International Journal of Environmental Analytical 2023

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

International Journal of Environmental Analytical

Chemistry

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/geac20

Influence of manganese-substitution on
adsorption properties of BiFeO3 onto the
elimination of methyl orange

Azia Wahida Binti Aziz, Noor Haida Mohd Kaus, Saifullahi Shehu Imam &
Takaomi Kobayashi

To cite this article: Azia Wahida Binti Aziz, Noor Haida Mohd Kaus, Saifullahi Shehu Imam &
Takaomi Kobayashi (2023): Influence of manganese-substitution on adsorption properties
of BiFeO3 onto the elimination of methyl orange, International Journal of Environmental
Analytical Chemistry, DOI: 10.1080/03067319.2023.2178916

To link to this article: https://doi.org/10.1080/03067319.2023.2178916

View supplementary material

Published online: 22 Feb 2023.

Submit your article to this journal

Article views: 83

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=geac20
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY
https://doi.org/10.1080/03067319.2023.2178916

Influence of manganese-substitution on adsorption


properties of BiFeO3 onto the elimination of methyl orange
Azia Wahida Binti Aziza, Noor Haida Mohd Kausa,b, Saifullahi Shehu Imamc
and Takaomi Kobayashib
a
School of Chemical Sciences, Universiti Sains Malaysia, Penang, Malaysia; bDepartment of Material Science
and Technology, Nagaoka University of Technology, Nagaoka, Japan; cDepartment of Pure and Industrial
Chemistry, Bayero University, Kano, Nigeria

ABSTRACT ARTICLE HISTORY


Herein, the BiFeO3 was modified by substituting varying manganese Received 19 December 2022
(Mn) concentration to improve the efficacy of eliminating methyl Accepted 1 February 2023
orange (MO), a water-soluble carcinogenic azo dye was presented. To KEYWORDS
examine the structure, morphology and band gap characteristics of Nanoporous materials;
the as-synthesised Mn-BiFeO3 nanocatalyst were characterised using adsorption; BiFeO3;
various techniques including XRD, FESEM with EDX, FTIR, UV-vis DRS, manganese; methyl orange
and BET. Subsequently, parameters affecting the removal perfor­
mance, such as adsorbent dosage (0.02–0.10 g), initial dye concen­
trations (10–60 mg/L), initial pH (1–13), percentage mol of
manganese substitution (0–30%), and reusability of the adsorbent,
were examined in detail. The specific surface area increases with
doping from 6.829 to 13.063 m2/g, leading to enhanced adsorption
capacity of manganese-substituted BiFeO3 as opposed to pristine
BiFeO3. Adsorption kinetics was found to follow the pseudo-second-
order kinetic model, while the adsorption equilibrium data obeyed
the Freundlich isotherm model with a maximum adsorption capacity
of 25.2 mg/g at 298 K. Interestingly, the catalyst showed more than
95% adsorption even after three consecutive cycles and therefore
could be a good candidate for wastewater remediation.

1. Introduction
Water pollution caused by organic pollutants such as pharmaceuticals, cosmetics, biome­
dical goods, personal care products, organic dyes, and pesticides is rapidly becoming one
of the world’s most pressing challenges [1]. A great demand on water resources has
emerged from the expansion of industry and the astronomical rise in human population.
Simultaneously, enormous quantities of harmful wastewaters are produced, endangering
the quality of these resources [2]. The wastewater discharged as a result of industrial
activity contains hazardous organic compounds that contribute to climate change and
negatively impact natural ecosystems and human life [3,4].
Methyl orange (MO) is one of the most often employed dyes in a variety of industrial
operations [5]. It is a water-soluble (5 g/L, H2O, 20 ºC) anionic dye that belongs to the azo

CONTACT Noor Haida Mohd Kaus noorhaida@usm.my


Supplemental data for this article can be accessed online at https://doi.org/10.1080/03067319.2023.2178916.
© 2023 Informa UK Limited, trading as Taylor & Francis Group
2 A. W. BINTI AZIZ ET AL.

group [6,7]. MO is used extensively in the cosmetic, paper, and plastic sector, with textile
industries being the largest consumer [8,9]. However, throughout the colouring process,
many dyes are lost to the waste stream since not all dyes adhere to the fabric. Its loss in
wastewaters may range from 2% for basic dyes to 50% for reactive dyes, resulting in severe
contamination of surface and ground waters in the dyeing industry’s area [10]. In the case of
MO, it is irritating, carcinogenic, poisonous and can lead to severe health damages such as
vomiting, diarrhoea, mental confusion, nausea, vertigo, profuse sweating, kidney damage,
and genetic mutation [11–14]. Consequently, dye-containing wastewater must be managed
responsibly, and the dye components must be removed prior to release into the aquatic
environment. Unfortunately, the aromatic structure and azo ( N ¼ N ) chromophore
makes it difficult to degrade MO by biological and chemical methods [15].
Nowadays, adsorption is one of the most feasible methods for treating water pollutants
[16]. The process is facile, rapid, cheap, convenient, recyclable, environmentally friendly,
and effective for removing dyes [17–20]. Nevertheless, selecting an appropriate adsorbent
for maximum adsorption performance is crucial.
Recently, bismuth ferrite (BiFeO3) is an interesting material that has received significant
attention in environmental remediation [21]. It is a cost-effective, non-toxic material with
high chemical stability, smaller energy band gap (2.2–2.7 eV), and simultaneously exhibit­
ing ferroelectric and magnetic properties at room temperature [22–24]. In our previous
study, BiFeO3 nanoparticles were successfully synthesised via biotemplated method using
brown algae [25]. Regrettably, even after 24 hr, the performance of removing MO dye via
adsorption using pristine BiFeO3 nanoparticles was negligible, necessitating modification.
The objective of the present study is to synthesise, characterise, and then subsequently
employ Mn-substituted BiFeO3 as a possible sorbent for the removal of MO dye. The
physical and chemical properties of the Mn-substituted BiFeO3 were analysed using X-ray
diffraction (XRD), Field emission scanning electron microscopy (FESEM), UV-vis diffuse
reflectance spectroscopy (DRS), Fourier transform infrared (FTIR) spectroscopy, N2 adsorp­
tion-desorption isotherms. Then the effectiveness of the Mn-substituted BiFeO3 was
evaluated with various working parameters such as the effects of pH, mol% of Mn
substitute in BiFeO3, adsorbent dosage, initial MO concentration, and temperature in
assisting the elimination of MO.

2. Materials and methods


2.1. Materials
Analytical grade chemicals were used in the present work without further purification, and
deionised water was used to prepare solutions. Bismuth (III) nitrate pentahydrate [Bi(NO3)3.
5 H2O, molecular weight: 485.07 g/mol] and iron (III) nitrate nonahydrate [Fe(NO3)3.9 H2O,
molecular weight: 404 g/mol] were obtained from Sigma-Aldrich. Manganese (II) nitrate
tetrahydrate, [Mn(NO3)2.4 H2O, molecular weight: 251 g/mol] was obtained from Merck
chemicals. Hydrochloric acid, [HCl, molecular weight: 36.46 g/mol] sodium hydroxide
[NaOH, molecular weight: 40 g/mol] and methyl orange, [C14H14N3NaO3S, molecular weight:
327.33 g/mol] were obtained from QReC Chemicals. Sodium alginate [C6H9NaO7, molecular
weight: 198.11 g/mol] was obtained from Acros Organics. Ammonium hydroxide [NH4OH,
molecular weight: 35.01 g/mol] was obtained from J.T.Baker.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 3

2.2. Synthesis of Mn-substituted BiFeO3 nanoparticles


The preparation process is similar to the synthesis of pristine BiFeO3 reported in our
previous study [25] except for adding various mol percentages of manganese (10, 20 and
30 mol%) into the Bi-Fe solution before the heating process of the precursor commenced.
Bismuth nitrate pentahydrate [Bi(NO3)3·5 H2O] was weighed and dissolved in deionised
water. Iron (III) nitrate nonahydrate [Fe(NO3)3·9 H2O] and manganese (II) nitrate tetrahy­
drate [Mn(NO3)2.4 H2O] were weighed, and the mixture was dissolved in deionised water
before adding into the bismuth (III) nitrate pentahydrate solution. Afterwards, 5 mL of Bi-
Fe-Mn-O solution was taken out and added into 20 mL of sodium alginate (0.5 weight %).
The precursor was heated at 80°C under vigorous stirring for 2 h. The pH of the precursor
was adjusted to reach pH 8 using 1.44 M NH4OH. The caped beaker, with solid deposition,
was kept overnight in the oven at 80°C. The product was calcined at 550°C for 2 h with
a heating rate of 5°C/min. The resulting powder was washed with distiled water to remove
all soluble salts/impurities and then dried in an oven at 80°C for 2 h. A schematic flow
chart of the process is shown in Scheme 1.

Scheme 1. Flow chart for the preparation of Mn-substituted BiFeO3 nanoparticles using sodium
alginate as a biotemplate.
4 A. W. BINTI AZIZ ET AL.

2.3. Characterisation techniques


X-ray diffraction (XRD) measurements were carried out using Panalytical X’pert Pro
diffractometer equipped with graphite monochromator with Cu-Kα X-rays source at
1.54060 nm operating at 40 kV. The morphology of the samples was examined using
Field emission scanning electron microscope (FESEM, FEI-QUANTA FEG 650). Functional
groups were determined using Fourier transform infrared spectroscopy (Perkin Elmer
2000 FTIR) in the range of 4000–400 cm−1. UV-Vis absorption spectra of the solid materials
were recorded by LAMBDA 25 UV/Vis Systems. The surface area and porosity of the
samples were determined using Micromeritic ASAP 2020 surface area and porosity
analyser.

2.4. pH of point of zero charge (pHpzc)


The surface of the materials has a great consequence on their behaviour towards
water pollutants. In this study, pHpzc is the pH that the solid surface materials tend to
have no net charge (zero charge). In a series of Erlenmeyer flasks, 40 mL of an
electrolyte (0.1 M NaNO3) was adjusted to specific values of pH in the range of 1 to
11 by using NaOH (0.1 and 1.0 M) and HNO3 (0.1 and 1.0 M). Subsequently, 0.1 g of the
adsorbents was loaded into the flasks. The suspensions were shaken at a shaking rate
of 250 rpm. After 24 h, the final pH of each suspension was recorded. The difference in
the pH value was computed and plotted against the initial pH of the suspension. The
point at which the curve passed through zero line was considered the point of zero
charge of the material.

2.5. Removal of MO dye using Mn-substituted BiFeO3


To investigate the MO dye removal efficiency by Mn-BiFeO3, MO sorption studies were
monitored at 463 nm via the Shimadzu UV 2600 spectrophotometer (version 1.03 operating
using UV probe 2.42) using a UV-visible spectrum recorded in a 200–800 nm wavelength
range. The process was conducted under different experimental conditions and parameters,
including variable pH, mol % of Mn-substituted BiFeO3 as a catalyst adsorbent, adsorbent
dosage, and time of contact on dye adsorption. The amount of dye adsorbed by the
adsorbent was measured using the following equation.
Co Ct
% Removal ¼ � 100% (1)
Co
where Co and Ct are the initial and the concentration of dye solution at different reaction
time t.

2.5.1. Adsorption kinetics


In order to find the suitable chemical removal model for describing the experimental
kinetic data, the obtained data were evaluated using pseudo-first and pseudo-second-
order reaction rates at various temperatures. The pseudo-first-order kinetic model is
known for describing the adsorption rate of liquid systems based on the adsorption
solid capacity [26]. The pseudo-first-order kinetic model is generally expressed as:
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 5

lnðqe qt Þ ¼ lnqe k1 t (2)


Meanwhile, the pseudo-second-order kinetic model deals with the sorption capacity and
mechanism, which involves chemisorption and physicochemical interactions between the
two phases. Thus, the linear form of the sorption kinetics following pseudo-second-order
kinetics, as reported in the literature, can be represented as follows [27]:
t t 1
¼ þ (3)
qt qe k2 q2e
Where qe (mg/g) is the amount of the solute adsorbed per unit weight of the catalyst at
equilibrium and qt is the amount adsorbed per unit time t (min), while k1 and k2 are the
pseudo-first-order and pseudo-second-order rate constants.
The activation energy, Ea (kJ/mol), for the adsorption process was obtained by using
the Arrhenius equation:
Ea
ln k2 ¼ ln A (4)
RT
Where Ea is the activation energy (kJ/mol), k2 is the pseudo-second-order rate constant,
A is the temperature-independent Arrhenius factor, R is the gas constant (8.314 J/mol K),
and T is the absolute temperature (K).

2.5.2. Adsorption isotherms


Langmuir and Freundlich models were applied and examined in detail to determine the
equilibrium nature of adsorption. The Langmuir isotherm model deals with the monolayer
coverage of the adsorbent, and adsorption occur over specific homogeneous sites on the
adsorbent [28]. The linear form of the Langmuir model is given as:
Ce 1 Ce
¼ þ (5)
qe qm kL qm
The Freundlich isotherm model assumes multilayer adsorption is used to investigate the
adsorption behaviour for heterogeneous surfaces [29]. The linear form of the Freundlich
model is given as:
� �
1
ln qe ¼ ln Kf þ lnCe (6)
n
Where Ce is the equilibrium concentration (mg/L) of methyl orange, qe is the amount
adsorbed per specified amount of adsorbent (mg/g) at equilibrium, qm is the maximum
amount adsorbed at monolayer saturation (mg/g), and KF (mg/g) is the empirical
Freundlich constant related to the adsorption capacity. The value of 1/n is dimensionless
with a value in the range of 0 to 1, which indicates the magnitude of the adsorption
driving force or surface heterogeneity.

2.6. Reusability of the adsorbent


For this purpose, the adsorbent was transferred into 50 mL of 50 ppm methyl orange at
pH 1. The flask was place inside the shaker at 350 rpm for 1 h before the solution was
centrifuged, and the absorbance of the solution was measured. The spent adsorbent was
6 A. W. BINTI AZIZ ET AL.

washed using 0.1 M NaOH and washed back with distilled water three times to neutralise
it. The washed adsorbent was dried in the oven before reusing it with the new dye
solution. The experiment was repeated until the sixth cycle using the same adsorbent.

3. Results and discussion


3.1. Characterisations
3.1.1. XRD analysis
Figure 1 compares the XRD patterns of pure and Mn-substituted BiFeO3 powders sintered
at various molar ratios and 550°C. It is challenging for manganese ions to occupy inter­
stitial sites of Bi3+, which has 103 pm ionic radii; instead, it will replace Fe3+ ions on the
substitutional sites. Due to the fact that Fe3+ and Mn2+ have comparable ionic radii of 65
pm and 66 pm, respectively. The rhombohedral structure of BiFeO3 is anticipated to cause
minor lattice distortions by Mn substitution. The peaks are attributed to BiFeO3 and low-
intensity peaks of Bi2Fe2Mn2O10 (JCPDS 81–317), Bi2Fe4O9, and Bi2O3 occurred in the
presence of a 10% Mn content. The impurity peaks of Bi2Fe2Mn2O10 and Bi2Fe4O9
intensified with increasing mol percent of Mn substitution. Additionally, it was discovered
that the Mn substitution in BiFeO3 nanoparticles causes the crystallite size to increase with
mol percent of the substituent. The diameters of the crystallite size were found to be 16.7
nm, 18.1 nm, and 19.4 nm for 10, 20, and 30 mol percent manganese, respectively.

3.1.2. FESEM/EDX analysis


FESEM confirmed the morphology and grain particles of BiFeO3 nanoparticles prepared in
the presence of various mol percent of Mn substituent (Figure 2(a–c)). The particle
diameter of the nanoparticles increases significantly as the amount of Mn concentration
increases. An average diameter of 54 nm was observed in the presence of 10 mol% Mn
substitution in bismuth ferrite nanoparticles, and further addition of Mn2+ at Fe3+ sites
revealed the variation in grain morphology with large particle sizes. Larger particle

Figure 1. XRD patterns of Mn substitution in bismuth ferrite nanoparticles at different mol%.


INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 7

Figure 2. FESEM images of (a) 10 mol% Mn, (b) 20 mol% Mn, (c) 30 mol% Mn substitute in BiFeO3 and
(d) particle distribution for 10 mol% Mn, (e) particle distribution for 20 mol% Mn and (f) particle
distribution for 30 mol% Mn.

diameters of Mn-BiFeO3 with 198 nm was observed in the presence of 20 mol% Mn.
Further, increase in the concentration of Mn to 30 mol%, the particles are agglomerated
with a diameter of 101 nm. The variation in grain morphology may be the evidence of the
formation of impurity phases as shown in the XRD patterns as discussed earlier.
Average particle sizes diameter observed from SEM and crystallite sizes calculated from
XRD peak (110) for 10, 20 and 30 mol % of Mn substituted BiFeO3 are summarised in
Table 1.
The EDX spectra of 10 mol % Mn, 20 mol % Mn, and 30 mol % Mn were measured and
shown in Figure 3. EDX images taken from selected areas of different grains confirmed the
presence of Bi, Fe, Mn, and O elements, which further supported the fact that Mn2+ was
present in the nanoparticles. Thus, the chemical reaction was completed in the precursor
8 A. W. BINTI AZIZ ET AL.

Table 1. Crystallites and particle sizes of Mn-BiFeO3 nanoparticles at various


mol% of Mn substitution.
Mol % of Mn substitution Crystallite size (nm) Particles size (nm)
10 16.7 54.3 ± 12.81
20 18.1 198.6 ± 36.24
30 19.4 101.5 ± 20.00

Figure 3. EDX spectra of Mn-substituted at B-site of BiFeO3: (a) 10% mol, (b) 20% mol, and (c) 30% mol.

to form the desired perovskite BiFeO3 and BiFe1-xMnxO3 nanoparticles without any loss of
ingredients during annealing.

3.1.3. Bandgap of pristine and Mn substituted BiFeO3


Figure 4 shows that the optical band gap (Eg) of the pristine and Mn-substituted BiFeO3
samples evaluated by UV – Vis absorbance spectroscopy at wavelengths from 200 nm to
800 nm. A significant decrease in the bandgap energy value of Mn substituted BiFeO3
samples was observed.

Figure 4. Dependence of (αhv)2 of composite Mn-BiFeO3 at (a)20%, (b) 10%, (c) 0%, and (d) 30% Mn of
Kubelka-Munk.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 9

The optical bandgap of the pristine BiFeO3, 10%, 20%, and 30% Mn-doped BFO were
found to be 2.05, 1.67, 1.54, and 1.26 eV, respectively. This reduction in band-gap energy
for all the Mn substituted samples could be ascribed to the distortion generated in the
Fe – O octahedral rearrangement of molecular orbitals [30]. The distortion was attributed
to Mn substitution in BiFeO3, whereby Mn2+ ions were substituted for lower valence Fe3+
ions on the B-sites, forming donor impurity energy levels and then caused the decline in
the conduction band of BiFeO3 hence narrowing the bandgap. The reduction in optical
bandgap energy with the increase in Mn concentration additionally cause Mn- clustering
and enhancement of surface area-to-volume ratio. The reduction in the bandgap of
BiFeO3 is beneficial to the transition of electrons from the valence band to the conductive
band and the improvement in photoelectric conversion efficiency of BiFeO3 on the
application of photoactive materials.

3.1.4. FTIR analysis


Room temperature FTIR spectra for the synthesised BiFe1-xMnxO3 at x = 10%, 20% and
30% mol were shown in Figure 5. The strong absorption peaks at 563 cm−1, 562 cm−1, and
567 cm−1 represent Fe-O stretching vibration of FeO6 group for 10% mol Mn, 20% mol Mn,
and 30% mol Mn, respectively. Meanwhile, peaks at 442 cm−1, 447 cm−1, and 448 cm−1
reveal the stretching vibration of Bi-O in the perovskite structure of 10% mol Mn, 20% mol
Mn, and 30% mol Mn, respectively. Peaks at the region of 840 cm−1 reveal the stretching
vibration of Mn-O. All the samples had a sharp peak at 1386 cm−1 which assigned the
presence of trapped nitrates. However, nitrogen was not detected in EDX analysis despite
the considerable nitrogen content in the material (Figure 4). This might mean that
nitrogen was removed from the surface during the EDX process. FTIR spectroscopy

Figure 5. FTIR spectra various mol % of Mn substitution in BiFeO3 nanoparticles:(a) 10 mol %, (b) 20
mol % and (c) 30 mol %.
10 A. W. BINTI AZIZ ET AL.

does not require a vacuum so nitrogen could not absorb infrared rays. The band around
1630 cm−1 and 3440 cm−1 present in all the compositions was assigned to the bending
vibrational and stretching mode of water molecules trapped from the surrounding.

3.1.5. Materials surface area


The nitrogen sorption isotherms of 10 mol % Mn-BiFeO3, 20 mol % Mn- BiFeO3 and 30 mol
% Mn-BiFeO3 were presented in Figure 6(a–c). Based on the graphs, the isotherm exhibits
type IV and possesses H3 hysteresis loop according to IUPAC classification. This indicates
that all the materials are mesoporous [31]. Isotherm type IV can also be attributed to
monolayer-multilayer adsorption. Isotherms reveal type H3 hysteresis represents plate-
like particles giving rise to slit-shape pores. Among all these samples, 30 mol % Mn-BiFeO3
has the highest nitrogen uptake at relative pressure 0.98 P/P0 with the quantity adsorbed
27.51 cm3/g and desorption began at the same point (27.51 cm3/g) since the adsorption
already saturated and ended at 1.53 cm3/g.
The surface area of Mn-BiFeO3 nanoparticles can be determined using Brunauer,
Emmet and Teller, (BET) method. BET theory explains the adsorption of nitrogen gas
molecules onto a solid surface. Condensation of adsorbate inside pores occurred within
the pressure range usually used in BET method in the range of 0.05–0.30. From the data

Figure 6. N2 linear sorption isotherm of (a) 10 mol % Mn-BiFeO3 (b) 20 mol % Mn-BiFeO3 and
(c) 30 mol % Mn-BiFeO3.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 11

obtained, BET surface area increase as percentage mol of manganese in a host BiFeO3
increase. BET surface area (SBET) of 30 mol % Mn-BiFeO3 was found to be the highest with
the value of 13.063 m2/g because the porosity of this material is high compared to 10 mol
% (6.829 m2/g) and 20 mol % (7.350 m2/g) of manganese.

3.2. Adsorption studies of methyl orange using Mn-BiFeO3 nanoparticles


Adsorption studies of methyl orange using Mn-BiFeO3 nanoparticles were conducted in
detail to examine its efficiency in the dye removal at various parameters such as the
influence of pH solutions, the effect of mol percentage of the manganese, catalyst dosage
in the system, influence of initial concentration of the dye as well as the reusability of the
material.

3.2.1. Operational parameters


The effect of mol percentage of Mn substituted in BiFeO3 nanoparticles was examined to
assess the dye removal as indicated in Figure 7(a). The removal of methyl orange signifi­
cantly depends on the % of Mn content. The percentage removal of methyl orange
increased from 63% to 98% as the content of Mn in BiFeO3 increased from 10% to 30%
mol, respectively. Interestingly, 30 mol % Mn-BiFeO3 has a higher surface area (13.063 m2/g),

Figure 7. Effect of (a) mol % of Mn in BiFeO3 (b) adsorbent dosage of 30 mol % Mn-BiFeO3 (c) initial
concentration of methyl orange, and (d) initial solution pH.
12 A. W. BINTI AZIZ ET AL.

making the adsorption capacities of this material towards methyl orange high compared
with 20 mol % (7.350 m2/g) and 10 mol % of Mn-BiFeO3 (6.829 m2/g) as discussed earlier.
The removal of methyl orange using different catalyst dosages (0.02 g to 0.10 g) was
conducted in 1 hour, and the results were presented in Figure 7(b). Removal of methyl
orange was observed to be dependent on the catalyst dosage. At initial dose of 0.02, 0.04,
0.06, 0.08 and 0.10 g, about 54%, 65%, 89%, 94% and 97% of methyl orange was adsorbed,
respectively, using 30% mol of Mn-BiFeO3 nanoparticles. The removal potential of methyl
orange is increased by a greater adsorbent dosage, which provides the adsorbate mole­
cule more active sites.
The initial dye concentration is another important parameter affecting the dye removal
process. This was studied for 1 hr by varying the initial concentration from 10 to 60 mg/L,
and the results are shown in Figure 7(c). From Table S1, an approximate efficiency of
about 98% was achieved for all initial dye concentrations, indicating the independence of
adsorption to the initial concentration of methyl orange.
Finally, the effect of pH was examined in the range of 1 to 12, and the results are shown
in Figure 7(d). The pH dependence of dyes adsorption can primarily be related to its effect
on the surface properties of the adsorbent and the degree of ionization/dissociation of
the adsorbate molecule. From the results obtained, the maximum removal of methyl
orange was more than 98% at pH 1. The removal efficiency of methyl orange decreased
from 98.56% to 2.82% upon changing the pH from pH 1 to pH 10, respectively, and the
removal of the dye slightly increased at pH 12 (37.07%), as tabulated in Table S2. Since the
removal of methyl orange was efficient at acidic conditions, electrostatic interaction
between the anionic dye molecules and the positively charged surface of Mn-BiFeO3
nanocomposite might be the predominant mechanism. A negatively charged surface site
on the 30 mol % Mn-BiFeO3 at pH 12 does not favour the adsorption of anionic dye due to
electrostatic repulsion. However, the percentage removal of methyl orange was slightly
increased at this pH. This suggests another mechanism involving important non-
electrostatic interactions between the delocalised π electrons on the adsorbent surface
and the free electrons of the dye molecule present in the aromatic ring [32]. 30 mol % Mn-
BiFeO3 has pHpzc = 7.80, as shown in Fig. S1. The surface of Mn-BiFeO3 nanocomposite is
positively charged below pHpzc, and the uptake of methyl orange anions increased with
the decrease in the pH value. Therefore, pH 1 was the optimised condition to remove
methyl orange.

3.2.2. Adsorption kinetic studies


The adsorption kinetic parameters are useful for predicting adsorption rate, giving con­
siderable information for designing and modelling the adsorption process, operation
control, and adsorbent evaluation [33]. From the result obtained in Fig. S2, the adsorption
rate occurred quickly at the beginning of the reaction due to the freely available active
sites on the adsorbent’s surface. As the equilibrium time increases, the amount of
adsorption does not drastically increase since the surface has been saturated, and finally
an equilibrium state is obtained. The maximum adsorption capacities of methyl orange on
30 mol% Mn-BiFeO3 reached to 25.2 mg g−1 at 298 K and achieved 23.8 mg g−1 at both
318 K and 328 K.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 13

The pseudo-first-order plot of ln (qe – qt) versus t is shown in Figure 8(a). The values for
k1 and qe were calculated from the slope and intercept derived from the linear equation
from the graph and the values were tabulated in Table 2. The pseudo-second plot of t/qt
versus t is shown in Figure 9(b).
The faster removal rate of methyl orange was achieved at 318 K since the value of h for
methyl orange sorption by 30 mol% Mn-BiFeO3 was higher than at other temperatures.
The amount of methyl orange adsorbed on the catalyst decreased with increasing
temperature, as stated in Table 2. This could be attributed to either reversible adsorption
(back diffusion) controlling mechanism or the exothermic effect of the surroundings
during the adsorption process.
As tabulated in Table 2, the values of the correlation coefficients (R2) for the pseudo-
first-order were very low when the dye adsorption occurred at high temperatures of 318
K, and 328 K and the large difference between calculated and experimental values of the
maximum amount adsorbed indicated that the adsorption of methyl orange onto 30 mol
% Mn-BiFeO3 did not follow the pseudo-first-order model and was not suitable to explain
the reaction mechanism of the dye sorption by the adsorbent. Higher R2 values with the

Figure 8. Pseudo-first-order (a) and pseudo-second-order (b) kinetic plots for methyl orange adsorp­
tion onto 30 mol% Mn-BiFeO3.

Table 2. Kinetic model and parameters for the adsorption of methyl orange onto
30 mol% Mn-BiFeO3 at various temperatures.
Temperature (K)
Kinetic models and its parameters 298 318 328
Pseudo-first-order kinetic
qe, exp (mg/g) 25.200 23.890 23.850
qe, cal (mg/g) 28.460 0.110 0.110
k1 (min−1) 0.083 0.007 0.018
2
R 0.974 0.088 0.230
Pseudo-second-order kinetic
qe, exp (mg/g) 25.200 23.890 23.850
qe, cal (mg/g) 27.100 23.866 23.810
k2 (g/mg min) 0.006 1.170 0.802
h (mg/g min) 4.406 666.000 454.667
2
R 0.999 1.000 1.000
14 A. W. BINTI AZIZ ET AL.

Figure 9. Langmuir (a) and Freundlich (b) isotherm plots for methyl orange adsorption onto 30 mol%
Mn-BiFeO3.

best-fit lines reflected the model’s accuracy applied in the system. A higher value for R2
was obtained in pseudo-second-order at various temperatures: 298 K, 318 K, and 328 K by
which R2>0.99 and qe, cal had a close agreement with the experimental (qe exp), which
indicated that the adsorption rate of methyl orange on the 30 mol% Mn-BiFeO3 could be
more appropriately described using the pseudo-second-order model rather than the
pseudo-first-order model.
The slope of the plot of ln k2 against. 1/T was used to evaluate Ea. The graph is shown in
Fig. S3. The value for the activation energy of the adsorption process could be used to
determine whether it involved a physical or chemical adsorption process. As reported by
other researchers, the physisorption process normally had an activation energy of 5–40
kJ/mol, while chemisorption had a higher activation energy (40–800 kJ/mol) [34]. In this
present experiment, the activation energy obtained from the graph was 144 kJ/mol,
which was more than 40 kJ/mol so it was inferred that the adsorption of methyl orange
onto 30 mol% Mn-BiFeO3 was chemisorption. The heterogeneous adsorbent reaction
involves activation of a reactant molecule by adsorption onto a catalyst surface, which
implies a fairly strong chemical bond formed with the catalyst surface; this mode of
adsorption is called chemisorption, and it is characterised by an enthalpy change typically
greater than 80 kJ/mol and sometimes greater than 400 kJ/mol [34].

3.2.3. Adsorption isotherm studies


Adsorption isotherm is generally used to describe the relationship between the initial
concentration of adsorbate in solution (liquid phase) and adsorbent (solid phase) at
a constant temperature under given conditions such as pH and ionic strength, the fixed
mass of the catalyst, and particles size of adsorbent. To investigate the adsorption
property of methyl orange onto 30 mol % Mn-BiFeO3 nanoparticles, Langmuir and
Freundlich models were applied and examined in detail to determine the equilibrium
nature of adsorption.
Figure 9(a) is the Langmuir plot for the adsorption of methyl orange onto 30 mol%
BiFeO3. The values of KL and qm were calculated from the intercept and slope from the
linear plot of Ce/qe against Ce for methyl orange sorption on 30 mol% Mn-BiFeO3 and
tabulated in Table 3. Negative and zero values for Langmuir isotherm constants for methyl
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 15

Table 3. Langmuir and Freundlich’s parameter for methyl orange adsorption at different
temperature.
Isotherm models and its parameters Temperature (K)
298 318 328
Langmuir
qm, exp (mg/g) 28.9270 29.6700 29.5000
qm, cal (mg/g) −50 × 103 33 × 103 20 × 103
KL (L/mg) −10 × 10−6 1.5 × 10−4 2.5 × 10−5
R2 0.6923 0.1073 0.1546
Freundlich
KF ((mg/g)(L/mg)1/n) 0.499 0.500 0.513
1/n 1.000 0.999 0.990
R2 0.999 0.999 0.998

Table 4. Comparison of adsorption capacity of MO dye onto various


adsorbents.
Adsorbent Adsorption capacity (mg/g) Reference
α-Fe2O3 NPs 28.90 [35]
Nitrogen-doped TiO2 14.10 [36]
Graphene oxide 16.83 [37]
Amorphous CNTs 21.39 [37]
Nanoporous carbon 18.80 [38]
30 mol% Mn-BiFeO3 25.2 This work

orange sorption were obtained. This indicates the inadequacy of the isotherm model to
explain the adsorption process since these constants are indicative of the surface binding
energy and monolayer coverage. Moreover, the amount of adsorption at equilibrium
experimentally is totally different from the amount of adsorption calculated. The correla­
tion coefficient R2 for a linear plot showed very low values, and thus it failed to obey the
Langmuir isotherm model.
Figure 9(b) is the Freundlich plot for the adsorption of methyl orange onto 30 mol%
BiFeO3. The Freundlich isotherm model obtained higher values for R2 at various tempera­
tures: 298 K, 318 K and 328 K by which R2>0.99 as compared to the Langmuir isotherm
model, which indicated that the sorption characteristics of methyl orange were well
described by the Freundlich model and indicated the heterogeneity of the surfaces of
30 mol% Mn-BiFeO3. According to high correlation coefficients R2 for the Freundlich
isotherm model, the Freundlich constant might be appropriate for calculating thermo­
dynamic parameters. Thus, the adsorption of methyl orange using 30 mol% of Mn-BiFeO3
nanoparticles is very well described and fitted by the Freundlich isotherm model.
A summary of MO removal by various adsorbents is shown in Table 4. The adsorption
capacity values for some adsorbents recently reported in the literature are comparable
with that of 30 mol% of Mn-BiFeO3. Thus, 30 mol% of Mn-BiFeO3 is a good adsorbent for
the removal of MO dye from wastewater.

3.2.4. Possible removal mechanism


Many factors play significant roles in the adsorption mechanisms, such as adsorbate
structure and functional group, textual properties and surface chemistry adsorbents
16 A. W. BINTI AZIZ ET AL.

and the specific interactions between the adsorbent surface and the adsorbate [39]. From
the results mentioned previously, Mn substituted BiFeO3 exhibits fast removal of methyl
orange, which contributed to the synergistic effect of manganese. The pH condition of the
solution is one of the basic parameters that have to be accounted for when analysing the
adsorption behaviour of sorbate-sorbent systems and outlining the mechanism of the
process as it affects both the aqueous chemistry and the surface binding-sites of the
adsorbents. In this work, the reaction is enhanced dominantly below pH of pHpzc. The
positive charge of Mn-BiFeO3 below pHpzc is due to the protonation reaction on the
surface-active site. This implies that the uptake of the dye is mainly by electrostatic
attraction between the dye and the Mn-BiFeO3 surface. The latter mechanism may arise
from the interactions between the active site of the Mn-BiFeO3 surface and the free
electron from the methyl orange present in the aromatic rings and multiple bonds. The
adsorption of the azo dye is associated with the oxygen atoms of the sulphonate group,
which is linked to the cation of the Mn-BiFeO3 surface as illustrated in Figure 10.
Methyl orange is a polar molecule, and therefore it could be adsorbed on the
surface of Mn-BiFeO3 via van der Waals attraction force. In addition, as evidenced by
FTIR experiments, OH groups are present on the Mn-BiFeO3 surface. When the
solution pH is low (Figure 10), high concentration H+ ions are available to protonate
the surface of Mn-BiFeO3, which results in high concentration OH groups on the Mn-
BiFeO3 surface, thereby leading to the formation of hydrogen bonds between OH
groups and methyl orange molecules. Besides van der Waals attraction force and
hydrogen bond formation, the charges of adsorbent and adsorbate also play an
important role in the adsorption reaction. At lower pH, it will be changed into its
anionic form of C14H15N3SO3-. On the other hand, the Mn-BiFeO3 surface will be
positively charged via the protonation process, increasing the electrostatic attraction
between methyl orange molecules and the Mn-BiFeO3 surface leading to a higher
adsorption capacity for methyl orange.
To demonstrate further proposed mechanism of methyl orange adsorption on 30 mol
% Mn-BiFeO3, Fourier transform infrared spectroscopy (FTIR) of original methyl orange,
30% mol Mn-BiFeO3 before and after sorption of methyl orange were measured and
compared in the range 4000–400 cm−1 (Figure 11). The original IR spectrum of methyl
orange exhibited a peak at 3508 cm−1 for the N – H stretching vibration. C – H stretching

Figure 10. Mechanism for the influence of pH on methyl orange adsorption: adsorption at low pH.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 17

Figure 11. FTIR spectra of (a) methyl orange (b) 30 mol% of Mn-BiFeO3 after adsorption of methyl
orange (c) 30 mol% of Mn-BiFeO3 before adsorption of methyl orange.

vibrations of – CH3 were located at 2962 cm−1, and the C – C vibration of the benzene
skeleton was observed at 1609 cm−1 and 1589 cm−1. Peaks obtained at 1425 cm−1 reveals
the N=N vibration. Meanwhile, the C – N vibrations were assigned at 1369 cm−1. S=O
vibration was located at 1122 cm−1, and C – H stretching vibrations of the benzene ring
were located at 1032 cm−1. Peaks observed between 703 cm−1 and 579 cm−1 were due to
the C – S stretching vibrations. Based on this figure, the adsorption ability of Mn-BiFeO3
nanoparticles for methyl orange dye ions from an aqueous solution showed a strong
interaction of active sites of heterogeneous Mn-BiFeO3 towards the dye. Extra bands
formed at 1111 cm−1 and 2979 cm−1, as highlighted in the figure showing the presence of
methyl orange anchored to the active sites of the adsorbent. The peak at 1111 cm−1
belongs to the stretching vibrations of S=O and – SO3-. Meanwhile, the absorption peak at
2979 cm−1 is attributed to C-H from CH3 of methyl orange [40]. The slight shift in peak
values may be due to chemical bonds between absorbent and functional groups of
methyl orange.
The representation of methyl orange sorption onto Mn-BiFeO3 nanoparticles is illu­
strated in Scheme 2.

3.2.5. Reusability and recyclability of the adsorbent


Nowadays, industrial wastewater treatment strategies focus on developing management
practices and green technologies for environmental benefit. Therefore, the reusability of
Mn-BiFeO3 is a significant and effective factor for sustainable wastewater treatment. In
this work, the reusability studies of 30 mol% Mn-BiFeO3 were evaluated and compared
between reusable cycles, as shown in Figure 12.
18 A. W. BINTI AZIZ ET AL.

Scheme 2. Representation of methyl orange sorption using Mn-BiFeO3.

100
(a)
90
80
Removal efficiency

70
60
50
40
30
20
10
0
1 2 3 4 5 6
No. of cycles

(b)

5μm

Figure 12. (a) Reusability studies of 30 mol % Mn-BiFeO3 on methyl orange removal and (b) FESEM of
used catalyst after 6th treatment cycle.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 19

From the results obtained, the first three cycles removed 99.36%, 99.16%, and
98.31% of the dye, respectively. Nevertheless, for the next fourth cycle, the efficiency
of the adsorbent marked a decrease, where the percentage removal was left to
70.35%. The percentage removal of methyl orange kept decreasing to 57.21% and
30.00% when the spent adsorbent was used for the fifth and sixth cycles. The
observed decrease in the percentage removal might be due to the mass loss of
adsorbent during the separation of adsorbent in methyl orange solution and when
the washing step process took place. It can be seen from Figure 12(b) the agglom­
eration of the catalyst surface that contribute to the reduction of adsorption activity.
However, heterogeneous Mn-BiFeO3 is an excellent adsorbent with good reusability
and stability for removing methyl orange.

4. Conclusion
The present investigation has shown that the modification of BiFeO3 via substituting variable
amounts of manganese could be a promising approach to remove high amount of MO via
adsorption over a wide dye concentration range. Processing parameters Including pH value
of the dye, % mol of Mn-substituted BiFeO3, dosage of the catalyst, and initial dye concen­
tration at various times and temperatures were found to affect the adsorption process. The
results showed that higher MO removal efficiency could be achieved using Mn substituted in
BiFeO3 as compared to pristine BiFeO3. A maximum adsorption capacity of 25.2 mg/g was
achieved using 30% mol Mn-BiFeO3 and the process was found to follow the pseudo-second
-order kinetic model. Moreover, the data obtained showed that, the adsorption of MO onto
the surface of the adsorbent is more fitted to the Freundlich isotherm than the Langmuir
isotherm. Interestingly, 30% Mn-BiFeO3 had shown consistently high MO removal efficiency
of up to 98% when used repeatedly for three cycles, and therefore the adsorbent may have
great potential for the elimination of MO from wastewater.

Acknowledgments
Special acknowledgment goes to the School of Chemical, Universiti Sains Malaysia, and Nagaoka
University of Technology, Japan, for the chemical analysis.

Disclosure statement
No potential conflict of interest was reported by the author(s).

Funding
This research was supported by the Fundamental Research Grant Scheme (FRGS) 203/PKIMIA/
6711792 from the Ministry of Higher Education and Research University Grant (RUI) 1001/PKIMIA/
8011069 allocated by Universiti Sains Malaysia.
20 A. W. BINTI AZIZ ET AL.

References
[1] F.E. Titchou, H. Zazou, H. Afanga, J. El Gaayda, R.A. Akbour, P.V. Nidheesh and M. Hamdani,
Chem. Eng. Pro. Inten. 169, 108631 (2021). doi:10.1016/j.cep.2021.108631.
[2] S.G. Poulopoulos, A. Yerkinova, G. Ulykbanova and V.J. Inglezakis, PLoS One. 14, 0216745
(2019). doi:10.1371/journal.pone.0216745.
[3] M. Ikram, E. Umar, A. Raza, A. Haider, S. Naz, D.A. Ul-Hamid, J. Haider, I. Shahzadi, J. Hassan and
S. Ali, RSC Adv. 10 (41), 24215–24233 (2020). doi:10.1039/D0RA04851H.
[4] S. Keerthana, R. Yuvakkumar, G. Ravi, S. Hong, A.G. Al-Sehemi and D. Velauthapillai,
Chemosphere. 293, 133540 (2022). doi:10.1016/j.chemosphere.2022.133540.
[5] R.R. Karri, M. Tanzifi, M.T. Yaraki and J. Sahu, J. Env. Man. 223, 517–529 (2018). doi:10.1016/j.
jenvman.2018.06.027.
[6] L. Zhai, Z. Bai, Y. Zhu, B. Wang and W. Luo, Chin. J. Chem. Eng. 26 (3), 657–666 (2018). doi:10.
1016/j.cjche.2017.08.015.
[7] A. Siyasukh, Y. Chimupala and N. Tonanon, Carbon. 134, 207–221 (2018). doi:10.1016/j.
carbon.2018.03.093.
[8] A.M. Zayed, M.S.A. Wahed, E.A. Mohamed and M. Sillanpää, App. Clay. Sci. 166, 49–60 (2018).
doi:10.1016/j.clay.2018.09.013.
[9] A. Boughelout, R. Macaluso, M. Kechouane and M. Trari, Reac. Kin. Mech. Cat. 129 (2),
1115–1130 (2020). doi:10.1007/s11144-020-01741-8.
[10] A. Pandey, P. Singh and L. Iyengar, Int. biodet. biodeg. 59 (2), 73–84 (2007). doi:10.1016/j.
ibiod.2006.08.006.
[11] Y. Tang, R. Yang, D. Ma, B. Zhou, L. Zhu and J. Yang, Polym. Polym. Comp. 26 (2), 161–168
(2018). doi:10.1177/096739111802600204.
[12] S. Chen, Y. Huang, X. Han, Z. Wu, C. Lai, J. Wang, Q. Deng, Z. Zeng and S. Deng, Chem. Eng. J.
352, 306–315 (2018). doi:10.1016/j.cej.2018.07.012.
[13] M. Hasanpour, S. Motahari, D. Jing and M. Hatami, Chemosphere. 284, 131320 (2021). doi:10.
1016/j.chemosphere.2021.131320.
[14] S. Bao, K. Li, P. Ning, J. Peng, X. Jin and L. Tang, J. Taiwan Inst. Chem. Eng. 87, 64–72 (2018).
doi:10.1016/j.jtice.2018.03.009.
[15] S. Hosseini, M.A. Khan, M.R. Malekbala, W. Cheah and T.S. Choong, Chem. Eng. J. 171 (3),
1124–1131 (2011). doi:10.1016/j.cej.2011.05.010.
[16] Z. Liu, W. He, Q. Zhang, H. Shapour and M.F. Bakhtari, ACS. Omega. 6, 4597–4608 (2021).
doi:10.1021/acsomega.0c05091.
[17] K. Tan and B. Hameed, J. Taiwan Inst. Chem. Eng. 74, 25–48 (2017). doi:10.1016/j.jtice.2017.01.
024.
[18] J. Shen, X. Wang, L. Zhang, Z. Yang, W. Yang, Z. Tian, J. Chen and T. Tao, J. Clean. Prod. 184,
949–958 (2018). doi:10.1016/j.jclepro.2018.03.015.
[19] Y. Jiang, B. Liu, J. Xu, K. Pan, H. Hou, J. Hu and J. Yang, Carbo. Polym. 182, 106–114 (2018).
doi:10.1016/j.carbpol.2017.10.097.
[20] A.O.A. El Naga, S.A. Shaban and F.Y. El Kady, J. Taiwan Inst. Chem. Eng. 93, 363–373 (2018).
doi:10.1016/j.jtice.2018.07.044.
[21] N.S.A. Satar, R. Adnan, H.L. Lee, S.R. Hall, T. Kobayashi, M.H.M. Kassim and N.H.M. Kaus, Cer.
Intern. 45 (13), 15964–15973 (2019). doi:10.1016/j.ceramint.2019.05.105.
[22] S.W.C. Chien, D.-Q. Ng, D. Kumar, S.-M. Lam and Z.H. Jaffari, J. Phy. Che. Sol. 160, 110342
(2022). doi:10.1016/j.jpcs.2021.110342.
[23] S.M. Lam, Z.H. Jaffari, J.C. Sin, H. Zeng, H. Lin, H. Li and A.R. Mohamed, Coll. Surf. A: Physico.
Eng. Asp. 614, 126138 (2021). doi:10.1016/j.colsurfa.2021.126138.
[24] Z.H. Jaffari, S.M. Lam, J.C. Sin, H. Zeng and A.R. Mohamed, Sep. Pur. Tech. 236, 116195 (2020).
doi:10.1016/j.seppur.2019.116195.
[25] N.H.M. Kaus, S.S. Imam, A.W. Aziz, H.L. Lee, R. Adnan and M.L. Ibrahim, Coll. Surf. A: Physico.
Eng. Asp. 615, 126294 (2021). doi:10.1016/j.colsurfa.2021.126294.
[26] S. Lagergren, Handlingar. 24, 1–39 (1898).
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 21

[27] Y.S. Ho and G. McKay, Chem. Eng. J. 70 (2), 115–124 (1998). doi:10.1016/S0923-0467(98)
00076-1.
[28] I. Langmuir, J. Amer. Chem. Soc. 40 (9), 1361–1403 (1918). doi:10.1021/ja02242a004.
[29] M.A. Darweesh, M.Y. Elgendy, M.I. Ayad, A.M.M. Ahmed, N.K. Elsayed and W. Hammad, S. Afri.
J. Chem. Eng. 40, 10–20 (2022). doi:10.1016/j.sajce.2022.01.002.
[30] S. Irfan, Y. Shen, S. Rizwan, H.C. Wang, S.B. Khan and C.W. Nan, J. Amer. Cer. Soc. 100 (1), 31–40
(2017). doi:10.1111/jace.14487.
[31] D. Ramimoghadam, M.Z. Bin Hussein and Y.H. Taufiq-Yap, Chem. Cent. J. 7 (1), 1–10 (2013).
doi:10.1186/1752-153X-7-1.
[32] M. Belhachemi and F. Addoun, App. Wat. Sci. 1 (3), 111–117 (2011). doi:10.1007/s13201-011-
0014-1.
[33] F. Deniz, Sci. Wor. J. 2013, 961671 (2013). doi:10.1155/2013/961671.
[34] C.H. Wu, J. Hazard. Mat. 144 (1–2), 93–100 (2007). doi:10.1016/j.jhazmat.2006.09.083.
[35] A. Debnath, K. Deb, K.K. Chattopadhyay and B. Saha, Desal. Wat. Treat. 57 (29), 13549–13560
(2016). doi:10.1080/19443994.2015.1060540.
[36] J. Fan, Z. Zhao, W. Liu, Y. Xue and S. Yin, J. Coll. Interf. Sci. 470, 229–236 (2016). doi:10.1016/j.
jcis.2016.02.045.
[37] D. Robati, B. Mirza, M. Rajabi, O. Moradi, I. Tyagi, S. Agarwal and V. Gupta, Chem. Eng. J. 284,
687–697 (2016). doi:10.1016/j.cej.2015.08.131.
[38] S. Kundu, I.H. Chowdhury and M.K. Naskar, J. Mol. Liq. 234, 417–423 (2017). doi:10.1016/j.
molliq.2017.03.090.
[39] A. Roy, S. Chakraborty, S.P. Kundu, B. Adhikari and S. Majumder, Ind. Eng. Chem. Res. 51 (37),
12095–12106 (2012). doi:10.1021/ie301708e.
[40] M. Arshadi, M. Mehravar, M. Amiri and A. Faraji, J. Coll. Inter. Sci. 440, 189–197 (2015). doi:10.
1016/j.jcis.2014.10.053.

You might also like