You are on page 1of 11

Fuel 286 (2021) 119447

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

A novel Citrus sinensis peel ash coated magnetic nanoparticles as an easily


recoverable solid catalyst for biodiesel production
Bishwajit Changmai a, Ruma Rano a, Chhangte Vanlalveni b, Lalthazuala Rokhum a, c, *
a
Department of Chemistry, National Institute of Technology Silchar, Silchar 788010, Assam, India
b
Department of Botany, Mizoram University, Tanhril, Aizawl, Mizoram 796001, India
c
Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge CB2 1EW, UK

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

Keywords: A magnetic nano-sized solid catalyst derived from bio-waste Citrus sinensis peel ash (CSPA)@Fe3O4 was devel­
Biodiesel oped for the synthesis of biodiesel from waste cooking oil (WCO). The core–shell structure of the catalyst
Transesterification enhanced the surface properties and maintained a controlled size and shape of the catalyst, thus improved the
Waste cooking oil
catalyst stability to a great extent. The synthesized catalyst was characterized by several analysis techniques. The
Green Fe3O4 nanoparticles
high amount of potassium and calcium in the Citrus sinensis peel ash (CSPA) catalyst makes it extremely basic and
Magnetic recoverability
Citrus sinensis peel ash played an important catalytic role in the transesterification of WCO. The CSPA@Fe3O4 catalyzed trans­
esterification afforded a maximum biodiesel yield of 98% under the optimized reaction conditions such as 6:1
methanol/oil molar ratio, 6 wt% catalyst loading, 65 ◦ C temperature and 3 h time. The activation energy of the
reaction was found to be 34. 41 KJ mol− 1, which indicated that the transesterification of WCO using the present
catalyst is chemically controlled reaction. Beneficially, the magnetic iron oxide core endorsed easily recoverable
of the catalyst from the reaction mixtures by using an external magnet. Moreover, the catalyst showed high
physical stability and reactivity up to the 9 consecutive cycles, demonstrating it as a promising solid base catalyst
for sustainable production of biodiesel.

and commercialization [9–11]. Biodiesel, primarily, is the trans­


1. Introduction esterification product of vegetable oil and animal fat to form fatty acid
methyl esters (FAME) [12]. First-generation biodiesel is generated from
Rapid population growth and increasing industrial revolution have the edible oils, which are already commercialized [13]. However, the
led to a tremendous increase in the consumption of fossil fuels that re­ use of animal fats and edible oils as a source for biodiesel production is
sults in the abnormal climate change and rapid depletion of fossil re­ not a good choice as it leads to a competition between edible oils de­
sources [1,2]. Therefore, to meet the rising energy crisis and minimize mand for food and fuel (food vrs fuel nexus); and also increases the cost
the environmental pollution due to excessive use of fossil fuels, an of the biodiesel production due to the high cost of edible oil feedstocks.
alternative source of energy which is renewable, sustainable and envi­ In this regard, WCO can potentially substitute the feedstocks like animal
ronmentally friendly is highly desirable. Over the last two decades, fats and edible oils for the biodiesel production as it can potentially
extensive research has been carried out for the development of renew­ reduce the cost of the raw feedstocks for biodiesel production and also
able and sustainable energy such as solar energy [3], tidal energy [4], could resolve the problem associated with the disposal of WCO [14]. The
biofuels [5], wind energy [6] etc. Recently, biofuels (bioethanol, bio­ major challenge associated with WCO in the production of biodiesel is
diesel etc.) are widely investigated by many researchers due to its clean the presence of water and higher free fatty acid (FFA) compared to the
combustion, potent substitute for fossil fuels and can be used as a fresh/virgin oil. Ironically, the high FFA contents in vegetable oil react
transportation fuel in heavy-duty vehicles [7,8]. with the alkali metals of the catalyst to form soap via saponification,
Biodiesel has many advantages compared to bioethanol as it has high which inhibits the separation of FAME from glycerol and reduce the
energy density due to the presence of fatty acids having a long carbon FAME yield [15]. In the meantime, researchers have highlighted that the
chain (C14-C20) in the triglycerides. Besides, the easy process of biodiesel FFA content should be held between 0.5 and 2 wt% for effective trans­
production compared to bioethanol is the main reason for its wide use esterification reaction using an alkaline metal catalyst [16].

* Corresponding author at: Department of Chemistry, National Institute of Technology Silchar, Silchar 788010, Assam, India.
E-mail addresses: lalthazualarokhum@gmail.com, rokhum@che.nits.ac.in (L. Rokhum).

https://doi.org/10.1016/j.fuel.2020.119447
Received 29 April 2020; Received in revised form 3 August 2020; Accepted 6 October 2020
Available online 21 October 2020
0016-2361/© 2020 Elsevier Ltd. All rights reserved.
B. Changmai et al. Fuel 286 (2021) 119447

strongly basic medium for the synthesis of magnetic Fe3O4 via co-
Nomenclature precipitation method. Moreover, CSPA produced from orange peel, a
bio-waste material, which not only complements the recycling of natural
WCO Waste cooking oil waste but also reduces the overall cost of the biodiesel production.
FFA Free fatty acid
CSPA Citrus sinesis peel ash 2. Materials and experimental methods
FAME Fatty acid methyl ester
NPs Nanoparticles 2.1. Materials used
TLC Thin layer chromatography
CALB Candida antarctica lipase FeSO4⋅7H2O was purchased from Merck, anhydrous FeCl3 was pur­
HPW Phosphotungstic acid chased from Merck and WCO originated mainly from palm oil was
PGMA Poly(glycidyl methacrylate) collected from the Boy’s Hostel-8, NIT Silchar, Assam, India. The fresh
ASTM American Society for Testing and Materials palm oil for reference was purchased from the Big Bazaar market of
FESEM Field Emission Scanning Electron Microscopy Guwahati, Assam, India.
HR-TEM High Resolution-Transmission Electron Microscopy
XPS X-ray Photoelectron Spectroscopy 2.2. Treatment of feedstocks
EDX Energy Dispersive X-ray Spectroscopy
XRD X-ray Diffraction spectroscopy Firstly, the insoluble impurities and the contaminants of WCO was
XRF X-ray Fluorescence removed via filtration. Then to remove the soluble impurities, hot water
GC-HRMS Gas Chromatography-High Resolution Mass washing was performed on the filtered WCO followed by addition of
Spectrometry anhydrous sodium sulfate to remove the water present. The properties of
NMR Nuclear Magnetic Resonance the WCO and the fresh palm oil were measured and compared using
ASTM (American Society for Testing and Materials) (Supplementary
information, SI, Table S1). The WCO has higher FFA content compared
to the fresh oil as the cooking of oil at high temperature led to the
At present, commercial biodiesel is mainly produced via trans­ breakage of some triglyceride molecule to FFA. As mentioned earlier,
esterification using homogeneous catalysts. Despite the excellent cata­ the presence of FFA reduces the yield of biodiesel and influence the
lytic activity of the commercially available basic/acid catalyst, separation process. In contrast, the WCO used in this report has very low
separation of the catalysts necessitate washing with water for several FFA content (1.1%) and thus has a minimum effect on the FAME yield
times which result in loss of FAME and generation of wastewater. and separation process.
Moreover, the corrosive nature and reusability problem of the catalyst
limits its wide applications [17,18]. To overcome these problems, 2.3. Preparation of the catalyst
nowadays heterogeneous catalysts are widely investigated as it is a
potent substitute for the homogeneous catalysts due to some excellent The fresh oranges (Citrus sinensis) were collected from the local
properties such as non-hazardous, easy recoverability and recyclability market of Aizawl, Mizoram, India. Then the separated peels were
etc. [19,20]. To this end, several novel solid catalysts such as washed properly three times with deionized water and kept under sun
M. acuminata (banana) peel ash [21] Citrus sinensis (orange) peel ash for 3 days to dry. Then the dried peels were burnt in the air for 30 min to
[22], snail shell [23] and Ag/ZnO nanoparticles (NPs) [24] for the facile form CSPA [22]. 20 g of the resulted ash was poured into a 250 mL round
production of biodiesel from vegetable oil have been reported in recent bottom (RB) flask containing 150 mL deionized water and stirred for 2 h
years. at 80 ◦ C to extract the basic components present in CSPA, which is
Over the past few decades, magnetic iron oxide nanoparticles (NPs) termed as CSPA extract in the present study.
have emerged as one of the primary nanoparticles in the field of Magnetic Fe3O4 NPs was synthesized via traditional co-precipitation
biomedical and heterogeneous catalysts [25–27]. Due to its highly method with slight modification. To prepare the magnetite nanoparticle
magnetic nature, they are a potential substitute for centrifugation and FeSO4⋅7H2O (1.94 g, 7 mmol) and FeCl3 (2.27 g, 14 mmol) (Fe2+/
filtration and hence reduces the loss of catalyst, increases the reusability, Fe3+=1/2 M ratio) were mixed in deionized water (140 mL) and stirred
makes the catalyst more profitable and can be used in industrial appli­ vigorously at 65 ◦ C. Then, CSPA extract was added dropwise until the
cations [25,28,29]. Several magnetic Fe3O4 supported catalysts are colour of the solution turns to dark brown. The solution was stirred for
already reported for the biodiesel production e.g., CALB/Fe3O4@SiO2 another 30 min and allowed to settle followed by decantation of the
[30], HPW-PGMA-Fe3O4 [31], MgFe2O4@CaO [32], Ca/Al/Fe3O4 [33], solution and collected the solid part, which was then washed with
sulfonic acid functionalized Fe/Fe3O4 core–shell [34], [SO3H-(CH2)3- deionized water to remove the unreacted iron salts and water-soluble
HIM][HSO4]/Fe3O4@SiO2 [35] etc. However, synthesis of these cata­ impurities. After that, 70 mL of CSPA extract was added to the resul­
lysts involves multi-steps and harsh reaction conditions that increase the ted magnetite nanoparticle and stirred for 3 h at room temperature (RT).
overall cost and reduced environmental compatibility of the biodiesel Finally, the catalyst was collected by evaporating the water using a ro­
production; thereby, limits their industrial application. In this context, tary evaporator. The amount of the recovered catalyst was measured
the synthesis of magnetic nanoparticles using bio-waste material could and was found to be 1.93 g. The process for catalyst synthesis is depicted
provide an alternative sustainable and cost-effective production of bio­ schematically in Scheme 1.
diesel. Several methods are already reported regarding the biogenic
synthesis of magnetic nanoparticles using fungi [36], yeast [37], bac­ 2.4. Characterization of catalyst
teria [38], plant leaves [39] and fruit peel [40] etc. A deep study of the
reported literature reveals that till now no methods are available for the It is well known that for any transesterification reaction, the solid
synthesis of Fe3O4 nanoparticles using Citrus sinensis peel ash (CSPA) catalyst should have various physicochemical properties such as high
extract. basicity to abstract proton [41], high surface area to provide numerous
The present study reported the bio-genic synthesis of basic magnetic active sites and large pore diameter to allow the reactant and product
Fe3O4 NPs using CSPA extract followed by a coating of CSPA on Fe3O4 molecules to diffuse in and out, respectively [25]. Hence, to study the
nanoparticle. The resulted catalyst (CSPA@Fe3O4) was used for the structure, morphology, composition, surface area, pore diameter and
transesterification of WCO. The highly basic nature of CSPA provides a crystallinity of the synthesized catalyst, various characterization

2
B. Changmai et al. Fuel 286 (2021) 119447

catalytic performance for transesterification reaction [14]. Thus, the


influence of the reaction parameters on the transesterification of WCO
using CSPA@Fe3O4 was investigated to find out the optimized reaction
conditions.

2.5. Transesterification of WCO

Transesterification of WCO using CSPA@Fe3O4 catalyst (Scheme 2)


was performed in a three-neck round bottom flask equipped with a
water condenser. The reaction parameters such as methanol-to-oil molar
ratio, catalyst loading, reaction temperature and reaction time were
varied from 1:3–1:12, 2–10 wt%, 30–110 ◦ C and 1–3.5 h, respectively to
examine their effects in the transesterification of WCO and also to find
the optimized reaction conditions. The reaction procedure at optimized
reaction conditions is described as follows: In an RB, methanol (4 mL)
and WCO (14 g) were mixed in a 6:1 methanol/oil molar ratio followed
by addition of 0.84 g catalyst (6 wt% with respect to oil) and stirred at
65 ◦ C for 3 h. The progress of the reaction was monitored via TLC (thin
layer chromatography). After completion of the reaction, the catalyst
Scheme 1. Catalyst preparation process.
was recovered from the reaction mixture by using an external magnet.
The excess methanol present in the reaction mixture was removed via
techniques have been used such as X-ray fluorescence (XRF), high- the rotary evaporator. Finally, the resultant product was separated from
resolution transmission electron microscopy (HR-TEM), X-ray photo­ the byproduct glycerol via centrifugation (3500 rpm for 25 min).
electron spectroscopy (XPS), Energy-dispersive X-ray spectroscopy
(EDX), X-ray diffraction (XRD), Fourier-transform infrared spectroscopy
(FT-IR) and Brunauer-Emmett-Teller (BET) analysis. 2.6. Characterization of FAME
XRF was performed in Bruker; model: S4 pioneer. Operating condi­
tions: the powdered sample was placed on a circular shape block having To examine the formation and chemical structure of the synthesized
a diameter of 38–40 mm and 10% methane, 90% argon were used as a FAME, 1H NMR (Nuclear magnetic resonance) and 13C NMR was per­
detector gas to study the presence of metallic and non-metallic oxides formed on JEOL ECZR series instrument with a spectrometer frequency
present in CSPA. To study the elemental composition present in the of 600 MHz and tetramethyl silane (TMS). GC-HRMS (Gas
catalyst, EDX analysis was performed in JEOL model JSM – 6390LV chromatography-high resolution mass spectrometry) of FAME was
instrument; operating conditions: 80 mA beam current, 20 kV and analyzed to investigate the chemical composition of the synthesized
1500X magnification. Similarly, HR-TEM was done on Jeol/JEM 2100 FAME. GC of GC-HRMS was performed on Agilent 7890 instrument
operated at 200 kV and LaB6 Electron gun was used to study the particle equipped with head space injector mode, a CPSIL 8CB capillary column
shape and size. In addition, XPS was performed on ESCALAB Xi + in­ (30 m × 0.25 mm × 0.25 µm) and GC-FID detector. The oven temper­
strument to investigate the major elements present in the catalyst. XRD ature was started from 55 ◦ C and increase with a rate of 10 ◦ C min-1 up
was performed on a Bruker D8 advance instrument to study the crys­ to a final temperature of 230 ◦ C. The temperature of the detector and the
tallinity of the catalyst. Apart from that, FT-IR analysis was performed injector were 300 ◦ C and 250 ◦ C respectively. For quantitative analysis
on BRUKER AVANCE II 400 spectrometer to study the presence of of the products, biphenyl was used as an internal standard to investigate
functional groups present in the catalyst. To study the surface area of the the composition of the resulted biodiesel. MS part of GC-HRMS was
catalyst BET analysis was performed on Micromeritics ASAP 2010 sur­ performed on AccuTOF GCV instrument with a mass range of 10–2000
face area and porosity analyzer. Operating conditions: bath temperature amu and mass resolution of 6000. Apart from that, the properties of the
77 K and degassed at 200 ◦ C for 10 h. synthesized FAME was determined by using various ASTM methods such
Apart from the basicity, surface area and pore diameter of the cata­ as D1980-87(1998), D445 and D2500 etc. The percentage yield of the
lyst, reaction parameters such as methanol/oil molar ratio, catalyst synthesized FAME was calculated by using the Eq. (1) [14]. On the other
loading, reaction temperature and time have a major influence on hand, the percentage conversion of FAME from the 1H NMR data was
calculated by using Knothe and Kenar Eq. (2) [23].

Scheme 2. Schematic representation of transesterification of WCO using CSPA@Fe3O4.

3
B. Changmai et al. Fuel 286 (2021) 119447

Weight of FAME CSPA, XRF analysis was performed. XRF data (SI, Table S2) shows the
FAMEyield(%) = × 100 (1)
Weight of WCO presence of K2O (51.9%), CaO (24%), SiO2 (13%) as a major component
and P2O5 (2.95%) and MgO (4.76%) etc., as a minor component. Thus,
Conversion(%) = 100 ×
2AMe
(2) the presence of these highly basic alkaline metal oxide plays a tremen­
3ACH2 dous role in the preparation of the magnetic Fe3O4 NPs [42]. Further­
more, coating of CSPA over Fe3O4 NPs makes the catalyst highly basic
here, C, AMe and ACH2 are the percentage oil conversion, integration and therefore can be utilized in the conversion of WCO biodiesel.
values of methoxy protons and methylene protons respectively. XPS was used to investigate the catalyst chemical composition as
shown in Fig. 1. The survey data (Fig. 1A) of XPS confirms the presence
of Ca, K, C, O and Mg, and the atomic wt. % of the respective elements
2.7. Determination of activation energy:
are shown in Table 3. Apart from Fe, as displayed by Table S3, the
catalyst consists the highly basic elements such as K (8.64%) and Ca
Determination of activation energy was carried out by investigating
(4.46%) in a major amount, which is mainly accountable for the
the temperature effect on the transesterification reaction. It is reported
excellent reactivity of the catalyst [43,44]. On deconvolution, K 2p
that transesterification of WCO using an excess amount of methanol to
spectra (Fig. 1B) show two peaks at 292.76 eV and 295.22 eV, which
produce biodiesel follows pseudo-first-order kinetics [14]. Therefore,
attributed to the presence of K2O and K2CO3 respectively [21]. The
the rate constant can be determined using Eq. (3):
deconvoluted spectra of Ca 2p (Fig. 1C) gives two humps at 346.76 eV
− ln (1 − Ct ) = kt (3) and 350.84 eV, which correspond to CaO and CaCO3 respectively. C 1s
spectra (Fig. 1D) show two humps at 258.18 eV and 293.31 eV, attribute
here, ‘Ct’ is the amount of FAME produced after the time ‘t’ and ‘k’ is the to C–C and C– – O bond respectively [22]. Moreover, O 1s spectra
rate constant. (Fig. 1E) shows one hump at 531.21 eV. EDX analysis was also per­
To calculate the activation energy, the rate constant of the trans­ formed to further verify the chemical composition of CSPA@Fe3O4
esterification reaction was measured at different reaction temperatures catalyst. The EDX data (Fig. 2A) displayed the presence of K, Ca, Mg etc.
(30 ◦ C, 45 ◦ C and 65 ◦ C) and finally, the respective rate constants are The data obtained from EDX analysis supplements the XPS data and
inserted in the Arrhenius equation (4). atomic wt. % of the elements are also presented in Fig. 2A indicates the
Ea presence of highly basic elements such as K (8.95%) and Ca (5.01%) in
lnk = − + lnP (4) major amount apart from Fe.
RT
To study the crystalline nature and chemical composition of the
here, ‘Ea’ is the activation energy, ‘R’ is the gas constant (8.314 × 10− 3 CSPA@Fe3O4, XRD was carried out. Fig. 2B shows peaks at 2θ = 29.72,
KJ mol− 1), ‘T’ is the reaction temperature and ‘P’ is the pre-exponential 40.68, 35.61 and 43.49, which attributed to the presence of K2O (JCPDS
factor or frequency factor. file no: 77-2178) and K2CO3 (JCPDS file no: 71-1466). The 2θ value at
28.46, 57.52, 21.44 and 50.37 confirms the presence of CaCO3 (JCPDS
3. Results and discussions file no. 28-0775) and CaO (JCPDS file no: 72-1214). Moreover, the peaks
at 66.50 and 73.90 confirms the presence of MgO (JCPDS file no. 57.83)
3.1. Catalyst characterization and SiO2 (JCPDS file no. 89-3609) [22].
To measure the catalyst basicity, Hammet indicator method [45] was
To investigate the presence of metallic and nonmetallic oxides in applied and found that the basicity of the catalyst lies between 11.0 ≤ H_

Fig. 1. XPS survey spectra of the catalyst (A) and deconvoluted spectra of K 2p (B), Ca 2p (C), C 1s (D) and O 1s (E).

4
B. Changmai et al. Fuel 286 (2021) 119447

Fig. 2. EDX spectra (A), XRD pattern (B), FT-IR spectra (C), TEM images at 10 nm (D) and 20 nm (E) scale, SAED pattern (F), N2 adsorption–desorption isotherm (G)
of CSPA@Fe3O4 and comparison of the magnetic hysteresis loop of bare Fe3O4 and CSPA@Fe3O4 (H). Fig. 3: Influence of reaction parameters (A) molar ratio, (B)
catalyst loading, (C) reaction temperature and (D) reaction time.

≤ 12.2. Furthermore, titration of benzoic acid was also performed to pattern (Fig. 2F) reveals the polycrystalline nature of CSPA@Fe3O4. [47]
measure the catalyst basicity [46], where the indicator used as bromo­ BET analysis was performed to examine the surface area, pore vol­
thymol blue. The calculated basicity of the catalyst is 0.170 mmol g− 1. ume and pore diameter of CSPA@Fe3O4. Fig. 2G displayed the N2
Owing to the high basicity of CSPA@Fe3O4, it shows remarkable reac­ adsorption–desorption isotherm, which exhibits a type IV isotherm with
tivity in the conversion of WCO to biodiesel. H1 hysteresis loop, suggesting the mesoporous nature of CSPA@Fe3O4
Functional groups present in the catalyst was tested by performing [22]. The catalyst has a pore diameter of 2.45 nm, which lies in the
FTIR analysis (Fig. 2C). The peak at around 3422.55 cm− 1 may be mesoporous range and thus supports the N2 adsorption–desorption
attributed to O–H group present in the catalyst due to the adsorption of isotherm. Moreover, the calculated surface area of CSPA@Fe3O4 was
moisture [14]. Other peaks at 1637.46, 1112.05 and 895 cm− 1 were found to be 15.55 m2/g. Owing to the high surface area of the catalyst,
observed due to the stretching and bending vibration of carbonate the available surface for the reaction to take place is very high and thus
groups respectively. In addition, two peaks at 794.57 and 615.52 cm− 1 the rate of the reaction is also improved [48].
was observed, which may be attributed to the stretching vibration fre­ To study the effect of impregnation on the magnetic moment of bare
quencies of K-O and Ca-O bonds respectively. [46] Fe3O4, VSM analysis was performed as shown in Fig. 2H displayed a
To study the structure and particle size of the catalyst TEM analysis closed M− H loop, which suggests the presence of high magnetic
was performed. Fig. 2 (D, E) displayed a core–shell like structure, sug­ behaviour of both bare Fe3O4 and CSPA@Fe3O4 [47]. However, the
gests the successful impregnation of metal oxides and carbonates on magnetic moment of bare Fe3O4 and CSPA@Fe3O4 are 37.74 and 31.56
Fe3O4 NPs. The average particle size of the nanoparticle is ̴ 12–13 nm as emu/g respectively. The drop in magnetism of CSPA@Fe3O4 is because
shown in Fig. 2D. The smaller particle size ascribed to high surface/ on impregnation the surface of the highly magnetic Fe3O4 was covered
volume ratio, which supports the BET analysis data. Moreover, the SAED by metal oxides and carbonates, which indicates that CSPA was

5
B. Changmai et al. Fuel 286 (2021) 119447

successfully impregnated on Fe3O4 surface. Table 2


Physico-chemical properties of FAME and WCO along with ASTM limits for
3.2. Characterization of FAME FAME.
Properties WCO FAME ASTM limits for Testing methods
1
H NMR was performed and compared with the 1H NMR of WCO to FAME [49] (ASTM)
confirm the formation of FAME (SI, Fig. S1). 1H NMR spectra of WCO (SI, Acid value (mg KOH 1.1 0.05 <0.79 D1980-87(1998)
Fig. S1A) shows two peaks at 4.28 and 5.34 ppm, which are responsible g− 1)
for the glyceridic and olefinic protons respectively. However, the gen­ Kinematic viscosity 32.5 4.34 1.9–6.0 D 445
at 40 ◦ C
eration of a peak at 3.67 ppm is attributed to the methoxy protons and
Cloud point (◦ C) 9.2 3 − 3 to 12 D 2500
complete dissipation of the peak at 4.28 ppm in the 1H NMR spectra of Pour point (◦ C) 0.9 − 2.5 − 15 to 6 D 97
FAME (SI, Fig. S1B) confirms the successful formation of FAME [24]. 13C Flash point (◦ C) 290 149 >93 D 93
NMR (SI, Fig. S2) of both WCO and FAME was performed to further Cetane number 53 57 ≥47 D 6890
Density (g cm− 3) 0.9 0.87 0.86 to 0.9 D 1448–1972
confirm the formation of FAME. 13C NMR of WCO (SI, Fig. S2A) shows
two peaks at 68.86 and 62.04 ppm, which are attributed to the glyceridic
carbons. In contrast, the generation of a peak at 51.46 ppm (attributed percentage oil conversion. The highest conversion of 98.7% was ob­
for methoxy carbon) and dissipation of peaks for glyceridic protons in tained when the catalyst concentration rises from 4 wt% to 6 wt% as
the (SI, Fig. S2B) affirms the successful formation of FAME, thus sup­ availability of active sites increases with an increase in catalyst amount.
ports the 1NMR data [22]. The peak at 174 ppm is responsible for the [51] However, a further increase in the catalyst amount led to reduction
carbonyl carbon. From the 1H NMR data, the percentage conversion of of biodiesel yield and conversion. This is due to the fact that, on crossing
WCO was calculated by using Eq. (2) and found to be 98.7%. Moreover, the optimum catalyst amount, the reaction mixture becomes more
Eq. (1) is utilized to calculate the percentage yield of synthesized FAME viscous and thus diffusion of reactants to oil-catalyst-methanol is
and found to be 98%. disturbed. [52] Therefore, the optimum catalyst amount is 6 wt%.
Chemical composition of the synthesized FAME was investigated by
complying GC–MS analysis (SI, Fig. S3) was performed. Five consider­ 3.3.3. Influence of reaction temperature
able peaks were attained in the total ion chromatogram. Quantitative The transesterification of triglyceride using a solid catalyst is a three-
measurement of GC–MS data revealed the presence of methyl oleate, species/phase system (methanol-catalyst-oil) and therefore diffusion of
methyl linoleate and methyl palmitate in major components (Table 1) reactant molecules from one phase to another is difficult. This diffusion
Physico-chemical properties of synthesized FAME was measured and resistance can be minimized by applying temperature to the reaction
compared with the properties of WCO (Table 2). It is well observed from system. The influence of reaction temperature (Fig. 3C) on the trans­
Table 2 that all the Physico-chemical properties of FAME meet the ASTM esterification of WCO has been investigated in the range 30–110 ◦ C
standards [49]. using the reaction conditions: 6:1 methanol/oil molar ratio, catalyst
amount of 6 wt% for 3 h. Fig. 3C shows that conversion increases with
3.3. Influence of reaction parameters increase in temperature and maximum conversion of 98.7% were ob­
tained when the temperature of the reaction system was increased from
3.3.1. Influence of methanol/oil molar ratio 45 to 65 ◦ C, as the transesterification reaction is endothermic in nature,
The reaction rate of the transesterification reaction can drive to the thus increases in temperature facilitates the reaction towards the for­
forward direction by using an excess amount of methanol as the reaction ward direction. Moreover, high temperature lead to a vigorous collision
is reversible in nature. The influence of methanol/oil molar ratio on between the three species/phase system and thus increases the misci­
biodiesel production in the range 4:1 to 12:1 (Fig. 3A) under the reaction bility [53]. However, conversion/yield decreases when the further in­
conditions such as 6 wt% catalyst amount, 65 ◦ C of temperature for 3 h crease in temperature i.e., beyond 65 ◦ C as at high-temperature
have been investigated. On increment of methanol/oil molar ratio from methanol vaporizes, thus the availability of methanol in the reaction
4:1 to 6:1, the reaction gives a maximum conversion of 98.7% of WCO. mixture decreases [54]. Therefore the optimum temperature is 65 ◦ C.
However, a further rise in the methanol concentration does not rise the
oil conversion rather it decreases the biodiesel yield and conversion as 3.3.4. Influence of reaction time
excess methanol diluted the reaction mixture, thus the efficiency of the The influence of reaction time on the conversion of WCO oil to
collision between the reactant molecules decreases. Moreover, a higher biodiesel (Fig. 3D) was tested to get a suitable/optimum reaction time in
amount of methanol facilitates of the glycerol solubility, results in dif­ the range 1–3.5 h using the reaction conditions: 6:1 methanol/oil molar
ficulty in separation of biodiesel from glycerol [50]. Thus, the optimized ratio, catalyst amount of 6 wt% and a temperature of 65 ◦ C. Fig. 3D
methanol/oil molar ratio is 6:1. depicted an increase in the percentage conversion with time and the
maximum conversion of 98.7% was obtained in 3 h. The conversion
3.3.2. Influence of CSPA@Fe3O4 loading remains unchanged after 3 h and therefore, 3 h is the optimum reaction
The influence of catalyst amount (Fig. 3B) in the transesterification time.
of WCO was investigated in the range 2 wt% to 10 wt%, using the re­
action conditions: 6:1 methanol/oil molar ratio, 65 ◦ C temperature and
3 h. Fig. 3B shows that with a rise in the catalyst amount also raise the 3.4. A probable mechanism for transesterification of WCO catalyzed by
CSPA@Fe3O4
Table 1
GC–MS data for the chemical composition of FAME. In a heterogeneous catalysis reaction, the reactant molecules diffuse
to the catalyst surface, where they get adsorbed via chemical bonding
Sl. Retention Chemical Corresponding Amount
No time composition acid (%) followed by interaction of the reactant molecules to give products.
Finally, the product gets desorbed from the catalyst surface. In the
1 22.27 Methyl palmitate C 16:0 20
2 24.15 Methyl linolenate C 18:3 4.1 present study, transesterification of WCO was carried out by using a
3 27.02 Methyl lenoleate C 18:2 30.5 solid base catalyst CSPA@Fe3O4. Since, the elemental analysis of the
4 30.40 Methyl oleate C 18:1 42.1 catalyst shows the presence of highly basic elements such as K, Ca and
5 32.11 Methyl cis-11- C 20:1 2.1 Mg in major amount, we consider the oxides of K, Ca and Mg in the
eicosenoate
plausible mechanism demonstrated in Scheme 3. At first, the O2−

6
B. Changmai et al. Fuel 286 (2021) 119447

Table 3
Comparison of the present catalyst with various reported catalyst.
Sl. No Catalyst Methanol/oil molar ratio Catalyst loading (wt%) Time (h) Temp. (◦ C) Yield (%) Ref.

1 TiO2-MgO 30:1 5 6 150 89.6 [57]


2 Mo-Mn/γ-Al2O3-MgO 27:1 15 4 100 91.4 [58]
3 KBr/CaO 12:1 4 2 65 78.9 [59]
4 CaO/MgO 15:1 6 2 90 96.47 [44]
4 Chicken manure (CaO) 15:1 7.5 4 65 90 [14]
5 Chicken bone 15:1 5 4 65 89.33 [43]
6 CaO/SiO2 14:1 8 1.5 60 91 [60]
7 Lipase/Fe3O4-Au 6:1 20 24 45 90 [48]
8 CSPA@Fe3O4 6:1 6 3 65 98 Present work

Fig. 3. Influence of reaction parameters (A) molar ratio, (B) catalyst loading, (C) reaction temperature and (D) reaction time.

present in the catalyst abstract a proton from methanol to generate In this present work, the efficiency of CSPA@Fe3O4 catalyst in lowering
methoxide ion (Nucleophile). In the second step, methoxide ion attacks the activation energy of the transesterification of WCO using methanol
the electron-deficient carbonyl carbon of the triglyceride to form a was examined. In this context, rate constant was measured at different
tetrahedral intermediate followed by rearrangement of the tetrahedral reaction temperatures such as 30 ◦ C, 45 ◦ C and 65 ◦ C and inserted the
intermediate to methyl ester (biodiesel) and diglyceride anion. In the data in the Arrhenius equation. Then the data of ‘ln k’ was plotted
fourth step, diglyceride anion abstracts a proton from the catalyst to against the respective inverse reaction temperature (1/T), which resul­
form diglyceride. Similarly, the second mole of methoxide ion reacts ted in a straight line as displayed in Fig. 4, fitted well in the pseudo-first-
with the diglyceride to form monoglyceride and methyl ester. Subse­ order kinetic model. Thus, the slope of the straight line (Eq. (5)) provides
quently, monoglyceride was converted to glycerol and methyl ester by the activation energy of the reaction and the intercept of the line gave
the third mole of methoxide ion. Hence, one mole of triglyceride gen­ the pre-exponential factor or the frequency factor.
erates three moles of methyl ester and one mole of glycerol.
Slope = Ea/R (5)
The slope was calculated from the straight line and it was found to be
3.5. Activation energy − 4.14 and correspondingly the value of activation energy was found to
be 34.41 KJ mol− 1, which lies well in between the range of activation
In a transesterification reaction, use of catalyst plays a vital role as energy (21 kJ mol− 1 to 84 KJ mol− 1) required for transesterification of
they increase the rate of the reaction by lowering the activation energy.

7
B. Changmai et al. Fuel 286 (2021) 119447

Scheme 3. Probable mechanism of CSPA@Fe3O4 catalyzed transesterification of WCO.

and Astrocaryum aculeatum (61.23 KJ mol− 1) [56]. Besides, the trans­


esterification of WCO using the present catalyst is a chemically
controlled reaction as the activation energy of this reaction is higher
than 25 KJ mol− 1. The frequency factor or the pre-exponential factor of
the reaction was also calculated from the intercept of the straight line
and found to be 1.58 × 108 min− 1.

3.6. Comparison of the present catalyst with the other reported catalyst

Various heterogeneous catalysts were already available reported


earlier for the conversion of WCO to biodiesel and are compared with
present catalyst as displayed in Table 3. TiO2-MgO, Mo-Mn/γ-Al2O3-
MgO, KBr/CaO, CaO/MgO etc., showed high yield and good stability
towards the transesterification reaction to form biodiesel, but the main
drawbacks of these catalysts are their complex preparation methods and
Fig. 4. Plot of ln k vs 1/T, Reaction temperature: 30 ◦ C, 45 ◦ C and 65 ◦ C. handling issue as preparation of such catalysts require hazardous
chemicals and drastic reaction conditions. However, such drawbacks
oil [55]. Moreover, the efficiency of lowering the activation of the can be minimized by using bio-waste derived heterogeneous catalysts
present catalyst CSPA@Fe3O4 is way better than the other reported such as chicken manure (CaO), chicken bone, CaO/SiO2 etc., as they are
catalyst such as CaO derived from chicken manure (78.8 KJ mol− 1) [14] easy to handle, no requirement of toxic chemicals and most importantly

8
B. Changmai et al. Fuel 286 (2021) 119447

they are widely available low-cost natural waste, thus reduced the
overall cost of FAME production. However, these catalysts are not
without shortcoming as preparation of such catalysts need very high
calcination temperature, longer preparation time which may deter their
applications in industrial scale. In the present work, preparation of the
catalyst (CSPA@Fe3O4) does not require any calcination process and
harsh reaction conditions. Moreover, CSPA@Fe3O4 resulted in very high
biodiesel conversion of 98.76% and showed great stability in FAME
synthesis from WCO as it can be reused for 9 successive cycles with a
minimal loss of catalyst reactivity.

4. Reusability of CSPA@Fe3O4 catalyst

Reusability of a catalyst can be considered as a very vital factor to


utilize the catalyst in industrial scale. To examine the reusability of
CSPA@Fe3O4, the catalyst was recovered from the reaction mixture via Fig. 5. Reusability test of CSPA@Fe3O4 catalyst up to 9th cycle.
an external magnet, washed with methanol for 2–3 times to remove any
organic compound after the reaction and dried in an oven for 3–4 h at biowaste-based catalyst can be reused for at least 9 successive reaction
80 ◦ C. The reusability of CSPA@Fe3O4 catalyst was tested for 9 cycles without much depreciation in the biodiesel yield. Moreover, the
consecutive cycles under the optimized reaction conditions such as use of bio-waste materials as a source of catalyst complements the po­
methanol/oil molar ratio of 6:1, catalyst amount of 6 wt%, 65 ◦ C of tential recycling of natural waste resources as well as reduce the total
temperature for 3 h (Fig. 5) and same recovery method. In each suc­ cost and negative environmental impact of biodiesel production for in­
cessive step, the conversion of WCO to FAME was found to decreases dustrial scale.
slightly in each successive cycles. A conversion of 91% was observed in
the 9th reaction cycle. The decrease in the conversion of WCO to bio­
CRediT authorship contribution statement
diesel can be attributed to the leaching of the basic components in the
catalyst [22].
Bishwajit Changmai: Conceptualization, Methodology, Validation,
After 9th cycle of the catalyst, the change in the particle size and
Investigation, Writing - original draft, Visualization. Ruma Rano:
structure of CSPA@Fe3O4 was tested by performing TEM analysis (SI,
Conceptualization, Validation, Writing - review & editing, Supervision.
Fig. S4), which showed no serious change in the structure and particle
Chhangte Vanlalveni: Validation, Resources, Writing - review & edit­
size. The particles retained its core–shell structure and the average
ing, Visualization. Lalthazuala Rokhum: Conceptualization, Valida­
particle size is around 16–17 nm.
tion, Investigation, Writing - review & editing, Supervision.
EDX analysis was used to study the change in the elemental
composition of CSPA@Fe3O4 after 9th cycle. The EDX spectrum and data
(SI, Fig. S5), showed a slight change in the atomic wt. % of catalytically Declaration of Competing Interest
active elements such as K and Ca from 8.95% and 5.01% to 8.42% and
4.75% respectively. The authors declare that they have no known competing financial
XRD (SI, Fig. S6) was executed to study the change in crystallinity of interests or personal relationships that could have appeared to influence
CSPA@Fe3O4 after the 9th cycle. Fig. S6 showed no change in the the work reported in this paper.
crystallinity of the recovered CSPA@Fe3O4 even after the 9th run. This
observation reinforced the high stability of the catalyst CSPA@Fe3O4 Acknowledgement
revealed by EDX analysis.
To study any changes in functional group in the regenerated SERB, New Delhi is acknowledged for the research fund (Grant No.
CSPA@Fe3O4 after the 9th cycle, FT-IR analysis was carried out (SI, SB/FT/CS-103/2013 and SB/EMEQ-076/2014). The authors also
Fig. S7). A stretching band was attained at around 3415.54 cm− 1, acknowledge SAIF NEHU, STIC Cochin, IIT Bombay and IIT Guwahati
ascribed to the O–H functional group. The attainment of peaks at for sample characterization.
1622.79 and 1118.96 cm− 1 ascribed to the stretching and bending vi­
bration of the carbonate group. In addition, two peaks were attained at
Appendix A. Supplementary data
794 and 615.12 cm− 1, ascribed to the K-O and Ca-O stretching vibration
respectively. Thus, the FT-IR spectra for the regenerated CSPA@Fe3O4
Supplementary data to this article can be found online at https://doi.
after the 9th run is similar to the FT-IR spectra of fresh regenerated
org/10.1016/j.fuel.2020.119447.
CSPA@Fe3O4 as all the functional groups remain intacked.

References
5. Conclusion
[1] Zhang S, Zu YG, Fu YJ, Luo M, Zhang DY, Efferth T. Rapid microwave-assisted
In this present work, novel iron oxide (Fe3O4) nanoparticle was transesterification of yellow horn oil to biodiesel using a heteropolyacid solid
successfully synthesized via bio-genic route and coated it with Citrus catalyst. Bioresour Technol 2010;101:931–6. https://doi.org/10.1016/j.
biortech.2009.08.069.
sinensis peel ash to form a highly basic CSPA@Fe3O4 nanocatalyst. The [2] Zhao C, Yang L, Xing S, Luo W, Wang Z, Lv P. Biodiesel production by a highly
exploitation of magnetic Fe3O4 as catalyst support enhanced the effective renewable catalyst from pyrolytic rice husk. J Clean Prod 2018;199:
dispersion and stability of the catalyst towards the WCO conversion to 772–80. https://doi.org/10.1016/j.jclepro.2018.07.242.
[3] Lewis NS. Solar Energy Use. Science (80) 2007;798:798–802. https://doi.org/
biodiesel. The prepared base catalyst showed excellent activity for the
10.1126/science.1137014.
transesterification of WCO to biodiesel with 98% yield under the opti­ [4] Khojasteh D, Khojasteh D, Kamali R, Beyene A, Iglesias G. Assessment of renewable
mized reaction conditions such as 6:1 methanol/oil molar ratio, 6 wt% energy resources in Iran; with a focus on wave and tidal energy. Renew Sustain
catalyst loading, 65 ◦ C reaction temperature and 3 h reaction time. The Energy Rev 2018;81:2992–3005. https://doi.org/10.1016/j.rser.2017.06.110.
[5] Molino A, Larocca V, Chianese S, Musmarra D. Biofuels production by biomass
presence of K and Ca in CSPA increases the basicity of the and thus gasification: a review. Energies 2018;11:1–31. https://doi.org/10.3390/
escalate its catalytic activity for the transesterification reaction. The en11040811.

9
B. Changmai et al. Fuel 286 (2021) 119447

[6] Carvalho D, Rocha A, Gómez-Gesteira M, Silva Santos C. Offshore winds and wind [31] Zillillah, Ngu TA, Li Z. Phosphotungstic acid-functionalized magnetic nanoparticles
energy production estimates derived from ASCAT, OSCAT, numerical weather as an efficient and recyclable catalyst for the one-pot production of biodiesel from
prediction models and buoys – a comparative study for the Iberian Peninsula grease via esterification and transesterification. Green Chem 2014;16:1202–10.
Atlantic coast. Renew Energy 2017;102:433–44. https://doi.org/10.1016/j. https://doi.org/10.1039/c3gc41379a.
renene.2016.10.063. [32] Liu Y, Zhang P, Fan M, Jiang P. Biodiesel production from soybean oil catalyzed by
[7] Ibarra-Gonzalez P, Rong BG. A review of the current state of biofuels production magnetic nanoparticle MgFe2O4@CaO. Fuel 2016;164:314–21. https://doi.org/
from lignocellulosic biomass using thermochemical conversion routes. Chinese J 10.1016/j.fuel.2015.10.008.
Chem Eng 2019;27:1523–35. https://doi.org/10.1016/j.cjche.2018.09.018. [33] Tang S, Wang L, Zhang Y, Li S, Tian S, Wang B. Study on preparation of Ca/Al/
[8] Menegazzo ML, Fonseca GG. Biomass recovery and lipid extraction processes for Fe3O4 magnetic composite solid catalyst and its application in biodiesel
microalgae biofuels production: a review. Renew Sustain Energy Rev 2019;107: transesterification. Fuel Process Technol 2012;95:84–9. https://doi.org/10.1016/j.
87–107. https://doi.org/10.1016/j.rser.2019.01.064. fuproc.2011.11.022.
[9] Hajjari M, Tabatabaei M, Aghbashlo M, Ghanavati H. A review on the prospects of [34] Wang H, Covarrubias J, Prock H, Wu X, Wang D, Bossmann SH. Acid-
sustainable biodiesel production: A global scenario with an emphasis on waste-oil functionalized magnetic nanoparticle as heterogeneous catalysts for biodiesel
biodiesel utilization. Renew Sustain Energy Rev 2017;72:445–64. https://doi.org/ synthesis. J Phys Chem C 2015;119:26020–8. https://doi.org/10.1021/acs.
10.1016/j.rser.2017.01.034. jpcc.5b08743.
[10] Chuah LF, Klemeš JJ, Yusup S, Bokhari A, Akbar MM. A review of cleaner [35] Wu Z, Li Z, Wu G, Wang L, Lu S, Wang L, et al. Brønsted acidic ionic liquid modified
intensification technologies in biodiesel production. J Clean Prod 2017;146: magnetic nanoparticle: An efficient and green catalyst for biodiesel production. Ind
181–93. https://doi.org/10.1016/j.jclepro.2016.05.017. Eng Chem Res 2014;53:3040–6. https://doi.org/10.1021/ie4040016.
[11] Abdullah SHYS, Hanapi NHM, Azid A, Umar R, Juahir H, Khatoon H, et al. [36] Sastry M, Ahmad A, Islam Khan M, Kumar R. Biosynthesis of metal nanoparticles
A review of biomass-derived heterogeneous catalyst for a sustainable biodiesel using fungi and actinomycete. Curr Sci 2003;85:162–70.
production. Renew Sustain Energy Rev 2017;70:1040–51. https://doi.org/ [37] Moghaddam AB, Namvar F, Moniri M, Tahir PM, Azizi S, Mohamad R.
10.1016/j.rser.2016.12.008. Nanoparticles biosynthesized by fungi and yeast: a review of their preparation,
[12] Ambat I, Srivastava V, Sillanpää M. Recent advancement in biodiesel production properties, and medical applications. Molecules 2015;20:16540–65. https://doi.
methodologies using various feedstock: a review. Renew Sustain Energy Rev 2018; org/10.3390/molecules200916540.
90:356–69. https://doi.org/10.1016/j.rser.2018.03.069. [38] Asmathunisha N, Kathiresan K. A review on biosynthesis of nanoparticles by
[13] Naik SN, Goud VV, Rout PK, Dalai AK. Production of first and second generation marine organisms. Colloids Surfaces B Biointerfaces 2013;103:283–7. https://doi.
biofuels: a comprehensive review. Renew Sustain Energy Rev 2010;14:578–97. org/10.1016/j.colsurfb.2012.10.030.
https://doi.org/10.1016/j.rser.2009.10.003. [39] Herlekar M, Barve S, Kumar R. Plant-mediated green synthesis of iron
[14] Maneerung T, Kawi S, Dai Y, Wang CH. Sustainable biodiesel production via nanoparticles. J Nanoparticles 2014;2014:1–9. https://doi.org/10.1155/2014/
transesterification of waste cooking oil by using CaO catalysts prepared from 140614.
chicken manure. Energy Convers Manag 2016;123:487–97. https://doi.org/ [40] Venkateswarlu S, Rao YS, Balaji T, Prathima B, Jyothi NVV. Biogenic synthesis of
10.1016/j.enconman.2016.06.071. Fe3O4 magnetic nanoparticles using plantain peel extract. Mater Lett 2013;100:
[15] Jung JM, Cho J, Kim KH, Kwon EE. Pseudo catalytic transformation of volatile fatty 241–4. https://doi.org/10.1016/j.matlet.2013.03.018.
acids into fatty acid methyl esters. Bioresour Technol 2016;203:26–31. https://doi. [41] Teo SH, Islam A, Chan ES, Thomas Choong SY, Alharthi NH, Taufiq-Yap YH, et al.
org/10.1016/j.biortech.2015.12.048. Efficient biodiesel production from Jatropha curcus using CaSO4/Fe2O3-SiO2
[16] Berchmans HJ, Hirata S. Biodiesel production from crude Jatropha curcas L. seed core-shell magnetic nanoparticles. J Clean Prod 2019;208:816–26. https://doi.org/
oil with a high content of free fatty acids. Bioresour Technol 2008;99:1716–21. 10.1016/j.jclepro.2018.10.107.
https://doi.org/10.1016/j.biortech.2007.03.051. [42] Vasantharaj S, Sathiyavimal S, Senthilkumar P, LewisOscar F, Pugazhendhi A.
[17] Vicente G, Martínez M, Aracil J. Integrated biodiesel production: a comparison of Biosynthesis of iron oxide nanoparticles using leaf extract of Ruellia tuberosa:
different homogeneous catalysts systems. Bioresour Technol 2004;92:297–305. Antimicrobial properties and their applications in photocatalytic degradation.
https://doi.org/10.1016/j.biortech.2003.08.014. J Photochem Photobiol B Biol 2019;192:74–82. https://doi.org/10.1016/j.
[18] Di Serio M, Cozzolino M, Giordano M, Tesser R, Patrono P, Santacesaria E. From jphotobiol.2018.12.025.
homogeneous to heterogeneous catalysts in biodiesel production. Ind Eng Chem [43] Farooq M, Ramli A. Biodiesel production from low FFA waste cooking oil using
Res 2007;46:6379–84. https://doi.org/10.1021/ie070663q. heterogeneous catalyst derived from chicken bones. Renew Energy 2015;76:362–8.
[19] Hara M. Environmentally benign production of biodiesel using heterogeneous https://doi.org/10.1016/j.renene.2014.11.042.
catalysts. ChemSusChem 2009;2:129–35. https://doi.org/10.1002/ [44] Rabie AM, Shaban M, Abukhadra MR, Hosny R, Ahmed SA, Negm NA. Diatomite
cssc.200800222. supported by CaO/MgO nanocomposite as heterogeneous catalyst for biodiesel
[20] Bournay L, Casanave D, Delfort B, Hillion G, Chodorge JA. New heterogeneous production from waste cooking oil. J Mol Liq 2019;279:224–31. https://doi.org/
process for biodiesel production: a way to improve the quality and the value of the 10.1016/j.molliq.2019.01.096.
crude glycerin produced by biodiesel plants. Catal Today 2005;106:190–2. https:// [45] Rajkumari K, Das D, Pathak G, Rokhum L. Waste-to-useful: a biowaste-derived
doi.org/10.1016/j.cattod.2005.07.181. heterogeneous catalyst for a green and sustainable Henry reaction. New J Chem
[21] Pathak G, Das D, Rajkumari K, Rokhum L. Exploiting waste: towards a sustainable 2019;43:2134–40. https://doi.org/10.1039/c8nj05029e.
production of biodiesel using: Musa acuminata peel ash as a heterogeneous [46] Changmai B, Laskar IB, Rokhum L. Microwave-assisted synthesis of glycerol
catalyst. Green Chem 2018;20:2365–73. https://doi.org/10.1039/c8gc00071a. carbonate by the transesterification of glycerol with dimethyl carbonate using
[22] Changmai B, Sudarsanam P, Rokhum L. Biodiesel production using a renewable Musa acuminata peel ash catalyst. J Taiwan Inst Chem Eng 2019;102:276–82.
mesoporous solid catalyst. Ind Crops Prod 2019:111911. https://doi.org/10.1016/ https://doi.org/10.1016/j.jtice.2019.06.014.
j.indcrop.2019.111911. [47] Pathak G, Rajkumari K, Rokhum L. Wealth from waste: M. acuminata peel waste-
[23] Laskar IB, Rajkumari K, Gupta R, Chatterjee S, Paul B, Rokhum L. Waste snail shell derived magnetic nanoparticles as a solid catalyst for the Henry reaction.
derived heterogeneous catalyst for biodiesel production by the transesterification Nanoscale Adv 2019;1:1013–20. https://doi.org/10.1039/c8na00321a.
of soybean oil. RSC Adv 2018;8:20131–42. https://doi.org/10.1039/c8ra02397b. [48] Liu C, Yuan J, Gao H, Liu C. Biodiesel production from waste cooking oil by
[24] Laskar IB, Rokhum L, Gupta R, Chatterjee S. Zinc oxide supported silver immobilized lipase on superparamagnetic Fe3O4 hollow sub-microspheres.
nanoparticles as a heterogeneous catalyst for production of biodiesel from palm oil. Biocatal Biotransform 2016;34:283–90. https://doi.org/10.1080/
Environ Prog Sustain Energy 2019:1–11. https://doi.org/10.1002/ep.13369. 10242422.2016.1265948.
[25] Rajkumari K, Kalita J, Das D, Rokhum L. Magnetic Fe3O4@silica sulfuric acid [49] Gohain M, Devi A, Deka D. Musa balbisiana Colla peel as highly effective
nanoparticles promoted regioselective protection/deprotection of alcohols with renewable heterogeneous base catalyst for biodiesel production. Ind Crops Prod
dihydropyran under solvent-free conditions. RSC Adv 2017;7:56559–65. https:// 2017;109:8–18. https://doi.org/10.1016/j.indcrop.2017.08.006.
doi.org/10.1039/c7ra12458a. [50] Lim BP, Maniam GP, Hamid SA. Biodiesel from adsorbed waste oil on spent
[26] Teja AS, Koh PY. Synthesis, properties, and applications of magnetic iron oxide bleaching clay using CaO as a heterogeneous catalyst. Eur J Sci Res 2009;33:
nanoparticles. Prog Cryst Growth Charact Mater 2009;55:22–45. https://doi.org/ 347–57.
10.1016/j.pcrysgrow.2008.08.003. [51] Mutreja V, Singh S, Ali A. Potassium impregnated nanocrystalline mixed oxides of
[27] Singamaneni S, Bliznyuk VN, Binek C, Tsymbal EY. Magnetic nanoparticles: recent La and Mg as heterogeneous catalysts for transesterification. Renew Energy 2014;
advances in synthesis, self-assembly and applications. J Mater Chem 2011;21: 62:226–33. https://doi.org/10.1016/j.renene.2013.07.015.
16819–45. https://doi.org/10.1039/c1jm11845e. [52] Xie W, Zhao L. Heterogeneous CaO-MoO3-SBA-15 catalysts for biodiesel
[28] Verma S, Verma D, Sinha AK, Jain SL. Palladium complex immobilized on production from soybean oil. Energy Convers Manag 2014;79:34–42. https://doi.
graphene oxide-magnetic nanoparticle composites for ester synthesis by aerobic org/10.1016/j.enconman.2013.11.041.
oxidative esterification of alcohols. Appl Catal A Gen 2015;489:17–23. https://doi. [53] Thinnakorn K, Tscheikuna J. Transesterification of palm olein using sodium
org/10.1016/j.apcata.2014.10.004. phosphate impregnated on an alumina support. Appl Catal A Gen 2014;484:
[29] Kiasat AR, Nazari S. Magnetic nanoparticles grafted with β-cyclodextrin- 122–33. https://doi.org/10.1016/j.apcata.2014.07.007.
polyurethane polymer as a novel nanomagnetic polymer brush catalyst for [54] Madhuvilakku R, Mariappan R, Jeyapal S, Sundar S, Piraman S. Transesterification
nucleophilic substitution reactions of benzyl halides in water. J Mol Catal A Chem of palm oil catalyzed by fresh water bivalve mollusk (margaritifera falcata) shell as
2012;365:80–6. https://doi.org/10.1016/j.molcata.2012.08.012. heterogeneous catalyst. Ind Eng Chem Res 2013;52:17407–13. https://doi.org/
[30] Mehrasbi MR, Mohammadi J, Peyda M, Mohammadi M. Covalent immobilization 10.1021/ie4025903.
of Candida antarctica lipase on core-shell magnetic nanoparticles for production of [55] Nath B, Das B, Kalita P, Basumatary S. Waste to value addition: utilization of waste
biodiesel from waste cooking oil. Renew Energy 2017;101:593–602. https://doi. Brassica nigra plant derived novel green heterogeneous base catalyst for effective
org/10.1016/j.renene.2016.09.022. synthesis of biodiesel. J Clean Prod 2019;239. https://doi.org/10.1016/j.
jclepro.2019.118112.

10
B. Changmai et al. Fuel 286 (2021) 119447

[56] Mendonça IM, Paes OARL, Maia PJS, Souza MP, Almeida RA, Silva CC, et al. New [59] Mahesh SE, Ramanathan A, Begum KMMS, Narayanan A. Biodiesel production
heterogeneous catalyst for biodiesel production from waste tucumã peels from waste cooking oil using KBr impregnated CaO as catalyst. Energy Convers
(Astrocaryum aculeatum Meyer): parameters optimization study. Renew Energy Manag 2015;91:442–50. https://doi.org/10.1016/j.enconman.2014.12.031.
2019;130:103–10. https://doi.org/10.1016/j.renene.2018.06.059. [60] Putra MD, Irawan C, Udiantoro, Ristianingsih Y, Nata IF. A cleaner process for
[57] Wen Z, Yu X, Tu ST, Yan J, Dahlquist E. Biodiesel production from waste cooking biodiesel production from waste cooking oil using waste materials as a
oil catalyzed by TiO2-MgO mixed oxides. Bioresour Technol 2010;101:9570–6. heterogeneous catalyst and its kinetic study. J Clean Prod 2018;195:1249–58.
https://doi.org/10.1016/j.biortech.2010.07.066. https://doi.org/10.1016/j.jclepro.2018.06.010.
[58] Farooq M, Ramli A, Subbarao D. Biodiesel production from waste cooking oil using
bifunctional heterogeneous solid catalysts. J Clean Prod 2013;59:131–40. https://
doi.org/10.1016/j.jclepro.2013.06.015.

11

You might also like