Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Journal Pre-proof

A biobased mixed metal oxide catalyst for biodiesel


production from waste cooking oil: reaction
conditions modeling, optimization and sensitivity
analysis study

Ibhadebhunuele Gabriel Okoduwa, Osamudiamhen


Oiwoh, Andrew Nosakhare Amenaghawon, Charity
O. Okieimen

PII: S2307-1877(24)00057-9
DOI: https://doi.org/10.1016/j.jer.2024.03.009
Reference: JER100372

To appear in: Journal of Engineering Research


Received date: 22 January 2024
Revised date: 29 February 2024
Accepted date: 12 March 2024
Please cite this article as: Ibhadebhunuele Gabriel Okoduwa, Osamudiamhen
Oiwoh, Andrew Nosakhare Amenaghawon and Charity O. Okieimen, A
biobased mixed metal oxide catalyst for biodiesel production from waste cooking
oil: reaction conditions modeling, optimization and sensitivity analysis study,
Journal of Engineering Research, (2024)
doi:https://doi.org/10.1016/j.jer.2024.03.009
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2024 Published by Elsevier.
A biobased mixed metal oxide catalyst for biodiesel production from waste cooking oil: reaction conditions
modeling, optimization and sensitivity analysis study

Ibhadebhunuele Gabriel Okoduwa, Osamudiamhen Oiwoh, *Andrew Nosakhare Amenaghawon and Charity O.
Okieimen

Bioresources Valorization Laboratory, Department of Chemical Engineering, University of Benin, Benin City, Edo
State, Nigeria

Email: *andrew.amenaghawon@uniben.edu

Phone: +234 806 927 5563

ORCID ID: 0000-0003-0433-0114

Abstract

This study assessed the utilization of a heterogeneous catalyst derived from waste marble tiles and cow horn for

of
waste cooking oil-derived biodiesel synthesis. The synthesized composite catalyst was prepared from waste marble
tiles and cow horn through sequential calcination and wet impregnation. The catalyst and its precursors were

ro
characterized to analyze their composition, structural, and surface properties. The biodiesel production process was
modeled and optimized via the response surface methodology (RSM) and artificial neural networks (ANN).

-p
Characterization of waste cooking oil revealed its suitability in biodiesel production. The catalyst had a surface area
of 301.510 m2/g, while the pore volume and pore diameter were 0.165 cm3/g and 2.110 nm respectively, which
re
contributed to the optimum biodiesel yield of 98% at a reaction time of 119.92 min, a catalyst concentration of 5.12
wt%, a reaction temperature of 74.86 °C, and methanol-to-oil ratio of 11.43:1. The catalyst was reusable for up to
seven cycles while retaining significant activity. The optimized biodiesel sample had properties that met standards.
lP

RSM and ANN demonstrated adequate representation of the biodiesel yield as shown by their high R2 values
although ANN was superior since it had higher R2 (0.9993) and adjusted R2 (0.9977) values. Global sensitivity
a

analysis results showed reaction temperature as the most important input variable. This work has provided insights
into the valorization of waste cooking oil for sustainable biodiesel production through heterogeneous catalysis.
rn

Keywords: Biodiesel, modeling, optimization, waste cooking oil, global sensitivity analysis
u

1. Introduction
Jo

Sustainable energy has been identified as one of the very important ingredients of development of a country and its
generation and consumption levels have been taken as an indication of the level of economic development and
growth of that country [1]. The global energy demand has been on a progressive increase in recent times and this
has been linked directly to the increasing global population and level of industrialization [2]. The transportation
sector has had a large share of the demand for energy in the form of liquid fuels like gasoline and diesel [3]. As it
stands, the global energy demand is mostly met from fossil-based sources mainly petroleum or crude oil [4].
However, recent evidence has shown that continued utilization of petroleum-based fuels is not sustainable for
several reasons [5]. The combustion of petroleum-based fuels like diesel, kerosene, and gasoline has been shown to
have seriously negative environmental implications [6].

An important biofuel that has been identified as a suitable liquid fuel for transportation is biodiesel. Biodiesel has
many positive attributes and has thus been described as a viable and sustainable alternative to petroleum-based
diesel. It is biodegradable, renewable, nontoxic, and environmentally benign. It also acts as a solvent and can clean
residue deposited by mineral diesel in the combustion chamber thereby maintaining the optimal operating efficiency
of the engine [7]. Furthermore, biodiesel can be blended with conventional petrodiesel in various proportions
1
because both have similar chemical properties and can be used in existing diesel engines without the need for
modification [8]. This makes it possible for biodiesel use to be deployed on a widescale using already existing
infrastructure for distributing petrodiesel.

Although biodiesel can be produced through synthesis routes such as pyrolysis, microemulsification, supercritical
fluid method, membrane process, ultrasound process, reactive distillation, transesterification of oils and fats is the
most common method of producing biodiesel [9]. It involves the reaction between the triglyceride present in the oil
or fat with a suitable alcohol such as methanol or ethanol in the presence of a catalyst to produce fatty acid alkyl
esters (FAAE), referred to as biodiesel [10]. During this process, glycerol is a byproduct that is removed from the
generated biodiesel, which is then purified further before storage and use.

Homogeneous and heterogeneous catalysts can be used to facilitate transesterification. The traditionally used
homogeneous catalysts are limited by their feedstock selectivity based on the fatty acid content, sensitivity to
feedstock moisture content, non-recoverability, and soap production, reducing biodiesel yield. The separation

of
challenges and potential corrosivity lead to increased production costs and negative environmental impacts [11].
This has necessitated the need for the use of heterogeneous catalysts. These catalysts exist in a different phase from

ro
the reaction components and are usually solid in nature. Solid base, solid acid, and bifunctional catalysts are types
of heterogeneous solid catalysts employed in biodiesel production. Specific examples include metal oxides, mixed
-p
oxides, and hydrotalcite for basic operation and transition metal oxides, ion exchange resin, calcium-based catalysts,
and zeolites for acidic operation [12].
re
Although calcium-based heterogeneous catalysts are effective, cheap, and improve the process economics of
biodiesel production, they are less active than homogeneous catalysts such as sodium hydroxide [13]. They also
lP

suffer from longer reaction times and leaching of the active phase into the reaction mixture which reduces the
efficiency of the catalyst when used in multiple cycles [14]. There have been attempts to overcome this in previous
work. For example, the use of a CaO-CeO2 mixed metal oxide solid catalyst to produce biodiesel from palm oil with
a

a yield of 95% was reported by Wong et al. [15]. The catalyst was stable over six cycles of use. In a similar work,
rn

Boro et al. [16] used a Ba-CaO-based catalyst to catalyze biodiesel production from waste cooking oil. They
obtained a biodiesel yield above 98%. Despite these results, metal oxide-based catalysts have been reported to have
u

a low surface area which limits their mass transfer characteristics and reduces their reuse capacity [17].
Jo

To address this, some authors have investigated the incorporation of biomass with calcium-based materials to
produce more active catalysts. It is thought that the high elemental content of some of these materials could enhance
the activity of the catalyst by increasing the basicity and reducing the leaching of calcium [18]. Furthermore, beyond
being inexpensive and readily available, the bio-based precursors are highly stable and possess desirable functional
groups for positive interactions with metals, carboxyl, hydroxyl, and aromatic groups which confers catalysts
produced from them with high functionality [19]. Additionally, their utilization avoids the need for significant
doping using expensive metal salts.

Biodiesel production is a multiparametric process dependent on multiple factors of which reaction time, reaction
temperature, type and amount of catalyst, alcohol-to-oil ratio, and stirring speed are the most significant factors
[20]. Biodiesel yield could potentially be improved if these factors are optimized. This can be done using statistical
methods like design of experiment (DOE). With DOE, the number of experimental runs is significantly lower for
the same number of input variables when compared with the one-factor-at-a-time method [21]. Recently, machine
learning and artificial intelligence tools like artificial neural networks (ANN) are now receiving attention for process

2
modeling as they are particularly suited for nonlinear chemical engineering processes and in situations where it is
difficult or impossible to develop normal empirical models to describe a process [22]. Furthermore, the model
parameter uncertainty is usually not taken into consideration in many biodiesel production studies. In this regard,
global sensitivity analysis (GSA) can be used to derive insights into critical model parameters that have a major
influence on model output and propose strategies to reduce the associated parametric uncertainty [23].

Based on the literature review done, no study has documented the combination of cow horn and waste marble tiles
to generate a heterogeneous catalyst to facilitate transesterification, modeled and optimized the process, and then
assessed model uncertainty and ranking of input variables using global sensitivity analysis. Cow horns are widely
accessible in Nigeria, because of the country's numerous abattoirs. It is also high in calcium, implying its potential
for catalyst synthesis. Waste marble tiles are abundant, particularly with the increase in construction activity in
Nigeria, and they now have little or no use. Combining two components to create the catalyst is likely to result in
some positive synergism. This multidimensional approach demonstrates the potential for efficient and sustainable

of
biodiesel production, with benefits including waste valorization and the advancement of green energy solutions.

2. Materials and Methods

ro
2.1. Feedstock and precursor collection and preparation

-p
The waste marble tiles were collected from a local tiles vendor in Benin City, Nigeria. The cow horn was obtained
from a local abattoir in Benin City. Both precursors were collected in July 2021. The waste marble tiles and cow
re
horn were cleaned to remove impurities and rinsed with warm water severally. They were then sun-dried for 4 days.
The waste cooking oil (WCO) used as feedstock was obtained from a diner in Benin City. It was filtered before use
lP

to eliminate solid contaminants. All other reagents used in this investigation were analytical grade and procured
from a vendor in Benin City.
2.2. Characterization of waste cooking oil
a

The suitability of the collected WCO as a feedstock for biodiesel production was examined by determining its
physical and chemical properties via appropriate characterization methods. The properties examined include
rn

moisture content, saponification value, acid value, peroxide value, viscosity, density, iodine value, and specific
gravity and fatty acid profile [24].
u

2.3. Catalyst preparation and characterization


Jo

The cleaned waste marble tiles and cow horn were crushed using a laboratory hammer mill and both were then
sieved to obtain 150 μm particles. The resulting waste marble tile and cow horn powder were separately kept covered
in a plastic container. They were later calcined separately at 900 oC for 4 hours in a muffle furnace (Fig. 1). The
composite catalyst made up of calcined marble tile and cow horn precursors which were then functionalized with
strontium and nickel to improve its reactivity was prepared by the wet impregnation method as reported by Chantara-
Arpornchai et al. [25]. For this procedure, the nickel nitrate, strontium nitrate, calcined cow horn, and calcined
waste marble tiles were mixed in the appropriate proportion and dissolved in 400 ml of deionized water followed
by constant stirring using a magnetic stirrer at a temperature of 80 oC for 5 hours. The resulting mixture was filtered
to obtain a solid residue which was dried at 110 oC for 12 hours in an oven. Calcination of the dried residue was
carried out at 1000 oC for four hours in a muffle furnace and then cooled in a desiccator. The calcined catalyst was
then crushed to obtain a fine powder which was kept in an air-tight container.

3
Fig. 1: Schematic of the catalyst preparation process
The evaluation of the properties of the produced catalyst was carried out through various characterization tests. A
scanning electron microscope (SEM) coupled with energy dispersive X-ray (EDX) detector was utilized to

of
investigate the morphology and elemental composition of the catalyst. The crystallinity of the catalyst was assessed
using X-ray diffraction with Cu-Kα radiation using an X-ray diffractometer. Infrared spectra were obtained using

ro
Fourier transform infrared spectroscopy in the range of 4000–400 cm−1 (Thermo-Nicolet FT-IR equipped with
attenuated total reflectance). X-ray fluorescence provided information on the oxides present in the catalyst. The

-p
Brunauer–Emmett–Teller (BET) analysis was used to determine the catalyst surface area, while the pore volume
and pore diameter were determined using the Barrett–Joyner–Halenda (BJH) method. The thermal stability of the
re
catalyst was assessed using thermo-gravimetric analysis (TGA) and differential thermal analysis (DTA). Similar
tests were carried out for the precusors and the spent catalyst.
lP

2.4. Biodiesel production and characterization


As a result of the observed bifunctionality of the catalyst, a one-pot process was used to prepare biodiesel in a sealed
250 ml Erlenmeyer flask. The catalyst was first dried in an oven at 100 °C for 2 h and then dispersed in methanol
a

at 50 °C under constant stirring for 30 min. This was followed by the addition of the required quantity of waste
cooking oil which had been preheated to the reaction temperature [26]. Yield-dependent variables such as
rn

temperature, time, catalyst concentration, and methanol-to-oil ratio were fixed according to the experimental design.
Upon completion of the reaction, the used catalyst was separated from the products using a filter cloth. A separating
u

funnel was used to separate the different products by allowing it to stand for 24 hours for separation to take place.
Jo

The yield of biodiesel produced was determined according to Equation 1 [27].


mass of biodiesel produced
Biodiesel yield  100 (1)
mass of waste cooking oil used

Some fuel properties like flash point, pour point, and cetane number and physicochemical properties like acid value,
viscosity, free fatty acid, iodine value, density, and fatty acid profile) were evaluated according to standards.

2.5. Design of experiment and RSM modeling


A Box-Behnken design (BBD) was used to plan and carry out the process to produce biodiesel. The factors
considered were reaction time, reaction temperature, catalyst concentration, and methanol-to-oil ratio, and their
levels were fixed based on preliminary experiments and information from the literature sources as shown in Table
1 [28]. The choice of the BBD was a result of its advantages which include a smaller number of experimental runs,
minimized experimental errors as a result of the distribution of the design points, capacity to fit quadratic response
surfaces, statistical efficiency in estimating model parameters, and adequate interpolation properties [22]. The

4
design was implemented in the Design Expert® 7.0.0 software and resulted in a total of 29 experiments that were
randomized to eliminate bias.

Table 1: Range of input variables


Levels
Variables Symbol Unit
-1 0 +1
o
Temperature X1 C 50 65 80
Catalyst concentration X2 % 3.0 6.5 10.0
Time X3 min 60 105 150
Methanol to oil ratio X4 - 3:1 8.5:1 14:1

Equation 2 was adopted to fit the experimental data to predict biodiesel yield. The significance of the model terms
was determined using analysis of variance (ANOVA).

of
N N N N
Y  bo   bi X i   bij X i X j   bii X i2   ei

ro
i 1 i , j 1 i 1 i 1
(2)
where Yi is the response, Xi and Xj are input factors, bi and bij are single and interaction term coefficients, and bo and
ei are offset and error terms.
2.6. ANN modeling
-p
re
Artificial neural networks (ANN) was used to model the biodiesel production process. This was accomplished using
NeuralPower, version 2.5. Since it is not always possible to know apriori, the most suitable network for modeling a
process, it is usually useful to investigate different network options. For this work, the dependent variable (biodiesel
lP

yield) was predicted using a multilayer full feed-forward (MFFF) and a multilayer normal feed-forward (MNFF)
neural network to determine which was more suitable. Furthermore, because it is also not possible to know before
a

hand, the best training algorithm for a network, it is also useful to evaluate several training algorithms to determine
which is most suitable for a particular process. Consequently, two network architectures were chosen and trained
rn

using different training algorithms using different transfer functions and number of neurons. For the data split, 70%
was used for training the model, while 15% was used for validating and testing the model. The training algorithm
u

that best describes the biodiesel production process was selected based on their R2 and RMSE values. [29].
Jo

2.7. Comparison of RSM and ANN performance

The accuracy of the RSM and ANN predictions was assessed using statistical metrics such as coefficient of
determination, adjusted coefficient of determination (adjusted R2), mean square error (MSE), root mean square error
(RMSE), mean absolute error (MAE) and average absolute deviation (AAD) as outlined in Equations 3 to 8 [30].

2
∑𝑛
𝑖=1(𝑥𝑎,𝑖 −𝑥𝑝,𝑖 )
𝑅2 = 1 − 2 (3)
∑𝑛
𝑖=1(𝑥𝑝,𝑖 −𝑥𝑎,𝑎𝑣𝑒 )

𝑛−1
𝐴𝑑𝑗𝑢𝑠𝑡𝑒𝑑 𝑅2 = 1 − [(1 − 𝑅2 ) × ] (4)
𝑛−𝑘−1

1 2
𝑀𝑆𝐸 = ∑𝑛𝑖=1(𝑥𝑝,𝑖 − 𝑥𝑎,𝑖 ) (5)
𝑛

1 2
𝑅𝑀𝑆𝐸 = √ ∑𝑛𝑖=1(𝑥𝑝,𝑖 − 𝑥𝑎,𝑖 ) (6)
𝑛

5
1
𝑀𝐴𝐸 = ∑𝑛𝑖=1|(𝑥𝑎,𝑖 − 𝑥𝑝,𝑖 )| (7)
𝑛

1 |(𝑥𝑎,𝑖 −𝑥𝑝,𝑖 )|
𝐴𝐴𝐷 = (∑𝑛𝑖=1 ( )) × 100 (8)
𝑛 𝑥𝑎,𝑖

where n, 𝑥𝑝,𝑖 , 𝑥𝑎,𝑖 , 𝑥𝑝,𝑎𝑣𝑒 , 𝑥𝑎,𝑎𝑣𝑒 , and k represent the number of experiments, predicted values, actual values, mean
predicted values, mean actual values, and number of input variables respectively.

2.8. Optimization of biodiesel production

Biodiesel yield was optimized using genetic algorithm (GA) which was done using MATLAB R2015a GA toolbox.

2.9. Global sensitivity analysis

Global sensitivity analysis was used to identify the impact of each input variable on the prediction of biodiesel

of
production. The high-dimensional model representation (HDMR) method implemented in SobolGSA software
(Centre for Process Systems Engineering, Imperial College London) was used for this purpose. Unlike local

ro
sensitivity analysis, which only computes the effects of input features at a specified point, global sensitivity analysis
computes the effects of input features over the whole space. This means that the magnitude of the range over which

-p
each input feature is varied has a direct influence on the sensitivity to that input in the global sensitivity analysis
[31]. Global sensitivity indices were calculated using the HDMR coefficients. For the HDMR process, the complete
re
response function is decomposed into a sum of functions that are dependent on subsets of the inputs, as shown in
Equation 9:
lP

y  f x
N N N (9)
 fo   fi  xi     fij  xi , x j   ...  f1,2...,N  x1 , x2 ...x N 
i 1 i 1 j i 1
a

where N, i, j, and fo represent the number of input features, index of input features, and average value of f(x),
rn

respectively. Following Equation 9, the contribution of each term, i.e.,  i2 and  ij2 , to the variance of the output

response is related to the total variance (  2f ) as shown in Equation 10:


u
Jo

N N N
 2f   1 fi2  xi dxi    1 1 fij2  xi , x j dxi dx j
1 1 1

i 1 i 1 j i 1
N N N
(10)
  i2     ij2
i 1 i 1 j i 1

The linear, interaction and total sensitivity indices were obtained as shown in Equations (11), (12), and (13) which
were used to assess the relative level of importance of the input variable to the variance of the output response.

 i2
Si  (11)
 2f

 ij2
Sij  (12)
 2f

STi  Si   Sij (13)


j i

6
2.10. Catalyst reusability studies

The catalyst was repeatedly used to catalyze several transesterification reactions for several cycles to assess its
reusability. At the end of each experiment, the used catalyst was recovered, washed with n-hexane before being
oven dried at 80 oC for 12 hours and then weighed. The yield of biodiesel produced for each successive reaction run
was recorded.

3. Results and Discussion

3.1. Characterization of waste cooking oil


The utilization of waste cooking oil as a feedstock for biodiesel production has great potential due to its abundance
and potential to mitigate environmental impact. Furthermore, it does not have the limitation of suffering from
competition associated with edible oil feedstocks. The results shown in Table 2 give a comprehensive overview of
the properties of waste cooking oil. The waste cooking oil had an acid value of 10.52 mg KOH/g oil, indicating the

of
presence of free fatty acids (FFA). This suggests a certain degree of hydrolysis and oxidation, which could hinder
the efficiency of biodiesel production [32]. The FFA content was obtained as 5.28%, and this further confirmed the

ro
presence of free fatty acids. High levels of FFA are not desirable as it can result in the formation of soap during
biodiesel production via transesterification, thus, reducing both the quality and yield of biodiesel [1]. The WCO

-p
moisture content was determined to be 0.61%. Moisture content is very important because excessive moisture levels
promotes microbial proliferation during the storage of oil and may interfere with biodiesel synthesis during
re
transesterification [33]. Viscosity, an important parameter for biodiesel production, was found to be 38.03 mPas at
40 °C. High viscosity values indicate increased resistance to flow, which can influence the reaction kinetics during
transesterification [34]. The saponification value, measured at 207.10 mg KOH/g oil, provides insights into the oil's
lP

potential for biodiesel production. High saponification values suggest a larger proportion of higher molecular weight
fatty acids, which can contribute to improved energy content and enhanced cold flow properties in the resulting
a

biodiesel [35]. The density was obtained as 931 kg/m3. The density of oil plays an important role in the separation
and purification processes of biodiesel production. These properties can influence phase separation and the removal
rn

of glycerol and impurities from biodiesel [36]. The iodine value (82.90 mg I2/100g oil) is a measure of the oil's level
of unsaturation. Unsaturated fatty acids contribute to the oxidative stability and cold-flow properties of biodiesel
u

[37]. Waste cooking oil with a high iodine value demonstrates a greater potential for blending with other feedstocks
Jo

to optimize the final biodiesel properties. For any oil feedstock and in particular, waste cooking oil, the fatty acid
composition is a crucial factor that can influence its suitability for biodiesel synthesis. Table 3 shows the fatty acid
composition of the WCO obtained via GC-MS analysis.

Table 2: Properties of waste cooking oil used for biodiesel production


Property Value
Moisture content (%) 0.61
Acid value (mg KOH/g oil) 10.52
FFA (%) 5.28
o
Dynamic viscosity @ 40 C (mPa.s) 38.03
Saponification value (mg KOH/g oil) 207.10
o 3
Density @ 40 C (kg/m ) 931.00
Average molecular weight (g/mol) 856.09
Specific gravity 0.931

7
Iodine value (mg I2/100 g oil) 82.90

Table 3: Fatty acid composition of waste cooking oil used for biodiesel production
Chemical
Fatty acid Nature Composition (%)
formula
Myristic C14H28O2 Saturated 1.10
Linoleic C18H32O2 Unsaturated 11.58
Oleic C18H34O2 Unsaturated 43.33
Palmitic C16H32O2 Saturated 36.01
Lauric C12H24O2 Saturated 0.20
Arachidic C20H40O2 Saturated 1.10
Stearic C18H36O2 Saturated 3.88
Palmitoleic C16H30O2 Unsaturated 0.47
Others 2.33
Total 100

of
Oleic acid, which accounts for 43.33% of the total composition, is a monounsaturated fatty acid. Its composition in
WCO is beneficial for biodiesel manufacturing. It improves fuel and cold flow properties. Overall, oleic acid

ro
enhances the quality of biodiesel and its performance in diesel engines [38]. The second most abundant fatty acid,
palmitic acid is a saturated fatty acid and accounted for 36.01% of the total composition. Like oleic acid, it also
-p
positively impacts the fuel properties of biodiesel made from waste cooking oil [39]. Linoleic acid accounted for
11.85% of the fatty acid profile of the oil. Its presence in waste cooking oil suggests a high degree of unsaturation,
re
which might impact the oxidative stability of the resultant biodiesel. Unsaturated fatty acids increase cold flow
properties and oxidation stability, making linoleic acid an important component of oils for biodiesel production [40].
lP

The other fatty acids, lauric acid, and arachidic acid are all saturated and in low proportions, and they also contribute
to the cold flow properties of biodiesel.

3.2. Catalyst characterization


a

3.2.1. SEM analysis


rn

The morphological characteristics of the catalyst precursor and the synthesized catalyst before and after use for
u

transesterification are shown in Fig. 2. These characteristics obtained from scanning electron microscopy analysis
provide valuable insights into the catalyst precursor (composed of cow horn and waste marble tiles) and the
Jo

synthesized catalyst, both before and after usage in the transesterification process. The SEM images clearly show
that both the catalyst precursor (Fig. 1a) and the calcined composite catalyst (Fig. 2b and c) have irregularities in
their catalytic pores. These irregularities are caused by the heterogeneous nature of the catalyst materials, which
vary in composition and surface properties. Despite this irregularity, the pore size of the catalysts remained generally
uniform, with the unused calcined bifunctional catalyst showing a modest increase in pore size (Fig. 2b). The
existence of residual carbonate, which could not be entirely removed during the calcination process, may be
responsible for the increase in pore size.

8
Fig. 2 SEM images of (a) raw precursors (b) catalyst before transesterification reaction (c) catalyst after
transesterification reaction

of
Comparing the SEM image of the catalyst precursor with that of the catalyst before usage, noticeable differences in
pore structure and overall morphology are observed. The calcination process, which involves high-temperature

ro
treatment, efficiently eliminates bound gases and moisture content from the pores of the precursor. This gives the
calcined catalyst larger and more well-defined pores compared to the raw precursor, increasing the specific surface
area. Increased specific surface area and pore size improve catalyst action by generating more active sites for
-p
reactive species to attach during the transesterification process. The calcination temperature is extremely important
in shaping the morphology of the catalyst [41]. The duration and temperature of calcination can influence the
re
development of primary catalytic characteristics such as basic and acidic site density, pore size and volume, surface
area, and the crystalline and molecular structure of the catalyst [42]. To derive the requisite catalytic characteristics
lP

for transesterification processes, careful and purposeful adjustment of the calcination process is required. When the
SEM images of the calcined composite catalyst before (Fig. 2b) and after usage (Fig. 2c) for transesterification
were compared, it was found that there was a considerable decrease in the catalyst's porosity. This reduction in
a

porosity can be possibly attributed to a decrease in catalyst surface area induced by reacting species occupying the
rn

active sites on the catalyst’s surface in the course of the transesterification process [43]. When a catalyst is utilized,
its active sites may get occupied, resulting in a loss of porosity and this can cause its activity to decrease. This
u

behavior is similar to that reported in the previous research by Amenaghawon et al. [26], who discovered decreased
catalyst activity and active site deactivation following reuse.
Jo

3.2.2. XRF analysis


XRF analysis was used to assess the chemical compositions of both the catalyst precursor and the produced catalyst
(before and after use). The results are shown in Table 4 and they show the presence of key oxides such as CaO,
SiO2, K2O, Al2O3, and P2O5 in the precursor and synthesized catalyst. The properties of these oxides enable them
to play important roles in catalyzing biodiesel production. For example, CaO and K2O are known for their basic
properties. The presence of SrO and NiO in the catalyst could be explained by the doping carried out during the
preparation stage. The goal of doping with nickel and strontium was to enhance the acidic and basic strength
respectively of the catalyst and the presence of oxides of both elements in the prepared catalyst suggests that the
doping process was successful. The deliberate functionalization of the catalyst enhances its performance by
introducing additional catalytic sites that promote both esterification and transesterification [44]. This is particularly
useful when using WCO that contains high FFA levels. It should be noted that SiO2 enhances the stability of the
catalyst. SiO2 can serve as a support and in this case, provide the catalyst with the needed mechanical stability and
the accompanying high surface area needed for solid-catalyzed reactions [6].

9
Table 4: XRF results of the precursor and synthesized catalyst (before and after use)
Concentration (%)
Component
Precursor Catalyst (before reaction) Catalyst (after reaction)

CaO 39.998 26.658 25.88

Fe2O3 4.010 4.172 4.077

SiO2 29.329 32.114 31.150

K2O 4.227 5.821 5.696

Al2O3 5.219 6.737 6.007


NiO 0.000 6.105 5.290
SrO 0.107 8.297 7.589
Co3O4 0.024 0.027 0.021

of
MoO3 0.006 0.011 0.007

P2O5 14.342 9.084 8.784

ro
CuO 0.071 0.082 0.079

Nb2O3 0.014 0.025 0.018

MgO

Cr2O3
0.000

0.036
-p
0.107

0.032
0.000

0.024
re
BaO 0.142 0.221 0.208

WO3 0.000 0.08 0.076


lP

TiO2 0.373 0.347 0.336

ZnO 0.022 0.000 0.008


a

Ag2O 0.019 0.025 0.055


rn

ZrO2 0.064 0.055 0.132

Others 1.997 0.000 4.563


u

3.2.3. XRD analysis


Jo

The crystallinity or otherwise as well as the phase transformation of both the precursor and the catalyst was assessed
using the XRD test. The results obtained are shown in Fig. 3 which shows that for the precursor, sharp peaks were
observed at diffraction angles of 20.96°, 24.92°, 26.73°, 27.64°, 31.75°, and 39.66° suggesting the presence of
specific crystalline phases. In the same vein, Fig. 2 shows that for the pristine catalyst (before the reaction), peaks
were found at 20.96°, 22.11°, 24.15°, 26.75°, 28.00°, 34.37°, 42.52°, and 51.25° suggesting the presence of CaO,
SiO2, K2O, NiO, and CaO respectively. In comparing the diffraction patterns of the precursor and that of the catalyst
before usage, it was noted that there was an increase in the intensity of peaks in the catalyst. This could be an
indication of an increase in the catalytic activity of the catalyst, potentially due to the generation of additional active
sites due to calcination. It was reported previously that one of the effects of calcination on the catalyst is the solid-
state transformation reactions which ultimately result in an improvement in catalyst activity [45]. A comparison
between the diffraction patterns of the calcined catalyst before and after usage showed that there was a decrease in
the intensity of the peaks of the used catalyst. This could have been caused by a reduction in the catalyst's activity

10
as the reaction progressed. Similar findings have been reported by other researchers, who noted that the deactivation
of active sites on the surface of the catalyst can occur in the course of the reaction [46].

of
ro
-p
Fig. 3 XRD pattern of the precursor, pristine catalyst before and after transesterification reaction

3.2.4. Surface and pore analysis


re
The surface and pore characteristics of the catalyst and its precursor are presented in Table 5. The precursor had a
lP

surface area, pore diameter, and pore volume of 121.548 m2/g, 2.122 nm, and 0.075 cc/g respectively. Conversely,
the catalyst had a higher surface area and pore volume of 301.510 m2/g and 0.165 cc/g respectively, with a slightly
reduced pore diameter of 2.110 nm. The improved surface characteristics recorded for the composite catalyst could
a

have resulted from the calcination process which essentially adsorbed gases and moisture content, thereby
generating more pores and increasing the available surface area and pore volume of the catalyst [47]. In a previous
rn

work, Ghampson et al. [48] reported that calcination can cause changes in particle size and consequently affect the
surface area. The higher surface area enhances catalytic activity by improving the diffusion of reactants to the active
u

sites. These observations are in line with those of previous works such as Amenaghawon et al. [26] who utilized
Jo

doped calcined cow-horn for biodiesel production from waste cooking oil, and Chen et al. [49] who used a
mesoporous silica-based catalyst to catalyze biodiesel production.

Table 5: Surface characteristics of the precursor and synthesized catalyst (before use)
Parameter Precursor Pristine catalyst (before reaction)
2
Surface area (m /g) 121.548 301.510
Pore diameter (nm) 2.122 2.110
Pore volume (cm3/g) 0.075 0.165

3.2.5. FTIR analysis


The FTIR spectra of the prepared catalyst and its precursor are shown in Fig. 4. The result shown in Fig. 4 for the
precursor indicated major peaks at 2851.4 and 2922.2 cm−1 which were attributed to the stretching of the C–H bond
while the peak at 3295.0 cm−1 was attributed to the vibration of the O–H bond. There were also peaks at 764.1 cm−1,
1457.4 cm−1, 1740.7 cm−1, which corresponded to the stretching of the C–Cl bond, bending of the C–H bond, and
stretching of the C=O bond respectively. For the fresh composite catalyst shown in Fig. 3, the major peaks were
found at 3630.4 cm−1 which corresponded to O–H stretching while the peaks at 1021.3 cm−1, 771.6 cm−1, and 726.8

11
cm−1 were attributed to the stretching of C–N, C-Cl and C–H bonds. For the used catalyst shown in Fig. 3, the peaks
were found at 2922.2 cm−1, 2855.1 cm−1, 1744.5 cm−1, 1017.6 cm−1, and 723.1 and these were attributed to the C–H
stretching, C–H stretching, C=O stretching, C–H bending and C–H rocking respectively. Similar observations were
reported by Yusuff et al. [50] for a solid catalyst prepared using egg shell and anthill and Amenaghawon et al. [26]
for a heterogeneous catalyst prepared from cow horn.

of
ro
-p
re
a lP
u rn
Jo

12
of
ro
-p
re
a lP
u rn
Jo

Fig. 4 FTIR spectra of the precursor, pristine catalyst before and after transesterification reaction

13
3.3. RSM model analysis
The actual statistical modeling was done by data fitting of the quadratic model to estimate the unknown model
coefficients/parameters and the process was achieved using multiple regression analysis [51]. When this was
completed, the estimated model coefficients were inputted into the general quadratic model to obtain the final
statistical model (Equation 14) which relates biodiesel yield to the independent variables. Table 6 shows a
comparison of the actual biodiesel yield with the model prediction and there was good agreement between them
showing the accuracy of the RSM model developed to predict biodiesel yield.

Biodiesel yield (%)  184.85  2.48 X1  24.65 X 2  1.69 X3  4.76 X 4


 0.078 X1 X 2  0.0049 X1 X3  0.012 X1 X 4  0.087 X 2 X3
2 2 2 (14)
 0.32 X 2 X 4  0.015 X3 X 4  0.0089 X1  0.64 X 2  0.0031 X3
2
 0.0021 X 4

of
Table 6: Experimental and model-predicted results for biodiesel yield

ro
Actual values of input factors Biodiesel yield (%)

Run Experimental RSM ANN


X1 X2 X3 X4

1 65 3.0 105 14.0


-p
observation

97.0
prediction

96.8
prediction

96.8
re
2 50 3.0 105 8.5 79.2 79.0 79.0
3 65 3.0 60 8.5 67.5 66.3 67.9
lP

4 80 6.5 105 3.0 96.6 93.4 96.7


5 65 6.5 105 8.5 93.4 93.9 94.0
6 65 10.0 150 8.5 66.2 66.0 66.1
a

7 50 6.5 60 8.5 77.4 77.4 77.4


rn

8 50 6.5 150 8.5 87.2 87.9 87.1


9 65 10.0 105 14.0 79.6 80.4 79.6
u

10 80 3.0 105 8.5 92.4 93.2 92.4


11 65 6.5 150 3.0 90.6 90.8 91.3
Jo

12 80 10.0 105 8.5 80.6 80.9 80.6


13 65 6.5 105 8.5 93.4 93.9 94.1
14 80 6.5 60 8.5 89.4 90.0 89.4
15 65 3.0 105 3.0 79.0 79.6 79.0
16 65 6.5 60 3.0 77.4 79.7 77.0
17 80 6.5 150 8.5 86.0 87.3 86.0
18 65 10.0 105 3.0 86.2 87.7 85.9
19 65 6.5 105 8.5 94.2 93.9 93.5
20 65 6.5 60 14.0 92.0 91.9 92.1
21 65 10.0 60 8.5 91.2 89.6 91.1
22 50 10.0 105 8.5 83.8 83.1 84.0
23 65 6.5 150 14.0 90.8 88.6 91.0

14
24 65 6.5 105 8.5 94.8 93.9 93.6
25 50 6.5 105 3.0 87.0 85.5 87.0
26 80 6.5 105 14.0 96.4 96.5 96.0
27 50 6.5 105 14.0 90.6 92.3 91.0
28 65 3.0 150 8.5 97.4 97.6 97.3
29 65 6.5 105 8.5 93.6 93.9 94.0

The ANOVA results presented in Table 7 showed a very significant model p-value (p < 0.0001) showing that it was
very useful for predicting biodiesel yield. The correlation between the lowest p-values (<0.0001) and the highest F-
values (49.63) in the ANOVA results highlights the statistical significance of the RSM model for biodiesel
production. A low p-value suggests that the observed effects are not likely due to random chance, while a high F-
value indicates significant differences among group means. In this study, variables with the lowest p-values and

of
highest F-values hold considerable significance, emphasizing their substantial influence on biodiesel yield and
affirming the reliability of the model. The model also showed a nonsignificant lack of fit (p = 0.1918). The low CV
value (1.91) indicates that the experimental runs are reliable and repeatable [52]. The adequate precision value

ro
obtained (26.341) was greater than the value of 4 typically recommended showing that the model has an adequate
signal-to-noise ratio. The model was characterized by a high R2 and adjusted R2 values of 0.9802 and 0.9605
-p
respectively showing good agreement between model predictions and actual experimental values (Table 8).
Table 7: ANOVA results for the RSM model representing biodiesel yield
re
Source SS df MS F value p-value
Model 1933.34 14 138.10 49.63 < 0.0001
lP

X1 109.20 1 109.20 39.25 < 0.0001


X2 51.67 1 51.67 18.57 0.0007
X3 45.24 1 45.24 16.26 0.0012
a

X4 73.01 1 73.01 26.24 0.0002


rn

X1X2 67.24 1 67.24 24.17 0.0002


X1X3 43.56 1 43.56 15.66 0.0014
u

X1X4 3.61 1 3.61 1.30 0.2738


X2X3 753.50 1 753.50 270.81 < 0.0001
Jo

X2X4 151.29 1 151.29 54.37 < 0.0001


X3X4 51.84 1 51.84 18.63 0.0007
X12 26.23 1 26.23 9.43 0.0083
2
X2 397.00 1 397.00 142.68 < 0.0001
X32 249.21 1 249.21 89.56 < 0.0001
2
X4 0.03 1 0.03 0.01 0.9233
Residual 38.95 14 2.78
Pure error 1.49 4 0.37
Lack of fit 37.47 10 3.75 10.07 0.1918
Corrected totals 1972.30 28
C.V 1.91
Adequate precision 26.34

15
Table 8: Performance comparison of RSM and ANN models

Parameter RSM ANN

R2 0.9802 0.9993

Adjusted R2 0.9605 0.9977

MSE 1.3183 0.1460

RMSE 1.1482 0.3821

MAE 0.8517 0.2534

AAD (%) 0.9814 0.2842

3.4. ANN model analysis


The choice of the most suitable transfer function is essential as it is responsible for mapping the input signals to the

of
output response and this process will be unsuccessful if the appropriate transfer function is not chosen [53]. In this
regard, the hyperbolic-tangent, Gaussian, sigmoid, linear, bipolar linear, and threshold linear transfer functions were

ro
assessed for their suitability in the ANN model for biodiesel production and the results are shown in Table 9. The
best transfer function for developing the ANN model was found to be the hyperbolic tangent because it was

-p
characterized by the highest R2 value (0.99368) and lowest RMSE value (0.6673) compared to the others. Some
previous studies have also corroborated the fact that the hyperbolic tangent transfer function is suitable for building
re
artificial neural networks [54, 55]. To identify the neural network's training algorithm, different training methods
were evaluated, including the Levenberg-Marquardt algorithm, genetic algorithm, batch back propagation,
incremental back propagation, and quick propagation, and the results are displayed in Table 10. The Levenberg-
lP

Marquardt algorithm was best for training the network (R2 = 0.99925; RMSE = 0.23053) when compared to the
other training algorithms. The determination of the optimal number of neurons is critical as using too many neurons
a

might lead to overtraining of the network while using too few neurons will lead to a situation referred to as
underfitting [56]. Thus, in this study, the ANN was built with varying numbers of neurons, and the influence on the
rn

network's predictive ability was evaluated using the R2 values. As seen in Fig. 5, increasing the number of neurons
from one to four enhanced the R2 value. This indicates that the network's prediction was improving. However,
u

increasing it beyond four did not improve the R2 value, making the optimal number of neurons to be four.
Jo

Table 9: Selection of best transfer function


Transfer function R2 RMSE
Sigmoid 0.34397 6.7978
Gaussian 0.95042 1.8688

Linear 0.02957 8.2678


Hyperbolic tangent 0.99368 0.6673
Threshold linear 0.01111 8.4393

Table 10: Selection of best network configuration and training algorithm


Network architecture Learning algorithm R2 RMSE
LM 0.99923 0.23217
MNFF IBP 0.96173 1.64200
QP 0.95554 1.76960

16
BBP 0.99196 0.75244
GA 0.80134 3.74080
LM 0.99925 0.23053
IBP 0.96822 1.49630
MFFF QP 0.9633 1.60790
BBP 0.92864 2.24200
GA 0.92327 2.32490

1.0

0.9

of
R2 value

0.8

ro
0.7

0.6
-p
re
0 1 2 3 4 5
Number of neurons
lP

Fig. 5 Selection of the optimum number of neurons

Two network types, multilayer normal feed-forward (MNFF) and multilayer full feed-forward (MFFF), with their
a

transfer function (hyperbolic tangent), training algorithm (Levenberg-Marquardt algorithm), and number of neurons
rn

(four) were evaluated to ascertain which was the best configuration. Both network types were specified with two
hidden layers, both with the same number of neurons. The MFFF network trained with the Levenberg-Marquardt
algorithm gave the best representation of the biodiesel production process (R2 = 0.99925; RMSE = 0.23053). Thus,
u

the chosen configuration for an ANN with four input factors, four neurons in both hidden layers, and one output
Jo

layer for predicting biodiesel yield is 4-4-4-1 as shown in Fig. 6.

17
of
ro
-p
Fig. 6 Optimum ANN configuration for modeling biodiesel yield

The ANN model's predictions were compared to the results of the actual experiment (Table 6) and there was marked
re
similarity between both, showing the ANN model's reliability and accuracy. The goodness of fit results for the ANN
model representing biodiesel yield are shown in Table 8. According to the results, the ANN prediction has a very
lP

high R2 and adjusted R2 (R2 and adjusted R2 > 0.99). This demonstrates that the ANN model predictions and
experimental values were nearly exact.

3.5. Comparison of the predictive performance of the RSM and ANN models
a

The prediction accuracy of the RSM and ANN models was evaluated and compared to determine which was better
rn

suited for modeling biodiesel yield. As shown in Table 8, the goodness of fit metrics used to assess the performance
of the models such as the R2 value and adjusted R2 value, as well as error functions such as MAE, MSE, AAD, and
u

RMSE showed their respective predictive capacities. Although the RSM model and ANN model had very high R2
and adjusted R2 values, the ANN model was superior since it had higher R2 (0.9993) and adjusted R2 (0.9977) values
Jo

compared to RSM, indicating a better fit with the experimental data. Furthermore, ANN had lower MSE (0.1460),
RMSE (0.3821), MAE (0.2534), and AAD (0.2842) values than RSM, indicating that the ANN predictions were
more accurate. The parity plots in Fig. 7 compare the RSM and ANN predictions to the data from the experiments.
ANN provided a superior fit since the data points were all practically on the 45o line, as compared to the RSM
model, which had some data dispersion about the 45o line. Many researchers have reported that ANN outperforms
RSM in modeling biodiesel production [2, 26, 57, 58]. This demonstrates that the ANN model can be implemented
for biodiesel production since it provides a more accurate picture of actual biodiesel production conditions. Although
artificial neural networks are commonly known for their ability to handle large datasets, their effectiveness extends
beyond the amount of data available. In this work, the dataset comprised 29 data points, but ANN still showed
exceptional predictive capabilities, capturing the complex relationships between the input and output variables. This
demonstrates the model's flexibility and its capacity to extract valuable insights even from relatively small datasets.
The results of this study attest to the efficacy of ANN in extrapolating meaningful information from available data.

18
100 100

Predicted biodiesel yield (%)

Predicted biodiesel yield (%)


90 90

80 80

70 70

60 60
60 70 80 90 100 60 70 80 90 100
Actual biodiesel yield (%) Actual biodiesel yield (%)

of
(a) (b)
Fig. 7 Parity plot comparing experimental biodiesel yield with (a) RSM prediction and (b) ANN prediction

ro
3.6. Effect of input variables using response surface plots
The combined interactive influence of temperature and time on the yield of biodiesel is presented in Fig. 8a. The
-p
trend observed shows that increasing the reaction temperature within the specified experimental range caused an
increase in the biodiesel yield. This observation would suggest that higher temperatures will favor biodiesel
re
production. This is in line with the principles of the Arrhenius equation, which predicts an increase in reaction rates
with increasing reaction temperatures [59]. This trend could also be attributed to the higher kinetic energy of the
lP

triglyceride molecules, which promotes more frequent and effective collisions, ultimately accelerating the reaction
rate [60]. A similar trend of increase in biodiesel yield with reaction temperature has also been reported by previous
researchers [61–63]. In addition, Fig. 8a also shows that increasing the reaction time caused an increase in biodiesel
a

yield. This observation can be attributed to the progressive conversion of triglycerides present in the oil to fatty acid
rn

methyl esters, which are the main constituents of biodiesel [64]. Previous studies have reported a similar trend and
suggested that allowing sufficient time for the transesterification reaction to take place facilitates a more complete
conversion of triglycerides and thus increases the biodiesel yield [65, 66].
u

Fig. 8b shows the influence of the methanol-to-oil ratio on the biodiesel yield within the range of 3 to 14. There is
Jo

a clear trend showing that increasing the methanol-to-oil ratio caused a progressive increase in biodiesel yield. When
the amount of methanol is increased relative to the oil, the availability of more alcohol molecules for reaction results
in a greater conversion of triglycerides and, consequently, a higher biodiesel yield [67]. However, it should be noted
that even though a higher methanol-to-oil ratio results in the production of more biodiesel, the use of excessive
amounts of methanol can also have certain drawbacks. For instance, excess methanol may cause separation
problems during the separation and purification steps, as well as higher energy requirements for the subsequent
removal of excess alcohol from the biodiesel product [68].

19
of
ro
-p
re
lP

Fig. 8 Effect of (a) reaction time and temperature, (b) methanol-oil ratio and temperature, and (c) reaction
temperature and catalyst concentration on biodiesel yield

Fig. 8c shows the dependence of biodiesel yield on the catalyst concentration in the range of 3 to 10 wt%. The trend
a

observed shows that the catalyst concentration produces a notable effect on the biodiesel yield, with an optimum
rn

yield achieved at a catalyst concentration of 5.12 wt% beyond which biodiesel production was not favored. This
decrease in yield can be attributed to the unfavorable mixing characteristics that arise in the reaction mixture
u

containing both solid catalyst and the polar (methanol) and nonpolar (oil and biodiesel) phases [69]. Furthermore,
with the use of higher catalyst concentrations, the solid catalyst particles tend to aggregate, leading to poor
Jo

dispersion and limited contact between the catalyst and the reactants. As a result, mass transfer between the separate
phases becomes limited and this hinders the efficient conversion of triglycerides to biodiesel [70].

3.7. Optimization of biodiesel production


The optimization results gave an optimum biodiesel yield of 98% for a temperature, catalyst concentration, time,
and methanol to oil ratio of 74.86 oC, 5.12 wt%, 119.92 min, and 11.42:1, respectively (Table 11). The optimized
results were confirmed by performing triplicate experiments under the optimal conditions. The mean biodiesel yield
of these three replicates was 98.13% which was not very different from the predicted optimum value of 98.0%. The
high biodiesel yield of 98% obtained in this study could be linked to the composition of the catalyst. Being composed
mainly of CaO and some appreciable amounts of K2O, Fe2O3, Al2O3, SiO2, SrO, and NiO, the catalyst exhibits a
balanced blend of acidic and basic components derived from the precursors used. This bifunctional quality enables
it to effectively catalyze both acid and base-catalyzed transesterification reactions, optimizing the conversion of
WCO into biodiesel. The synergy between the catalyst's different components creates an ideal environment for the

20
transesterification process, resulting in a high biodiesel yield of 98%. There was general agreement between the
optimized conditions and previous research (Table 12).
Table 11: Optimization results
Variable Value

Reaction temperature 74.86 oC

Catalyst concentration 5.12 wt%

Reaction time 119.92 min

Methanol-oil ratio 11.43:1

Optimum biodiesel yield 98.0%

Table 12: Comparison of biodiesel production conditions with previous studies


Reaction Catalyst

of
Catalyst Reaction Methanol Biodiesel
Feedstock temperature loading Reference
source o time (min) to oil ratio yield (%)
( C) (wt%)
Waste

ro
Waste Farooq and Ramli
chicken 65 5 240.00 15:1 89.3
cooking oil [71]
bone
Waste Mahesh et al.
KBr/CaO 65 3 108.00 12:1 83.0
cooking oil
Waste palm
oil
Chicken
egg shell
80 3
-p
180.0 15:1 90.1
[72]
Mansir et al. [73]
re
Crab shell
Waste and Amenaghawon et
60 5 149.94 13.01:1 93.0
cooking oil plantain al. [2]
peels
lP

Rubber
Egg shell - 5 240.00 9:1 97.8 Sai et al. [74]
seed oil
Waste Ni-Sr-
74.86 5.12 119.92 11.41:1 98.0 This study
cooking oil doped CaO
a

3.8. Global sensitivity analysis


rn

The first- and total-order Sobol sensitivity indices (SI) were computed and are shown in Fig. 9a to rank the relevance
of these input variables. The second-order sensitivity indices, which assess the importance of the interaction between
u

the input variables, were also calculated and are given in Fig. 9b. The first-order sensitivity indices assess the
Jo

variation in response attributable to first-order variables (single effect terms). The total-order sensitivity indices, on
the other hand, quantify the variance in response ascribed to the combination of first- and second-order variables.
Zhang et al. [75] have previously reported that a parameter is deemed sensitive if its total order sensitivity index is
greater than 0.1. Thus, Fig. 9a shows that apart from the single effect term representing reaction time (X 3), all the
input parameters were sensitive. When the input variables were ranked in order of relevance, X1, which reflects
reaction temperature, was found to be the most sensitive, (SI = 0.21; 21% of the total variance in biodiesel yield).
This was followed by X2, which represented the catalyst concentration, and X4, which represented the methanol-to-
oil ratio. With a sensitivity index of 0.05, reaction time (X3) was the least sensitive input variable. For the second-
order terms, Fig. 9b shows that they were important when adopting a sensitivity threshold of 0.01, apart from X1X3
and X1X4. Fig. 9b demonstrates that X2X3 representing the interaction between catalyst concentration and reaction
time was the most sensitive, accounting for 7.8% of the variation in the biodiesel yield. The reaction temperature
and methanol-to-oil ratio (X1X4) contributed the least to the interaction between the input variables, with an overall
contribution of 0.5%. The GSA results showed that reaction temperature exhibited the highest sensitivity among the
other variables investigated, while reaction time showed the least sensitivity. This has significant importance and

21
implications for optimizing the biodiesel production process [76]. The high sensitivity of the reaction temperature
is an indication that small changes in the reaction temperature could have a significant impact on the biodiesel yield.
This emphasizes the important role of temperature in assessing the kinetics and efficiency of the transesterification
reaction.

of
ro
Fig. 9 Plots showing (a) first- and total order and (b) second-order global sensitivity indices for biodiesel

3.9. Catalyst reusability studies


production -p
re
Fig. 10 shows the results obtained after reusing the catalyst over seven runs. There was a reduction from 98.13 to
93.10% between the first and second cycles. There were further progressive reductions in yield as the number of
lP

cycles increased. The overall reduction in yield between the first (98.13%) and last cycle (75%) was 23.13%.
Although not very significant, the reduction in biodiesel output might be attributed to the increasing occupation of
the active sites on the surface of the catalyst by unreacted triglycerides, which has the effect of gradually reducing
a

its efficacy [2]. Similar observations have also been reported by other researchers who investigated the reusability
rn

of heterogeneous catalysts for biodiesel production [77–79].

100
u
Jo

80
Biodiesel yield (%)

60

40

20

0
1 2 3 4 5 6 7
Number of cycles
Fig. 10 Results of catalyst reusability studies

22
3.10. Biodiesel characterization
Characterization of the properties of the produced biodiesel at the optimum conditions was carried out and the
properties were compared with the two widely recognized benchmarks for biodiesel quality, i.e., the ASTM D6571
and EN 14214 standards as shown in Table 13. It was found that the acid value of the biodiesel sample was 0.22 mg
KOH/g oil which suggests the unlikelihood of the biodiesel to cause corrosion or damage to the fuel system when
used in engines [80]. The biodiesel sample had a low kinematic viscosity of 5.79 mm2/s which was responsible for
the fact that it was a liquid at room temperature which ensures easy handling and smooth flow [81]. Furthermore,
the biodiesel sample demonstrated very good fuel characteristics as seen in the high cetane number of 54.09 and a
higher heating value of 38.60 MJ/kg. An iodine value of 83 g/100g exhibited by the biodiesel sample indicates that
it possesses excellent oxidative stability, ensuring its long-term storage without significant deterioration [37]. In
summary, the biodiesel sample produced at the optimized conditions displayed very desirable properties which all
met the ASTM D6571 and EN 14214 standards.

of
Table 13: Summary of biodiesel properties
Properties Biodiesel ASTM D6751 EN 14214

ro
3
Density 858.2 kg/m Not specified Not specified
Pour point -7 oC <0 <0
Acid value 0.22 mg KOH/g oil <0.5 <0.5
FFA
Specific gravity
0.11%
0.85
-p Not specified
0.86 to 0.90
Not specified
Not specified
re
Cetane number 54.09 ≥47 ≥51
Saponification value 121 mg KOH/g oil Not specified Not specified
lP

Flash point 175 oC >130 >120


o 2
Kinematic viscosity @ 40 C 5.79 mm /s 1.9 to 6.0 3.5 to 5.0
o
Cloud point -18 C Not specified Not specified
a

Iodine value 83 gI2/100 g oil Not specified <120


rn

Higher heating value 38.60 MJ/kg Not specified ≥35

The GC-MS analysis shown in Table 14 revealed that the dominant fatty acids were oleic (38.74%), palmitic
u

(25.41%), stearic (13.41%), and linoleic (9.54%). Oleic acid, the most abundant fatty acid in the biodiesel sample,
Jo

is an example of a monounsaturated fatty acid. Accounting for the highest composition in the biodiesel sample, it
contributes to the favorable fuel properties of the biodiesel. This is because oleic acid is known for its excellent cold
flow properties and low-temperature operability of diesel engines [82]. This is very important for the efficient
functioning of diesel engines fueled with biodiesel in colder regions. Palmitic acid, which was the second most
abundant fatty acid in the biodiesel sample is a saturated fatty acid. Saturated fatty acids provide stability and
improve the overall quality of the fuel. Beyond that, they increase the oxidative stability of biodiesel, exhibiting
oxidation resistance, and thus, reduce the likelihood of the biodiesel undergoing unwanted chemical reactions, such
as polymerization or degradation, during storage or usage [83]. Stearic acid, which was the third most abundant
fatty acid in the biodiesel sample is also a saturated fatty acid, and its presence further contributes to the oxidative
stability of the biodiesel, ensuring that the fuel can be stored for long periods without suffering deterioration.

23
Table 14: Fatty acid profile of biodiesel produced under optimum conditions

Chemical
Fatty acid Nature Composition (%)
formula
Linoleic C18H32O2 Unsaturated 9.54
Stearic C18H36O2 Saturated 13.41
Oleic C18H34O2 Unsaturated 38.74
Arachidic C20H40O2 Saturated 0.62
Palmitoleic C16H30O2 Unsaturated 7.12
Myristic C14H28O2 Saturated 3.12
Palmitic C16H32O2 Saturated 25.41
Others 2.04
Total 100

4. Conclusion

WCO is a is suitable for biodiesel production due to its excellent physical properties and fatty acid compositional

of
profile. A highly active heterogeneous catalyst was successfully synthesized from fused marble waste and cow horn
and used for efficient biodiesel production. The biodiesel production process was established as a multiparametric

ro
process and biodiesel yield was enhanced by high levels of reaction time, reaction temperature, and methanol to oil
ratio while only moderate levels of catalyst concentration are desirable. The optimum conditions for biodiesel
-p
production were successfully established as an optimum biodiesel yield of 98% at a reaction temperature, catalyst
concentration, reaction time, and methanol to oil ratio of 74.86 oC, 5.12 wt%, 119.92 min, and 11.42:1, respectively.
re
Based on its better goodness of fit metrics (R2 value and adjusted R2 value) as well as lower error values (MAE,
MSE, AAD, and RMSE), artificial neural networks is a superior modeling tool compared to response surface
lP

methodology. The optimally prepared biodiesel sample had very good fuel and cold flow properties and they all
satisfied the ASTM D6571 and EN 14214 requirements. In ranking the input variables in order of importance via
global sensitivity analysis, it is seen that X1, which represents reaction temperature while X3 represents reaction
a

time was the least sensitive. The prepared catalyst can be reused for up to seven cycles without very significant
rn

compromise in biodiesel yield.

Availability of data and material


u

All data generated or analyzed during this study are included in this published article or the supplementary material.
Jo

Funding

This research did not receive any funding.

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Author contributions
Ibhadebhunuele Gabriel Okoduwa: Investigation, Writing - original draft. Osamudiamhen Oiwoh: Writing -
original draft, Reviewing and Editing. Andrew Amenaghawon: Conceptualization, Methodology, Writing-
original draft, Reviewing and Editing, Supervision. Charity O. Okieimen: Reviewing and Editing, Supervision

References

1. Maheshwari, P., Haider, M.B., Yusuf, M., Klemeš, J.J., Bokhari, A., Beg, M., Al-Othman, A., Kumar, R.,
24
Jaiswal, A.K.: A review on latest trends in cleaner biodiesel production: Role of feedstock, production
methods, and catalysts. J. Clean. Prod. 355, 131588 (2022). doi:10.1016/J.JCLEPRO.2022.131588

2. Amenaghawon, A.N., Obahiagbon, K., Isesele, V., Usman, F.: Optimized biodiesel production from waste
cooking oil using a functionalized bio-based heterogeneous catalyst. Clean. Eng. Technol. 8, 100501
(2022). doi:10.1016/J.CLET.2022.100501

3. Borugadda, V.B., Dalai, A.K.: Current Developments in the Production of Liquid Transportation Fuels
through the Fischer-Tropsch Synthesis. Fuel Process. Energy Util. 109–121 (2019).
doi:10.1201/9780429489594-7

4. Agrawal, D., Awani, K., Nabavi, S.A., Balan, V., Jin, M., Aminabhavi, T.M., Dubey, K.K., Kumar, V.:
Carbon emissions and decarbonisation: The role and relevance of fermentation industry in chemical
sector. Chem. Eng. J. 475, 146308 (2023). doi:10.1016/J.CEJ.2023.146308

5. Wang, W., Sun, W., Awan, U., A. Nassani, A., H. Binsaeed, R., Zaman, K.: Green investing in China’s air
cargo industry: Opportunities and challenges for sustainable transportation. Heliyon. 9, (2023).
doi:10.1016/J.HELIYON.2023.E19013

of
6. Chen, X., Rahaman, M.A., Murshed, M., Mahmood, H., Hossain, M.A.: Causality analysis of the impacts
of petroleum use, economic growth, and technological innovation on carbon emissions in Bangladesh.

ro
Energy. 267, 126565 (2023). doi:10.1016/J.ENERGY.2022.126565

7. Wirawan, S.S., Solikhah, M.D., Setiapraja, H., Sugiyono, A.: Biodiesel implementation in Indonesia:
Experiences and future perspectives. Renew. Sustain. Energy Rev. 189, 113911 (2024).

8.
doi:10.1016/J.RSER.2023.113911 -p
Rozina, Ahmad, M., Zafar, M.: Synthesis of green and non-toxic biodiesel from non-edible seed oil of
re
Cichorium intybus using recyclable nanoparticles of MgO. Mater. Today Commun. 35, 105611 (2023).
doi:10.1016/J.MTCOMM.2023.105611
lP

9. Wang, L., Wang, H., Fan, J., Han, Z.: Synthesis, catalysts and enhancement technologies of biodiesel from
oil feedstock – A review. Sci. Total Environ. 904, 166982 (2023).
doi:10.1016/J.SCITOTENV.2023.166982
a

10. Ghosh, N., Rhithuparna, D., Rokhum, S.L., Halder, G.: Ethical issues pertaining to sustainable biodiesel
synthesis over trans/esterification process. Sustain. Chem. Pharm. 33, 101123 (2023).
rn

doi:10.1016/J.SCP.2023.101123

11. Gupta, V., Pal Singh, K.: The impact of heterogeneous catalyst on biodiesel production; a review. Mater.
Today Proc. 78, 364–371 (2023). doi:10.1016/J.MATPR.2022.10.175
u

12. Hanif, M.A., Nisar, S., Rashid, U.: Supported solid and heteropoly acid catalysts for production of
Jo

biodiesel. Catal. Rev. - Sci. Eng. 59, 165–188 (2017). doi:10.1080/01614940.2017.1321452

13. Etim, A.O., Musonge, P., Eloka-Eboka, A.C.: An effective green and renewable heterogeneous catalyst
derived from the fusion of bi-component biowaste materials for the optimized transesterification of
linseed oil methyl ester. Biofuels, Bioprod. Biorefining. 15, 1461–1472 (2021). doi:10.1002/BBB.2252

14. Mazaheri, H., Ong, H.C., Amini, Z., Masjuki, H.H., Mofijur, M., Su, C.H., Badruddin, I.A., Yunus Khan,
T.M.: An Overview of Biodiesel Production via Calcium Oxide Based Catalysts: Current State and
Perspective. Energies 2021, Vol. 14, Page 3950. 14, 3950 (2021). doi:10.3390/EN14133950

15. Wong, Y.C., Tan, Y.P., Taufiq-Yap, Y.H., Ramli, I., Tee, H.S.: Biodiesel production via transesterification
of palm oil by using CaO–CeO2 mixed oxide catalysts. Fuel. 162, 288–293 (2015).
doi:10.1016/J.FUEL.2015.09.012

16. Boro, J., Deka, D., Thakur, A.J.: A review on solid oxide derived from waste shells as catalyst for
biodiesel production. Renew. Sustain. Energy Rev. 16, 904–910 (2012). doi:10.1016/j.rser.2011.09.011

17. Nazloo, E.K., Moheimani, N.R., Ennaceri, H.: Graphene-based catalysts for biodiesel production:
Characteristics and performance. Sci. Total Environ. 859, 160000 (2023).
doi:10.1016/J.SCITOTENV.2022.160000
25
18. Chanthon, N., Munbupphachart, N., Ngaosuwan, K., Kiatkittipong, W., Wongsawaeng, D., Mens, W.,
Rokhum, S.L., Assabumrungrat, S.: Metal loading on CaO/Al2O3 pellet catalyst as a booster for
transesterification in biodiesel production. Renew. Energy. 218, 119336 (2023).
doi:10.1016/J.RENENE.2023.119336

19. Li, Y., Zhang, S., Li, Z., Zhang, H., Li, H., Yang, S.: Green synthesis of heterogeneous polymeric bio-
based acid decorated with hydrophobic regulator for efficient catalytic production of biodiesel at low
temperatures. Fuel. 329, 125467 (2022). doi:10.1016/J.FUEL.2022.125467

20. Ong, H.C., Milano, J., Silitonga, A.S., Hassan, M.H., Shamsuddin, A.H., Wang, C.T., Indra Mahlia, T.M.,
Siswantoro, J., Kusumo, F., Sutrisno, J.: Biodiesel production from Calophyllum inophyllum-Ceiba
pentandra oil mixture: Optimization and characterization. J. Clean. Prod. 219, 183–198 (2019).
doi:10.1016/j.jclepro.2019.02.048

21. Jisieike, C.F., Ishola, N.B., Latinwo, L.M., Betiku, E.: Crude rubber seed oil esterification using a solid
catalyst: Optimization by hybrid adaptive neuro-fuzzy inference system and response surface
methodology. Energy. 263, 125734 (2023). doi:10.1016/J.ENERGY.2022.125734

of
22. Kusuma, H.S., Amenaghawon, A.N., Darmokoesoemo, H., Neolaka, Y.A.B., Widyaningrum, B.A.,
Onowise, S.U., Anyalewechi, C.L.: A comparative evaluation of statistical empirical and neural
intelligence modeling of Manihot esculenta-derived leaves extract for optimized bio-coagulation-

ro
flocculation of turbid water. Ind. Crops Prod. 186, 115194 (2022). doi:10.1016/J.INDCROP.2022.115194

23. Todri, E., Amenaghawon, A.N., del Val, I.J., Leak, D.J., Kontoravdi, C., Kucherenko, S., Shah, N.: Global
sensitivity analysis and meta-modeling of an ethanol production process. Chem. Eng. Sci. 114, 114–127

24.
(2014). doi:10.1016/J.CES.2014.04.027 -p
Association of Official Analytical Chemists (AOAC): Official methods of analysis of the Association of
re
Official Analytical Chemists (Vol. 1). (1990)

25. Chantara-arpornchai, S., Luengnaruemitchai, A., Jai-in, S.: Biodiesel Production from Palm Oil using
lP

Heterogeneous Base catalyst. 536–541 (2012)

26. Amenaghawon, A.N., Evbarunegbe, N.I., Obahiagbon, K.: Optimum biodiesel production from waste
vegetable oil using functionalized cow horn catalyst: A comparative evaluation of some expert systems.
a

Clean. Eng. Technol. 4, 100184 (2021). doi:10.1016/J.CLET.2021.100184


rn

27. Sani, J., Samir, S., Rikoto, I., Tambuwal, A.D., Sanda, A., Maishanu, S.M., Ladan, M.M.: Production and
characterization of heterogeneous catalyst (CaO) from snail shell for biodiesel production using waste
cooking oil. Innov. Energy Res. 6, 1–4 (2017)
u

28. Falowo, O.A., Oladipo, B., Taiwo, A.E., Olaiya, A.T., Oyekola, O.O., Betiku, E.: Green heterogeneous
base catalyst from ripe and unripe plantain peels mixture for the transesterification of waste cooking oil.
Jo

Chem. Eng. J. Adv. 10, 100293 (2022). doi:10.1016/J.CEJA.2022.100293

29. Ekpenyong, M., Asitok, A., Ben, U., Amenaghawon, A., Kusuma, H., Akpan, A., Antai, S.: Application of
the novel manta-ray foraging algorithm to optimize acidic peptidase production in solid-state fermentation
using binary agro-industrial waste. Prep. Biochem. Biotechnol. 1–13 (2023).
doi:10.1080/10826068.2023.2214936

30. Kusuma, H.S., Amenaghawon, A.N., Darmokoesoemo, H., Neolaka, Y.A.B., Widyaningrum, B.A.,
Anyalewechi, C.L., Orukpe, P.I.: Evaluation of extract of Ipomoea batatas leaves as a green coagulant–
flocculant for turbid water treatment: Parametric modelling and optimization using response surface
methodology and artificial neural networks. Environ. Technol. Innov. 24, 102005 (2021).
doi:10.1016/J.ETI.2021.102005

31. Saltelli, A., Jakeman, A., Razavi, S., Wu, Q.: Sensitivity analysis: A discipline coming of age. Environ.
Model. Softw. 146, 105226 (2021). doi:10.1016/J.ENVSOFT.2021.105226

32. Vieitez, I., Callejas, N., Irigaray, B., Pinchak, Y., Merlinski, N., Jachmanián, I., Grompone, M.A.: Acid
value, polar compounds and polymers as determinants of the efficient conversion of waste frying oils to
biodiesel. JAOCS, J. Am. Oil Chem. Soc. 91, 655–664 (2014). doi:10.1007/s11746-013-2393-y

26
33. Komariah, L.N., Arita, S., Rendana, M., Ramayanti, C., Suriani, N.L., Erisna, D.: Microbial
contamination of diesel-biodiesel blends in storage tank; an analysis of colony morphology. Heliyon. 8,
(2022). doi:10.1016/J.HELIYON.2022.E09264

34. Ávila Vázquez, V., Díaz Estrada, R.A., Aguilera Flores, M.M., Escamilla Alvarado, C., Correa Aguado,
H.C.: Transesterification of non-edible castor oil (Ricinus communis L.) from Mexico for biodiesel
production: a physicochemical characterization. Biofuels. 11, 753–762 (2020).
doi:10.1080/17597269.2020.1787700

35. Folayan, A.J., Anawe, P.A.L., Aladejare, A.E., Ayeni, A.O.: Experimental investigation of the effect of
fatty acids configuration, chain length, branching and degree of unsaturation on biodiesel fuel properties
obtained from lauric oils, high-oleic and high-linoleic vegetable oil biomass. Energy Reports. 5, 793–806
(2019). doi:10.1016/j.egyr.2019.06.013

36. Adama, K.K., Anani, O.A.: Experimental and theoretical assessment of phenomena linked with separation
and purification of biodiesel from Ricinus communis seed oil. Heliyon. 9, e16536 (2023).
doi:10.1016/J.HELIYON.2023.E16536

of
37. Huang, Y., Li, F., Bao, G., Li, M., Wang, H.: Qualitative and quantitative analysis of the influence of
biodiesel fatty acid methyl esters on iodine value. Environ. Sci. Pollut. Res. 29, 2432–2447 (2022).
doi:10.1007/s11356-021-15762-w

ro
38. Lanjekar, R.D., Deshmukh, D.: A review of the effect of the composition of biodiesel on NOx emission,
oxidative stability and cold flow properties. Renew. Sustain. Energy Rev. 54, 1401–1411 (2016).
doi:10.1016/j.rser.2015.10.034

39.
-p
Ni, Z., Zhai, Y., Li, F., Wang, H., Yang, K., Wang, B., Chen, Y.: Reaction kinetics analysis of branched-
chain alkyl esters of palmitic acid and cold flow properties. Renew. Energy. 147, 719–729 (2020).
re
doi:10.1016/j.renene.2019.08.138

40. Kumbhar, V., Pandey, A., Sonawane, C.R., El-Shafay, A.S., Panchal, H., Chamkha, A.J.: Statistical
lP

analysis on prediction of biodiesel properties from its fatty acid composition. Case Stud. Therm. Eng. 30,
101775 (2022). doi:10.1016/J.CSITE.2022.101775

41. Nahas, L., Dahdah, E., Aouad, S., El Khoury, B., Gennequin, C., Abi Aad, E., Estephane, J.: Highly
a

efficient scallop seashell-derived catalyst for biodiesel production from sunflower and waste cooking oils:
Reaction kinetics and effect of calcination temperature studies. Renew. Energy. 202, 1086–1095 (2023).
rn

doi:10.1016/J.RENENE.2022.12.020

42. Chooi, C.Y., Sim, J.H., Tee, S.F., Lee, Z.H.: Waste-derived green nanocatalyst for biodiesel production:
Kinetic-mechanism deduction and optimization studies. Sustain. 13, (2021). doi:10.3390/su13115849
u

43. Ajala, E.O., Ajala, M.A., Odetoye, T.E., Aderibigbe, F.A., Osanyinpeju, H.O., Ayanshola, M.A.: Thermal
Jo

modification of chicken eggshell as heterogeneous catalyst for palm kernel biodiesel production in an
optimization process. Biomass Convers. Biorefinery. 11, 2599–2615 (2021). doi:10.1007/s13399-020-
00636-x

44. Li, M., Chen, J., Li, L., Ye, C., Lin, X., Qiu, T.: Novel multi–SO3H functionalized ionic liquids as highly
efficient catalyst for synthesis of biodiesel. Green Energy Environ. 6, 271–282 (2021).
doi:10.1016/j.gee.2020.05.004

45. Al-Fatesh, A.S.A., Fakeeha, A.H.: Effects of calcination and activation temperature on dry reforming
catalysts. J. Saudi Chem. Soc. 16, 55–61 (2012). doi:10.1016/J.JSCS.2010.10.020

46. Pandit, P.R., Fulekar, M.H.: Egg shell waste as heterogeneous nanocatalyst for biodiesel production:
Optimized by response surface methodology. J. Environ. Manage. 198, 319–329 (2017).
doi:10.1016/j.jenvman.2017.04.100

47. Zhang, H., An, Q., Su, Y., Quan, X., Chen, S.: Co3O4 with upshifted d-band center and enlarged specific
surface area by single-atom Zr doping for enhanced PMS activation. J. Hazard. Mater. 448, 130987
(2023). doi:10.1016/J.JHAZMAT.2023.130987

48. Ghampson, I.T., Newman, C., Kong, L., Pier, E., Hurley, K.D., Pollock, R.A., Walsh, B.R., Goundie, B.,
27
Wright, J., Wheeler, M.C., Meulenberg, R.W., Desisto, W.J., Frederick, B.G., Austin, R.N.: Effects of pore
diameter on particle size, phase, and turnover frequency in mesoporous silica supported cobalt Fischer-
Tropsch catalysts. Appl. Catal. A Gen. 388, 57–67 (2010). doi:10.1016/j.apcata.2010.08.028

49. Chen, S.Y., Mochizuki, T., Abe, Y., Toba, M., Yoshimura, Y.: Ti-incorporated SBA-15 mesoporous silica
as an efficient and robust Lewis solid acid catalyst for the production of high-quality biodiesel fuels. Appl.
Catal. B Environ. 148–149, 344–356 (2014). doi:10.1016/j.apcatb.2013.11.009

50. Yusuff, A.S., Adeniyi, O.D., Azeez, S.O., Olutoye, M.A., Akpan, U.G.: Synthesis and characterization of
anthill-eggshell-Ni-Co mixed oxides composite catalyst for biodiesel production from waste frying oil.
Biofuels, Bioprod. Biorefining. 13, 37–47 (2019). doi:10.1002/bbb.1914

51. Yu, P., Low, M.Y., Zhou, W.: Design of experiments and regression modelling in food flavour and sensory
analysis: A review. Trends Food Sci. Technol. 71, 202–215 (2018). doi:10.1016/j.tifs.2017.11.013

52. Ilomuanya, M.O., Amenaghawon, N.A., Odimegwu, J., Okubanjo, O.O., Aghaizu, C., Oluwatobiloba, A.,
Akimien, T., Ajayi, T.: Formulation and optimization of gentamicin hydrogel infused with tetracarpidium
conophorum extract via a central composite design for topical delivery. Turkish J. Pharm. Sci. 15, 319–

of
327 (2018). doi:10.4274/TJPS.33042

53. Cabaneros, S.M., Calautit, J.K., Hughes, B.R.: A review of artificial neural network models for ambient

ro
air pollution prediction. Environ. Model. Softw. 119, 285–304 (2019). doi:10.1016/j.envsoft.2019.06.014

54. Solís-Pérez, J.E., Hernández, J.A., Parrales, A., Gómez-Aguilar, J.F., Huicochea, A.: Artificial neural
networks with conformable transfer function for improving the performance in thermal and environmental

55.
-p
processes. Neural Networks. 152, 44–56 (2022). doi:10.1016/J.NEUNET.2022.04.016

Sharma, U., Gupta, N., Verma, M.: Prediction of the compressive strength of Flyash and GGBS
re
incorporated geopolymer concrete using artificial neural network. Asian J. Civ. Eng. 24, 2837–2850
(2023). doi:10.1007/S42107-023-00678-2/METRICS
lP

56. Madhiarasan, M., Tipaldi, M., Siano, P.: Analysis of artificial neural network performance based on
influencing factors for temperature forecasting applications. J. High Speed Networks. 26, 209–223 (2020).
doi:10.3233/JHS-200639
a

57. Liyanaarachchi, V.C., Nishshanka, G.K.S.H., Sakarika, M., Nimarshana, P.H.V., Ariyadasa, T.U.,
Kornaros, M.: Artificial neural network (ANN) approach to optimize cultivation conditions of microalga
rn

Chlorella vulgaris in view of biodiesel production. Biochem. Eng. J. 173, 108072 (2021).
doi:10.1016/j.bej.2021.108072

58. Ayoola, A.A., Hymore, F.K., Omonhinmin, C.A., Olawole, O.C., Fayomi, O.S.I., Babatunde, D., Fagbiele,
u

O.: Analysis of waste groundnut oil biodiesel production using response surface methodology and
artificial neural network. Chem. Data Collect. 22, 100238 (2019). doi:10.1016/j.cdc.2019.100238
Jo

59. Zhao, S., Han, X., Liu, B., Wang, S., Guan, W., Wu, Z., Theodorakis, P.E.: Shelf-life prediction model of
fresh-cut potato at different storage temperatures. J. Food Eng. 317, 110867 (2022).
doi:10.1016/J.JFOODENG.2021.110867

60. Nguyen, T.T., Lam, M.K., Cheng, Y.W., Uemura, Y., Mansor, N., Lim, J.W., Show, P.L., Tan, I.S., Lim, S.:
Reaction kinetic and thermodynamics studies for in-situ transesterification of wet microalgae paste to
biodiesel. Chem. Eng. Res. Des. 169, 250–264 (2021). doi:10.1016/j.cherd.2021.03.021

61. Degfie, T.A., Mamo, T.T., Mekonnen, Y.S.: Optimized Biodiesel Production from Waste Cooking Oil
(WCO) using Calcium Oxide (CaO) Nano-catalyst. Sci. Rep. 9, 1–8 (2019). doi:10.1038/s41598-019-
55403-4

62. Khoobbakht, G., Kheiralipour, K., Rasouli, H., Rafiee, M., Hadipour, M., Karimi, M.: Experimental
exergy analysis of transesterification in biodiesel production. Energy. 196, (2020).
doi:10.1016/j.energy.2020.117092

63. Rezania, S., Korrani, Z.S., Gabris, M.A., Cho, J., Yadav, K.K., Cabral-Pinto, M.M.S., Alam, J., Ahamed,
M., Nodeh, H.R.: Lanthanum phosphate foam as novel heterogeneous nanocatalyst for biodiesel
production from waste cooking oil. Renew. Energy. 176, 228–236 (2021).
28
doi:10.1016/j.renene.2021.05.060

64. Anand Kumar, S.A., Sakthinathan, G., Vignesh, R., Rajesh Banu, J., Al-Muhtaseb, A.H.: Optimized
transesterification reaction for efficient biodiesel production using Indian oil sardine fish as feedstock.
Fuel. 253, 921–929 (2019). doi:10.1016/J.FUEL.2019.04.172

65. Yaakob, Z., Mohammad, M., Alherbawi, M., Alam, Z., Sopian, K.: Overview of the production of
biodiesel from Waste cooking oil. Renew. Sustain. Energy Rev. 18, 184–193 (2013).
doi:10.1016/j.rser.2012.10.016

66. Attari, A., Abbaszadeh-Mayvan, A., Taghizadeh-Alisaraie, A.: Process optimization of ultrasonic-assisted
biodiesel production from waste cooking oil using waste chicken eggshell-derived CaO as a green
heterogeneous catalyst. Biomass and Bioenergy. 158, 106357 (2022).
doi:10.1016/J.BIOMBIOE.2022.106357

67. Elkelawy, M., Alm-Eldin Bastawissi, H., Esmaeil, K.K., Radwan, A.M., Panchal, H., Sadasivuni, K.K.,
Ponnamma, D., Walvekar, R.: Experimental studies on the biodiesel production parameters optimization
of sunflower and soybean oil mixture and DI engine combustion, performance, and emission analysis

of
fueled with diesel/biodiesel blends. Fuel. 255, 115791 (2019). doi:10.1016/j.fuel.2019.115791

68. Babadi, A.A., Rahmati, S., Fakhlaei, R., Barati, B., Wang, S., Doherty, W., Ostrikov, K.: Emerging

ro
technologies for biodiesel production: Processes, challenges, and opportunities. Biomass and Bioenergy.
163, 106521 (2022). doi:10.1016/J.BIOMBIOE.2022.106521

69. Jazie, A.A., Pramanik, H., Sinha, A.S.K.: Special Issue of International Journal of Sustainable
-p
Development and Green Economics (IJSDGE), EGG SHELL AS ECO-FRIENDLY CATALYST FOR
TRANSESTERIFICATION OF RAPESEED OIL: OPTIMIZATION FOR BIODIESEL PRODUCTION.
(2013)
re
70. Ho, W.W.S., Ng, H.K., Gan, S.: Advances in ultrasound-assisted transesterification for biodiesel
production. Appl. Therm. Eng. 100, 553–563 (2016). doi:10.1016/J.APPLTHERMALENG.2016.02.058
lP

71. Farooq, M., Ramli, A.: Biodiesel production from low FFA waste cooking oil using heterogeneous catalyst
derived from chicken bones. Renew. Energy. 76, 362–368 (2015). doi:10.1016/j.renene.2014.11.042
a

72. Mahesh, S.E., Ramanathan, A., Begum, K.M.M.S., Narayanan, A.: Biodiesel production from waste
cooking oil using KBr impregnated CaO as catalyst. Energy Convers. Manag. 91, 442–450 (2015).
rn

doi:10.1016/j.enconman.2014.12.031

73. Mansir, N., Hwa, S., Rashid, U., Hin, Y.: Efficient waste Gallus domesticus shell derived calcium-based
catalyst for biodiesel production E ffi cient waste Gallus domesticus shell derived calcium-based catalyst
u

for biodiesel production. Fuel. 211, 67–75 (2018). doi:10.1016/j.fuel.2017.09.014


Jo

74. Sai Bharadwaj, A.V.S.L., Singh, M., Niju, S., Meera Sheriffa Begum, K.M., Anantharaman, N.: Biodiesel
production from rubber seed oil using calcium oxide derived from eggshell as catalyst-optimization and
modeling studies. Green Process. Synth. 8, 430–442 (2019). doi:10.1515/gps-2019-0011

75. Zhang, W., Cho, C., Piao, C., Choi, H.: Sobol’s sensitivity analysis for a fuel cell stack assembly model
with the aid of structure-selection techniques. J. Power Sources. 301, 1–10 (2016).
doi:10.1016/J.JPOWSOUR.2015.08.076

76. Jaliliantabar, F., Ghobadian, B., Najafi, G., Yusaf, T.: Artificial Neural Network Modeling and Sensitivity
Analysis of Performance and Emissions in a Compression Ignition Engine Using Biodiesel Fuel. Energies.
11, 2410 (2018). doi:10.3390/EN11092410

77. Yusuff, A.S., Gbadamosi, A.O., Popoola, L.T.: Biodiesel production from transesterified waste cooking oil
by zinc-modified anthill catalyst: Parametric optimization and biodiesel properties improvement. J.
Environ. Chem. Eng. 9, 104955 (2021). doi:10.1016/J.JECE.2020.104955

78. Zhang, Y., Niu, S., Han, K., Li, Y., Lu, C.: Synthesis of the SrO–CaO–Al2O3 trimetallic oxide catalyst for
transesterification to produce biodiesel. Renew. Energy. 168, 981–990 (2021).
doi:10.1016/j.renene.2020.12.132

29
79. Gouda, S.P., Ngaosuwan, K., Assabumrungrat, S., Selvaraj, M., Halder, G., Rokhum, S.L.: Microwave
assisted biodiesel production using sulfonic acid-functionalized metal-organic frameworks UiO-66 as a
heterogeneous catalyst. Renew. Energy. 197, 161–169 (2022). doi:10.1016/J.RENENE.2022.07.061

80. Alves, S.M., Dutra-pereira, F.K., Bicudo, T.C.: Influence of stainless steel corrosion on biodiesel oxidative
stability during storage. Fuel. 249, 73–79 (2019). doi:10.1016/J.FUEL.2019.03.097

81. Wilkanowicz, S.I., Hollingsworth, N.R., Saud, K., Kadiyala, U., Larson, R.G.: Immobilization of calcium
oxide onto polyacrylonitrile (PAN) fibers as a heterogeneous catalyst for biodiesel production. Fuel
Process. Technol. 197, 106214 (2020). doi:10.1016/j.fuproc.2019.106214

82. Yeong, S.P., Chan, Y.S., Law, M.C., Ling, J.K.U.: Improving cold flow properties of palm fatty acid
distillate biodiesel through vacuum distillation. J. Bioresour. Bioprod. 7, 43–51 (2022).
doi:10.1016/J.JOBAB.2021.09.002

83. Sia, C.B., Kansedo, J., Tan, Y.H., Lee, K.T.: Evaluation on biodiesel cold flow properties, oxidative
stability and enhancement strategies: A review. Biocatal. Agric. Biotechnol. 24, 101514 (2020).
doi:10.1016/J.BCAB.2020.101514

of
ro
Declaration of interests

-p
☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.
re
Highlights
lP

 Waste marble tiles and cow horn composite catalyst was used for biodiesel production.

 RSM and ANN models were used to predict and optimize biodiesel yield.
a


rn

Global sensitivity analysis showed that reaction temperature was the most influential

input.
u

 The produced biodiesel met recommended standards ensuring quality compliance.


Jo

 The catalyst was stable and reusable for seven cycles.

30

You might also like