1 s2.0 S187853522400193X Main

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

[Short title + Author Name - P&H title] 17 (2024) 105791

Contents lists available at ScienceDirect

Arabian Journal of Chemistry


journal homepage: www.ksu.edu.sa

Original article

Waste para-rubber wood ash and iron scrap for the sustainable preparation
of magnetic Fenton catalyst for efficient degradation of tetracycline
Natthanan Rattanachueskul a, Parichart Onsri a, Waralee Watcharin b, Arthit Makarasen c,
Supanna Techasakul c, Decha Dechtrirat c, d, e, *, Laemthong Chuenchom a, f, *
a
Division of Physical Science, Faculty of Science, Prince of Songkla University, Songkhla 90110, Thailand
b
Faculty of Biotechnology (Agro-Industry), Assumption University, Hua Mak Campus, Bangkok 10240, Thailand
c
Laboratory of Organic Synthesis, Chulabhorn Research Institute, Bangkok 10210, Thailand
d
Department of Materials Science, Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
e
Specialized Center of Rubber and Polymer Materials for Agriculture and Industry (RPM), Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
f
Center of Excellence for Innovation in Chemistry, Faculty of Science, Prince of Songkla University, Songkhla 90110, Thailand

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

Keywords: This study introduces a sustainable method of synthesizing a magnetic Fenton catalyst. The use of readily
Fenton reaction available iron scrap waste and para-rubber wood ash from heating and power plants is innovative and cost-
Zero-waste effective. The catalyst is obtained from the precursor materials by pyrolysis. The physicochemical properties
Iron scrap
of the catalyst were characterized to investigate the effects of different iron-to-ash ratios and pyrolysis tem­
Para-rubber wood ash
peratures. The Fenton process was investigated through the degradation of tetracycline (TC). A catalyst produced
with an iron–ash ratio of 1:2 exhibited exceptional performances with a TC removal rate of up to 90 % within 15
min under optimized conditions. The Fenton catalyst also possessed desirable magnetic properties, enabling easy
separation. The stable catalyst could be regenerated and was successfully recycled four times while retaining its
efficiency. This research not only addresses environmental concerns but also highlights the potential value of two
waste materials in the synthesis of advanced Fenton-like catalysts.

1. Introduction advantages of technical feasibility, complete mineralization, high effi­


ciency, and the ability to operate at ambient temperatures and pressures.
Antibiotic resistance acquired from antibiotic contamination poses a One of the more familiar advanced oxidation processes is the Fenton
risk to aquatic ecosystems and human health (Wang & Zhuan, 2020; reaction, which uses H2O2 and an iron-based catalyst. Iron-based ma­
Xiang et al., 2020; Zhang et al., 2018). A substantial volume of anti­ terials such as zero-valent iron (Chen et al., 2023; Jiang et al., 2020; Li
microbial agents is released annually into the environment, with tetra­ et al., 2023; Yoon & Bae, 2019), iron oxides (Khataee et al., 2017; Li
cycline (TC) emerging as one of the most frequently identified antibiotic et al., 2022; Radoń et al., 2019; Rashmishree et al., 2023) and iron
compounds within aquatic ecosystems. (Wang et al., 2021). The elimi­ compounds (Agú et al., 2020; Soufi et al., 2021) have been extensively
nation of TC from aquatic environments by adsorption (Dutta & Mala, studied for the Fenton-like degradation of TC. Incorporating magnetic
2020; Krasucka et al., 2021; Priya & Radha, 2017) and coagulation characteristics into Fenton catalysts during the preparation process is an
(Choi et al., 2008; Saitoh et al., 2017) creates TC-contaminated solids approach that increases their applicability and facilitates the separation
which are considered hazardous waste. The suitable procedure for the of catalysts. The acquired magnetic properties can be tuned to allow the
collection and disposal of this waste has become a crucial issue since rapid and effective separation of the catalyst with an external magnet
unsuitable disposal may discharge the contamination back into the and suitably magnetized iron particles can serve as active sites for
environment. This issue can be addressed by the use of advanced Fenton-like reactions. This combination of tunable magnetic properties
oxidation processes, which break down organic contaminants with and catalytic functionality offers significant advantages in the field of
highly reactive radicals. These promising chemical processes offer wastewater treatment and pollutant degradation but the commercial

Peer review under responsibility of King Saud University.


* Corresponding authors.
E-mail addresses: [email protected] (D. Dechtrirat), [email protected] (L. Chuenchom).

https://doi.org/10.1016/j.arabjc.2024.105791
Received 1 March 2024; Accepted 9 April 2024
Available online 11 April 2024
1878-5352/© 2024 The Author(s). Published by Elsevier B.V. on behalf of King Saud University. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

chemicals used as precursors increase production costs. 2.2. Preparation of MPRA catalyst
Alternative precursors that are less expensive than the traditional
iron salts are available in abundance and could lead to a sustainable A solution of iron scrap was prepared by dissolving iron scrap in
method of reducing costs. Steel, for instance, is used in large quantities concentrated HCl (1 g/mL) under stirring for 20 min. The obtained so­
in fields such as transportation, building construction and lution was diluted with DI water at a ratio of 1:3. Various precursor
manufacturing. The large volume of steel used creates waste materials, mixtures were then prepared by mixing 15 g of PRA under stirring for 1 h
known collectively as iron scrap, that can be recycled by smelting at high with different amounts of iron precursor solution to give weight ratios of
temperatures, which is a costly and energy-intensive process (Koch iron scrap–PRA of 1:1, 1:2, and 1:4. In the optimal catalyst ultimately
et al., 2002). It is, however, feasible to use iron scrap in a more used to impregnate PRA, 15 g of PRA were mixed with 300 mL of the
economical way as a precursor for advanced, novel, composite materials diluted iron scrap solution to give an iron scrap–PRA weight ratio of 1:2
(Babar et al., 2018; Chandane et al., 2019; Deganello et al., 2019; and a total iron concentration of ~ 6250 mg/L. The various mixtures
Fatimah et al., 2023; Priyadarshini et al., 2023) produced in processes were dried at 110 ◦ C and dried samples were pyrolyzed for 1.5 h under
that uphold the principles of sustainable chemistry. nitrogen gas at different temperatures from 600 to 800 ◦ C, ramped at
To improve the efficiency of iron-based catalysts, many studies have 2 ◦ C/min to obtain magnetic iron oxide@PRA composites (MPRA).
attached the catalysts to solid supports such as silica (Do et al., 2018; The obtained PRA catalysts were labeled MPRAx-y-T, where x-y
Farhadian et al., 2021; Fatimah et al., 2019; Munoz et al., 2015; Šuligoj represents the weight ratio of iron scrap–PRA and T is the pyrolysis
et al., 2023; Wang & Zhuan, 2020) and carbon-based materials (Chen temperature, and a control catalyst, denoted Scrap-800, was prepared
et al., 2020; Li et al., 2020; Lu et al., 2023; Priyadarshini et al., 2023; Tao using the same procedure without the inclusion of PRA. All prepared
et al., 2023; Wang & Zhuan, 2020; Wu et al., 2024; Yoo et al., 2017). materials were ground to a fine powder with a particle size of around 20
These works were motivated by the stability, low toxicity, widespread μm measured by field emission scanning electron microscope (FESEM)
availability and cost-effectiveness of the supporting materials. Para- analysis. Five replications of each sample were produced and the overall
rubber wood ash (PRA) is the residual solid waste produced by the preparation yield is shown in Table S1.
combustion of para-rubber wood for electricity generation in wood
substitute industries (Hasan et al., 2000; Masae, 2013; Saosee et al., 2.3. Characterization of MPRA catalysts
2020). In Thailand, the world’s major producer of para-rubber wood,
this practice produces a large amount of ash (Masae, 2013), which is The surface morphology of the particles was examined by field
classed as particulate matter and causes air pollution and respiratory emission scanning electron microscopy integrated with energy-
diseases (Le Blond et al., 2017; Rodríguez-Díaz et al., 2015). Many at­ dispersive X-ray spectroscopy (FESEM-EDX, Apreo, FEI). Pore volume,
tempts have been made to find a use for PRA, whether as a component of pore size, and surface area were determined using a nitrogen adsorp­
building materials (Madurwar et al., 2014; Sales & Lima, 2010; Xu et al., tion/desorption analyzer (ASAP2460, Micromeritics). Prior to N2
2018) or as an acidic soil neutralizer to increase agricultural output adsorption/desorption examination, the samples were degassed at
(Purnomo et al., 2018; Webber et al., 2017). PRA is also a promising 120 ◦ C for 16 h. Functional groups and iron species at the surface of the
material as a catalyst support, due to its large surface area, stability and catalyst were identified by Fourier transform infrared spectroscopy (FT-
silica-based structure (Abdul Mutalib et al., 2020; Fatimah et al., 2019; IR, Spectrum GX, Perkin Elmer) from 4000 to 400 cm− 1 utilizing a KBr
Ma et al., 2020). Therefore, iron scrap and PRA are potentially intriguing pellet. Surface functionalities were quantitatively and qualitatively
starting materials for the synthesis of a magnetic Fenton catalyst. analyzed by X-ray photoelectron spectroscopy (XPS, AXIS Ultra DLD,
In this work, a magnetic PRA catalyst (MPRA) is obtained by the Kratos Analytical Ltd.). The crystal structure and iron speciation of
simple pyrolysis of PRA previously impregnated with an iron scrap so­ samples was analyzed by X-ray diffraction analysis (XRD, Philips, X’Pert
lution. PRA possesses favorable chemical properties that make it a MPD), with a 2θ scanning range of 5◦ to 90◦ and a step size of 0.05◦ . The
promising catalyst support for iron oxide. Notably, its inert silica surface magnetization of samples was studied at 298 K using a vibrating sample
can hinder the aggregation of magnetic particles in liquid media during magnetometer (VSM, Lake Shore). X-ray fluorescence spectrometry
synthesis, thereby enhancing stability and mitigating concerns related to (XRF, Zetium, PANalytical) was employed to identify the chemical
biodegradability and toxicity (Thangaraj et al., 2019). At the same time, composition of PRA.
hydroxyl groups on the surface of PRA facilitate strong interactions The stability of the as-prepared catalysts was assessed by quantifying
between the support and iron oxide particles (Vinayagam et al., 2022). total iron ion contents using ICP-OES (AVIO 500, Perkin Elmer) to
Additionally, synthesizing a Fenton catalyst from PRA and iron scrap not determine iron release from the composites. In the catalytic experi­
only adds value to these waste materials but also contributes to the ments, the concentration of TC in solution (C0 = 80 mg/L) after the
development of sustainable catalytic processes. Harnessing these experiments was measured by UV–Vis spectrophotometry. The TC so­
chemical advantages holds significant potential for advancing catalytic lution (C0 = 80 mg/L) treated with MPRA1-2-800 was also analyzed by
systems. liquid chromatography-mass spectrometry (ESI– mode, Agilent Tech­
nologies). A TOC analyzer (multi N/C 3100, Analytik Jena) was used to
2. Experimental quantify the degree of mineralization by measuring the total organic
carbon (TOC) removal.
2.1. Materials and chemicals
2.4. Experimental procedure for TC degradation
Sodium hydroxide (97.0 %) was from RCI Labscan (Bangkok,
Thailand); hydrochloric acid (36.5–38.0 %) from J.T. Baker (Phillips­ TC degradation by Fenton reaction: Using 250 mL conical flasks, 0.2 g
burg, NJ, USA); and tetracycline hydrochloride (>98.0 %) from TCI of MPRA was mixed with 200 mL of 80 mg/L TC solution. The flasks
America (Portland, OR, USA). PRA was obtained from the wood sub­ were placed in a thermostat water bath (Model TOL09-FTSH- 01, SCI­
stitute plant of Panel Plus Public Company Limited in Songkhla, FINETECH) at a temperature of 28 ± 2 ◦ C, and shaken at 200 rpm for 1 h
Thailand, and was used as a catalyst support. Iron scrap was obtained while maintaining the pH of the mixture at 3.7. After 1 h, the TC solution
from a local junkyard in Songkhla Province. Experiments were carried was sampled to measure concentration, and H2O2 (1–10 mM) was added
out with AR-grade chemicals of high purity and deionized (DI) water. to initiate the reaction. The influence of solution pH on the degradation
of TC was tested by altering pH with 0.1 M HCl or 0.1 M NaOH. These
experiments were conducted in darkness. After specific time periods
(0–4 h), a magnet was used to separate the MPRA from solution, and TC

2
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

concentration was determined. TC concentration before and after the scrap powder was red. The magnetic properties of MPRA1-1–800,
initiation of the Fenton reaction was measured by UV–Vis spectropho­ MPRA1-2-800, MPRA1-4–800, and Scrap-800 were also measured by
tometry (UV 2600, Shimadzu) via a calibrated approach at λ = 357 nm. VSM. All four samples produced hysteresis loops within the ± 10 kOe
Degradation efficiency was calculated from formula (1): range at a temperature of 300 K (Fig. 1), indicating their ferromagnetic
nature. It was noted that the saturation magnetization values (Ms) of the
(C0 − Ct )
%Efficiency = × 100 (1) sample series were well correlated to the weight of iron scrap in the
C0
samples. The magnetization of MPRA1-4–800 was only 0.31 emu/g. The
where Ct denotes the concentration of TC at specific time periods, sample was barely attracted to the magnet due to the relatively low
measured in mg/L. amount of iron scrap used in the preparation process. All the other
Catalyst regeneration: A 0.2 g sample was combined with 200 mL of an samples were attracted to the magnet. Scrap-800 was prepared using
80 mg/L TC solution. The mixture was then shaken in a thermostat only iron scrap but the magnetization of Scrap-800 was lower than
water bath for 4 h at 28 ± 2 ◦ C. The pH value was carefully controlled at MPRA1-1–800, indicating that the use of PRA as a catalyst support
3.7. Subsequently, 5 mM of H2O2 was added to the mixture. An external prevented the aggregation of magnetic iron oxide particles and main­
magnet was used to isolate the catalyst. The collected catalyst was then tained magnetic properties (Castillo et al., 2021; Chen et al., 2022;
heated in air at 300 ◦ C to be reactivated. A further four experimental Thangaraj et al., 2019). Thus, the inclusion of PRA improved the mag­
runs were then carried out using the same sample of magnetic catalyst. netic properties of the composite. After pyrolysis, the MPRA samples
were darker in color. The degree of darkening was proportional to the
3. Results and discussion ratio of iron scrap in the sample and was attributed to the presence of
iron oxide, probably magnetite, in the materials (Fig. S2).
3.1. Optimization of the preparation conditions
3.2. Characterization of MPRA catalysts
The pyrolysis temperature was preliminarily optimized in the range
of 600–800 ◦ C under a N2 atmosphere. The preliminary results showed The FESEM-EDX results for all samples in Fig. S3 show that the
that the samples pyrolyzed at 600 and 700 ◦ C (MPRA1-2–600 and particles of MPRA were rougher than the particles of PRA. The increased
MPRA1-2–700) displayed very low magnetization values (Fig. 1 and surface roughness of MPRA was due to the iron oxide particles precip­
Table S1) and cannot be attracted by an external magnet. Only samples itated on the PRA surface. EDX mapping of all MPRA samples revealed a
pyrolyzed at 800 ◦ C were attracted to a magnet (Fig. S1A-C). In com­ uniform distribution of iron particles, along with calcium, oxygen, and
parison to the samples calcined at 600 and 700 ◦ C, the one calcined at silicon elements. The average iron contents of the MPRA samples were
800 ◦ C (MPRA1-2-800) showed stronger peak intensities at 35.6◦ (3 1 1) 6.6 wt% for MPRA-1–4-800, 14.7 wt% for MPRA-1–2-800, and 43.3 wt%
and 62.5◦ (4 4 0) corresponding to magnetite (Fe3O4, JCPDS No. 01–076- for MPRA-1–1-800 (Table S2). These results exhibited a strong corre­
5948). Specifically, only the diffraction pattern of MPRA1-2-800 lation with their respective magnetization values.
revealed a distinct peak of magnetite at 30.33◦ (2 2 0) as shown in The presence of iron oxide particles was also confirmed by XRD
Fig. S1D. These findings indicate a greater extent of magnetite formation analysis (Fig. 2). The patterns of magnetite (Fe3O4, JCPDS No. 01–076-
at 800 ◦ C than at 600 ◦ C and 700 ◦ C. Therefore, the pyrolysis temper­ 5948) and hematite (Fe2O3, JCPDS No. 01–079-1741) were observed,
ature of 800 ◦ C was suitable and was used for the synthesis of further indicating that mixed phases of iron were formed during pyrolysis due to
samples. the uncontrolled mole ratio between Fe2+ and Fe3+ in the iron scrap
The effect of the iron scrap–PRA weight ratio on the physical solution (Yi et al., 2019). Both forms of iron oxide were also detected in
appearance of the catalyst was investigated. Images of MPRA1-1–800, the Scrap-800 sample. Additionally, silica (SiO2, JCPDS No. 03–065-
MPRA1-2-800, MPRA1-4–800, and Scrap-800 are shown in Fig. S2. 0466) from PRA was detected in all MPRA samples. Notably, in agree­
Their yields are listed in Table S1. PRA was dark grey in color while iron ment with the EDX results, calcite (CaCO3, JCPDS No. 01–078-4614)
was detected in the PRA sample, but the maximum intensity of the
calcite peaks was very low compared to the maximum intensity of the
silica peak at 26.64◦ 2θ. Additionally, sylvite (KCl, JCPDS No. 01–075-
0296) was also detected in all MPRA samples. XRF spectrometry was
used to confirm the presence of calcite and silica in the PRA sample. The
XRF analysis revealed that the composition of PRA included 48.48 %wt
SiO2, 22.03 %wt CaO, 6.96 %wt K2O, and 5.81 %wt Al2O3, along with
3.48 %wt CHNO. This analysis indicated that silica and calcite were the
main constituents of PRA.
The functional groups on the surface of the catalysts were examined
using FTIR spectroscopy. Detailed band assignments were listed in
Table S3. The absorption bands around 3435 cm− 1 were attributed to
the stretching vibrations of O − H of adsorbed water and silanol groups
(Si-OH). The absorption bands in the range of 1700–1400 cm− 1 indi­
cated stretching vibrations of C-O bonds of carbonate groups and C = C
bonds, suggesting the presence of CaCO3 and carbon materials in the
samples. Additionally, the peak at ~ 1076 cm− 1 was attributed to the
stretching vibrations of Si-O-Si bonds, indicating the presence of silica in
PRA (Fig. S4). Therefore, PRA contained silica, CaCO3, and carbon
materials. After the preparation process, a significant amount of CaCO3
had been eliminated by HCl, and a new absorption band was present in
the spectra of MPRA samples at 563 cm− 1, which was associated with
Fig. 1. The magnetic hysteresis loops for MPRA1-1–800, MPRA1-2-800, the stretching vibrations of Fe-O bonds from iron oxides.
MPRA1-4–800, MPRA1-2–600, MPRA1-2–700 and Scrap-800 samples. The Fig. 3 presents the wide-scan XPS spectrum of MPRA1-2-800 and the
inset displays a zoomed-in view of the magnetic hysteresis loops for the MPRA- high-resolution C 1 s, O 1 s, Si 2p, and Fe 2p spectra of the same sample.
1–4–800, MPRA1-2–600, and MPRA1-2–700 samples. The XPS survey scan confirmed the presence of C, O, Si, and Fe in

3
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

Subsequently, iron oxide particles were precipitated and deposited


onto the surface of PRA, leading to the acquisition of magnetic
properties.

3.3. Effect of reaction conditions on the catalytic activity of MPRA

This section focuses on examining the MPRA composite as a het­


erogeneous Fenton catalyst, specifically investigating the catalytic
degradation of TC. First, an 80 ppm TC solution at pH 3.7 was shaken in
darkness for 1 h on MPRA at a catalyst concentration of 1 g/L. The re­
action was then initiated by introducing 5 mM of H2O2. MPRA1-2-800
showed the highest removal efficiency at ~ 95 % (Fig. 4A). MPRA1-
1–800 removed ~ 66 % of TC after a 4-hour reaction time, even though
the iron content in MPRA1-1–800 was higher. The Scrap-800 control
removed only ~ 40 % of TC. In the absence of the catalyst (only 5 mM of
H2O2 with TC), only 5.66 % of TC was removed. This finding shows that
the presence of PRA as a catalyst support considerably enhanced
removal efficiency. The degree of Fe leaching was 1.65, 2.74, and 4.45
mg/L for MPRA1-1–800, MPRA1-2-800, and Scrap-800, respectively.
This result indicated the lower stability of Scrap-800 due to the lack of
support material. Although MPRA1-1–800 had a higher magnetization
value with better stability than MPRA1-2-800, MPRA1-2-800 showed a
significantly higher efficiency toward TC removal. BET surface area
analysis (Table S5) revealed that the BET surface area of MPRA1-2-800
was bigger, which was a beneficial characteristic because it allowed TC
molecules greater accessibility to active sites. As a result, MPRA1-2-800
was selected as the most promising Fenton catalyst for the catalytic
degradation of TC. This choice was based on its capacity to remove TC,
its satisfactory level of iron leaching, and its strong magnetic
characteristics.
The pH level is a crucial determinant in the Fenton-like reaction (Lai
et al., 2019; Wang et al., 2016). The effect of initial solution pH on the
catalytic degradation of TC using MPRA1-2-800 is shown in Fig. 4B. At a
pH solution of 3.7, up to 90 % of TC was removed within 15 min, and
95.61 % in 4 h. This result is consistent with the fact that an acidic so­
Fig. 2. XRD patterns of MPRA1-1–800, MPRA1-2-800, MPRA1-4–800, Scrap- lution is often the suitable condition for most Fenton-like catalysts
800, and PRA samples. (Subramanian et al., 2016). Moreover, TC degradation remained nearly
the same when the pH was increased from 3.7 to 9.0, indicating that this
MPRA1-2-800, and the high-resolution C 1 s and O 1 s spectra showed Fenton system can be applied in a wide pH range, therefore overcoming
the presence of oxygenated functional groups that included carboxylic, the costly requirement of classical Fenton systems for an acidic condi­
hydroxyl, and ether groups. These groups were responsible for the hy­ tion that also results in secondary pollution of iron sludge. However, the
drophilic properties of the catalyst. The peak at around 530 eV binding effectiveness of the reaction notably decreased as the pH of the solution
energy was attributed to magnetite (Fe3O4) (Zhu et al., 2016). This approached 12 due to a reduction in the reactivity of the radicals and the
finding was consistent with the presence of the band associated with Fe- precipitation of iron. (Subramanian et al., 2016). Hence, the pH of 3.7
O bonding at around 559 cm− 1 in the FTIR spectra and the XRD data. was selected as the optimal pH for subsequent experiments. The results
Moreover, CaCO3 and SiO2 were found in all MPRA samples. These were were analyzed using the BMG kinetic model to assess the kinetics of TC
the inorganic compounds left from the pyrolysis process. The Fe 2p catalytic degradation (Behnajady et al., 2007; Sidney Santana et al.,
spectrum confirmed the presence of Fe(II) and Fe(III) in MPRA1-2-800. 2019) as in Equation (2),
More specifically, the peaks at 710.66 and 723.76 eV verified the t
( ) = m + bt (2)
presence of Fe(II) in magnetite form. Table S4 contains the compre­ C
1−
hensive peak assignments for the deconvoluted XPS peaks derived from C0

the XPS spectra, along with their corresponding binding energies. The
high-resolution Fe 2p spectrum confirmed the presence of both Fe2+ and where C is the TC concentration at time t (min) and C0 is the initial
Fe3+ (36.5:63.5 %at) from the magnetite structure in the materials. concentration of TC at time t = 0. Additionally, b and m are two char­
These findings corroborated the presence of magnetite particles on the acteristic constants associated with the reaction kinetics and oxidation
surface of PRA. capacities. The kinetics of TC degradation on MPRA1-2-800 fitted well
During the synthesis of MPRA by pyrolysis, the residual carbon in with the BMG kinetic model with an R2 close to 1 (Fig. S5). The recip­
PRA reduced Fe3+ to Fe2+ and then both Fe2+ and Fe3+ formed rocal of the b value is the theoretical maximum TC fraction that can be
magnetite and hematite. The process is described in reactions R(1)–R(3) removed, which is equal to the maximum oxidation capacity at the end
(Liu et al., 2020; Wang et al., 2021). of the reaction. Meanwhile, 1/m is the initial rate of TC removal. In this
work, the value of 1/b was 0.944, which is close to 1 (maximum ca­
4FeCl3 + C + 18H2O → 4 FeCl2⋅4H2O + CO2 + 4HCl (R1) pacity) and the value of 1/m was 1.491, which is very high. This model
Fe 2+
+ 2Fe 3+ -
+8OH → Fe3O4 + 4H2O (R2) suggests that the oxidation of TC by Fenton-like processes, utilizing iron
as a catalyst, typically proceeds in two stages: a rapid initial stage and a
4Fe3O4 + O2 → 6Fe2O3 (R3) slower secondary stage. The initial rapidity of the process was ascribed

4
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

Fig. 3. (A) The wide-scan XPS and (B-E) high-resolution C 1 s, O 1 s, Si 2p, and Fe 2p spectra of MPRA1-2-800.

Fig. 4. (A) Results from the preliminary test of the catalytic heterogeneous Fenton-like reaction and the catalytic degradation of TC: (B) effect of initial pH; (C) effect
of H2O2 concentration; and (D) scavenger test. (Conditions: TC initial concentration (C0) is 80 mg/L, pH is 3.7 (natural), 5 mM H2O2, catalyst dosage of 1 g/L, and
temperature at 28 ± 2 ◦ C).

5
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

to the reaction between dissolved iron ions and H2O2, while the subse­ cleavage and mineralization into inorganic compounds such as CO2 and
quent slower rate was caused by the accumulation of Fe3+ and the H2O.
limited regeneration of Fe2+ by H2O2. This finding was consistent with The mechanism of the Fenton-like reaction on the MPRA1-2-800
the rate constant of the reaction between Fe2+ and H2O2, which is catalyst is proposed as shown in R6-R10 (Ameta et al., 2018; Gan
generally higher than that of the reaction between Fe3+ and H2O2 in et al., 2020; Tang & Wang, 2018). Initially, the iron oxide on the surface
homogeneous reactions. When these two competing processes consume of MPRA1-2-800 dissolves rapidly in the acidic condition. Subsequently,
H2O2, the concentration of H2O2 in the solution is gradually reduced to a iron ions (Fe2+/Fe3+) engage in a reaction with H2O2, resulting in the
low level, the retardation stage begins, and the small amount of oxidant formation of hydroxyl radicals (⋅OH), perhydroxyl radicals (⋅OOH), and
in the solution becomes a limiting factor. superoxide radicals (⋅O–2), as shown in R6-R8. To maintain the catalytic
A succession of trials was carried out using H2O2 concentrations of 0, cycle, Fe2+ is regenerated through the reduction of Fe3+, as described by
1, 2.5, 5, and 10 mM to investigate the influence of different quantities R9-R10.
on the catalytic degradation of TC (Fig. 4C). An absence of H2O2 in the
(R6)
[Ox]
adsorption experiment restricted the elimination of TC to 22.8 %, owing Fe2+ + H2 O2 ̅̅→ Fe3+ + OH − + • OHfree
to the shortage of adsorption sites and the low BET surface area.
Nonetheless, as the concentration of H2O2 increased, the effectiveness of [Red]
Fe3+ + H2 O2 ̅̅
̅→ Fe2+ + • OOH + H + (R7)
MPRA1-2-800 increased. After 4 h, 5 mM H2O2 had removed 95.61 % of
TC. However, the removal of TC in 10 mM H2O2 was relatively un­
(R8)
[Red]
Fe3+ + H2 O2 ̅̅
̅→ Fe2+ + ⋅O−2 + 2H +
changed because the surplus H2O2 competed with TC for ⋅OH in the
following reactions (Lai et al., 2019):
(R9)
[Red]
Fe3+ + • OOH ̅̅
̅→ Fe2+ + O2 + H +
H2O2 + ⋅OH → HO2⋅ + H2O (R4)
(R10)
[Red]
HO2⋅ + ⋅OH → O2 + H2O (R5) Fe3+ + ⋅O−2 ̅̅
̅→ Fe2+ + O2

Thus, the optimal concentration of H2O2 was 5 mM. The dissolution of the iron ions into the solution, due to the acidic
To identify the free radicals involved in the Fenton-like reaction, free condition, was considered inevitable, leading to the leaching of Fe into
radical scavengers were incorporated into the reaction solution. Hy­ the solution. However, the Fe concentration determined by ICP-OES was
droxyl radicals (⋅OH) were trapped with 0.3 M tert-butyl alcohol (TBA), 2.74 mg/L, which does not exceed the maximum concentration
and superoxide radicals (⋅O–2) were trapped with 20 mM p-benzoquinone permitted in industrial areas of Thailand (<10 mg/L).
(BQ) (Hou et al., 2016; Ma et al., 2018). TBA significantly hindered TC
degradation but 20 mM BQ had only a small effect on TC degradation 3.4. The regeneration of MPRA1-2-800
(Fig. 4D). It was concluded that ⋅OH served as the main reactive oxygen
species in this heterogeneous Fenton-like system and to confirm that the Catalyst regeneration was studied because it is a crucial criterion of
Fenton-like reaction was induced by ⋅OH radicals, 0.3 M isopropanol practicality. The reusability of MPRA1-2-800 was tested by degrading
(IPA) was added to the reaction to differentiate between adsorbed and 200 mL of 80 mg/L TC solution for 4 h, at pH 3.7 with the addition of 5
unbound hydroxyl radicals. Generally, TBA can suppress both free hy­ mM of H2O2. The used sample was then regenerated by heat treatment at
droxyl radicals in solution (⋅OHfree) and hydroxyl radicals adsorbed onto 300 ◦ C for 1 h in air to expose more iron oxide as active sites. The
the catalyst surface (⋅OHads), whereas IPA can only suppress ⋅OHfree. The reactivated catalyst was then used again to degrade TC. As shown in
degradation of TC was substantially inhibited by the addition of IPA, and Fig. 5A, the degradation of TC on the fresh catalyst was 95.61 %. After
removal efficiency was comparable to that achieved in the presence of four cycles of use, degradation efficiency was 67.40 % which is an
TBA. These results suggested that ⋅ OHfree contributed to the degradation acceptable level. The degree of Fe leaching was only high during the first
of TC, while ⋅ OHads and ⋅ O–2 had very minor impacts. use and, afterwards was less than 0.5 mg/L until the fourth cycle
The total organic carbon (TOC) content of the TC solution was (Fig. 5B). Despite the Fe leaching, the magnetic properties of the catalyst
determined both before and after catalytic degradation. This evaluation were maintained, indicating the good stability of the catalyst. Moreover,
provided additional evidence that the primary mechanism for TC the high-resolution XPS spectra of the used catalyst in Fig. S8 showed
removal was catalytic degradation and not adsorption. The removal of that the surface chemistry of used MPRA1-2-800 after four cycles was
TOC was found to be 90.81 %, indicating the mineralization of TC similar to the surface chemistry of fresh catalyst (Fig. 3). Interestingly,
during the Fenton-like reaction. Moreover, a solution of 80 ppm TC was the Fe(II)–Fe(III) ratio from Fe 2p in Table S4 was changed from 1:2 to
analyzed by LC-MS before and after degradation on MPRA1-2-800. The 1:1, indicating that circulation of Fe(II) and Fe(III) occurred during the
loss of the m/z peak at 443.16 after degradation (Fig. S6) suggested that Fenton-like reaction. This behavior could enhance degradation effi­
the intermediates quickly underwent oxidation to inorganic compounds. ciency because the rate constant of the Fenton reaction is higher when
The LC-MS analysis was used to explore the mechanism of degra­ Fe2+ is involved (R6) than Fe3+ (R7) (Wang et al., 2021). Furthermore,
dation. A possible degradation pathway was proposed in Fig. S7. The the used MPRA1-2-800 was characterized by VSM analysis (Fig. 5C). The
peak at m/z 325.18 has been associated previously with TC degradation saturation magnetization of used MPRA1-2-800 was lower but the
and confirms the presence of the intermediate P4 (C19H18O5) (Huang catalyst was still attracted to the magnet. The decrease in magnetic
et al., 2020), suggesting that fresh TC molecules (m/z = 443) first strength may be attributed to a combination of factors, including the
degraded into P1 (m/z = 461) via hydroxylation at position 14, and then leaching of iron, and the air oxidation during the air-calcination
P1 was further decomposed into P2 (m/z = 429) by losing the methyl in regeneration process. Following four cycles of catalyst use, the %Fe
the amino group. After that, the amide of P2 was attacked by reactive content from the EDX analysis decreased slightly, from 14.70 to 13.50 %
radicals, which resulted in the production of P3 (m/z = 371). P3 (Table S2). Fe leaching should thus not be a primary cause of the loss of
continuously converted to smaller structures of P4 (m/z = 325) through magnetic strength. The primary cause should be the air-calcination
dehydroxylation and the oxidation of the –NHCH3 group at positions 1 process, since oxygen gas in the air may gradually oxidize the mag­
and 2, respectively. Subsequently, the double bond at position 1 was netic phase to the non-magnetic phase.
attacked, leading to a ring cleavage reaction, and an aldehyde group was
formed at position 17, producing P5 (m/z = 272) which was further 3.5. Comparative analyses
decomposed to P6 (m/z = 179). Finally, the molecules P6 were further
oxidized into smaller products via subsequent reactions such as ring To assess the performance of MPRA1-2-800 as a Fenton-like catalyst

6
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

Fig. 5. The regeneration experiments determined (A) the degradation of TC on MPRA1-2-800 over four cycles; (B) the amount of iron leaching after each cycle
obtained from ICP-OES analysis; and (C) the magnetic hysteresis loop of the fresh MPRA1-2-800 sample and the used MPRA1-2-800 sample. (Conditions: TC initial
concentration (C0) of 80 mg/L, pH of 3.7 (natural), 5 mM H2O2, catalyst dosage of 1 g/L, and temperature at 28 ± 2 ◦ C).

for TC degradation, a comparison was made with previous studies iron scrap by pyrolysis. The catalyst was used for the degradation of
(Table S6). In terms of efficiency, the MPRA1-2-800 removed 90 % of TC tetracycline in a Fenton system. Extensive characterizations confirmed
within 15 min and up to 94.35 % in 4 h. Most previous studies used quite the uniform distribution of iron oxide species on the para-rubber wood
low concentrations or a small volume of TC solution, leading to a high ash support material. The most effective catalyst removed more than 90
removal percentage. This is reflected by the actual amount of removed % of tetracycline within 15 min under optimal conditions (1 g/L catalyst
TC in mg which confirms the high performance of the proposed Fenton- dose, 5 mM H2O2, pH 3.7, and 80 mg/L tetracycline solution), while
like catalyst compared to other catalysts. Although some studies used demonstrating excellent magnetic properties and good stability. The
hybrid systems between Fenton and other techniques such as UV-Fenton removal percentage exceeded those previously reported, highlighting
degradation, ultrasound-assisted Fenton-like degradation, and photo- the potential of the proposed catalyst for the Fenton-like degradation of
Fenton degradation, their actual removal values were low compared antibiotics in wastewater. Scavenging experiments indicated that hy­
to this work. Also, the H2O2 concentration is a key factor that affects droxyl radicals were the major reactive species in the catalytic Fenton
operating costs. The H2O2 concentration used in this work was quite low system. LC–MS analysis identified reaction intermediates, enabling the
compared to previous reports, leading to better cost effectiveness for a proposal of a possible pathway of tetracycline degradation. Addition­
real operation. Additionally, the high TOC removal in the present work ally, the catalyst could be recycled with minimal Fe leaching, retaining
indicates the effective degradation of TC by mineralization and oxida­ stability over four cycles. Further research might prioritize the expan­
tion. Secondary contamination from intermediates is therefore minimal, sion of synthesis on a larger scale, investigating its effectiveness against
which cannot be said for processes in some other works that reported different contaminants, and evaluating its real-world performance to
low TOC removal values. Furthermore, the Fenton-like catalyst in this promote sustainable wastewater treatment.
work was prepared by a simple pyrolysis method using only para-rubber
wood ash and iron scrap as precursors, which are waste industrial ma­ CRediT authorship contribution statement
terials. On the contrary, previous studies described complicated pro­
cedures and used harmful chemicals, in particular some highly toxic Natthanan Rattanachueskul: Formal analysis, Investigation,
heavy metals were utilized to accelerate Fe(II)/Fe(III) circulation, which Methodology, Validation, Writing – original draft, Writing – review &
must lead to secondary contamination. Therefore, MPRA1-2-800 is a editing. Parichart Onsri: Formal analysis, Investigation. Waralee
promising Fenton-like catalyst for water remediation. Watcharin: Investigation. Arthit Makarasen: Investigation. Supanna
Techasakul: Supervision, Visualization, Writing – review & editing.
4. Conclusion Decha Dechtrirat: Conceptualization, Funding acquisition, Methodol­
ogy, Project administration, Validation, Writing – original draft, Writing
A magnetic catalyst was synthesized from para-rubber wood ash and – review & editing. Laemthong Chuenchom: Conceptualization,

7
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

Formal analysis, Funding acquisition, Methodology, Validation, Writing copper. J. Clean. Prod. 172, 1243–1253. https://doi.org/10.1016/j.
jclepro.2017.10.246.
– original draft, Writing – review & editing.
Dutta, J., Mala, A.A., 2020. Removal of antibiotic from the water environment by the
adsorption technologies: A review. Water Sci. Technol. 82 (3), 401–426. https://doi.
org/10.2166/wst.2020.335.
Declaration of competing interest Farhadian, N., Liu, S., Asadi, A., Shahlaei, M., Moradi, S., 2021. Enhanced heterogeneous
Fenton oxidation of organic pollutant via fe-containing mesoporous silica
composites: A review. J. Mol. Liq. 321, 114896 https://doi.org/10.1016/j.
The authors declare that they have no known competing financial
molliq.2020.114896.
interests or personal relationships that could have appeared to influence Fatimah, I., Amaliah, S.N., Andrian, M.F., Handayani, T.P., Nurillahi, R., Prakoso, N.I.,
the work reported in this paper. Wicaksono, W.P., Chuenchom, L., 2019. Iron oxide nanoparticles supported on
biogenic silica derived from bamboo leaf ash for rhodamine B photodegradation.
Sustain. Chem. Pharm. 13, 100149 https://doi.org/10.1016/j.scp.2019.100149.
Acknowledgements Fatimah, I., Yanti, I., Wijayanti, H.K., Ramanda, G.D., Sagadevan, S., Tamyiz, M.,
Doong, R.-A., 2023. One-pot synthesis of Fe3O4/NiFe2O4 nanocomposite from iron
rust waste as reusable catalyst for methyl violet oxidation. Case Stud. Chem.
The authors would like to thank the Kasetsart University Research Environ. Eng. 8, 100369 https://doi.org/10.1016/j.cscee.2023.100369.
and Development Institute (grant No. FF(KU)27.67) for financial sup­ Gan, Q., Hou, H., Liang, S., Qiu, J., Tao, S., Yang, L., Yu, W., Xiao, K., Liu, B., Hu, J.,
port. N. Rattanachueskul also thanks the Graduate School, Prince of Wang, Y., Yang, J., 2020. Sludge-derived biochar with multivalent iron as an
efficient Fenton catalyst for degradation of 4-chlorophenol. Sci. Total Environ. 725,
Songkla University, for PSU-PhD. scholarship (Contract No.
138299 https://doi.org/10.1016/j.scitotenv.2020.138299.
PSU_PHD2562-003) and financial support, and the Division of Physical Hasan, S., Hashim, M.A., Gupta, B.S., 2000. Adsorption of Ni(SO4) on Malaysian rubber-
Science, Faculty of Science, Prince of Songkla University, Thailand, for wood ash. Bioresour. Technol. 72 (2), 153–158. https://doi.org/10.1016/S0960-
8524(99)00101-7.
support with facilities. L. Chuenchom acknowledges partial support
Hou, L., Wang, L., Royer, S., Zhang, H., 2016. Ultrasound-assisted heterogeneous Fenton-
from the Center of Excellence for Innovation in Chemistry (PERCH-CIC), like degradation of tetracycline over a magnetite catalyst. J. Hazard. Mater. 302,
Ministry of Higher Education, Science, Research, and Innovation, 458–467. https://doi.org/10.1016/j.jhazmat.2015.09.033.
Thailand. The authors are also indebted to Thomas Duncan Coyne and Huang, X., Zhu, N., Mao, F., Ding, Y., Zhang, S., Liu, H., Li, F., Wu, P., Dang, Z., Ke, Y.,
2020. Enhanced heterogeneous photo-Fenton catalytic degradation of tetracycline
Dr. Titilope John Jayeoye for English proof-reading and editing. over yCeO2/Fh composites: Performance, degradation pathways, Fe2+ regeneration
and mechanism. Chem. Eng. J. 392, 123636 https://doi.org/10.1016/j.
cej.2019.123636.
Appendix A. Supplementary material
Jiang, W., Dionysiou, D.D., Kong, M., Liu, Z., Sui, Q., Lyu, S., 2020. Utilization of formic
acid in nanoscale zero valent iron-catalyzed Fenton system for carbon tetrachloride
Supplementary data to this article can be found online at https://doi. degradation. Chem. Eng. J. 380, 122537 https://doi.org/10.1016/j.
cej.2019.122537.
org/10.1016/j.arabjc.2024.105791.
Khataee, A., Gholami, P., Vahid, B., 2017. Catalytic performance of hematite
nanostructures prepared by N2 glow discharge plasma in heterogeneous Fenton-like
References process for acid red 17 degradation. J. Ind. Eng. Chem. 50, 86–95. https://doi.org/
10.1016/j.jiec.2017.01.035.
Koch, G.H., Brongers, M.P.H., Thompson, N.G., Virmani, Y.P., Payer, J.H. 2002.
Abdul Mutalib, A.A., Ibrahim, M.L., Matmin, J., Kassim, M.F., Mastuli, M.S., Taufiq-
Corrosion Costs and Preventive Strategies in the United States: Report by CC
Yap, Y.H., Shohaimi, N.A.M., Islam, A., Tan, Y.H., Kaus, N.H.M., 2020. SiO2-rich
Technologies Laboratories, Inc. to Federal Highway Administration (FHWA);
sugar cane bagasse ash catalyst for transesterification of palm oil. Bioenergy Res. 13,
National Technical Information Service: Alexandria, VA.
986–997. https://doi.org/10.1007/s12155-020-10119-6.
Krasucka, P., Pan, B., Sik Ok, Y., Mohan, D., Sarkar, B., Oleszczuk, P., 2021. Engineered
Agú, U.A., Mendieta, S.N., Gerbaldo, M.V., Crivello, M.E., Casuscelli, S.G., 2020. Highly
biochar – A sustainable solution for the removal of antibiotics from water. Chem.
active heterogeneous fenton-like system based on cobalt ferrite. Ind. Eng. Chem. 59
Eng. J. 405, 126926 https://doi.org/10.1016/j.cej.2020.126926.
(4), 1702–1711. https://doi.org/10.1021/acs.iecr.9b04042.
Lai, C., Huang, F., Zeng, G., Huang, D., Qin, L., Cheng, M., Zhang, C., Li, B., Yi, H., Liu, S.,
Ameta, R., K. Chohadia, A., Jain, A., Punjabi, P.B. 2018. Chapter 3 - Fenton and photo-
Li, L., Chen, L., 2019. Fabrication of novel magnetic MnFe2O4/bio-char composite
Fenton processes. in: Adv. Oxid. Processes Wastewater Treat. (Eds.) S.C. Ameta, R.
and heterogeneous photo-Fenton degradation of tetracycline in near neutral pH.
Ameta, Academic Press, pp. 49-87.
Chemosphere 224, 910–921. https://doi.org/10.1016/j.chemosphere.2019.02.193.
Babar, S., Gavade, N., Shinde, H., Mahajan, P., Lee, K.H., Mane, N., Deshmukh, A.,
Le Blond, J.S., Woskie, S., Horwell, C.J., Williamson, B.J., 2017. Particulate matter
Garadkar, K., Bhuse, V., 2018. Evolution of waste iron rust into magnetically
produced during commercial sugarcane harvesting and processing: A respiratory
separable g-C3N4–Fe2O3 photocatalyst: an efficient and economical waste
health hazard? Atmos. Environ. 149, 34–46. https://doi.org/10.1016/j.
management approach. ACS Appl. Nano Mater. 1 (9), 4682–4694. https://doi.org/
atmosenv.2016.11.012.
10.1021/acsanm.8b00936.
Li, X., Cui, K., Guo, Z., Yang, T., Cao, Y., Xiang, Y., Chen, H., Xi, M., 2020. Heterogeneous
Behnajady, M.A., Modirshahla, N., Ghanbary, F., 2007. A kinetic model for the
Fenton-like degradation of tetracyclines using porous magnetic chitosan
decolorization of C.I. acid yellow 23 by Fenton process. J. Hazard. Mater. 148 (1),
microspheres as an efficient catalyst compared with two preparation methods.
98–102. https://doi.org/10.1016/j.jhazmat.2007.02.003.
Chem. Eng. J. 379, 122324 https://doi.org/10.1016/j.cej.2019.122324.
Castillo, J., Vargas, V., Macero, D., Le Beulze, A., Ruiz, W., Bouyssiere, B., 2021. One-step
Li, H., Shi, B., Fu, X., Zhang, H., Yang, H., 2023. Preparation and application of red mud-
synthesis of SiO2 α− Fe2O3 / Fe3O4 composite nanoparticles with magnetic properties
based zero-valent iron heterogeneous Fenton catalyst: A new idea for red mud
from rice husks. Phys. B: Condens. Matter. 605, 412799 https://doi.org/10.1016/j.
recycling. J. Environ. Chem. Eng. 11 (3), 109998 https://doi.org/10.1016/j.
physb.2020.412799.
jece.2023.109998.
Chandane, P., Ladke, J., Jori, C., Deshmukh, S., Zinjarde, S., Chakankar, M., Hocheng, H.,
Li, J., You, J., Wang, Z., Zhao, Y., Xu, J., Li, X., Zhang, H., 2022. Application of α-Fe2O3-
Jadhav, U., 2019. Synthesis of magnetic Fe3O4 nanoparticles from scrap iron and use
based heterogeneous photo-Fenton catalyst in wastewater treatment: A review of
of their peroxidase like activity for phenol detection. J. Environ. Chem. Eng. 7 (3),
recent advances. J. Environ. Chem. Eng. 10 (5), 108329 https://doi.org/10.1016/j.
103083 https://doi.org/10.1016/j.jece.2019.103083.
jece.2022.108329.
Chen, L., Costa, E., Kileti, P., Tannenbaum, R., Lindberg, J., Mahajan, D., 2022.
Liu, Y., Huang, J., Xu, H., Zhang, Y., Hu, T., Chen, W., Hu, H., Wu, J., Li, Y., Jiang, G.,
Sonochemical synthesis of silica-supported iron oxide nanostructures and their
2020. A magnetic macro-porous biochar sphere as vehicle for the activation and
application as catalysts in Fischer-Tropsch synthesis. Micro 2, 632–648.
removal of heavy metals from contaminated agricultural soil. Chem. Eng. J. 390,
Chen, W.-H., Huang, J.-R., Lin, C.-H., Huang, C.-P., 2020. Catalytic degradation of
124638 https://doi.org/10.1016/j.cej.2020.124638.
chlorpheniramine over GO-Fe3O4 in the presence of H2O2 in water: The synergistic
Lu, Y., Qin, X., Wang, K., Chen, S., Quan, X., 2023. Accelerated Fe(II) regeneration on Fe-
effect of adsorption. Sci. Total Environ. 736, 139468 https://doi.org/10.1016/j.
doped oxidized carbon nanotube enabling highly-efficient electro-Fenton process for
scitotenv.2020.139468.
pollutants removal. Sep. Purif. Technol. 320, 124196 https://doi.org/10.1016/j.
Chen, J.-Q., Zhou, G.-N., Ding, R.-R., Li, Q., Zhao, H.-Q., Mu, Y., 2023. Ferrous ion
seppur.2023.124196.
enhanced Fenton-like degradation of emerging contaminants by sulfidated
Ma, X., Cheng, Y., Ge, Y., Wu, H., Li, Q., Gao, N., Deng, J., 2018. Ultrasound-enhanced
nanosized zero-valent iron with pH insensitivity. J. Hazard. Mater. 459, 132229
nanosized zero-valent copper activation of hydrogen peroxide for the degradation of
https://doi.org/10.1016/j.jhazmat.2023.132229.
norfloxacin. Ultrason. Sonochem. 40, 763–772. https://doi.org/10.1016/j.
Choi, K.-J., Kim, S.-G., Kim, S.-H., 2008. Removal of antibiotics by coagulation and
ultsonch.2017.08.025.
granular activated carbon filtration. J. Hazard. Mater. 151 (1), 38–43. https://doi.
Ma, C., Jia, S., Yuan, P., He, Z., 2020. Catalytic ozonation of 2, 2′-methylenebis (4-
org/10.1016/j.jhazmat.2007.05.059.
methyl-6-tert-butylphenol) over nano-Fe3O4@cow dung ash composites:
Deganello, F., Joshi, M., Liotta, L.F., La Parola, V., Marcì, G., Pantaleo, G., 2019.
Optimization, toxicity, and degradation mechanisms. Environ. Pollut. 265, 114597
Sustainable recycling of insoluble rust waste for the synthesis of iron-containing
https://doi.org/10.1016/j.envpol.2020.114597.
perovskite-type catalysts. ACS Omega 4 (4), 6994–7004. https://doi.org/10.1021/
Madurwar, M., Mandavgane, S., Ralegaonkar, R., 2014. Use of sugarcane bagasse ash as
acsomega.8b03522.
brick material. Curr. Sci. 117, 1044–1051.
Do, Q.C., Kim, D.-G., Ko, S.-O., 2018. Catalytic activity enhancement of a Fe3O4@SiO2
yolk-shell structure for oxidative degradation of acetaminophen by decoration with

8
N. Rattanachueskul et al. Arabian Journal of Chemistry 17 (2024) 105791

Masae, M.S., L.; Kongsong, P.; Phoempoon, P.; Rawangwong, S.; Sririkun, W. 2013. Tao, S., Liang, S., Chen, X., Zhu, Y., Yu, W., Hou, H., Hu, J., Xiao, K., Yuan, S., Yang, J.,
Application of rubber wood ash for removal nickel and copper from aqueous 2023. Enhanced sludge dewatering by PDMDAAC coupled with Fenton-like reaction
solution. Environ. Nat. Resour. J. 11(No.2), 17-27. initiated by Fe-rich sludge biochar with in-situ generation of H2O2: Fe/C structure as
Munoz, M., de Pedro, Z.M., Casas, J.A., Rodriguez, J.J., 2015. Preparation of magnetite- an electron shuttle. Resour. Conserv. Recycl. 198, 107184 https://doi.org/10.1016/
based catalysts and their application in heterogeneous Fenton oxidation – A review. j.resconrec.2023.107184.
Appl. Catal. b: Environ. 176–177, 249–265. https://doi.org/10.1016/j. Thangaraj, B., Jia, Z., Dai, L., Liu, D., Du, W., 2019. Effect of silica coating on Fe3O4
apcatb.2015.04.003. magnetic nanoparticles for lipase immobilization and their application for biodiesel
Priya, S.S., Radha, K.V., 2017. A review on the adsorption studies of tetracycline onto production. Arab. J. Chem. 12 (8), 4694–4706. https://doi.org/10.1016/j.
various types of adsorbents. Chem. Eng. Commun. 204 (8), 821–839. https://doi. arabjc.2016.09.004.
org/10.1080/00986445.2015.1065820. Vinayagam, R., Pai, S., Murugesan, G., Varadavenkatesan, T., Narayanasamy, S.,
Priyadarshini, M., Ahmad, A., Ghangrekar, M.M., 2023. Efficient upcycling of iron scrap Selvaraj, R., 2022. Magnetic activated charcoal/Fe2O3 nanocomposite for the
and waste polyethylene terephthalate plastic into Fe3O4@C incorporated MIL-53 adsorptive removal of 2,4-dichlorophenoxyacetic acid (2,4-D) from aqueous
(Fe) as a novel electro-Fenton catalyst for the degradation of salicylic acid. Environ. solutions: Synthesis, characterization, optimization, kinetic and isotherm studies.
Pollut. 322, 121242 https://doi.org/10.1016/j.envpol.2023.121242. Chemosphere 286, 131938. https://doi.org/10.1016/j.chemosphere.2021.131938.
Purnomo, C.W., Respito, A., Sitanggang, E.P., Mulyono, P., 2018. Slow release fertilizer Wang, C., Sun, R., Huang, R., Wang, H., 2021. Superior fenton-like degradation of
preparation from sugar cane industrial waste. Environ. Technol. Inno. 10, 275–280. tetracycline by iron loaded graphitic carbon derived from microplastics: Synthesis,
https://doi.org/10.1016/j.eti.2018.02.010. catalytic performance, and mechanism. Sep. Purif. Technol. 270, 118773 https://
Radoń, A., Łoński, S., Warski, T., Babilas, R., Tański, T., Dudziak, M., Łukowiec, D., doi.org/10.1016/j.seppur.2021.118773.
2019. Catalytic activity of non-spherical shaped magnetite nanoparticles in Wang, N., Zheng, T., Zhang, G., Wang, P., 2016. A review on Fenton-like processes for
degradation of Sudan I, rhodamine B and methylene blue dyes. Appl. Surf. Sci. 487, organic wastewater treatment. J. Environ. Chem. Eng. 4 (1), 762–787. https://doi.
1018–1025. https://doi.org/10.1016/j.apsusc.2019.05.091. org/10.1016/j.jece.2015.12.016.
Rashmishree, K.N., Bhaskar, S., Shri Hari, S., Thalla, A.K. 2023. Green synthesis of Wang, J., Zhuan, R., 2020. Degradation of antibiotics by advanced oxidation processes:
laterite iron-based nanocatalysts using Psidium guajava and Macaranga peltata plant An overview. Sci. Total Environ. 701, 135023 https://doi.org/10.1016/j.
extract for its catalytic application in Fenton’s oxidation of triclosan. Clean Technol. scitotenv.2019.135023.
Environ. Policy. 10.1007/s10098-023-02507-1. Webber, C., Jr, P., Spaunhorst, D., Petrie, E. 2017. Impact of sugarcane bagasse ash as an
Rodríguez-Díaz, J., García, J., Sánchez, L., Silva, M., Silva, V., Arteaga-Pérez, L., 2015. amendment on the physical properties, nutrient content and seedling growth of a
Comprehensive Characterization of sugarcane bagasse ssh for its use as an adsorbent. certified organic greenhouse growing media. J. Agric. Sci. 9, 1, 10.5539/jas.v9n7p1.
Bioenergy Res. 8, 1885–1895. https://doi.org/10.1007/s12155-015-9646-6. Wu, C., Guo, T., Chen, Y., Tian, Q., Zhang, Y., Huang, Z., Hu, H., Gan, T., 2024. Facile
Saitoh, T., Shibata, K., Fujimori, K., Ohtani, Y., 2017. Rapid removal of tetracycline synthesis of excellent Fe3O4@starch-derived carbon photo-Fenton catalyst for
antibiotics from water by coagulation-flotation of sodium dodecyl sulfate and poly tetracycline degradation: Rapid Fe3+/Fe2+ circulation under visible light condition.
(allylamine hydrochloride) in the presence of Al(III) ions. Sep. Purif. Technol. 187, Sep. Purif. Technol. 329, 125174 https://doi.org/10.1016/j.seppur.2023.125174.
76–83. https://doi.org/10.1016/j.seppur.2017.06.036. Xiang, Y., Huang, Y., Xiao, B., Wu, X., Zhang, G., 2020. Magnetic yolk-shell structure of
Sales, A., Lima, S.A., 2010. Use of Brazilian sugarcane bagasse ash in concrete as sand ZnFe2O4 nanoparticles for enhanced visible light photo-Fenton degradation towards
replacement. Waste Manage. 30 (6), 1114–1122. https://doi.org/10.1016/j. antibiotics and mechanism study. Appl. Surf. Sci. 513, 145820 https://doi.org/
wasman.2010.01.026. 10.1016/j.apsusc.2020.145820.
Saosee, P., Sajjakulnukit, B., Gheewala, S.H., 2020. Life cycle assessment of wood pellet Xu, Q., Ji, T., Gao, S.-J., Yang, Z., Wu, N., 2018. Characteristics and applications of sugar
production in Thailand. Sustainability 12 (17). https://doi.org/10.3390/ cane bagasse ash waste in cementitious materials. Mater. 12, 39. https://doi.org/
su12176996. 10.3390/ma12010039.
Sidney Santana, C., Nicodemos Ramos, M.D., Vieira Velloso, C.C., Aguiar, A., 2019. Yi, Y., Tu, G., Zhao, D., Tsang, P.E., Fang, Z., 2019. Biomass waste components
Kinetic evaluation of dye decolorization by Fenton processes in the presence of 3- significantly influence the removal of Cr(VI) using magnetic biochar derived from
hydroxyanthranilic ccid. Int. J. Environ. Res. Public Health 16 (9). https://doi.org/ four types of feedstocks and steel pickling waste liquor. Chem. Eng. J. 360, 212–220.
10.3390/ijerph16091602. https://doi.org/10.1016/j.cej.2018.11.205.
Soufi, A., Hajjaoui, H., Elmoubarki, R., Abdennouri, M., Qourzal, S., Barka, N., 2021. Yoo, S.H., Jang, D., Joh, H.-I., Lee, S., 2017. Iron oxide/porous carbon as a
Spinel ferrites nanoparticles: Synthesis methods and application in heterogeneous heterogeneous Fenton catalyst for fast decomposition of hydrogen peroxide and
Fenton oxidation of organic pollutants – A review. Appl. Surf. Sci. Adv. 6, 100145 efficient removal of methylene blue. J. Mater. Chem. a. 5 (2), 748–755. https://doi.
https://doi.org/10.1016/j.apsadv.2021.100145. org/10.1039/C6TA07457J.
Subramanian, V., Ordomsky, V.V., Legras, B., Cheng, K., Cordier, C., Chernavskii, P.A., Yoon, S., Bae, S., 2019. Novel synthesis of nanoscale zerovalent iron from coal fly ash and
Khodakov, A.Y., 2016. Design of iron catalysts supported on carbon–silica its application in oxidative degradation of methyl orange by Fenton reaction.
composites with enhanced catalytic performance in high-temperature Fischer- J. Hazard. Mater. 365, 751–758. https://doi.org/10.1016/j.jhazmat.2018.11.073.
Tropsch synthesis. Catal. Sci. Technol. 6 (13), 4953–4961. https://doi.org/10.1039/ Zhang, Y., Shi, J., Xu, Z., Chen, Y., Song, D., 2018. Degradation of tetracycline in a
C6CY00060F. schorl/H2O2 system: Proposed mechanism and intermediates. Chemosphere 202,
Šuligoj, A., Trendafilova, I., Maver, K., Pintar, A., Ristić, A., Dražić, G., Abdelraheem, W. 661–668. https://doi.org/10.1016/j.chemosphere.2018.03.116.
H.M., Jagličić, Z., Arčon, I., Logar, N.Z., Dionysiou, D.D., Novak Tušar, N., 2023. Zhu, X., Qian, F., Liu, Y., Matera, D., Wu, G., Zhang, S., Chen, J., 2016. Controllable
Multicomponent Cu-Mn-Fe silica supported catalysts to stimulate photo-Fenton-like synthesis of magnetic carbon composites with high porosity and strong acid
water treatment under sunlight. J. Environ. Chem. Eng. 11 (5), 110369 https://doi. resistance from hydrochar for efficient removal of organic pollutants: An overlooked
org/10.1016/j.jece.2023.110369. influence. Carbon 99, 338–347. https://doi.org/10.1016/j.carbon.2015.12.044.
Tang, J., Wang, J., 2018. Fenton-like degradation of sulfamethoxazole using Fe-based
magnetic nanoparticles embedded into mesoporous carbon hybrid as an efficient
catalyst. Chem. Eng. J. 351, 1085–1094. https://doi.org/10.1016/j.cej.2018.06.169.

You might also like