You are on page 1of 18

Catalysis Letters

https://doi.org/10.1007/s10562-019-02906-4

Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective


Biodiesel Production from Waste Cooking Oil
A. Ashok1 · L. John Kennedy1

Received: 26 September 2018 / Accepted: 21 July 2019


© Springer Science+Business Media, LLC, part of Springer Nature 2019

Abstract
ZnFe2O4 nanocatalysts synthesized by microwave combustion method is employed for biodiesel production from waste cook-
ing oil (WCO). The zinc ferrite samples are prepared by varying the microwave power from 500 to 1500 W. The nanocatalysts
are characterized by XRD, FTIR, DRS, HR-SEM and VSM techniques. Transesterification of WCO are investigated and
maximum biodiesel yield of 98.6% is achieved with 4 wt% of Z ­ nFe2O4 nanocatalyst (ZF-1500 sample), methanol/oil molar
ratio of 21:1, reaction temperature about 60 °C and reaction time 1 h. The nanocatalyst ­(ZnFe2O4) was reused at least for 10
times. The activation energy (Ea) and frequency factor (A) is calculated to be 59.4 kJ mol−1 and 1.66 × 108 min−1 respectively.
The thermodynamic parameters ∆H and ∆S were found to be 88.76 kJ mol−1 and − 0.096 kJ mol−1 K−1. The positive values
of ∆G for transesterification process is found to be non-spontaneous and endergonic.
Graphic Abstract

Keywords  Biodiesel production · Transesterification · Heterogeneous catalysis · Spinels

Extended author information available on the last page of the article

13
Vol.:(0123456789)
A. Ashok, L. J. Kennedy

Abbreviations 250/L, Rs. 300/L, Rs. 140/L, Rs. 260/L respectively. Due
WCO Waste cooking oil to the increasing human population, fast food culture, the
ZF-500 Zinc ferrite using 500 W amount of WCO generated at homes and food restaurants
ZF-600 Zinc ferrite using 600 W are increasing rapidly. Thus from economic and environ-
ZF-900 Zinc ferrite using 900 W mental point of view, WCO can be a promising biodiesel
ZF-1200 Zinc ferrite using 1200 W (Fatty Acid Methyl Esters) feedstock.
ZF-1500 Zinc ferrite using 1500 W Biodiesel is generally produced by the transesterification
XRD X-ray diffraction of vegetable oil (triglyceride) or animal fat with an alcohol
FTIR Fourier transformed infrared (mostly methanol) to form fatty acid methyl ester (FAME)
HRSEM High resolution scanning electron microscopy and glycerol in the presence of suitable catalyst. Three major
DRS Diffuse reflectance spectroscopy reaction steps involved in the transesterification are, triglyc-
VSM Vibrating sample magnetometer eride to diglyceride with one ester in step one, diglyceride
GCMS Gas chromatography mass spectroscopy to monoglyceride with another ester in step two and finally
FAME Fatty acid methyl ester monoglyceride to glycerol with one more ester in step three
[4]. Thus each step gives one ester and finally leading to
three esters (FAME). Therefore the reaction kinetics in every
1 Introduction step becomes dominant for efficient biodiesel production.
The reaction kinetics can be catalyzed either by using
Diesel is a widely used fuel in combustion engines besides homogeneous or heterogeneous catalyst and biocatalyst
petrol. Despite these being classified to be clean fuel, still otherwise [5–7]. Generally, heterogeneous catalyst are pre-
we suffer from the problem of pollution and toxic exhaust. ferred for biodiesel production, as they have high catalytic
The rise in vehicle population and increased industriali- activity sites, longer life time, easy separation, recyclable,
zation lead to emission problems and depletion of fossil selectivity and does not dissolve in methanol. Even though,
fuels. Therefore to overcome the fossil fuel depletion, one alkaline earth metal oxide like calcium oxide (CaO) are
of the alternate fuels to match the demands of mankind are reported [8] to be an effective catalyst for biodiesel pro-
the fuels from bio-ethanol, biodiesel and biomass. Among duction, they suffer from the problem of water and carbon
these, biodiesel is one of the preferred alternative fuel di-oxide absorption which further leads to separation dif-
because, it is non-toxic, biodegradable and environmental ficulty by the increase in viscosity and thereby retarding the
friendly with low emission profiles [1]. In addition, bio- reaction kinetics [7]. ZnO based heterogeneous nanocatalyst
diesel possess other technical advantages like, have high are also used for biodiesel production, and appeared to have
flash point, higher cetane number and better lubrication good catalytic activity for maximum biodiesel conversion
than petroleum fuel [2]. Although biodiesel is produced [9] but still suffered from catalyst separation in the biodiesel
from vegetable oils, the cost is much higher than the regu- process. Therefore recovery or separation of catalyst from
lar diesel. Hence research is concentrated in producing the biodiesel/glycerol mixture still remains a great challenge
biodiesel from cheaper raw materials and one among those in the biodiesel industry, especially when nanostructured
is from waste cooking oils (WCO). Moreover, producing catalyst are employed.
biodiesel from WCO solves the problem of (i) oil contami- The combustion based synthesis is one of the most acces-
nation in soils and water bodies as they are just thrown sible, fast and low energy soft method for synthesis of ferrite
or poured after repeated use and (ii) cheap feed stock for materials. Microwave assisted combustion method recently
biodiesel production from oil wastes. WCO cannot be used gained importance over the conventional method because,
again as a substitute for fresh vegetable oils because the microwave interacts with the reactant at the molecular level.
physical and chemical properties are different from those In the microwave mechanism, the electromagnetic energy is
of fresh vegetable oils. Despite the change in properties, converted to heat energy using propellant chemistry princi-
WCO is still a preferred feed stock as the cost of WCO ple. The choice of nitrate precursor as oxidizer and organic
is 2 to 3 times less than fresh vegetable oils and does not compound as a fuel at suitable stoichiometric ratios, control
affect the quality of the biodiesel produced [3]. The WCO the combustion process. The microwave energy is trans-
is obtained as a waste product from many food processing formed into heat energy by strong intermolecular friction
plants, restaurants and local fast food shops. Mostly as a and raises the temperature of the precursor materials sud-
matter of disposal, the price involves only in collection denly. As a result the crystal size, morphology, other physi-
and transportation. Therefore the WCO is comparatively cal and chemical properties are altered. This results in the
cheap than the fresh edible oils. In addition the price of formation of nanoparticles, early phase formation, different
WCO is only Rs. 35/L compared to the price of groundnut, morphologies and high efficiencies with in few minutes of
gingerly, coconut and mustard oils which cost around Rs. time [10, 11].

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

Hence in the present study, herein the attempt is to use Table 1  Physicochemical properties of waste cooking oil used in the
zinc ferrite spinel as heterogeneous catalyst to overcome the present study
problem of separation and efficient biodiesel production. Parameters Values
­ZnFe2O4 spinel type catalyst have more advantages such as,
Acid value (mg KOH/g) 0.33
better catalytic activity, allow a great cationic mobility, non-
Density (g/m3) 0.929
toxic, inexpensive, thermal and chemical stability, catalytic
Kinematic viscosity at 40 °C (­ mm2/s) 10.89
resistance and attraction by magnets by superior magnetiza-
Saponification value (mg KOH/g of oil) 193.5
tion. Therefore the focus of this study is to prepare ­ZnFe2O4
Molecular weight (g/mol) 871.0
nanoparticles by microwave combustion method and employ
Calorific value (MJ/kg) 32.98
for the transesterification of WCO to biodiesel using metha-
Cetane number 33
nol. l-arginine is found to facilitate the homogeneous mixing
Iodine value (mg KOH/g) 8.65
of cations in solution and it undergoes decomposition and
Flash point (°C) 209
promotes combustion. The literature on the preparation of
FFA (%) 0.165
zinc ferrite using l-arginine is found to be scarce. We have
Oxidation stability (h at 110 °C) 5.8
investigated the effect of this fuel to evaluate the character-
istics like surface morphology, size distribution, and mag-
netic properties of pure zinc ferrite nanoparticles [12]. The
prepared ­ZnFe2O4 nanocatalyst are characterized by X-ray ferric nitrate (0.808 g), zinc nitrate (0.297 g) in mole ratio
diffraction (XRD), Fourier transform infrared spectroscopy (Fe:Zn) 2:1 and l-Arginine (0.204 g) were dissolved, each
(FTIR), High resolution scanning electron microscopy (HR- in 10 ml of demineralized water separately and then mixed.
SEM), Diffuse reflectance spectroscopy (DRS), Vibrating The mixture is sonicated using 3  mm titanium probe at
sample magnetometer (VSM) analysis to study the physico- 25 kHz and approximately 528 W (44% of total power) using
chemical properties. Maximum biodiesel yield is optimized a probe sonicator (Esquire Biotech – EBT 1200, India) for
by varying the process parameters. Reusability, kinetic and 5 min at room temperature until a clear transparent solu-
thermodynamic parameters are also evaluated. The purity of tion is obtained. This clear solution is poured in silica cru-
synthesized biodiesel were characterized using Gas chroma- cible and placed inside and industrial microwave oven at
tography–Mass spectroscopy. 2.45 GHz with the different power values (500 W, 600 W,
900 W, 1200 W and 1500 W) for 10 min. The solution is
initially boiled and underwent spontaneous combustion with
2 Materials and Methods evolution of gases and finally resulting in the formation of a
brown colored puffy powder. The incident microwave power
2.1 Materials varied from 500 to 1500 W to study the effect of nanostruc-
ture formation and is labelled as ZF-500, ZF-600, ZF-900,
All the chemicals used, such as zinc nitrate, ferric nitrate ZF-1200 and ZF-1500 respectively. The final product is
and l-Arginine are analytical grade purchased from Merck, washed well with alcohol and dried at 110 °C for further
India and is used as received without further purification. use. During the combustion process, the chemical reaction
Waste cooking oil (WCO) is collected from a local restau- involved in the formation of zinc ferrite using l-Arginine as
rant in Chennai, India. The collected oil is subjected to pre- fuel is as follows:
treatment to remove the suspended solid materials by filtra- ( ) ( )
Zn NO3 2 6H2 O(s) + 2Fe NO3 3 9H2 O(s)
tion. The physicochemical properties of WCO are listed in
Table 1. The impurities in WCO was removed by continuous + 1.176 C6 H14 N4 O2(s) → ZnFe2 O4(s)
filtration and repeated washing with water so that the soluble + 8.235H2 O(g) ↑ + 7.058CO2(g) ↑ + 6.352N2(g) ↑
metal ion and other impurities are eliminated. In our present
study, the acid value of WCO is 0.33 mg KOH/g (Table 1)
which is less than 1 mg KOH/g oil (permissible limit) and 2.3 Characterization
hence favorable for alkali catalyzed transesterification.
Therefore the oil was directly used without any deacidifica- The microwave (MW) system operating at 2.45 GHz fre-
tion treatment for transesterification reaction [4, 13]. quency is supplied by INFURR Super heat furnaces, Chen-
nai India. The microwave oven as inner chamber area
2.2 Zinc Ferrite Nanocatalyst Preparation 12″L × 12″B × 8″H, heating achieved by two number of
magnetrons of power rating about 1200 W each for alternate
Zinc ferrite nanoparticles are prepared by microwave option automatically on a preselected time cycle, so as to
assisted combustion method. Stoichiometric amounts of operate the unit continuously for a minimum of 120 min

13
A. Ashok, L. J. Kennedy

at a stretch. Maximum temperature 1000 °C. The MW sys- ln(1 − X) = −kt (2)


tem has provisions like rotation table, exhaust through top, where, X is the conversion of triglyceride at time t.
power selection process, timer, insert Teflon bush etc. The The pseudo second order rate is given in Eq. 3 and from
structural characterization of ­ZnFe2O4 is performed using the plot of (X/1 − X) versus time (t) yielded the second order
a Rigaku (Smart Lab) X-ray diffractometer with Cu Kα rate constants.
radiation at λ = 1.5406 Å. The surface functional groups
are analyzed by Perkin Elmer FT-IR spectrometer. Sur- X
= CTG ⋅ k ⋅ t (3)
face morphologies are observed with a Joel JSM6360 high 1−X
resolution scanning electron microscope. The UV–visible where, CTG is the initial triglyceride molar concentration
diffuse reflectance (DRS) spectra of samples are recorded (mol L−1), X is the conversion of triglyceride at time t.
using Thermo scientific evolution 300 UV–visible spectro- The activation energy (Ea) required for the transesterifi-
photometer. Magnetic measurements are carried out at room cation process is calculated according to Arrhenius Eq. 4,
temperature using the PMC MicroMag3900 model vibrating ( )
sample magnetometer (VSM) equipped with 1 T magnet. k = A. exp − a
E
(4)
For TPD-NH3 and ­CO2 analysis, the zinc ferrite sample was RT
pre-treated under He flow of 10 ml/min at 400 °C. ­NH3/
where, k is the reaction rate constant (­ min−1), A is the Arrhe-
CO2 adsorption was carried out under standard condition
nius constant or frequency factor, Ea is the activation energy
by flowing of 10% ­NH3/He and 10% ­CO2/He over the fer-
(J/mol), T is the temperature and R is the universal gas con-
­ H3/
rite nanoparticle still saturation and then desorption of N
stant (8.3145 J mol−1 k−1). The slope of the graph between
CO2 by temperature-programmed analysis under constant
ln (k) versus 1/T (Eq. 5) gives the value of activation energy
He flow from room temperature to 800 °C with a heating
[15].
rate of 10 °C/min.
Ea
ln k = ln A − (5)
RT
2.4 Synthesis of Biodiesel

Transesterification reaction are carried in a 3-neck jacketed 3 Result and Discussion


type glass reactor of 500 mL capacity attached with a water
cooled reflux condenser. The reaction mixtures, methanol, 3.1 Catalyst Characterization
WCO and zinc ferrite are fed into the reactor and subjected
to constant mixing using magnetic stirrer. Transesterification 3.1.1 X‑Ray Diffraction (XRD)
reaction is carried out by varying the catalyst concentration
(1 to 8 wt%), methanol to oil molar ratio (3:1 to 30:1), tem- The X-ray diffraction patterns of zinc ferrite nanostruc-
perature (35 °C to 60 °C), and reaction time up to 80 min. tures are shown in Fig. 1. All diffraction peaks are matched
Consequently after transesterification reaction, the mixture
was left to settle and the catalyst was recovered by using
a magnet. Using a separating funnel, the biodiesel (lighter
phase) and glycerol (denser phase) was separated and the
biodiesel product is analyzed by gas chromatography–mass
spectroscopy. The percentage yield of biodiesel is calculated
using the formula [14],
total weight of biodiesel produced
%yield = × 100 (1)
total weight of oils used

2.5 Reaction Kinetics

The reaction kinetics are evaluated by considering the


pseudo first order and pseudo second order kinetic model.
The pseudo first order rate kinetics is evaluated using the
Eq. 2, where the rate constants are obtained from the slope
of plot of ln(1 − X) versus time (t). Fig. 1  XRD patterns of (a) ZF-500, (b) ZF-600, (c) ZF-900, (d)
ZF-1200, (e) ZF-1500

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

with standard JCPDS card no: 82-1042. The characteristics almost similar to the values reported by Ferrari et al. and
peaks at 2θ values of 30.16°, 35.57°, 43.19°, 53.52°, 57.01°, Shoushtari et al. [19, 20].
62.69° and 74.32° are indexed to (220), (311), (400), (422), Williamson–Hall (W–H) method gives the effective par-
(511), (440) and (533) planes of the zinc ferrite spinel type ticles size (D) and strain component using the formula given
cubic structure with Fd3m space group [16]. The average in Eq. 8,
crystallite size of synthesized ­ZnFe2O4 nanostructures from
k𝜆
the most intense diffraction peak (311) plane were estimated 𝛽 cos 𝜃 = + 4𝜀 sin 𝜃 (8)
by the Scherer formula [17], D
where, β is the full width half maximum (FWHM) of the
0.89𝜆
L= (6) X-ray diffraction peak, θ is the diffraction angle, k is a con-
𝛽cos𝜃
stant equal to 0.89, λ is the wavelength of the X-ray source
where, L, is the crystallite size; λ, is the X-ray wavelength (λ = 0.15406 nm), and ε is the strain associated with the
(0.15406 nm); β, is the full width at half maximum (FWHM) nanoparticles. From Eq. 8, W–H plot between 4sinθ versus
of diffraction peak; and 2θ, is the diffraction angle. The βcosθ (Fig. 2) gives the strain (ε) and the effective particle
average crystallite size for Z­ nFe2O4 samples are shown in size (D) from slope and intercept (k/D) respectively. The
Table 2. The crystallite size were found in the range 8.55 nm effective particle size values (Table 2) calculated from W–H
to 13.04 nm [18]. From our previous studies [18], zinc fer- method is little lesser than the average crystallite size calcu-
rite nanoparticles prepared by conventional method showed lated by using Debye–scherrer’s formula, may be due to the
a larger crystallite size around 74 nm, while the present reason that former is associated with the strain components,
study shows an average crystallite size in the range from while the later is not. A negative slope in W–H plot indi-
8.55 to 13.04 nm. Lower the size, better would be the sur- cates the presence of compressive strain in the zinc ferrite
face area and available active sites. More over the magneti-
zation properties were also better than the conventionally
prepared zinc ferrites, as it would facilitate the separation
of the catalyst using external magnet. The results shows,
varying the microwave power (Watts) from 500 to 1500 W
doesn’t seem to have any remarkable impact in controlling
­ nFe2O4. However, ZF-500 possessed
the crystallite size of Z
lesser size (8.55 nm) in comparison to the other samples.
Therefore it is suggestive that minimum microwave energy
of about 500 W is sufficient enough to synthesis small sized
zinc ferrite nanoparticles. The lattice parameter ‘a’ of zinc
ferrite spinel’s is calculated from the X-ray diffraction data
using the formula [10],

a = dhkl h2 + k2 + l2 (7)

where, a is the lattice parameter, dhkl is the interplanar spac-


ing corresponding to the miller indices, h, k, and l are the
miller indices. The calculated lattice parameters are shown
in Table 2. When increased Watts from 500 to 1500 W of Fig. 2  Williamson-Hall plots of zinc ferrite (ZF-500 to ZF-1500)
­ZnFe2O4 samples, the lattice parameter of all samples are samples

Table 2  Crystallite size, strain, lattice constant and band gap values prepared zinc ferrite nanoparticles
Sample code Average crystallite size, L (nm) Effective particle size, D (nm) by Band gap, Eg Lattice constant Strain (× 10−3)
by Scherrer’s formula the Williamson–Hall plot (eV) a (nm)

ZF-500 8.55 7.61 1.90 0.8384 − 4.0


ZF-600 13.04 8.21 1.88 0.8361 − 2.7
ZF-900 10.46 7.79 1.87 0.8358 − 3.5
ZF-1200 10.15 7.83 1.94 0.8346 − 3.3
ZF-1500 10.15 8.06 1.91 0.8304 − 3.2

13
A. Ashok, L. J. Kennedy

Table 3  Bulk density, X-ray density and porosity percentage of zinc


ferrite spinels
Sample code Molecular Bulk den- X-ray den- Porosity, P
weight, M sity, 𝜌b (g/ sity, 𝜌x (g/ (%)
(g) cm3) cm3)

ZF-500 241.07 2.17 5.441 60.12


ZF-600 241.07 2.19 5.477 60.01
ZF-900 241.07 2.15 5.484 60.80
ZF-1200 241.07 2.10 5.508 61.87
ZF-1500 241.07 2.18 5.453 60.02

samples. The X-ray density [21] and percentage porosity


(P) [22] is calculated using the formula according to Eqs. 9
and 10,
8M
𝜌x = (9)
Na3
[ ]
Bulk density
P= 1− × 100 (10)
X − ray density

where, M is the molecular weight of zinc ferrite (241.07 g/


mol), N is the Avogadro constant (N = 6.02214 × 10 23
­mol−1), a is the lattice parameter. It is observed that the
values of X-ray density, bulk density and porosity param-
eter varies minimal, despite increasing the microwave Watts
(Table 3). The bulk densities along with molecular weight
Fig. 3  FT-IR spectra of (a) ZF-500, (b) ZF-600, (c) ZF-900, (d)
calculated are listed in Table 3. The values of X-ray density
ZF-1200, (e) ZF-1500
is found to be greater than that of bulk densities which are
attributed to the presence of pores that is formed during the
microwave combustion synthesis adopted. ZF-500, well defined crystallite grains with distinct grain
boundaries with inter and intra angular pores are observed.
In addition, all the grain wall surfaces are not smooth or
3.1.2 Fourier Transform Infrared (FTIR) plain but are made of tiny spherical particles. Thus a unit
area of grain surfaces are made of numerous spherical parti-
The FT-IR spectra of zinc ferrite samples are shown in cles, which would lead to inherit high surface area followed
Fig. 3. The broad band at 3414 cm−1 exhibit the O–H group by high catalytic activities. However for ZF-600, ZF-900,
stretching vibrations of water molecule [23]. The band at ZF-1200 and ZF-1500, the grains are broken down and the
2357 cm−1 indicates the presence of H–O–H bending vibra- spherical particles on the grain surfaces get fused, as the
tion of absorbed water [12]. The bending modes of absorbed microwave power is increased. This effect is well observed
water is assigned to the band at around 1635 cm−1 [23]. The in ZF-1500 sample.
sharp band at around 1377 cm−1 signifies adsorbed moisture
in the system of FeOOH and adsorbed nitrate ions [24]. The 3.1.4 Energy Dispersive X‑Ray (EDX) Analysis
characteristic bands at 419 cm−1 and 550 cm−1 confirms
the intrinsic stretching vibrations of tetrahedral (­ Zn2+–O2−) The prepared zinc ferrite nanoparticles were composed of
bond and octahedral ­(Fe3+–O2−) bond and confirms the for- Zn, Fe and O confirmed by EDX analysis at room tempera-
mation of ­ZnFe2O4 spinel structure [18]. ture as show in Fig. 5. The variation in the values shall be
due to the presence of defects and oxygen vacancies. Gen-
3.1.3 High Resolution Scanning Electron Microscope erally, HR-SEM and EDX analysis was performed using
(HR‑SEM) thin layer of gold particle on the surface of the zinc ferrite
nanoparticles. The gold (Au) peak at 2.1 to 2.2 keV for all
The surface morphology of synthesized zinc ferrite nano- the samples, is due to the coating of gold on the surface of
structures observed by the HR-SEM are shown in Fig. 4. For the zinc ferrite sample during HR-SEM and EDX analysis.

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

Fig. 4  HR-SEM images of a ZF-500, b ZF-600, c ZF-900, d ZF-1200, e ZF-1500

13
A. Ashok, L. J. Kennedy

Fig. 5  Energy dispersive X-ray analysis spectra of a ZnF-500, b ZnF-600, c ZnF-900, d ZnF-1200, e ZnF-1500

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

3.1.5 Diffuse Reflectance Spectroscopy (DRS) 3.2 Acid Base Properties of Zinc Ferrite


Nanoparticles by Temperature Programmed
The band gap, Eg of the zinc ferrite nanoparticles was Desorption Studies (TPD‑NH3/CO2)
evaluated from the diffuse reflectance spectra using the
Kubelka–Munk function (Eq. 11) [25] in the Tauc relation The TPD analysis of pure zinc ferrite nanocatalsyt (ZF-
as in Eq. 12, 1500) using N ­ H3 and C­ O2 as probe gases are shown in
Fig. 8. Generally, the peak temperature and peak area can
(1 − R)2
𝛼 = F(R) = (11) be used to estimate the acid strength and acidity, respectively
2R (Fig. 8a). For ferrite samples, the ammonia desorbed from
where, R is the diffuse reflectance and F(R) is the room temperature to 200 °C as treated as Bronsted (Weak
Kubelka–Munk function. sites), medium acid sites for 201 °C to 500 °C, and strong
acid sites (Lewis sites) for 501 °C to 800 °C. Figure 8a
F(R)h𝜈 = A(h𝜈 − Eg )n (12) shows clearly the zinc ferrite sample possess both weak and
where, n = 1/2 and 2 represents direct and indirect transi- strong acid sites. The total amount of acidity was found to
tions, thereby giving direct and indirect band gaps. A graph be 0.010 mmol/g.
(Tauc plot) is plotted between (F(R)hʋ)2 versus hʋ (eV) The surface basicity exhibited by zinc ferrite nanopar-
and the extrapolation of the linear region of plots, until the ticle was studied and characterized by TPD-CO2 as shown
intersection with x-axis gives the band gap energy values as in Fig. 8b. The desorption of zinc ferrite gives two intense
shown in Fig. 6. The calculated band gap ranges from 1.87 peaks were observed between 30 and 150  °C treated as
to 1.94 eV (Table 4) and are closer to the previous reported weak basic sites and > 500 °C treated as strong basic sites
value, 2.08 eV [18]. The Eg values obtained in our study respectively. The zinc ferrite total basicity was found to be
is less than the bulk zinc ferrite nanoparticles and can be 0.014 mmol/g. From the TPD–NH3/CO2 result shows the
attributed to the quantum confinement phenomenon taking zinc ferrite nanocatalyst possesses both strong acid and
place at the nano regime. strong basic sites which is main advantages for catalytic
reactions.
3.1.6 Vibrating Sample Magnetometer (VSM)
3.3 Reactions Variables Effect
The magnetic properties of the zinc ferrite nanoparticles
were recorded at room temperature using vibrating sample 3.3.1 Catalyst Comparison on the Transesterification
magnetometer (VSM) with the applied field ranging from of WCO
− 15 kOe to + 15 kOe (Fig. 7). The magnetic moment (μB)
[26] is calculated from the Eq. 13 All the zinc ferrite nanocatalysts synthesized at different
microwave power are used in the transesterification process
M × MS
𝜇B = (13) for optimization. In this process the zinc ferrite catalysts are
5585
compared under the following reaction conditions: 4 wt%
where, Ms is the saturation magnetization, M is the molecu- of catalyst amount, 21:1 methanol to oil molar ratio, 60 °C
lar weight of zinc ferrite and 5585 is the magnetic factor. reaction temperature and 1  h reaction time as shown in
From the hysteresis curves, the magnetic properties like sat- Fig. 9. For a catalyst loading at 4 wt% for ZF-500, ZF-600,
uration magnetization (Ms), coercivity (Hc), remanence mag- ZF-900, ZF-1200 and ZF-1500 the biodiesel yield were
netization (Mr) and magnetic moment (μB) are calculated 97%, 97.2%, 97.7%, 98.1% and 98.6% respectively. The bio-
and are reported in Table 4. The magnetization of all the diesel yield values are invariably close to the same value,
samples are in the range 16.04 emu/g to 61.46 emu/g, cor- despite varying the incident microwave power for catalyst
responding to the results obtained by Wiriya et al. [17]. The preparation.
M–H plot of magnetic hysteresis confirms, all the samples
have ferromagnetic performance with soft magnetic nature. 3.3.2 Ferrites Amount Effect
Figure 7 shows that all the samples possessed a satisfactory
magnetic responsiveness with an external magnetic field During the transesterifiction reaction, one the most impor-
which can be effective during the magnetic separation of the tant factors that affect the biodiesel yield is catalyst loading
catalyst. In comparison to the other samples, ZF-1500 sam- and therefore, it needs to be optimized. Figure 10a shows the
ple have high magnetization (61.46 emu/g) and high mag- loading of zinc ferrite samples, ranging from 1 to 8 wt% pre-
netic moment (2.6528) and hence can be easily separated in pared between 500 and 1500 W at identified reaction condi-
quicker time from the reaction mixture when employed for tions like methanol to oil molar ratio 21:1, reaction tempera-
any applications. ture 60 °C and 1 h reaction time. The result showed that the

13
A. Ashok, L. J. Kennedy

Fig. 6  (F(R)hʋ)2 versus hʋ plot of a ZF-500, b ZF-600, c ZF-900, d ZF-1200, e ZF-1500

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

Table 4  Magnetic properties (coercivity, remanent magnetization,


saturation magnetization and magnetic moment) of zinc ferrite nano-
particles

Sample code Coerciv- Rema- Magnetiza- Magnetic


ity (Hc) nence (Mr) tion (Ms) moment
(Oe) (emu/g) (emu/g) (μB)

ZF-500 410 5.69 16.04 0.692


ZF-600 400 18.77 47.26 2.039
ZF-900 383 20.73 57.56 2.484
ZF-1200 404 11.78 28.78 1.242
ZF-1500 408 21.96 61.46 2.652

Fig. 7  Magnetic hysteresis loops of (a) ZF-500, (b) ZF-600, (c)


ZF-900, (d) ZF-1200, (e) ZF-1500

transesterification strongly depends on the catalyst amount Fig. 8  TPD-NH3/CO2 profile for zinc ferrite (ZF-1500) nanoparticles
with maximum biodiesel yield of 98.6% at 4 wt% using
ZF-1500 catalyst. While increasing the catalyst amount from
1 to 4 wt%, the initial reagents get adsorbed on the catalysts and Mazaheri et al. reported similar trends indicating that
active centre effectively, leading to the formation of reaction the catalyst concentration above a threshold value can have
products. Also increase in catalyst concentration, leads to negative impact during mixing in the transesterification reac-
higher exposure of the catalyst surface area which in turn, tion [28, 29]. Therefore, 4 wt% of the catalyst amount is
increases the product yield in the reaction mixture. Catalyst optimized for transesterification process.
loading beyond 4 wt% however decreased the biodiesel yield
for all the samples. This could be due to the difficulty in 3.3.3 Methanol‑to‑Oil Molar Ratio Effect
high viscous mixing of liquid reactants at high solid catalyst
concentration [15]. Moreover, an excess amount of catalyst Generally, 3 mol of alcohol and 1 mol of triglyceride react
would lead to the saponification of biodiesel and also occa- together producing 3 mol of alkyl ester and 1 mol of glyc-
sionally, biodiesel may get adsorbed on the catalyst surface erol. An excess of alcohol is required to shift the equi-
which would be the possible reason for decreased yield of librium towards the formation of biodiesel (alkyl esters)
biodiesel during the reaction of transesterification process. because the transesterification of triglyceride is a revers-
Additionally, some amount of the catalyst would remain ible reaction [30]. Figure 10b shows the ratio of metha-
unused due to more mass transfer resistance and poor mix- nol to oil on the yield of biodiesel for all the zinc ferrite
ing, resulting in low biodiesel yield [27]. Deshmane et al. nanocatalyst. In the present study, in order to shift the

13
A. Ashok, L. J. Kennedy

3.3.5 Reaction Temperature Effect

The reaction temperature are varied from 35 to 60 °C for all


samples (1 h duration) using 21:1 methanol to oil molar ratio
and 4 wt% of catalyst as shown in Fig. 10d. It is observed
that maximum biodiesel yield of about 98.6% is reached at
60 °C. The highest biodiesel yield at 60 °C can be attributed
to accelerated chemical reaction by increasing the collision
among the reactant molecules and thus enhancing the mis-
cibility and mass transfer [32]. As the methanol ebullition
temperature is 64.7 °C, the experiments were not conducted
at T > 60 °C. Moreover as the reaction takes place into the
liquid phase, when the temperature exceeds the ebullition
temperature of the methanol, the concentration of the metha-
nol in the liquid phase decreases leading in a decrease of the
reaction yield. The condenser avoid the evaporation of the
methanol into the atmosphere and allows its condensation.
Fig. 9  Catalyst comparison for the yield of biodiesel (methanol/oil Nevertheless, the concentration in methanol into the liquid
molar ratio 21:1, reaction temperature 60 °C, reaction time 1 h) phase is not constant. Above 60 °C, no methanol remains
into the liquid phase, all methanol is under gas form. Hence
equilibrium of the reaction to get maximum FAME, the the temperature studies were restricted between 35 and
methanol to oil molar ratio is varied in different propor- 60 °C.
tions like 3:1, 6:1, 9:1, 12:1, 15:1, 18:1, 21:1, 24:1, 27:1
and 30:1. The maximum biodiesel yield, 98.6% is achieved
at methanol to oil molar ratio of 21:1 under the reaction 3.3.6 Nanocatalyst Recovery and Reuse
conditions catalyst concentration 4 wt%, reaction tempera-
ture 60 °C and 1 h reaction time. Beyond this ratio and up The reusability of the optimized, zinc ferrite (ZF-1500)
to 30:1, the yield of the biodiesel declined. A least ratio nanocatalyst is tested for 10 cycles of transesterification
of triglyceride to catalyst mass and subsequent increase in process of WCO under optimized reaction parameters like
contact time for the triglyceride with the surface shall be reaction temperature of 60 °C, reaction time 60 min, metha-
reason for it. Excess of methanol favors triglycerides to nol to oil molar ratio of 21:1 and catalyst loading of 4 wt%.
monoglycerides conversion and in turn, would facilitate The biodiesel yield after 10th cycle decreased to 93.6% from
the solubility of glycerol in FAME resulting in glycerolize initial 98.6% (Fig. 11). This marginal decrease show that the
of FAME and finally leading to a low product yield [31]. activity and the stability of the catalyst is not much affected.
This reduction in the yield shall be due to mass loss of the
catalyst by repeated washing after magnetic separation. The
3.3.4 Reaction Time Effect recovered catalyst is generally washed with methanol to
remove any organic compounds like unreacted oil, biodiesel
The experiment are carried out to investigate the perfor- and glycerol, followed by drying at 110 °C for 1 h before
mance of zinc ferrite catalyst with the reaction time vary- next use. The structural composition of the catalyst before
ing from 10 to 80 min, at the optimal reaction conditions and after 10 runs were evaluated using X-ray diffraction
like methanol to oil molar ratio 21:1 and reaction tem- studies as shown in Fig. 12. For both the samples, the peak
perature 60 °C with 4 wt% catalyst loading as shown in corresponding to 2θ values remain to be same. However
Fig. 10c, increasing the reaction time has influence on the distortion or noisy like pattern is observed in the recovered
biodiesel yield is produced. The conversion percentage is catalyst shall be due to the deposition of oil, biodiesel and
increased as the reaction time is increased to 60 min but other organic matters on the surface of the catalyst. The lat-
no significant change in biodiesel yield is observed beyond tice parameters d-spacing for the fresh and the recovered cat-
60 min until 80 min. The surface energy and the catalytic alyst is almost same with values 0.2504 nm and 0.2530 nm
sites available on the zinc ferrite catalysts are sufficient respectively. The leaching test were also performed in the
enough to complete the transesterification process with a biodiesel by XRF analysis. The results indicate that there
stipulated time of 60 min favoring a maximum biodiesel are Zn and Fe ions were leached. This may be due to ineffi-
yield. cient filtering process of the catalyst after transesterification
reaction.

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

Fig. 10  Effect of process variables on transesterification of WCO for tion time; c reaction time. Catalyst concentration 4 wt%, methanol to
a catalyst concentration. Methanol to oil molar ratio 21:1, reaction oil molar ratio 21:1, reaction temperature 60 °C; d reaction tempera-
temperature 60 °C, 1 h reaction time; b methanol to oil molar ratio. ture. Methanol to oil molar ratio 21:1, catalyst concentration 4 wt%,
Catalyst concentration 4  wt%, reaction temperature 60  °C, 1  h reac- 1 h reaction time

3.3.7 Reaction Kinetics shown). Thus it is concluded that the transesterification pro-


cess of WCO using ZF-1500 follows the first order kinetic
Kinetic study for the transesterification process is deter- model. The obtained rate constants, ­k1 were 0.0132 min−1,
mined under optimal conditions for ZF-1500 at six different 0.0205 min−1, 0.0256 min−1, 0.0431 min−1, 0.0587 min−1
temperatures (35 °C, 40 °C, 45 °C, 50 °C, 55 °C and 60 °C). and 0.0731 min−1 for reactions carried out at 35 °C, 40 °C,
The overall rate constant calculated from this study gave an 45 °C, 50 °C, 55 °C and 60 °C respectively. The reaction
insight to the transesterification of WCO. Using the plots rate rises with rise in temperature with a comparatively high
of the pseudo first order and second order equation, given value at 60 °C.
in Eqs. 2 and 3, the best fitting kinetic model is determined The activation energy is calculated from the plot of ln (k)
from the regression value coefficient R2. It is found that the versus (1/T) as shown in Fig. 14. The microwave power add
R2 value of the first order reaction plot − ln(1 − X) versus an influence on the catalytic performance that was reflected in
time (Fig. 13) is greater than 0.98 which is much higher the activation energy on the samples with increase in micro-
than the R2 value of the pseudo second order plot (figure not wave power during sample synthesis from 500 to 1500 W the

13
A. Ashok, L. J. Kennedy

Fig. 11  Effect of reusability on ZF-1500 catalyst. Reaction


condition:methanol/oil molar ratio 21:1, reaction time 1  h, reaction
temperature 60 °C

activation energy (Ea) varied from 49 to 59.40 kJ/mol. The


highest activation energy was achieved for ZF-1500 (59.40 kJ/
mol) indicating the limitation mass transfer or diffusion pro-
cess while the reaction is chemically controlled. The activa-
tion energy for transesterification of WCO using ZF-1500 was
found to be 59.40 kJ/mol and the value of A (frequency factor)
was 1.66 × 108 ­min−1 [33]. The Ea value obtained in the pre-
sent study is in agreement with other reported value [34–36],
generally lying between 33.6 and 84 kJ/mol.

3.3.8 Thermodynamic Analysis ­ nFe2O4 catalyst a before and b


Fig. 12  X-ray diffraction pattern of Z
after transesterification reaction
The most important thermodynamic parameters to be
determined for the transesterification reactions are ∆H
(enthalpy), ∆S (entropy) and ∆G (Gibbs free energy). From
the Eyring–Polanyi equation expressed in Eq. 14, Gibbs free
energy value can be calculated,
K T (
ΔG
)
k = 𝜅 B exp − (14)
h RT
where, 𝜅 (equal to one) is the transmission coefficient, kB
is the Boltzmann constant (1.38 × 10−23 J K−1), and h is the
Planck’s constant (6.63 × 10−34 J s). The general Gibbs free
energy equation can be written in Eq. 15
ΔG = ΔH − TΔS (15)
Taking the logarithm of Eq. 14 and substituting the Gibbs
free energy Eq. 15 gives,
( ) [ ( )
kB
]
k −ΔH 1 ΔS
( )
ln = + ln 𝜅 + ln + (16)
T R T h R
Fig. 13  Determination of kinetic rate constants on the basis of the
pseudo first-order kinetic model

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

to be 56.80 kJ mol−1 and − 0.096 kJ mol−1 K−1, respectively.


The kinetic and thermodynamics values are listed in Table 5.
The positive values of ∆H in our present study indicates,
that the transesterification reaction is an endothermic pro-
cess. The value of ΔS (entropy) is an important thermo-
dynamic property suggesting the degree of disorder of the
system. The negative value of ΔS in the transesterification
process indicates that either the ordered geometry degree or
the transition state alignment is higher than reactants in the
ground state. The Gibbs free energy gives an idea on, either
spontaneity or non-spontaneity of the reaction depending on
the negative or positive values of ∆G. The positive values of
∆G obtained indicates the nonspontaneous and endergonic
nature of transesterification process.

3.3.9 Physicochemical Properties of Synthesized Biodiesel


Fig. 14  Arrhenius graph for estimation of activation energy
The physicochemical properties of synthesized biodiesel
from WCO are determined and compared with previous
reported literatures and ASTM D-6751 standards as shown
in Table 6. The obtained results matching with the standards
suggest that zinc ferrite catalyst had good catalytic activity
for the large scale biodiesel production from WCO.

3.3.10 Characterization of Prepared Biodiesel

The fatty acid methyl ester composition are analysed by


GC–MS technique with the help of NIST library. Figure 16
shows the peaks corresponding to the presence of methyl
esters. The maximum peaks at retention time at 18.5 indi-
cates the presence of 9-octa decanoic acid methyl ester. The
fatty acid methyl esters corresponding to different retention
times are listed in Table 7.

3.3.11 Mechanism of Zinc Ferrite Catalyzed Reaction

Fig. 15  Plot of ln k versus 1/T ­(K−1) for biodiesel production with


The reaction mechanism of ­ZnFe2O4 catalyzed transesteri-
optimal zinc ferrite (ZF-1500) nanocatalyst for WCO
fication are proposed as in Fig. 17. The active site of the
prepared samples may have different functions co-existing
Equation 16 shows the interrelationship between enthalpy, on the surface, and shall be composed of surface ensem-
entropy and rate constant. The slope and intercept of the plot ­ n+/Mn+1 ion
bles of atoms to permit the reaction to occur. M
of 1/T versus ln (k/T), yields the thermodynamic parameters. pairs (M = Zn, Fe) and acid–base sites in Z­ nFe2O4 behaves
From Fig. 15, the enthalpy and entropy values are calculated as the active sites promoting a structure sensitive catalytic

Table 5  Kinetics and T (K) ∆G (kJ mol−1) ∆H (kJ mol−1) ∆S (kJ mol−1 k−1) Ea (kJ mol−1) A ­(min−1)


thermodynamic parameters
involved in biodiesel production 308 86.36 56.80 − 0.096 59.40 1.66 × 108
313 86.84
318 87.32
323 87.80
328 88.28
333 88.76

13
A. Ashok, L. J. Kennedy

Table 6  Fuel properties of Property ASTM specification for standard Measured value of


biodiesel produced from WCO biodiesel (D-6751) biodiesel in present
work

Kinematic viscosity at 40 °C (­ mm2/s) 1.9 to 6.0 3.79


Density (g/cm3) 0.860 to 0.894 0.836
Flash point (°C) 120 minimum 176
Cetane number 47 minimum 46
Fire point (°C) – 210
Calorific value (MJ/kg) – 43.29
Iodine value (mg KOH/g) – 0.19
Acid number (mg KOH/g) 0.5 maximum 0.03
Oxidation stability (h at 110 °C) – 3

Fig. 16  GC-MS chromatogram for biodiesel produced using prepared zinc ferrite (ZF-1500) catalyst

reactions [37]. Initially, the reaction between the catalyst the surface of the catalyst to form diglyceride in the active
and methanol produces methoxide ion, which is attached site, and the surface of the catalyst is regenerated at the same
to the surface of the catalyst and the carbonyl carbon of time. Later, the methoxide anion attacks another carbonyl
triglyceride molecule to forms a tetrahedral intermediate carbon atom in diglyceride, forming another mole of FAME
[34]. In the next step, rearrangement of the unstable tetra- and monoglyceride. These three steps are repeated until the
hedral intermediate occurs, which is later broken down to three triglyceride carbonyl centers are attacked by the meth-
form diglyceride anion and fatty acid methyl ester (FAME). oxide ions to give three moles of methyl esters and one mole
Then, the diglyceride anion is stabilized by a proton from of glycerol [35]. It is clearly depicted in Fig. 17.

13
Magnetically Separable Zinc Ferrite Nanocatalyst for an Effective Biodiesel Production from…

Table 7  Fatty acid methyl esters composition of biodiesel synthesized energy is found to be non-spontaneous. The mechanism
form WCO for the formation of fatty acid methyl esters are discussed.
Peak no Retention Fatty acid methyl ester
time (min) Acknowledgements  All the authors sincerely thank VIT Chennai Cam-
pus, Chennai, Tamil Nadu, India, for providing financial assistance
1 15.03 MethylZ-11-tetradecenoate through research associateship to the first author.
2 16.87 Hexadecanoic acid, methyl ester
3 17.7 Pentadecanoic acid, methyl ester Compliance with Ethical Standards 
4 18.5 9-Octadecenoic acid, methyl ester
5 18.73 Heptadecanoic acid, methyl ester Conflict of interest  The authors declare no conflict of interest.
6 19.23 Oleic acid
7 21.82 Hexadecanoic acid, butyl ester
8 23.27 Methoxyacetic acid, octadecyl ester
9 23.85 9-Octadecenoic acid (Z); 2-butoxyethyl ester References
1. Xiang Y, Xiang Y, Wang L (2017) Microwave radiation improves
biodiesel yields from waste cooking oil in the presence of modi-
fied coal fly ash. J Taibah Univ Sci 11:1019–1029
2. Silitonga AS, Masjuki HH, Mahlia TMI et al (2013) Experimental
study on performance and exhaust emissions of a diesel engine
fuelled with Ceiba pentandra biodiesel blends. Energy Conversat
Manag 76:828–836
3. Yaakob Z, Mohammad M, Alherbawi M et al (2013) Overview of
the production of biodiesel from waste cooking oil. Renew Sustain
Energy Rev 18:184–193
4. Pukale DD, Maddikeri GL, Gogate PR et al (2015) Ultrasound
assisted transesterification of waste cooking oil using heterogene-
ous solid catalyst. Ultrason Sonochem 22:278–286
5. Thitsartarn W, Maneerung T, Kawi S (2015) Highly active and
Fig. 17  Mechanism of Z
­ nFe2O4-catalyszed transesterification durable Ca-doped Ce-SBA-15 catalyst for biodiesel production.
Energy 89:946–956
6. Torres-Rodríguez DA, Romero-Ibarra IC, Ibarra IA et al (2016)
Biodiesel production from soybean and jatropha oils using cesium
4 Conclusions impregnated sodium zirconate as a heterogeneous base catalyst.
Renew Energy 93:323–331
The zinc ferrite nanocatalysts are successfully prepared 7. Duarte JG, Leone-Ignacio K, Da Silva JAC et al (2016) Rapid
by microwave combustion method. X-ray diffraction study determination of the synthetic activity of lipases/esterases
via transesterification and esterification zymography. Fuel
revealed the formation of cubic spinel structure with Fd3m 177:123–129
space group. FTIR spectra confirm the formation of spi- 8. Thanh LT, Okitsu K, Van Boi L et al (2012) Catalytic technologies
nel structure by the appearance of bands at 550 cm−1 and for biodiesel fuel production and utilization of glycerol: a review.
419  cm −1 corresponding to octahedral and tetrahedral Catalysts 2:191–222
9. Varghese R, Henry JP, Irudayaraj J (2017) Ultrasonication-assisted
bonds respectively. The direct band gap energy of zinc transesterification for biodiesel production by using heterogeneous
ferrite nanocatalysts ware in the range, 1.87 eV to 1.94 eV. ZnO nanocatalyst. Env Program Sustain Energy 37:1176–1182
The magnetic study confirm the ferromagnetic in nature of 10. Anand GT, Kennedy LJ, Vijaya JJ et al (2014) Structural, optical
zinc ferrite nanocatalyst. A maximum yield of biodiesel and magnetic characterization of Z ­ n1−xNixAl2O4 (0 ≤ x ≤ 5) spinel
nanostructures synthesized by microwave combustion technique.
is obtained at the optimized conditions of 60 °C for 1 h Ceram Int 41:603–615
with molar ratio of methanol to oil is 21:1 and 4  wt% 11. Theophil Anand G, John Kennedy L, Aruldoss U et al (2015)
of catalyst. For ZF-500, ZF-600, ZF-900, ZF-1200 and Structural, optical and magnetic properties of ­Zn1−xMnxAl2O4
ZF-1500 the biodiesel yield is nearly the same with 97%, (0 ≤ x ≤ 0.5) spinel nanostructures by one-pot microwave com-
bustion technique. J Mol Struct 1084:244–253
97.2%, 97.7%, 98.1% and 98.6% respectively. ZF-1500 12. Sundararajan M, Kennedy LJ, Aruldoss U et al (2015) Microwave
catalyst that exhibited catalytic activity was stable even combustion synthesis of zinc substituted nanocrystalline spinel
after reused for at least 10 times. The transesterification of cobalt ferrite: structural and magnetic studies. Mater Sci Semi-
WCO followed psesudo first order rate kinetics with maxi- cond Process 40:1–10
13. Ding J, Xia Z, Lu J (2012) Esterification and deacidification of a
mum rate constant, k­ 1 = 0.0731 min−1. The positive values waste cooking oil (TAN 68.81 mg KOH/g) for biodiesel produc-
of ∆H in our present study indicates, that the transesterifi- tion. Energies 5:2683–2691
cation reaction is an endothermic process while Gibbs free 14. Umdu ES, Tuncer M, Seker E (2009) Transesterification of
Nannochloropsis oculata microalga’s lipid to biodiesel on

13
A. Ashok, L. J. Kennedy

­Al2O3 supported CaO and MgO catalysts. Bioresour Technol 26. Prabhakaran T, Hemalatha J (2014) Chemical control on the size
100:2828–2831 and properties of nano N ­ iFe2O4 synthesized by sol-gel autocom-
15. Baskar G, Gurugulladevi A, Nishanthini T et al (2017) Opti- bustion method. Ceram Int 40:3315–3324
mization and kinetics of biodiesel production from Mahua oil 27. Chen GY, Shan R, Yan BB et al (2016) Remarkably enhancing
using manganese doped zinc oxide nanocatalyst. Renew Energy the biodiesel yield from palm oil upon abalone shell-derived CaO
103:641–646 catalysts treated by ethanol. Fuel Process Technol 143:110–117
16. Mueller F, Bresser D, Paillard E et al (2013) Influence of the 28. Mazaheri H, Ong HC, Masjuki HH et al (2018) Rice bran oil
carbonaceous conductive network on the electrochemical perfor- based biodiesel production using calcium oxide catalyst derived
mance of ­ZnFe2O4 nanoparticles. J Power Sources 236:87–94 from Chicoreus brunneus shell. Energy 144:10–19
17. Wiriya N, Bootchanont A, Maensiri S et al (2014) Magnetic prop- 29. Deshmane VG, Adewuyi YG (2013) Synthesis and kinetics of
erties of Z
­ n1−xMnxFe2O4 nanoparticles prepared by hydrothermal biodiesel formation via calcium methoxide base catalyzed trans-
method. Microelectron Eng 126:1–8 esterification reaction in the absence and presence of ultrasound.
18. Kombaiah K, Vijaya JJ, Kennedy LJ et al (2017) Optical, magnetic Fuel 107:474–482
and structural properties of Z­ nFe2O4 nanoparticles synthesized 30. Tang Y, Ren H, Chang F et al (2017) Nano KF/Al2O3 particles
by conventional and microwave assisted combustion method: a as an efficient catalyst for no-glycerol biodiesel production by
comparative investigation. Optik (Stuttg) 129:57–68 coupling transesterification. RSC Adv 7:5694–5700
19. Ferrari S, Kumar RS, Grinblat F et al (2016) In-situ high-pressure 31. Balakrishnan K, Olutoye MA, Hameed BH (2013) Synthesis of
X-ray diffraction study of zinc ferrite nanoparticles. Solid State methyl esters from waste cooking oil using construction waste
Sci 56:68–72 material as solid base catalyst. Bioresour Technol 128:788–791
20. Shoushtari MZ, Emami A, Ghahfarokhi SEM (2016) Effect of 32. Farooq M, Ramli A, Subbarao D (2013) Biodiesel production
bismuth doping on the structural and magnetic properties of from waste cooking oil using bifunctional heterogeneous solid
zinc-ferrite nanoparticles prepared by a microwave combustion catalysts. J Clean Prod 59:131–140
method. J Magn Magn Mater 419:572–579 33. Mazubert A, Taylor C, Aubin J et al (2014) Key role of tempera-
21. Iqubal MA, Sharma R, Kamaluddin (2015) Studies on interac- ture monitoring in interpretation of microwave effect on transes-
tion of ribonucleotides with zinc ferrite nanoparticles using terification and esterification reactions for biodiesel production.
spectroscopic and microscopic techniques. Karbala Int J Mod Sci Bioresour Technol 161:270–279
1:49–59 34. Feyzi M, Shahbazi Z (2017) Preparation, kinetic and thermody-
22. Vinosha PA, Mely LA, Jeronsia JE et al (2017) Synthesis and namic studies of Al–Sr nanocatalysts for biodiesel production. J
properties of spinel ­ZnFe2O4 nanoparticles by facile co-precipi- Taiwan Inst Chem Eng 71:145–155
tation route. Optik (Stuttg) 134:99–108 35. Leung DYC, Wu X, Leung MKH (2010) A review on biodiesel
23. López F, López-Delgado A, Martı́n de Vidales J et al (1998) production using catalyzed transesterification. Appl Energy
Synthesis of nanocrystalline zinc ferrite powders from sulphuric 87:1083–1095
pickling waste water. J Alloys Compd 265:291–296 36. Endalew AK, Kiros Y, Zanzi R (2011) Inorganic heterogeneous
24. Habibi MH, Parhizkar HJ (2014) FTIR and UV-vis diffuse reflec- catalysts for biodiesel production from vegetable oils. Biomass
tance spectroscopy studies of the wet chemical (WC) route synthe- Bioenerg 35:3787–3809
sized nano-structure ­CoFe2O4 from ­CoCl2 and ­FeCl3. Spectrochim 37. Scurrell MS (1973) Heterogeneous catalysis on metal oxides.
Acta Part A Mol Biomol Spectrosc 127:102–106 Annu Reports Prog Chem Sect A Phys Inorg Chem 70:87–122
25. Sundararajan M, Kennedy LJ, Vijaya JJ et al (2015) Microwave
combustion synthesis of C ­ o1−xZnxFe2O4 (0 ≤ x ≤ 0.5): structural, Publisher’s Note Springer Nature remains neutral with regard to
magnetic, optical and vibrational spectroscopic studies. Spectro- jurisdictional claims in published maps and institutional affiliations.
chim Acta Part A Mol Biomol Spectrosc 140:421–430

Affiliations

A. Ashok1 · L. John Kennedy1

1
* L. John Kennedy Materials Division, School of Advanced Sciences,
l.johnkennedy@vit.ac.in; jklsac14@yahoomail.co.in Vellore Institute of Technology (VIT) Chennai Campus,
Chennai 600 127, India

13

You might also like