You are on page 1of 10

Groundwater for Sustainable Development 14 (2021) 100585

Contents lists available at ScienceDirect

Groundwater for Sustainable Development


journal homepage: http://www.elsevier.com/locate/gsd

Efficient low-cost magnetic composite based on eucalyptus wood biochar


for arsenic removal from groundwater
Amalia Lara Bursztyn Fuentes a, Facundo Barraqué a, Roberto Carlos Mercader b,
Alberto Néstor Scian a, c, María Luciana Montes b, *
a
CETMIC - Centro de Tecnología de Recursos Minerales y Cerámica (CIC-CONICET La Plata), Cno. Centenario y 506, M.B. Gonnet, 1897, Argentina
b
IFLP - Instituto de Física La Plata (CONICET La Plata), Facultad de Ciencias Exactas, Universidad Nacional de La Plata, Diagonal 113 y 64, La Plata, 1900, Argentina
c
Departamento de Química, Facultad de Ciencias Exactas, Universidad Nacional de La Plata, Calle 1 y 47, La Plata, 1900, Argentina

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

Keywords: Low-cost, eco-friendly and efficient charcoal-based composites have been developed to remove arsenic from two
Magnetic charcoal Argentine naturally contaminated groundwater samples. Iron compounds have been added to the charcoals both
Fe oxides to enhance the composites’ sorption capacities and to provide permanent magnetization to the materials. The
Sorbent materials
raw charcoals, one made from eucalyptus wood charcoal (EWC) and a commercial activated carbon (CSC), as
Arsenic
well as their magnetic variants (MEWC and MCSC, respectively), have been characterized by several techniques
Groundwater
Argentina and tested for arsenic sorption. In the magnetic analogues, Mössbauer spectroscopy enabled the identification of
magnetite as the main Fe oxide together with some paramagnetic Fe phases and goethite. The magnetization
saturation values of 11.2 ± 0.5 Am2 kg− 1 for MEWC and 10.7 ± 0.5 Am2 kg− 1 for MCSC agree with Mössbauer
results. These saturation values indicate that it is possible to separate the sorbents from the liquid phase with an
external magnetic field. Arsenic removal was higher for MEWC than for MCSC, and arsenic sorption from
naturally contaminated water samples resulted lower than in the synthetic one. For natural waters, As residual
concentrations resulted lower than 50 μg L− 1, which is the limit for drinking water in Argentina. The satisfactory
results achieved for the homemade eucalyptus-derived composite indicate that it is a promising material to be
used in simple technological systems for water treatment.

1. Introduction et al., 2017a), which make them expensive. Alternative methodologies


are required to attain efficient and low-cost carbonaceous sorbents, and
Environmental monitoring in Argentina reveals that four million to improve their sorption capacity by modifying their internal and
people are chronically exposed to arsenic (As) through drinking water, external surfaces.
and it has been documented that up to 30% of patients with arsenicosis Supporting iron oxide particles onto carbon has proven to be an
will likely die of skin, liver, lung, bladder, stomach, or pancreas cancer efficient way to remove pollutants, such as arsenic (Dhanasekaran and
(Bardach et al., 2015). The current situation demands the development Sahu, 2020; Singh et al., 2018), avoiding the frequent problems of
of eco-friendly, simple, and low-cost As sorbents to provide potable agglomeration and oxidation that result from using these particles in
water sources to these communities. isolation (Shaikh et al., 2020; Zhang et al., 2013). Also, the magnetic
Carbonaceous materials, such as activated carbons, charcoals, and properties that those particles confer to the composite allow quick
biochars, derived from lignocellulosic biomass have been proposed as separation from the liquid phase after the pollutant removal by applying
promising materials to remove pollutants from water (Danish and an external magnetic field (Ahmad et al., 2019; Hu et al., 2018). How­
Ahmad, 2018; Li et al., 2017; Oliveira et al., 2017; Shaheen et al., 2019). ever, the syntheses of magnetic carbons generally include multiple steps,
However, biomass pyrolysis and activation require specific and sophis­ the vast majority of which involve specialized procedures and labora­
ticated equipment. For instance, commercial carbons are usually phys­ tory equipment, like high-temperature muffle furnaces with controlled
ically activated with steam, CO2, air or gas mixtures, or are chemically N2 atmosphere (Reddy and Lee, 2014; Yang et al., 2008), chemical
activated by adding zinc salts, KOH or H3PO4 (Adib et al., 2015; B. Wang co-precipitation with N2 bubbling (Han et al., 2015), or microwave

* Corresponding author. ,
E-mail addresses: lmontes@fisica.unlp.edu.ar, mlucianamontes@gmail.com (M.L. Montes).

https://doi.org/10.1016/j.gsd.2021.100585
Received 2 December 2020; Received in revised form 12 February 2021; Accepted 11 April 2021
Available online 23 April 2021
2352-801X/© 2021 Elsevier B.V. All rights reserved.
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

systems (Yap et al., 2017). loaded iron particles. Specific surface area (SBET) was estimated with the
The aim of this paper is to synthesize a magnetic composite with a BET equation, in the adequate pressure range as recommended by
high As sorption capacity, using locally available biomass and relatively IUPAC (Cychosz and Thommes, 2018; Thommes et al., 2015), and the
easy and low-cost techniques. Toward this goal, local eucalyptus wood total pore volume, VT (cm3 g− 1), at p/p0 = 0.97. The micropore volume,
has been turned into charcoal using an easy and inexpensive charring VDR (cm3 g− 1), was determined from the application of the
system, and Fe-oxide particles were synthesized onto the charcoal using Dubinin-Radushkevich (DR) equation (Dubinin, 1989). The mesopore
a versatile and simple procedure (a commercial carbon has been also volume, VMESO (cm3 g− 1), was calculated as the difference between VT
used as contrast material). In addition, a thorough characterization of and VDR.
the synthesized composites was carried out to better understand the As The morphology of the magnetic composites was studied with a
removal performance, being a crucial stage in the development of sor­ transmission electron microscope (CM200, Philips). The samples were
bent materials aimed at providing environmental technological dispersed in ethanol, using an ultrasonic bath, and they were further
solutions. deposited on a carbon-coated copper grid. Micrographs were taken at an
accelerating voltage of 200 kV in bright-field mode, and acquired using a
2. Materials and methods Gatan Orius CCD camera with an exposure time of 1 s field. The particle
size distribution of iron oxide was determined from these images using
2.1. Materials and reagents the software ImageJ (Rasband, 1997), with n = 90 particles per
material.
Two carbon precursors were used for the composites’ synthesis. One X-ray diffraction (XRD) patterns were collected using a Philips PW
charcoal was produced from eucalyptus wood, a residual biomass which 1710 diffractometer with CuKα radiation, operated at 40 kV and 30 mA,
is abundantly available in Argentina, following the procedure presented with counting time 10 s/step and 0.02◦ (2θ) step size. The main re­
in detail in a previous work (Bursztyn Fuentes et al., 2020). The flections were defined and contrasted with reference patterns. The
carbonization was performed inside an open kiln. The fire was started Scherrer equation was used to estimate the mean crystalline size of Fe
with small branches (diameter d about 1 cm), and it was continuously oxide particles:
fed with the addition of new larger branches (d < 4 cm). When all the

available biomass was used, distilled water was poured to put out the Dhkl = (1)
β cos(θ)
fire. The charcoal thus obtained was then filtered, dried at 105 ◦ C for 48
h and labeled EWC (for Eucalyptus Wood Charcoal). The other precursor where K is the shape factor (0.89), λ the Cu radiation wavelength (0.154
was a commercial steam-activated carbon (Clarimex S.A) produced from nm), and β and θ the full width at half maximum (FWHM) at half
coconut shell, labeled CSC (for Coconut Shell Charcoal), and it was used maximum and the diffraction angle of the reflection peak, respectively.
here as a contrast to the homemade charcoal. Both carbons were ground X-Ray Photoelectron Spectrometry (XPS) of the magnetic composites
and sieved between mesh #30 and #60 (ASTM), to obtain particles with was performed using a Thermo Scientific Nexsa system equipped with a
sizes in the range 600–250 μm. monochromatic Al Kα X-ray source (E = 1486.6 eV). Measurements were
The magnetic composites were synthesized using analytical grade done in the constant analyzer energy mode, with a spot size of 400 μm.
reagents: KNO3 and KOH (99%, Biopack) and FeSO4.7H2O (analytical The survey spectra were taken at an energy pass of 200 eV, while the
degree, Cicarelli). For arsenic removal tests, the control synthetic solu­ high resolution (HR) core level spectra were taken at a 50 eV pass energy
tion was prepared by dilution with ultrapure water of a 1000 mg L− 1 As (0.100 eV/step).
(V) standard solution (Merck, analytical grade). The room temperature 57Fe Mössbauer spectra of the magnetic var­
iants were taken in transmission geometry using a conventional constant
2.2. Magnetic composite synthesis acceleration spectrometer of 512 channels with a 5 mCi nominal activity
57
Co source in Rh matrix. The absorbers were powdered samples of a
The magnetic composites were synthesized according to the method suitable thickness to apply the approximate method by Montes et al.
described by Bartonkova et al. (2007). This procedure was chosen (2016) that enables estimating the total Fe content from the Mössbauer
because of its simplicity and versatility, i.e.: it allows to achieve mag­ spectra, which were collected between − 11 and + 11 mm s− 1. The hy­
netic adsorbents for the removal of pollutants, with an adequate mag­ perfine parameters are the result of fitting the data to hyperfine mag­
netic response, in a reproducible way, and without the need to use netic fields and quadrupole splitting distributions using a least-squares
nitrogen gas to avoid the oxidation of materials (Barraqué et al., 2018, computer code with Voigt lineshape. Isomer shift was calibrated against
2020, 2021; Montes et al., 2020; Zelaya Soulé et al., 2021). Briefly, the an α-Fe foil at room temperature.
precursor EWC or CSC was suspended in distilled water under magnetic The high-field susceptibility (χhifi), saturation magnetization (Ms),
stirring (2.5 g/425 mL water). Then, 25 mL of FeSO4⋅7H2O (0.3 M) were coercitive field (Hc), and remnant magnetization (Mr) were determined
added and left to interact for 2 h. Finally, 25 mL of KNO3 (0.49 M) and from the hysteresis loops acquired in a VSM magnetometer LakeShore
25 mL of KOH (1.25 M) were added sequentially, and the temperature of 7404 using external magnetic fields between − 1.9 T and 1.9 T. Samples
the suspension was raised to 90 ◦ C with the hotplate stirrer. Then, the were supported in a diamagnetic sample holder with a negligible mag­
suspension was cooled at room temperature. The supernatant was netic response.
filtered and 100 mL of distilled water was used to wash the solid phase
retained in the filter. Afterward, the solid was placed in a crucible and
oven-dried at 40 ◦ C overnight. The obtained products were labeled 2.4. Arsenic removal essays
MEWC (for Magnetic Eucalyptus Wood Charcoal) and MCSC (for Mag­
netic Coconut Shell Charcoal), respectively. Arsenic removal from aqueous solution was performed in batch
conditions. A fixed amount of adsorbent was weighed in amber stop­
2.3. Characterization of the materials pered bottles, and the corresponding volume of test solution (100 μg L− 1
As(V), pH = 7) was added so as to keep a solid-to-liquid ratio of 5 g L− 1.
Nitrogen sorption isotherms were obtained at − 196 ◦ C using an The tightly closed bottles were placed in a platform shaker. After 24 h of
automatic equipment (ASAP 2020; Micromeritics). The raw charcoals contact time at room temperature (25 ◦ C), the liquid and solid phases
were previously outgassed at 110 ◦ C under secondary vacuum for at were separated with a neodymium magnet. Solid phases were dried at
least 48 h while the magnetic variants were outgassed at 80 ◦ C for at 50 ◦ C and reserved for further analysis. Total arsenic concentration in
least 48 h, not exceeding the synthesis temperature so as not to alter the the liquid phase was measured using inductively coupled plasma mass

2
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

spectrometry (7800 ICP-MS, Agilent). Other magnetic materials synthesized with the same procedure had a
The same procedure was repeated with natural groundwater samples similar mean particle size (Barraqué et al., 2018; Montes et al., 2020),
gathered in October 2019 from private wells located in Castelli and 9 de indicating that Fe oxide particle sizes do not depend strongly on the
Julio cities, Province of Buenos Aires, Argentina. These cities are known characteristics of the support material.
to have high As concentrations in groundwaters (Litter et al., 2019). The Diffraction patterns in Fig. 2 showed differences between the carbon
water samples were characterized using standardized procedures precursors used. While CSC did not display crystalline structures, EWC
(American Public Health Association, 2018). exhibited the characteristic peaks of calcite (PDF 00-047-1743), a
The As removal percentage (%) was calculated according to equation compound that has been observed for other wood-derived charcoals
(2): (Khanna et al., 1994). The main identified calcite diffraction peaks, with
the corresponding Miller Indices, were: 23.04◦ (012), 29.32◦ (104),
Ci − Cf
As removal (%) = 100 (2) 35.92◦ (110), 39.36◦ (113), 43.08◦ (202), 47.24◦ (024 and 018) and
Ci
48.40◦ (116). In addition, all the materials exhibited a broad peak at 2θ
where Ci is the initial As concentration (μg L− 1) and Cf is the remaining = 22.7◦ , characteristic of carbons and biochars, which belongs to the
As concentration (μg L− 1) after 24 h of contact time. diffraction from the conversion of the lignin-cellulose crystal structure
The sorption capacity, Q (μg g− 1), was calculated according to to a more amorphous structure after pyrolysis (Regmi et al., 2012), but it
equation (3): decreases significantly after iron particles’ synthesis, as observed for a
( ) magnetic Douglas fir biochar (Karunanayake et al., 2019). The magnetic
Ci − Cf composites showed the presence of iron oxides, such as magnetite (PDF
Q= V (3)
W 01-075-0033) or maghemite (PDF 00-39-1346), which cannot be
distinguished from each other with this technique (Pecharromán et al.,
where V is the volume of test solution (L) and W is the sorbent material 1995). Eight diffraction peaks were identified for both diffraction pat­
mass (g). terns, with the following positions and Miller Indices: 18.29◦ (111),
For removal experiments, two independent batch systems were 30.11◦ (220), 35.45◦ (311), 37.00◦ (222), 43.10◦ (400), 53.41◦ (422),
performed, including a control test with the test solution and no solid. 56.99◦ (511) and 63.59◦ (440). The most intense line is that located at
Results are expressed as the mean of the independent experiments and 35.45◦ . MEWC also maintained the reflections that correspond to
plotted as mean ± standard deviation (SD). calcite, the crystalline structure that is present in the precursor, indi­
cating that magnetic composite synthesis did not disturb significantly
3. Results and discussion the structural properties of the precursors, as also observed for mont­
morillonite and beidellites composites synthesized with the same pro­
3.1. Magnetic composite characterization cedure (Barraqué et al., 2018; Montes et al., 2020).
The Fe oxide particles mean crystalline size was estimated using the
The synthesis of magnetic particles lowered the surface area (ABET), Scherrer equation (equation (1)). The peaks used for the estimation
in comparison with that of the raw charcoals, EWC and CSC (Table 1). were: 30.11◦ (220), 35.45◦ (311), 43.10◦ (400), 56.99◦ (511) and 63.59◦
For EWC, the surface area went from 392 m2 g− 1 to 73 m2 g− 1, and for (440). The mean value and the respective standard deviation resulted in
CSC from 1029 m2 g− 1 to 836 m2 g− 1. This may be related to the 23 ± 1 nm for MCSC and 21 ± 3 nm for MEWC. The average particle
blockage of carbon pores by the iron particles, which do not offer sig­ sizes calculated considering the 4 crystallographic directions mentioned
nificant surface area themselves, being lower than 20 m2 g− 1 (Zelaya were lower than the particle size determined by TEM. This could indi­
Soulé et al., 2021), in agreement with other magnetic carbons (Almeida cate that the FWHM diffraction peacks corresponding to magnetite/
et al., 2017; Cazzeta et al., 2016; Spessato et al., 2020). However, ABET maghemite could have other contributions, as microstrains and small
values for the magnetic materials resulted higher than those of other crystallite sizes (Maniammal et al., 2017).
activated carbons made from biomass residues, such as peanut hulls Low Resolution XPS of MEWC and MCSC gave information about the
(Almeida et al., 2017) and coconut shells (Cazzeta et al., 2016). surface chemical composition (Table S1) and High Resolution (HR) XPS
Micropores were the major contribution to the porosity of the raw enabled to go deeper into the analysis of C, O and Fe oxidation states.
material. A decrease in microporosity was observed for the composites, The low resolution XPS results indicated that oxygen, carbon, and iron
related to the iron particles deposition on the material surface, blocking are the principal components of magnetic carbons, with trace concen­
the micropores connected to the surface (Spessato et al., 2020). trations of K and Ca. The C1s, O1s, and Fe2p spectra collected and fitted
TEM images of both composites exhibited an abundant and homo­ to Gaussian curves are shown in Fig. 3. Their binding energies, FWHMs,
geneous surface deposit of Fe oxide particles, with a rhombic shape, and atomic percentages are included in the Supplementary material
characteristic of magnetite (Fig. 1). (Tables S2, S3 and S4.
The particle size distribution was estimated from the TEM images. In both magnetic composites, C1s spectra were deconvoluted into 3
Histograms were fitted to a normal distribution to estimate the mean peaks which are listed in Table S2, together with the corresponding
particle size and the standard deviation. MEWC presented a mean value parameters. The fittings show the binding energies characteristic of
of 61 ± 20 nm, while MCSC presented 52 ± 13 nm, with sizes ranging bondings in carbonaceous structures: the peak at 284.0 eV was assigned
between 26 and 118 nm and between 25 and 75 nm, respectively. The to C–C and C–H, while the peak at 285.5 eV could be associated to
particle size range resulted narrower for MCSC than MEWC, indicating C–O–C and C–OH. The binding energy peak between 290.2 and 289.4 eV
that Fe oxide particles were more homogeneously formed on the com­ corresponds to C– – O/CO2−3 (Alchouron et al., 2020; Navarathna et al.,
mercial carbon, probably as a consequence of a more regular surface. 2019; Spessato et al., 2020). It could also be associated to –COOH, but it
is less probable given the alkaline nature of the synthesis performed.
O1s spectra were deconvoluted into 4 peaks which are listed in
Table 1 Table S3, together with the corresponding parameters. While the main
Textural parameters of the raw charcoals and the magnetic variants.
peak at 529.66 eV is due to Fe–O–Fe (Fe2+-O-Fe2+, Fe2+-O-Fe3+ or Fe3+-
Sample ABET (m2 g− 1) VDR (cm3 g− 1) V0.97 (cm3 g− 1) VMESO (cm3 g− 1) O-Fe3+), the peak centered at 531.07 eV could belong to Fe–OH (octa­
EWC 392 0.15 0.17 0.02 hedral 111 plane of magnetite) and C–O compounds. In addition, the
CSC 1029 0.40 0.42 0.02 peaks at 532.38 and 533.63 eV can be related to C–O–Fe and CO2− 3 /-
MEWC 73 0.04 0.12 0.08 COOH(R) (Navarathna et al., 2019).
MCSC 836 0.35 0.50 0.15
The Fe2p HR-XPS spectra of MEWC and MCSC resulted relatively

3
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Fig. 1. TEM images of MEWC sample (above) and MCSC (below). The particle size distribution of the observed Fe oxide particles and the normal distribution fitting
are shown on the right.

complex, with many possible fittings and multiple deconvolutions. The mentioned should reveal several peaks and their respective XPS spectra,
information about the proposed fittings is listed in Table S4. Considering including their satellite peaks (Frau et al., 2010; Grosvenor et al., 2004;
that many Fe compounds can contribute to the peaks observed in the Zhang and Jia, 2014). Thus, the following peaks should be noticed in the
XPS spectra, the Fe-phases identification attempt was completed using XPS spectra of ferrihydrite 724.5, 732.2, 710.8 and 719.3 eV. Peaks at
Mössbauer spectroscopy. Fig. 4 shows the Mössbauer spectra collected 711.0, 718.9, 724.7, and 733.5 eV must show up for akaganeite, and
for both composites. Table 3 exhibits the hyperfine parameters deter­ 711.8 eV, 725.7 eV and 720.5 eV for lepidocrocite. Although not
mined for each proposed Fe site, together with the relative spectral area. unambiguously, because of the broadening and overlap of peaks, the
The estimated total Fe concentration is also presented. current Mössbauer and XPS results are evidence of the existence of these
The Mössbauer spectra of both magnetic composites were fitted to Fe compounds in both composites. Considering the magnetic phases,
two ferric paramagnetic sites (Fe3+), two magnetic sites of hyperfine since magnetite and goethite formation was confirmed by Mössbauer
parameters similar to those of magnetite, and a third magnetic envi­ spectroscopy, the XPS spectra fitting at lower energies can be associated
ronment that can be assigned to goethite. to the overlap of signals coming from these Fe oxides: magnetite (708.3,
The XPS spectra revealed only the existence of C, O and Fe atoms. In 709.3 and 710.4 eV for Fe2+, with a satellite signal around 715 eV;
addition, they showed that the C atoms are bound only to O atoms and 710.2, 711.3, 712.4 and 713.6 eV for Fe3+), and goethite (710.2, 711.2,
not to Fe ones. Hence, the components of the Mössbauer spectra can 712.1, 713.2 eV with surface and satellite peaks around 714 and 720 eV,
originate in Fe oxides like β-FeOOH (akaganeite), Fe5HO8.4H2O (ferri­ respectively) (Grosvenor et al., 2004).
hydrite) or γ-FeOOH (lepidocrocite). The Mössbauer spectrum of aka­ The Mössbauer results showed that, although MEWC and MCSC
ganeite displays two paramagnetic Fe3+ signals with hyperfine spectra displayed equivalent sites, the relative areas assigned to
parameters δ = 0.39 mm s− 1 and Δ = 0.95; and δ = 0.38 mm s− 1 and Δ = magnetite differ. Taking into account the total Fe concentration esti­
0.55. The spectrum of ferrihydrite is made up of only one Fe3+ singlet of mated and the RSA for both magnetite sites, the approximate magnetite
δ = 0.35 mm s− 1 and Δ = 0.62 mm s− 1 and that of lepidocrocite exhibits concentrations were roughly 81 ± 11 g kg− 1 and 108 ± 15 g kg− 1 for
a single line of parameters δ = 0.30 mm s− 1 and Δ = 0.55 mm s− 1 MCSC and MEWC, respectively. This difference might arise from
(Mccammon, 1995). Therefore, the Fe3+ paramagnetic environments of different magnetite degrees of oxidation. For pure magnetite the RSA
MEWC might belong to akaganeite, but because of the overlapping lines, (%) of site B should be 1.8 times that of site A (Vandenberghe and Grave,
the presence of lepidocrocite or ferrihydrite cannot be fully discarded. 2013). The ratio of areas for the signals assigned to magnetite in MCSC
The hyperfine parameters of Fe3+ paramagnetic environments corre­ lie within the range of non-oxidized magnetite. MEWC, instead,
sponding to MCSC cannot be unequivocally assigned. However, taking exhibited areas that depart from the ideal ratio. If the current assump­
into account that the sample is synthesized in the same way as MEWC, tions are valid, the degree of oxidation of magnetite in MEWC would
the observed signals might be associated to not-well-formed akaganeite. lead to a structural formula of Fe2.87O4 (Murad and Cashion, 2004).
If the identification described above is correct, the three Fe compounds The samples also differed in the concentration of goethite; that in

4
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Fig. 2. Diffraction patterns of the carbon precursors EWC and CSC and of their magnetic variants. Symbols: calcite (C), iron oxides (+) and cellulose (Cell). Numbers
in brackets represent the Miller indices for calcite on EWC and for magnetite/maghemite on MEWC and MCSC.

Fig. 3. HR-XPS deconvolutions of MEWC and MCSC C1s, O1s and Fe2p. Lines and numbers refer to the assigned deconvolutions (Tables S2, S3 and S4).

MCSC was about twice that in MEWC (20 g kg− 1 and 8 g kg− 1). Some­ indicating that the paramagnetic behavior may be associated to
thing similar happened to the relative abundance of Fe3+ environments. magnetite particles with sizes lower than 36 nm while the hysteresis
These differences might be relevant in the removal stage for pollutants may be originated in the magnetite particles with larger size (Dunlop,
that have high affinity to iron compounds, such as arsenic. 1973).
Fig. 5 displays the hysteresis loops and Table 3 presents the magnetic The saturation magnetization, Ms, is related to the attraction ca­
parameters extracted from them. The magnetic response of the com­ pacity of materials by a magnet: the higher the Ms value, the higher the
posites revealed that they can be manipulated by external magnetic magnetic interaction. MEWC exhibited a Ms value of 11.2 ± 0.5 Am2
fields (photograph in Fig. 5), enabling the easy recovery from the kg− 1, while MCSC had 10.7 ± 0.5 Am2 kg− 1. The saturation magneti­
suspension. zation of these composites resulted higher than the values achieved for
Both composites revealed Mr and Hc values different from zero, i.e., magnetic natural montmorillonite (Barraqué et al., 2018) and some
they have hysteresis behavior. The samples are also paramagnetic, with synthetic beidellite (Montes et al., 2020) subjected to the same synthesis
saturation not achieved even for the highest external magnetic field. procedure. Besides, these values resulted lower than for magnetic acti­
These results are supported by the Mössbauer spectra and TEM results vated carbons derived from coconut shells (Cazzeta et al., 2016) and

5
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

the WHO recommended drinking water limit.


Sorption on raw charcoals may be related to the existence of
carbonyl and carboxyl groups, but magnetite seems to play a key role for
the magnetic variants: the Fe and O atoms could act as Lewis acids in the
arsenate sorption on magnetic adsorbents and then arsenate may be
sorbed selectively by inner-sphere complex formation (Zhang et al.,
2010).

3.2.2. Natural water samples


Table 4 displays the physicochemical properties of the natural
groundwater samples. In general, the sample from 9 de Julio presented
higher absolute values than those determined in Castelli water sample,
except for pH, Ca2+ and arsenic. Both water samples exhibit arsenic
concentrations exceeding the limit established by the WHO (10 μg L− 1)
(OMS, 2004), and the local limit of 50 μg L− 1 (ANMAT, 2002).
Fig. 6 shows the arsenic sorption capacity of the materials as well as
the removal percentage from naturally contaminated groundwater
samples. Similarly to synthetic water trials, sorption capacities were
higher for MEWC and MCSC than for their respective raw charcoals.
Fig. 4. Mössbauer spectra for MEWC and MCSC obtained at room temperature. Likewise, MEWC presented the highest As removal capacity. The
adsorption efficiency of MCSC resulted satisfactorily for 9 de Julio water
wild rice (Wang et al., 2017b), but higher than for those materials ob­ sample, considering the Argentine limit, but insufficient for Castelli
tained from rice husks (Yang et al., 2008). From a magnetic point of sample, for which the remaining concentration was 62 μg L− 1. Arsenic
view, both materials can be manipulated efficiently by an external concentration after sorption on EWC and CSC remained higher than
magnetic field. WHO and Argentine suggested limits (102 and 97 μg L− 1 for Castelli
Considering the Fe environments revealed by Mössbauer spectros­ sample and 78 and 76 μg L− 1 for 9 de Julio sample). Besides, As
copy, magnetite is the major responsible for the achieved saturation remaining concentration after sorption on MEWC resulted lower than
magnetization values. The fact that both samples exhibited similar Ms the Argentine limit (30 and 16 μg L− 1 for Castelli and 9 de Julio samples,
values, despite the higher content of magnetite in MEWC, could be respectively), but higher than the WHO suggested limit.
related to the different oxidation degree of magnetite (Spessato et al., Given that MCSC had a higher surface area, micropores and particle
2020). size than MEWC, these parameters would not account for the differences
The above results indicate that the use of local eucalyptus wood observed in As sorption. In addition, the total iron concentration esti­
subjected to a relatively easy procedure to achieve Fe particle nucleation mated by Mössbauer showed that there were no significant differences
seems to be adequate for magnetic composite production. between the composites, considering the experimental uncertainty
(Table 2). The distribution of iron among the different detected phases
and the level of oxidation of magnetite may explain the different per­
3.2. Arsenic removal trials formance in the removal essays, in agreement with the different As af­
finity to Fe oxides (Giménez et al., 2007). The higher As sorption
3.2.1. Synthetic water sample capacity of MEWC than MCSC in all cases might be related to both its
Fig. 6 shows the As sorption capacity and removal percentage of
EWC, CSC, MEWC and MCSC from a synthetic water sample. The
Table 3
nucleation of Fe-oxide particles onto charcoals produced a positive ef­
Parameters extracted from the hysteresis loops before and after As removal. Xhf,
fect on the As removal, as reported for magnetic biochars derived from Ms, Mr, and Hc represent the susceptibility of high field, the saturation
bamboo residues (Alchouron et al., 2020), timber industry waste wood magnetization, the remanent magnetization and the coercive field, respectively.
(Navarathna et al., 2019), and pinewood (Wang et al., 2015), among 7
Xhf (10− m3 Ms (Am2 Mr (Am2 Hc (mT)
others. A removal efficiency increase from 7% to 99% for EWC and from kg− 1) kg− 1) kg− 1)
4% to 69% for CSC was observed. In addition, a higher increase was
MEWC 2.6 ± 0.3 11.2 ± 0.5 1.4 0.5 6.3 ± 0.3
observed for EWC than for CSC. The final As concentration after sorption
± −
MCSC 3.8 ± 0.3 10.7 ± 0.5 1.4 ± 0.5 − 5.0 ± 0.2
experiments with EWC and MEWC were 93 μg L− 1 and 1 μg L− 1, MEWC -As 3.0 ± 0.3 10.4 ± 0.5 1.4 ± 0.1 − 5.0 ± 0.2
respectively, while for CSC and MCSC resulted in 96 μg L− 1 and 31 μg MCSC -As 3.0 ± 0.3 9.9 ± 0.4 1.3 ± 0.2 − 5.0 ± 0.2
L− 1. Only in the case of MEWC the As concentration resulted lower than

Table 2
Magnetic and paramagnetic Fe phases assigned after the fittings described in the text. δ, ε and H denote the isomer shift in mm s− 1, the quadrupole shift in mm/s and the
hyperfine magnetic field in Tesla, respectively. RSA is the relative spectral area for each Fe phase (%). Estimation of the total Fe content is also presented.
Paramagnetic phases

Sample Fe3þ(I) Fe3þ(II) Fe (g kg¡1)

δ Δ RSA δ Δ RSA

MCSC 0.10 0.55 5±1 0.68 0.64 5±1 108 ± 11


MEWC 0.32 0.86 6±1 0.37 0.46 5±1 130 ± 15

Magnetic phases
Sample Fe3O4(A) Fe3O4(B) α-FeOOH
δ ε H RSA δ ε H RSA δ ε H RSA

MCSC − 0.10 − 0.07 48.5 22 ± 3 0.62 − 0.04 48.0 53 ± 5 0.49 − 0.12 40.5 15 ± 2
MEWC 0.32 0.00 48.8 50 ± 6 0.51 0.06 45.3 33 ± 5 0.49 − 0.11 36.3 6±1

6
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Fig. 5. a) Hysteresis loops of MCSC (red squares) and MEWC (black circles); b) central part of the hysteresis loops for both samples.

Fig. 6. Arsenic sorption capacity of the four materials for synthetic water (A) (S/L = 5 g L− 1, contact time = 24 h, T = 25 ◦ C, pH = 7.0, Ci = 100 μg L− 1 As(V)) and for
the two naturally contaminated groundwater samples: Castelli (B) and 9 de Julio (C). Removal percentage is expressed in each case, as well as the residual As
concentration after the removal tests. Red values represent concentrations higher than the limits suggested by the WHO and Argentina, blue ones concentrations
higher than WHO limit but lower than the Argentine limit, and green concentrations lower than both, the WHO and Argentine limits.

higher magnetite concentration and the higher oxidation degree of (2010), decreasing the sorption sites available for As(V). In particular,
magnetite. given the chemical similarity of phosphate and arsenate, the presence of
Besides, Q values obtained for the water sample from Castelli resulted the former in the natural water samples could account for the lower
higher than those obtained for 9 de Julio. This could be associated to two removal exhibited (in comparison to synthetic water) In fact, the water
facts: the higher As initial concentration and the lower concentration of sample from 9 de Julio had the highest concentration of phosphate ions
other water constituents (Table 4). and the performance of the magnetic composites was diminished.
It is interesting to highlight that the sorption capacities of the com­ Finally, it is interesting to note that the magnetic response of the
posites were lower in natural samples than in synthetic ones. This could dried samples after the arsenic removal trials did not seem to change
be related to the As sorption mechanism and the presence of other an­ significantly (Table 3). This result is crucial for the later retrieval of the
ions (nitrate, phosphate, sulfate, chloride) in high concentrations in the sorbent.
groundwater samples (Alchouron et al., 2020). These anions could be These results indicate that the use of natural biomass such as euca­
adsorbed on magnetite sites by the mechanism proposed by Zhang et al. lyptus wood as raw material for the development of magnetic charcoal is

7
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Table 4 Table 5
Physicochemical characterization of the natural groundwater samples. Arsenic removal performance of similar composite materials.
Parameter Sample 1 Sample 2 CAA Adsorbent Ci (μg S/L pH T q (μg Reference
Castelli 9 de Julio Guideline L− 1 ) (g (◦ C) g− 1)
L− 1)
pH 8.19 7.81 6.5–8.5
EC (μS cm− 1) 506 1840 – Pine wood char 100 10 3.5 25 775 Mohan et al.
CO2−
3 (mg L )
− 1
nd nd – As (2007)
HCO−3 (mg L− 1) 396.5 693 – (III)
Cl− (mg L− 1) 42.5 167 <350 Oak wood char 100 10 3.5 25 2573 Mohan et al.
SO2−
4 (mg L )
− 1
3.9 7.4 <400 As (2007)
PO3−
4 (μg L )
− 1
23.0 30.5 – (III)
Ca2+ (mg L− 1) 16.8 11.0 – Biochar/γ 5000 2 – RT 243 Zhang et al.
Mg2+ (mg L− 1) 11.3 22.4 – -Fe2O3 As(V) (2013)
Na+ (mg L− 1) 162.5 350 – Bagasse fly ash- 100 6 7.0 RT 16.33 Singh Yadav
K+ (mg L− 1) 7.0 14.5 – iron coated et al. (2014)
Alkalinity (mg CaCO3 L− 1) 325 568 – Sponge iron char 100 6 7.0 RT 16.58 Singh Yadav
Hardness (mg CaCO3 L− 1) 89.3 121 <400 As (V) et al. (2014)
Arsenic (μg L− 1) 107 79 <50 Pinewood/ 1000 2.5 7.0 RT 92 Wang et al.
hematite As (V) (2015)
EC: electrical conductivity; CAA: Argentine Alimentary Codex for potable water; magnetite/ 1000 2 7.0 25 250 Navarathna
nd: not detected. Douglas fir As et al. (2019)
biochar (III)
bamboo 2500 2 7.0 25 3900 Alchouron et al.
possible, especially in places where well-established technologies are
biochar/ As (V) (2020)
not available. magnetite
Table 5 presents sorption capacities, q (μg g− 1), determined on pre­ activated 2500 2 7.0 25 18,000 Alchouron et al.
vious relevant articles. A wide range of q values is observed. The results bamboo As (V) (2020)
achieved in this study are not directly comparable with the ones re­ biochar/
magnetite
ported, given the variable experimental conditions. However, the
MEWC 100 5 7.0 25 21.45 present work
sorption capacity is within the reported range which indicates that the As (V) (synthetic
synthesized composite is a promising material for As removal from water)
naturally polluted waters. MCSC 100 5 7.0 25 14.8 present work
References: RT, Room temperature; Ci, initial As concentration; S/L, As (V) (synthetic
water)
solid/liquid ratio; T, temperature; q, sorption capacity for the consid­
ered initial concentration estimated from the sorption isotherms re­
ported by the authors. waters were reduced from about 100 μg L− 1 to less than 50 μg L− 1, which
It must be highlighted that the immobilization of the retrieved is the Argentine established limit for drinking water.
pollutant is compatible with local ceramics production (Simón et al.,
2019). The saturated As-loaded composite may be included in bricks; the Declaration of competing interest
As presence might contribute to the sintering of the final ceramic and the
decomposition of the charcoal might promote the development of pores, The authors declare that they have no known competing financial
which is a property usually achieved by the incorporation of organic interests or personal relationships that could have appeared to influence
matter into the brick mixture. It is likely that As leaching be negligible the work reported in this paper.
because this type of ceramics are able to immobilize compounds in the
amorphous phases and red clays used in brick production have a high Acknowledgements
iron content. However, further trials have to be performed to fully study
the final product. A. L. Bursztyn Fuentes and F. Barraqué acknowledge CONICET
doctoral fellowship. In addition, A. L. Bursztyn Fuentes also acknowl­
4. Conclusions edges the Ernst Mach Grant-worldwide from the Austrian Agency for
International Cooperation in Education and Research (OeAD-GmbH),
Two charcoal-based composites with magnetic response were syn­ Centre for International Cooperation & Mobility (ICM) which allowed
thesized by simple procedures and tested for arsenic removal from completing characterizations in the University of Vienna, Austria. Cor­
synthetic and naturally contaminated Argentine water samples. The nelia von Baeckmann is acknowledged for TEM images and Andreas
products were carefully characterized by several techniques. Their likely Mautner for XPS measurements. Financial support of Argentine Ministry
magnetic oxide composition could be derived by Mössbauer spectros­ of Science (ANPCyT - PICT 2018-01536), UNLP School of Exact Sciences
copy. Its results indicated that the main Fe oxide formed in both mag­ (11X/847) and CONICET (PUE 066) are gratefully acknowledged by M.
netic composites was magnetite, although goethite and Fe3+ in L. Montes and R.C. Mercader. M. L. Montes, R.C. Mercader, and A. N.
paramagnetic environments -like akaganeite, ferrihydrite or Scian are members of the National Council for Scientific and Techno­
lepidocrocite-might also be present. The iron content estimated with logical Research (CONICET).
Mössbauer spectroscopy was similar in both composites, but the
eucalyptus-derived composite showed the highest magnetite concen­
Appendix A. Supplementary data
tration as well as the highest saturation magnetization, which facilitates
its removal after As sorption with a neodymium magnet. The presence of
Supplementary data to this article can be found online at https://doi.
Fe oxide particles significantly enhanced the arsenic removal ability of
org/10.1016/j.gsd.2021.100585.
the composites, in comparison to the raw charcoals and the magnetic
response of magnetic variants did not change after the arsenic sorption.
References
The magnetic composite synthesis reported here, which is relatively
easy, inexpensive, and makes use of local biomass, gave promising re­ Adib, M., Al-qodah, Z., Ngah, C.W.Z., 2015. Agricultural bio-waste materials as potential
sults since after the composite’s use, arsenic concentrations in naturals sustainable precursors used for activated carbon production : a review. Renew.
Sustain. Energy Rev. 46, 218–235. https://doi.org/10.1016/j.rser.2015.02.051.

8
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Ahmad, I., Ahmad, W., Ahmad, T., 2019. Synthesis and characterization of molecularly Li, H., Dong, X., da Silva, E.B., de Oliveira, L.M., Chen, Y., Ma, L.Q., 2017. Mechanisms of
imprinted magnetite nanomaterials as a novel adsorbent for the removal of heavy metal sorption by biochars: biochar characteristics and modifications. Chemosphere
metals from aqueous solution. J. Mater. Res. Technol. 8, 4239–4252. https://doi. 178, 466–478. https://doi.org/10.1016/j.chemosphere.2017.03.072.
org/10.1016/j.jmrt.2019.07.034. Litter, M.I., Ingallinella, A.M., Olmos, V., Savio, M., Difeo, G., Botto, L., Mónica, E.,
Alchouron, J., Navarathna, C., Chludil, H.D., Dewage, N.B., Perez, F., Barbary, E., Torres, F., Taylor, S., Frangie, S., Herkovits, J., Schalamuk, I., José, M.,
Pittman, C.U., Vega, A.S., Mlsna, T.E., 2020. Assessing South American Guadua Berardozzi, E., García, F.S., Bhattacharya, P., Ahmad, A., 2019. Arsenic in Argentina:
chacoensis bamboo biochar and Fe 3 O 4 nanoparticle dispersed analogues for occurrence, human health, legislation and determination. Sci. Total Environ. 676,
aqueous arsenic (V) remediation. Sci. Total Environ. 706, 135943. https://doi.org/ 756–766. https://doi.org/10.1016/j.scitotenv.2019.04.262.
10.1016/j.scitotenv.2019.135943. Maniammal, K., Madhu, G., Biju, V., 2017. X-ray diffraction line profile analysis of
Almeida, L., Maria, S., Borges, S., Nogueira, P., Fraga, M.A., Telles, S., Oliva, D., nanostructured nickel oxide: shape factor and convolution of crystallite size and
Gustavo, S., Rangel, C., 2017. Methylene blue oxidation over iron oxide supported microstrain contributions. Physica E 85, 214–222. https://doi.org/10.1016/j.
on activated carbon derived from peanut hulls. Catal. Today 289, 237–248. https:// physe.2016.08.035.
doi.org/10.1016/j.cattod.2016.11.036. Mccammon, C., 1995. Mossbauer spectroscopy of minerals. In: Mineral Physics and
American Public Health Association, 2018. Standard Methods for the Examination of Crystallography. A Handbook of Physical Constants, pp. 332–347.
Water and Wastewater [WWW Document]. URL, 5.3.19. https://www.standa Mohan, D., Pittman Jr., C.U., Bricka, M., Smith, F., Yancey, B., Mohammad, J., Steele, P.
rdmethods.org/action/showTopic?taxonomyUri=part&topicCode=part2000. H., Alexandre-Franco, M.F., Gómez-Serrano, V., Gong, H., 2007. Sorption of arsenic,
ANMAT, 2002. Capítulo XII. Código Aliment, Argentino, pp. 7–21. cadmium, and lead by chars produced from fast pyrolysis of wood and bark during
Bardach, A.E., Ciapponi, A., Soto, N., Chaparro, M.R., Calderon, M., Briatore, A., bio-oil production. J. Colloid Interface Sci. 310, 57–73. https://doi.org/10.1016/j.
Cadoppi, N., Tassara, R., Litter, M.I., 2015. Epidemiology of chronic disease related jcis.2007.01.020.
to arsenic in Argentina: a systematic review. Sci. Total Environ. 538, 802–816. Montes, M.L., Rivas, P.C., Taylor, M.A., Mercader, R.C., 2016. Approximate total Fe
https://doi.org/10.1016/j.scitotenv.2015.08.070. content determined by Mossbauer spectrometry: application to determine the
Barraqué, F., Montes, M.L., Fernández, M.A., Mercader, R.C., Candal, R.J., Torres correlation between gamma-ray-emitter activities and total content of Fe phases in
Sánchez, R.M., 2018. Synthesis and characterization of magnetic-montmorillonite soils of the Province of Buenos Aires, Argentina. J. Environ. Radioact. 162–163,
and magnetic-organo-montmorillonite: surface sites involved on cobalt sorption. 113–117. https://doi.org/10.1016/j.jenvrad.2016.05.016.
J. Magn. Magn Mater. 466, 376–384. https://doi.org/10.1016/j. Montes, M.L., Barraqué, F., Fuentes, A.L.B., Taylor, M.A., Mercader, R.C., 2020. Effect of
jmmm.2018.07.052. synthetic beidellite structural characteristics on the properties of beidellite/Fe oxides
Barraqué, F., Montes, M.L., Fernández, M.A., Mercader, R.C., Candal, R.J., Torres magnetic composites as Sr and Cs adsorbent materials. Mater. Chem. Phys. 245,
Sánchez, R.M., 2020. Synthesis of high-saturation magnetization composites by 3–10. https://doi.org/10.1016/j.matchemphys.2020.122760.
montmorillonite loading with hexadecyl trimethyl ammonium ions and magnetite Murad, E., Cashion, J., 2004. Mossbauer Spectroscopy of Environmental Materials and
nucleation for improved effluent sludge handling and dye removal. Appl. Phys. A Their Industrial Utilization. Springer. https://doi.org/10.1007/978-1-4419-9040-2.
126, 736. https://doi.org/10.1007/s00339-020-03834-6. Navarathna, C.M., Karunanayake, A.G., Gunatilake, S.R., Pittman, C.U., Perez, F.,
Barraqué, F., Montes, M.L., Fernández, M.A., Mercader, R.C., Candal, R.J., Torres Mohan, D., Mlsna, T., 2019. Removal of Arsenic ( III ) from water using magnetite
Sánchez, R.M., Marco-Brown, J.L., 2021. Arsenate removal from aqueous solution by precipitated onto Douglas fir biochar. J. Environ. Manag. 250, 109429. https://doi.
montmorillonite and organo-montmorillonite magnetic materials. Environ. Res. 192, org/10.1016/j.jenvman.2019.109429.
110247. https://doi.org/10.1016/j.envres.2020.110247. Oliveira, F.R., Patel, A.K., Jaisi, D.P., Adhikari, S., Lu, H., Khanal, S.K., 2017.
Bartonkova, H., Mashlan, M., Medrik, I., Jancik, D., Zboril, R., 2007. Magnetically Environmental application of biochar: current status and perspectives. Bioresour.
modified bentonite as a possible contrast agent in MRI of gastrointestinal tract. Technol. 246, 110–122. https://doi.org/10.1016/j.biortech.2017.08.122.
Chem. Pap. 61, 413–416. https://doi.org/10.2478/s11696-007-0057-9. OMS, 2004. Guidelines for Drinking-Water Quality, third ed. WHO, Geneva.
Bursztyn Fuentes, A.L., Canevesi, R.L.S., Gadonneix, P., Mathieu, S., Celzard, A., Pecharromán, C., González-Carreño, T., Iglesias, J.E., 1995. The infrared dielectric
Fierro, V., 2020. Paracetamol removal by Kon-Tiki kiln-derived biochar and properties of maghemite, γ-Fe2O3 , from reflectance measurement on pressed
activated carbons. Ind. Crop. Prod. 155, 112740. https://doi.org/10.1016/j. powders. Phys. Chem. Miner. 21–29.
indcrop.2020.112740. Rasband, W., 1997. ImageJ - Image Processing and Analysis in Java.
Cazzeta, A.L., Pezoti, O., Bedin, K.C., Silva, T.L., Paesano Junior, A., Asefa, T., Reddy, D., Lee, S.M., 2014. Magnetic biochar composite: facile synthesis,
Almeida, V.C., 2016. Magnetic activated carbon derived from biomass waste by characterization, and application for heavy metal removal. Colloids Surfaces A
concurrent synthesis: efficient adsorbent for toxic dyes. ACS Sustain. Chem. Eng. 4, Physicochem. Eng. Asp. 454, 96–103. https://doi.org/10.1016/j.
1058–1068. https://doi.org/10.1021/acssuschemeng.5b01141. colsurfa.2014.03.105.
Cychosz, K.A., Thommes, M., 2018. Progress in the physisorption characterization of Regmi, P., Luis, J., Moscoso, G., Kumar, S., Cao, X., Mao, J., Schafran, G., 2012. Removal
nanoporous gas storage materials. Engineering 4, 559–566. https://doi.org/ of copper and cadmium from aqueous solution using switchgrass biochar produced
10.1016/j.eng.2018.06.001. via hydrothermal carbonization process. J. Environ. Manag. 109, 61–69. https://doi.
Danish, M., Ahmad, T., 2018. A review on utilization of wood biomass as a sustainable org/10.1016/j.jenvman.2012.04.047.
precursor for activated carbon production and application. Renew. Sustain. Energy Shaheen, S.M., Niazi, N.K., Hassan, N.E.E., Bibi, I., Wang, H., Tsang, D.C.W., Ok, Y.S.,
Rev. 87, 1–21. https://doi.org/10.1016/j.rser.2018.02.003. Bolan, N., Rinklebe, J., 2019. Wood-based biochar for the removal of potentially
Dhanasekaran, P., Sahu, O., 2020. Arsenate and fluoride removal from groundwater by toxic elements in water and wastewater : a critical review. Int. Mater. Rev. 6608
sawdust impregnated ferric hydroxide and activated alumina (SFAA). Groundw. https://doi.org/10.1080/09506608.2018.1473096.
Sustain. Dev. https://doi.org/10.1016/j.gsd.2020.100490. Shaikh, W.A., Alam, A., Alam, O., Chakraborty, S., Owens, G., Bhattacharya, T., Kumar
Dubinin, M.M., 1989. Fundamentals of the theory of adsorption in micropores of carbon Mondal, N., 2020. Enhanced aqueous phase arsenic removal by a biochar based iron
adsorbents: characteristics of their adsorption properties and microporous nanocomposite. Environ. Technol. Innov. 19, 100936. https://doi.org/10.1016/j.
structures. Carbon N. Y. 27, 457–467. https://doi.org/10.1016/0008-6223(89) eti.2020.100936.
90078-X. Simón, D., Quaranta, N., Medici, S., Costas, A., Cristóbal, A., 2019. Immobilization of Zn
Dunlop, D.J., 1973. Superparamagnetic and single-domain threshold sizes in magnetite. (II) ions from contaminated biomass using ceramic matrices. J. Hazard Mater. 373,
J. Geophys. Res. 78, 1780–1793. https://doi.org/10.1029/JB078i011p01780. 687–697. https://doi.org/10.1016/j.jhazmat.2019.03.123.
Frau, F., Addari, D., Atzei, D., Biddau, R., 2010. Influence of major anions on as ( V ) Singh, N.B., Nagpal, G., Agrawal, S., 2018. Water purification by using adsorbents: a
adsorption by synthetic 2-line ferrihydrite. Kinetic Investigation and XPS Study of review. Environ. Technol. Innov. 11, 187–240. https://doi.org/10.1016/j.
the Competitive Effect of Bicarbonate 25–41. https://doi.org/10.1007/s11270-009- eti.2018.05.006.
0054-4. Singh Yadav, L., Kumar Mishra, B., Kumar, A., Karar Paul, K., 2014. Arsenic removal
Giménez, J., Pablo, J. De, Rovira, M., Duro, L., 2007. Arsenic sorption onto natural using bagasse fly ash-iron coated and sponge iron char. Journal of Environmental
hematite, magnetite, and goethite. J. Hazard Mater. 141, 575–580. https://doi.org/ Chemical Engineering 2 (3), 1467–1473. https://doi.org/10.1016/j.
10.1016/j.jhazmat.2006.07.020. jece.2014.06.019.
Grosvenor, A.P., Kobe, B.A., Biesinger, M.C., Mcintyre, N.S., 2004. Investigation of Spessato, L., Cazetta, A.L., Melo, S., Pezoti, O., Tami, J., Ronix, A., Fonseca, J.M.,
Multiplet Splitting of Fe 2p XPS Spectra and Bonding in Iron Compounds, Martins, A.F., Silva, T.L., Almeida, V.C., 2020. Synthesis of superparamagnetic
pp. 1564–1574. https://doi.org/10.1002/sia.1984. activated carbon for paracetamol removal from aqueous solution. J. Mol. Liq. 300
Han, Z., Sani, B., Mrozik, W., Obst, M., Beckingham, B., Karapanagioti, H.K., Werner, D., https://doi.org/10.1016/j.molliq.2019.112282.
2015. Magnetite impregnation effects on the sorbent properties of activated carbons Thommes, M., Kaneko, K., Neimark, A.V., Olivier, J.P., Rodriguez-Reinoso, F.,
and biochars. Water Res. 70, 394–403. https://doi.org/10.1016/j. Rouquerol, J., Sing, K.S.W., 2015. Physisorption of gases, with special reference to
watres.2014.12.016. the evaluation of surface area and pore size distribution (IUPAC Technical Report).
Hu, Q., Zhu, Y., Hu, B., Lu, S., Sheng, G., 2018. Mechanistic insights into sequestration of Pure Appl. Chem. 87, 1051–1069. https://doi.org/10.1515/pac-2014-1117.
U (VI) toward magnetic biochar: batch, XPS and EXAFS techniques. J. Environ. Sci. Vandenberghe, R.E., Grave, E. De, 2013. Application of mössbauer spectroscopy in earth
70, 217–225. https://doi.org/10.1016/j.jes.2018.01.013. Sciences. In: Yoshida, Y., Langouche, G. (Eds.), Mössbauer Spectroscopy. Springer,
Karunanayake, A.G., Navarathna, C.M., Gunatilake, S.R., Crowley, M., Anderson, R., Berlin, Heidelberg, pp. 91–185. https://doi.org/10.1007/978-3-642-32220-4_3.
Mohan, D., Perez, F., Pittman Jr., C.U., Mlsna, T., 2019. Fe3O4 nanoparticles Wang, S., Gao, B., Zimmerman, A.R., Li, Y., Ma, L., Harris, W.G., Migliaccio, K.W., 2015.
dispersed on Douglas fir biochar for phosphate sorption. ACS Appl. Nano Mater. 2 Removal of arsenic by magnetic biochar prepared from pinewood and natural
(6), 3467–3479. https://doi.org/10.1021/acsanm.9b00430. hematite. Bioresour. Technol. 175, 391–395. https://doi.org/10.1016/j.
Khanna, P.K., Raison, R.J., Falkiner, R.A., 1994. Chemical properties of ash derived from biortech.2014.10.104.
Eucalyptus litter and its effects on forest soils. For. Ecol. Manage. 66, 107–125. Wang, B., Gao, B., Fang, J., 2017a. Recent advances in engineered biochar productions
https://doi.org/10.1016/0378-1127(94)90151-1. and applications. Crit. Rev. Environ. Sci. Technol. 3389 https://doi.org/10.1080/
10643389.2017.1418580.

9
A.L. Bursztyn Fuentes et al. Groundwater for Sustainable Development 14 (2021) 100585

Wang, F., Liu, L.Y., Liu, F., Wang, L.G., Ouyang, T., Chang, C.T., 2017b. Facile one-step montmorillonite, Fe oxides, and hydrothermal carbon: synthesis, characterization
synthesis of magnetically modified biochar with enhanced removal capacity for and pollutants adsorption tests. Materialia 15, 100973. https://doi.org/10.1016/j.
hexavalent chromium from aqueous solution. J. Taiwan Inst. Chem. Eng. 81, mtla.2020.100973.
414–418. https://doi.org/10.1016/j.jtice.2017.09.035. Zhang, Y., Jia, Y., 2014. A facile solution approach for the synthesis of akaganéite ( ␤
Yang, N., Zhu, S., Zhang, D., Xu, S., 2008. Synthesis and properties of magnetic Fe3O4- -FeOOH ) nanorods and their ion-exchange mechanism toward as ( V ) ions. Appl.
activated carbon nanocomposite particles for dye removal. Mater. Lett. 62, 645–647. Surf. Sci. 290, 102–106. https://doi.org/10.1016/j.apsusc.2013.11.007.
https://doi.org/10.1016/j.matlet.2007.06.049. Zhang, S., Li, X.-y., Chen, J.P., 2010. An XPS study for mechanisms of arsenate
Yap, M.W., Mubarak, N.M., Sahu, J.N., Abdullah, E.C., 2017. Microwave induced adsorption onto a magnetite-doped activated carbon fiber. J. Colloid Interface Sci.
synthesis of magnetic biochar from agricultural biomass for removal of lead and 343, 232–238. https://doi.org/10.1016/j.jcis.2009.11.001.
cadmium from wastewater. J. Ind. Eng. Chem. 45, 287–295. https://doi.org/ Zhang, M., Gao, B., Varnoosfaderani, S., Hebard, A., Yao, Y., Inyang, M., 2013.
10.1016/j.jiec.2016.09.036. Preparation and characterization of a novel magnetic biochar for arsenic removal.
Zelaya Soulé, M.E., Barraqué, F., Fernández Morantes, C., Flores, F.M., Fernández, M.A., Bioresour. Technol. 130, 457–462. https://doi.org/10.1016/j.biortech.2012.11.132.
Torres Sánchez, R.M., Montes, M.L., 2021. Magnetic nanocomposite based on

10

You might also like