Fe Ni SB

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

The relative performance of gelatin hydrogels doped with nanostructured


transition-metal ferrites for chromium (VI) removal: Packed bed column
studies and insights into reduction mechanisms
M. Priyadarshini a, b, E. Rekha a, b, Asha Sathish a, b, K. Nithya c, d, *
a
Department of Sciences, Amrita School of Physical Sciences, Coimbatore, Amrita Vishwa Vidyapeetham 641112, India
b
Functional Materials Laboratory, Amrita School of Engineering, Coimbatore, Amrita Vishwa Vidyapeetham, 641112, India
c
Department of Chemical Engineering and Materials Science, Amrita School of Engineering, Coimbatore, Amrita Vishwa Vidyapeetham, 641112, India
d
Centre of Excellence in Advanced Materials and Green Technologies (CoE-AMGT), Amrita School of Engineering, Coimbatore, Amrita Vishwa Vidyapeetham, 641112,
India

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

Keywords: The article comprehends investigations using gelatin hydrogels reinforced with cobalt and nickel ferrites to
Chromium remove chromium ions. The synthesized ferrite nanoparticles are mesoporous with high surface areas of 132.059
Nickel m2/g and 141.792 m2/g. The deposition of chromium ions on the dissected surfaces is more evident from the
Cobalt
morphological studies. Supporting evidence from thermal studies showed how the nanofiller encapsulation
Hydrogels
increased hydrogels’ thermal stability. The bed capacity in the packed bed column is 32.48 mg/g and 31.81 mg/g
Nanoparticles
Column for cobalt ferrite and nickel ferrite nanoparticles entrapped hydrogels, respectively. X-ray Photoelectron spec­
troscopy findings indicated that the presence of characteristic peaks at 578.68 eV and 591.16 eV is assigned to
hexavalent chromium, and the peaks observed at 575.56 and 583.71 eV correspond to the trivalent state of
chromium. The electron transfer from the iron centre to chromium (VI) favoured the reduction of chromium (VI)
to less toxic chromium (III)

Introduction compared to Cr (VI).


For decades, there has been a significant investment in research on
Water is one of the foremost resources available on this planet, and the permanent reduction of hexavalent chromium to trivalent chromium
every life form on Mother Earth has a basic entitlement to clean, pure [1]. However, a substantial solution has not yet been arrived at to
water. However, one-third of the world’s biodiversity has already been combat the widespread chromium pollution brought on by industries
destroyed due to water quality degradation. Industrial exploitation of worldwide [2]. Even though there are several procedures available for
freshwater reservoirs without adequate wastewater recycling increases the removal of heavy metal ions, such as ion exchange, chemical pre­
the exploitation index and decreases the water table the day after the cipitation, chemical coagulation, membrane filtration, and advanced
day. Moreover, heavy metal contamination caused by industries con­ oxidation processes, each has its limitations. Among all these methods,
tinues to be a severe threat and a significant contributor to water Adsorption: a mass-transfer phenomenon, appears to be a successful and
pollution. Heavy metals are poisonous, carcinogenic, and highly toxic environmentally compatible approach to removing heavy metal ions. It
and cause significant health problems such as cancer, gastrointestinal is a viable technique that is also economical and does not demand highly
disorders, skin abrasions, cardiovascular ailments, and neurological skilled operation. There are different types of adsorbents, such as silica
problems. Chromium is a human carcinogen in the − 2 to + 6 oxidation gel [3,4] activated carbons, zeolites, clay minerals, nanostructured
state, which is mainly stable in three forms, i.e., Cr (0), Cr (III), and Cr materials[5], biochar[6], agricultural waste etc.[7–9]. Numerous
(VI). Of all three states, hexavalent chromium is highly toxic and research articles have been published using these adsorbents in heavy
mutagenic. On the other hand, Cr (0) and Cr (III) are not very toxic when metal removal applications [10]. For instance, Jyotikusum Acharya and

* Corresponding author at: Department of Chemical Engineering and Materials Science, Amrita School of Engineering, Coimbatore, Amrita Vishwa Vidyapeetham,
641112, India.
E-mail address: [email protected] (K. Nithya).

https://doi.org/10.1016/j.jiec.2024.03.001
Received 19 October 2023; Received in revised form 20 February 2024; Accepted 2 March 2024
Available online 3 March 2024
1226-086X/© 2024 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

the team’s research used agricultural wastes, such as rice husk, to the soft ferrite materials primarily used as catalysts, gas-sensors, and
remove heavy metal ions from wastewater. Although silica gels, acti­ magnetic fluids.
vated carbons[11], and zeolites[12,13] are frequently used as adsor­ Despite their extraordinary surface properties, nanoparticles are
bents [14]and are effective in the treatment of wastewater, they are a bit known to have poor chemical stability in some cases due to their ten­
more expensive[15]. However, it is impressive that bioadsorbents are dency to form agglomerates, resulting in unfavorable conditions. To
readily available, plentiful, affordable, non-toxic, and most preferable overcome this problem, nanomaterials are either blended or embedded
for environmental cleanup[16–18].In addition, other distinctive on the surface of biopolymer-based hydrogels[26] to obtain a core–shell-
adsorption features include high sorption capacity, reusability, and like structure [27,28]. Three generations of biopolymers are utilized in
recyclability. In the modern world, the importance of nanotechnology is hydrogels[29,30] 1. Chitosan, 2. Alginate, and 3. Gelatin[31,32].
well-known, and its concepts are used in every branch of science. When Several nanofillers were incorporated into the polymer matrix for water
nano adsorbents replace conventional adsorbents, the process is called purification applications all these years[33]. Notably, metal oxide
nano adsorption. Indeed, the adsorption capacity of nanosorbents is nanoparticles have been widely used in wastewater treatment applica­
much higher than conventional adsorbents. tions in the last few years. There are immense scientific contributions
Recently, prominent research has occurred in the field of nanotech­ worldwide in using metal ferrites and composites for water treatment.
nology for the removal of contaminants from wastewater[19]. This is However, to the author’s knowledge, hardly any reports compare the
because nanomaterials have a larger adsorption capacity owing to the adsorption performance of nanocomposites when two different nano­
high surface area. Further, the availability of tiny pores distributed fillers of the same kind are encapsulated in the polymer matrix. Indeed,
throughout their surface and a well-defined crystal structure are added the work on gelatin polymers as hydrogels for wastewater treatment
features[20]. Due to their intriguing characteristics and outstanding applications is also limited. Understanding the unique interface between
performance efficiency, nanocomposites have also seen a boom in the nanofillers and the polymer matrix is essential for successfully
development as adsorbents in recent years[21]. Carbon-based nano­ impregnating nanofillers into the polymer matrix. This intricate
fillers, such as CNT and graphene, nanoparticles of metal oxides or hy­ polymer-nanofiller relationship also decides the improved performance
droxides, magnetic nanoparticles such as ferrites, Fe2O3, and Fe3O4, of the overall composite structure[34].Key attributes that determine the
nano cellulosic materials, zerovalent metal nanoparticles, and engi­ effective impregnation of nanofillers include the specific surface area of
neered nanoparticles[22] are few examples of nanomaterials that have nanoparticles, filler loading, size, availability of interface interaction
been recently reported in adsorption applications [23].Among these sites, interfacial free energy, orientation and wetting coefficient. With
nanoparticles, so many publications have reported the application of reference to research on hydrogels and their applications, gelatin
iron-oxide nanoparticles due to their remarkable surface characteristics hydrogels are extensively used in drug delivery, wound dressing, and the
and magnetic properties [24]. Recently cobalt-ferrite and nickel ferrite development of bone-graft materials in biomedical research, and only a
nanoparticles also received considerable attention in materials research. few investigations have happened in adsorption.
This is because of their chemical stability, excellent photocatalytic Compared to biopolymers such as starch, chitosan, and alginates,
ability, anti-microbial characteristics, ferromagnetic properties, high little work has been explored using gelatin-based hydrogels as nano­
electrochemical stability and promising electromagnetic features. sorbents for chromium removal applications. Due to gelatin hydrogels’
Notably, cobalt ferrite nanoparticles are used in drug delivery, catalysis, low mechanical and thermal resistance, stable and functional re­
magnetic resonance imaging, ferrofluids, gas sensors, and microwave inforcements are indispensable to deploy in water treatment applica­
devices [25].On the other hand, nickel ferrite nanoparticles are one of tions. Gelatin hydrogels’ stability and binding capacity can be enhanced

Fig. 1. Synthesis of CoFe2O4 and NiFe2O4 nanoparticles.

145
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 2. Optimization of weight % composition of CoFe2O4 and NiFe2O4 hydrogels.

by reinforcing the bio-polymer framework with nano-fillers [30]. hexahydrate are mixed using a magnetic stirrer. At the same time, in
Although this may be true, only limited works have experimented with another beaker, 0.1 M cobalt nitrate hexahydrate and 0.2 M ferric ni­
different nano-based reinforcements in gelatin hydrogels [35]. More­ trate nonahydrate solutions were mixed uniformly with a magnetic
over, good binding capacity and excellent thermal and mechanical stirrer. 3 M sodium hydroxide solution was slowly added dropwise to the
resistance are crucial when these hydrogels are packing mediums in stirring solution as a precipitating agent in both beakers. The solution’s
continuously packed bed columns. Published reports on the nanodoped pH was frequently measured while continuously adding NaOH until the
gelatin hydrogels for continuous packed bed columns are also scarce pH became 12. The liquid precipitate was then heated at about 80◦ C and
[36]. In light of the above discussions, the novelty of the proposed work constantly agitated for 40 min, allowing it to stand at room temperature.
is pointed toward the impregnation of nanostructured transition-metal Different methods were adopted to wash the precipitates of CoFe2O4 and
ferrites onto gelatin hydrogels. This would make the gelatin hydrogels NiFe2O4 nanoparticles based on their physical and chemical properties.
tough, resistant, long-lasting, and highly adsorptive. To the best of our In the case of nickel ferrite, the precipitate was centrifuged and washed
knowledge, only a few works have reported the entrapment of ferrite- with deionized water until the pH of the solution reached neutral. On the
based nanofillers onto gelatin hydrogels. Also, the use of such hydro­ other hand, since cobalt ferrite is a hard magnetic material, the pre­
gels for chromium removal is limited. Thus, the work undertaken is cipitate was magnetically separated. Finally, both precipitates were
devoted for the first time to evaluating the difference in adsorption ca­ washed with ethanol to eliminate undesired contaminants, followed by
pacity between neighbouring transition metal elements, i.e., Cobalt (Z overnight drying at around 80◦ C. The product was finally milled into a
= 27) and Nickel (Z = 28) with and without entrapping gelatin for fine powder and calcined at a temperature of 600◦ C for 6 h, and thus
chromium removal. pure cobalt ferrite (CoFe2O4) and nickel ferrite (NiFe2O4)) nanoparticles
For large-scale applications, continuous moving bed and fluidized were obtained. [Fig. 1].
bed sorption systems are more expensive and complicated when To synthesize the cobalt-ferrite and nickel ferrite gelatin hydrogels,
compared to continuous fixed bed adsorption studies. For this reason, four different weights % composition (1–4) of filler i.e., cobalt ferrite
batch studies are scaled up using a fixed bed column. Considerable and nickel ferrite, respectively, were tried. The performance efficiency
attention is shown in the design and tuning of hydrogel synthesis for of each hydrogel obtained at different weight ratios is tested [Fig. 2]. As
selective removal of chromium ions. The study also delves more into seen in the figure, the efficiency of gelatin hydrogels decreased with an
studies related to the binding mechanism of chromium and to under­ increase in the CoFe2O4 and NiFe2O4 weight ratios. The same trend is
stand the influence of nanofillers on adsorption. The work also provides observed for cobalt ferrite and nickel ferrite hydrogels; therefore, cobalt
more perspectives on studies related to the regeneration of the adsorbent and nickel ferrite nanoparticles of one weight percentage are entrapped
bed after repeated cycles. The findings of the proposed work could lead with NiFe2O4 and CoFe2O4 hydrogels.
to effective wastewater management from industries that release The synthesis of CoFe2O4 and NiFe2O4 hydrogels is as follows. The
wastewater containing chromium and hold potential implications in synthesized cobalt ferrite and nickel ferrite nanoparticles were sonicated
wastewater recycling. separately into a suspension using deionized water. A homogeneous
dispersion was obtained, and an appropriate amount of gelatin was
Methodology added to each of these CoFe2O4 and the NiFe2O4 dispersions. As dis­
cussed above, the weight ratio of the nanofillers used throughout the
Preparation of gelatin hydrogels doped with nanostructured transition- synthesis is set at 1 % for both cobalt and nickel ferrite nanoparticles.
metal ferrites This mixture was magnetically stirred for six to seven hours at a tem­
perature of 50 to 60 ◦ C. A portion of this mixture was dropped at a
Firstly, 0.2 M ferric chloride hexahydrate and 0.1 M nickel chloride controlled flow rate in a beaker containing sunflower oil in a cold-ice

146
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 3. Synthesis of CoFe2O4 and NiFe2O4 hydrogels.

Fig. 4. CoFe2O4 and NiFe2O4 hydrogels after washing and drying.

bath. The resultant hydrogels were undisturbed in the refrigerator for 24 information on the thermal stability of the NiFe2O4 and CoFe2O4
to 48 h and then transferred to a 1 % glutaraldehyde solution for cross- hydrogels. Further, the elemental composition of the nanoparticles and
linking, increasing the beads’ mechanical strength. The cross-linked NiFe2O4, CoFe2O4 hydrogels after adsorption is predicted using X-Ray
beads were kept undisturbed for drying for 24 h and the beads were photoelectron spectroscopy. The XPS analysis also investigated how
then rinsed with deionized water to wash the residuals of glutaraldehyde chromium ions are bound onto the hydrogel surfaces.
[Fig. 3]. The resultant NiFe2O4 and CoFe2O4 gelatin hydrogels [Fig. 4]
were used to conduct further experiments. Studies on the determination of binding capacity of cobalt and nickel
ferrite hydrogels for the uptake of chromium ions
Nanostructured Transition-Metal ferrites and gelatin hydrogels
Characterization The CoFe2O4 and NiFe2O4 nanoparticles and the nanostructured
transition metal ferrites doped hydrogels were brought into contact with
Prominent adsorbent features, such as surface morphology, and the chromium solution of different concentrations. First, the hydrogels’
functional, structural, elemental, and chemical properties, are evaluated performance efficiency is studied in a static system. Process parameters
using modern analytical tools. X-Ray diffraction studies investigated the related to batch studies were optimized, including pH, temperature,
phase identification of crystalline cobalt and nickel ferrite nanoparticles. contact time, and concentration. The working concentrations of chro­
Fourier Transform Infra Red spectroscopy gave deep insights into the mium were prepared using 1000 mg/L chromium solution. The zero-
nature of the functional groups on the hydrogel’s surfaces. On the other point charge of the hydrogels was determined by the pH drift method.
hand, Field Emission Scanning electron Microscopy predicted the sur­ Different pH solutions from 1 to 12 were prepared using 0.1 N HCl and
face morphology and texture, and Energy dispersive X-ray spectroscopy 0.1 N NaOH solution. The NiFe2O4 and CoFe2O4 hydrogels were added
determined the nanocomposite’s elemental composition. Another crit­ to 50 mL of each NaCl solution in a conical flask and agitated for 24 h.
ical characterization is the Brunauer, Emmett and Teller [BET] analysis, After 24 h, the final pH of the solution was tested using a pH meter. The
which helped in understanding the surface area and porosity of the difference between the initial pH and final pH was noted as ΔpH. The
nanoparticles. In addition, the thermogravimetric analysis provided ΔpH values were plotted against the solution’s initial pH, and the

147
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 5. XRD profiles of CoFe2O4 and NiFe2O4 nanoparticles.

intersection point is marked as the hydrogel’s zero-point charge. Fig. 6. FTIR Spectrum of CoFe2O4, NiFe2O4 nanoparticles and hydrogels.
Studies on fixed bed columns are vital to find the hydrogels’ suit­
ability for industrial or large-scale applications. In view of that, the with the JCPDS file (89–4927), and the reflections at 30.2◦ , 35.6◦ , 37.1◦ ,
batch results were scaled up, and a fixed bed column was designed. A 43.3◦ , 54.0◦ , 57.3◦ , and 63.0◦ correspond to the crystallographic planes
cylindrical acrylic column of length 30 cm and height 2.5 cm was of [h, k, l] of nickel ferrite nanoparticles, respectively. At the same time,
fabricated. The column packing was supported with glass wool to pre­ the diffractograms of CoFe2O4 nanoparticles were also compared with
vent the loss of hydrogels. Flow rate (mL/min) and bed depth (cm) the JCPDS file of cobalt ferrite nanoparticles. The obtained peaks align
varied at 2 mL/min, 3 mL/min, and 1 cm and 2 cm, respectively. The with the actual JCPDS card no (22–1086), and the reflections corre­
composite material’s weight was varied according to the column’s bed spond to crystallographic planes [h,k,l] of cobalt-ferrite nanoparticles.
depth. Breakthrough curves were obtained from the results obtained The sharp peaks in the XRD diffractograms of both CoFe2O4 and NiFe2O4
from the fixed bed column. This is vital to estimate the mass of the nanoparticles indicate the crystalline nature of the sample. The crys­
chromium adsorbed on the column. The effluent was collected every 60 tallite size of the nanoparticles was determined using the well-known
min, and the concentration of the solution was checked using a UV–vi­ Debye-Scherrer equation.
sible spectrophotometer.
0.9λ
d= (1)
βCOSθ
Swelling studies and desorption and regeneration analysis of CoFe2O4 and
0.9 refers to the Debye-Scherrer constant represented by the symbol
NiFe2O4 Hydrogels
’k’, and λ corresponds to a wavelength of Kα radiation of the Cu atom.
“β” is obtained from the FWHIM value (peak width having maximum
The swelling behaviour of the hydrogels was studied by immersing
intensity) and θ is half the value of Bragg’s angle. After applying the
both bare gelatin hydrogels and nanostructured transition metal ferrites
above equation, the crystallite size of CoFe2O4 and NiFe2O4 nano­
doped hydrogels for 24 h in cross-linking agents with different weight
particles was 11.47 nm and 12.08 nm, respectively.
percentage compositions (0.5 to 1.5 %). The swelling studies are also
Another interesting surface characterization tool is FTIR, which
carried out in de-ionized water. The initial weight of the beads was
studies the reactive surface sites responsible for adsorption. The infrared
measured and denoted as Wi. The final weight after swelling was taken
waves of the mid-IR range were allowed to pass through the desired
as Wf.. The % ratio of the difference between final and initial weight and
sample of interest with the help of an FT-IR spectrophotometer [Alpha II
that of initial weight gave the swelling %.
Compact, Bruker USA]. After interacting with characteristic infrared
The regeneration of hydrogels is very significant and of practical
light radiation, the sample’s transmittance was measured and recorded
importance. For this reason, the reusability of the hydrogel was also
as a spectrum. Fig. 6 represents the Infrared spectrum of cobalt ferrite
tested by carrying out desorption and re-adsorption processes for five
and nickel ferrite nanoparticles as well as the CoFe2O4 and NiFe2O4
cycles. The desorption process was carried out by using sodium hy­
hydrogels. In the IR spectrum of the cobalt ferrite nanoparticles, the
droxide (0.1 N) as the eluting agent. The chromium-loaded beads and
peaks at 573.99 cm− 1 and 421.11 cm− 1 represent stretching vibrations
40 mL of 0.1 N NaOH solution are added and agitated for almost an
of the Iron-oxygen [ Fe3+–O2–] and Cobalt-oxygen [Co2+–O2–] bond
hour. The filtrate obtained was then tested for the presence of chromium
respectively. On the other hand, the peak at 579.87 cm− 1 corresponds to
and desorption efficiency was calculated.
Fe3+- O2– bond, and the peak at 424.57 cm− 1 is attributed to Ni2+- O2–
bond for nickel ferrite nanoparticles. Summarising these findings, it
Results and Discussions becomes evident that the synthesized nanoparticles are cobalt ferrite
and nickel ferrite nanoparticles. Furthermore, it is essential to analyse
Structural characterisation and FTIR spectrum Analysis the IR spectrum of CoFe2O4 and NiFe2O4 entrapped hydrogels. The
characteristic peaks obtained at 3290.80 cm− 1 and 1635.08 cm− 1 are
The CoFe2O4 and NiFe2O4 nanoparticles were subjected to X-ray related to gelatin’s N–H and carbonyl stretching band. Additionally,
diffraction analysis to draw inferences about structural characterization gelatin’s aliphatic C–H bond stretching is observed at 2926.11.
and crystallite size. Fig. 5 shows the XRD pattern of cobalt ferrite and cm− 1 wavenumbers. These results are indications for gelatin
nickel ferrite nanoparticles. The nickel ferrite’s XRD pattern is in line

148
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 7. Scanning electron micrographs of [a] and [b] NiFe2O4 nanoparticles, [c] and [d] CoFe2O4 nanoparticle. [e] and [f] cross-sectional view of NiFe2O4 hydrogel
before adsorption, [g] and [h] cross-sectional view of NiFe2O4 hydrogel after adsorption [i] and [j] cross-sectional view of CoFe2O4 hydrogel before adsorption, [k]
and [l] cross-sectional view of CoFe2O4 hydrogel after adsorption.

encapsulation. Another critical observation is the recurrence of iron- of hydrogels. The cross-sectional SEM analysis of hydrogels shows how
oxygen, cobalt-oxygen, and nickel-oxygen stretching vibration at these nanoparticles are entrapped in the gelatin matrix (Fig. 7e-i and
551.58 cm-1, 418.15 cm-1, and 420.06 cm− 1, respectively. The suc­ Fig. 7h and 7 L). The cross-sectional views of NiFe2O4 and CoFe2O4
cessful impregnation of CoFe2O4 and NiFe2O4 nanoparticles in the hydrogels, before adsorption are shown in Fig. 7e and 7f, and Fig. 7i and
gelatin hydrogels is thus validated. 7j respectively. Branch-like projections on the microscale [Fig. 7e and
7i] demonstrate that the hydrogels are smooth and tender. This obser­
Surface topography through FE-SEM analysis and EDX studies vation is seen in both CoFe2O4 and NiFe2O4 hydrogels. Another obser­
vation is the appearance of some spherical clusters at nanoscale
Hydrogels and nanoparticles’ surface morphology and topography dimensions [Fig. 7f and 7j]. These observations signify the encapsula­
are inferred using Scanning Electron Microscopy. The FE-SEM micro­ tion of CoFe2O4 and NiFe2O4 nanoparticles into gelatin hydrogels.
graphs of NiFe2O4, CoFe2O4 nanoparticles, and hydrogels are illustrated Compared to CoFe2O4 hydrogels, the encapsulation in the case of
in Fig. 7. Fig. 7a- 7d represents the SEM images of nanoparticles. The NiFe2O4 hydrogels is localized and concentrated over a particular region
images’ a’ and ’b’ denote NiFe2O4 nanoparticles, and ’c’ and ’d’ are the of space. This localization may result in improved interactions between
pictures of CoFe2O4 nanoparticles. Fig. 7a and 7b both demonstrate adsorbate and adsorbent molecules facilitating the uptake of more
clustered spherical NiFe2O4 nanoparticles. The agglomeration is due to chromium ions by NiFe2O4 hydrogels on par with CoFe2O4 hydrogels
the high surface energy possessed by these nanoparticles [37]. Alter­ [40].
natively, SEM images of CoFe2O4 nanoparticles exhibited cubical and Interesting findings have also been observed in the SEM micrograph
spinel geometry [Fig. 7c and 7d]. CoFe2O4 nanoparticles are less of hydrogels after adsorption (Fig. 7g and 7 k). The cross-sections of the
agglomerated than NiFe2O4 nanoparticles, and nanocrystalline features hydrogels post-adsorption exhibit some tiny bubble-like structures in the
are distributed throughout the surface [38]. This is due to the decreased NiFe2O4 hydrogel (Fig. 7g). This shows how the NiFe2O4 hydrogel is
crystallite size of CoFe2O4 nanoparticles over NiFe2O4 nanoparticles. modified structurally after adsorption. In contrast, no such bubbles are
This intriguing structural feature also adds to CoFe2O4 nanoparticles’ seen in CoFe2O4 hydrogels [Fig. 7k], but some noticeable structural
better binding capacity over NiFe2O4 nanoparticles [39]. deformations are still seen (Fig. 7e). Moreover, the smooth and tender
The SEM studies also delve into investigations on the surface of the surface of CoFe2O4 hydrogels disappeared after adsorption. The SEM
NiFe2O4 hydrogels. The hydrogel’s surface is comprised of numerous analysis performed on hydrogels after adsorption suggests a favourable
cracks and crevices in nanoscale dimensions. Apart from cracks and chromium removal in NiFe2O4 hydrogels compared to CoFe2O4
crevices, hollow cavities and small openings are also witnessed in the hydrogels.
case of NiFe2O4 hydrogel [Fig. 7h]. Nevertheless, this porous structure is The chemical characterisation of the NiFe2O4, CoFe2O4 nano­
absent in CoFe2O4 hydrogel. These structural features enable the un­ particles and hydrogels were analysed using EDX analysis. Fig. 8 rep­
derstanding that the passage of nanoparticles into the NiFe2O4 hydrogel resents the EDX spectrum of CoFe2O4, NiFe2O4 nanoparticles, and
is very much possible. This factor contributes to the better entrapment of hydrogels. In the case of CoFe2O4 nanoparticles [Fig. 8a], 52.60 % of the
nanoparticles by NiFe2O4 hydrogels than CoFe2O4 hydrogels, leading to total elemental composition is Iron, 24.47 % is oxygen, and cobalt, on
high-performance efficiency. the other hand, is 22.93 % of the constituent elemental composition.
Equally important is the analysis of cross-sectional SEM micrographs Subsequent analysis of NiFe2O4 nanoparticles shows 28.32 % Iron,

149
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 8. Energy dispersive X-ray spectrum of [a] CoFe2O4 nanoparticles [b] NiFe2O4 nanoparticles, CoFe2O4 hydrogels: [c] before adsorption [d] after adsorption,
NiFe2O4 hydrogels: [e] before adsorption [d] after adsorption.

16.49 % nickel, and 55.18 % oxygen. (Fig. 7b) The synthesis of NF and thermogram. Fig. 9a and 9b depict the TGA and DTA thermogram of
CoFe2O4 nanoparticles is thus validated using EDX analysis. CoFe2O4 and NiFe2O4 nanoparticles and hydrogels. In the case of
The hydrogels of NiFe2O4 and CoFe2O4 before adsorption are also CoFe2O4 nanoparticles, the overall material loss is 11.33 %, and no in­
subjected to EDX analysis (Fig. 8b, Fig. 8c). The images of the CoFe2O4 termediate decomposition are noticed. Dehydration is the only signifi­
hydrogels showed elements including cobalt, iron, carbon, nitrogen, and cant weight loss, corresponding to 11.33 %. The results of [41] support
oxygen. Alternatively, after adsorption (Fig. 8d)., the CoFe2O4 hydro­ this. Also, owing to the well-dispersible nature of NiFe2O4 nanoparticles
gel’s EDX indicated some additional peaks of chromium. Similarly, in compared to CoFe2O4 nanoparticles, the amount of water molecules and
the EDX images of NiFe2O4 hydrogels, the presence of nickel, iron, ox­ hydroxyl groups on the surface of these nanoparticles will be higher than
ygen, nitrogen, and carbon is noticed [Fig. 8e] As expected, the EDX the former, due to which the total weight % loss in NiFe2O4 nano­
images of NiFe2O4 hydrogels after adsorption indicated the presence of particles is 21.11 %. Unlike CoFe2O4 nanoparticles, there are three
chromium [Fig. 8f]. This explains the structural changes that occurred in distinct stages of decomposition in NiFe2O4 nanoparticles here:
the hydrogels after adsorption. 30–200 ◦ C is the region of dehydration, 200––300 ◦ C refers to the
initiation phase of crystallization, and 300–600 ◦ C is the region of
complete crystallization [42]. These findings show that the crystalliza­
Thermal stability Analysis tion rate is higher in CoFe2O4 nanoparticles than NiFe2O4 nanoparticles,
giving rise to different response patterns over the same temperature
The thermal stability, moisture and the percentage of volatile matter range.
in the ferrite nanoparticles and hydrogels are analysed using TGA In addition, the peak at 260 ◦ C and 280 ◦ C in the DTA curve of
analysis. The CoFe2O4, NiFe2O4 nanoparticles, and hydrogels were CoFe2O4 and NiFe2O4 nanoparticles, respectively [Fig. 9b], indicates the
heated from room temperature to a temperature of 600 ◦ C, and the exothermic process involved in the ferrite crystal formation. The weight
weight loss was recorded as a function of temperature in the form of a

150
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

CoFe2O4 and NiFe2O4 hydrogels thus indicate that a significant fraction


of hydrogel constitutes gelatin bio-polymer. Three corresponding py­
rolysis zones exist: i. moisture vaporization [ 30–300 ◦ C], ii. active zone
[300––400 ◦ C], iii. Passive zone [400––600 ◦ C]. The moisture vapor­
ization is related to the evaporation of H2O molecules trapped in the
crevices of the hydrogel polymer matrix. The active zone is the region
where the decomposition process occurs at a rapid phase, which is
evident from the slope of the curve [Fig. 9a]. This is also a region where
the constituent cellulose fibres of the polysaccharide framework of
gelatin biopolymer disintegrate. Alternatively, the passive zone is the
phase where molecules are slowly disintegrated. As prominent and
sharp peaks are seen in the DTA curves of CoFe2O4 and NiFe2O4
hydrogels at 500 ◦ C and 320 ◦ C respectively, the exothermic nature of
denaturation of protein molecules is confirmed [43]. Also, the minimal
difference observed in weight loss percentage between NiFe2O4 and
CoFe2O4 hydrogel is co-related with the amount of the substance
encapsulated in each hydrogel. As demonstrated in the TGA results,
increasing the encapsulation of the nanofillers in the polymer hydrogels
increases the thermal stability of hydrogels [44].

Bet-surface area analysis and swelling studies

The BET gives the multi-point measurement of nanocomposites’


specific surface area and pore size distribution. BET analysis is more
reliable for nanomaterials with Type II or Type IV isotherms than ma­
terials with other isotherm behaviour. It is widely known that the pore
structure and surface area predominantly affect the selectivity of a
material. Mesoporous hydrogels have versatile delivery and attractive
features in terms of ease of nanofiller encapsulation. Materials with a
Fig. 9. [a] Thermogram of CoFe2O4, NiFe2O4 nanoparticles and hydrogels 9
pore diameter of less than 2 nm are microporous; when a pore diameter
[b] Differential thermogram of CoFe2O4 and NiFe2O4 nanoparticles
and hydrogels.
is between 2 and 50 nm, they are categorised as mesoporous materials.
Macroporous materials, on the other hand, are those whose pore
diameter is greater than 50 nm. In this analysis, a material whose pore
loss behaviour with increased temperature also provides significant
volume is to be measured is heated and degassed to remove any gas
characteristic observation of the hydrogels. When the hydrogels’ and the
molecules if adsorbed already. After this procedure, the adsorbent is
nanoparticles’ TGA results are compared, the overall weight loss per­
exposed to liquid nitrogen or any other suitable inert gas of defined
centage is higher in hydrogels than in nanoparticles. This is due to
volume. The volume of inert gas adsorbed by the material is continu­
organic functional groups of gelatin biopolymer that comprise amino
ously traced to find the pore volume of the adsorbent material. The
acids of protein entities bearing peptide bonds. These peptide bonds on
Fig. 10a and 10b are the characteristics of type IV isotherm that depict
denaturation contribute extensively to weight loss. The TGA curves of
mesoporous materials’ adsorption behaviour. The summary of the

Fig. 10. [a] BET adsorption and desorption isotherm of CoFe2O4, NiFe2O4 nanoparticles [b] BJH desorption summary of CoFe2O4 and NiFe2O4 nanoparticles.

151
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Table 1 Analysis of point of zero charge and variations in pH on the uptake


BJH desorption summary parameters of Cobalt Ferrite and nickel Ferrite NPs. capacity
COFe2O4 NPs NiFe2O4 NPs
2 Zero-point charge (pzc) is the pH value at which the total net charge
Surface area 141.792 m /g 132.059 m2/g
Pore Volume 0.214 cc/g 0.120 cc/g on the surface of the adsorbent is zero. The determination of Pzc gives the
Pore diameter 3.823 nm 3.403 nm electroneutrality of a material. When an adsorbent is in contact with an
aqueous medium, H + and OH– ions will be distributed over the
adsorbent surfaces. If the concentration of H+ ions exceeds OH– ions, the
textural characteristics of the adsorbent are also listed in Table 1. The adsorbent is positively charged and alternatively, when OH– ions are
pore diameters of CoFe2O4 and NiFe2O4 nanoparticles are 3.8 nm and more significant than H + ions, the adsorbent is negatively charged.
3.4 nm, respectively. These values obtained for CoFe2O4 and NiFe2o4
nanoparticles align well with those reported in the literature [45].The
marginal difference in the porosity of CoFe2O4 and NiFe2O4 nano­
particles is due to their varying degrees of crystallinity. The higher
surface area obtained in CoFe2O4 and NiFe2O4 nanoparticles is due to
the smaller size and large surface-to-volume ratio.
The swelling capacity of bare gelatin and NiFe2O4, CoFe2O4 hydro­
gels is investigated by varying the cross-linking agent concentration
from 0.5 to 1.5 %. Initially, the known quantity of hydrogels is weighed
and stirred with 50 mL deionized water for about 24 h. Subsequently,
after stirring, the weight of the beads was measured and the swelling
percentage was calculated. Fig. 12 shows the graph of crosslinker vs
swelling %. It is found that the ferrite encapsulated in the composite
hydrogels hinders the diffusion of water molecules into the hydrogels.
For this reason, bare gelatin beads exhibited more swelling than
NiFe2O4, CoFe2O4 hydrogels. Another finding is that the swelling per­
centage decreases as the crosslinking rate increases due to the tighter
hydrogel structure. To put it another way, the penetration of water
molecules into the surface of the hydrogel becomes difficult with
increased cross-linking. [Fig. 11]. Fig. 13. Zero-point charge of CoFe2O4 and NiFe2O4 hydrogels.

Fig. 11. Effect of Cross-linking of CoFe2O4 and NiFe2O4 hydrogels upon swelling.

Fig. 12. Swelling % of a. Bare gelatin hydrogels b. CoFe2O4 embedded hydrogels c. NiFe2O4 embedded hydrogels.

152
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 14. Batch Optimization parameters: [a] studies on pH using nanoparticles as adsorbent [b] studies on pH using hydrogels as adsorbent [c] studies on con­
centration using nanoparticles and [d] studies on concentration using hydrogels as the adsorbent [e] studies on contact time using nanoparticles as adsorbent and [f]
studies on contact time using hydrogels as adsorbent [g] studies on temperature using nanoparticles as adsorbent and [h] studies on temperature using hydrogels
as adsorbent.

Apparently, an equilibrium is established at a particular pH such that the be an affinity between chromate anion and H+ ions leading to maximum
concentration of H+ ions equals OH– ions. This theory assumes that the efficiency at pH = 2. At pH = 1, chromium exists in the form of chromic
interaction between constituent ions in water and the adsorbent surface acid; thus, the efficiency is lower at pH = 1 compared to pH = 2 even
is electrostatic. Below and above zero-point charge, the adsorbent is said though pH = 1 is strongly acidic.
to be positively and negatively charged, respectively. In this case, the
zero-point charge of cobalt ferrite gelatin composite is 5.2, whereas it is
6.2 for nickel ferrite gelatin composite (Fig. 13). Understanding the Modelling and interpretation of adsorption isotherms
point of zero charges is crucial because, based on the adsorbent’s zero-
point charge, the adsorbate medium’s pH has to be modified. The study systematically investigates the relative performance of
The response of the effect of pH in the uptake of chromium ions is nanoparticles and nanostructured transition metal ferrites doped
studied by varying the pH of the solution from 1 to 10 [Fig. 14a ]. In both hydrogels. The experimental conditions were optimized and set at pH 2
cases [Fig. 14a ], chromium removal is pronounced more in an acidic for a contact time of two hours. The adsorbent dose is set as 0.2 g for
medium with the highest adsorption efficiency at pH 2. Alternatively, CoFe2O4 and NiFe2O4 nanoparticles and alternatively, the dose for
the sorption reduces gradually when the solution approaches alkalinity. composite hydrogels is optimized at 0.4 g for 50 ppm metal ions con­
Thus, it is found that the adsorption efficiency decreases with an in­ centration. Throughout the studies, isotherm, kinetic modelling, and
crease in solution pH. This is because when the solution’s pH is reduced, thermodynamic investigations were conducted in these optimized pro­
more H + ions are present, and there will be an electrostatic attraction cess conditions.
between the negative ions of the solution and the positive ions on the The adsorption efficiency of the CoFe2O4 and NiFe2O4 nanoparticles
adsorbent surfaces [46]. It is also important to note that CoFe2O4 for different initial concentrations (10 to 250 ppm) of chromium solu­
nanoparticles outperformed NiFe2O4 nanoparticles in terms of the tion is shown in Fig. 14c. Chromium solutions of different concentra­
highest uptake. tions of 10, 50, 100, 150, 200 and 250 mg/L are tested. It is observed
As seen above, the pH vs adsorption efficiency in the case of that the efficiency continuously dropped with an increase in chromium
hydrogels also exhibited the highest adsorption efficiency at pH 2, fol­ solution concentration for both ferrite nanoparticles and hydrogels. This
lowed by a decrease in efficiency at increased pH [Fig. 14b]. However, can be associated with the number of available H + ions interacting with
contrary to nanoparticle sorption behaviour, NiFe2O4 hydrogel showed chromate anion, which will diminish continuously after a particular
better adsorption than CoFe2O4 hydrogels. This is due to the zero-point point [48]. When CoFe2O4 and NiFe2O4 nanoparticle’s uptake efficiency
charge of the CoFe2O4 and NiFe2O4 hydrogels, which is 5.2 and 6.2, are compared, CoFe2O4 has the highest uptake efficiency. This is due to
respectively. Therefore, at pH values below 5.2 and 6.2, the hydrogels the high porosity of CoFe2O4 nanoparticles, as seen in BET studies. The
are positively charged; in other words, there will be an accumulation of above findings demonstrate that the initial metal ion concentration af­
H+ ions on the surface of the sorbent. Also, the existence of chromium fects the adsorbent efficiency [49]. In general, more chromium ions are
(VI) as chromate anion CrO2- 4 at pH = 2 is well-known [47], so there will
present in the solution as the concentration rises, and as a result, more
ions tend to adhere to the adsorbent surface. But the number of

153
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 15. Adsorption Isotherm models and Kinetic models of CoFe2O4 and NiFe2O4 hydrogels: [a] Langmuir [b] Freundlich [c] Pseudo-first order [d] Pseudo-
second order.

adsorbent sites is fixed for a certain amount of adsorbent. Beyond this adsorbed on the surface of the adsorbent. The homogeneity of surface
limit, the adsorbent cannot adsorb a more significant number of ions. sites is an intriguing feature of this adsorption isotherm model, and the
Some contradicting findings are also seen in the investigations. Although main significance of this model is to predict whether the adsorption
the above results indicate that CoFe2O4 nanoparticles performed better, process will take place or not. Based on the separation factor ‘RL’ value,
CoFe2O4 hydrogels efficiency is slightly lower than NiFe2O4 hydrogels. it is also helpful in understanding the mechanism behind the proposed
This can be co-related with the greater hydrophilic nature of cobalt adsorption phenomenon.
ferrite compared to nickel ferrite nanoparticles.
1 1 1
The experimental and theoretical validation of system variables are = + (2)
qe qm qmKLCx
evaluated using isotherm models (Fig. 15). Langmuir adsorption
isotherm is based on the theoretical assumption that the interaction
Equation (2) is the mathematical way of representing the Langmuir
between adsorbate and adsorbent is more likely a physisorption mech­
adsorption isotherm model.
anism. Moreover, it supports the idea of monolayer adsorption, i.e., at a
‘RL’ expresses the equilibrium parameter. ‘RL’ can take any values
particular given time, only a single layer of adsorbate molecules can get

Table 2
Parameters of different Adsorption Isotherm models.
Adsorption Isotherm model Parameters CoFe2O4 nanoparticles NiFe2O4 CoFe2O4 NiFe2O4
nanoparticles hydrogels hydrogels

R2 0.9975 0.9854 0.9899 0.9962


Langmuir qm 17.543 mg/g 4.5126 mg/g 7.0077 mg/g 9.2421 mg/g
KL 2.666 L/mg 0.0663 L/mg 9.871 L/mg 0.21723 L/mg
RL 0.0004531 [Ci = 50 ppm] 0.231610 [Ci = 50 ppm] 0.0016768 [Ci = 50 ppm] 0.08430 [Ci = 50 ppm]
R2 0.9700 0.9881 0.9877 0.9810
Freundlich Kf 0.061122 L/g 0.17644 L/g 1.10383 L/g 1.45982 L/g
n 0.4442 0.4849 1.6650 2.0462
R2 0.9311 0.8931 0.9784 0.9553
Temkin bT 6.38302 kJ/mol 5.8614 kJ/mol 7.76874 kJ/mol 6.4367 kJ/mol
BT 1.3484 L/mg 1.2673 L/mg 1.3976 L/mg 1.3091 L/mg
qm 17.3256 mg/g 3.9751 mg/g 7.2367 mg/g 9.1134 mg/g
Dubinin Radushkevich model R2 0.8891 0.8503 0.8202 0.8793

154
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Table 3
Parameters of different Kinetic models.
Kinetic model Parameters CoFe2O4 nanoparticles NiFe2O4 nanoparticles CoFe2O4 hydrogels NiFe2O4 hydrogels
− 1 − 1 − 1
Rate Constant 0.0082908 min 0.0087514 min 0.02418 min 0.005296x10-3 min− 1

Pseudo first order


qe (calculated) 0.1832 mg/g 0.9124 mg/g 3.8636 mg/g 0.05108 mg/g

qe (experimental) 9.1184 mg/g 3.743689 mg/g 4.083495 mg/g 4.379612 mg/g

Pseudo second order Rate constant 6.723x10-2 3.2334x10-2 g.mg-1min− 1


6.378x10-2 min-1gmg− 1
6.531x10-2
min-1gmg− 1 g.mg− 1.min− 1

qe (calculated) 8.012 mg/g 5.46448 mg/g 4.8756 mg/g 5.3763 mg/g

qe (experimental) 9.1184 mg/g 3.743689 mg/g 4.083495 mg/g 4.379612 mg/g

Intraparticle diffusion model R2 0.9798 0.9905 0.9797 0.98


1/2
Kdiff 1.1083 0.9413 mg.g.min− 0.9372 0.9241
mg.g.min-1/2 mg.g.min− 1/2
mg.g.min− 1/2

R2 0.6279 0.9376 0.9894 0.926


Elovich model a 0.6504 0.6341 2.8595 0.3217
b 0.6321 0.4123 2.1134 0.3451

between 0 and 1 for the adsorption process to be favourable, and a value


greater than 1 means that the process is unlikely to occur finally, if ‘RL’ ln(qe) = lnqm − βE2 (5)
=1, then it represents linear adsorption. Table 2 shows that the 2
The high variation in R values obtained in the Temkin and Dubinin
adsorption process favours nanoparticles and hydrogels based on the Radushkevich model shows that the adsorption process does not fit the
‘RL’ separation factor. Compared to nickel ferrite nanoparticles, the model.
maximum adsorption capacity is higher for cobalt ferrite nanoparticles
with respect to physisorption. This can be attributed to the higher sur­ Modelling of adsorption kinetic processes
face area, pore volume and pore diameter [ Table 1] of cobalt ferrite
nanoparticles. The information obtained from studying the adsorption kinetics at
Freundlich adsorption isotherm is an empirical relation that pre­ different contact times is essential while designing adsorption columns.
sumes that the interaction between the adsorbent and adsorbate mole­ This is because the adsorption kinetics of hydrogels strongly influence
cules is of chemisorption type. It relies on heterogeneous sites for the release of solute from chromium solution onto hydrogel surfaces.
adsorption. It also assumes that both cations and anions simultaneously However, besides contact time, solute concentration, pore diffusivity,
absorb on the adsorbent surfaces. and surface complexity of adsorbents also affect the kinetic behaviour.
Equation (3) is the mathematical formula of the Freundlich adsorp­ In the present study, the duration of the adsorption process varied from
tion isotherm model. 10 to 140 min. Fig. 14e shows that the adsorption efficiency increased
1 with contact time for both CoFe2O4 and NiFe2O4 nanoparticles. Even
logqe = logKf + logCx (3) though both depicted the same trend, CoFe2O4 nanoparticles are more
n
efficient than NiFe2O4 nanoparticles. Nevertheless, maximum adsorp­
The corresponding parameters derived from the Freundlich adsorp­ tion efficiency is observed at 120 min for both CoFe2O4 and NiFe2O4
tion isotherm model are given in Table 2. ‘n’ denotes the strength of the nanoparticles. With increased agitation time, more ions came into
adsorption process and has a numerical value greater than 1. Any value contact with the adsorbent surface, and as a result, the interaction be­
between 0.1 and 1 tells us that the adsorption process is favourable. If tween H+ ions and the adsorbate medium increased. From 10 to 120
the n value is less than 0.1, then it indicates that the adsorption process min, adsorption efficiency increased, and after 120 min, the efficiency
is less favourable. If we examine the R2 value of each adsorption became constant, which is seen as the adsorbent’s saturation region
isotherm in Table 2, the nanoparticles and the hydrogels showed a better [50]. It is also inferred that the void spaces on adsorbent surfaces to
fit for Freundlich models supporting heterogeneous adsorption. The uptake chromium ions are occupied completely [51]. Similar sorption
greater removal efficiency of NiFe2O4 hydrogel compared to CoFe2O4 kinetic studies were also evaluated with the hydrogels. Every 15 to 20
hydrogel is consistent with the higher Kf value of NiFe2O4 hydrogel min, the adsorption efficiency is investigated for 120 min [Fig. 14f].
compared to CoFe2O4 hydrogel. In contrast, CoFe2O4 nanoparticles Here, it is also evident that adsorption efficiency increased with contact
exhibit a higher ‘KL’ value than NiFe2O4 nanoparticles because of their time. However, as noted in the previous section, it is evident that
higher surface area and porosity, which enable them to interact physi­ NiFe2O4 hydrogels showed higher efficiency than CoFe2O4 hydrogels.
cally with chromium anions to a greater extent. The Temkin model The reaction order is a preliminary parameter to be determined
states that heat given out by molecules after adsorption is inversely before optimizing the process conditions such as temperature, pressure
proportional to surface coverage. The dependence on both physisorption and catalyst [52,53]. We can quantify the amount of product obtained
and chemisorption tells that the model has intermediate characteristics based on the reaction rate. Rate and concentration are well related, and
of both Freundlich and Langmuir isotherms, and the results obtained by adjusting any one parameter, the other one will show some changes.
[Table 2] correlate. Adsorption kinetics data are thus modelled by assuming that the system
qe = Blog(Kt) + BlogCe (4) could follow pseudo-first-order or pseudo-second-order rate laws
[5457]). The corresponding equation of each model is given below
In addition, the Dubinin Radushkevich model was also studied to respectively:
interpret the nature of the adsorption process. The ‘qm’ values obtained K1
log(qe − qt) = logqe -2.303 t (6)
in this case match well with that of Langmuir isotherm values supporting
the physisorption nature of the process. t 1 1
= t+ (7)
qt qe K2 qe2

155
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

On interpreting Table 3, it is found that the pseudo-first-order kinetic Table 4


model did not yield satisfactory results for most of the studied adsor­ Energy parameters of the adsorption process.
bents. This is evident from the regression coefficient values seen in the Nanoparticles Hydrogels
tables. The pseudo-first-order kinetic plots showed R2 values signifi­
Cobalt Ferrite Nickel Ferrite CoFe2O4 NF Hydrogels
cantly lower than their pseudo-second-order models’ R2 values. How­ NPs NPs Hydrogels
ever, CoFe2O4 hydrogels are an exception. An R2 value of 0.98 and a
ΔHo 520.80 kJ/mol
good match of the experimental and theoretical adsorption capacity of 222 kJ/mol 709.53 kJ/mol 252.32 kJ/mol
the CoFe2O4 hydrogels explains the validity of the pseudo-first-order ΔGo
model. − 5568 kJ/mol − 4871.39 kJ/ − 4378.47 kJ/ − 5876.78 kJ/
The same kind of observation is noticed while comparing the validity mol mol mol
o
26.024 kJ/mol
of the pseudo-second-order model for the studied adsorbents. CoFe2O4 ΔS
12.29 kJ/mol 32.010 kJ/mol 24.415 kJ/mol
hydrogels showed a better fit (R2 = 0.98) for the pseudo-second-order
model than all the studied adsorbents. A close match of experimental
and theoretical adsorption capacity is also noticed here. With relevance adsorption is also investigated at various temperatures, as shown in
to R2 values, for other adsorbents, the coefficient of the determination Fig. 14h. Both CoFe2O4 and NiFe2O4 hydrogels exhibited endothermic
value falls between 0.94 and 0.97. Moreover, experimental and theo­ adsorption with higher efficiency for NiFe2O4 hydrogels.
retical adsorption capacity values do not precisely match the other The study also comprehends thermodynamic investigations using
studied adsorbents in the pseudo-second-order model. However, their Vant Hoff’s plot, which gives an insight into reaction thermodynamics.
differences are not so huge. Considering the above results, it is inferred Thermodynamics of reaction plays a crucial role in determining the
that the studied adsorption system demonstrated better validity with the reaction’s spontaneity [57]. The feasibility of any given reaction in
pseudo-second-order than the pseudo-first-order model. This indicates chemistry is evaluated by calculating energy parameters like change in
that the studied adsorbents are more inclined towards chemisorption enthalpy(ΔH), change in entropy (ΔS) and Gibbs free energy parameter
than physisorption. (ΔG) using the following equations given below.
From the pseudo-second-order model results, it is found that the rate
ΔGo = ΔH o − TΔSo (10)
constant is higher for CoFe2O4 nanoparticles compared to NiFe2O4
nanoparticles. On the contrary, hydrogels entrapping cobalt ferrite Also,
witnessed a lower rate constant than NiFe2O4 hydrogels. When a reac­ o
ΔG = − RTlnKc (11)
tion proceeds slowly, the rate constant will be low, indicating a less
favourable interaction between the adsorbent and adsorbate. As a result, Where Kc represents the adsorption distribution coefficient given by the
it suggested that the NiFe2O4-based nanoparticles and CoFe2O4 based formula:
hydrogels adsorbate -adsorbent interaction is found to be poor
Ca
compared to CoFe2O4-based nanoparticles and NiFe2O4 based hydrogels Kc = (12)
Cx
respectively. Moreover, CoFe2O4 hydrogel’s experimental and theoret­
ical adsorption capacities also correlate well. These results indicate that Ca = concentration of chromium ions on the surface of the
the studied system is more inclined toward a pseudo-second-order adsorbent.
model supporting chemisorption. The intraparticle diffusion model Cx = concentration of chromium ions remaining in solution after
[IPD] was also studied to understand the influence of neighbouring adsorption.
particles in the adsorbent bound to each other upon adsorption. The
ΔS0 ΔH O
equation for the IPD model is as follows: ln Kc = − (13)
R RT
qt = Kdif t1/2 + c (8) Equation (10) is called the Van’t Hoff plot.
‘Kdif’ is the rate constant, and ‘c’ is the intercept. This process mainly Table 4 gives the thermodynamic analysis of all the studied adsor­
describes the initial adsorption behaviour of the process. Elovich’s bents. The nature of the adsorption process is inferred by examining the
model describes the direct relationship between chemisorption activa­ enthalpy values. The table shows the positive enthalpy (ΔH) change for
tion and the adsorption process’s energy. The energy of the chemi­ all the studied adsorbents supporting endothermic adsorption. The re­
sorption process decreases with the coverage of active sites on the sults of Gibbs free energy, on the other hand, give information on
surface of the adsorbent. The governing equation of the model is given studying the spontaneity of the adsorption process. The greater the
below. negative Gibbs free energy value, the more the reaction is predicted to
be very spontaneous. This is possible only if the affinity between the
1 1
qt ln(ab) + ln(t) (9) adsorbent and adsorbate is high. From table 4, it is clear that all the
a a adsorbents present a negative Gibbs free energy, which shows that the
The adsorption process does not obey the Elovich model, which is adsorption is spontaneous in all the cases. Among all the studied ad­
evident from the R2 values obtained. sorbents, NF hydrogels show more spontaneity. The table also indicates
that all the studied adsorbents give a positive entropy value. The greater
Temperature studies and interpretation of adsorption thermodynamics degree of freedom available for the adsorbent molecules to move freely
is evident from the positive magnitude of change in entropy (ΔS).
Temperature is a critical parameter in adsorption reactions, which
decides whether the adsorption is physical or chemical. The influence of Chromium adsorption in the fixed bed column using ferrites entrapped
temperature on adsorption is studied by varying the adsorption tem­ gelatin hydrogels
perature, say 30 ◦ C, 40 ◦ C, 50 ◦ C, and 60 ◦ C respectively. The corre­
sponding response is depicted in Fig. 14g. The adsorption efficiency of The batch results are scaled up by conducting fixed bed column
both CoFe2O4 and NiFe2O4 nanoparticles increased with temperature., studies to understand the applicability of the synthesized hydrogels in
The greatest adsorption efficiency was observed at 60 ◦ C, implying that large-scale applications. A column made of acrylic material with a
the process is endothermic [56]. Compared to NiFe2O4 nanoparticles, length of 30 cm and an inner diameter of 2.5 cm was used in the
CoFe2O4 nanoparticles depicted higher efficiency due to the higher experiment. The column was filled with composite hydrogels of known
surface area of these particles. The effectiveness of the hydrogels’ weight, and the adsorbate was allowed to flow through from the top of

156
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 16. Fixed-bed column parameters: [a] Flow-rate variation and [b] Bed-depth variation:

the column. The column bottom was supported with glass wool to pre­
Table 5
vent the loss of adsorbent and adsorbate. After adsorption, a peristaltic
Packed-bed column parameters of the CoFe2O4 and NiFe2O4 hydrogels.
pump was used to adjust the flow rate and the effluent was collected in
the outlet.Two different parameters were varied: 1. Bed-depth and 2. Adsorbent Flow rate (mL/ qtotal qbed Volume of
min) and bed (mg) (mg/g) chromium solution
Flow rate of the inlet solution. The bed depth was varied between 1 cm
depth (cm) treated
and 2 cm. The flow rate, on the other hand, was set at 2 mL/min and 3
mL/min. The breakthrough experiments were performed using both 2 mL/min and 1 101.36 32.48 1.32 L
Cobalt ferrite cm mg mg/g
CoFe2O4 and NiFe2O4 hydrogels and the corresponding break-through hydrogels 3 mL/min and 1 92.24 29.36 1.62 L
curves were plotted. cm mg mg/g
Fixed bed column studies were conducted by varying flow rates of 2 3 mL/min and 2 119.42 19.0 2.16 L
mL/min and 3 mL/min for a fixed bed height of 1 cm. The breakthrough cm mg mg/g
2 mL/min and 1 100.62 31.81 1.44 L
curves are given in Fig. 16a. It was found that the saturation point
Nickel ferrite cm mg mg/g
approached earlier as the flow rate increased [58]. At a 2 mL/min flow hydrogels 3 mL/min and 1 91.65 27.39 1.88 L
rate, the saturation point reached 660 min for cobalt ferrite hydrogels, cm mg mg/g
whereas for 3 mL/min, the saturation point reached 540 min. The same 3 mL/min and 2 92.76 13.84 2.34 L
kind of pattern is observed for nickel ferrite hydrogels. At a 2 mL/min cm mg mg/g

flow rate and 3 mL/min, the saturation values are 720 min and 600 min,
respectively (Fig. 16a). This is because as the flow rate rises, more ions bed represented in mg. ‘Q’ is the system’s flow rate in mL/min. ‘Cad’ is
tend to be adsorbed on the surface of the adsorbent, causing the satu­ the total chromium ions adsorbed in the column in mg/L. ‘qbed’ is the
ration point to be attained more quickly[59]. bed adsorption capacity in mg/g. ‘W’ is the adsorbent dosage. Table 5
The column was filled with the composite hydrogels to different bed gives values of the parameters of column studies.
depths of 1 and 2 cm and allowed to run at a fixed flow rate of 3 mL/min. It is clear from the preceding table that the qtotal value of both cobalt
Fig. 16b depicts the breakthrough curves of the column using cobalt ferrite and nickel ferrite hydrogels declined as the flow rate increased.
ferrite hydrogels and nickel ferrite hydrogels with varied bed depths. This is due to a rise in adsorbed chromium ions with an increasing flow
From the breakthrough curves, we can understand that with an increase rate. Thereby, bed adsorption capacity also witnessed a fall. In contrast,
in the bed height, there is an increase in the breakthrough time [60]. At a cobalt ferrite and nickel ferrite hydrogels qtotal values increased sub­
bed height of 1 cm, the breakthrough is only at 540 min, and at a bed stantially with increased bed height. This is because more active sites are
height of 2 cm, the saturation point went up to 720 min. This is the case available for adsorption with increased adsorbent loading. The results
with cobalt ferrite hydrogels. For nickel ferrite hydrogels, at a bed height thus infer that at a flow rate of 2 mL/min, 2 cm bed depth was optimum
of 1 cm, the saturation point is observed at 600 min; at a 2 cm bed for 50 ppm chromium solution. For CoFe2O4 hydrogel and NiFe2O4
height, the breakthrough is at 780 min. This is because, with an increase hydrogel, respectively, qbed was found to be 32.48 mg/g and 31.81 mg/g
in bed height, the adsorbent dosage will increase, and more sites will be (Table 6). Despite the distinct performance patterns of all the studied
available for the adsorption of ions[61]. adsorbents witnessed in batch systems, since the qbed values do not
Different parameters of the fixed bed adsorption system are calcu­ significantly differ, it is proposed that both CoFe2O4 and NiFe2O4
lated using equations (10) and (11). hydrogels have potential applications in large-scale fixed bed columns.
t=t
∫ total The pzc of the hydrogels was measured to be 5.2 and 6.2 for cobalt
qtotal =
Q
Cad dt (14) ferrite- and nickel ferrite-hydrogels, indicating that the hydrogels are
1000
t=0
positively charged below the pH values 5.2 and 6.2. As a result, the
batch experimental studies were conducted at an optimum pH of 2; the
qbed =
qtotal
(15) hydrogels are positively charged at this pH value. The existence of
W chromium (VI) in the form of HCrO-4 at a pH value equal to 2 is well
‘qtotal’ is the total number of chromium ions adsorbed in the packed proven [62]. Therefore, the positively charged protonated hydrogels

157
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Table 6
Comparison between adsorption performance of CoFe2O4,NiFe2O4 hydrogel for
Cr(VI) removal with existing literature reports
Parameters
S. Adsorbent
pH Temperature Adsorption Reference
NO material
capacity (mg/g)

1 Cellulose doped 2 – 12–13 mg/g [33]


gelatin hydrogels
2 Leather waste – 60 C◦
17 mg/g
derived gelatin [46]
3 ZnO-polyacrylic 2 – 11.71 mg/g [31]
acid/PEG hydrogel
4 Cellulose-gelatin 2 70 ◦ C 12 mg/g [49]
composite
5 Starch- 2 35 ◦ C 23.47 mg/g [67]
polyacrylic/
cellulose
Hydrogels
6. Chitosan beads 2 – 9.5 mg/g [68]
modified with SDS
7. Alginate/ – – 29.89 mg/g [69]
Polyaniline
composite
8. Avocado seeds 2 25 ◦ C 133.33 mg/g [9]
(NASs)
9. Bimetallic MOFs 3 – 24.81 mg/g [22]
(Fe0.75 Cu0.25 Fig. 17. Mechanism of Adsorption process of Hexavalent Chromium ion on the
BDC)- bound
surface of CoFe2O4 and NiFe2O4 hydrogels.
Alginate -MoO3/
Graphene Oxide
10. NanoBch-EDA 2 – 79.37 mg/g [6] This is supported by XPS spectrum [Fig. 18]. The peaks observed at
11. MOF 1 792.1–1157.32 [20] 285.22 eV and 284.68 eV are aligned with sp2 hybridized aliphatic
(nFe3O4@MIL- mg/g
carbon atoms, and sp3 hybridized carbon atoms belonging to gelatin
88A(Fe)/APTMS)
12. N-doped Graphene 1 30 ◦ C 350.42 mg/g [26] (Fig. 18a). In the XPS spectrum of oxygen (Fig. 18b), the peak at 533.96
Oxide hydrogel eV is assigned to surface hydroxyl groups on the adsorbent. This is
with shrimp shell responsible for ion exchange with chromate anions to form an inner
13. CoFe2O4 hydrogels 2 30 ◦ C 32.48 mg/g Present
sphere complex. At 532.27 eV, the presence of a metal–oxygen bond
study
14. NiFe2O4 hydrogels 2 30 ◦ C 31.81 mg/g Present
confirms the impregnation of ferrite nanoparticles into the adsorbent.
study The characteristic peaks at 400.26 eV, 400.37 eV and 401.71 eV are
related to the amide groups, primary amine and secondary amine groups
of gelatin biopolymer, respectively (Fig. 18c). The peak observed at
interact with these negatively charged chromate anions through elec­ 736.02 eV indicates the presence of Iron (Fig. 18d) in the elemental state
trostatic attractions1*, and HCrO-4 anions will adsorb on the hydrogel’s of + 3 valency. This also shows that the oxidation change of Iron from +
surfaces. Ferrites are a class of mixed valent metal oxides constituting 2 to + 3 has occurred. Further, the presence of characteristic peaks at
mainly Iron center. The distribution of hydroxyl groups on these ferrite 578.68 eV and 591.16 eV is assigned to hexavalent chromium, and the
nanoparticles’ surface when dispersed into an aqueous medium is also a peaks observed at 575.56 and 583.71 eV correspond to the trivalent
proven concept [63]. These hydroxyl groups exhibit amphoteric state of chromium. This confirms the reduction of Cr (VI) to Cr (III), as
behaviour, i.e., they have both basic and acidic natures. When they mentioned in the mechanism (Fig. 17e).
exhibit basic behaviour, they tend to show some affinity towards other *- Numbering given in Fig. 17.
anions with an acidic nature, such as HCrO4. Thus, it can be inferred
that the adsorbed chromate anions on the composite surfaces will get
exchanged with these hydroxyl groups to form inner sphere complexes2* Comparison between ferrite nanoparticles and gelatin hydrogels adsorption
as shown in the equation:
In the case of nanoparticles, cobalt ferrite exhibited a higher per­
S-OH- + HCrO-4 + H+ S-HCrO-4 + H 2O
formance efficiency of 90.03 % at an optimum pH of 2 with a maximum
[Inner sphere complex] adsorption capacity of 17.543 mg/g compared to nickel ferrite nano­
The formation of these inner-sphere complexes leads to the adsorp­ particles whose efficiency is 76 % and adsorption capacity of 4.5126
tion of hydrogen chromate anions on the surface of ferrite nanoparticles mg/g. This is attributed to the higher surface area, pore volume and pore
loaded into the composite, creating a pathway for electron transfer re­ diameter of cobalt ferrite nanoparticles obtained from BJH desorption
action between the Iron-oxide center to chromium (VI) centre [64]. It is summary table of BET Isotherms. The distribution coefficient of divalent
also known that the synthesis of cobalt ferrite nanoparticles through the and trivalent cations in the inverse spinel structure is also essential in
co-precipitation method will usually result in a mixture of Iron oxides exerting electrostatic attractions. The distribution coefficient (x) can be
and hydroxides, ranging from magnetite [Fe3O4], Fe (OH)2, Fe-O-OH determined by Mossbauer spectroscopy [66]. Based on this inversion
and γ-Fe2O3. In addition, the rapid conversion of Fe2+ to Fe3+ in parameter, the concentration of Fe3+, Co2+, and Ni2+ distributed in
spinel ferrites could occur by pronounced oxidation and an electron octahedral and tetrahedral holes will vary. According to Fajan’s rule, the
hopping mechanism [65]. The electrons lost while oxidation of Fe2+ to smaller the size of a cation, the more excellent its polarizing power.
Fe3+ are gained by chromium (VI) to form less toxic trivalent chromium Therefore, Fe3+ ions smaller than corresponding divalent metal ions will
ions. The electron transfer from the Iron center to chromium (VI) thus exert more polarising power upon the electron cloud of chromate an­
favours the reduction of Chromium (VI) to less toxic Chromium (III) 3*. ions. Due to the larger size of octahedral holes compared to tetrahedral
holes, the greater concentration of Fe3+ ions in these sites means more

158
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 18. XPS spectrum of hydrogel [a] Carbon [b] Oxygen [c] Nitrogen [d] Iron [e] Chromium.

chromate anions will attract the adsorbent. This variation in the distri­
bution coefficient (x) of cobalt ferrite and nickel ferrite can contribute to
the superior performance of one over the other.
When moving to hydrogels, nickel ferrite hydrogels exhibited a
higher efficiency of 86.17 % and adsorption capacity of 9.2421 mg/g
compared to the CoFe2O4 hydrogels, whose efficiency and adsorption
capacity were 80.14 % and 7.0077 mg/g, respectively. This can be co-
related with the greater hydrophilic nature of cobalt ferrite compared
to nickel ferrite nanoparticles. Because of this nature, not all the cobalt
ferrite particles could thoroughly blend with gelatin biopolymer, which
is hydrophobic. Even though the exact amounts of nanoparticles are
entrapped in the hydrogel, encapsulation was much more pronounced in
nickel ferrite than cobalt ferrite, giving rise to higher performance ef­
ficiency of NiFe2O4 hydrogels. However, it is also important to note that
the difference in the adsorption capacity does not vary to a more sig­
nificant extent.

Desorption and regeneration studies of ferrites nanoparticles and gelatin


hydrogels

The chromium-loaded hydrogels, after adsorption, were washed


using deionized water and dried. After drying, the hydrogels were Fig. 19. Desorption % of CoFe2O4 and NiFe2O4 hydrogels.
transferred to a conical flask containing 40 mL of 0.1 N NaOH solution.
The hydrogels were subjected to agitation in a rotary shaker. At regular
periodic intervals such as 30 min, 60 min, 120 min and 240 min,
respectively, the chromium concentration of the desorbed solution was

159
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

Fig. 20. Reusability of the adsorbent: [a] CoFe2O4 hydrogels and [b] NiFe2O4 hydrogels.

examined. It was observed that the chromium concentration in the adsorption efficiency is noticed for cobalt ferrite and nickel ferrite
desorbed solution increased tremendously with an increase in time hydrogels, respectively. This suggests the reusability of hydrogels in the
[Fig. 19]. The desorption-readsorption process was carried out using the adsorption process. The characterization results indicated that the
same initially employed composite hydrogels to evaluate the adsor­ remediation mechanism of Cr (VI) ions is electrostatic attraction fol­
bent’s reusability. This process was repeated five times to test the lowed by inner-sphere complex formation and reduction. Continuous
hydrogel’s reusability. Fig. 20 shows a drastic difference in the sorption studies were also conducted to understand the reliability of the
adsorption efficiency after each cycle. It is also important to note that adsorbents for large-scale applications. For CoFe2O4 hydrogel and
not much difference is seen in the magnitude of the desorption efficiency NiFe2O4 hydrogel, respectively, qbed was found to be 32.48 mg/g and
of CoFe2O4 and NiFe2O4 hydrogels. In the first cycle, the CoFe2O4 31.81 mg/g. Despite the distinct performance patterns of all the studied
hydrogel’s adsorption efficiency is about 80 %; during the fifth cycle, the adsorbents witnessed in batch systems, since the qbed values do not
adsorption efficiency decreased to 55 %. Similarly, an adsorption effi­ significantly differ, it is proposed that both CoFe2O4 and NiFe2O4
ciency of 86.7 % for NiFe2O4 hydrogels was found during the first cycle, hydrogels have potential applications in large-scale fixed bed columns.
and it gradually reduced to 70 % during the fifth cycle. The reusability of
cobalt ferrite and nickel ferrite hydrogels, even after the fifth cycle, CRediT authorship contribution statement
confirmed the hydrogel’s reusability.
M. Priyadarshini: Conceptualization, Data curation, Writing –
Conclusions original draft. E. Rekha: Conceptualization, Data curation, Writing –
original draft. Asha Sathish: Conceptualization, Funding acquisition,
The nickel and cobalt ferrite nanoparticles and composite hydrogels Data curation, Writing – review & editing, Visualization, Investigation,
were synthesized in the present study to remove hexavalent chromium Validation, Formal analysis, Methodology, Supervision, Resources. K.
ions. The results indicated a good affinity between the entrapped Nithya: Conceptualization, Funding acquisition, Data curation, Writing
hydrogels and the chromium ions. The cross-sectional examination of – review & editing, Visualization, Investigation, Validation, Formal
hydrogels after adsorption, signalled structural changes in the hydro­ analysis, Methodology, Supervision, Resources, Project administration.
gels. The deposition of chromium ions on the dissected surfaces was
further confirmed by characteristic peaks observed in energy-dispersive
X-ray analysis. Further, the prominent peaks of metal ferrite nano­ Declaration of competing interest
particles are confirmed in the infrared spectrum, and the degree of
crystallinity is analysed using XRD characterization. The observed XRD The authors declare that they have no known competing financial
pattern of metal ferrite nanoparticles is completely aligned with JCPDS. interests or personal relationships that could have appeared to influence
XPS studies revealed the characteristic peaks at 400.26 eV, 400.37 eV the work reported in this paper.
and 401.71 eV are related to the amide groups, primary amine and
secondary amine groups of gelatin biopolymer, respectively. Batch References
adsorption studies were performed using both nanoparticles and
hydrogels, while hydrogels entrapped with CoFe2O4 and NiFe2O4 [1] L. Li, Q. Liao, B. Hou, C. He, J. Liu, B. Li, M. Yu, Y. Liu, B. Lai, B. Yang, Sep Purif
Technol 304 (2023) 122363, https://doi.org/10.1016/j.seppur.2022.122363.
nanoparticles were used in the continuous fixed bed column. When
[2] K. GracePavithra, V. Jaikumar, P.S. Kumar, P. SundarRajan, J Clean Prod 228
CoFe2O4 and NiFe2O4 hydrogels’ efficiency are compared at batch-scale, (2019) 580–593, https://doi.org/10.1016/j.jclepro.2019.04.117.
nickel ferrite hydrogels performed better than cobalt ferrite hydrogels. [3] Y. Zheng, T. Shengyang, Y. Jinxiang, L. Guangtao, J. Mater. Chem. 16 (2006) 2347.
This can be co-related with the greater hydrophilic nature of cobalt [4] D. Duranoğlu, I.G.B. Kaya, U. Beker, B.F. Senkal, Chem. Eng. J. 181–182 (2012)
103.
ferrite compared to nickel ferrite nanoparticles. Both adsorption and [5] A. Nagar, T. Pradeep, ACS Nano 14 (6) (2020) 6420–6435, https://doi.org/
desorption studies indicated that, after five cycles, 55 % and 70 % 10.1021/acsnano.9b01730.

160
M. Priyadarshini et al. Journal of Industrial and Engineering Chemistry 137 (2024) 144–161

[6] M.E. Mahmoud, A.M. El-Ghanam, S.R. Saad, Biomass Convers Biorefin (2022), [38] A. Mohd. Hashim, S. Kumar, S. Ali, B.H. Koo, H. Chung, R. Kumar, J Alloys Compd
https://doi.org/10.1007/s13399-022-02888-1. 511 (1) (2012) 107–114, https://doi.org/10.1016/j.jallcom.2011.08.096.
[7] P.L. Tang, C.K. Lee, K.S. Low, Z. Zainal, Environ. Technol. 24 (2003) 1243. [39] A.K. Thakur, B. Lin, F.H. Nowrin, M. Malmali, ACS ES&T Water 2 (1) (2022)
[8] A.V. K.K., D. Gangadharan, Nano-Structures & Nano-Objects 29 (2022) 100837, 75–87, https://doi.org/10.1021/acsestwater.1c00259.
https://doi.org/10.1016/j.nanoso.2022.100837. [40] W. Zhai, J. He, P. Han, M. Zeng, X. Gao, Q. He, Vacuum 195 (2022) 110634,
[9] M.E. Mahmoud, G.F. El-Said, G.A.A. Ibrahim, A.A.S. Elnashar, Biomass Convers https://doi.org/10.1016/j.vacuum.2021.110634.
Biorefin (2022), https://doi.org/10.1007/s13399-022-03619-2. [41] A. Hussain, A. Naeem, G. Bai, M. Yan, Journal of Materials Science: Materials in
[10] R. Khalili, M.M. Sabzehmeidani, M. Parvinnia, M. Ghaedi, Environ Nanotechnol Electronics 29 (24) (2018) 20783–20789, https://doi.org/10.1007/s10854-018-
Monit Manag 18 (2022) 100750, https://doi.org/10.1016/j.enmm.2022.100750. 0220-9.
[11] D. Mohan, C.U. Pittman, J. Hazard. Mater. 137 (2006) 762. [42] A.B. Mapossa, J. Dantas, M.R. Silva, R.H.G.A. Kiminami, A.C.F.M. Costa, M.
[12] A.A. Shyaa, O.A. Hasan, A.M. Abbas, J. Saudi Chem. Soc. (2012). O. Daramola, Arabian Journal of Chemistry 13 (2) (2020) 4462–4476, https://doi.
[13] Y.A.B. Neolaka, Y. Lawa, J. Naat, A.A.P. Riwu, A.W. Mango, H. Darmokoesoemo, B. org/10.1016/j.arabjc.2019.09.003.
A. Widyaningrum, M. Iqbal, H.S. Kusuma, Journal of Materials Research and [43] A. Neelam, S. Junaid Mahmood, S. Ishteyaq, (2018). 10.15436/2377-
Technology 18 (2022) 2896–2909, https://doi.org/10.1016/j.jmrt.2022.03.153. 0619.18.1847.
[14] A. El Nemr, R.M. Aboughaly, A. El Sikaily, S. Ragab, M.S. Masoud, M. Shafik [44] M. Khairy, M.E. Gouda, J Adv Res 6 (4) (2015) 555–562, https://doi.org/10.1016/
Ramadan, (n.d.). 10.1007/s13399-020-00995-5/Published. j.jare.2014.01.009.
[15] P. Luo, J.S. Zhang, B. Zhang, J.H. Wang, Y.F. Zhao, J.D. Liu, Ind. Eng. Chem. Res. [45] D. Uzunoğlu, M. Ergüt, P. Karacabey, A. Özer, Desalination Water Treat 172 (2019)
50 (2011) 10246. 96–105, https://doi.org/10.5004/dwt.2019.24942.
[16] M.R. Samani, S.M. Borghei, A. Olad, M.J. Chaichi, J. Hazard. Mater. 184 (2010) [46] B. Ghanim, J.J. Leahy, T.F. OʼDwyer, W. Kwapinski, J.T. Pembroke, J.G. Murnane,
248. Journal of Chemical Technology and Biotechnology 97(1) (2022) 55–66. 10.1002/
[17] M. Dakiky, M. Khamis, A. Manassra, M. Mereb, Adv. Environ. Res. 6 (2002) 533. jctb.6904.
[18] K.G. Bhattacharyya, S.S. Gupta, Ind. Eng. Chem. Res. 45 (2006) 7232. [47] N. Unceta, F. Séby, J. Malherbe, O.F.X. Donard, Anal Bioanal Chem (2010)
[19] C.V.T. Rigueto, M.T. Nazari, M. Rosseto, L.A. Massuda, I. Alessandretti, J.S. Piccin, 1097–1111, https://doi.org/10.1007/s00216-009-3417-1.
A. Dettmer, J Mol Liq 330 (2021) 115638, https://doi.org/10.1016/j. [48] S. Kutluay, O. Baytar, Ö. Şahin, Research on Engineering Structures and Materials 5
molliq.2021.115638. (3) (2019) 279–298, https://doi.org/10.17515/resm2019.73en1122.
[20] M.E. Mahmoud, S.M. Elsayed, S. El, M.E. Mahmoud, R.O. Aljedaani, M.A. Salam, [49] X. Yang, H. Zhang, S. Cheng, B. Zhou, RSC Adv 12 (7) (2022) 4101–4112, https://
J Mol Liq 347 (2022) 118274, https://doi.org/10.1016/j.molliq.2021.118274. doi.org/10.1039/d1ra08169a.
[21] G. G, A. Sathish, P.S. Kumar, K. Nithya, G. Rangasamy, Chemosphere 309 (2022) [50] D. Kurniawati, Bahrizal, T.K. Sari, F. Adella, S. Sy, Journal of Physics: Conference
136617, https://doi.org/10.1016/j.chemosphere.2022.136617. Series, Vol. 1788, IOP Publishing Ltd, 2021, p. .
[22] M.E. Mahmoud, M.F. Amira, M.M.H.M. Azab, A.M. Abdelfattah, Sci Rep 12 (1) [51] S. Saber-Samandari, H.O. Gulcan, S. Saber-Samandari, M. Gazi, Water Air Soil
(2022) 19108, https://doi.org/10.1038/s41598-022-23508-y. Pollut 225 (2014) 11, https://doi.org/10.1007/s11270-014-2177-5.
[23] A. Rasaie, M.M. Sabzehmeidani, M. Ghaedi, M. Ghane-Jahromi, A. Sedaratian- [52] W.J. Weber, J.C. Morris, J. Sanit, Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 31.
Jahromi, Mater Chem Phys 262 (2021) 124296, https://doi.org/10.1016/j. [53] S.H. Chien, W.R. Clayton, Soil Sci. Soc. Am. J. 44 (1980) 265.
matchemphys.2021.124296. [54] Y.S. Ho, G. McKay, Process Biochem. 34 (1999) 451.
[24] P.N. Nirenjan Shenoy, N.M. Arjun, P. Senthil Kumar, A.B. Sree Hari, K. Nithya, P., [55] M. Nigam, S. Rajoriya, S. Rani Singh, P. Kumar, J Environ Chem Eng 7 (3) (2019)
Asha sathish, Chemosphere 292 (2022), https://doi.org/10.1016/j. 103188, https://doi.org/10.1016/j.jece.2019.103188.
chemosphere.2021.133375. [56] A. Elkhaleefa, I.H. Ali, E.I. Brima, I. Shigidi, A.B. Elhag, B. Karama, Processes 9
[25] H. Mohammadi, M. Ghaedi, M. Fazeli, M.M. Sabzehmeidani, Microporous and (2021) 3, https://doi.org/10.3390/pr9030559.
Mesoporous Materials 324 (2021) 111275, https://doi.org/10.1016/j. [57] C. Raji, T.S. Anirudhan, Water Res. 32 (1998) 3772.
micromeso.2021.111275. [58] T. Ataei-Germi, A. Nematollahzadeh, J Colloid Interface Sci 470 (2016) 172–182,
[26] M.E. Mahmoud, A.K. Mohamed, M.A. Salam, J Hazard Mater 408 (2021) 124951, https://doi.org/10.1016/j.jcis.2016.02.057.
https://doi.org/10.1016/j.jhazmat.2020.124951. [59] D.M. Saad, E. Cukrowska, H. Tutu, Appl Water Sci 5 (1) (2015) 57–63, https://doi.
[27] S. Singh, E. Arputharaj, H.-U. Dahms, A.K. Patel, Y.-L. Huang, Bioresour Technol org/10.1007/s13201-014-0162-1.
351 (2022) 127018, https://doi.org/10.1016/j.biortech.2022.127018. [60] A.P. Lim, A.Z. Aris, Biochem Eng J 87 (2014) 50–61, https://doi.org/10.1016/j.
[28] G.S. Chauhan, K. Chauhan, S. Chauhan, S. Kumar, A. Kumari, Carbohydr. Polym. bej.2014.03.019.
70 (2007) 415. [61] M. Calero, I. Iáñez-Rodríguez, A. Pérez, M.A. Martín-Lara, G. Blázquez, Bioresour
[29] H. Li, Z. Li, T. Liu, X. Xiao, Z. Peng, L. Deng, Bioresour. Technol. 99 (2008) 6271. Technol 252 (2018) 100–109, https://doi.org/10.1016/j.biortech.2017.12.074.
[30] T.A. Khan, M. Nazir, I. Ali, A. Kumar, Arabian J. Chem. (2013). [62] X. Guo, A. Liu, J. Lu, X. Niu, M. Jiang, Y. Ma, X. Liu, M. Li, ACS Omega 5 (42)
[31] F. Venditti, F. Cuomo, A. Ceglie, L. Ambrosone, F. Lopez, J. Hazard. Mater. 173 (2020) 27323–27331, https://doi.org/10.1021/acsomega.0c03652.
(2010) 552. [63] H.P. Boehm, Discuss Faraday Soc 52 (1971) 264–275, https://doi.org/10.1039/
[32] G.S. Chauhan, S. Kumar, A. Kumari, R. Sharma, J. Appl. Polym. Sci. 90 (2003) DF9715200264.
3856. [64] V. Vinodhini, N. Das, Am-eurasian, J. Sci. Res. 4 (2009) 324.
[33] J.S. Marciano, R.R. Ferreira, A.G. de Souza, R.F.S. Barbosa, A.J. de Moura Junior, [65] F.F.H. Aragón, L. Villegas-Lelovsky, J.G. Parizaka, E.G. Zela, R. Bendezu, R.
D.S. Rosa, Int J Biol Macromol 181 (2021) 112–124, https://doi.org/10.1016/j. O. Gallegos, D.G. Pacheco-Salazar, S.W. da Silva, R. Cohen, L.C.C.M. Nagamine, J.
ijbiomac.2021.03.117. A.H. Coaquira, P.C. Morais, Mater Adv 4 (5) (2023) 1389–1402, https://doi.org/
[34] M.E. Mahmoud, G.A.A. Ibrahim, Int J Biol Macromol 253 (2023), https://doi.org/ 10.1039/D3MA00053B.
10.1016/j.ijbiomac.2023.126489. [66] K.R. Sanchez-Lievanos, J.L. Stair, K.E. Knowles, Inorg Chem 60 (7) (2021)
[35] R. Jobby, P. Jha, A.K. Yadav, N. Desai, Chemosphere 207 (2018) 255–266, https:// 4291–4305, https://doi.org/10.1021/acs.inorgchem.1c00040.
doi.org/10.1016/j.chemosphere.2018.05.050. [67] M. Heidarzadeh-Samani, T. Behzad, A. Mehrabani-Zeinabad, N.B. Baghbadorani,
[36] Q. Li, X. Zhuang, G. Zhou, Z. Yang, T. Yang, H. Xiao, T. Xu, W. Wang, J Environ (2023). 10.21203/rs.3.rs-2604460/v1.
Chem Eng 11 (5) (2023) 110390, https://doi.org/10.1016/j.jece.2023.110390. [68] X. Du, C. Kishima, H. Zhang, N. Miyamoto, N. Kano, Applied Sciences (switzerland)
[37] H. Zhang, M. Zuo, X. Zhang, X. Shi, L. Yang, S. Sun, J. Zhong, Y. Song, Q. Zheng, 10 (2020) 14, https://doi.org/10.3390/app10144745.
Compos Part A Appl Sci Manuf 149 (2021) 106590, https://doi.org/10.1016/j. [69] A. Olad, F. Farshi Azhar, Desalination Water Treat 52(13–15) (2014) 2548–2559,
compositesa.2021.106590. https://doi.org/10.1080/19443994.2013.794711.

161

You might also like