Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Results in Engineering 16 (2022) 100645

Contents lists available at ScienceDirect

Results in Engineering
journal homepage: www.sciencedirect.com/journal/results-in-engineering

A green process synthesis of bio-composite heterogeneous catalyst for the


transesterification of linseed-marula bi-oil methyl ester
A.O. Etim a, P. Musonge a, b, A.C. Eloka-Eboka c, *
a
Institute of Systems Science, Durban University of Technology, Durban, South Africa
b
Faculty of Engineering, Mangosuthu University of Technology, Umlazi, South Africa
c
Centre of Excellence in Carbon-based Fuels, School of Chemical and Minerals Engineering, North-West University, Potchefstroom, South Africa

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

Keywords: Waste biomass materials contribute immensely in the environmentally friendly and cost-effective clean energy
Biomass waste production. In this study, a bio-composite catalyst was developed from waste biogenic materials (waste chicken
Biodiesel eggshells and banana peels) and was used in the transesterification of linseed–marula bi-oil methyl ester. The
Bi-oil
synthesized bio-composite catalyst was characterized to determine its properties. The catalytic potential of the
Heterogeneous catalyst
Transesterification
bio-composite catalyst was tested in the transesterification of linseed-marula bi-oil for effectiveness. Trans­
Taguchi design esterification process was optimized using Taguchi design (TD) of the response surface methodology (RSM),
which was used to generate 9 experimental conditions for the investigation of the effect of methanol to oil molar
ratio, catalyst loading, reaction time and temperature. Result of the characterization shows that Ca and K are the
dominant elements present in the synthesized bio-composite catalyst. Optimum experimental condition for the
transesterification of the bi-oil was established at methanol-to-oil ratio of 15:1, catalyst amount of 3.5 wt%,
reaction time of 60 min and reaction temperature of 65 ◦ C with 95.03% of biodiesel yield. The quality of the
biodiesel produced was within the ASTM D6751, EN 14,214 and SANS 833 standards. This study shows that the
combined effect of biogenic waste materials and linseed-marula bi-oil is a potential factor in an effective and
sustainable feedstock source(es) for biodiesel production.

biodiesel production. Catalysts for biodiesel can also be produced using


locally available materials and waste resources. This can in turn reduce
1. Introduction
the waste disposal inconvenience as well as production cost of biodiesel.
Feedstock blends or mixtures are majorly designed to address the
Globally, scientific contributions in the exploration of sustainable
challenges of feedstock scarcity for biodiesel production. The fact that
materials for biofuel production is increasing daily either to address the
not all the feedstocks could be found in a particular region indicates that
challenging issues of environmental pollution or the higher rate of pe­
they are different in their compositions because of climatic, geograph­
troleum fuel consumption. Biodiesel among liquid biofuels stands as the
ical, and territorial differences. In this case, co-mingling more than one
most auspicious and suitable fuel alternative to petroleum fuel due to
feedstock is considered a viable approach to examine the effects on
their similarities in properties. The indigenous nature of biodiesel and its
physicochemical properties based on the chemical and biological mod­
potency in greenhouse gases emission reduction makes it a desired fuel
ifications. This process thus creates new feedstock with improve prop­
substitute option. Biodiesel is a long chain fatty acid – alcohol - ester
erties capable of enhancing effective production and biodiesel products.
triglyceride produced in the presence of a catalyst through a chemical
In transesterification reaction, alcohol is an important reactant,
reaction called transesterification. Most biodiesel available in the mar­
which brings about alkylation of esters during the reaction process. It is
ket comes from edible oil sources. However, deploying edible oils in
quicker to dissolve in base catalyst than in acid catalyst. Catalysts
biodiesel production does not only increase the cost of production but
employed in transesterification reaction can be homogeneous or het­
also provide avenue for food versus oil contention. For this reason,
erogeneous. Homogenous catalysts are associated with soap formation,
attention has been diverted to exploiting of food processing waste oil,
difficulty in product isolation, catalyst recovery and spent water gen­
non-edible, and underutilized plant oils, which are available and of less
eration. Heterogeneous catalyst has extensively been reported as a
economic value to develop a long-term sustainable feedback basis for

* Corresponding author.
E-mail address: fatherfounder@yahoo.com (A.C. Eloka-Eboka).

https://doi.org/10.1016/j.rineng.2022.100645
Received 6 July 2022; Received in revised form 10 September 2022; Accepted 13 September 2022
Available online 20 September 2022
2590-1230/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Unlike other RSM modelling such as central composite design (CCD)


Nomenclature and Box Behnken design (BBD) which involves a large number of ex­
periments, Taguchi method involves the smallest number of experi­
ANOVA Analysis of variance ments. Only nine experimental runs are employed to perform its
CaO Calcium oxide optimization function. It makes use of orthogonal arrays for optimiza­
CE-CaO Calcined eggshells Calcium oxide tion of different parameters controlling the process and their variation
CBP Calcined banana peels magnitude. It allows for the assembling of data to decide on the most
CV Coefficient of variance influential factor on the response from the least number of experiments
Cf Contribution factor [16]. It has been applied in many esterification and transesterification
EDX Energy dispersive X-ray spectroscopy experiments and reported to be effective in optimizing the process
FFA Free fatty acid variables [17–20]. Although linseed and marula seed oil has been uti­
FAME Fatty acid methyl ester lised, as single feedstock oil for biodiesel production, their combination
FT-IR Fourier transform infrared spectroscopy (bi-oil) in transesterification process has not been investigated.
LOMO Linseed oil-marula oil This study hence investigates the transesterification of the bi-oil of
LOMOB Linseed oil-marula oil biodiesel linseed and marula (LOMO) at (50:50 w/w) using waste biomass
TD Taguchi Design modified functional base catalyst derived from eggshells and banana
RSM Response surface Methodology peels, which is an aspect of the novelty of this study. The process con­
R2 coefficient of determination ditions were also modelled and optimized using Taguchi design, the
SNR Signal to oil ratio reusability of the synthesized catalyst was investigated. The overall
SEM Scanning electron microscopy protocol is a contribution to an existing knowledge of waste develop­
XRD X-ray diffraction ment and effective biodiesel production approach.

2. Experimental

solution to most of the homogenous catalysts’ draw back being that they 2.1. Materials
can be recovered and reused. They are grouped into two, acid and base.
Acid heterogeneous catalysts are slow in reactions and are linked to Commercial linseed and marula seed oils were supplied by United
longer reaction time and higher operation conditions. Base catalysts Scientific SA cc, Congella, South Africa. The chemicals and reagents
have become attractive due to their fast reaction time and mild opera­ used for experimentation such as methanol (98%), ethanol (98%),
tion conditions. Aside these characteristics, base heterogeneous cata­ diethyl ether, cyclohexane, potassium hydroxide, phenolphthalein (in­
lysts can be sourced from mineral-rich waste renewable resources. dicator), wij’s solution were all analytical grades.
Wastes such as eggshells among other animal residues have been re­
ported as the chief source of natural CaO, with the highest content of 2.2. Preparation of the bio-composite catalysts and characterization
80–90%; and they has been used extensively in biodiesel production
[1–4]. Agricultural wastes such as banana peels, Carica papaya peels and The preparation method adopted for the bio-composite catalyst in
stem, cocoa pod husk have all been studied and reported as the major this study is similar to the one used in in the recent study of Etim et al
source of potassium based compounds such as K2CO3, K2O, KCl, K2SO4 [15] but with slight changes. The waste eggshells and banana peels were
and so on, with high catalytic performance converting up to 90–99% washed using distilled water to remove all debris at the surface of the
biodiesel yield [5–7]. materials. The washed materials were dried in an oven for 48 h at 80 ◦ C
Recently, several biochemical modification approaches applicable in to constant weight. The dried eggshells were crushed to powder, sieved
the development of heterogeneous catalyst for biodiesel production has to obtain fine particle size of less 75 μm, and further calcined in a pro­
been established and reported by Etim et al [8]. The purpose of which is grammable muffle furnace [Havells, B 25, DHMCBSPF025] at 900 ◦ C for
to develop improve physicochemical characteristics of the heteroge­ 3 h to generate natural CaO, labelled as calcined eggshell (CE-CaO). The
neous catalyst for effective transesterification and enhancement of pure dried banana peels were burnt in open air and ground into fine powder
biodiesel products. Biomass modified functional heterogeneous cata­ using a porcelain mortar and pestle and then sieved to obtain fine char of
lysts, with attachment of various homogenous catalysts at the surface of similar particle size. The resultant powder was calcined at 700 ◦ C for 3 h
either commercial or natural derived CaO such as CaO–KOH, to obtain calcined banana peels (CBP). The calcined char of CBP and
CaO–NaOH, CaO–ZnO–K2CO3, CaO–Na2SO4 and CaO–H2SO4 to CE-CaO were bonded via wet impregnation method.
enhanced effective catalytic activities and purity product [9–11] has Equal proportion of the CE-CaO and CBP were bonded by dissolving
been the most attractive area in heterogeneous catalyst development. 1 g of CE-CaO in a 100 ml of warm distilled water followed by the
Attachment of support enhance increase of the active spores and the addition of 1 g of CBP. The suspension was subjected to vigorous
surface area of the synthesized catalyst [8,12]. However, these modifi­ agitation for 1 h using a magnetic hot plate set at 60 ◦ C, to obtain a
cations although very effective are linked with homogeneous com­ slurry. The resultant slurry obtained were dried over 120 ◦ C for 24 h to
pounds, which are not environmentally friendly. remove moisture. The dried sample of CE-CaO:CBP hybrid were pul­
In order to make the biomass derived catalyst a standalone venture verized and further activated via calcination at 800 ◦ C for 2 h in the
and completely green process, co-mingling of several agro-waste mate­ furnace. The synthesized bio-composite nanoparticles catalyst obtained
rials have to be investigated: cocoa pod husk-kola nut husk and fluted was coded-termed as CE-CaO/CBP-800 NPs. The bio-char obtained from
pumpkin waste has been reacted together to produce a heterogeneous the each material and hybrids; CE-CaO, CBP and CE-CaO/CBP-800 NPs
catalyst for the transesterification of yellow oleander-rubber oil blend were packed separately into airtight bottles and kept in a desiccator for
and was reported effective by Falowo and Betiku, [13]. Mixed animal characterization and further use.
bones was also used to synthesize heterogeneous catalyst for the trans­
esterification of waste cooking oil [14]. A case of a green biogenic 2.3. Catalyst characterization
modification using calcium and potassium rich sources was proposed in
the last study of this authors using eggshells and Carica papaya peels The surface morphology and elemental composition of the catalysts
[15]. The current study therefore is the continuation of previous re­ were was determined using scanning electron microscope (SEM) at a
ported investigation using the same novel approach of development. high resolution by Auriga (Zeiss Germany) fitted to the energy dispersive

2
A.O. Etim et al. Results in Engineering 16 (2022) 100645

X-ray (EDX) detector. The crystalline phase of the composite catalyst ( )


n ( )2
sample was confirmed by XRD, using D-8 Advanced diffractometer (Cu- 1∑ 1
SNR = − 10 log (2)
Kα radiation) fitted to PSD detector (Bruker AXS Karlsruhe, Germany) n i=1 yi
using electron beam generated at 40 kV and 30 mA at ambient tem­
perature. The identification of the active functional groups present in where n is the number of replicates of each experimental condition, yi is
the developed catalyst sample was done by Fourier transform infrared the HOME yield of each experiment. The contribution of each variable to
(FT-IR) spectra, PerkinElmer spectrum (version 10.5.4) at the range of the response was determined by Eq. (3).
4000–500 cm− 1. The basic strength of the synthesized catalyst was
SSf
determined by Hammett indicator method. 25 mg of the catalyst was Cf (%) = × 100 (3)
SST
dispersed in 10 ml of methanol solution in a Hammett indicator and left
to equilibrate for 3 h and the colour changes was observed. The Ham­ where Cf is the influence of each factor, SSf is the sum of square of each
mett indicator used was phenolphthalein (H_9.8) [4,7]. parameter and SST is the total sum of squares of all the parameters. The
linear regression model equation was developed to correlates the pre­
dicted response for all the process parameters.
2.4. Preparation and characterization of the bi-oil
2.6. Transesterification process of the bi-oil using the bio-composite
The bi-oil of linseed-marula seed oil (LOMO) at equal amount was
(CaO/CBPA -800 NPs) catalyst
found to have an acid value of 5.61 mg KOH/g with the corresponding
FFA of 2.81% and was converted using one-step pathway trans­
The transesterification process was conducted in a set up system
esterification method. The bi-oil was then heated at 60 ◦ C and stirred
consisting of 250 mL three-necked round bottom glass reactor equipped
using a magnetic hot plate, for homogeneity and to remove all moisture
with a condenser, thermometer, and magnetic stirrer. A specified
present in the oil. The physicochemical properties of the bi-oil and the
amount of the bi-oil was placed in the reactor and preheated to a desired
biodiesel produced namely the density, viscosity, acid value, iodine
temperature. A certain quantity of methanol as justified by the meth­
value, cetane number, calorific value were estimated according to the
anol/oil ratio displayed in Table 2 was added to the oil in the reactor and
standard protocol [21]. The values estimated were compared with the
allowed to stir for 5 min, followed by the addition of specified amount of
ASTM D6751, EN 14,214 and SAN 833 standards.
bio-composite catalyst in weight %. The reaction time commenced
immediately after the catalyst was added. A magnetic hotplate allows
for adjustment of plate temperature to keep the mixture temperature
2.5. Empirical design and statistical analysis
constant for each run. At the end of the reaction, the content was
transferred into a separating funnel and allowed to settle overnight. A
Taguchi design was used to investigate the effect of the process pa­
three-phase layer separation was observed which include the biodiesel
rameters viz a viz methanol-to-oil molar ratio, catalyst loading, reaction
layer at the top, the glycerol and the catalyst layer at the bottom
time and temperature. L9 orthogonal array method with three different
respectively. This was separated accordingly starting from the catalyst
levels and four parameters was considered in the study as presented in
and then glycerol layer, the remaining crude biodiesel product in the
Table 1. This was selected to generate the nine number of experiments
separating funnel was washed thrice with warm distilled water at 50 ◦ C
(L9) in order to achieve the desired results as shown in Table 2. The
and further dried by heating over Na2SO4. The yield of the purified
number of parameters and the variation levels determine the orthogonal
biodiesel was determine gravimetrically using Eq. (4).
array selection [20]. The triplicates of each experiment are to minimise
the error. The performance characteristic of each parameter was weight of purified LOMOB (g)
Biodiesel yield (%) = X100 (4)
assessed to predict the most significant parameter. The relationship with weight of LOMO sample (g)
which the number of experiments can be evaluated is given by Eq. (1).
3. Result and discussion
N = 1 + (L − 1)Z (1)

where N denotes the number of experiments, Z is the total number of 3.1. Catalyst characterization
control parameters studied and L is the corresponding levels.
The design, modelling, optimization, and statistical analysis of data The properties of bio-composite catalyst were investigated alongside
in this study was done using the design expert software (version 11.0). with the individual catalyst produced from the two waste materials for
Mathematical quadratic model equation was developed using regression comparison.
analysis by correlating the independent variables with the responses.
The performance characteristics of the obtained experimental result 3.1.1. Basic strength
could be computed using signal-to-noise ratio. Signal function of the Hammet indicator test conducted for the basicity determination for
expected outcome are serving as the objective optimization problem the synthesized catalysts showed the basic strength of H_10.9 for CE-
therefore, they are used to determine the extent of deviation of the CaO, H_12.0 for CBP and H_22.8 for CE-CaO/CBP-800 NPs were taken
quality function to the expected response [22]. A signal to noise ratio to be higher than the weakest indicator (H_9.8). The increase of the basic
(SNR) is the mean value of the quality characteristics of FAME yield to strength of these catalysts are attributed to the major concentration of
its standard deviation, as expressed in Eq. (2). the alkaline metal oxides present [23,24]. The higher basic strength
observed for the composite catalyst is due to the combined strength of
the two major active basic ions (K+ and Ca+). The efficiency of the
Table 1
Process parameters and their values at different levels. heterogeneous base-catalysed transesterification reaction is enhanced
due to the strong basic sites on the surface of these catalysts, which is
Symbols Parameters unit Levels
responsible for their high catalytic activities in the conversion process.
1 2 3

A Methanol/oil – 9:1 12:1 15:1 3.1.2. FT-IR analysis


B Catalyst loading wt% 2.5 3.5 4.5 The FT-IR analysis is important to determine the active surface
C Time min 50 60 70 functional groups of compound present in the catalysts. The combined
o
D Temp C 55 65 75
FT-IR plots for CE-CaO, CBP and CE-CaO/CBP-800 NPs are shown in

3
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Table 2
Taguchi L9 Experimental matrix for LOMO transesterification.
Std Order Run Meth/Oil ratio Cat. Loading (wt%) Time (min) Temp (oC) LOMOB (%) Predicted Value SNR

7 1 15 2.5 70 65 93.12 93.77 39.38


6 2 12 4.5 50 65 87.72 86.58 38.86
9 3 15 4.5 60 55 87.10 87.59 38.80
2 4 9 3.5 60 65 89.60 90.09 39.04
8 5 15 3.5 50 75 92.23 91.09 39.29
3 6 9 4.5 70 75 72.53 73.18 37.21
5 7 12 3.5 70 55 82.89 83.54 38.37
1 8 9 2.5 50 55 73.97 72.83 37.38
4 9 12 2.5 60 75 72.33 72.82 37.18

Fig. 2. Combined XRD spectra of CE-CaO, CBP and CE-CaO/CBP-800 NPs.


Fig. 1. Combined FT-IR plot for CBP, CE-CaO and CE-CaO:CBP-800 NPs.
to the decomposition of compounds present in CBP to KCl, K2CO3 and
Fig. 1. The bands at 3642 cm-1 are attributed to (▬OH) of absorbed MgCaSiO4.
moisture on the surface of catalyst samples. This was obvious in CE-CaO The XRD pattern indicates the dominant peak of the face centered
and more intense in CE-CaO/CBP-800 NPs samples due to water cubic (FCC) lattice structure of KCl corresponding to 2θo at 28.3, 40.5
absorbed during the synthesis as a result of the hygroscopic nature of and 50.2. A crystal monoclinic structure of K2CO3 observed at the degree
CaO. Although not clearly visible on CBP samples, but can still be of 12.8, 25.7, 26.8, 32.6, 39 and 41.8; and the orthorhombic structure of
observed around 3500 - 3000 cm-1 [24]. The characteristic band around MgCaSiO4 was also detected as corresponding to 15.9, 21.2, 24.47, 27.9,
1411 and 873 cm-1 in CBP and CE-CaO/CBP-800 samples can be 30.4, 33.5 and 34.6. The result showed that the new compounds
accredited to the carbonyl (O–C–O) vibration of metal carbonate entities HN4Al3(SO4)2(OH)6 and Na K Cl were formed as a result of the
resulting in the K–O and Ca–O vibrations at high calcination tempera­ impregnation in the X-ray diffraction pattern of CE-CaO/CBP-800 NPs
tures [25]. The band around 600 in CBP are attributed to K–O [52]. The hybrid. The peaks at 17.65, 29.57, 47.43, 52.16 can be ascribed to the
band around 500 cm-1 could be accredited to vibratory stretching of presence of rhombohedral structure of Ammoniolunite – Ammonium
Ca–O and are clearly visible in CE-CaO, and CE-CaO/CBP-800 NPs Aluminium Sulphate (HN4Al3(SO4)2(OH)6). While the peaks at 28.6,
spectra [26]. The EDX result also corroborates the observed functional 40.9, 50.7 and 67 indicates the presence of FCC lattice structure of so­
groups identified in the IR spectra of each sample. dium potassium chloride (Na K Cl). The figure clearly showed that
during the process of impregnation, minerals from CBP have been
3.1.3. The XRD analysis absorbed on the surface of CaO, leading to the present and improvement
The XRD analysis was performed on the catalyst samples to deter­ of other minor mineral components contained in the two materials
mine their mineral compositions at the crystalline phase. The combined through the high heat of calcination. Thus, the catalytic activity of the
XRD spectra of CE-CaO, CBP and CE-CaO/CBP-800 NPs are shown in developed catalyst in the transesterification process was improved due
Fig. 2. The X-ray diffraction (XRD) intensity plot shows the crystalline to the support of other mineral components present alongside with CaO.
characteristics of the catalyst structure of CE-CaO, CBP and CE-CaO/
CBP-800 NPs catalysts. The obtained peaks were compared with the 3.1.4. SEM-EDX analysis
Joint Committee on Powder Diffraction Standards (JCPDS) database. In The SEM analysis was conducted on the synthesized catalyst samples
CE-CaO, CBP and CE-CaO/CBP-800 NPs samples, distinct peaks were to examine their morphological structure. The micrograph displaying
observed between 14 and 55◦ and this can be attributed to the efficacy of the morphology of the powdered CE-CaO, CBP and CE-CaO/CBP-800
the calcination temperature. The calcined eggshells (CE) diffraction NPs is depicted in Fig. 3. The calcined eggshells consist of irregular
peak showed CaO as the dominant and active compounds present in micro morphology of stone-like structures while calcined banana peels
eggshells (JCPDS: 00-037-1487). However, there are other minor peaks consist of a number of aggregates of microporous structure due to
indicating the presence of Ca (OH)2. This is due to the hygroscopic na­ calcination effect. The composite nano particles catalyst CE-CaO/CBP-
ture of CaO when exposed to air and moisture. The peak intensity was 800 shows the aggregate of numerous porous and spongy nature of
found to decrease at high temperature of calcination attributable to the the ash mineral particles as a result of high temperature of calcination.
increasing formation of CaO. Thus CaO peaks increase as a result of the The particle size distribution of the non-uniform structure on the surface
greater crystallinity of the catalyst [14]. In CBP sample, calcination led

4
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Fig. 3. The SEM images of CE-CaO, CBP and CE-CaO:CBP-800.

of the composite catalyst might be attributed to the merging effect of the parameters viz methanol to oil, catalyst loading and temperature. The
two biogenic materials during synthesis, which clearly indicate the insignificant parameter (reaction time) was eliminated from the model
modified surface structure of pure CaO as a result of the attachment of equation. The plot of experimental values and the predicted values is
other mineral particles from CBP catalyst [15]. depicted in Fig. 4. The plot shows that both values are in good agree­
Moreover, the EDS analysis indicates that CE contains only two ment, which indicate the significance and accuracy of the model in
major elements, which are Ca (50.8%) and O (49.2%) while CBP was evaluating the yield(s).
found to be mainly composed of high content of K (44.1), and Cl The result of the transesterification process of LOMO together with
(14.17%) with trace amounts of P (0.38%), Mg (0.72%), Na (0.15%) and the predicted yields are presented in Table 2. The analysis by ANOVA
Si (0.77%). However, the hybrids of the two biogenic materials CE-CaO- was used to determine the significance of the model for the optimization
CBP-800 was found to contain majorly of high content of O (44.39%), Ca process together with individual parametric effect on the response.
(34.93%), K (18.62%) with small amounts of Cl (1.41%) and Si (0.65%). Table 3 shows the ANOVA of the transesterification process of the
The content of Ca was observed to increase which was minimal in CBP. linseed oil-marula oil biodiesel (LOMOB). The result majorly focus on
The existence of the increased content of K and Cl was found present in
pure CaO surface, indicating the effectiveness of impregnation and
calcination processes.

3.2. Mathematical modelling equation

The multiple regression model analysis was done to determine the


effect of each parameter on LOMOB yield. The numerical value obtained
for each significant parameter coefficient at 95% confident level are
expressed in terms of mathematical equation for prediction of FAME
yield and is presented in Eq. (5).

Y = 83.50 – 4.80A [1]– 2.52A [2]– 3.69B [1] + 4.74B [2]– 2.18D [1]
+ 6.65D [2] (5)

Where Y is the biodiesel yield, A [1] and A [2] is methanol to oil ratio
at first and second level, B [1] and B [2] is catalyst loading at first and
second level and D [1] and D [2] is the reaction temperature at first and
second level respectively as shown in Table 1. The equation is useful for
identifying the relative impact of the factors by comparing the factor
coefficients. The model equation only comprised the significant Fig. 4. Predicted versus actual values of LOMOB yield.

5
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Table 3 Table 5
ANOVA for the selected factorial model. Percentage contribution of each process parameter on the yield.
Source Sum of Squares df Mean Square F-value p-value Parameter Contribution factor (%)

Model 567.14 6 94.52 32.04 0.0306 Methanol-to-oil molar ratio (A) 43.86
A-Methanol/oil 248.77 2 124.39 42.17 0.0232 Catalyst Loading (B) 19.68
B-Catalyst 111.63 2 55.82 18.92 0.0502 Temperature (D) 36.45
D-Temp 206.73 2 103.37 35.04 0.0277 Reaction time (C) Insignificant
Residual 5.90 2 2.95
Cor Total 573.04 8
3.3. Effect of process parameters on LOMOB yield

the mean performance characteristics established by Fischer’s test value The effect of the process variables on LOMOB production can be
(F-value) estimation. The F-value and the sum of squares are used to adjudged based on the experimental result. From the ANOVA in­
explain the impact and the status of each parameter on the yield. The p- vestigations and the contribution factor of individual parameter (Ta­
value measures the probability of obtaining the F-value due to noise and bles 3 and 4), it is observed that methanol to oil molar ratio, catalyst
the value below 0.05 indicate the significance of a specific parameter. loading, and reaction temperature were the significant parameters that
The high sum of square value of 567.14 obtained for the model in the influences LOMOB yield and hence their impacts are studied. The effect
study was enough to indicate that the model is significant. The ANOVA of time is not discussed because it appears insignificant with a negligible
also shows that out of the four selected parameters for the study, the impact on LOMOB yield. The reaction time selected were within the
significant effect of the two parameters is well pronounced on the yield. range studied for the maximum yield conversion of LOMO. Although the
Nevertheless, the effect of catalyst loading with the p-value of 0.05 is less reaction time is a vital factor in the conversion reaction process but in
significant while the effect of reaction time is completely negligible in this case, it was insignificant since there was less variation in FAME yield
the chosen range of parameters. Among the three parameters in Table 3, within the range of time chosen.
methanol/oil ratio has the highest influence on FAME yield with F-value
and sum of square of 42.17 and 248.77 respectively. 3.3.1. Effect of methanol to bi-oil molar ratio
Reaction temperature also have an influence on FAME yield with an Methanol to oil molar ratio is observed to be the most influential
F-value of 35.04 and sum of square of 206.73 and 2.7% chance to obtain parameter affecting LOMOB yield (Table 3). The investigation of the
its high value due to noise. The catalyst loading also have effect on impact of this parameter was done by varying the three levels at 9:1,
FAME yield but with less significant than methanol/oil ratio and tem­ 12:1 and 15:1. The optimum LOMOB yield was observed at methanol-to-
perature. The significance of each parameter could also be demonstrated oil ratio of 15:1 as confirmed by Fig. 5a. As the transesterification re­
with the estimation of the contribution factors of individual parameter action is reversible in nature, large amount of methanol is required to
in the process. The contribution factors are evaluated by Eq. (3) and the sustain the reaction in the forward direction which favours the product
results obtained is displayed in Table 5. The contribution factor of formation [8]. However, precaution must be taken to avoid dilution of
methanol/oil molar ratio was found to be 43.86, which significantly catalyst with excess methanol, which may lead to deactivation of the
influenced the process with strong impact, reaction temperature was activity of the catalyst in the reaction process. Thus, excessive increase
found with 36.45 contribution factor which also shows strong impact on of methanol to oil ratio hinders glycerol separation leading to decrease
the yield and the catalyst loading with only 18.92 contribution factor in biodiesel yield [19].
which indicate less influence on FAME yield as compared to the other
two parameters. 3.3.2. Effect of bio-composite (CE-CaO/CBP-800) catalyst loading
The analyses of the process that was also used to evaluate the sig­ Catalyst concentration is another important variable that influence
nificance and the applicability of the chosen model for the trans­ biodiesel yield. The effect of CE-CaO/CBP-800 loading was studied by
esterification of the LOMO are displayed in Table 4. This include the varying the weight from 2.5 to 4.5 wt%. As displayed in Fig. 5b, by
correlation coefficient R2, adjusted R2 and predicted R2. The value of increasing the catalyst from the lowest level (2.5 wt%), leads to gradual
correlation coefficient R2 (0.9897) obtained signifies the accuracy and increase in LOMOB till 3.5 wt%. Further increase at this drastically re­
the good fit of data in the chosen model. The adjusted R2 (0.9588) and duces the yield. This is because excessive use of CE-CaO/CBP-800 is
predicted R2 (0.7915) indicates good agreement between each other found to form emulsion resulting on higher viscosity of the reaction
with acceptable difference of 0.16 which is less than the maximum mixture, thereby promoting poor agitation and hinders biodiesel for­
allowable difference of 0.2. This shows the extent of variation explained mation [27]. In this study, the impact of catalyst loading was less pro­
by the model about the mean and the prediction aptitude of the response nounced on LOMOB maximum yield as compared with that of
within the satisfactory range. Adequate precision ratio of 13.833 greater methanol/oil ratio and temperature. The catalytic activity of the chosen
than the allowable ratio of 4 is desirable and indicate adequate response range of catalyst loading during the transesterification reaction was able
to predict and optimise the yield. The low standard deviation value of to convert all LOMO to LOMOB.
1.72 and the coefficient of variance of 2.06% implies the validity of the
model and its capability to predict the best condition with high 3.3.3. Effect of reaction temperature
precision. Reaction temperature is observed to be another most influential
factor aside methanol to oil ratio on LOMOB yield. Temperature effect
on FAME yield was studied by varying the values in the order of 55, 65
and 75 ◦ C at three different levels (Table 1). The maximum yield was
observed at the temperature of 65 ◦ C. Fig. 5c shows that LOMOB yield
Table 4
Fit statistical of the model for LOMOB production process. increases from the temperature of 55 ◦ C to maximum at 65 ◦ C, thereafter
decreases drastically at further increase in temperature. Although, in­
R2 0.9897 Standard deviation 1.72
crease in temperature increases the rate of the reaction, excessive in­
Adjusted R2 0.9588 Mean 83.50 crease in temperature above the boiling point of methanol during
Predicted R2 0.7915 C.V. (%) 2.06
production do not favour the molar ratio of methanol in the reaction
Adeq Precision 13.8334 PRESS 119.47
BIC 37.12 − 2 Log Likelihood 21.74 mixture due to evaporation. The resultant effect may lead to a shift of the
AICc 147.74 reversible equilibrium toward the backward reaction, which favour

6
A.O. Etim et al. Results in Engineering 16 (2022) 100645

3.4. Optimization of the process parameters and model validation

The process optimization was achieved by solving the regression


equation (Eq. (4)). The optimum predicted values obtained are meth­
anol to oil ratio of 15:1, CE-CaO/CBP-800 loading of 3.5 wt%, reaction
time of 50 min, and temperature 65 ◦ C with LOMOB yield of 95.53 wt%.
The predicted condition was validated by conducting a triplicate
confirmatory experiment with an average yield of 95.03 wt%. Table 6
showed a brief review of the performance of the single and blended
biomass developed heterogeneous catalysts together with that of the
present study in transesterification reactions. Thus, comparing the op­
timum operating condition of this study using the developed bio-
composite catalyst as shown in Table 6, it can be observed that it is
more economical in terms of time and also compete well in yield with
other study using similar biomass catalysts.

3.5. Reusability test of the bio-composite catalyst

Eco-friendliness and economic viability of heterogeneous catalysts


are based on their recyclability potential. The reusability test of CE-CaO/
CBP-800 was investigated in biodiesel production by repeated reused of
the recovered catalyst in subsequent transesterification reaction. This
was achieved under optimum condition by decanting the spent catalyst
after each round of experiment and then washed with hexane, filtered
and dried in the oven for 6 h at 120 ◦ C and further activated at 700 ◦ C for
2 h. The activated catalyst was then used in subsequent cycle of pro­
duction. FAME yield of 85.2% was observed at the 6th cycle (Fig. 6),
indicating the decrease in process efficiency. Similar method has also
been reported in literature with a minimal decrease of 6.2% after 5th
cycle when gold-core-shell was used to transesterify sunflower oil [42].
The decrease in FAME yield however might be attributed to the weak­
ening of the catalyst activity as a result of leaching and blockages of the
catalyst active sites by oil, water molecules and glycerol [43–45].
However, CaO catalyst is noted for its severe leaching characteristics
due to its hygroscopic nature. Attachment of K compound and other
mineral elements from banana peels on the surface of CaO (eggshells)
coupled with high temperature calcination prevent leaching [46]. A
reported by Papargyriou et al. [47]; in their experiment using pure CaO
and CaO–Ca3Al2O6 composite to transesterified fish oil also showed that
CaO was fast to deactivate than CaO–Ca3Al3O6 due to severe leaching.
The reusability process of this study was investigated up to 6 cycles with
infinitesimal decrease in LOMOB yield (Fig. 6). Thus, the result confirms
that the developed catalyst has appreciable heterogeneous qualities,
which can be adopted at industrial level.

4. Fuel characterization

The quality of LOMOB produced was ascertained through its physi­


ochemical properties. The obtained result compared with the ASTM
D6751, EN 14,214 and SAN 833 standards (Table 7). LOMOB is a good
substitute fuel for diesel engines, due to all its properties are within the
standard specifications for biodiesel. the Cetane number of 58 is above
the minimum required, which shows that LOMOB produced has a short
ignition delay, which will lead to effective cold start-up of engine [30].
The high calorific value of LOMOB suggest it improve energy capacity
[50]. The low acid value of LOMOB suggest that the metal part of the
engine will not be damaged and the associated problem such as injector
Fig. 5. a: The plot of FAME yield versus methanol-to-bi-oil molar ratio, b: The fouling, fuel filter plugging and deposit formation will be minimised
plot of FAME yield versus bio-composite catalyst, c: The plot of FAME yield when LOMOB is applied in engine [30].
versus temperature.
5. Conclusion
triglyceride formation (soap). Moreover, the selected temperature
within the study range was capable to convert LOMO to LOMOB without A new bio-composite catalyst was successfully developed from waste
much loss of methanol. biomass material of eggshells and banana peels. The synthesized catalyst
showed the presence of high alkali elements. The crystal structure of the
catalyst contains majorly of calcium and potassium compounds, which

7
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Table 6
Review of single and mixed biomass heterogeneous catalysts for biodiesel production.
Lipid feedstock Waste Biomass materials Calcination Transesterification conditions Yield References
conditions (%)
Cat amount Oil/methanol ratio Reaction Time Reaction Temp
(wt%) (molar) (min) (oC)

Linseed oil Musa accuminata peels 700 ◦ C, 4 h 2.7 1:11 51.42 65 96.50 [28]
Used cooking oil Carica papaya peels 700, 4 h 3.5 1:12 60 65 97.5 [8]
Waste cooking oil Tectona grandis leaves 700, 4 h 2.5 6:1 180 25 100 [29]
Marula oil Banana peels 700, 4 h 2.0 6:1 50 65 96.45 [30]
Waste cooking oil Tectona grandis leaves 700, 4 h 2.5 6:1 180 25 100 [29]
Jatropha Heteropanax fragrans 550, 2 h 7.0 12:1 65 65 97.75 [31]
(Kesseru)
Neem Plantain peels 700, 4 h 0.65 0.73:1v/v 57 65 99.2 [32]
Waste cooking oil Carica papaya stem 700, 4 h 2.0 9:1 180 60 95.23 [23]
Soy beans oil Moringa leaves 500, 2 h 6.0 6:1 120 65 86.7 [33]
Soy beans oil Tucumá peels 800, 4 h 1.0 15:1 240 80 97.3 [34]
Neem Coco pod husks 700,4 h 0.6 0.73:1v/v 57 65 99.3 [35]
Ceiba pentandra Snail shell 850, 4 h 1.0 9:1 60 60 56.7 [36]
seed oil
Waste cooking oil Chicken bones 900, 4 h 5.0 15:1 240 65 89.33 [37]
Rubber seed oil Cockle shell 900, 4 h 9 15.57:1 204 65 88.06 [38]
Yello oleander- Cocoa/kola nut/fluted 500, 4 h 1.5 9:1 40 55 95.02 [13]
rubber pumkin
Honne Cocoa husk/plantain 500, 4 h 4.5 15:1 90 65 98.98 [39]
peels
Honne/neem, Plantain/cocoa/kola nut 500, 4 h 1.15 12:1 6 150 W 98.40 [40]
rubber
Palm oil Rice husk/eggshells 800 ◦ C, 4 h 7.0 9:1 240 65 91.50 [41]
Used cooking oil Waste chicken/fish 1000 ◦ C,4 h 1.98 10:1 114 65 89.50 [14]
bones
Used vegetable oil Eggshell/papaya peel 900/700 ◦ C 3 h 3.78 14.9:1 80 65 91.20 [15]
Linseed-marula oil Eggshells/banana peels 900/700 ◦ C, 3 h 3.5 15:1 60 65 95.03 This Study

combined strength of the two major active basic ions (K+ and Ca+). The
transesterification process was designed and modelled by Taguchi. The
optimum values obtained was methanol to oil ratio 15:1, CE-CaO/CBP-
800 of 3.5 wt%, reaction temperature of 65 ◦ C and time of 60 min, which
led to a validated average LOMOB yield of 95.03 wt%. The study in­
dicates that the CE-CaO/CBP-800 catalysed transesterification with
methanol was effective.
The properties of the final product obtained were within the bio­
diesel specification standards ASTM D6751, EN 14,214 and SAN 833.
The CE-CaO/CBP-800 could be reused for six reaction cycles with high
yield. Thus, the combined effect of biogenic waste materials and the
blend linseed-marula oil are potential source of an effective and sus­
tainable feedstock for biodiesel development.

Credit author statement

Author AO Etim: jointly with the supervisors conceptualized the


Fig. 6. Reusability test of CE-CaO/CBP-800. study, carried out experiments, wrote the draft and interpreted the re­
sults. Author P. Musonge: jointly conceptualized the study, reviewed
are responsible for the high catalytic function of the catalyst in the the manuscript and supervised the study. Author AC Eloka-Eboka:
transesterification of LOMO bi-oil. The properties characterization in­ jointly conceptualized and supervised the study, reviewed the manu­
vestigations revealed that the basic strength of the composite catalyst script, managed the editorials and the corresponding authorship of the
was higher than the individual produced catalysts. This is due to the final publication.

Table 7
Physiochemical properties of LOMOB.
Units LOMOB Method Limits
Property
ASTM D6751 EN 14214 SAN 833

Physical state/colour – Liquid/light yellow Visual inspection – –


Density at 25 ◦ C g/cm3 0.864 ASTM 4052 0.850 0.86–0.90 0.86–0.90
Kinematic Viscosity at 40 ◦ C mm2/s 3.69 ASTM D445 1.9–6.0 3.5–5.0 3.5–5.0
Acid Value mg KOH/g 0.42 ASTM D664 0.5 max 0.5 max 0.5 max
Saponification value 175.31 – –
Iodine Value g I2/100 g 79.27 AOAC – 120 max 140 max
Calorific value MJ/kg 40.4 [48] – 35 –
Cetane number 58 [49] 47 min 51 min 51 min

8
A.O. Etim et al. Results in Engineering 16 (2022) 100645

Declaration of competing interest Taguchi approach, Renew. Energy 105 (2017) 616–624, https://doi.org/10.1016/
j.renene.2016.12.096.
[19] S.H. Dhawane, T. Kumar, G. Halder, Biodiesel synthesis from Hevea brasiliensis oil
The authors declare that they have no known competing financial employing carbon supported heterogeneous catalyst: optimization by Taguchi
interests or personal relationships that could have appeared to influence method, Renew. Energy 89 (2016) 506–514, https://doi.org/10.1016/j.
the work reported in this paper. renene.2015.12.027.
[20] P. Kumar, M. Aslam, N. Singh, S. Mittal, A. Bansal, M.K. Jha, A.K. Sarma,
Characterization, activity and process optimization with a biomass–based thermal
Data availability power plant’s fly ash as a potential catalyst for biodiesel production, RSC Adv. 5
(2015) 9946–9954, https://doi.org/10.1039/c4ra13475c.
[21] D. Kumar, T. Das, B.S. Giri, E.R. Rene, B. Verma, Biodiesel production from hybrid
Data will be made available on request. non-edible oil using bio-support beads immobilized with lipase from Pseudomonas
cepacia, Fuel 255 (2019), 115801, https://doi.org/10.1016/j.fuel.2019.115801.
[22] R. Kumar, K. Sureshkumar, R. Velraj, Optimization of biodiesel production from
Acknowledgement
Manilkara zapota (L.) seed oil using Taguchi method, Fuel 140 (2015) 90–96,
https://doi.org/10.1016/j.fuel.2014.09.103.
Authors are grateful for the the financial support from the National [23] M. Gohain, K. Laskar, A.K. Paul, N. Daimary, M. Maharana, I.K. Goswami,
Research Foundation of South Africa, under grant number 130427. A. Hazarika, U. Bora, D. Deka, Carica papaya stem: a source of versatile
heterogeneous catalyst for biodiesel production and C–C bond formation, Renew.
Energy 147 (2020) 541–555, https://doi.org/10.1016/j.renene.2019.09.016.
References [24] K. Rajkumari, L. Rokhum, A sustainable protocol for production of biodiesel by
transesterification of soybean oil using banana trunk ash as a heterogeneous
catalyst, Biomass Convers. Biorefinery (2020), https://doi.org/10.1007/s13399-
[1] A. Marwaha, P. Rosha, S.K. Mohapatra, S.K. Mahla, A. Dhir, Biodiesel production
020-00647-8.
from Terminalia bellerica using eggshell-based green catalyst: an optimization
[25] T.F. Adepoju, M.A. Ibeh, E.O. Babatunde, A.J. Asquo, Methanolysis of CaO based
study with response surface methodology, Energy Rep. 5 (2019) 1580–1588,
catalyst derived from egg shell-snail shell-wood ash mixed for fatty acid
https://doi.org/10.1016/j.egyr.2019.10.022.
methylester (FAME) synthesis from a ternary mixture of Irvingia gabonensis
[2] P.R. Pandit, M.H. Fulekar, Egg shell waste as heterogeneous nanocatalyst for
-Pentaclethra macrophylla - elais guineensis oil blend: an application of simplex
biodiesel production: optimized by response surface methodology, J. Environ.
lattice and central composite design optimization, Fuel 275 (2020), 117997,
Manag. 198 (2017) 319–329, https://doi.org/10.1016/j.jenvman.2017.04.100.
https://doi.org/10.1016/j.fuel.2020.117997.
[3] W.U. Rahman, A. Fatima, A.H. Anwer, M. Athar, M.Z. Khan, N.A. Khan, G. Halder,
[26] A.O. Etim, A.C. Eloka-Eboka, P. Musonge, Potential of Carica papaya peels as
Biodiesel synthesis from eucalyptus oil by utilizing waste egg shell derived calcium
effective biocatalyst in the optimized parametric transesterification of used
based metal oxide catalyst, Process Saf. Environ. Protect. 122 (2019) 313–319,
vegetable oil, Environ. Eng. Res. 26 (2021) 200–299, https://doi.org/10.4491/
https://doi.org/10.1016/j.psep.2018.12.015.
eer.2020.299.
[4] N. Tshizanga, E.F. Aransiola, O. Oyekola, Optimisation of biodiesel production
[27] P. Verma, M.P. Sharma, Review of process parameters for biodiesel production
from waste vegetable oil and eggshell ash, S. Afr. J. Chem. Eng. 23 (2017)
from different feedstocks, Renew. Sustain. Energy Rev. 62 (2016) 1063–1071,
145–156, https://doi.org/10.1016/j.sajce.2017.05.003.
https://doi.org/10.1016/j.rser.2016.04.054.
[5] M. Gohain, A. Devi, D. Deka, Musa balbisiana Colla peel as highly effective
[28] A.O. Etim, P. Musonge, A.C. Eloka-eboka, Process optimization of bio-alkaline
renewable heterogeneous base catalyst for biodiesel production, Ind. Crop. Prod.
catalysed transesterification of flax seed oil methyl ester, Sci. African 16 (2022),
109 (2017) 8–18, https://doi.org/10.1016/j.indcrop.2017.08.006.
e01275, https://doi.org/10.1016/j.sciaf.2022.e01275.
[6] V.O. Odude, A.J. Adesina, O.O. Oyetunde, O.O. Adeyemi, N.B. Ishola, A. Etim,
[29] M. Gohain, K. Laskar, H. Phukon, U. Bora, D. Kalita, D. Deka, Towards sustainable
E. Betiku, Application of agricultural waste-based catalysts to transesterification of
biodiesel and chemical production: multifunctional use of heterogeneous catalyst
esterified palm kernel oil into biodiesel : a case of banana fruit peel versus cocoa
from littered Tectona grandis leaves, Waste Manag. 102 (2020) 212–221, https://
pod husk, Waste and Biomass Valorization 10 (2019) 877–888, https://doi.org/
doi.org/10.1016/j.wasman.2019.10.049.
10.1007/s12649-017-0152-2.
[30] Anietie Okon Etim, P. Musonge, A.C. Eloka-eboka, Transesterification via
[7] G. Pathak, D. Das, K. Rajkumari, L. Rokhum, Exploiting waste: towards a
parametric modelling and optimization of marula (sclerocarya birrea) seed oil
sustainable production of biodiesel using: musa acuminata peel ash as a
methyl ester synthesis, J. Oleo Sci. 93 (2021) 77–93.
heterogeneous catalyst, Green Chem. 20 (2018) 2365–2373, https://doi.org/
[31] S. Basumatary, B. Nath, B. Das, P. Kalita, B. Basumatary, Utilization of renewable
10.1039/c8gc00071a.
and sustainable basic heterogeneous catalyst from Heteropanax fragrans (Kesseru)
[8] A.O. Etim, P. Musonge, A.C. Eloka-Eboka, Effectiveness of biogenic waste-derived
for effective synthesis of biodiesel from Jatropha curcas oil, Fuel 286 (2021),
heterogeneous catalysts and feedstock hybridization techniques in biodiesel
119357, https://doi.org/10.1016/j.fuel.2020.119357.
production, Biofuels, Bioprod. Biorefining 14 (2020) 620–649, https://doi.org/
[32] A. Etim, E. Betiku, S. Ajala, P. Olaniyi, T. Ojumu, Potential of ripe plantain fruit
10.1002/bbb.2094.
peels as an ecofriendly catalyst for biodiesel synthesis: optimization by artificial
[9] R. Foroutan, S.J. Peighambardoust, R. Mohammadi, S.H. Peighambardoust,
neural network integrated with genetic algorithm, Sustainability 10 (2018) 707,
B. Ramavandi, Application of walnut shell ash/ZnO/K2CO3 as a new composite
https://doi.org/10.3390/su10030707.
catalyst for biodiesel generation from Moringa oleifera oil, Fuel 311 (2022),
[33] J.L. Aleman-Ramirez, J. Moreira, S. Torres-Arellano, A. Longoria, P.U. Okoye, P.
122624, https://doi.org/10.1016/j.fuel.2021.122624.
J. Sebastian, Preparation of a heterogeneous catalyst from moringa leaves as a
[10] J. Nisar, R. Razaq, M. Farooq, M. Iqbal, R.A. Khan, M. Sayed, A. Shah, I.
sustainable precursor for biodiesel production, Fuel 284 (2021), 118983, https://
ur Rahman, Enhanced biodiesel production from Jatropha oil using calcined waste
doi.org/10.1016/j.fuel.2020.118983.
animal bones as catalyst, Renew. Energy 101 (2017) 111–119, https://doi.org/
[34] I.M. Mendonça, O.A.R.L. Paes, P.J.S. Maia, M.P. Souza, R.A. Almeida, C.C. Silva,
10.1016/j.renene.2016.08.048.
S. Duvoisin, F.A. de Freitas, New heterogeneous catalyst for biodiesel production
[11] I.M. Ogbu, V.I.E. Ajiwe, C.P. Okoli, Performance evaluation of carbon-based
from waste tucumã peels (Astrocaryum aculeatum Meyer): parameters
heterogeneous acid catalyst derived from hura crepitans seed pod for esterification
optimization study, Renew. Energy 130 (2019) 103–110, https://doi.org/10.1016/
of high FFA vegetable oil, Bioenergy Res 11 (2018) 772–783, https://doi.org/
j.renene.2018.06.059.
10.1007/s12155-018-9938-8.
[35] E. Betiku, A.O. Etim, O. Pereao, T.V. Ojumu, Two-step conversion of neem
[12] S.O. Bitire, T.C. Jen, M. Belaid, Synthesis of beta-tricalcium phosphate catalyst
(Azadirachta indica) seed oil into fatty methyl esters using a heterogeneous
from Herring fishbone for the transesterification of parsley seed oil, Environ.
biomass-based catalyst: an example of cocoa pod husk, Energy Fuel. 31 (2017)
Technol. (2021), https://doi.org/10.1080/09593330.2021.1916094.
6182–6193, https://doi.org/10.1021/acs.energyfuels.7b00604.
[13] A.O. Falowo, E. Betiku, A novel heterogeneous catalyst synthesis from agrowastes
[36] C. Muhammad, Z. Usman, F. Agada, Biodiesel production from ceiba pentandra
mixture and application in transesterification of yellow oleander-rubber oil :
seed oil using CaO derived from snail shell as catalyst, Pet. Sci. Eng. 2 (2018) 7–16,
optimization by Taguchi approach, Fuel 312 (2022), 122999, https://doi.org/
https://doi.org/10.11648/j.pse.20180201.12.
10.1016/j.fuel.2021.122999.
[37] M. Farooq, A. Ramli, A. Naeem, Biodiesel production from low FFA waste cooking
[14] Y.H. Tan, M.O. Abdullah, J. Kansedo, N.M. Mubarak, Y.S. Chan, C. Nolasco-
oil using heterogeneous catalyst derived from chicken bones, Renew. Energy 76
Hipolito, Biodiesel production from used cooking oil using green solid catalyst
(2015) 362–368, https://doi.org/10.1016/j.renene.2014.11.042.
derived from calcined fusion waste chicken and fish bones, Renew. Energy 139
[38] M.M. Zamberi, F.N. Ani, M. Fadzli, B. Abdollah, Heterogeneous transesterification
(2019) 696–706, https://doi.org/10.1016/j.renene.2019.02.110.
of heterogeneous transesterification of rubber seed oil biodiesel production,
[15] A.O. Etim, P. Musonge, A.C. Eloka-Eboka, An effective green and renewable from
J. Teknol. (2016) 8–14.
the fusion of bi-component transesterification of linseed oil methyl ester, Biofuels,
[39] E. Adedayo, O. Oluwatumininu, E. Betiku, Cocoa pod husk-plantain peel blend as a
Bioprod. Biorefining (2021) 1–12, https://doi.org/10.1002/bbb.2252.
novel green heterogeneous catalyst for renewable and sustainable honne oil
[16] A.P. Bora, S.H. Dhawane, K. Anupam, G. Halder, Biodiesel synthesis from Mesua
biodiesel synthesis : a case of biowastes-to-wealth, Renew. Energy 166 (2020)
ferrea oil using waste shell derived carbon catalyst, Renew. Energy 121 (2018)
163–175, https://doi.org/10.1016/j.renene.2020.11.131.
195–204, https://doi.org/10.1016/j.renene.2018.01.036.
[40] O.A. Falowo, T.V. Ojumu, O. Pereao, E. Betiku, Sustainable biodiesel synthesis
[17] S.H. Dhawane, B. Karmakar, S. Ghosh, G. Halder, Parametric optimisation of
from honne-rubber-neem oil blend with a novel mesoporous base catalyst
biodiesel synthesis from waste cooking oil via Taguchi approach, J. Environ. Chem.
synthesized from a mixture of three agrowastes, Catalysts 10 (2020) 1–24, https://
Eng. 6 (2018) 3971–3980, https://doi.org/10.1016/j.jece.2018.05.053.
doi.org/10.3390/catal10020190.
[18] S.H. Dhawane, A.P. Bora, T. Kumar, G. Halder, Parametric optimization of
biodiesel synthesis from rubber seed oil using iron doped carbon catalyst by

9
A.O. Etim et al. Results in Engineering 16 (2022) 100645

[41] G.Y. Chen, R. Shan, J.F. Shi, B.B. Yan, Transesterification of palm oil to biodiesel [47] D. Papargyriou, E. Broumidis, M. de Vere-Tucker, S. Gavrielides, P. Hilditch, J.T.
using rice husk ash-based catalysts, Fuel Process. Technol. 133 (2015) 8–13, S. Irvine, A.D. Bonaccorso, Investigation of solid base catalysts for biodiesel
https://doi.org/10.1016/j.fuproc.2015.01.005. production from fish oil, Renew. Energy 139 (2019) 661–669, https://doi.org/
[42] M. Banerjee, B. Dey, J. Talukdar, M. Chandra, Production of biodiesel from 10.1016/j.renene.2019.02.124.
sunflower oil using highly catalytic bimetallic gold-silver core-shell nanoparticle, [48] T. Ahmad, M. Danish, P. Kale, B. Geremew, S.B. Adeloju, M. Nizami, M. Ayoub,
Energy 69 (2014) 695–699, https://doi.org/10.1016/j.energy.2014.03.065. Optimization of process variables for biodiesel production by transesterification of
[43] R. Foroutan, S.J. Peighambardoust, R. Mohammadi, S.H. Peighambardoust, flaxseed oil and produced biodiesel characterizations, Renew. Energy 139 (2019)
B. Ramavandi, Generation of biodiesel from edible waste oil using ZIF-67-KOH 1272–1280, https://doi.org/10.1016/j.renene.2019.03.036.
modified Luffa cylindrica biomass catalyst, Fuel 322 (2022), 124181, https://doi. [49] O.A. Falowo, I.M. Oloko-Oba, E. Betiku, Biodiesel production intensification via
org/10.1016/j.fuel.2022.124181. microwave irradiation-assisted transesterification of oil blend using nanoparticles
[44] N. Mansir, S.H. Teo, U. Rashid, M.I. Saiman, Y.P. Tan, G.A. Alsultan, Y.H. Taufiq- from elephant-ear tree pod husk as a base heterogeneous catalyst, Chem. Eng.
Yap, Modified waste egg shell derived bifunctional catalyst for biodiesel Process. - Process Intensif. 140 (2019) 157–170, https://doi.org/10.1016/j.
production from high FFA waste cooking oil. A review, Renew. Sustain. Energy cep.2019.04.010.
Rev. 82 (2018) 3645–3655, https://doi.org/10.1016/j.rser.2017.10.098. [50] M.K. Yesilyurt, C. Cesur, V. Aslan, Z. Yilbasi, The production of biodiesel from
[45] N. Mansir, S. Hwa Teo, M. Lokman Ibrahim, T.Y. Yun Hin, Synthesis and safflower (Carthamus tinctorius L.) oil as a potential feedstock and its usage in
application of waste egg shell derived CaO supported W-Mo mixed oxide catalysts compression ignition engine: a comprehensive review, Renew. Sustain. Energy
for FAME production from waste cooking oil: effect of stoichiometry, Energy Rev. 119 (2020), 109574, https://doi.org/10.1016/j.rser.2019.109574.
Convers. Manag. 151 (2017), https://doi.org/10.1016/j.enconman.2017.08.069. [52] A.O. Etim, P. Musonge, A.C. Eloka-Eboka, Evaluation of in-situ and ex-situ
[46] A.S. Yusuff, O.D. Adeniyi, M.A. Olutoye, U.G. Akpan, A review on application of hybridization in the optimized transesterification of waste and pure vegetable oils,
heterogeneous catalyst in the production of biodiesel from vegetable oils, J. Appl. Biofuels, Bioproducts and Biorefining (2022), https://doi.org/10.1002/bbb.2360.
Sci. Process Eng. 4 (2017) 142–157, https://doi.org/10.33736/jaspe.432.2017.

10

You might also like