ArticlePDF Available

ScienceDirect Photo-Fenton Degradation of Methyl Orange Using Hematite (α-Fe 2 O 3 ) of Various Morphologies

Authors:

Abstract

In this work, the effects of type of additives on the morphology and corresponding photocatalytic property of various hematite (α-Fe2O3) nanostructures were investigated. α-Fe2O3 croissant-like structures, urchin-like structures, and textured microspheres were formed by hydrothermal treatment at 120 °C for 6 h in the presence of NaCl, Na2SO4 , and Na2C2O4 as additives, respectively. After heat treatment in air, the photocatalytic activity of the α-Fe2O3 powders was assessed by degrading methyl orange (MO) under UV-C lamp with hydrogen peroxide (H2O2) as activator. The urchin-like α-Fe2O3 hierarchical structures exhibited the best photocatalytic behavior with a 76.5% removal of initial 2.5 ppm MO after 2 h irradiation. This is attributed to the high surface area of the urchin-like morphology which provides more active sites for degradation of MO. The reaction kinetics correspond well to the Langmuir-Hinshelwood model, which indicates that adsorption of MO on the surface of catalyst occurs as a monolayer.
Available online at www.sciencedirect.com
ScienceDirect
Materials Today: Proceedings 22 (2020) 248–254
www.materialstoday.com/proceedings
1876-6102 © 2019 Elsevier Ltd. All rights reserved.
Peer-review under responsibility of the scientific committee of the 2018 2nd International Conference on Nanomaterials and Biomaterials.
2018 2nd International Conference on Nanomaterials and Biomaterials, ICNB 2018,
10–12 December 2018, Barcelona, Spain
Photo-Fenton Degradation of Methyl Orange Using Hematite
(α-Fe2O3) of Various Morphologies
Anthea Marika G. Domacena, Christian Laurence E. Aquino,
Mary Donnabelle L. Balela*
Sustainable Electronic Materials Group, University of the Philippines, Diliman, Quezon City 1101, Philippines
Abstract
In this work, the effects of type of additives on the morphology and corresponding photocatalytic property of various hematite (α-
Fe2O3) nanostructures were investigated. α-Fe2O3 croissant-like structures, urchin-like structures, and textured microspheres were
formed by hydrothermal treatment at 120 °C for 6 h in the presence of NaCl, Na2SO4, and Na2C2O4 as additives, respectively.
After heat treatment in air, the photocatalytic activity of the α-Fe2O3 powders was assessed by degrading methyl orange (MO)
under UV-C lamp with hydrogen peroxide (H2O2) as activator. The urchin-like α-Fe2O3 hierarchical structures exhibited the best
photocatalytic behavior with a 76.5% removal of initial 2.5 ppm MO after 2 h irradiation. This is attributed to the high surface
area of the urchin-like morphology which provides more active sites for degradation of MO. The reaction kinetics correspond
well to the Langmuir-Hinshelwood model, which indicates that adsorption of MO on the surface of catalyst occurs as a
monolayer.
© 2019 Elsevier Ltd. All rights reserved.
Peer-review under responsibility of the scientific committee of the 2018 2nd International Conference on Nanomaterials and
Biomaterials.
Keywords: photo-fenton; hematite; photocatalysis, urchin-like nanostructures
* Corresponding author. Tel.: +63-02-9818500 local 3132; fax: +63-02-9818500 local 3164.
E-mail address: mlbalela1@up.edu.ph
A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254 249
1. Introduction
Dyes are used to provide color to various materials such as papers, textiles, and plastics [1]. In the past, dyes were
made in small scale from natural sources like plants and insects. When synthetic dyes were discovered, their use
became more prevalent due to their low cost and ease of production. However, synthetic dyes are composed of
hazardous compounds, which cannot be easily degraded due to their stability [2-3]. There are currently around
10,000 dyes and pigments in production, yielding more than 7x108 kg per year worldwide [2]. This results to a
significant amount of discharged colored wastewater, containing around 10-20% of the used dyes [2-3]. The textile
industry, by itself, consumes around 80-200 m3 of water per ton of product and produces 1650 m3 of wastewater
each day. Consequently, it discharges around 280,000 tons of dyes in its effluent per year [4].
Conventional methods of treating dye-laden wastewater that are currently being employed include coagulation
and flocculation which require physical separation methods through screening due to the production of sludges.
Photocatalytic degradation is a promising method in treating these contaminated wastewaters [5-7]. The process is
considerably easier to handle and does not produce secondary sludges. Photocatalytic degradation is the
decomposition of a substance using a semiconductor photocatalyst by harnessing light energy to accelerate the
reaction. In the photodegradation of dyes, light energy excites the electrons of the catalyst thereby producing
hydroxyl radicals which can oxidize organic substances, such as dyes, causing degradation [4].
Hematite (α-Fe2O3) is a promising semiconductor photocatalyst that has the advantage of being abundant, cheap,
and environment-friendly [8-12]. It is the most thermodynamically stable iron oxide with a band gap of about 2.1
eV, which theoretically allows it to absorb light energy on a wide range of wavelengths. However, its photocatalytic
efficiency is limited by its very fast photo-excited state lifetime (10-12 s) [13] and short diffusion length (<10 nm)
[14,15]. Morphological engineering and doping have been effective in addressing these issues by providing shorter
paths for the electron-hole pairs to migrate to the surface and inhibiting charge recombination [6, 13, 16].
Solution-based methods offer a simple and economical way of synthesizing α-Fe2O3 nanostructures.
Hydrothermal synthesis, in particular, is also more advantageous as it only involves the use of water as solvent
compared to other wet chemical methods that employ potentially harmful organic substances. α-Fe2O3 with varying
morphologies have been successfully prepared by hydrothermal process [6,11,16]. However, the effect of various
morphologies in the photocatalytic degradation of dyes, specifically methyl orange, has not been investigated to the
best of our knowledge. In this work, the effect of different additives on the morphology and structure of α-Fe2O3 was
investigated. Assessment of the photocatalytic property of the as-synthesized powders was also done through
degradation of methyl orange in the presence of UV light.
2. Methodology
2.1. Materials
Chemical reagents in the study were used without further purification. Anhydrous iron chloride (FeCl3; HiMedia
Laboratories; 98% purity) was used as Fe source. The additives were sodium chloride (NaCl; EMD Millipore; 99.5%
purity), sodium sulfate (Na2SO4; Ajax Finechem; 99% purity), and sodium oxalate (Na2C2O4; B. E. Chemicals;
assumed to be of 99% purity). Absolute ethanol (C2H6O; Univar; Analytical Grade) was used for washing of
precipitates. Meanwhile, methyl orange (C14H14N3NaO3S; Loba Chemie; 95% purity) was used as model pollutant.
Hydrogen peroxide (H2O2; Rhea, 10 volumes) was used as activator for α-Fe2O3 during photocatalysis.
2.2. Synthesis
2 mmol of anhydrous FeCl3 was mixed with 2 mmol of sodium salt (NaCl, Na
2SO4, Na2C2O4) in 40 mL of
deionized (DI) water. The solutions were magnetically stirred at 400 rpm for 30 min until complete dissolution and
then transferred to a Teflon-lined stainless-steel autoclave reactor. The reactor was heated at 120 °C for 6 h. After
cooling naturally, the powders were collected through centrifugation at 5000 rpm for 30 min and washed with DI
water three times followed by absolute ethanol. Drying of powders was done at 80 °C for 1 h before annealing in air
250 A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254
at 400 °C for 2 h, with a ramp rate of 2 °C/min. The samples were labelled as Samples A, B, and C for the powders
synthesized with NaCl, Na2SO4, and Na2C2O4, respectively.
2.3. Photocatalytic Degradation
The photocatalytic properties of the α-Fe2O3 microstructures were investigated through the photocatalytic
degradation of methyl orange. 30 mg of the α-Fe2O3 samples was added to 30 mL of 5 ppm methyl orange. The
mixtures were magnetically stirred for 30 min in the dark to allow for adsorption equilibrium before the addition of
30 μL of H2O2. Photocatalytic degradation was done by irradiation of the mixtures for 2 h under two 10 W UV-C
lamps (Sankyo Denki G10T8, λ=253.7 nm) inside a blackbox set-up. Filtration and centrifugation were used to
separate solids from the solution. The sample which exhibited the greatest degradation capability was used for
kinetic study. To investigate the kinetics of the degradation, 30 mg of the chosen sample was added to different
concentrations (2.5, 5.0, and 7.5 ppm) of methyl orange solution. The mixtures were similarly stirred in the dark for
30 min. Then, 30 μL H2O2 was added to each mixture before irradiation. Aliquots of about 2 mL each were obtained
at set time intervals (0, 15, 30, 45, 60, 75, 90, 105, 120 min.) from each mixture.
2.4 Characterization
The synthesized α-Fe2O3 powders were observed in a scanning electron microscope (SEM, Jeol JSM-5310). The
phase composition and crystal structure of the products were characterized by X-ray diffraction (XRD, Shimadzu
7000) using Cu Kα radiation. The concentration of methyl orange was determined using a UV-Visible
spectrophotometer (Ocean Optics). The absorbance of the MO solutions was taken at a peak wavelength of 465 nm.
3. Results and Discussion
3.1. Structural and Morphological Characterization
Fig. 1 shows the XRD patterns of α-Fe2O3 structures prepared with different additives after annealing at 400 °C.
All patterns show major peaks at 2θ = 33.2, 35.7, 49.5, 54.1, 62.5, 64.1° which correspond to the (104), (110), (024),
(116), (214), and (300) peaks of α-Fe2O3. The diffraction peaks of all samples were found to correspond to
rhombohedral α-Fe2O3 (PDF No. 84-0307). Difference in peak intensities can be observed from the XRD patterns,
indicating the influence of additives on the crystallinity of the α-Fe2O3 structures. The XRD peaks of sample A
(NaCl) appear sharper compared to samples B (Na2SO4) and C (Na2C2O4) as seen in Fig. 1a. Additionally, the
maximum peak is indexed to the (104) plane, which corresponds well with the standard diffraction pattern of α-
Fe2O3. On the other hand, the (110) plane was the strongest peak in the diffraction pattern of the sample with
Na2SO4 [11,16]. For sample C, the (104) and (110) planes exhibit almost similar intensity. The deviation of the
intensity ratios from the standard diffraction pattern may be attributed to the apparent anisotropic growth of the
structures as shown in Fig 2 (b-c). The formation of nanorods as seen in Fig. 2 (b) suggests a preferential growth in a
particular direction. Peak broadening was observed for samples B and C. This indicates smaller crystallite sizes
compared to sample A. Using the Scherrer equation, the crystallite size for samples A, B, and C were calculated to
be about 33.95, 21.30, and 14.60 nm respectively. In addition, the XRD pattern of sample A show an extra minor
peak at 2θ=44.0°. This is attributed to NaCl (PDF 5-628), which indicates that some of the additive in the synthesis
might also have crystallized during the process.
Fig. 2 shows SEM images of the annealed α-Fe2O3 samples. α-Fe2O3 croissant-like structures with a mean length
of about 419 nm and an average width of 264 nm were formed for sample A, as seen in Fig. 2 (a). The presence of
cleavages at the surface of the grown particles are strongly suggestive of agglomeration. After the nucleation of the
individual nanoparticles, their aggregation was facilitated by the Cl- ions in the solution. It is known that Cl- has a
very strong coordinative property with Fe3+ which enables the aggregation of fine primary particles [17]. Meanwhile,
the urchinlike structures formed in sample B were similar to those shown in previous researches [11, 16]. These
structures have a spherical body from which nanorods point outward like needles, creating a morphology resembling
an urchin. The hydrolysis of Fe3+ ions results to the formation of nanoparticles which interconnect into spheres. An
A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254 251
electrical double layer around the spheres is formed by the acidic solution and the SO4
2- ions [16]. These anions
attract the other Fe3+ ions in the solution and allow for the epitaxial growth of nanorods onto the surface of the
microsphere [16]. The microspheres have a mean diameter of 4.13 μm, while the needles have an average length of
321 nm. Sample C show textured microspheres with a mean diameter of 1.63 μm as seen in Fig. 2 (c). The textured
surface is due to short and thick rods growing on the surface of the spheres. The C2O4
2- ions possibly coordinate with
Fe3+ similar to the effect of SO4
2-. However, it is apparent that coordination between C2O4
2- and Fe3+ ions are not
strong, leading to thick and short rods compared to that in Fig. 2 (b).
Fig. 1. XRD Patterns of α-Fe2O3 structures prepared with different additives: (a) A [NaCl], (b) B [Na2SO4,] and (c) C [Na2C2O4.].
Fig. 2. SEM images of α-Fe2O3 structures prepared with different additives: (a) A [NaCl], (b) B [Na2SO4,] and (c) C [Na2C2O4.].
3.2. Photocatalytic Property of α-Fe2O3 structures
The absorbance spectra of MO with and without H2O2, and with α-Fe2O3/H2O2 samples are shown in Fig. 3. The
H2O2 was used as an activator for the α-Fe2O3 catalysts. α-Fe2O3 cannot degrade MO by itself efficiently because it
degrades through the photo-Fenton process. The mechanism for the initiation of degradation is as follows: [11, 16].
𝐹𝑒𝑂 → 𝐹𝑒𝑂 𝑒,ℎ (1)
𝐹𝑒  𝑒 → 𝐹𝑒 (2)
𝐹𝑒 𝐻𝑂→𝐹𝑒 𝑂𝐻•𝑂𝐻 (3)
252 A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254
Upon irradiation of the catalyst, photoexcitation happens which produces electron hole pairs (Eq.1). The excited
electron then reduces Fe3+ to Fe2+ which consequently reacts with H2O2 to produce the Fenton reagent, OH• (Eq. 2
and 3). The produced radical species is responsible for the degradation of the dye [11, 16].
It can be observed from Fig. 3 that Sample B was able to substantially degrade the MO dye after 2 h of exposure
to UV-C light. This is possibly due to its high surface area [19]. The nanorods radiating out of the spheres could
have provided more sites for adsorption of H2O2, thus producing more OH•. Meanwhile, samples A and C were able
to degrade the dye by only a small degree. It is interesting to note that H
2O2 alone actually registered a higher
degradation of MO than with the presence of samples A and C. H2O2 in the presence of UV light can also dissociate
and produce OH• radicals. Intuitively, samples A and C in the presence of H2O2 would most likely degrade more
MO than with H2O2 alone does. However, after irradiation, it was observed that only sample B settled at the bottom
of the container of MO while samples A and C were left dispersed. The dispersed particles may have occluded the
whole mixture thereby impeding efficient penetration of light throughout the vessel. Only those at the top most
accessible to UV irradiation may have successfully absorbed photon energy and thus contribute to the degradation of
MO. Having different morphologies and sizes, the samples exhibited different properties when mixed in a solution.
An excess dose of catalyst promotes aggregation, light scattering, and turbidity, thereby decreasing light penetration
in the solution [7].
Fig. 3. Absorbance spectra of 5 ppm MO containing the three samples and the control set-ups after 2 h irradiation in UV light
3.3. Degradation Kinetics
The time-dependent degradation of 2.5, 5.0, and 7.5 ppm MO solutions containing α-Fe2O3 Sample B was
investigated. Shown in Fig. 4 and 5(a) are the absorption spectra and percent removal vs. time for the three
concentrations. The degraded fraction was computed using Eq. 4. C0 was taken to be the concentration of the system
at time zero, while C was taken to be the concentration at the actual time of measurement.
Percent removal 100  
100 (4)
It can be observed from the figures that all solutions were degraded over time. The solution concentration
decreased as irradiation time is increased. Kinetic investigation of the degradation of the different systems was also
done to determine the fit of the reaction to the Langmuir-Hinshelwood model. A good fit was obtained for all
systems, evidenced by the R2 values ranging from 0.93 to 0.97 as seen in Fig. 5(b). This strongly suggests that the
adsorption of OH• and MO molecules occurs as a monolayer on the α-Fe
2O3 surface. A monolayer adsorption
highlights the advantage of the large surface area of the urchin-like structure. However, the Langmuir-Hinshelwood
A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254 253
model is limited to the description of the surface reaction on the catalyst. It does not consider interactions involving
free molecules, i.e. the degradation of adsorbed MO by free OH• [20].
Fig. 4. Absorbance spectra of the photodegradation of MO using sample B at different initial concentrations.
The rate constants for the 2.5, 5.0, and 7.5 ppm MO systems were calculated to be 0.0118, 0.0071, and 0.0058
min-1 respectively. This shows that the α-Fe2O3 urchin-like structures are able to degrade MO faster at lower dye
concentrations. This is attributed to fewer organic molecules to be oxidized and degraded by OH• radicals produced
from the photocatalytic reaction. Conversely, a higher amount of organic molecules in solution also impedes the
efficient reaction between H2O2 and Fe2+ resulting to less OH• available for the degradation of MO. In addition, this
also decreases photon absorption by the photocatalyst [7]. Nevertheless, in all systems, the dye was degraded
substantially by the urchin-like α-Fe2O3 photocatalysts. This demonstrates the promising capability of such α-Fe2O3
structures in the degradation of organic pollutants in water.
Fig. 5. (a) Percent removal of MO at various initial concentrations and (b) Langmuir-Hinshelwood Model for the three systems.
4. Conclusion
This work presents the synthesis of α-Fe2O3 structures with different morphologies by varying the type of additive
during hydrothermal synthesis. Croissant-like microstructures, urchin-like structures, and textured microspheres
were formed using NaCl, Na2SO4, and Na2C2O4 as additives, respectively. It was shown that photocatalytic property
is greatly affected by the morphology of the samples with urchin-like hierarchical structures exhibiting the best
photocatalytic activity for the degradation of MO. The urchin-like structures provide higher surface areas than the
(a) (b)
254 A.M.G. Domacena et al. / Materials Today: Proceedings 22 (2020) 248–254
croissant-like structures and textured microspheres. Kinetic studies show that photocatalytic degradation of MO is
more enhanced at lower concentrations due to greater photon absorption of the photocatalyst and efficient interaction
between H2O2 and Fe2+.
Acknowledgements
The authors would like to acknowledge the Engineering Research and Development for Technology (ERDT), the
UP Diliman College of Engineering, and the Department of Science and Technology (DOST) through the National
Academy of Science and Technology (NAST) for funding this project through its Grants for Outstanding
Achievements in Science entitled “Hydrothermal Synthesis of Hierarchical Hematite (α-Fe2O3) Nanostructures for
Environmental Cleaning”.
References
[1] N. P. Cheremisinoff, P. Rosenfeld and A. R. Davletshin. “Responsible Care - A New Strategy for Pollution Prevention and Waste Reduction
through Environmental Management”, Texas: Gulf Publishing Company, 2008.
[2] E. Bazrafshan, A. A. Zarei, H. Nadi and M. A. Zazouli.Adsorptive removal of Methyl Orange and Reactive Red 198 dyes by Moringa
peregrina ash.” Indian Journal of Chemical Technology 21 (2014), 105-113.
[3] A. Zaman, P. Das and P. Banerjee. “Biosorption of Dye Molecules, Toxicity and Waste Management Using Bioremediation” (2016), 51-74.
[4] S. Dwivedi and T. Vats. “Remidiation of Dye Containing Wastewater Using Viable Algal Biomass” in Green Materials for Sustainable Water
Remediation and Treatment, The Royal Society of Chemistry (2013), 212-228.
[5] I. Abdulkadir and A. B. Aliyu. “Some wet routes for synthesis of hematite nanostructures,” African Journal of Pure and Applied Chemistry (7)
no. 3 (2013), 114-121.
[6] S. Zeng, K. Tang, T. Li, Z. Liang, D. Wang, Y. Wang and W. Zhou. “Hematite Hollow Spindles and Microspheres: Selective Synthesis,
Growth Mechanisms, and Application in Lithium Ion Battery and Water Treatment,” Journal of Physical Chemistry C 111, no. 28 (2007),
10217-10225.
[7] S. K. Kansal, M. Singh and D. Sud. “Studies on photodegradation of two commercial dyes in aqueous phase using different photocatalysts,”
Journal of Hazardous Materials 141 no. 3 (2007), 581-590.
[8] T. Wang, M.C. Huang, Y.K. Hsieh, W.S. Chang, J.C. Lin, C.H. Lee and C.F. Wang.“Influence of Sodium Halides (NaF, NaCl, NaBr, NaI) on
the Photocatalytic Performance of Hydrothermally Synthesized Hematite Photoanodes,” Applied Materials & Interfaces (2013), 7937-7949.
[9] C.E. Aquino, M. Bongar, A. Silvestre, M.D. Balela. “Synthesis of Hematite (α-Fe2O3) Nanostructures by Thermal Oxidation of Iron Sheet for
Cr (VI) Adsorption,” Key Engineering Materials 775 (2018), 395-401.
[10] M. Gili., G. Latag, and M. Balela. “In-situ deposition of hematite (α-Fe2O3) microcubes on cotton cellulose via hydrothermal method,”
Journal of Physics: Conference Series 985 (2018), 012027.
[11] N.J. Rapadas and M.D. Balela. “Hydrothermal Synthesis of Hierarchical Hematite (α-Fe2O3) Microstructures for Photocatalytic Degradation
of Methyl Orange,” Philippine Journal of Science 146 (2017), 395-402.
[12] M.L. Canaria, J.L. Ramos, C.M. Sayson, and M.D. Balela. “Growth of Hematite Nanostructures in Iron Foil for Environmental Cleaning,”
Solid State Phenomena 266 (2017), 101-104
[13] Cao, Zhiqin, Mingli Qin, Yueru Gu, Baorui Jia, Pengqi Chen, and Xuanhui Qu. "Synthesis and characterization of Sn-doped hematite as
visible light photocatalyst." Materials Research Bulletin 77 (2016), 41-47.
[14] Satheesh, R., K. Vignesh, A. Suganthi, and M. Rajarajan. "Visible light responsive photocatalytic applications of transition metal (M=Cu, Ni
and Co) doped α-Fe2O3 nanoparticles." Journal of Environmental Chemical Engineering 2, no. 4 (2014), 1956-1968.
[15] Tsege, Ermias L., Timur S. Atabaev, Md. A. Hossain, Dongyun Lee, Hyung-Kook Kim, and Yoon-Hwae Hwang. "Cu-doped flower-like
hematite nanostructures for efficient water splitting applications." Journal of Physics and Chemistry of Solids 98 (2016), 283-289.
[16] S. Zeng, K. Tang, T. Li and Z. Liang. “Hematite with the Urchinlike Structure: Its Shape-Selective Synthesis, Magnetism, and Enhanced
Photocatalytic Performance after TiO2 Encapsulation,” Journal of Physical Chemistry C (2010), 274-283.
[17] Kandori, Kazuhiko, Junko Sakai, and Tatsuo Ishikawa. "Definitive effects of chloride ions on the formation of spherical hematite particles in
a forced hydrolysis reaction." Physical Chemistry Chemical Physics 2, no. 14 (2000), 3293-3299.
[18] X. Li, X. Yu, J. He and Z. Xu. “Controllable Fabrication, Growth Mechanisms, and Photocatalytic Properties of Hematite Hollow Spindles,”
Journal of Physical Chemistry C (2009), 2837-2845.
[19] M. M. Khan, S. F. Adil and A. Al-Mayouf. “Metal oxides as photocatalysts,” Journal of Saudi Chemical Society (2015), 462-464.
[20] B. Liu, X. Zhao, C. Terashima, A. Fujishima and K. Nakata. “Thermodynamic and kinetic analysis of heterogeneous photocatalysis for
semiconductor systems,” Physical Chemistry Chemical Physics 16 (2014), 8751-8760.
... The Fe 3+ ions on the surface of the Fe 2 O 3 particles are also reduced to Fe 2+ ions by accepting electrons (Eq. 8) [13,48]: ...
... This rapid discoloration is due to the oxidation of CR by hydroxyl radicals resulting from the reaction of Fe 2+ ions with H 2 O 2 (Eq. 9) [48]: ...
... A gradual decrease of the most intense band located at 500 nm was observed from the first minutes of treatment until almost complete disappearance after 30 min of reaction. This disappearance of color can be explained by the destruction of the azo group in the dye structure as a result of attack by hydroxyl radicals [22,48]. ...
Article
Full-text available
Natural hematite supported cobalt (Co/Nat-Hem), a new heterogeneous catalyst, was successfully prepared by a simple impregnation technique. Textural, structural, functional groups and morphological aspects of the prepared catalyst were analyzed by BET, XRD, FT-IR, XRF, pHpzc, SEM-EDX methods. BET analysis shows that the Co/Nat-Hem catalyst has a specific surface area of 13.44 m²/g and a pore volume of 0.062 cm³/g. The SEM-EDX method showed that 7.19 wt% of the cobalt species were dispersed on the surface of the prepared catalyst. The catalytic activity of Co/Nat-Hem was evaluated by photo-Fenton oxidation of the dye Cong red (CR). The effect of reaction parameters on CR oxidation efficiency was investigated. Under optimal reaction conditions (1 g/L catalyst concentration, 0.2 mol/L H2O2 initial concentration and pH 3), the percentage of CR discoloration reaches 97% after 30 min. The Co/Nat-Hem catalyst achieved a synergistic ratio of 28.5%, which was added to the oxidation rate of the Nat-Hem catalyst. The stability of the Co/Nat-Hem catalyst was demonstrated by the limited activity of leachates in the homogeneous Fenton oxidation of CR. The phytotoxicity of the CR dye was tested during the irrigation of the bean (Vicia faba L.).
... Due to the absence of adequate wastewater treatment methods, different dangerous materials constantly enter natural water resources [3]. Every year, tons of chemically stable synthetic dyes are wasted from the pharmaceutical [4,5], garment [6], textile [7], leather, ink, paper [8], and plastic industries [9], negatively affecting aquatic life and also threatening human health, as many of these chemicals are highly hazardous and have been classified as mutagenic, carcinogenic, or genotoxic [10]. Furthermore, these water contaminants decrease sunlight penetration into bodies of water, which influences photosynthetic processes and affects aquatic flora and fauna [11]. ...
... log(q e − q t ) = logq e − k 1 t/2.303 (4) t/q t = 1/k 2 q e 2 + t/q e (5) ...
Article
Full-text available
Water pollution by organic dyes represents a major health and environmental issue. Despite the fact that peptide-based hydrogels are considered to be optimal absorbents for removing such contaminants, hydrogel systems often suffer from a lack of mechanical stability and complex recovery. Recently, we developed an enzymatic approach for the preparation of a new peptide-based magnetogel containing polyacrylic acid-modified γ-Fe2O3 nanoparticles (γ-Fe2O3NPs) that showed the promising ability to remove cationic metal ions from aqueous phases. In the present work, we tested the ability of the magnetogel formulation to remove three model organic dyes: methyl orange, methylene blue, and rhodamine 6G. Three different hydrogel-based systems were studied, including: (1) Fmoc-Phe3 hydrogel; (2) γ-Fe2O3NPs dispersed in the peptide-based gel (Fe2O3NPs@gel); and (3) Fe2O3NPs@gel with the application of a magnetic field. The removal efficiencies of such adsorbents were evaluated using two different experimental set-ups, by placing the hydrogel sample inside cuvettes or, alternatively, by placing them inside syringes. The obtained peptide magnetogel formulation could represent a valuable and environmentally friendly alternative to currently employed adsorbents.
... The use of clays as supports provides an economical and simple alternative due to their abundance, availability, and incorporation of iron with attainable techniques [35,36] [12,23]. Hematite (α-Fe 2 O 3 ) is a promising semiconductor photocatalyst, with the advantage that it is abundant, inexpensive and environmentally friendly, has a band gap of approximately 2.1 eV, which allows it to adsorb light energy over a wide wavelength range, and is thermodynamically stable [37,38]. Moreover, hematite is a promising material for photocatalytic systems due to its high electrical conductivity and magnetic nature, and therefore, it can be beneficially used in the system by adding UV to the heterogeneous Fenton system to take advantage of the appropriate band gap and increase the degradation efficiency [39]. ...
Conference Paper
Inadequate handling and disposal of contaminated industrial waste significantly contribute to environmental pollution. The presence of pollutants, including dyes, in wastewater necessitates the development of innovative remediation techniques. Metal oxide-catalyzed photodegradation capitalizes on the capacity of a dye to absorb light energy, offering a rapid method to break down the dye into less harmful, colorless byproducts. In this work, bimetallic copper-iron oxides with various copper to iron were synthesized for the photodegradation of fuchsine. The photocatalysts were prepared through oxalate precipitation followed by thermal decomposition. Structural analysis revealed a MOF-like structure of the bimetallic oxalate precursors. Thermal decomposition of the oxalates yielded photocatalytic bimetallic copper-iron oxides. Photodegradation studies demonstrated that the addition of copper-iron oxides accelerated the degradation of fuchsine and a higher concentration of CuO enhances the performance of the photocatalyst. Notably, the copper-iron oxide with a 1:1 (CuFe) ratio proved to be the most effective catalyst for the photodegradation of fuchsine. Furthermore, the photodegradation of fuchsine conforms to a pseudo-first order model and exhibits characteristics of a first-order reaction. Our findings emphasize that simple and high-efficiency bimetallic oxide catalysts can be used for water decontamination applications.
Article
Synthetic dyes play a significant role in various industries, including textiles, paper, food, leather, cosmetics, and pharmaceuticals, with azo dyes being the predominant category. However, their widespread use has led to a pressing environmental issue due to the release of these dyes into wastewater, posing concerns for public health. To address this problem, numerous advanced oxidation processes (AOPs) have been devised, among which Fenton oxidation has gained prominence for its ability to break down these dyes. This review compiles the most recent research findings on the photo-Fenton and electro-Fenton degradation of methylorange (MO).
Article
Herein, in the current study, Fe2O3/Mn2O3 nanocomposite was fabricated using the gel-thermal decomposition method and investigated for the removal of Acid Yellow 23 (AY23) via a Fenton-like reaction. The fabricated nanocomposite was characterized by employing XRD, FTIR, SEM, EDX, and pH PZC analysis. The XRD/FTIR results confirmed the presence of Mn2O3 and Fe2O3 with no impurities. According to SEM results, the nano-composite exhibits coral-like architectural morphology, and the presence of the required elements is confirmed by the EDX analysis. The influence of operating parameters including initial pH values, temperature, dosages of Fe2O3/Mn2O3 catalyst, and H2O2 on the removal efficiency was investigated. The nanocomposite showed excellent performance toward the removal of AY23. Under optimized experimental conditions, complete removal was observed in 60 min. Furthermore, scavenger studies demonstrate the presence of HO• radicals as the primary active species involved in the dye removal process. Based on the results of the experiments, a possible mechanism for AY23 removal was proposed. At the three recycling runs, the removal efficiency of AY23 dye remained around 80 %, which indicates that Fe2O3/Mn2O3 has good stability and can be used repeatedly. These findings suggested that the Fe2O3/Mn2O3 nanocomposite as formed could have excellent application prospects in the Fenton-like field.
Article
Full-text available
In this study, hematite (α-Fe 2 O 3 ) nanostructures were synthesized via thermal oxidation of Fe sheet in dry air and in water vapor. SEM images show nanoblades and nanowires growing on the surface of the sheet. Samples synthesized in water vapor generally produced larger nanostructures while samples oxidized in higher temperatures formed taller and slender nanostructures. The α-Fe 2 O 3 nanostructures were used as adsorbent for Cr (VI) in acidic medium. Chromium removal was highest with the samples synthesized at 650°C in water vapor with 95% efficiency. Kinetic and thermodynamic studies revealed that the adsorption process strongly followed pseudo-second order kinetics model and is endothermic. The process also follows the Langmuir adsorption isotherm model, suggesting that the process is described by homogeneous, monolayer adsorption. Adsorption of Cr (VI) onto hematite may be attributed to the electrostatic reaction between the positively charged hematite adsorbent and negative chromium ion.
Article
Full-text available
Hematite microcubes with truncated edges have been successfully deposited on cotton cellulose via one-step hydrothermal process using anhydrous FeCl3 and glycine as Fe(III) precursor and chelating agent, respectively. The amount of glycine significantly affects the morphology and yield of hematite. The addition of 0.495 g of glycine to 50 ml of 0.1 M FeCl3 solution with 0.400 g of cotton resulted to hematite-deposited cellulose having ~15% hematite content. The reduction of glycine to 0.247 g increased the amount of hematite on the surface of the cotton cellulose to ~20% by weight. However, the hematite microcubes have a wide size distribution, with particle size in the range of 0.684 μm to 1.520 μm. Without glycine, hematite cannot be formed in the solution.
Article
Full-text available
Hematite (Fe2O3) nanostructures were successfully grown on the surface of iron foil by simple thermal oxidation in air. The iron foil was heated at 300 – 700 °C for 8 h to produce Fe2O3 nanosheets and nanowires. The oxide layer is primarily composed of Fe2O3 with a small amount of magnetite (Fe3O4) and wustite (FeO). The photocatalytic activity of the Fe2O3 nanowires was investigated by degradation of methyl orange dye under 30 W UV light in the presence of a minute amount of hydrogen peroxide. The Fe2O3 nanowires exhibit excellent photocatalytic activity, wherein 2.5 ppm of the methyl orange was fully decomposed in 30 min.
Article
Full-text available
In the present study, a low-cost adsorbent derived from Moringa peregrina plant has been used for its ability to remove two common textile dyes, namely Methyl Orange (MO) and Reactive Red 198 (RR-198) from aqueous solutions. The effect of various operating parameters such as initial concentration of dye (10-150 mg L-1), contact time (10-150 min), adsorbent dosage (1-13 g L-1) and pH (2-11) is investigated. At an optimum pH of 2, approximately 96% removal of dye (50 mg L-1) is obtained for an adsorbent dose of 7.0 g L-1 after a 70 min contact time. The equilibrium assessment reveals that the Langmuir model is better than Freundlich model for the experimental data. The thermodynamic properties (λG°, λH° and λS°) show that the adsorption of RR-198 and MO dyes onto Moringa peregrina tree shell ash (MPTA) is spontaneous, endothermic and feasible in the temperature range of 293-313 K. Also, the magnitude of enthalpy change indicates that the adsorption is physical in nature. Finally it can be concluded that the MPTA can be used as a low-cost adsorbent for the removal of different concentrations of anionic dyes from aqueous solutions.
Article
Full-text available
Increased attention has been paid to the synthesis of hematite nanoparticles recently due to its properties and application in various fields of modern technology. Various methods (aqueous and non-aqueous) have been applied in order to produce hematite nanostructures of required size and morphology suitable for these various applications. The aqueous route has been found to be simpler and more versatile than the non-aqueous routes. This paper reviews some of the most commonly used wet route for the synthesis of hematite nanoparticles and shows that the sol-gel and hydrothermal methods are simple and versatile, capable of producing narrow particle size ranges with good control of morphology and high yield in contrast to precipitation and microemulsion methods.
Article
Full-text available
Metal oxides are of great technological importance in environmental remediation and electronics because of their capability to generate charge carriers when stimulated with required amount of energy. The promising arrangement of electronic structure, light absorption properties, and charge transport characteristics of most of the metal oxides has made possible its application as photocatalyst. In this article definition of metal oxides as photocatalyst, structural characteristics, requirements of the photocatalyst, classification of photocatalysts and the mechanism of the photocatalytic process are discussed.
Article
Sn-doped hematite nanoparticles are prepared by solution combustion synthesis. The products are characterized with various analytical and spectroscopic techniques to determine their structural, morphological, light absorption and photocatalytic properties. The results reveal that all the samples consist of nanocrystalline hematite with mesoporous structures, and Sn has the ability to inhibit the growth of hematite particle. Compared to pure hematite, the doped hematite samples with appropriate amount of Sn show better activities for degradation of methylene blue under visible light irradiation. The highest activity is observed for 5% Sn doped hematite and this product has long-term stability and no selectivity for dye degradation. The enhanced performance of 5% Sn doped hematite is ascribed to the smaller particle size, increased ability to absorb in visible light, efficient charge separation as well as improved e− transfer associated with the effects of appropriate amount of Sn doped sample.
Article
Inadequate access to clean water afflicts people throughout the world, and in developing countries any solution to this challenge must be achieved at a low cost and low energy demand. At the same time, the use of chemicals, and subsequent environmental impact must also be reduced. Green and sustainable water remediation is a rapidly growing field of interest to governments and corporations alike, with considerable input from academics, environmental consultants and public interest groups. This book presents a focused set of articles covering a range of topics in the field, examining not only the adoption of natural products for water remediation, but also the synthesis of new materials and emerging clean technologies. Contributors from across the globe (including some "on the ground" in the developing world) present a comprehensive digest in the form of review-style articles highlighting the current thinking and direction in the field. Interested stakeholders from all sectors will find this book invaluable, and postgraduate students of chemical engineering or environmental science will benefit from the real-world applications presented.
Article
Environmental regulations provide protection to the public, workers and the environment. To protect themselves from long-term liabilities, however, companies have to do more than just comply with the basic responsibilities. This handbook is designed to introduce terminology, methodology, tools, procedures and practical guidance for incorporating efficient pollution prevention strategies into the overall business plan. It is a company's responsibility to protect and control its management of waste and pollution, and a company that fails to do so will ultimately inflict a negative impact on its bottom line, especially in financial performance. Responsible Care delivers critical guidelines and rules of thumb required for industrial managers to improve their companies' profitability through waste reduction, cleaner production technologies and sound management practices.