbed antimonate.


Interesting alterations occurred in the XRD pattern of post-adsorbed
Fe-biochar

Fe-biochar
Al-biochar

Al-biochar

Fe-biochar (Fig. 1b). After the Sb(V) interaction with Fe-biochar, new
Biochar

Biochar
Table 1

peaks emerged at two thetas of 60.45◦ and 80.53◦ (in PFe-biochar). The
senarmontite mineral (Sb2O3) had a characteristic peak around 80.53◦

6
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Fig. 4. Fittings of (a) pseudo-second-order, (b) Elovich, and (c) power function models for Sb adsorption onto pristine biochar, Fe-biochar, and Al-biochar. Solid lines
indicate the respective models fittings.

two theta degrees. Furthermore, the peak at 35.6◦ two theta degrees in 3.6. Effect of biochars on Sb immobilization in soil
Fe-biochar shifted to 34.6◦ two theta degrees in PFe-biochar, and the
peaks at 43.8◦ and 57.6◦ two theta degrees decreased after Sb(V) Considering good results of Sb adsorption onto Fe-biochar in aqueous
adsorption onto Fe-biochar. The mineral phase of Sb2O3 developed medium, it was further used to remediate the Sb-contaminated shooting
when Sb(V) decreased onto the solid surface during sorption by Fe- range soil in comparison with pristine biochar. Fig. 5 shows that both
biochar (Rahman et al., 2021b). pristine and modified biochar application gradually decreased the
Antimonates should bind to biochar primarily through an ion- various extracted forms of Sb in shooting range soil, including water-
exchange mechanism because it mainly exists as an anionic species in soluble, exchangeable, PBET, and TCLP-extracted fractions of Sb. The
solution. The Sb(V) in an aqueous solution by sorption onto biochar is water-soluble fraction was the most readily available form for plant
generally caused by one or more of the following factors: (i) electrostatic uptake and likely had the highest soil bioavailability. Here, both the
attraction; (ii) formation of nodules through hydrogen bonding; (iii) treatments of applied biochars significantly decreased the water-soluble
surface complexation and multilayer adsorption; (iv) ligand exchange of Sb level (Fig. 5a) from 0.12 mg kg− 1 (control soil) to almost zero for both
Sb(V) with biochar; and (v) co-precipitation. Specifically, for Fe- biochars at all application rates. The sorption of Sb through carbonyl
biochar, the unoccupied orbitals on Fe could aid Sb(V) complexation functional groups of biochars behaving as proton donors could be one
by forming inner Fe–O–Sb complex hydrogen bonds (Rahman et al., reason for Sb immobilization in soil. Likewise, the exchangeable fraction
2021a). Moreover, the protonated functional groups on the engineered significantly decreased with biochar addition (Fig. 5b), suggesting that
biochar surface could facilitate antimonate adsorption. The following Sb bioavailability to plants decreased.
dissociation reaction controls Sb(V) existence as an oxyanion or neutral The PBET test was also used to examine the effects of biochar on
species in the aqueous solution (Rahman et al., 2021a): mammals’ direct consumption of Sb-contaminated soil. Fig. 5c repre­
sents the findings of the PBET test for Sb content in the polluted-soil
Sb(OH)5 + H2O → Sb(OH)–6 + H+ treated with biochars. The PBET test is usually performed in acidic
The FTIR and XRD analyses confirmed the presence of O–H, N–H, conditions, hence, the control soil without biochar had the highest PBET
Fe–O–OH, Fe–O, and CO/C–O–C groups on the Fe-biochar surface, Sb content (7.33 mg kg− 1). Both biochars significantly decreased Sb bio-
which could enhance its capacity for complexing with Sb(V), resulting in accessibility, further decreasing with increasing biochar application.
ligand exchange at the Sb(V) center. The peak at 587.8 cm− 1 related to Increased soil Sb immobilization was evident in both biochars, reaching
the Fe–O–Sb band in the FTIR spectra, and the peak at 80.53◦ two theta reduction rates of 33.33% for pristine biochar and 37.50% for Fe-
related to senarmontite minerals in the XRD spectra confirmed that Sb biochar.
(V) adsorbed onto the Fe-biochar surface through surface complexation The TCLP test is used to reveal the metal toxicity of soil, especially for
and electrostatic attraction (Han et al., 2017; McComb et al., 2007; metal leaching from soil to groundwater due to acid rain. The control
Rahman et al., 2021a; Yin et al., 2020; Zhang et al., 2022). soil (without biochar) had a high level of TCLP Sb (3.20 mg kg− 1;
Fig. 5d) and thus greater susceptibility to Sb toxicity driven by leaching

7
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Fig. 5. (a) Water-soluble, (b) exchangeable, (c) physiologically based extraction test (PBET), and (d) toxicity characteristic leaching procedure (TCLP) extracted Sb
concentrations in shooting range soil amended with pristine biochar and Fe-biochar. ** not detected.

into groundwater. The biochars at various applications significantly


reduced Sb leaching. The amount of Sb immobilized in the soil increased
with increasing biochar application, possibly reducing groundwater
toxicity. Similar to bioavailable Sb, Fe-biochar decreased the leachable
Sb fraction to a greater extent (up to 54%) than pristine biochar (up to
52%), suggesting that Fe-biochar had a stronger binding affinity with Sb
than pristine biochar. Overall, both biochars successfully immobilized
Sb, reduced its ability to induce toxicity, and remediated the Sb-
contaminated soil. The greater immobilization effect of the Fe-biochar
as compared to the pristine biochar is due to the strong chemical
interaction between the Sb and Fe-biochar surface. The results of the Sb
removal from water were well-aligned with those for Sb remediation in
contaminated soil.
We reported good efficiencies for the pristine and Fe-biochars
compared to the limited research on Sb immobilization in soil using
engineered biochar. For example, according to some investigations,
ferrous sulfate immobilizes Sb by reducing Sb(V) to Sb(III), bonding
with iron oxide/hydroxide (Wang et al., 2019a). Similarly, other re­
searchers used Fe- and Fe–Mn-modified biochars for Sb immobilization
in soil (Almås et al., 2019; Tandy et al., 2017; Wang et al., 2019b). Fig. 6. Sequentially extracted fractions of Sb in shooting range soil amended
with pristine biochar and Fe-biochar. F1: non-specifically sorbed, F2: specif­
ically sorbed fraction, F3: interconnected with amorphous and imperfectly
3.7. Sequential extractions of Sb in soil crystalline ferrous and aluminum hydrous oxides, F4: linked with well-
crystallized Fe- and Al-hydrous oxides, and F5: residual soil fraction.
Fig. 6 shows Sb fractions in soil obtained from sequential extractions,
with the most predominant well-crystallized hydrous oxides of Fe/Al significantly changed the distribution of Sb fractions in the soil,
and residual phases, occupying 53.94% and 9.42%, respectively in the increasing the residual Sb fraction to 45.30% (pristine biochar) and
unamended control soil. The soil contained 0.75% and 1.25% of Sb 61.23% (Fe-biochar) and the well-crystallized hydrous oxides of the Fe/
associated with non-specifically and specifically sorbed fractions, Al fraction of Sb to 31.47% (pristine biochar) and 24.23% (Fe-biochar).
respectively. A minor amount of Sb was present in the amorphous and In contrast, in the Sb fractions associated with specifically sorbed and
poorly crystalline hydrous oxide of Fe/Al (0.54%). Biochar addition non-specifically sorbed decreased in the biochar-amended soils and

8
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

those associated with the poorly crystalline hydrous oxides of Fe/Al functional groups) directly influence Sb immobilization in soil, where it
fraction of Sb did not significantly differ. The biochars, particularly Fe- is most commonly found as oxides, hydroxides, or oxyanions. The sur­
biochar, transformed the pre-existing Sb into the residual form, the most face of Fe-biochar is abundant in oxygen-containing functional groups,
stable form of any metal in soil. The success of an immobilizing agent in such as hydroxyl, carboxyl, amine, and phenolic , to be involved in the
soil depends on the increase in the residual fraction of contaminant. conversion of Sb into a residual fraction. Sb immobilization by iron
Here, the biochar was an efficient immobilizing agent for Sb in the oxide adsorbents could be due to strong iron-associated ions resulting in
shooting range soil. direct precipitation, co-precipitation, and adsorption (Guo et al., 2009,
Other soil parameters, such as pH and EC, were examined to assess 2014; Wu et al., 2010). In this context, non-destructive elemental dot
how biochars affect soil quality, with no significant difference in these mapping may provide information on the association between target
parameters before or after incubation, indicating that treating Sb- contaminants and other elements in soil. Microstructural variations and
contaminated soil with biochar did not adversely impact plant growth Sb prevalence in soils were identified using SEM and EDX (Fig. 7). The
under these soil conditions . EDX spectrum of Fe-biochar amended soil showed the occurrence of Si,
Pb, Fe, and Sb elements. Elemental dot maps further revealed that Sb
was associated with Fe, Si, and O in the soil, which assumed the for­
3.8. Mechanism of Sb immobilization by biochars in soil
mation of some type of stable Sb compounds with Fe, O and Si elements
in the Fe-biochar amended soil. Pb was also predicted because shooting
The characteristics of biochar itself (e.g., presence of abundant

Fig. 7. SEM, EDX, and elemental dot maps of shooting range soil amended with Fe-biochar.

9
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

range soils generally contain substantial amount of Pb. Appendix A. Supplementary data
Most of the elemental forms in the residual state are found in sili­
cates; thus, silicates may be responsible for the elevated residual fraction Supplementary data to this article can be found online at https://doi.
of Sb in the soil. On the other hand, Sb(V) occurs as H2SbO4 and Sb(OH)–6 org/10.1016/j.envpol.2022.120723.
and is the predominant oxidation state at pH 2.7–12 in soils (Okkenhaug
et al., 2011). Complexation, reduction, and electrostatic interactions References
may be the main driving forces behind Sb immobilization in soils.
Additionally, the soil sequential extraction data revealed that the addi­ Ahmad, M., Hashimoto, Y., Moon, D.H., Lee, S.S., Ok, Y.S., 2012a. Immobilization of lead
in a Korean military shooting range soil using eggshell waste: an integrated
tion of Fe-biochar considerably enhanced the residual fraction of Sb in mechanistic approach. J. Hazard Mater. 392–401. https://doi.org/10.1016/J.
soils, which could be due in part to the occurrence of geochemically JHAZMAT.2012.01.047, 209–210.
stable Fe(III) antimonate mineral (tripuhyite, FeSbO4) (Wang et al., Ahmad, M., Lee, S.S., Yang, J.E., Ro, H.M., Lee, Y.H., Ok, Y.S., 2012b. Effects of soil
dilution and amendments (mussel shell, cow bone, and biochar) on Pb availability
2019a; Zhang et al., 2019). Hence, it can be speculated that Sb in the and phytotoxicity in military shooting range soil. Ecotoxicol. Environ. Saf. 79,
shooting range soil transformed into the stable Fe–O–Sb(OH)5 complex 225–231. https://doi.org/10.1016/J.ECOENV.2012.01.003.
by addition of Fe-biochar. In the natural environment, tripuhyite and Akhtar, L., Ahmad, M., Iqbal, S., Abdelhafez, A.A., Mehran, M.T., 2021. Biochars’
adsorption performance towards moxifloxacin and ofloxacin in aqueous solution:
schafarzikite are geochemically stable forms of Sb in soil (Leverett et al., role of pyrolysis temperature and biomass type. Environ. Technol. Innovat. 24,
2012; Teng et al., 2020). 101912 https://doi.org/10.1016/J.ETI.2021.101912.
Alaboudi, K.A., Ahmed, B., Brodie, G., 2019. Effect of biochar on Pb, Cd and Cr
availability and maize growth in artificial contaminated soil. Ann. Agric. Sci. (Cairo)
4. Conclusions
64 (1), 95–102. https://doi.org/10.1016/J.AOAS.2019.04.002.
Almås, Å.R., Pironin, E., Okkenhaug, G., 2019. The partitioning of Sb in contaminated
This study investigated pinecone-derived pristine biochar and its soils after being immobilization by Fe-based amendments is more dynamic
engineered biochars (Fe- and Al-biochars) as potential adsorbents for Sb compared to Pb. Appl. Geochem. 108, 104378 https://doi.org/10.1016/J.
APGEOCHEM.2019.104378.
(V) removal from aqueous solution. The experimental adsorption data Amin, M., Ahmad, M., Farooqi, A., Hussain, Q., Ahmad, M., Al-Wabel, M.I., Saleem, H.,
revealed that Fe-biochar had better adsorption capacity than the other 2020. Arsenic release in contaminated soil amended with unmodified and modified
two biochars due to the high affinity of Sb(V) to iron-associated func­ biochars derived from sawdust and rice husk. J. Soils Sediments 20 (9), 3358–3367.
https://doi.org/10.1007/S11368-020-02661-9/TABLES/3.
tional groups on its surface. The post-adsorption experiments with FTIR Amin, M.T., Alazba, A.A., Shafiq, M., 2021. Successful application of Eucalyptus
and XRD further confirmed the association of minerals like Fe, O–H, camdulensis biochar in the batch Adsorption of crystal violet and methylene blue
N–H, FeO–OH, C–O and C– – C with Sb(V) adsorption onto Fe-biochar. dyes from aqueous solution. Sustainability 2021 13 (7), 3600. https://doi.org/

10.3390/SU13073600.
Furthermore, isotherm and kinetics models revealed that Sb adsorp­ Ashley, P.M., Craw, D., Graham, B.P., Chappell, D.A., 2003. Environmental mobility of
tion onto biochars followed the chemisorption phenomena. Similarly, antimony around mesothermal stibnite deposits, New South Wales, Australia and
the pristine and Fe-biochars immobilized Sb in a shooting range soil southern New Zealand. J. Geochem. Explor. 77 (1), 1–14. https://doi.org/10.1016/
S0375-6742(02)00251-0.
(particularly the Fe-biochar) as indicated by the decline in water- Ayoob, S., Gupta, A.K., Bhakat, P.B., Bhat, V.T., 2008. Investigations on the kinetics and
soluble, exchangeable, PBET, and TCLP-extracted Sb levels. The pri­ mechanisms of sorptive removal of fluoride from water using alumina cement
mary mechanism for Sb immobilization and stabilization was attributed granules. Chem. Eng. J. 140 (1–3), 6–14. https://doi.org/10.1016/J.
CEJ.2007.08.029.
to the formation of more stable forms of Sb bound to O− , Fe− and Si-
Bagherifam, S., Brown, T.C., Fellows, C.M., Naidu, R., 2019. Derivation methods of soils,
associated minerals. water and sediments toxicity guidelines: a brief review with a focus on antimony.
J. Geochem. Explor. 205 https://doi.org/10.1016/J.GEXPLO.2019.106348.
Beesley, L., Marmiroli, M., 2011. The immobilisation and retention of soluble arsenic,
Credit author statement
cadmium and zinc by biochar. Environ. Pollut. 159, 474–480. https://doi.org/
10.1016/j.envpol.2010.10.016.
Basit Ahmed Khan: Conceptualisation, Experimentation, Writing – Bolan, N., Kumar, M., Singh, E., Kumar, A., Singh, L., Kumar, S., Keerthanan, S.,
original draft, Methodology, Investigation, Formal analysis, Data cura­ Hoang, S.A., El-Naggar, A., Vithanage, M., Sarkar, B., Wijesekara, H.,
Diyabalanage, S., Sooriyakumar, P., Vinu, A., Wang, H., Kirkham, M.B., Shaheen, S.
tion. Mahtab Ahmad: Conceptualisation, Supervision, Formal analysis, M., Rinklebe, J., Siddique, K.H.M., 2022. Antimony contamination and its risk
Data curation, Writing – original draft, and editing. Sajid Iqbal: Formal management in complex environmental settings: a review. Environ. Int. 158, 106908
analysis, Data curation, Investigation, Results interpretation. Fath Ullah: https://doi.org/10.1016/J.ENVINT.2021.106908.
Campos, T., Chaer, G., Leles, P.D.S., Silva, M., Santos, F., 2019. Leaching of heavy metals
Experimentation, Investigation, Formal analysis. Nanthi Bolan: Results in soils conditioned with biosolids from sewage sludge. Floresta e Ambiente 26
interpretation, Writing, Reviewing and editing. Zakaria Solaiman: Re­ (Specialissue1). https://doi.org/10.1590/2179-8087.039918.
sults interpretation, Writing, Reviewing and editing. Munib Ahmed Cao, D., Guo, T., Zhao, X., 2019. Treatment of Sb(V) and Co(II) containing wastewater by
electrocoagulation and enhanced Sb(V) removal with Co(II) presence. Separ. Purif.
Shafique: Formal analysis, Investigation, Results interpretation. Technol. 227 https://doi.org/10.1016/J.SEPPUR.2019.05.091.
Kadambot H. M. Siddique: Results interpretation, Writing, Reviewing Chen, Z.L., Zhang, J.Q., Huang, L., Yuan, Z.H., Li, Z.J., Liu, M.C., 2019. Removal of Cd
and editing and Pb with biochar made from dairy manure at low temperature. J. Integr. Agric.
18 (1), 201–210. https://doi.org/10.1016/S2095-3119(18)61987-2.
Chu, J., Hu, X., Kong, L., Wang, N., Zhang, S., He, M., Ouyang, W., Liu, X., Lin, C., 2021.
Declaration of competing interest Dynamic flow and pollution of antimony from polyethylene terephthalate (PET)
fibers in China. Sci. Total Environ. 771, 144643. https://doi.org/10.1016/J.
SCITOTENV.2020.144643.
The authors declare that they have no known competing financial
Cidu, R., Biddau, R., Dore, E., Vacca, A., Marini, L., 2014. Antimony in the
interests or personal relationships that could have appeared to influence soil–water–plant system at the Su Suergiu abandoned mine (Sardinia, Italy):
the work reported in this paper. strategies to mitigate contamination. Sci. Total Environ. 319–331. https://doi.org/
10.1016/J.SCITOTENV.2014.07.117.
Din, S.U., Azeez, A., Zain-ul-Abdin, Haq, S., Hafeez, M., Imran, M., Hussain, S.,
Data availability Alarfaji, S.S., 2021. Investigation on cadmium ions removal from water by a
nanomagnetite based biochar derived from Eleocharis dulcis. J. Inorg. Organomet.
No data was used for the research described in the article. Polym. Mater. 31 (1), 415–425. https://doi.org/10.1007/S10904-020-01758-5.
Diquattro, S., Garau, G., Mangia, N.P., Drigo, B., Lombi, E., Vasileiadis, S., Castaldi, P.,
2020. Mobility and potential bioavailability of antimony in contaminated soils:
Acknowledgements short-term impact on microbial community and soil biochemical functioning.
Ecotoxicol. Environ. Saf. 196 https://doi.org/10.1016/J.ECOENV.2020.110576.
Du, X., Wang, Y., Su, X., Li, J., 2009. Influences of pH value on the microstructure and
The study was supported by the International Research Support phase transformation of aluminum hydroxide. Powder Technol. 192 (1), 40–46.
Initiative Program of the Higher Education Commission of Pakistan (No: https://doi.org/10.1016/J.POWTEC.2008.11.008.
1–8/HEC/HRD/2022/11965), and by the University Research Fund
(URF) of Quaid-i-Azam University.

10
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Dupont, D., Arnout, S., Jones, P.T., Binnemans, K., 2016. Antimony recovery from end- Mariussen, E., Johnsen, I.V., Strømseng, A.E., 2017. Distribution and mobility of lead
of-life products and industrial process residues: a critical review. J. Sustain. Metall. (Pb), copper (Cu), zinc (Zn), and antimony (Sb) from ammunition residues on
2, 79–103. https://doi.org/10.1007/S40831-016-0043-Y. shooting ranges for small arms located on mires. Environ. Sci. Pollut. Res. Int. 24
El-Naggar, A., Shaheen, S.M., Ok, Y.S., Rinklebe, J., 2018. Biochar affects the dissolved (11), 10182–10196. https://doi.org/10.1007/S11356-017-8647-8.
and colloidal concentrations of Cd, Cu, Ni, and Zn and their phytoavailability and McComb, K.A., Craw, D., McQuillan, A.J., 2007. ATR-IR spectroscopic study of
potential mobility in a mining soil under dynamic redox-conditions. Sci. Total antimonate adsorption to iron oxide. Langmuir 23 (24), 12125–12130. https://doi.
Environ. 624, 1059–1071. https://doi.org/10.1016/J.SCITOTENV.2017.12.190. org/10.1021/LA7012667.
Fan, J., Xu, X., Ni, Q., Lin, Q., Fang, J., Chen, Q., Shen, X., Lou, L., 2018. Enhanced As(V) Murad, H.A., Ahmad, M., Bundschuh, J., Hashimoto, Y., Zhang, M., Sarkar, B., Ok, Y.S.,
removal from aqueous solution by biochar prepared from iron-impregnated corn 2022. A remediation approach to chromium-contaminated water and soil using
straw. J. Chem. https://doi.org/10.1155/2018/5137694, 2018. engineered biochar derived from peanut shell. Environ. Res. 204, 112125 https://
Fort, M., Grimalt, J.O., Querol, X., Casas, M., Sunyer, J., 2016. Evaluation of atmospheric doi.org/10.1016/J.ENVRES.2021.112125.
inputs as possible sources of antimony in pregnant women from urban areas. Sci. Okkenhaug, G., Zhu, Y.G., Luo, L., Lei, M., Li, X., Mulder, J., 2011. Distribution,
Total Environ. 544, 391–399. https://doi.org/10.1016/J.SCITOTENV.2015.11.095. speciation and availability of antimony (Sb) in soils and terrestrial plants from an
Guo, X., Wu, Z., He, M., 2009. Removal of antimony(V) and antimony(III) from drinking active Sb mining area. Environ. Pollut. 159 (10), 2427–2434. https://doi.org/
water by coagulation-flocculation-sedimentation (CFS). Water Res. 43 (17), 10.1016/J.ENVPOL.2011.06.028.
4327–4335. https://doi.org/10.1016/J.WATRES.2009.06.033. Palansooriya, K.N., Shaheen, S.M., Chen, S.S., Tsang, D.C.W., Hashimoto, Y., Hou, D.,
Guo, X., Wu, Z., He, M., Meng, X., Jin, X., Qiu, N., Zhang, J., 2014. Adsorption of Bolan, N.S., Rinklebe, J., Ok, Y.S., 2020. Soil amendments for immobilization of
antimony onto iron oxyhydroxides: adsorption behavior and surface structure. potentially toxic elements in contaminated soils: a critical review. Environ. Int. 134,
J. Hazard Mater. 276, 339–345. https://doi.org/10.1016/J.JHAZMAT.2014.05.025. 105046 https://doi.org/10.1016/J.ENVINT.2019.105046.
Hafeez, A., Pan, T., Tian, J., Cai, K., 2022. 2022. Modified biochars and their effects on Poletto, M., Omaghi Junior, H.L., Zattera, A.J., 2014. Native cellulose: structure,
soil quality: a review. Environments 9 (5), 60. https://doi.org/10.3390/ characterization and thermal properties. Materials 7 (9), 6105. https://doi.org/
ENVIRONMENTS9050060. 10.3390/MA7096105.
Han, L., Sun, H., Ro, K.S., Sun, K., Libra, J.A., Xing, B., 2017. Removal of antimony (III) Rahman, M., Rahman, M.M., Bahar, M., Sanderson, P., Lamb, D., 2021a. Antimonate
and cadmium (II) from aqueous solution using animal manure-derived hydrochars sequestration from aqueous solution using zirconium, iron and zirconium-iron
and pyrochars. Bioresour. Technol. 234, 77–85. https://doi.org/10.1016/J. modified biochars. Sci. Rep. 11 (1), 1–11. https://doi.org/10.1038/S41598-021-
BIORTECH.2017.02.130. 86978-6.
He, Z., Wei, Z., Zhang, Q., Zou, J., Pan, X., 2019. Metal oxyanion removal from Rahman, M.A., Rahman, M.M., Bahar, M., Sanderson, P., Lamb, D., 2021b.
wastewater using manganese-oxidizing aerobic granular sludge. Chemosphere 236, Transformation of antimonate at the biochar–solution interface. ACS ES&T Water 1
124353. https://doi.org/10.1016/J.CHEMOSPHERE.2019.124353. (9), 2029–2036. https://doi.org/10.1021/ACSESTWATER.1C00115.
Herath, I., Vithanage, M., Bundschuh, J., 2017. Antimony as a global dilemma: Saka, C., 2012. BET, TG–DTG, FT-IR, SEM, iodine number analysis and preparation of
geochemistry, mobility, fate and transport. Environ. Pollut. 223, 545–559. https:// activated carbon from acorn shell by chemical activation with ZnCl2. J. Anal. Appl.
doi.org/10.1016/J.ENVPOL.2017.01.057. Pyrol. 95, 21–24. https://doi.org/10.1016/J.JAAP.2011.12.020.
Hongtao, L., Shuxia, L., Hua, Z., Yanling, Q., Daqiang, Y., Jianfu, Z., Zhiliang, Z., 2018. Shetty, R., Prakash, N.B., 2020. Effect of different biochars on acid soil and growth
Comparative study on synchronous adsorption of arsenate and fluoride in aqueous parameters of rice plants under aluminium toxicity. Sci. Rep. 10 (1), 1–10. https://
solution onto MgAlFe-LDHs with different intercalating anions. RSC Adv. 8 (58), doi.org/10.1038/s41598-020-69262.
33301–33313. https://doi.org/10.1039/C8RA05968C. Tandy, S., Meier, N., Schulin, R., 2017. Use of soil amendments to immobilize antimony
Hu, L., Fu, J., Wang, S., Xiang, Y., Pan, X., 2020. Microplastics generated under simulated and lead in moderately contaminated shooting range soils. J. Hazard Mater. 324,
fire scenarios: characteristics, antimony leaching, and toxicity. Environ. Pollut. 269, 617–625. https://doi.org/10.1016/J.JHAZMAT.2016.11.034.
115905 https://doi.org/10.1016/J.ENVPOL.2020.115905. Teng, F., Zhang, Y., Wang, D., Shen, M., Hu, D., 2020. Iron-modified rice husk hydrochar
Hua, L., Wu, C., Zhang, H., Cao, L., Wei, T., Guo, J., 2021. Biochar-induced changes in and its immobilization effect for Pb and Sb in contaminated soil. J. Hazard Mater.
soil microbial affect species of antimony in contaminated soils. Chemosphere 263, 398, 122977 https://doi.org/10.1016/J.JHAZMAT.2020.122977.
127795. https://doi.org/10.1016/J.CHEMOSPHERE.2020.127795. Tessier, A., Campbell, P.G.C., Bisson, M., 1979. Sequential extraction procedure for the
Intrakamhaeng, V., Clavier, K.A., Liu, Y., Townsend, T.G., 2020. Antimony mobility from speciation of particulate trace metals. Anal. Chem. 51 (7), 844–851. https://doi.org/
E-waste plastic in simulated municipal solid waste landfills. Chemosphere 241, 10.1021/AC50043A017/ASSET/AC50043A017.FP.PNG_V03.
125042. https://doi.org/10.1016/J.CHEMOSPHERE.2019.125042. United States Environmental Protection Agency, 1992. Toxicity Characteristic Leaching
Ji, Y., Sarret, G., Schulin, R., Tandy, S., 2017. Fate and chemical speciation of antimony Procedure. Method 1311.
(Sb) during uptake, translocation and storage by rye grass using XANES Wang, L., Wang, J., Wang, Z., He, C., Lyu, W., Yan, W., Yang, L., 2018. Enhanced
spectroscopy. Environ. Pollut. 231, 1322–1329. https://doi.org/10.1016/J. antimonate (Sb(V)) removal from aqueous solution by La-doped magnetic biochars.
ENVPOL.2017.08.105. Chem. Eng. J. 354, 623–632. https://doi.org/10.1016/j.cej.2018.08.074.
Jia, X., Zhou, J., Liu, J., Liu, P., Yu, L., Wen, B., Feng, Y., 2020. The antimony sorption Wang, H., Lv, Z., Wang, B., Wang, Y.N., Sun, Y., Tsang, Y.F., Zhao, J., Zhan, M., 2019a.
and transport mechanisms in removal experiment by Mn-coated biochar. Sci. Total Effective stabilization of antimony in Waste-to-Energy fly ash with recycled
Environ. 724, 138158 https://doi.org/10.1016/J.SCITOTENV.2020.138158. laboratory iron–rich residuals. J. Clean. Prod. 230, 685–693. https://doi.org/
Johnson, C.A., Moench, H., Wersin, P., Kugler, P., Wenger, C., 2005. Solubility of 10.1016/J.JCLEPRO.2019.05.128.
antimony and other elements in samples taken from shooting ranges. J. Environ. Wang, Y.Y., Ji, H.Y., Lyu, H.H., Liu, Y.X., He, L.L., You, L.C., Zhou, C.H., Yang, S.M.,
Qual. 34 (1), 248–254. https://doi.org/10.2134/JEQ2005.0248. 2019b. Simultaneous alleviation of Sb and Cd availability in contaminated soil and
Karimipour, M., Moradi, N., Molaei, M., Dargahzadeh, M., 2019. Fabrication of single- accumulation in Lolium multiflorum Lam. After amendment with Fe–Mn-Modified
phase superparamagnetic iron oxide nanoparticles from factory waste soil. Sci. Iran. biochar. J. Clean. Prod. 231, 556–564. https://doi.org/10.1016/J.
26 (6), 3938–3945. https://doi.org/10.24200/SCI.2019.51960.2448. JCLEPRO.2019.04.407.
Khan, B.A., Ahmad, M., Iqbal, S., Bolan, N., Zubair, S., Shafique, M.A., Shah, A., 2022. Wang, N., Deng, N., Qiu, Y., Su, Z., Huang, C., Hu, K., Wang, J., Ma, L., Xiao, E., Xiao, T.,
Effectiveness of the engineered pinecone-derived biochar for the removal of fluoride 2020. Efficient removal of antimony with natural iron minerals. Environ. Chem. 17
from water. Environ. Res. 212, 113540 https://doi.org/10.1016/J. (4), 332–344. https://doi.org/10.1071/EN20002_AC.
ENVRES.2022.113540. Wu, Z., He, M., Guo, X., Zhou, R., 2010. Removal of antimony (III) and antimony (V)
Laporte-Saumure, M., Martel, R., Mercier, G., 2011. Characterization and metal from drinking water by ferric chloride coagulation: competing ion effect and the
availability of copper, lead, antimony and zinc contamination at four Canadian small mechanism analysis. Separ. Purif. Technol. 76 (2), 184–190. https://doi.org/
arms firing ranges. Environ. Technol. 32 (7), 767–781. https://doi.org/10.1080/ 10.1016/J.SEPPUR.2010.10.006.
09593330.2010.512298. Wu, F., Fu, Z., Liu, B., Mo, C., Chen, B., Corns, W., Liao, H., 2011. Health risk associated
Le Bozec, N., Persson, D., Nazarov, A., Thierry, D., 2002. Investigation of filiform with dietary co-exposure to high levels of antimony and arsenic in the world’s largest
corrosion on coated aluminum alloys by FTIR microspectroscopy and scanning antimony mine area. Sci. Total Environ. 409 (18), 3344–3351. https://doi.org/
kelvin probe. J. Electrochem. Soc. 149 (9), B403. https://doi.org/10.1149/ 10.1016/J.SCITOTENV.2011.05.033.
1.1497172. Wu, Q., Xian, Y., He, Z., Zhang, Q., Wu, J., Yang, G., Zhang, X., Qi, H., Ma, J., Xiao, Y.,
Leverett, P., Reynolds, J.K., Roper, A.J., Williams, P.A., 2012. Tripuhyite and Long, L., 2019. Adsorption characteristics of Pb(II) using biochar derived from spent
schafarzikite : two of the ultimate sinks for antimony in the natural environment. mushroom substrate. Sci. Rep. 9 (1), 1–11. https://doi.org/10.1038/S41598-019-
Mineral. Mag. 76 (4), 891–902. https://doi.org/10.1180/MINMAG.2012.076.4.06. 52554-2.
Lito, P.F., Aniceto, J.P.S., Silva, C.M., 2012. Removal of anionic pollutants from waters Xu, X., Gao, B., Tan, X., Zhang, X., Yue, D., Yue, Q., 2013. Uptake of perchlorate from
and wastewaters and materials perspective for their selective sorption. Water Air Soil aqueous solutions by amine-crosslinked cotton stalk. Carbohydr. Polym. 98 (1),
Pollut. 223 (9), 6133–6155. https://doi.org/10.1007/S11270-012-1346-7. 132–138. https://doi.org/10.1016/J.CARBPOL.2013.05.058.
Liu, S., 2015. Cooperative adsorption on solid surfaces. J. Colloid Interface Sci. 450, Yin, G., Song, X., Tao, L., Sarkar, B., Sarmah, A.K., Zhang, W., Lin, Q., Xiao, R., Liu, Q.,
224–238. https://doi.org/10.1016/J.JCIS.2015.03.013. Wang, H., 2020. Novel Fe-Mn binary oxide-biochar as an adsorbent for removing Cd
Liu, Z., Zhang, F.S., Sasai, R., 2010. Arsenate removal from water using Fe3O4-loaded (II) from aqueous solutions. Chem. Eng. J. 389 https://doi.org/10.1016/J.
activated carbon prepared from waste biomass. Chem. Eng. J. 160 (1), 57–62. CEJ.2020.124465.
https://doi.org/10.1016/J.CEJ.2010.03.003.

11
B.A. Khan et al. Environmental Pollution 317 (2023) 120723

Zhang, C., Jiang, H., Deng, Y., Wang, A., 2019. Adsorption performance of antimony by Bioresources 11 (4), 9068–9078. https://doi.org/10.15376/BIORES.11.4.9068-
modified iron powder. RSC Adv. 9 (54), 31645–31653. https://doi.org/10.1039/ 9078.
C9RA05646G. Zhu, H., Huang, Q., Fu, S., Zhang, X., Yang, Z., Lu, J., et al., 2021. Removal of antimony
Zhang, L., Dong, Y., Liu, J., Liu, C., Liu, W., Lin, H., 2022. The effect of co-pyrolysis (V) from drinking water using nZVI/AC: optimization of batch and fix bed
temperature for iron-biochar composites on their adsorption behavior of antimonite conditions. Toxics 9 (10), 266. https://doi.org/10.3390/toxics9100266.
and antimonate in aqueous solution. Bioresour. Technol. 347, 126362 https://doi. Zhuang, W., Lai, X., Wang, Q., Liu, Y., Chen, Q., Liu, C., 2018. Distribution
org/10.1016/J.BIORTECH.2021.126362. characteristics, sources and ecological risk of antimony in the surface sediments of
Zhu, C., Guo, F., Guo, X., Li, X., 2016. In situ saccharification of cellulose using a Changjiang Estuary and the adjacent sea, East China. Mar. Pollut. Bull. 137,
cellulase mixture and supplemental β-glucosidase in aqueous- ionic liquid media. 474–480. https://doi.org/10.1016/J.MARPOLBUL.2018.10.049.

12

You might also like