1 s2.0 S0045653523002709 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Chemosphere 319 (2023) 138003

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

A comprehensive review on nanocatalysts and nanobiocatalysts for


biodiesel production in Indonesia, Malaysia, Brazil and USA
Hilman Ibnu Mahdi a, b, 1, **, Nurfadhila Nasya Ramlee c, 1, José Leandro da Silva Duarte m,
Yu-Shen Cheng a, e, Rangabhashiyam Selvasembian f, ***, Faisal Amir g, h,
Leonardo Hadlich de Oliveira i, Nur Izyan Wan Azelee c, j, ****, Lucas Meili d, *****,
Gayathri Rangasamy k, l, *
a
Department of Chemical and Materials Engineering, National Yunlin University of Science and Technology, Yunlin, 64002, Taiwan
b
Future Technology Research Center, National Yunlin University of Science and Technology, 123 University Road, Section 3, Douliou, Yunlin, 64002, Taiwan
c
Faculty of Chemical and Energy Engineering, Universiti Teknologi Malaysia (UTM), 81310, Johor Bahru, Johor, Malaysia
d
Laboratory of Processes (LAPRO), Center of Technology, Federal University of Alagoas, Campus A. C. Simões, Lourival Melo Mota Avenue, Tabuleiro Dos Martins,
57072-970, Maceió, AL, Brazil
e
College of Future, National Yunlin University of Science and Technology, 123 University Road, Section 3, Douliou, Yunlin, 64002, Taiwan
f
Department of Biotechnology, School of Chemical & Biotechnology, SASTRA Deemed University, Thanjavur, 613401, India
g
Department of Mechanical Engineering, National Yunlin University of Science and Technology, 123 University Road, Section 3, Douliou, Yunlin, 64002, Taiwan
h
Department of Mechanical Engineering, Universitas Mercu Buana (UMB), Jl. Raya, RT.4/RW.1, Meruya Sel., Kec. Kembangan, Jakarta, Daerah Khusus Ibukota
Jakarta, 11650, Indonesia
i
Laboratory of Adsorption and Ion Exchange (LATI), Chemical Engineering Department (DEQ), State University of Maringá, Maringá (UEM), 5790 Colombo Avenue,
Zone 7, 87020-900, Maringá, PR, Brazil
j
Institute of Bioproduct Development (IBD), Universiti Teknologi Malaysia (UTM), UTM Skudai, 81310, Skudai Johor Bahru, Johor, Malaysia
k
School of Engineering, Lebanese American University, Byblos, Lebanon
l
Department of Sustainable Engineering, Institute of Biotechnology, Saveetha School of Engineering, SIMATS, Chennai, 602105, India
m
Laboratory of Applied Electrochemistry, Institute of Chemistry and Biotechnology, Federal University of Alagoas, Maceió, Alagoas, 57072-900, Brazil

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Worldwide energy demand for diesel


fuel has dramatically been enhanced.
• Biodiesel is an alternative of renewable
biofuel.
• The role of the nanocatalysts and nano­
biocatalysts in biodiesel production was
discussed.
• Key factors for large scale biodiesel
production were pointed out.
• Large scale production in Indonesia,
Malysia and Brazil using heterogeneous
catalysts.

* Corresponding author. School of Engineering, Lebanese American University, Byblos, Lebanon.


** Corresponding author. Department of Chemical and Materials Engineering, National Yunlin University of Science and Technology, Yunlin, 64002, Taiwan.
*** Corresponding author.
**** Corresponding author. Faculty of Chemical and Energy Engineering, Universiti Teknologi Malaysia (UTM), 81310, Johor Bahru, Johor, Malaysia.
***** Corresponding author.
E-mail addresses: dave087.ronel@gamil.com (H.I. Mahdi), rambhashiyam@gmail.com (R. Selvasembian), nur.izyan@utm.my (N.I. Wan Azelee), lucas.meili@
ctec.ufal.br (L. Meili), granga1983@gmail.com (G. Rangasamy).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.chemosphere.2023.138003
Received 23 July 2022; Received in revised form 24 December 2022; Accepted 27 January 2023
Available online 30 January 2023
0045-6535/© 2023 Elsevier Ltd. All rights reserved.
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

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

Handling Editor: Pau-Loke SHOW Biodiesel is an alternative to fossil-derived diesel with similar properties and several environmental benefits.
Biodiesel production using conventional catalysts such as homogeneous, heterogeneous, or enzymatic catalysts
Keywords: faces a problem regarding catalysts deactivation after repeated reaction cycles. Heterogeneous nanocatalysts and
Biotechnology nanobiocatalysts (enzymes) have shown better advantages due to higher activity, recyclability, larger surface
Nanotechnology
area, and improved active sites. Despite a large number of studies on this subject, there are still challenges
Nanobiocatalysts
regarding its stability, recyclability, and scale-up processes for biodiesel production. Therefore, the purpose of
Vegetable oils
Biodiesel this study is to review current modifications and role of nanocatalysts and nanobiocatalysts and also to observe
effect of various parameters on biodiesel production. Nanocatalysts and nanobiocatalysts demonstrate long-term
stability due to strong Brønsted-Lewis acidity, larger active spots and better accessibility leading to enhancethe
biodiesel production. Incorporation of metal supporting positively contributes to shorten the reaction time and
enhance the longer reusability. Furthermore, proper operating parameters play a vital role to optimize the
biodiesel productivity in the commercial scale process due to higher conversion, yield and selectivity with the
lower process cost. This article also analyses the relationship between different types of feedstocks towards the
quality and quantity of biodiesel production. Crude palm oil is convinced as the most prospective and promising
feedstock due to massive production, low cost, and easily available. It also evaluates key factors and technologies
for biodiesel production in Indonesia, Malaysia, Brazil, and the USA as the biggest biodiesel production supply.

Abbreviations NTVL Nile tilapia viscera lipase


PCL Pseudomonas cepacia lipase
Feedstocks PCYL Penicillium cyclopium lipase
CIO Calophyllum inophyllum oil PCLB Pseudomonas cepacia lipase B
CNO canola oil PEL Penicillium expansum lipase
CSO castor oil PFL Pseudomonas fluorescens lipase
CPO crude palm oil PPL porcine pancreas lipase
FFA free fatty acid ROL Rhizopus orizae lipase
FWO fish waste oil RML Rhizomucor miehei lipase
FPO frying palm oil TLL Thermomyces lanuginosus lipase
JTO Jatropha oil
JUO Jupati oil Reactors
KSO kapok seed oil BR Batch Reactor
KRO karanja oil CPSR Cylindrical Pressurized Stainless Reactor
MPO Macaw palm oil EF Erlenmeyer Flask
MFO Mesua ferrea oil FBR Fluidized Bed Reactor
PFAD palm fatty acid distillate GR Glass Reactor
RBO rubber seed oil MR Multiple Reactor
RSO rapeseed oil MW Microwave Reactor
SBO soybean oil MSCR Multistage stirred column reactor
SFO sunflower oil PAR Parr Autoclave Reactor
SPO sludge palm oil PAR Pressure Autoclave Reactor
TG triglycerides PBR Packed Bed Reactor
TSO tallow seed oil PCTR Pseudo-Catalytic Transesterification Reactor
WCO waste cooking oil PR Parr Reactor
WCSO waste cottonseed oil PTR Pressure Tube Reactor
WEO waste edible oil SI Shaker Incubator
WFO waste frying oil TNR Three Necked Reactor

Enzymes Variables
ANL Aspergillus niger lipase X biodiesel conversion percent (%)
BCL Burkholderia cepacia lipase Y biodiesel yield (%)
CALA Candida antarctica lipase A T temperature (◦ C)
CALB Candida antarctica lipase B t time (h)
CRL Candida rugosa lipase dp catalyst particle size (nm)
FHL Fusarium heterosporum lipase

that gradually reduces oil production by 1 million barrels per day by


2025, resulting in a dramatic effect on the oil price over time (Arezki
1. Introduction et al., 2017). Atmospheric pollution created by fossil fuel combustion
becomes the primary effect of increased greenhouse gas emissions and
The depletion of oil reserves declined its production and caused concern about global warming (Azelee et al., 2022). Fossil fuel com­
shortages for many society sectors, thus leading to an uncertain future bustion generates 85% of airborne nocive compounds by emitting sulfur
for the oil market. The pessimistic scenario assumes an exogenous shock

2
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 1
Physical and chemical properties of biodiesel and diesel fuel standard range (based on ASTM-D 6751 and ASTM-D 975).
Detailed Analysis method Standard Operating Procedure (SOP) Biodiesel Diesel ASTM References
properties Fuel No.

Acid number Colorimetric titration A mixture of oil and solvent was dissolved 0.198–1.167 0.03 ASTM-D (Mahajan et al., 2006; Gupta et al., 2020)
(mg KOH/g and then titrated with KOH 0.1 N as titrant 664
oil) and p-naphtholbenzein as an indicator
Specific gravity Density measurement Biodiesel samples were measured in density 0.80–0.89 0.83–0.85 ASTM-D (Tat and Van Gerpen, 2000; Baroutian
(mg/mL) flask (25 ml of oil/volume-25 ml) at 6751 et al., 2008; Benjumea et al., 2008; Huber
5–100 ◦ C and 1 bar et al., 2009; Veny et al., 2009; Nogueira
et al., 2010; Pratas et al., 2011)
Kinematic Brookfield b 1.9–6.0 1.3–4.1 ASTM-D
Kinematic viscosity (μ) formula = a.t − ;a MEHRA (2012)
viscosity at viscometer t 445
40 ◦ C (mm2/s) = 0.097 and b = 244.89
Viscosity at ASTM standard Tetrachloroethylene (4–14%) was applied in 5.51 3.60 ASTM-D MEHRA (2012)
30 ◦ C (cp) method the analysis as an additive 6751
Cetane number Spark-ignition The fuel starts to auto-ignite when squirted 48–67 47.8–59 ASTM-D (Van Gerpen, 1996; Xing-cai et al., 2004)
engines into the engine 613
Sulfur content ASTM standard Tetrachloroethylene (4–14%) was applied in 0.13–0.16 <1.0–1.2 ASTM-D MEHRA (2012)
(%) method the analysis as an additive 6751
Flash Point Closed Cup Pensky The lowest temperature at which an 100–170 60–80 ASTM-D MEHRA (2012)
Martens apparatus ignitable mixture is formed 93
Cloud Point Refrigerated cloud The temperature point at which wax starts to − 3 to +12 − 15 to +5 ASTM-D MEHRA (2012)
point and pour point be a sediment 2500
apparatus
Copper ASTM standard Tetrachloroethylene (4–14%) was applied in 3 – ASTM-D MEHRA (2012)
corrosion method the analysis as an additive 130
Pour point ASTM standard Tetrachloroethylene (4–14%) was applied in − 15 to 10 − 35 to ISO (MEHRA, 2012; Yaakob et al., 2014)
method the analysis as an additive − 15 3016
Oxidation Rancimat induction According to quicken the aging process by >3 – ASTM-D Yaakob et al. (2014)
stability period (RIP) increasing the temperature 6751
(hours)

Note: ASTM = American Society of Testing of Materials standard; μ = Kinematic viscosity in centipoise (cp); t = time in second.

oxides (SOx), nitrogen oxides (NOx), polycyclic aromatic hydrocarbons


(PAH), mercury (Hg), and volatile chemicals (VOCs) to the atmosphere.
All the airborne poisonous compounds are associated with adverse ef­
fects on human health (IEA, 2016; Perera, 2017) including microplastic
negatively contributing to environmental and ecological health risks
(Gwenzi et al., 2022).
Hence, there is an urge for searching alternative fuels that compen­
sate for oil shortage and mitigate the atmospheric damage, especially for
the energy and transportation sectors. Furthermore, with technological
advancements, increasing global population, and economic growth, the
search for alternative fuels is intensifying in the short term (Ali et al.,
2018).

1.1. Biodiesel as the alternative fuel Fig. 1. Esterification and transesterification reactions for biodiesel production.

As a potential alternative to petrodiesel, biodiesel fulfills sustain­ acidity and oxygenated compounds as well as water content.
ability criteria such as readily available feedstocks throughout the world The esterification reaction involves the conversion of free fatty acid
(binti Ramlee et al., 2022), ease to implement technology (technically (FFA) to fatty acid esters. In contrast, the transesterification reaction
feasible), non-toxic and biodegradable (social and environmentally uses triglycerides (TG) as a substrate (Hidayat et al., 2015; Andrade
friendly), low cost and competitive (economically affordable) (Gomiero, et al., 2017). The mass balance streams of both reactions for biodiesel
2015). Besides, the side-product, glycerol, can be used to produce an production using alcohol and a catalyst (Gupta et al., 2020; Vasić et al.,
intermediate product to mitigate the waste generated from biodiesel 2020) are shown in Fig. 1. The presence of catalyst and alcohol is crucial
production (Mahdi et al., 2016). As a result, global biodiesel production to perform both reactions.
is expected to reach 41.4 billion liters by 2025, corresponding to a 33% Various catalysts are commonly used for biodiesel production,
increase from 2015, which can compensate for the crude oil scarcity including homogeneous, heterogeneous, and enzymatic catalysts. For
(OECD/FAO, 2016). Furthermore, like petroleum diesel, the physical example, conventional transesterification utilizes homogeneous alkaline
and chemical properties of biodiesel produced from renewable feed­ catalysts to convert oils into biodiesel and presents some advantages,
stocks should comply with ISO/ASTM standards, as shown in Table 1. such as i) fast reaction in a short amount of time; ii) cost-effective pro­
Many alternative fuels are produced through thermochemical cess; iii) economically available; iv) and mild conditions of temperature
transformations such as pyrolysis, gasification, and liquefaction (Akia and pressure (Talha and Sulaiman, 2016).
et al., 2014a). However, the obtained fuels have some adverse proper­ Despite these advantages, the reaction requires low FFA content
ties, such as high acidity and oxygenated compounds, high water con­ feedstocks because high FFA content favors saponification reaction that
tent, less burning efficiency, and time-consuming processes lowers biodiesel production (Chhetri et al., 2008). Besides, the presence
(Badoei-dalfard et al., 2019). Therefore, esterification and trans­ of water can promote oil hydrolysis, which increases FFA at the end of
esterification reactions for biodiesel production are seen as highly effi­ the reaction. Apart from that, homogeneous reactions generate waste­
cient methods to improve the fuel quality produced by reducing high water, which is environmentally unfriendly and require extensive

3
H.I. Mahdi et al.
4

Chemosphere 319 (2023) 138003


Fig. 2. Roadmap to achieve commercial readiness level in biodiesel production using heterogeneous catalysts.
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

separation operation, complicating the downstream process, thus mak­ nanocatalysts is the large surface area to volume ratio, which enhances
ing it difficult to recover the alkaline soluble catalysts (Zahan and Kano, the mixing performance with reactants (Mahdi and Muraza, 2019) and
2018). On the other hand, homogeneous acid catalysts are introduced to eases the separation process (Jiang et al., 2014; Saoud, 2018).
counteract limitations of homogeneous alkaline catalysts. The main On the other hand, nanobiocatalysts employed in the biodiesel
advantage of utilizing homogeneous acid catalysts lies in their capability transesterification process are mostly consisted of two components
to esterify high FFA content feedstocks to produce biodiesel (Mazumdar which are nanocatalysts and biological components. The use of nano­
et al., 2013). biocatalysts in the biodiesel production is similar to the use of nano­
However, this method produces acidic wastewater, causing corro­ catalysts by adding alcohol, water, vegetable oils and promotes the
sion to the equipment, requires a high oil/methanol ratio and high acid biodiesel productivity and catalyst reusability. Gao et al. optimized
concentration, and complicates the catalyst recovery process (Citra biodiesel produced under transesterification process using Jatropha oil
Dewi and Slamet, 2019). Generally, the main problems associated with and nanobiocatalysts by resulting in the maximum biodiesel yield of
homogeneous acid and alkaline catalysts lie in the complicated separa­ 95%. Furthermore, the modified nanobiocatalysts demonstrated long-
tion process, wastewater generations, unable to recover and reuse the term stability and high thermal stability (Gao et al., 2016).
catalysts, thus increasing the overall operational cost. Using heteroge­ Various conventional catalyst synthesis methods (see Fig. 2), such as
neous catalysts is a promising solution to overcome these problems combustion, precipitation, co-precipitation, impregnation,
associated with homogeneous catalysts. precipitation-hydrothermal, precipitation-ultrasonic (Alaei et al., 2020),
Biodiesel production processes using heterogeneous catalysts can be aiming for high catalytic performance and stability were confirmed to
schematically represented as portrayed in Fig. 2, where all stages, from produce high biodiesel yield. Therefore, several reports employed
catalyst synthesis, and process development up to the commercialization nanocatalysts for biomass liquefaction and gasification, and esterifica­
process, are depicted. Therefore, Fig. 2 can be considered a guide to tion and transesterification reactions. This synergistic interaction be­
comparing any biodiesel production process using heterogeneous tween nanotechnology and biotechnology has become a driving force for
catalysts. the emergence of nanobiocatalyst in biodiesel production.
Different from homogeneous, heterogeneous catalysts are easy to Despite the advantages mentioned that: i) nanobiocatalysts, which
recover and can be reused for the next cycle (Mahdi et al., 2021), making involve the incorporation of enzymes in nanostructured materials,
the separation process easier and environmentally friendly (Ling et al., promise exciting, innovative advances in many areas of enzyme tech­
2019; Iqbal et al., 2020). Because of the small surface area (critical nology (Kim et al., 2008; Verma et al., 2016a); ii) in the prospect of
factor), heterogeneous catalysts require high temperatures, pressures, enzyme-catalyzed biodiesel production, lipases can be easily obtained
and ratios of oil/methanol as operation factors, low activity (observation from Candida rugosa, Themomyces lanuginosus, Aspergillus terreus and
factor), and low stability (results factor), all of which increase opera­ Rhizopus oryzae (Nematian et al., 2020b); iii) recent catalyst de­
tional costs (OPEX) (Mahdi and Muraza, 2016; Saoud, 2018). velopments have led to the improvement of esterification/transester­
Besides chemical catalysts, enzymatic catalysts (lipases) are ification reactions to increase biodiesel production; there is a gap in the
becoming preferred because of their high substrate specificity and can literature regarding the readiness of the technology of biodiesel pro­
catalyze reactions at ambient temperatures and pressure (Afifah et al.; duction using nanobiocatalysts in several countries. Moreover, what is
Knezevic et al., 2004; Ramlee et al., 2022). In addition, enzymes may the best technology for great vegetal oil producers such as Indonesia,
prevent saponification and reduce the number of reaction steps and Malaysia, Brazil, and the USA?
hazardous solvents, thus making the process less expensive and much Therefore, this manuscript reviews the recent development and
more environmentally friendly (Zdarta et al., 2018; Afifah et al., 2019). progress of nanocatalysts and nanobiocatalysts for the catalytic con­
Enzyme immobilization is a method to escalate the performance, capa­ version of triglycerides in biodiesel production processes are reviewed.
bility, and recyclability of lipase in biodiesel production. Then, the role of nanocatalysts and nanobiocatalysts was evaluated in
Researchers proposed several immobilization strategies such as terms of synthesis and characterization, operational conditions, exper­
adsorption, covalent bonding (Ramlee et al., 2022), entrapment, imental observations, downstream treatments, and engineering in­
encapsulation, cross-linking, and others (Fig. 2) to improve lipase ac­ dicators to propose the most effective production technologies using
tivity and stability (Jambulingam et al., 2019). Immobilization strate­ nanobiocatalysts to be implemented in Indonesia, Malaysia, Brazil, and
gies can solve the problems associated with the single use of free the USA. Exploration of the mostly used technology in biodiesel pro­
enzymes in the reaction. Nevertheless, the leaching problems and low duction in countries supplying the biggest biodiesel production in the
enzymes loading associated with the immobilization of enzymes on a world is proposed to discover the best biodiesel technology thus can be
carrier are still a significant concern. Both chem- and bio-catalysts useful for other countries to apply in the commercial biodiesel process.
possess limitations that reduce the efficiency of biodiesel production. Finally, this article concludes with the advancement of the latest tech­
In this sense, integrating different knowledge areas become essential to nology applied for biodiesel production, and the future perspectives of
enhance the catalytic performance of biodiesel production. nanobiocatalysts are discussed to give insight into process optimization.

1.2. Nanotechnology in biodiesel production 2. Heterogeneous solid catalysts in biodiesel production:


capability and drawbacks
The unprecedented developments in science and technology have
bridged the gap between sophisticated innovations and research using Esterification and transesterification reactions for biodiesel produc­
nanoscale techniques, leading to the establishment of a new era known tion are mainly processed at a large scale by utilizing the conventional
as nanotechnology. In addition, the developments in nanoscience and homogeneous catalysts, including KOH, NaOH, H2SO4, K3PO4, CH3ONa,
nanotechnology have stimulated new fields of nanocatalysts (Mahdi and CaO, CH3OK, for years (Agarwal et al., 2012; Atadashi et al., 2013).
Muraza, 2019; Hosseini, 2022) and nanobiocatalysts (Najeeb et al., Pretreatment of waste oils is required to reduce the FFA content by
2021). adding sulfuric acid and methanol in the reaction at 70 ◦ C and 4 bar
Nanocatalysts bridge the gap between homogeneous and heteroge­ (Knothe, 2010; Knothe and Razon, 2017). Nevertheless, the process has
neous catalysis, offering advantages in terms of activity, selectivity, ef­ many drawbacks, such as high energy consumption in purification and
ficiency, and reusability, helping to overcome some limitations using separation of products, catalyst removal in the downstream process, and
conventional catalysts. Nanoparticles are spherical dots, rods, thin non-reusability of the catalysts (Saha and Woodward, 1997; Konwar
plates, or any irregular shape with a cross-section of less than 100 nm et al., 2014; Long et al., 2014). Moreover, using homogeneous acid and
(Khan et al., 2019; Prinsen and Luque, 2019). The excellent property of base catalysts can cause corrosion in the reactor and soap formation

5
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 2
The ability of heterogeneous solid acid catalysts in transesterification for biodiesel productions.
Catalysts Reactants Treatment of catalysts Reactor Operating reactions Results Reference
Types o
T ( C) Catalyst X (%) Y (%) Cycles
loading
(wt.%)

CaO-dopped CAO Electro hydrodynamic TNR 55 2.5 g – 84–93 6 Wilkanowicz et al.


polyacrylonitrile processing (EHD) (2020)
(PAN)
Al2O3-assisted RBO Calcination and TNR 65 3 98.9 – – Lakshmi et al.
calcined eggshells impregnation (2020)
Carbon Oleic acid and Carbon nanoparticles FBR 100 2.5 95 – 4 Ballotin et al.
methanol synthesis (2020a)
Sulfonated carbon JUO Carbonization and MR 135 6 – 91.8 4 Bastos et al. (2020)
sulfonation
KOH WCO – BR 45 2 100 – – Outili et al. (2020)
A – 45 Vegetable oil Washing, drying, and PTR 150–170 5–15 – 80–84 5 Cabral et al. (2020)
crushing
Sulfuric acid (H2SO4) Corncob Sulfuric acid treatment PR 60–100 5 – >80 – Dechakhumwat
and Sulfonation et al. (2020)
An acid ion- WCO Drying TNR 115 2.5 96.73–99.26 – – Li et al. (2020a)
exchange resin
Biochar WCO – PCTR – 1 – 90 – Lee et al. (2020)
SO2−
4 /Al2O3–SnO2 Sewage sludge Co-precipitation BR 130 8 – 73.3 – Zhang et al. (2020c)
Fe3O4/SiO2 SBO Hydrothermal CPSR 120 99 93.3 – Several Xie and Wang
(2020)
CaO coming from RBO Washing and calcination TNR – 5 97.84 – – Sai et al. (2020)
calcined eggshells
ZrSiW/Fe-BTC and Oleic acid Hydrothermal BR 140–160 0.24 g 90 and 98 – 6 Zhang et al. (2020b)
ZrSiW/UiO-66
Zn–Mg–Al Lauric acid Calcination MR 190 5 98.2 – 6 Abas et al. (2020)
Tin (II) oxalate Lauric acid Calcination MR 190 5 65.4 – 6 Abas et al. (2020)
A-15 Lauric acid Calcination MR 190 5 31.6 – 3 Abas et al. (2020)
CaO Waste biomass Impregnation BR 200–400 10 – 96 – Chua et al. (2020)
Carica Papaya stem WCO and Crushing, drying, and TNR 100 2 95.23 and – 6 Gohain et al. (2020)
Scenedesmus washing 93.33
obliquus (SO) lipid
Bismuth silicate Oleic acid Synthesized over TNR 80 0.3 g 90 – – Mahmoud (2019)
ultrasonication-
supported hydrothermal
BSY-SO3H PFAD Activation (with TNR 65 8 87.8 – – Arumugamurthy
phosphoric acid), et al. (2019)
carbonization, and
sulfonation
ASCC WFO Impregnation BR 65 1.5 – 91.05 5 Yusuff and Owolabi
(2019)
CES-Fe3O4 Pongamia pinnatta Calcination at 900 C and

TNR 65 2 – 98 7 Chingakham et al.
impregnation (2019)
4-BDS-assisted PFAD Calcination at 200 ◦ C and BR 200 20 – 98.1 – Lim et al. (2019)
carbon sulfonation
rGO-SO3H SBO Sulfonation PAR 80 3 – 99 5 dos Santos et al.
(2019)
MgO-urea-800/ CSO A sol-gel and calcination BR 75 6 – 96.5 5 Du et al. (2019)
Carbon
AENiCo WFO Co-precipitation BR 70 3 – 90.23 4 Yusuff et al. (2019)
Nano-particle of Pork fat Synthesis and dopping BR 65 8 – 92.5 – Cherian et al. (2019)
Al–CaO
Nano-size of ZnO and Ulva lactuca Co-precipitation BR 55 8 – 97.43 – Kalavathy and
clay Baskar (2019)
Nano particle of Calophyllum Calcination – 55 6 – 89 – Naveenkumar and
Zinc-assisted CaO inophyllum oil Baskar (2019)
Tungsten-dopped Ti/ WCSO, CSO, KRO A sol-gel (no template) TNR 65 5 – >98 4 Kaur et al. (2018)
SiO2
Mg/MgAl2O4 Vegetable oil Impregnation BR 110 3 – 95.7 – Sahota and Tiwari
(2017)
Carbon Jatropha curcas Drying and sulfonation TNR 60 7.5 99.13 – 4 Mardhiah et al.
(2017)
HPMo MPO Impregnation PR 210 20 99.65 – 2 da Conceição et al.
(2017)
Graphene oxide Microalgae oil Sulfonation MR 90 5 95.1 – – Cheng et al. (2016)
MPD–SO3H–IL JTO Synthesis and ion PAR 100 5 98 94 4 Pan et al. (2016)
exchange
SiO2-Pr-SO3H Acid oil Sulfonation – 100 4 96.78 97 – Shah et al. (2015)
AlCl3⋅6H20 Brown grease – – 47 2 >90 65.3 – Pastore et al. (2014)
CsHPW Mixture of oleic Impregnation PAR 200 3 90.4 91 – Sheikh et al. (2013)
acid and soybean
CaO Palm olein BR 60 7 – 95.7 5 Yoosuk et al. (2010)
(continued on next page)

6
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 2 (continued )
Catalysts Reactants Treatment of catalysts Reactor Operating reactions Results Reference
Types
T (oC) Catalyst X (%) Y (%) Cycles
loading
(wt.%)

Calcination and water


treatment
Cs-modified Palmitic acid Drying and ethanol BR 60 0.05 g 100 S= – Pesaresi et al.
H4SiW12O40 treatment 87% (2009)
H+-exchanged SBO Hydrothermal with BR 60 0.5 80 80 – Chung and Park
ZSM-5 and different Si/Al ratios (2009)
mordenite (MOR)
SO2−
4 /ZrO2 and Vegetable oils Impregnation and BR 75 0.2 g 65–70 85 – Park et al. (2008)
WO3/ZrO2 calcination
ZrO2–SO4 Benzoic acid A sol-gel and calcination BR 160 2.5 g >60 – – Ardizzone et al.
at 617 ◦ C (2004)

Note:X% = percent conversion; Y% = percentage of biodiesel yield; A = Amberlyst; FBR = Fluidized Bed Reactor; MR = Multiple Reactor; PTR = Pressure Tube
Reactor; PR = Parr Reactor; CALA = Candida Antarctica lipase A; GR = Glass Reactor; TNR = Three Necked Reactor; BR = Batch Reactor; PCTR = Pseudo-Catalytic
Transesterification Reactor; CPSR = Cylindrical Pressurized Stainless Reactor; CES-Fe3O4 = Nano-size Fe3O4-supported catalytic eggshell; ASCC = Alumina-assisted
coconut chaff; 4-BDS = 4-Benzene Diazonium Sulfonate; rGO-SO3H = Sulfonic-Reduced Graphene Oxide; PAR = Parr Autoclave Reactor; AENiCo = Anthill-eggshell-
Ni-Co; BSY-SO3H = Brewer’s spent yeast; Bismuth silicate = Bi4Si3O12 and Bi2SiO5; HPMo = H3PMo12O40; MR = Microwave Reactor; MPD–SO3H–IL = Mesoporous
Polymer hydrothermally synthesized from copolymerization of D with V functionalized by ionic liquids; PAR = Pressure Autoclave Reactor; SiO2–Pr–SO3H = Sulfonic
acid-functionalized silica; CsHPW = cesium-doped heteropoly tungstate.

when employing vegetable oils, which comprise high water and FFA the heterogeneous catalysts can reduce biodiesel contamination and the
content and lead to low biodiesel yield (Sani et al., 2014; Mansir et al., catalyst deactivation rate caused by the solvation of the catalyst and
2017). According to these shortfalls, researchers were driven to study water content (Nusterer et al., 1996; Veljković et al., 2006).
other catalysts that could be applied in biodiesel synthesis through Regarding these remarkable capabilities, several developments in
esterification and transesterification, as reported in the recent literature biodiesel production through esterification and transesterification re­
(Akhabue et al., 2020; Karmakar et al., 2020). actions using heterogeneous solid acid and base catalysts are nowadays
Among the promising and potential catalysts applied to biodiesel exploited in many kinds of research to improve the reaction rate and to
reactions with a higher biodiesel yield, there are the heterogeneous base eliminate the undesired by-products (Hernández-Hipólito et al., 2015;
and acid catalysts (Sahota and Tiwari, 2017; Kulkarni et al., 2020). Osorio-González et al., 2020). Downstream and upstream variables
Heterogeneous base catalysts have alkaline earth metals cations (i.e., Ca, studied in biodiesel production using heterogeneous catalysts are shown
Na, Mg, and K), which can improve the catalysts’ active sites and further in Table 2. Data raised in Table 2 were used to provide a map shown in
enhance the catalytic performance (Mansir et al., 2017). Additionally, Fig. 3.
some authors show that heterogeneous solid acid catalysts can be It is possible to see in Table 2 that several heterogeneous catalysts
recycled up to several reaction cycles with less catalyst deactivation were studied with different reactants giving different conversion and
(Borges and Díaz, 2012; Galadima and Muraza, 2014; Teo et al., 2015). yield results, depending on the catalyst treatment and reactor type. It is
In this sense, the relationship between catalytic activity, stability, and also verified that literature has gaps regarding information catalyst
reusability of the heterogeneous catalysts has been continuously loading, conversion achieved, biodiesel yield, and the number of cycles.
developed. It should be pointed out that many papers just performed only a single
A potential method for enhancing catalyst reusability is to test cycle.
different molybdenum contents (i.e., 1–5 wt%) supported on calcium Results of Table 2 are depicted in Fig. 3a, where it can be verified that
oxide (Mo–CaO) and calcination temperatures (i.e. 300 - 800 ◦ C) (Kaur higher biodiesel yields are reported in the literature using catalyst
and Ali, 2015). The experiment resulted in an optimum yield of biodiesel loadings between 3 and 5 wt% and 4 to 5 catalyst cycles. Regions with
(>99%) at 65 ◦ C, 5 wt%, and 12:1 ethanol:oil ratio, where the catalysts the lowest yields are shown in 2–4 cycles, where it is also noticed that
could be reused for five cycles without deactivation of the catalyst literature needs more data on these regions.
activity.
The presence of Brønsted and Lewis acid, determined by Si/Al ratio, 3. The role of nanocatalyst in biodiesel production
remarkably induces the heterogeneous catalyst ability (Chouhan and
Sarma, 2011) and constitutes simultaneous esterification and trans­ Considering the advantages is shown in Section 2, using heteroge­
esterification reactions (Vyas et al., 2010; Yan et al., 2012). Moreover, it neous acid and base catalysts increases biodiesel yield. Furthermore, to
could be applied in low-quality feedstocks which are inexpensive and further upgrade the activity, capability, and stability of the catalysts,
contain high water and FFA contents without requiring any pretreat­ such catalysts was synthesized as nanocrystals (nanocatalysts), which
ment methods (Canakci, 2007; Zafiropoulos et al., 2007; Di Serio et al., played a significant and vital role because their morphology presents a
2008; Yan et al., 2009). Rattanaphra et al. have performed a simulta­ larger surface area and more spacious active sites (Islam et al., 2012;
neous esterification and transesterification reactions on low-grade Mansir et al., 2017).
feedstock (i.e., rapeseed oil) with methanol using sulfated zirconia The carbon nanocatalysts with various nanostructures such as gra­
(ZrO2–SO4) as the heterogeneous catalysts, at various reaction temper­ phene, nanotube, nano-onions, and nanographite were studied for the
atures (120, 150, and 170 ◦ C) and FFA contents (0, 10, and 20 wt%) esterification of oleic acid and methanol (Ballotin et al., 2020b). The
(Rattanaphra et al., 2010). The catalyst indicated better capability in the result exhibited that the nanocatalysts successfully resulted in high
simultaneous reactions than the conventional method, with 10 wt% of conversion (95%) at 100 ◦ C, 2.5 wt% of catalyst loading within 3 h and
FFA content showing the highest biodiesel production without solvent could be reused during 4 cycles without catalyst deactivation. Mean­
addition. Therefore, the washing method in the final process and while, Foroutan and his colleagues used a MgO-assisted CaO nano­
corrosion problem can be easily tackled using simultaneous esterifica­ catalyst with a surface area of 6.50 m2/g and a pore volume of 0.022
tion and transesterification reaction (Guo et al., 2012). On the contrary, cm3/g for the transesterification of waste edible oils (WEO) (Foroutan

7
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

et al., 2020), and the highest biodiesel yield was 98.37% at 69.37 ◦ C,
methanol:oil ratio of 16.7:1, 4.571 wt% of catalyst loading for 7.08 h
and the nanocatalyst was reused during 6 cycles and showed impressive
catalytic performance and stability, showing a higher yield and longer
reusability than the carbon nanocatalyst (Verziu et al., 2008). On the
other hand, the authors also found that the Zn-assisted CaO nanocatalyst
generated the lowest biodiesel yield (89%) (Naveenkumar and Baskar,
2019).
Borah et al. c utilized a Co-impregnated ZnO nanocatalyst with a
particle size of 27.8 nm. The highest biodiesel yield was 98.03%,
reached at 60 ◦ C, 2.5 wt% of catalyst loading, methanol:oil ratio of 9:1
within 6 h. However, the nanocatalyst is not stable after repeated cycles,
thus requiring improvements to upgrade the catalyst stability. The
transesterification of palm oil using ZnO–Ag-doped nanocatalyst with a
particle size of 23 nm gave a better activity with high biodiesel yield and
reusability of 5 cycles, as reported by Laskar et al. (2020).
Together with the novelties recently proposed in the literature, the
advantages of the using nanocatalysts for transesterification regarding
biodiesel yield and conversion are summarized in Table 3. In these
works, several nanobiocatalysts were prepared with different particle
diameters and studied using different oil feedstocks, reaction tempera­
tures, and times, giving different biodiesel yield and reusability (number
or cycles). As in Table 2, it should also be noted that most of the studies
only performed a single reaction cycle attributed to the leaching of the
active sites to the reaction medium and deactivation of the catalyst after
a repeated reaction cycle.
Results of Table 3 are depicted in Fig. 3b, where it can be verified
that higher biodiesel yields are reported in the literature using nano­
catalyst loadings between 4 and 6 wt% and 3 up to 9 catalyst cycles.
Regions with low and high yields are shown for 9 to 13 cycles, where the
literature needs other information.

3.1. Incorporation of metal oxides as nanocatalysts

Metal oxides based on zirconium (Zr), zinc (Zn), titanium (Ti), iron
(Fe), molybdenum (Mo), cerium (Ce), and tungsten (W) have a signifi­
cant role in the biodiesel production using nanocatalysts (Gawande
et al., 2012). In addition, these metal oxides were considered sustainable
and environmentally friendly since they presented a simple separation
method (Gawande et al., 2012). They improved the catalytic activity
with longer reusability and shorter reaction time as a consequence of
more significant active sites (Gawande et al., 2012; Vasić et al., 2020).
Furthermore, using metal oxides can enhance the stability of catalysts
and decrease the catalyst synthesis cost (Sulaiman et al., 2019).
Mohebbi et al. utilized MoO3 metal oxide supported on B-ZSM-5
nanocatalyst to esterify FFA to biodiesel (Mohebbi et al., 2020). This
nanostructure reduced acid site strength and increased the concentra­
tion of acid sites of the nanocatalyst. It also promoted higher surface
area, better crystalline size, and massive mesoporous structure. The
reaction using 25%-MoO3/B-ZSM-5 nanocatalysts dramatically
enhanced the catalytic performance and offered the optimum yield of
98%. Thus, this nanocatalyst presented excellent stability and reus­
ability (up to 6 cycles). It denotes that metal oxides positively influence
the activity and stability of nanocatalysts.
Meanwhile, Niu et al. optimized the development of dolomite
nanocatalysts for biodiesel production by incorporating cerium metal
Fig. 3. The yield obtained for different operation conditions of catalyst/
enzyme loading and catalyst cycles: (a) results for heterogeneous catalysts
oxide (CeO2) to ameliorate the activity and stability of the catalyst. The
prospection in Table 2; (b) results for nanocatalysts prospection in Table 3 cerium-doped dolomite nanocatalyst had increased activity and robust
(Section 3); (c) results for nanobiocatalysts prospection in Table 4 (Section 4). synergistic, leading to the reusability of the nanocatalyst up to 5 cycles.
Data boundary contour plots constructed in Microcal Origin Pro software, using The study obtained the highest biodiesel yield of 97.21% at 65 ◦ C, 0.05
103 total points increase factor and 10-3 smoothing parameter. Points are wt% catalyst loading within 2 h.
shown as - symbols. Fe3O4–Al2O3 was synthesized with 2 M sulfuric acid since it is widely
used as an active phase in the catalysts to improve the property of acid
strength (Mansir et al., 2017; Kazemifard et al., 2018; Wang et al.,
2019), leading to better catalyst activity and higher surface area during
transesterification reaction of microalgae (Safakish et al., 2020).

8
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 3
The role of nanocatalysts in transesterification for biodiesel production.
Catalysts Oils dp (nm) Process parameters Biodiesel Novelty Ref.
o
T ( C) t (h) Catalyst Y (%) Catalyst
loading cycles
(wt.%)

SO2−
4 / Microalgae 9.25–12.44 m2/ 120 4 8 87.6 5 Fe:Al ratio gave a good Safakish et al.
Fe3O4–Al2O3 g (surface area) impact on enhancing (2020)
catalytic activity and
stability
Sugar cane Palm fatty acid 1.95–3.92 60 1.5 2 98.6 7 The catalyst was very stable Akinfalabi et al.
bagasse (SCB) distillate (PFAD) up to 500 ◦ C, had good (2020b)
structure surface, and was
able to supply the optimum
yield of biodiesel
AC/CuFe2O4 on Waste chicken fat <50 65 4 3 95.6 – The properties of resulting Seffati et al.
CaO biodiesel were conformable (2020)
with international biodiesel
standard (ASTM D-6751 and
EN 14214), but it was not
sensibly used in cold
weather due to bad pour
point (3 ◦ C)
NaAlO2/γ-Al2O3 Palm oil 167–589 64.72 – 10.89 97.65 6 The deactivated catalyst Zhang et al.
could be regenerated with (2020d)
small amount of active sites
and further maintained the
yield of 93.29%. Moreover,
the reaction was
environmentally friendly
and low cost due to the use
of green and cheap catalyst
MoO3/B-ZSM-5 Free fatty acid 2.40 160 6 3 98 6 The catalyst had excellent Mohebbi et al.
(FFA) stability and reusability and (2020)
25% of MoO3/B-ZSM-5
dramatically enhanced the
catalytic performance
CaO from snail Hydnocarpus 50,000 61.6 and 2.4 0.892 98.93 and 5 The first order kinetic Krishnamurthy
shell wightiana oil and 58.56 and and 96.929 reaction was found with et al. (2020)
dairy waste scum 2 0.866 activation energy of 73.15
kJ/mol and 67.21 kJ/mol
and with frequency factor of
4.59 × 109/min and 5.182
× 108/min
Na2O doped CNTs WCO 250 65 3 3 97 3 Lewis acidity on the CNTs Ibrahim et al.
catalyst had a very positive (2020)
impact and became one of
the mostly used catalyst in
the biodiesel process
Fe3O4/Cs2O Tannery waste oil 10–3,500,000 65 5 7 97.1 9 The first order kinetic Booramurthy
reaction was appointed with et al. (2020)
activation energy and
frequency factor of 43.8 kJ/
mol and 7.5 × 104/min
Cu-supported Palm oil 80–150 45 <1 2–3 90.93 – TiO2 impregnated by Cu was De and Boxi
TiO2 potentially trusted to (2020)
become novel nanocatalyst
and produced biodiesel
having properties similar
with ASTM standard
CaO-doped MgO WEO 6.50 m2/g 69.37 7.08 4.571 98.37 6 CaO-doped MgO was Foroutan et al.
(surface area) appropriately harnessed in (2020)
and 0.022 cm3/ the biodiesel production due
g (porous to having crystalline
volume) structure on the surface area
catalyst with meso porosity
leading to producing the
best yield
Ca(OCH3)2 Nannochloropsis 250,000 80 3 3 99 5 Nanocatalyst of Ca(OCH3)2 Teo et al. (2016)
sp gave perfect catalytic
activity
CaO Algal oil Powder 55 – 1.25 96.3 – Algal oil and nano CaO Siva and
catalyst obtained from Marimuthu
calcination of egg shell (2015)
could effectively used in
transesterification with
solvent hexane
(continued on next page)

9
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 3 (continued )
Catalysts Oils dp (nm) Process parameters Biodiesel Novelty Ref.

T (oC) t (h) Catalyst Y (%) Catalyst


loading cycles
(wt.%)

CaO–Al2O3 Jatropha curcas oil 29.9 100 3 0.1 mg 82.3 3 Methanol:oil ratio, 5:1, was Hashmi et al.
implemented in the reaction (2016)
to optimize the resulting
yield of biodiesel
Li–CaO Jatropha oil 1.7 m2/g 65 2 5 >99 Non- 1.75 wt% of lithium (Li) was Kaur and Ali
(surface area) reusability impregnated into nano CaO (2011)
to convert jatropha oil
Li–CaO Karanja oil 1.7 m2/g 65 1 5 >99 Non- 1.75 wt% of lithium (Li) was Kaur and Ali
(surface area) reusability impregnated into nano CaO (2011)
to convert karanja oil
Heteropoly acid Madhua indica oil 5–29 50–60 5 2g 98 Several Heteropoly acid coated nano Thangaraj and
coated ZnO ZnO could be commercially Piraman (2016)
applied in the future process
due to a cheap catalyst
Cu–ZnO (CZO) Neem oil – 55 1 10 97.18 6 The reaction was optimized Gurunathan and
over exceptionally porous Ravi (2015b)
and non-uniform surface
area of nano CZO and the
first order kinetic reaction
studied was suitable with
experimental data (R2 =
0.9452) with activation
energy of 233.88 kJ/mol
Cs–MgO Olive oil 12.2–22.8 90 24 2.8 93 – Cs doped into MgO Woodford et al.
synthesized over (2014)
supercritical sol gel
treatment was trusted to
escalate the catalytic
transesterification
TiO2–ZnO Palm oil 34.2 60 5 200 mg 92.2 – TiO2–ZnO was a stable Madhuvilakku
catalyst, had a good catalytic and Piraman
activity, and could be (2013)
potentially applied in the
industry scale
ZnO Palm oil 28.4 60 5 250 mg 83.2 – ZnO contributed a worse Madhuvilakku
performance than that using and Piraman
TiO2–ZnO (2013)
Fe/ZnO Pongamia oil – 55 <1 12 93 4 Fe/ZnO was a promising Baskar et al.
candidate in biodiesel (2016)
reaction to reduce the
production cost because the
properties of resulting
biodiesel were similar with
ASTM-D6751 of commercial
biodiesel
MgO/MgFe2O4 Sunflower oil 10 110 4 4 91.2 5 MgO/MgFe2O4 synthesized Alaei et al.
over fuel ratio of 1.5 played (2018)
a significant role in the best
activity and was very
compatible in biodiesel
process
KOH/CaAl2O4 CAO 38.9–48.1 m2/g 65 4 3 96.7 3 Surface area and properties Nayebzadeh
(surface area) of catalyst were significantly et al. (2016)
enhanced by KOH doping
Al-MCM-41 Sunflower oil 50–200 70 8 10 84.2 4 Ultrasound irradiation and Vardast et al.
higher calcium content were (2019)
implemented in the reaction
to upgrade the surface area
and the yield of biodiesel
ZrO2/C4H4O6HK SBO 10–40 60 2 6% 98.03 5 Nano C4H4O6HK assisted Qiu et al. (2011)
with ZrO2 had longer
durability and produced
biodiesel that could applied
in the engine due to similar
properties with commercial
diesel fuel
MgO modified SBO 24.7–30.2 150–225 <1 0.1–7 95 – The higher reaction (Mguni et al.,
with Titania temperature did not cause 2012; Akia et al.,
MgO loss but increased the 2014b)
conversion
Sr3Al2O6 SBO – 60 1 1.3 95.2–96.2 – The highest yield of Rashtizadeh
biodiesel was achieved over et al. (2014)
(continued on next page)

10
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 3 (continued )
Catalysts Oils dp (nm) Process parameters Biodiesel Novelty Ref.

T (oC) t (h) Catalyst Y (%) Catalyst


loading cycles
(wt.%)

the appropriately operating


conditions
Hydrotalcites SBO – 65 4 2.5 97.1 – The highest yield of Rashtizadeh
Mg/Al biodiesel was achieved over et al. (2014)
the appropriately operating
conditions
Sr(NO3)2/ZnO SBO – 65 5 5 94.7 – The highest yield of Rashtizadeh
biodiesel was achieved over et al. (2014)
the appropriately operating
conditions
Benzyl bromide- RSO 0.07–0.09 65 3 5 99.2 – Benzyl bromide-doped CaO Tang et al.
assisted CaO produced an excellent yield (2013)
of biodiesel which was much
higher than the one obtained
from literature data and had
a perfect stability due to
preferable diffusion to
surface area of the catalyst
Pseudomonas SBO 50 37 12 200 mg 93 3 This recyclable and Andrade et al.
cepacia- environmentally-friendly (2016)
modified with catalyst was very suitably
Fe3O4 coated by used in the biodiesel
polydopamine production to mitigate the
environment issue
KF/CaO–Fe3O4 Stillingia oil 50 65 3 4 >95 14 KF/CaO–Fe3O4 was one of Hu et al. (2011)
the promising catalysts to
use in the reaction because it
had specific porous structure
leading to the better
performance
Ti(SO4)O WCO 44.45 m2/g 75 3 1.5 97.1 8 Ti(SO4)O had better Gardy et al.
(surface area) catalytic activity and (2016)
stability, due to the presence
of SO2−
4 species, than that
explored by other
experiments using titania
catalyst
CsH2PW12O40/ Sunflower oil 38–42 60 4 4 81 – Deactivation of catalyst Feyzi et al.
FeSiO2 could be easily solved by (2014)
using external magnet in
very rapid time in which did
not require high price
ultracentrifugation
CaO/NaX-Zeolite Sunflower oil – 60 6 10 93.5 – The obtained biodiesel Luz Martinez
satisfied the properties (i.e. et al. (2011)
viscosity, flash point, and
acid number) of European
Norm standard (EN-14214
KF/CaO TSO 30–100 65 2 3 96.8 – Novel crystal of KCaF3 in the Wen et al.
catalyst highly upgraded the (2010a)
catalytic activity and
stability leading to
achieving the best yield of
biodiesel
Cs–Ca/SiO2–TiO2 Vegetable oil 45 60 2 2 98 – Particular surface area, pore Feyzi and
diameter size, pore volume Shahbazi (2015)
and active site content over
the surface of catalyst had a
positive effect toward the
stability and capability of
catalyst
MgO/MgAl2O4 SFO 2.4–21.3 110 3 3 95 6 Urea utilized in the biodiesel Vahid and
production contributed well Haghighi (2017)
in the obtained yield of
biodiesel with stable
reaction up to 6 cycles
CaO–MgO WCO 36–300 m2/g 180 6 3 98.95 – Nano-size CaO played an Tahvildari et al.
(surface area) important role to support (2015)
MgO in the
transesterification due to its
specific structure surface
area and pore diameter size
(continued on next page)

11
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 3 (continued )
Catalysts Oils dp (nm) Process parameters Biodiesel Novelty Ref.

T (oC) t (h) Catalyst Y (%) Catalyst


loading cycles
(wt.%)

Cu–ZnO (CZO) WCO 80 55 <1 12 97.71 5 CZO was efficiently and Gurunathan and
effectively harnessed to Ravi (2015a)
enhance the yield by dint of
a larger surface area and
highly active site
MgO–La2O3 SFO – 65 5 3 97.7 4 Activation energy (77.6 kJ/ Feyzi et al.
mol) was obtained over the (2017)
best nanocatalyst with high
yield of biodiesel and the
reaction was a second order
reaction according to
pseudo-first and pseudo-
second order models
Co–ZnO (CZO) MFO – 60 3 2.5 98.03 4 The CZO denoted Borah et al.
impeccable catalytic activity (2019)
and stability and could be
utilized in wide-scale
biodiesel production with
high yield

Note: AC = Activated Carbon; CNT = Carbon Nano Tube.

Different Fe to Al ratios of 0.15 and 0.35 positively impacted catalyst acetate. It produces a higher surface area (from 4.76 m2/g to 8.39 m2/g)
efficiency (Safakish et al., 2020). The synthesized catalyst resulted in with a particle size of 12 nm, leading to increase biodiesel yield. The
87.6% biodiesel yield at 120 ◦ C, 8 wt% catalyst loading within 4 h and highest biodiesel yield produced was 98.2% obtained at 180 ◦ C, with
could be reused up to 5 runs with slight catalyst deactivation, which was catalyst loading of 9.3 wt% within 6 h. Evaluation of the mixed metal
14.2%. Since poor catalyst reusability is generally caused by Groups 1 oxides such as Mg–Fe on Mg0⋅25Fe2⋅75O4 nanocatalyst synthesized over
and 2 metal cations (e.g., Na+, K+, Ca2+, Mg2+) (Shao et al., 2013; Rad the combustion method using urea-nitrate was also investigated (Sar­
et al., 2018), aluminum is considered a suitable metal to be combined vestani et al., 2020). From the study, different urea contents dramati­
with the sulfate group due to abundance, inexpensive, higher thermal cally influenced the properties of nanocatalysts, such as textural,
stability and surface area, and broader pore size (Vahid et al., 2018; crystalline size, particle size, morphology, and stability. Using a higher
Ghalandari et al., 2019). Hence, Al is applied in synthesizing nano­ amount of urea reduced crystalline size causing a higher surface area of
catalysts to enhance the activity (Gardy et al., 2018, 2019) and spinel 23.6 m2/g, a higher pore volume of 0.0950 cm3/g, and particle size
ferric content doped in Al metal (Ashok et al., 2019; Naderi and ranges of 25–35 nm leading to better catalytic performance.
Nayebzadeh, 2019). Metal oxide of CoO–NiO doped on sulfated ZrO2 (CN/SZ) nano­
Graphene oxide (GO) was synthesized with sulfuric acid to convert catalyst synthesized under the co-precipitation impregnation method
microalgae into biodiesel at 90 ◦ C using 5 wt% catalyst loading within was performed by Singh et al.. Two nanocatalysts, sulfated ZrO2 (SZ)
40 min. The highest obtained yield was reported at 95.1% (Cheng et al., and pure ZrO2 (Z), were compared in the transesterification of triglyc­
2016, 2017). Ibrahim et al. employed sodium oxide (i.e., Na2O) eride to evaluate the role of metal oxide in the reaction. The report
impregnated on carbon nanotubes (CNTs) catalyst followed by calci­ claimed that the CN/SZ nanocatalyst had high acid and base sites and
nation at 500 ◦ C within 3 h (Ibrahim et al., 2020). The metal oxide of produced the optimum biodiesel yield (98.8%) at 65 ◦ C, catalyst loading
Na2O escalates the performance of CNTs nanocatalysts in consequence of 0.2 wt%, and methanol:oil ratio of 3:1 within 2 h. The nanocatalyst
of Na + cation as a positive metal ion having Lewis acid sites and O2− could be further reused until five reaction cycles with 96.8% biodiesel
anion as a negative metal ion possessing Brønsted acid sites. Thus, the yield were achieved after the last run. The metal oxide of CoO–NiO
Na2O-doped CNTs nanocatalyst resulted in a high biodiesel yield of 97% positivly affects the biodiesel yield and maintains the stability of the
and was reused for up to 3 runs. In addition, Lewis acidity on the CNTs nanocatalyst.
catalyst showed a positive impact and became one of the most used In addition, Kaur and Ali studied the impregnation of 1.75 wt%
catalysts in the biodiesel process. lithium (Li) onto a CaO nanocatalyst for the transesterification of Kar­
Another metal oxide, namely copper oxide (CuO), was used for the anja and Jatropha oils (Kaur and Ali, 2011). The Li-doped CaO nano­
transesterification of palm oil. De and Boxi impregnated TiO2 with CuO, catalyst gave a high yield for both oils at 65 ◦ C with catalyst loading of 5
resulting in nanocatalysts with particle sizes of 80–150 nm (De and Boxi, wt%. In another study by Gurunathn and Ravi, zinc oxide (ZnO) nano­
2020). High biodiesel yield (90.93%) was achieved at 45 ◦ C, using a catalyst was impregnated with copper salt named CZO nanocatalys with
methanol to oil ratio of 20:1 within 45 min. In their work, catalyst a particle size of 80 nm. The highest biodiesel yield (97.71%) was ob­
loading had a significant effect in increasing the biodiesel yield. The tained at 55 ◦ C, with catalyst loading of 12 wt% and methanol:oil ratio
biodiesel yield without adding Cu content was less than 60% of biodiesel of 1:8 for 50 h (Gurunathan and Ravi, 2015a). The CZO nanocatalyst has
production (De and Boxi, 2020). In contrast, the biodiesel yield slowly also been optimized over the highly porous and non-uniform surface
dropped to less than 75% when a higher Cu content of 3–5 wt% was area of the nanocatalyst (Gurunathan and Ravi, 2015b). The optimized
employed. This phenomenon occurred because of the incompatibility of CZO nanocatalyst could be reused up to six reaction times with an
the TiO2 nanocatalyst with Cu, and the stoichiometry imbalance due to activation energy of 233.88 kJ/mol (Gurunathan and Ravi, 2015b).
higher Cu contents (De and Boxi, 2020), leading to structural instability Laskar et al. impregnated zinc oxide (ZnO) with silver nanoparticles
and discrepancy in the nanocatalyst framework (Boxi and Paria, 2014). (Ag NPs) for the transesterification of palm oil into biodiesel (Laskar
On the other hand, Hashemzehi et al. evaluated mixed metal oxides et al., 2020). The ZnO–Ag NPs catalyst produce 97% of biodiesel yield
of Cu0⋅4Zn0⋅6Al2O4 catalyst (Cu–Zn–Al) synthesized over microwave reached at 60 ◦ C, 10 wt% of catalyst content, methanol:oil ratio of 1:10
combustion method (MCM) using 67% urea and 45% ammonium within 1 h, and could be recycled during five reaction runs (Laskar et al.,

12
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 4
The potentiality of enzymes as nanobiocatalysts in biodiesel production.
Lipase Lipid source Treatment of enzymes Reactor Reaction conditions Results Ref.
Types o
T ( C) Enzyme X (%) Y (%) Cycles
loading (wt.
%)

PFL SBO Suspension and immobilization BR 37 – – 70–95 – Ferrero et al. (2020)


CALA Waste frying palm Received from Novozymes A/S GR 30 5.5 – ±96 – Guo et al. (2020)
oil (FPO) (bagsvaerd, Denmark).
Serratia sp. ISTD04 Microbial oil Immobilization TNR 30 0.11 g – 91.50–97.41 5 and Kumar and Thakur
lipids and 10 (2020)
Pseudomonas sp.
ISTPL3 lipase
CRL Crude eruca sativa Provided by Sigma Chemical SI 21 5 mg – >60 – Aghababaie et al.
oil company (2020)
TLL immobilized on FWO Provided by Novozyme BR 35 10 – 75.3 – Ching-Velasquez
octadecyl (Mexico) et al. (2020)
metacrylate
beads
Lipase enzyme WCO Washing, reacted by n-hexane, – 35–60 1.5 – 20, 40, and – Jayaraman et al.
drying, and immobilization 60 (2020)
PCL Castor oil Provided by Sigma-Aldrich TNR 50 10 78 75 6 Ravi et al. (2020)
company
A recombinant FHL Crude plant oils Immobilization with an BR 30 0.2 g – 98.8 9 Quayson et al.
activated carbon (2020b)
Immobilized ROL SPO Immobilization with alginate- EF 30–50 2 – 91.30 15 Muanruksa and
polyvinyl alcohol (PVA) beads Kaewkannetra
(2020)
ROL Chlorella vulgaris Fe3O4 superparamagnetic BR 30 35 69.8 – 5 Nematian et al.
microalgae nanoparticles (MNPs) was (2020a)
assisted by ROL
PPL SBO Immobilization with p- BR 45 5.16 96.5 – 10 Rial et al. (2020)
nitrobenzyl cellulose xanthate
Immobilized CALB Yellow horn seed Immobilization with MR 35–60 1–6 – 92.3 5 Zhang et al.
oil polyporous magnetic cellulose (2020a)
beads (PMCBs)
NTVL WCO Extraction and purification EF 45 30 kUnit 96.5 – – Patchimpet et al.
(2020)
Lipases Microbial oil Multi treatments – – – – >90 – Hossain et al.
(2020)
TLL (Lipozyme TL RSO Provided by the global ER 30 5 99.92 99.89 – Santaraite et al.
IM) biotechnology company (2020)
Novozymes A/S (Copenhagen,
Denmark)
TLL Nannochloropsis N.oculata cells were BR 25–40 0.5–2 – 90.24 2 He et al. (2020)
strains (N. oculata synthesized over four mixed
cells) enzymes namely cellulase,
hemicellulase, papain and
pectinase
TLL (Lipozyme RM Crude palm oil Supplied by Novozymes Latin BR 50 3 – 25 – Matassoli et al.
IM and Lipozyme America Ltda (Parana, Brazil) (2009)
TL IM)
TLL Palm oil Synthesized over protein- BR 45 20 – 82.1 8 Raita et al. (2010)
coated microcrystals (PCMCs)
TLL (Lipozyme TL Crude palm oil Immobilization with silica gel BR 30 6.67 – 85.01 – Sim et al. (2010b)
IM)
TLL (Lipozyme TL Crude palm oil Provided by Novozymes BR 40 6.65 – 100 – Sim et al. (2009)
IM) (Bagsvaerd, Denmark).
TLL (Lipozyme TL Crude palm oil Immobilization with silica gel BR 30 6.67 – 81.73 and – Sim et al. (2010a)
IM) (supplied by Novozymes, and 96.15
Bagsvaerd, Denmark). 40
Chlorococcum sp. Microalgae Treated over supercritical – 60 7.1 – 98 – Halim et al. (2011)
carbon dioxide (SCCO2) and
extraction 80
Burkholderia sp. Chlorella vulgaris Immobilization with alkyl- BR 25–50 240–1804 – 97.3 6 Tran et al. (2012)
C20 ESP-31 grafted Fe3O4– SiO2 U/g
CALB (Novozym Chlorella sp. KR-1 Immobilization with BR 60 38.9 – 90 10 Lee et al. (2013)
435) macroporous resin and DMC
Candida sp. 99− 125 Chlorella Immobilization with BR – 75 98.15 – – Li et al. (2007)
protothecoides adsorption method
CALB (Novozym WCO Immobilization with ionic BR 50 40 – 72 – Taher et al. (2019)
435) liquid of [bmim][PF6]
CALB (Novozym Microalgae Immobilization with ionic BR 50 5 – 86.2 – Lai et al. (2012)
435) (C. pyrenoidosa) liquid of [bmim][PF6]
PEL Microalgae Immobilization with ionic BR 50 10 – 90.7 – Lai et al. (2012)
(C. pyrenoidosa) liquid of [bmim][PF6]
PEL Corn oil BR 40 20 – 86 – Zhang et al. (2011)
(continued on next page)

13
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 4 (continued )
Lipase Lipid source Treatment of enzymes Reactor Reaction conditions Results Ref.
Types
T (oC) Enzyme X (%) Y (%) Cycles
loading (wt.
%)

Immobilization with ionic


liquid of [bmim][PF6]
PCLB SBO Immobilization with ionic BR 40 5 – >80 – Abrahamsson et al.
liquid of [bmim][NTf6] (2015)
CALB (Novozym SBO Immobilization with acrylic PBR 52 5 – 75.2 – Shaw et al. (2008)
435) resin beads
CALB (Novozym Triolein oil Immobilization with ionic BR 60 28 mg – 96 3 Lozano et al. (2010)
435) liquid of [omim][NTf2]
CALB (Novozym Triolein oil Immobilization with ionic BR 60 10 – 98 – De Diego et al.
435) liquid of [C16mim] [NTf2] (2011)
CALB (Novozym Triolein oil Immobilization with ionic BR 60 10 – 98 – De Diego et al.
435) liquid of [C18mim] [NTf2] (2011)
Candida sp. lipase Microalgae Immobilization with MR 35 10 66.55 90.81 4 Guldhe et al. (2015)
and PFL (Scenedesmus immobead 150
obliquus)
Candidia sp. 99–125 Microalgae Immobilization with BR 38 30 – 98 – Xiong et al. (2008)
lipase (Chlorella macroporous resin
protothecoides)
ANL and CALB Schizochytrium sp Immobilization with BR 45 1.5 – 95 – Tian et al. (2016)
macroporous acrylic resin
Candida sp lipase WCO Immobilization with co-fixing FBR 45 5–35 – 91.08 – Chen et al. (2009)
agents
TLL SBO Immobilization with BR 30 15 – 75–100 – Rodrigues et al.
aldehyde− lewatit (Lew-TLL) (2010)
RML and PCYL SBO Immobilization with two BR 30 – – 68.5–95 – Guan et al. (2010)
different primers
PEL WCO Immobilization with resin BR 35 – – 92.8 10 Li et al. (2009)
D4020
BCL Jatropha oil Immobilization with n-butyl- BR and 40 – – 90 and 95 – Kawakami et al.
substituted hydrophobic silica CSTR (2011)
monolith
PCL Jatropha oil Immobilization with celit BR 50 4–5 – 98 4 Shah and Gupta
(2007)
TLL (Lipozyme TL WCO Immobilization with BR 25 4 – 93 7 Yagiz et al. (2007)
IM) hydrotalcite and four types
zeolites
PCL Madhuca indica Immobilization with cross- SI 40 25–50 mg 96–99 – – Kumari et al.
linked enzyme aggregates (2007)
(CLEAs) and protein-coated
microcrystals (PCMCs)
TLL (Lipozyme TL SBO Immobilization with amino- EF 50 40 90 – 4 Xie and Ma (2009)
IM) functionalized magnetic
nanoparticles (Fe3O4)

Note: SI = Shaker Incubator; EF = Erlenmeyer Flask; DMC = Dimethyl Carbonate; [BMIm][PF6] = 1-butyl-3-methylimidazolium hexafluorophosphate; [bmim][NTf6]
= 1-butyl-3-methylimidazolium trifluoromethanesulfonyl; PBR = Packed Bed Reactor; [omim][NTf2] = 1-methyl-3-octylimidazolium bis(trifluoromethylsul-fonyl)
imide; [C16MIM][NTf2] = 1-hexadecyl-3-methylimidazolium triflimide; [C18MIM][NTf2] = 1-octadecyl-3-methylimidazolium triflimide; X(%) = percent biodiesel
conversion; Y(%) = percent biodiesel yield.

2020). Meanwhile, Kumar and Ali implemented a Zn-doped CaO catalyst can improve catalyst performance. Unfortunately, in these
nanocrystalline catalyst for the transesterification of cotton seed oil and works, few studies discussed the stability and reusability as well as the
reported that the Zn-doped CaO catalyst was the most efficient for the concern regarding scale-up processes from bench-scale experiments to
biodiesel production using low-grade oils with high FFAs content in the pilot scale. Thus, in the following two sections, both subjects are
presence of Zn metal (Kumar and Ali, 2013). addressed.
The mixed metal oxide of TiO–ZnO assisted with ZnO nanocatalyst
was carried out by Madhuvilakku and Piraman for the transesterification
of palm oil into biodiesel (Madhuvilakku and Piraman, 2013). In their 3.2. Nanocatalyst synthesis method towards stability and reusability
work, the mixed metal oxide TiO–ZnO nanocatalyst revealed a highly
stable and effective activity with a resulting conversion of 98%. On the The synthesis method of the nanocatalyst is a crucial strategy to
contrary, Wen et al. (2010b) utilized TiO2-assisted MgO catalyst to improve its activity toward efficient biodiesel production. Accordingly,
convert waste cooking oil (WCO) into biodiesel. The TiO2–MgO catalyst different synthesis methods play an essential role in the resulting bio­
exhibited a high performance due to improved stability caused by a diesel yield and the cycle of nanocatalyst reusability, as shown in Fig. 2.
synergistic effect of Ti ions and Mg ions. Therefore, the metal oxides are Akinfalabi et al. transformed sugar cane bagasse (SCB) into biochar
frequently harnessed as an essential agent on the nanocatalysts to uning chlorosulfonic acid. They tested four different sulfonation pa­
improve the catalytic performance, activity, and stability (Alaei et al., rameters: sulfonation temperature, time, volume, and SCB-biochar
2018; Vardast et al., 2019; Ingle et al., 2020). mass. The authors obtained the best sulfonation parameters at 300 ◦ C,
Regarding the achievements presented in the last paragraphs, it 5 h, 200 ml, using 2 g of SCB-biochar. The surface area and pore volume
should be noted that incorporating inorganic metal oxides into the of nanocatalysts were 98–298 m2/g and 0.09–1.17 cm3/g, respectively,
with particle sizes ranging from 1.95 to 3.92 nm. The sulfonation-

14
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

synthesized nanocatalyst was used in a palm fatty acid distillate (PFAD)


reaction for biodiesel production. It generated the optimum biodiesel
yield (98.6%) at 60 ◦ C, catalyst loading of 2 wt%, and methanol:oil ratio
of 10:1 within 1.5 h and could be reused until seven reaction runs. This
outstanding performance of the nanocatalyst was correlated with the
best-chosen synthesis method.
Alaei et al. compared two different synthesis methods, namely con­
ventional and hybrid, for the transesterification of SFO using MgO/
MgFe2O4 nanocatalyst. Methods used were impregnation, precipitation,
precipitation-hydrothermal, precipitation-ultrasonic, and combustion.
The morphology of the nanocatalyst synthesized over precipitation and
precipitation-hydrothermal methods shows a similarity to the sheet. In
contrast, the one synthesized over the combustion method yielded
smaller particle sizes, such as low specific surface area, porosity, and
pore volume. Among the types of synthesis methods, the combustion
method was superior because it contributed well with the highest bio­
diesel yield (92.9%). In addition, the catalyst synthesized under the
combustion method was reused up to 5 cycles and was quickly separated
in the last process. The effect of fuel ratio in combustion synthesis was
also observed. The applied pore size (10 nm) and surface area (97.8 m2/ Fig. 4. Effect of different nanocatalyst synthesis methods towards bio­
g) gave the highest conversion of 91.2% achieved at 110 ◦ C, methanol: diesel yield.
oil ratio of 12, and catalyst loading of 4 wt% within 4 h (Alaei et al.,
2018). and was reused up to 5 runs with slight catalyst deactivation. Mean­
Hashemzehi et al. studied the effect of a microwave combustion while, Siva and Marimuthu investigated the transesterification of
method (MCM) on the activity of zinc-copper aluminate (ZCA) nano­ non-edible oils using CaO derived from eggshell powder calcined under
catalyst for the esterification reaction. The MCM indicated quicker water a sonochemical reactor with hexane solvent (Siva and Marimuthu,
evaporation, leading to a high crystalline structure with a larger surface 2015). The best yield (96.3%) was obtained at 55 ◦ C and 1.25 wt%
area which increased the catalytic performance. The pore diameter of catalyst loading.
the ZCA nanocatalyst increased from 2.1 nm to 9.5 nm and showed A top-down method of nanocatalyst synthesis was studied by Hashmi
optimum activity. The MCM-synthesized ZCA nanocatalyst produced the et al. for the transesterification of Jatropha oil using CaO–Al2O3 nano­
highest yield of 99.1% at 180 ◦ C, 9:1 methanol:oil ratio, 3 wt% catalyst catalyst with a particle size of 29.9 nm. This nanocatalyst generated a
loading, and up to 9 cycles reusability. It exhibits that the microwave lower yield of 82.3% at a 5:1 methanol:oil ratio and was reused 3 times.
combustion method (MCM) had a significant role in nanocatalyst syn­ In another study, the ZnO nanocatalyst synthesized over co-
thesis. The combustion method was also employed to synthesize MgO/ precipitation followed by calcination was implemented (Thangaraj
MgAl2O4 nanocatalyst using glycerine, urea, ethylene glycol, and citric and Piraman, 2016). The transesterification of Madhuca indica oil suc­
acid fto transesterify SFO as executed by Vahid and Haghighi (2017). cessfully achieved 95% conversion after 5 h. Feyzi et al. also imple­
The nanocatalyst treated by urea presented a remarkable capability, mented the co-precipitation method in synthesizing the MgO–La2O3
resulting in high biodiesel yield (95%), and up to 6 cycles of reusability. nanocatalyst. Different Mg:La ratios were studied for the trans­
The sono-hydrothermal method with carbon and polymeric tem­ esterification of sunflower oil (SFO). The higher Mg:La ratio enhanced
plates using mesoporous SAPO-34 was explored by Ebadinezhad et al. in the catalytic performance, and the best Mg:La ratio of the MgO–La2O3
biodiesel production. SAPO-34 synthesized over activated carbon indi­ nanocatalyst was 60 wt%.
cated better catalytic performance and produced a higher conversion of The effect of thermal treatment for catalyst synthesis was executed
oleic acid of 83%. The conversion increased to 94% after the nano­ successfully by Dahdah et al. using Ca–Mg–Al catalyst synthesized over
catalyst was impregnated with CeO2 (5%) using the sono-hydrothermal two different techniques, namely co-precipitation and impregnation
method. The CeO2 (5%)-synthesized SAPO-34 catalyst showed the methods (Dahdah et al., 2020). However, the thermal treatments could
highest reaction rate and further resulted in the highest conversion due destroy the structural framework of the catalyst and can hinder active
to the penetration of wide oil particles and escalated acidity. catalyst sites leading to a decrease in the biodiesel yield and the catalyst
On the other hand, Dehghani and Haghighi employed impregnation performance (lower Lewis acid sites, poor reusability, lower conversion,
and sono-dispersion to synthesize MCM-41 impregnated with Zr at yield, and selectivity) (Helwani et al., 2016; Ma et al., 2016). However,
different Si:Zr ratios for the esterification and transesterification of the Ca-impregnated Mg–Al catalyst resulted in the highest biodiesel
WCO. The catalyst particle size synthesized via sono-dispersion had a yield of 95% reached at 60 ◦ C, 400 rpm, 2.5 wt% of catalyst content
nanoscale surface area, and this method successfully doped CaO effi­ within 6 h. In addition, Wen et al. used the impregnation method to
ciently compared to the impregnation method. Moreover, the sono- synthesize KF-CaO nanocatalysts. As a result, the impregnated KF-CaO
dispersion-synthesized catalyst presented better reusability than the nanocatalyst showed an increasing capability due to the synergistic ef­
other (Dehghani and Haghighi, 2020). The sono-dispersion method of fect of KF and CaO after impregnation, resulting in a 96.8% biodiesel
Al-MCM-41 synthesis impregnated with calcium (Ca) was likewise yield (Wen et al., 2010a).
implemented by Vardast et al. (2019). The effect of ultrasound waves on Previously, Mahmoud employed a hydrothermal method for the
the catalytic performance was observed. The ultrasound irradiation used catalyst synthesis method assisted over ultrasonic irradiation at various
in Ca/Al-MCM-41 nanocatalyst synthesis upgraded the surface area of times. The enlarged active surface sites, porosity, acidity, and activation
the nanocatalyst and the yield of biodiesel. energies were dramatically increased under the longer ultrasonication
Apart from the previous synthesis methods, hydrothermal, top- times. The optimum esterification conversion obtained was 90%
down, and co-precipitation methods have been implemented. For reached at 30 min of the ultrasonication time (Mahmoud, 2019).
instance, the hydrothermal method was implemented by Teo et al. for Meanwhile, Yee et al. investigated the influence of catalyst calcination
the transesterification of Nannochloropsis sp. Using calcium methoxide on the resulting biodiesel yield. The best calcination temperature
nano (Ca [OCH3]2) (Teo et al., 2016). This catalyst resulted in a yield of (490 ◦ C) and duration (4 h) had a crucial effect in increasing the
99% at 80 ◦ C, 3 wt% catalyst loading, 30:1 methanol:oil ratio within 3 h

15
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

biodiesel yield obtained (Yee et al., 2011). using green and cheap nanocatalysts. RSM method was also applied by
The effect of various synthesis methods for the nanocatalysts on the Krishnamurthy et al. to optimize operating parameters for trans­
catalytic performance and stability of nanocatalysts has been discussed esterification of Hydnocarpus wightiana oil (HWO) and scum oil (SO)
in recent literature, emphasizing that the appropriate catalyst synthesis using CaO nanocatalyst derived from snail shells (Krishnamurthy et al.,
method plays a crucial role in the catalyst activity and the high biodiesel 2020). The best yields produced were 98.93% (for HWO) and 96.929%
yield obtained (Ambat et al., 2018; Yousefi et al., 2018). Fig. 4 was (for SO) obtained at 61.6 and 58.56 ◦ C, methanol:oil ratios of 12.4:1 and
constructed based on case study in order to compare the biodiesel yield 12.7:1, catalyst loadings of 0.892 and 0.866 wt%, within 2.4 and 2 h, for
using various nanocatalyst synthesis methods. HWO and SO, respectively. The catalyst could be reused for up to 5
Fig. 4 indicates that microwave combustion can be considered the cycles. The yield reached by Zhang et al. was slightly higher than that
best method with the highest efficiency for biodiesel production and that gained by Krishnamurthy et al. using SO because Zhang et al. used
the top-down method needs further improvement to achieve industrial higher catalyst content and methanol:oil ratio. In contrast, the yield
goals despite being technologically tested. obtained from HWO was higher because this reactant might interact
well with the catalyst.
3.3. Bench scale operational conditions towards biodiesel yield Meanwhile, the CCD was also utilized by Foroutan et al. to trans­
esterify WEO into biodiesel using a CaO-doped MgO nanocatalyst
Operating conditions such as methanol:oil ratio, catalyst content, (Foroutan et al., 2020). The properties of the produced biodiesel were
reaction temperature, and process time can have a mutual positive or compared with the commercial international biodiesel standard. The
negative impact on oil conversion, biodiesel yield, and catalyst reus­ applied reaction parameters were various reaction times (1–9 h), tem­
ability. Therefore researchers always seek the best operating parameters peratures (40–80 ◦ C), methanol:oil ratios (6:1–18:1), and catalyst con­
to gain the best performance resulting in the maximum yield with longer tents (2–6 wt%), where the best parameter took place at 69.37 ◦ C,
catalyst reusability (see Fig. 3). catalyst loading of 4.571 wt%, methanol:oil ratio of 16.7:1 within 7.08 h
The reaction time correlates with the capability of the catalyst to producing the highest yield of 98.37% and up to 6 times reusability. The
produce the optimum biodiesel yield and also influences the deactiva­ CaO-doped MgO nanocatalyst was appropriately applied for biodiesel
tion of the catalyst. Generally, higher biodiesel yield is achieved at a production because of its mesoporous and crystalline structure.
longer reaction time due to the sufficient hydrolysis of ester groups. In conclusion, the best reaction parameters for the esterification and
However, catalyst deactivation occurs after a specific reaction time transesterification of vegetable oils with methanol were in the temper­
(Uzun et al., 2012). The effect of operating conditions such as methanol: ature range of 50–75 ◦ C, catalyst loading of 3–5 wt%, methanol:oil ratios
oil molar ratio, catalyst content, reaction temperature, and agitation of 12:1–20:1, and reaction times between 1 and 7 h as were also tested
speed toward the biodiesel yield produced is portrayed in Fig. 2. by others (Feyzi and Shahbazi, 2015; Gardy et al., 2016).
One of the essential factors in the esterification and trans­
esterification of vegetable oils is the reaction temperature. Lower reac­ 4. The role of nanobiocatalyst in biodiesel production
tion temperature generates a lower biodiesel yield due to the catalyst’s
lower activation energy. Meanwhile, high temperatures may induce Acid, alkaline and heterogeneous nanocatalysts are considered
deactivation of the biocatalyst, leading to a drop in the catalytic activity suitable for the biodiesel production through transesterification reaction
and stability. Based on the current literature, the optimum reaction (Atadashi et al., 2013; Guo et al., 2020). Nevertheless, the activity and
temperature was discovered at 65 ◦ C (Keihani et al., 2018; Sahani et al., stability of those catalysts are negatively affected by the high FFAs and
2019). water contents (Vieitez et al., 2014; Nayak et al., 2016; Zhou et al.,
Catalyst concentration is another crucial factor in biodiesel produc­ 2020), leading to more soap formation, as mentioned before. In addi­
tion. Optimum catalyst concentration results in a faster reaction rate due tion, high reaction temperature, longer reaction time, and high meth­
to the efficient hydrolysis of the esters group and saponification of tri­ anol:oil ratio were applied in the reaction, which can cause corrosion in
glycerides (Esmaeili and Foroutan, 2018). The higher catalyst loading in the reactor (Lotero et al., 2005, 2006; Romdhane et al., 2013). Conse­
the reaction provides more contact between the reactant and nano­ quently, high-grade oil possessing low FFAs (<1%) and water (<0.3%)
catalysts, thus enhancing catalytic performance. However, the incre­ contents must be precisely implemented to lower the soap formation
ment of catalyst loading favoring saponification reaction leads to the (Encinar et al., 2005; Amini et al., 2017; Adsul et al., 2020; Mihajlovski
constant yield of biodiesel (Kim et al., 2004; Yesilyurt et al., 2019). et al., 2020) in these processes.
Several researchers have employed 3 wt% of catalyst loading in their Alternatively, enzymes, especially immobilized, have an excellent
study (Dai et al., 2015; Tan et al., 2015), as seen in Fig. 3. prospect as nanobiocatalysts not only for the dye adsorption process
Seffati et al. have investigated several parameters for an effective and (Veeramalai et al., 2022) but also for biodiesel production because they
efficient reaction condition. Diverse methanol:oil ratios (4:1–24:1), can be utilized for oils with high FFAs and water contents, particularly
catalyst loadings (0.5–7 wt%), reaction temperatures (50–80 ◦ C), and immobilized enzymes (Boviatsi et al., 2020; Malakar et al., 2020; Melani
process times (0.5–6 h) has been performed, and the nanosized CuFe2O4- et al., 2020). For example, Guo et al. used Candida antarctica lipase A
impregnated activated carbon (AC) catalyst with a particle size of less (CALA) enzyme to produce biodiesel from waste frying palm oil. This
than 50 nm followed by encapsulating CaO was employed. The most enzyme was used as a green and inexpensive biocatalyst for high FFAs
effective and efficient reaction conditions were discovered at methanol: and water contents in the waste frying palm oil. As a result, the high
oil ratio of 12:1, 3 wt% catalyst loading, for 4 h reaction time at 65 ◦ C, biodiesel yield of 94.6% was attained at 30 ◦ C, 5.5 wt% enzyme loading,
and the highest biodiesel yield achieved was 95.6%. The properties of 7:1 methanol:oil ratio within 22 h.
the resulting biodiesel were comparable with international biodiesel It should also be pointed out that several immobilized enzymes are
standards (ASTM D-6751 and EN 14214). However, it was not suitable cheaper than conventional chemical catalysts. In addition, they have
for cold weather due to the low pour point (3 ◦ C)). excellent water solubility (Jegannathan et al., 2008; Lima et al., 2015;
In order to optimize biodiesel yields, Zhang et al. exploited central Andrade et al., 2019; Ching-Velasquez et al., 2020), which can decrease
composite design (CCD) correlated with a response surface methodology the mass transfer limitations and solve the problem of by-product
(RSM) in their study (Zhang et al., 2020d). The highest yield produced adsorption like glycerol (Andrade et al., 2016, 2019; Pollardo et al.,
was 97.65% (less than the predicted 98.47%) achieved at 64.72 ◦ C with 2018). In this sense, the list of potential enzymes used as a biocatalyst in
10.89 wt% catalyst loading and methanol:oil of 20.79. NaAlO2/γ-Al2O3 biodiesel production is summarized in Table 4.
nanocatalyst with particle sizes ranging from 167 to 589 nm was used up Table 4 tabulated several lipases studied using different lipid sources
to 6 cycles. The reaction is environmentally friendly and low-cost due to and treated using different treatments. These studies applied different

16
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

reactor types, enzyme loadings, and temperatures, giving different et al., 2020; Santos et al., 2020; Zhong et al., 2020). The effect of enzyme
conversion and yield results. Tables 2 and 3 also verified that biodiesel immobilization on its activity was exhibited by Santaraite et al. . The
production using nanobiocatalysts has gaps regarding information authors compared two immobilized enzymes (Lipozyme RM IM and
catalyst loading, conversion achieved, biodiesel yield, and the number Lipozyme TL IM) with three free enzymes (Lipolase 100 L, Lipozyme TL
of cycles. 100 L, and Lipex 100 L) in the transesterification of RSO. Among them,
Results of Table 4 are depicted in Fig. 3c, where it can be verified that the Lipozyme TL IM was the most effective, producing the highest de­
higher biodiesel yields are reported in the literature using enzyme gree of transesterification (97.74%), followed by the Lipozyme RM IM
loadings smaller than 10 wt% with 4 times catalyst cycles. On the other (84.40%), Lipex 100 L (73.35%), Lipolase 100 L (50.82%), and Lip­
hand, regions with lowest yields are shown for 10 cycles and above 40 ozyme TL 100 L (30.10%). Additionally, the biodiesel produced by the
wt% of enzyme loading, indicating that new research could fill these Lipozyme TL IM conceived the lowest glyceride content. Similar results
gaps in the future. employing Lipozyme TL IM enzyme that gave the best performance were
also reported previously (Matassoli et al., 2009; Thliveros et al., 2014;
4.1. Enzymes as nanobiocatalysts Musa, 2016).
Ferrero et al. utilized Pseudomonas fluorescens lipase immobilized
Apart from troubleshooting for high FFAs and water contents and the with sodium-assisted SBA-15 to produce biodiesel from SBO, JTO, and
adsorption of by-products, the biodiesel produced using enzymes can be SFO in a batch reactor at 37 ◦ C. This immobilized enzyme produced
likewise carried out at mild reaction temperature (Rial et al., 2020), 70–95% of biodiesel yield. Many authors have widely explored the
without contributions to environmental pollution (Tan et al., 2010; Li immobilization of Pseudomonas lipase with different immobilization
et al., 2011b; Atabani et al., 2012; Verma et al., 2013) and achieving methods (Noureddini et al., 2005; Mak et al., 2009; Li and Yan, 2010; Li
high efficiency (Andrade et al., 2016). Meanwhile, Taher et al. et al., 2011a, 2011b; Lima et al., 2015; Chen et al., 2016). For instance,
compared the economic and environmental effects of biodiesel produced Ravi et al. immobilized the Pseudomonas lipase using a hybrid
using freeze-dried algae Nannochloropsis gaditana as the lipid source bio-support material (BSM) method in three neckle reactors at 50 ◦ C, 6:1
extracted via the supercritical CO2 method and using immobilized lipase methanol:oil ratio, 10 wt% catalyst loading for 24 h. The enzyme
from Pseudomonas cepacia and compared to compressed natural gas immobilized by the BSM method could be reused for up to 6 cycles to
(CNG) method to produce 1000 ton/year of biodiesel in 16 years (Taher reduce high lipase price with 70% biodiesel yield and 78% conversion.
et al., 2020). The result indicated that the biodiesel process using en­ Furthermore, the immobilized enzyme had good stability, keeping bio­
zymes via the supercritical CO2 method was profitable and produced diesel yield from run 1 up to 12 with less enzyme deactivation at con­
fewer emissions than those using the CNG. stant operating parameters. Nevertheless, a decrease in the biodiesel
However, the free enzymes can be easily denatured, has a short yield occurred after the 12 cycles (Nayak et al., 2019). Reusability of the
lifetime, and are very expensive when applied in large-scale production immobilized enzyme up to 6 cycles was also obtained via the polyvinyl
(Andrade et al., 2016). Thus, nanosized enzyme particles have an alcohol (PVA)-boric acid method (Busto et al., 2007).
essential function toward enzyme activity and stability. Vijayalakshmi Muanruksa and Kaewkannetra optimized the Rhizopus oryzae lipase
et al. developed that nanobiocatalysts were able to enhance the effi­ (ROL) immobilized with polyvinyl alcohol (PVA) in the enzymatic
ciency of biodiesel production. Besides, immobilized lipase on a nano­ esterification of sludge palm oil (SPO) at 40 ◦ C, 3:1 methanol:oil ratio,
sized particle called nano-immobilized lipase played a significant role in 200 rpm, and 2% enzyme loading within 4 h. The immobilized enzyme
biodiesel production due to thermal stability, high enzyme loading, produced the maximum yield of biodiesel (91.30%) and could even be
reusability, and effective and efficient security from enzyme denatur­ reused for up to 15 cycles. This experiment rectified the observation
ation (Vijayalakshmi et al., 2020). However, the nanobiocatalyst had studied by Busto et al. in the previous literature who immobilized the
poor mass transfer and uneven dispersibility, as reported by Zhong et al. enzyme with PVA, and the reusability of the enzyme was only reused for
. Therefore, it necessitates some researchers to make the enzymes worth 6 cycles (Busto et al., 2007). On the other hand, Nematian et al.
applying in biodiesel production, such as the improvement in immobi­ immobilized ROL on Fe3O4 superparamagnetic nanosize particle,
lization methods (Ferrero et al., 2020; Kumar and Thakur, 2020; Ravi whereby covalent bonding on the enzyme was observed toward the
et al., 2020), the selection of a suitable carrier (Rahman Talukder et al., enzymatic performance and stability of the enzyme. The immobilized
2006), the suitability of feedstocks (Agarwal et al., 2006; Shahid and lipase generated the optimum conversion of 69.8 wt% and was reused
Jamal, 2008; Singh and Singh, 2010; Sharma et al., 2012; Atabani et al., for up to 5 batches. The enzyme immobilized with PVA gave better
2013; Can, 2014; Hirkude and Padalkar, 2014; Singh et al., 2015) and enzymatic performance, activity, and stability than those with Fe3O4.
the appropriate operating parameters (Aghababaie et al., 2020). Recently, Kumar and Thakur enhanced the reusability and stability
of Serratia sp. ISTD04 lipids by Pseudomonas sp. ISTPL3 lipase enzymes
4.2. Nanobiocatalyst synthesis via immobilization of biocomposite substances in which the lipase was
immobilized onto calcite-assisted activated biochar (Kumar and Thakur,
4.2.1. Immobilization methods 2020). The optimum biodiesel yield (85.63%) was obtained from the
Enzyme immobilization methods to produce nanobiocatalyst for lipase immobilized with NaOH, 91.50% was achieved from the
biodiesel production are used to overcome the problems associated with glass-ceramic-immobilized lipase, 94.91% was achieved from the acti­
non-renewability of enzymes and to reduce the enzyme cost (Katch­ vated biochar-immobilized lipase, and 97.41% was achieved under the
alski-Katzir, 1993; Straathof et al., 2002; Mateo et al., 2007; Dizge and calcite-impregnated activated biochar-immobilized lipase. Moreover,
Keskinler, 2008; Hanefeld et al., 2009; Bose et al., 2010; Rebelo et al., the activated biochar-immobilized lipase impregnated with calcite could
2010; Barbosa et al., 2011, 2015; Garcia-Galan et al., 2011; Wang et al., be reused up to 10 cycles in the transesterification reaction. Biodiesel
2011; Yücel et al., 2011; Rodrigues et al., 2013; Firdaus et al., 2016; yield was strongly affected by different immobilization methods causing
Hama et al., 2018). In addition, nanosized ezyme particle gives advan­ improvement in the activity, stability, specificity, and reusability of the
tages in activity, thermal stability, more substantial mass transfer and enzyme (Babaki et al., 2016; Marín-Suárez et al., 2019).
reaction rates, and simple enzyme recovery (Verma et al., 2013). Immobilized Thermomyces lanuginosus lipase (TLL) significantly
In literature, various immobilization methods with application in contributed positively impact toward the yield and stability of nano­
biodiesel production are reported (Badgujar et al., 2019; Guo et al., biocatalyst. Ching-Velasquez et al. employed immobilized TLL with
2020). The latest inventions through immobilization methods reported octadecyl methacrylate beads to convert fish waste oil (FWO) into bio­
in 2020 provide an excellent reference to study due to the attractive and diesel (Ching-Velasquez et al., 2020). The reaction was carried out in a
novel findings (Ashjari et al., 2020; Bartha-Vári et al., 2020; Fatima batch reactor at 35 ◦ C, 10 wt% enzyme loading, and 216 rpm for 24 h.

17
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

The maximum yield was 75.3%, and the biodiesel properties were friendly, and inert for the enzymes (Cao, 2006), preferably possessing
aligned with the ASTM D-6751 standard. On the other hand, Ashjari hydrophobic characteristics (Schmid and Verger, 1998).
et al. immobilized TLL and RML with silica core-shell magnetic nano­ The carriers have two functions: interfacial activation while the first
particles (Fe3O4–SiO2) (Ashjari et al., 2020). The immobilized TLL immobilization takes place and provide interactions created by other
showed a much better performance than the RML, with the optimum groups for the second immobilization (Barbosa et al., 2013; Bernal et al.,
yield of 93.1% for TLL and 57.5% for RML at the same operating pa­ 2014; Santos et al., 2015; Rueda et al., 2016). In addition, carriers have
rameters. Both enzymes could be reused for up to 5 cycles. advantages such as enlarged porosity and active surface sites. Moreover,
Despite the immobilized TLL possessing perfect activity and stability, porous carrier structures can be created over acid pretreatment, as
Lipozyme TL IM (Thermomyces lanuginosus) enzyme was crowned as the investigated by Xu et al. . In this sense, different carriers can be found,
best enzyme among 10 enzymes for biodiesel production (Santaraite such as THF, choline citrate, activated carbon (AC), tosylated cloisite,
et al., 2020). The best results achieved were a 99.92% conversion and divinyl sulfone, n-heptane, iso-octane, etc. (Rios et al., 2019; Thangaraj
99.89% degree of transesterification under 30 ◦ C, and enzyme loading of et al., 2020), as was further discussed in the book of Cao (2006).
5 wt% within 7 h of reaction. Bajaj et al. employed the Lipozyme TL IM Talukder et al. reported THF as the most effective carrier to immo­
in the transesterification of Jatropha seeds at 45 ◦ C, 15% enzyme bilize CAL as the biocatalyst on acrylic resin, because THF was able to
loading, and 1:6 methanol:oil ratio for 12 h of reaction and positively prevent either inhibition of reaction rate caused by water content or
contributed to 90.6% biodiesel yield. poisoning palm oil transesterification caused by the addition of high
The role of the Lipozyme TL IM enzyme in the transesterification of amounts of methanol without using a polar organic solvent. Meanwhile,
Nannochloropsis strains (i.e., N. oculata, Nannochloropsis sp., and Gorke et al. employed ionic liquids containing choline citrate as a
N. oceanica) and ethanol in a batch reactor at 25–40 ◦ C with 0.5–2% biodegradable, cost-efficient, and less poisonous carrier in consequence
enzyme loading was explored by He et al. (2019). Their results disclosed of good water-solubility without damage to the enzymes. The choline
that the biodiesel production using the Lipozyme TL IM gave the highest citrate-assisted enzyme could enhance the solubility and even selectivity
yield than that without the enzyme. The yields produced by Nanno­ of biodiesel.
chloropsis sp. and N. oceanica were 85.12% and 76.33%, respectively. In Activated carbon (AC) was used as a carrier in the immobilization of
contrast, the yield achieved by N. oculata cells synthesized by four a recombinant Fusarium heterosporum lipase (FHL) in the reaction of
combined enzymes was 92.59% with the reusability of 2 cycles, while crude plant oils, as reported by Quayson et al. . The activated carbon-
the Lipozyme TL IM gave the 90.24% yield. The biodiesel production immobilized enzyme was trusted to overcome the problem of biodiesel
using the enzyme was the most effective, efficient, cost-effective, and melting point increase caused by the existence of phosphorus and other
green process. The higher conversion rate was produced by the Lip­ impurities in the reaction. The immobilized enzyme successfully pro­
ozyme TL IM in the transesterification of Nannochloropsis strains, fish oil, duced 98.8 wt% biodiesel yield. The carrier played a significant role in
and Schizochytrium oil compared with other enzymes has also been re­ the enzymatic activity and stability of enzymes that could be reused for
ported by others (Amoah et al., 2016; Hama et al., 2018; He et al., 2019). up to 9 cycles. The activated carbon carrier was also investigated and
Based on those above examples, the variety of enzyme immobiliza­ showed phosphorus content was significantly reduced from 565 ppm to
tion methods play a crucial role in enzymatic capability, activity, and 13 ppm (98% reduction) and reduced FFA content from 5.94% to 0.15%
stability that influence the biodiesel yield and the reusability of the (97% removal) (Sulihatimarsyila et al., 2019). These contaminants
enzymes (Rial et al., 2020; Zhang et al., 2020a). could inactivate the active sites of enzymes that hinder reaction.
Nezhad and Aghaei employed a novel heterofunctional carrier
4.2.2. Selection of carrier for immobilization of nanobiocatalyst in biodiesel namely tosylated cloisite in covalent immobilization of CRL called TCLL
production for biodiesel production from WFO (Nezhad and Aghaei, 2020). The
Development of enzyme immobilization procedures was carried out, hydrolytic TCLL activity used was around 1.96 U/mg and the immobi­
aiming to reduce the production costs and provide reusable enzymatic lization yield obtained was 93.6%. TCLL revealed a stunning result and
catalysts (Sheldon and Pelt, 2013; Rios et al., 2019; Andreo-Martínez stability whereabouts the yield obtained by free lipase was 33.4% then
et al., 2020) and thus enhancing process efficiency (Barbosa et al., 2015; significantly upgraded to 70.6% when using the TCLL during 30 days.
Pinheiro et al., 2019). Numerous nanomaterials have been developed for Even, the yield was, 61.3%, maintained after 10 batch reaction runs. On
enzyme immobilization, such as nanometals, gold nanoparticles, silver the other hand, the novel heterofunctional carrier used by Pinheiro et al.
nanoparticles (Dumri and Hung Anh, 2014), nanodiamond, nanofiber, (2019) was divinyl sulfone (DVS)-activated chitosan for the covalent
nanographene, and carbon nanotubes (Fatima, 2021). immobilization of CALB with the hydrolytic DVS activity of 14.52 U/g.
Some procedures have been deeply explored for enzyme immobili­ In this study, the DVS-activated chitosan was a potential carrier to use in
zation, including physical adsorption of lipase on the macroporous resin the immobilization process in order to enhance the enzymes and lipases
(Yang et al., 2006), lipase entrapment with tetramethoxysilane and stability.
iso-butyltrimethoxysilane (Noureddini et al., 2005), the use of bovine Inexpensive, abundant, and potential carriers have been explored
serum albumin (BSA) as a proteic feeder on crosslinked enzyme aggre­ recently using mycelium coming from the industrial waste as a new
gates (CLEAs) (Shah et al., 2006), crystallization of crosslinked enzyme carrier (Xu et al., 2020). The mycelium adsorption activity was
crystals (CLECs) (Lee et al., 2000; Verma et al., 2020), protein-coated enhanced over the acid pretreatment caused by the enlarged porosity. In
microcrystals (PCMCs) of the enzyme (Kumari et al., 2007; Raita the work, the highly-stable enzyme immobilization process was ach­
et al., 2010), and using an organic solvent with tetrahydrofuran (THF) as ieved because of the modified mycelium carrier. At the same time,
a carrier (Rahman Talukder et al., 2006). It was also observed that the agroindustrial wastes namely lignocellulosic and non-lignocellulosic as
using organic solvents can pressurize the ionization and lower the water the new carriers. The lignocellulosic carrier (i.e. coconut) outperform
content (Illanes and Fajardo, 2001; Subileau et al., 2017) and can the non-lignocellulosic carriers (Girelli et al., 2020).
effectively minimize enzyme leaching (Cantone et al., 2013; Ding et al., Despite these advantages, development of nanomaterials require
2019). multiple steps and involve laborious process. The preparation of
Carrier properties are essential role in the effective immobilization supermagnetic graphene oxide nanocatalyst required four main steps
process and affect enzyme capability. These materials must be insoluble starting with the preparation of magnetic nanoparticle using co-
in the solvent and be thermally, chemically, and mechanically stable precipitation method, preparation of graphene oxide using modified
during the process (Hartmann and Kostrov, 2013); instead they can Hummers’ method, preparation of magnetic-graphene oxide via in situ
destroy the framework of the enzyme leading to the weakening of its deposition method and functionalization prior immobilization proced­
activity. In addition, carriers should be cost-efficiently, environmentally ure (Nematian et al., 2020b). These steps are time consuming, requires

18
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Fig. 5. Effect of operating conditions on biodiesel yield: (a) methanol-to-oil molar ratio, (b) catalyst contents, (c) temperatures, and (d) agitation speeds toward the
biodiesel yield produced.

lots of chemicals and complicated instrumental setup that are less that protects the active lipase from bacterial degradation and improve
attractive in an economical point of view. Physical immobilization the binding between support and enzymes (Dumri and Hung Anh,
method also prone to enzymes leaching reducing its efficiency in 2014).
repeated reaction cycles. Functionalization of nanomaterials is a crucial Another concern of enzyme immobilization on nanometals materials
procedure prior immobilization in order to introduce reactive functional such as metal organic frameworks is prone to metal leaching and
groups on the support assuring strong bonding between support and enzyme poisoning. One of the strategy to minimize enzymes and metal
enzymes. Naked nanomaterials which are lacked of fuctional groups leaching is by synthesizing polymeric or inorganic frameworks in
make the immobilized enzymes susceptible to leaching. The hydro­ concomitant with enzymes that can lead to crosslinked networks that
phobicity and surface charge of graphene oxides can be altered by have good mechanical strength and stability. This strategy involves mild
activating the support’s surface using different types of amino acids reaction condition in order to prevent enzyme deactivation (Cirujano
(phenylalanine, glutamic acid, tryptophan, cysteine, lysine, and argi­ and Dhakshinamoorthy, 2021). Metal complexes involving metal ions
nine) (Zhou et al., 2019). Glutamic acid-graphene oxide-lipase shows and ligand have been explored having the inhibitory effect towards
the highest relative enzymatic activity up to 200% correspond to 172 enzymes by binding at the enzymes and interacting with amino acids
mg/g of protein among other immobilized samples due to its charged (Kilpin and Dyson, 2013). Hence, biocompatibility between the mate­
and polar characteristics that matched to the surface properties of rials and enzymes must be considered in order to maintain the effec­
Thermomyces lanuginosus lipase. tiveness and functionality of immobilized enzymes to reduce the
Addition of cross-linker such as glycerol diglycidyl ether (GDE) and i, inhibitory effect of metal complexes towards biocatalysts. Besides,
ii, bis-N-succinimidyl-(pentaethylene glycol) ester (BS(PEG)5) on carbon during long reaction time, declining of enzyme’s activity was observed
nanotubes minimized leaching issue and overcome low diffusion resis­ due to its long exposure and deactivation by methanol (Badoei-Dalfard
tance against reactants favoring the mass transfer processes (Bencze et al., 2021). Another drawbacks of nanobiocatalyst is tedious separa­
et al., 2016). The shorter crosslinker, GDE shows high biodiesel con­ tion process especially when using conventional separation method such
version up to 79.9% compared to the longer BS(PEG)5 crosslinker as centrifugation, extraction, vacuum distillation and filtration which
(50.8%). BS(PEG)5 increase flexibility of immobilized enzymes, favors can lead to catalyst loss and high operational cost. Hence, tailoring the
the multipoint fixation of the enzyme onto the nanotube surface, materials properties by adding the magnetic properties ease nano­
resulting a strongly anchored, but impaired biocatalyst. Kanchana biocatalysts separation using an external magnetic field. A tailored
Dumri and Dau Hung Anh reported that the addition of polydopamine graphene oxide was fabricated by adding magnetite using in situ
on a silver nanoparticles as an adhesive creates an antibacterial surfaces deposition method (Nematian et al., 2020b). The catalyst successfully

19
H.I. Mahdi et al.
20

Fig. 6. Advantages and drawbacks for most studied operational parameters for biodiesel production.

Chemosphere 319 (2023) 138003


H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Fig. 7. The minimum and maximum indicators for developments in biodiesel production.

improved its catalytic efficiency, thermal stability, storage stability and toward the biodiesel yield from frying palm oil (FPO) using CALA. The
can be recycled up to 5 cycle with 58.77% of biodiesel production. highest biodiesel yield (94.6%) was achieved at 30 ◦ C, 16.6 wt% water
It is seen that the carriers for enzymes immobilization play a very content, 5.5 wt% enzyme loading, 7:1 methanol:oil ratio in 22 h. Be­
crucial part to upgrade the capability and stability of the nano­ sides, the influence of operating conditions was likewise observed by
biocatalysts, which brings the increase in biodiesel yield. Otherwise, Jayaraman et al. (2020) in the enzymatic transesterification of WCO
cheap and renewable carriers are preferred, since it can affect the employing 1%, 1.5%, 2%, 2.5%, and 3% enzyme loadings of, 2:1, 3:1,
readiness of commercialization of such processes. After the choice of a 4:1, 5:1, and 6:1 methanol:oil ratios and 2, 3, 4, and 5 h reaction times.
proper immobilization method, reaction parameters should be taken The work represented that its best reaction conditions at 1.5% enzyme
into account in order to gather information to scale and sizing biodiesel loading, 3:1 methanol:oil ratio of and 4 h reaction time. More elevated
production using nanobiocatalysts, as shown in the next section. transformation rate was reached under higher enzyme content that
could escalate the biodiesel yield. The produced biodiesel was then
5. Appropriate operational parameters for enzymatic reactions tested in the engine by combining biodiesel with pure diesel fuel with
in biodiesel production the portions of biodiesel-diesel fuel of 20%–80%, 40%–60%, and 60%–
40%, respectively. However, the fuel mixture with less biodiesel content
In previous sections, it was showed that several works utilized provided good brake fuel consumption and brake thermal efficiency
different operational conditions in order to improve catalysts perfor­ with lower CO, hydrocarbons, and NOX emissions.
mance. In Fig. 6, these parameters aresummarize. It is shown that Thus, operating parameters including temperature, reaction time,
appropriate operational parameters should be studied and the results of catalyst loading, alcohol-to-oil molar ratio, various alcohols, different
biodiesel production are evaluated based on several indicators that solvents and water content can cause, when combined or not, both
should be optimized to ensure that the procedures can be successfully positive and negative impact enzymatic performance and capability for
applied in the real industries, achieving commercial scale production biodiesel production. Hence, the use of these parameters with compat­
(CRL 9, see Fig. 2). In this sense, some indicators that should be maxi­ ible concentration can maximally optimize the desired results. In the
mized or minimized are depicted in Fig. 7. next sections these parameters are briefly discussed in separate.
It can be notifced from Fig. 5 (Pasupulety et al., 2013) that the
objective function for the optimization of a biodiesel production process 5.1. Temperature, reaction time and catalyst loading
using nanobiocatalyst depends on maximize bench scale information,
such as biodiesel yield (show in Fig. 3), and minimize the enzyme costs It was reported that the higher reaction temperature can induced
as well as other challenges like contaminantion with impurities. Thus, higher biodiesel yield due to the good collision frequency between the
this optimization starts in the synthesis of the nanobiocatalyst then goes enzyme and reactants. Additionally, temperature may establish the
through bench scale experiments, where different parameters can be maximal lipase-reactants complex as well as lower the viscosity, mass
evaluated. transfer limitations (Guo et al., 2020) and biodiesel yield due to evap­
In this sense, nanobiocatalysts capability, activity and stability are orating alcohol during the process (Gomes Filho et al., 2015). Moreover,
outstandingly affected by many factors in nano and micro scales, which lower viscosity system turns faster the stirring speed and aids in efficient
can be controlled during synthesis. In other hand, other factors are also mixing solution, whus increasing the biodiesel yield obtained (Peiter
important for biodiesel production in meso (bench) and macro (indus­ et al., 2020). These reports have concluded that as longer reaction time,
trial) scale, such as the operating parameters which lead to the optimum highest the biodiesel yield. Furthermore, the more enzyme loading
conversion, selectivity, biodiesel yield as well as the longest reusability strongly augmented more actives sites leading to more superior reaction
of a nanobiocatalyst (Aghababaie et al., 2020; Guo et al., 2020). In rate, which produce higher biodiesel yield (Rathod and Pandit, 2010;
literature, Guo et al. evaluated the effect of operating conditions such as Tavares et al., 2017).
water contents (5–25 wt%), enzyme loadings (1 – 9 wt%), methanol:oil
ratios (3:1–12:1), reaction times (5–50 h) and temperatures (25–60 ◦ C)

21
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Fig. 8. Routes for biodiesel production using different generation of feedstocks.

5.2. Alcohol-to-oil molar ratio et al., 2020a). To prove, it was shown that adding excess methanol:oil
ratio contributed with a bad impact toward the biodiesel yield because
Based on the stoichiometric calculation, to produce 3 mol of bio­ the excess of methanol could destroy the structure of protein on the
diesel and 1 mol of glycerol, it needs 1 mol of triglyceride and 3 mol of enzyme resulting in poor activity and stability of the enzyme (Liu et al.,
methanol in the reaction. Theoretically, transesterification of vegetable 2013; Pollardo et al., 2018).
oils and methanol cannot occur at a less amount of methanol. In other
hand, an excessive amount of methanol:oil ratio is also not a good
5.3. Types of alcohol
practice considering the high methanol recovery process cost due to the
difficulties in the separation (Pasupulety et al., 2013; Yesilyurt et al.,
Various types of alcohols used in the biodiesel production are
2019).
methanol, ethanol, propanol and butanol, where methanol is widely
Deactivation of nanobiocatalysts was observed for higher alcohol-to-
applied due to the low-cost (Lam et al., 2010; Borges and Díaz, 2012).
oil molar ratio due to hindrance of the biocatalyst active surface area
Shorter hydrocarbons chains namely methanol and ethanol are
(Deive and Rodríguez, 2020; Hossain et al., 2020) and for insoluble al­
commonly harnessed in the biodiesel process due to more environ­
cohols (Al-Zuhair et al., 2007; Guan et al., 2010), whose droplets in­
mentally friendly and lower alcohols price (Li et al., 2009; Kumar and
fluence on mixing reduced the biodiesel yield obtained (Ngoie et al.,
Saravanan, 2016) as well as resulting in higher biodiesel yield while
2020). Consequently, the appropriate usage of alcohol-to-oil molar ratio
longer chains of the hydrocarbons such as propanol and butanol produce
is implemented to minimize the biocatalysts deactivation (Shimada
a lower yield despite the faster reaction rate (Coggon et al., 2007).
et al., 2002) and to maximize the biodiesel yield obtained (Chatto­
Methanol has such a lower price than ethanol, but it has an elevated
padhyay and Sen, 2020). In general, it was found that 1:3 methanol:oil
detonation risk due to poor boiling point and is more hazardous and
ratio is crowned as the best molar ratio (Law et al., 2018; Jayaraman
poisonous for human than ethanol (Leung et al., 2010). Furthermore,
et al., 2020). However, olther claim that the best methanol:oil ratio was
methanol is highly poisonous alcohol for the enzymes active sites and
obtained at 7:1 as reported by others (Halim and Kamaruddin, 2008; Yu
can fastly deactivate the enzyme performance (Ranganathan et al.,
et al., 2013; Mehrasbi et al., 2017).
2008) whereas ethanol is more cost-environmentally alcohol obtained
To prove this result, different alcohol-to-oil molar ratios has been
from renewable resources (Fjerbaek et al., 2009) and not as poisonous as
recently carried out (Muanruksa et al., 2020; Muanruksa and Kaew­
methanol (Shah and Gupta, 2007; Nikhom and Tongurai, 2014). Yet, in
kannetra, 2020; Ngoie et al., 2020; Patchimpet et al., 2020; Rocha et al.,
the last stage, a difficult separation process becomes a disadvantage of
2020) to observe the nanobiocatalysts ability and stability and biodiesel
using ethanol due to the formation of emulsion (Kim et al., 2010;
yield obtained. The higher alcohol-to-oil molar ratio contributes to shift
Stamenković et al., 2011).
the reaction equilibrium into biodiesel formation causing increasing the
Even though, methanol is irritant for the skin and enzymes, biodiesel
biodiesel yield (Shimada et al., 2002), yet the excess alcohols content
production using methanol as the reactant has more stable reaction
precisely reduces the yield, which is caused by denaturation of the en­
condition, better activity, higher stability, and easy separation process in
zymes and lipases proteins structures, leading to attenuate the bio­
the last process (Thoai et al., 2019). Excess methanol negatively con­
catalysts capability (Hama et al., 2018; Aghababaie et al., 2020; Zhang
tributes to the enzyme performance thus the appropriate methanol

22
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

content is responsible for the better enzyme activity and the yields. Application of the whole generations feedstocks for biodiesel production
From this overview, every alcohol has their advantages and draw­ was deeply reviewed by Mahlia et al. .
backs toward the enzyme activity and the biodiesel yield based on the The first-generation oils are edible vegetable oils derived from
types of enzymes and lipases applied in the reaction. In this part, authors rapeseed, coconut, soybean, corn, palm, olive oils and many more
highlight the effect of different types of alcohols and provide a recom­ (Mahdavi et al., 2015). Despite the high yield of biodiesel produced, as
mendation to use methanol as the reactant due to less budget, more shown in the previous sections, the process has some challenges for
stable process, better ability, longer stability, simple separation process, efficient implementation, such as competition with food production cost
and the higher biodiesel yield than the other alcohols in order to achieve and restricted cultivation area.
a sustainable process in the future. Due to these reasons, the second-generation biodiesel is produced
from non-edible vegetable oils derived from neem, jatropha, nag­
5.4. Solvent effect champa, karanja, calophyllum inophyllum, rubber seed, mahua indica
oils and others (Peer et al., 2017; Shameer and Ramesh, 2017). The
Solubility of alcohols into oil can be enhanced using a solvent rein­ non-edible oils-derived biodiesel is considerably produced these days by
forcing the nanobiocatalysts stability (Modi et al., 2007; Hama et al., dint of more environmentally friendly, cost-efficient, annihilating food
2011; Lee et al., 2013) and facilitate the reaction rate (Fjerbaek et al., dissimilitude and larger cultivation area (Tariq et al., 2012; Aransiola
2009; Li et al., 2009; Raita et al., 2010) as well as to degrade viscosity et al., 2014; Mahdavi et al., 2015; Chua et al., 2020). Nonetheless, the
and mass transfer limitation (Kumari et al., 2007; Andrade et al., 2020). yields of several plants such as jatropha, jojoba, and karanja oils can
Furthermore, the solvents can dissolve the glycerol (Akoh et al., 2007; drastically be reduced leading to the negative affect towards the econ­
Chen et al., 2009; Chattopadhyay and Sen, 2020) that can block the omy of the community and also the food production (Singh et al., 2020).
active biocatalysts sites (Kumari et al., 2009) and can protect the The oils from these plants also require more alcohols during the process
deactivation of the enzyme caused by more alcohols content (Hossain (Tariq et al., 2012).
et al., 2020). Hence, the new resources oils are introduced namely from micro­
The solvent consists of two types which are hydrophobic organic algae and waste cooking oils such as animal, biomass, chicken, chlorella,
solvents such as hexane, iso-octane, n-heptane, petroleum ether, tert- fish, and others called the third-generation of biodiesel (Verma et al.,
butanol etc (Royon et al., 2007; López et al., 2016; Chattopadhyay and 2016b; Arif et al., 2020). The third generation biodiesel is a highly su­
Sen, 2020) and hydrophylic organic solvents like methanol, ethanol, perior process in consequence of less pollution, high growth and pro­
propanol and butanol (Guldhe et al., 2015; Tran et al., 2016; Selvakumar ductivity rates, larger cultivation area, higher oil content and no impact
and Sivashanmugam, 2017; Ashjari et al., 2020). The hydrophobic on the food production (Singh et al., 2020). Even though it can resolve
organic solvents contribute to the increase in enzyme capability and the the problems of the first and second generation biodiesel, the third
biodiesel yield in which the iso-octane was proven by some studies as the generation requires high investment, enough sunlight (Agarwal et al.,
most effective and efficient solvent (Bartha-Vári et al., 2020). 2006), wide-scale production, and overcome the difficulty in oil
The poisonous characteristic of methanol causing the deactivation of extraction (Hannon et al., 2010; Tariq et al., 2012; Paul Abishek et al.,
the enzyme can be overcome by using the organic solvents like hexane, 2014). Alternatively, the fourth generation biodiesel pays great atten­
propan-2-ol, and ethyl acetate etc as acyl acceptor (Venkat Reddy et al., tion wherein it is produced from advance solar oils (i.e. solar energy)
2006; Pollardo et al., 2018; Nematian et al., 2020a; Ravi et al., 2020). and electro fuels which are in a new area of research because of abun­
dant availability, inexhaustible and inexpensive price (Singh et al.,
5.5. Water content 2020). All of these generations are reacted with methanol/ethanol by
using the same process, which is transesterification to produce
Water content is the main factor to obtain good enzyme stability high-grade biodiesel product (Singh et al., 2019, 2020).
because the water content can enhance nanobiocatalysts performance The influence of different kind of vegetable oils toward the quality
and stability using organic solvents (Lu et al., 2009; Li and Yan, 2010; and quantity of produced biodiesel as well as reducing pollution has
Babaki et al., 2015) and also can shift the reaction equilibrium into been widely explored by researchers (Awais et al., 2020; Simsek, 2020).
biodiesel formation (Adnan et al., 2018). The enlarged water-oil in­ Foteinis et al. compared the biodiesel produced from the first and
terfaces active site is trusted to fasten the reaction rate (Maruyama et al., third-generation oils. Third generation biodiesel was much cost-efficient
2000; Ghaly et al., 2010). Higher water content gives a significant role in because it used inexpensive waste cooking oils coming from house and
escalating the biocatalysts activity and the yield (Chen et al., 2008; Yu restaurant waste oils. Moreover, the industrial third generation biodiesel
et al., 2013; Ferrero et al., 2016). However, the excess water content can produced total carbon emission of around 0.55 ton CO2 equivalent,
cause the enzymes active sites to destroy leading to descending enzyme which was 40% lower than that from the first-generation biodiesel. The
capability (Guo et al., 2020). first generation biodiesel decreased hydrocarbon and CO emissions,
whereas NOX (80.50%) and CO2 (42.62%) emissions were enhanced
6. Different feedstocks: quality and quantity of biodiesel (Simsek, 2020). It means that the third generation biodiesel is much
product better than the first due to cost-efficient, reduced CO2 emission, and
guaranteed human health.
Based on the ASTM D-6751 standard, biodiesel can be produced by The second-generation biodiesel produced from Roselle (RB-10) and
different routes related to types of vegetable oils categorized into four Karanja (KB-10) toward biodiesel quality and reduction of environ­
generations, namely: edible oils as the first-generation, non-edible oils mental impact (Shrivastava et al., 2020). The result obtained indicated
as the second-generation, WCO as the third-generation and the fourth is that brake thermal efficiency produced from RB-10 was 1.5%, smaller
the advance solar oils (Can, 2014; Hirkude and Padalkar, 2014; Singh than that from KB-10 (2.8%) and the brake fuel consumptions were
et al., 2015), as conceived in Fig. 8. 5.5% for RB-10 and 2.0% for KB-10 which was higher than diesel fuel.
One of the developments carried out in biodiesel production is by Besides, the ignition delay period of RB-10 was decreased. Additionally,
choosing the high-grade vegetable oils (Kumar et al., 2010). At this either RB-10 or KB-10 were able to decrease NOX and smoke emissions
point, biodiesel produce from the first, second, and third generations with a little bit higher CO2 emission. Furthermore, the second genera­
utilize the abundant natural resources, which provides raw (unused) and tion biodiesel produced from Melia azedarach and Ricinus communis to­
waste (used) oils. In contrast, the fourth generation biodiesel is based on ward biodiesel yield (Awais et al., 2020). Biodiesel yield obtained from
human-engineering biological equipment, whose state-of-art is contin­ Ricinus communis oil (94%) was higher than Melia azedarach oil (90%).
uously been developed and increased in literature (Singh et al., 2019). This is due to the higher oil content in Ricinus communis (43.6%)

23
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

compared to Melia azedarach (35.6%). best technology for producing biodiesel in Indonesia, Malaysia, Brazil
The effect of variety of FFAs contents (0%, 40%, and 80%) was and USA cases is considered because these countries are recognized
observed by Chang et al. (Chang et al.) using Aspergillus oryzae lipase among the biggest biodiesel producers in the world:
(Eversa® Transform 2.0) and refined palm oil containing various FFAs
contents where these different FFAs contents affected the use of • Indonesia. Biodiesel production is also heavily dependent on policies
methanol-to-oil molar ratio and the best FAME conversion obtained. The in palm oil producing countries, such as Indonesia and its’ neigh­
highest FAME conversion (80%) was obtained from 0% of FFAs content bourhood, Malaysia. Since 2006, Indonesian government has made
using 4:1 methanol-to-oil molar ratio. Approximately, 90% of the opti­ regulations to use mixture of biodiesel and fossil diesel fuel in the
mum FAME conversion was obtained over 40% of FFAs content at 5:1 transportation sector. Moreover, the Indonesian government has
methanol-to-oil molar ratio while the maximum FAME conversion targeted 5% of the total Indonesia energy consumption coming from
(97%) was achieved over 80% of FFAs content using 4:1 methanol-to-oil biodiesel in 2025 (Putrasari et al., 2016) by employing 100% of
molar ratio. The best FAME conversion obtained from the highest FFAs biodiesel in the transportation sector (B100) (Sugiarto et al., 2018).
content was correlated with the faster reaction rate that could be esti­ In 2015, a financial mechanism was created to support domestic
mated using Ping-Pong Bi–Bi method (Chang et al.). On the other hand, biodiesel consumption focusing on offsetting the price gap between
the effect of vegetable oils containing different molecular structures (i.e. biodiesel and fossil diesel. The monetary funds were managed by the
C12 to C20) toward the biodiesel yield produced was reviewed by Arif Oil Palm Plantation Fund Management Agency Government that
et al. (2020). utilize the funds for research and development, replanting and palm
An opposite result exhibiting lower FFAs contents resulting in higher promotion activities that are located in Indonesia (Rahmanulloh,
biodiesel yields was reported by Ferrero et al. who observed application 2019). Consequently, the capacity of Indonesia biodiesel production
of various vegetable oils possessing different FFAs contents namely SFO was expanded from 4.9 billion liters in 2012 to 11.5 billion liters in
(0.05 wt%), used frying oil (0.11 wt%), residual SBO (76.91 wt%), and 2017 in which the refineries plants were increased from 22 plants to
J. hieronymi oil (4.07 wt%) (Ferrero et al., 2020). The highest yield 31 plants (Silalahi et al., 2020). Increasing the biodiesel production
(95.6 wt%) was accomplished over SFO followed by waste frying oil capacity is related to the Indonesian government policy by applying
(91.4 wt%), and J. hieronymi oil (89.7 wt%) and the lowest yield was in mixture 20% and 30% of biodiesel with 80% and 70% of fossil
obtained by residual SBO (76 wt%) (Ferrero et al., 2020). A decrease in diesel fuel, respectively. This policy has given a positive contribution
the biodiesel yield was caused by the higher FFAs contents because on the economic growth, social welfare and clean environment.
increasing FFAs content had excess water contents that could extermi­ • Malaysia. Malaysian government allocated US$79 million to set up
nate the enzymatic active sites framework (Salis et al., 2009; Ferrero blending facilities and infrastructure to accommodate the country’s
et al., 2016), as shown in Fig. 7, and even destroy the enzyme structure biodiesel mandates (Wahab, 2019) because the crude palm oil pro­
framework. Noteworthy to mention, the high FFAs content negatively duction in Malaysia significantly increased from 2.6 MT in 1960 to
affects the enzyme performance and the biodiesel yield obtained due to 19.7 MT by 2013 (Lam et al., 2009). Consequently, Malaysia
the production of more soap (Encinar et al., 2005; Atadashi et al., 2012; expanded the biodiesel industries by producing 10 MT of the palm
Atapour et al., 2014; Vieitez et al., 2014; Nayak et al., 2016) that could oil-derived biodiesel product (Johari et al., 2015). Malaysia gov­
block the enzymatic active sites. To overcome this problem, it is ernment enacted the fifth fuel policy by supplying 5% of Malaysia’s
encourage to apply in a two-step reaction wherein the esterification future energy from crude palm oil. One of the most prospective en­
reaction in the first step is purposed to bring down the FFA content to ergies in Malaysia is by using crude palm oil in the biodiesel pro­
less 2 wt% (Canakci and Van Gerpen, 2001), thus lesser soap formation duction (Abdullah et al., 2007; Yee et al., 2009) thus the palm
in the second step-transesterification process can be reached (Math and oil-based biodiesel production has dramatically increased annually
Irfan, 2007; Thoai et al., 2019) that will facilitate the separation process (Johari et al., 2015). On the other hand, utilization of palm oil in the
(Thoai et al., 2017). Thus, is was proposed that FFAs content must be biodiesel production provides substantial advantages if compared to
dramatically reduced to less than 1 wt% to lower the soap production other vegetable oils (Yee et al., 2009) such as simple technology, low
(Encinar et al., 2005; Amini et al., 2017). production cost, and abundant raw materials (Mahdi et al., 2021).
In consonance with the reports above, it means that different kind of Unfortunately, competition with food industries becomes a huge
vegetable oils strongly affects the quantity and quality of the biodiesel challenge (Azelee et al., 2022). Biofuel Industry Act (BIA) was built
produced. Hence, choosing the best vegetable oils based on demands for in 2007 to develop Malaysia biodiesel industries by adopting Euro­
each country is a crucial factor towards the optimum production cost, pean EN 14214 regulatory as standard in the testing methods for
high biodiesel yield, less environmental impact, protects human health Malaysian biodiesel fulfilling international specifications. Recently,
as well as provide a technology that can be suitable for an specific palm oil-based biodiesel technology developed in Malaysia is the
community. transesterification process using NaOH (sodium hydroxide) as the
catalyst resulting in yield of 98% as well as meeting EN 14214 and
7. Best technologies for Indonesia, Malaysia, Brazil and USA ASTM D 6751 standards for International biodiesel product (Johari
et al., 2015).
To accomplish the targeted percentage of biodiesel usage in many • Brazil. Reflecting social inclusion and regional development con­
countries, government support policies were established to promote the cerns, the “Social Fuel Stamp” programme was established in Brazil
healthy growth of domestic biodiesel sector and replace the imported to provide incentives to family farms in poor regions while regulating
diesel fuel. In this sense, expansion of biodiesel production worldwide the biodiesel market through a public auction system (Barros, 2019).
was driven by government mandates and incentives in place at various Brazil is known as the second largest biodiesel producer in the world,
countries. One of the government-established policies in 2018 autho­ the highest soybean production, and one of the biggest biodiesel
rized the use of 20% of biodiesel mixed with fossil-based diesel fuel exporters (Costa and Oliveira, 2022). Also, Brazil is the first country
called B20 (Silalahi et al., 2020). B20 mandatory policy donates much where the biodiesel incentive policy is established in order to fulfill
benefit for the economic growth in every country because around 30% demand of the National biodiesel production and its application
of global energy consumption is coming from the transportation sector called as PNPB (Stattman and Mol, 2014). In Brazil, biodiesel is
(Sukarno et al., 2016). Particularly in Indonesia, the total diesel fuel mostly produced from carbohydrates (sucrose and starch), vegetable
consumption in Indonesia by 2018 was around 32,000 Kilo Liters (KL) oils and animal fats (Karp et al., 2021) and substitutes around 12% of
thus the B20 mandatory policy would enhance the biodiesel demand and fossil-based diesel oil (Ramos et al., 2022). The common biodiesel
the welfare of the local citizen (Silalahi et al., 2020). In this section, the technology is the transesterification process called alcoholysis with

24
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 5
Compilation of best actual technology for biodiesel production using nanobiocatalysts in four big countries.
Feedstock Alcohol Catalyst Technology Results Problem solved Ref

Indonesia
Palm oil Methanol Commercial heterogeneous Biorefinery technology Biorefinery process was Biorefinery process has lower Harahap et al.
and catalysts the most cost-effective production cost than the (2019)
ethanol technology due to the conventional process.
high net income and
NPV as well as the low
BBP.
CPO NA NA Conventional technology Biodiesel produced Indonesia has large palm Simaremare
from CPO is economic plantation thus the low CPO oil et al. (2021)
feasible due to its low price has potential to provide
price. profit. Biodiesel produced
fulfill inside country energy
consumption and can also be
exported to other countries.
Palm oil NA NA Conventional technology Biodiesel yield Environmental issues Khatiwada
improvements could regarding biodiesel production et al. (2021)
reduce the using palm oil could be
environmental addressed over improvements
pollution and fulfill in yield.
biodiesel demand either
for Indonesia or for
export.
Palm oil NA NA Conventional technology LCA was used as a tool Environmental effect caused by Siregar et al.
equipped with LCA to analyze the biodiesel production using (2021)
environmental issue palm oil could be decreased
47% under LCA method if
compared to diesel fuel
production using palm oil.
Non edible Methanol Commercial catalysts Conventional technology Biodiesel produced Using waste feedstocks Zhou et al.
animal fats, and from waste feedstock is presents advantages such as (2021)
waste fish oil, ethanol cheap, effective, decreasing environmental
SPO, and tall innovative, and pollution, reducing fuel import
oil promising process. for Indonesia, and contributing
to narrowing Indonesia’s trade
deficit.
Microalgaes in Methanol Extraction of dried microalgae Conventional process by Chlorella sp PKB is the The abundant microalgaes in Jumiarni and
South cells biomass utilizing the local natural best microalgae. South Sumatera like Chlorella Anggraini
Sumatera feedstocks sp PKB, Chlorella sp PPP, (2021)
Chlorella sp SB, Crucigenia
quadrata PTA and
Scenedesmus sudetica PTA
became a potential feedstock in
biodiesel production in order to
utilize the plenteous natural
sources to be the sustainable
biofuel.
KSO Methanol Indonesia natural zeolite Transesterification Y = 84% Use of low cost feedstock could Sutrisno et al.
(clinoptilolite and modernite) technology solve the problem of high (2021)
was stirredly activated using biodiesel production cost while
sulfuric acid solutions 0.5 N at the use of local natural
100 ◦ C for 4 h catalysts could optimize the
potential of abundant natural
source
Reutealis Methanol Bukit Jaddih Madura limestone Transesterification Y = 9.45 for 1% of ZnO Indonesia natural catalyst Shalihah et al.
trisperma oil was modified with ZnO under technology in a batch content and 52.34 for performance was improved (2020)
wet impregnation synthesis reactor 7% of ZnO content over modification using ZnO.
(ZnO/CL) The higher ZnO content
mitigated crystallinity and
basicity but escalated the
biodiesel yield
WCO Methanol Graphite electrode and waste Co-solvent free Y = 78.51 with 100% of Higher voltage and longer Wicaksono
concrete heterogeneous catalyst electrochemical method FAME content reaction time offered a good et al. (2021)
result, but the highest voltage
and longest reaction
contributed to the poor
biodiesel yield
Malaysia
CPO Methanol Potassium hydroxide Batch reactor Y = 93.6% This method minimizes the Alkabbashi
processing cost and achieved et al. (2009)
the best yield of oil conversion
CPO, industrial Methanol Esterification using sulfuric acid Degumming and two step Y methyl hexadecanoate = CaO catalyst has good activities Maulidiyah
waste oil (H2SO4) and transesterfication esterification and 12.87% and produces an optimal yield et al. (2017)
process using CaO transesterification Y methyl 9-octadecanoate = of biodiesel
reactions 19.98%
(continued on next page)

25
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 5 (continued )
Feedstock Alcohol Catalyst Technology Results Problem solved Ref

Y methyl octadecanoate =
5.71%
Y methyl 8,11-
octadecadienoate =
10.22%.
Refined, Ethanol Commercial potassium MSCR Performance of the MSC Liquid− liquid extraction using (Mohd Noor
bleached and hydroxide reactor was promising, excess ethanol as the solvent et al., 2019)
deodorised as indicated by the high increased the rate of
(RBD) palm oil Da value 27 ≤ Da ≤ transesterification reaction
1688.9. with pure palm oil in a
X = ~100% multistage stirred column
under counter-current
Palm oil mill Methanol Microwave heating Microwave heating X methyl palmitate = This process is economically Waudby and
effluent occupied with recycling ~100% feasible with the inclusion of Zein (2021)
(POME) streams to recover significant recycle streams and
heptane, methanol, and additional separation
sulfuric acid equipment
RBD Methanol Commercial potassium Stirred tank reactor Y NaOH = 88% Reactor was equipped with Shahbazi et al.
hydroxide and sodium Y KOH = 93% reflux condenser to recylcle (2012)
hydroxide methanol escaping from the
reaction mixture
PFAD Methanol Eggshell-derived CaO catalyst Glycerolysis X = 98.7% Pretreatment of PFAD via Idris et al.
were prepared by washing and gycerolysis reaction reduces (2020)
drying the eggshells in an oven FFA content from 81 to 1.5% at
at 100 ◦ C. Then, the eggshell lower temperature (180 ◦ C)
was crushed into smaller pieces and catalyst loading of 1.5 wt%
and calcined at 600–1000 ◦ C
and 1 atm for 4 h
PFAD Methanol Sugarcane bagasse was dried Biobased nanocatalyst Y = 98.6% Sugarcane bagasse shown Akinfalabi et al.
and ground to powder form sustainable potential as (2020a)
prior soaking in phosphoric acid heterogeneous catalyst for
at (1:1) mass ratio for 24 h. The repeated esterification reaction
acidified sample was calcined in up to six cycles
a tube furnace under nitrogen
flow at 400 ◦ C for 2 h. Then, the
sample was washed with warm
water (80 ◦ C) and dried
(100 ◦ C) for 24 h
Oleic acid and Methanol CALB immobilized on Continuous and X oleic acid = 93.5% Esterification process using Mulalee (2020)
PFAD macroporous acrylic resin circulated batch X PFAD = 88.5% circulated batch and
(Novozyme 435) processes continuous processes in
expanded packed bed reactor
was performed to improve the
reusability of catalyst without
separation and prevented
mechanical shear stress
PFAD Methanol MnO–NiO–SO−4 2/ZrO2 was Esterification was Y = 97.70% Reusability of the catalyst was Al-Jaberi et al.
prepared using wetness performed using solid tested for at least five times (2017)
impregnation method and acid catalyst in without significant reduction
sulfonated with chlorosulfonic conventional normal in activity and the catalyst
acid before calcined at 600 ◦ C reflux reactor suitable for low grade
for 3 h feedstock
WCO of palm oil Methanol Molybdenum and zirconium Molybdenum and Y = 90.1% High catalytic activity in Mansir et al.
precursors were impregnated zirconium modified biodiesel conversion was (2018)
on the shells and calcined at calcium-based catalysts achieved due to the synergy of
650 ◦ C for 4 h in an opened-tube derived from waste shells weak and strong basic sites
furnace over the catalyst surfaces. The
incorporation of molybdenum
and zirconium on CaO surface
reduced Ca+2 leaching in FAME
product
WCO Methanol Magnetic biochar catalyst was Magnetic biochar catalyst Y = 90.2% Incorporation of iron layer on Quah et al.
synthesized with concentrated the magnetic biochar catalyst (2020)
sulfuric acid through the improves catalyst separation
sulfonation process of the performance
impregnated waste palm kernel
shell (PKS) biochar with ferrite
Fe3O4
WCO Methanol Potassium hydroxide Semi-industrial plant Y = 74.3% Feasibility of industrial-scale Farid et al.
(model design BIODIESEL biodiesel production from (2020)
BFD EOSYS 100 M) WCO was suggestively
lucrative by having projected
values for NPV of USD 1.47
million and IRR with 60%
Methanol Degumming and two step Y = 98.85% The antioxidant content of Damanik et al.
esterification and C. inophyllum boosts the (2017)
(continued on next page)

26
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 5 (continued )
Feedstock Alcohol Catalyst Technology Results Problem solved Ref

C. inophyllum Degumming using H3PO4, transesterification oxidation stability of the mixed


and palm esterification using H2SO4, and reaction biodiesel
mixed oil transesterification using KOH
Brucea javanica Methanol Sodium hydroxide Esterification with 1% (v/ Y = 94.34% Propyl gallate acts as a good Hasni et al.
seeds v) H2SO4 followed by antioxidant in improving the (2017)
transesterification using oxidative stability of Brucea
sodium hydroxide as base javanica biodiesel
catalyst
Pongamia Methanol Sodium hydroxide Transesterification Y = 84% Transesterification of KRO in Harreh et al.
pinnata (Crude process in an orbital an orbital shaker produce (2018)
KRO) shaker greater biodiesel yield (86%)
compared to magnetic stirrer
(72%)
Elite Jatropha Methanol Potassium hydroxide Transesterification using NA The new hybrid consumed Hanif et al.
curcas hybrid base catalyst 25.32 MJ of energy to produce (2019)
1 kg of biodiesel
Brazil
SBO Ethanol Immobilized lipase on resin Heterogeneous Y = 93% Brasil produces high amount of Beck et al.
acrylic (Novozym-435) transesterification X = 89% ethanol which avoid the need (2014)
T = 40 ◦ C to import methanol. This can
Catal./oil = 9.5% turn the production of biodiesel
Co-solvent = t-butanol totally national
SBO Methanol Burkholderia cepacia lipase Heterogeneous Y = 96% The best yield was achieved Fan et al.
immobilized by covalent transesterification X = 89% over less water content and (2017)
bonding 10 cycles lower reaction temperature
Waste SBO Methanol PEL immobilized on blue silica Heterogeneous Y > 90% Reduce the dependence on Li et al. (2009)
gel and 5A molecular sieve transesterification Co-solvent = t-amyl SBO. Brazil has greate potential
alcohol to produce PEL from wheat
bran
Bovine, pig and Methanol Bacillum megaterium lipase Heterogeneous Y > 89% Lipase obtained with bacillus (de Castro and
chichen fat or ethanol obtained with sugar cane transesterifications megaterium using wheat bran de Castro,
bagasse, immobilized showed high maximum 2012;
butkhoderia cepacian lipase activity. Since bovine fat is Toldrá-Reig
largely produced in Brazil, such et al., 2020)
combinations can provide a
best technology to produce
biodiesel in small farms
CSO Methanol CRL immobilized Heterogeneous Y > 98.4% Cottonseed is another Köse et al.
transesterification Solvent free important crop in Brazil and (2002)
together with lipase production
from Candida rugosa can
provide a
Macaúba oil Ethanol TLL and BCL immobilized in Heterogeneous 90% until 60% in 6th Better results with combination Araki et al.
ZSM-5 transesterification cycle of immobilized lipases (2018)
United State of America
Soybean Oil Methanol Magnetic responsive lipase Enzymatic Y = 82.2 MRL allowed 10 cycles of use Li et al. (2020b)
(MRL) from Thermomyces transesterification with only 10.97% of activity
lanuginosus reaction loss which showed better
operational stability
Waste cooking Methanol Thermomyces lanuginose lipase Enzymatic Y = 80.1 The design variables set at the Banerjee et al.
oil immobilized onto Transesterification methanol to WCO molar ratio (2010)
Superparamagnetic mesoporous of 6.7, a catalyst concentration
silica-SPION core-shell of 35%, a water content of 12%
nanoparticles and a reaction time of 20 h
were identified as the optimal
condition
Soybean Oil Methanol Eversa Transform Lipase Enzymatic Y > 95 Technology for using liquid Nielsen et al.
Transesterification using formulated lipases for (2016)
lipase liquid formulation enzymatic biodiesel production
is new and, since enzyme prices
have been reduced, it is now
possible to simplify the process
considerably and apply it for
very low-quality oils.
Mustard Oil Methanol P. aeruginosa lipase Enzymatic Y = 100 The first study to report Rana et al.
Transesterification transesterification of mustard (2019)
oil via biocatalysis
Soybean Oil Methanol Nanoparticulate Rhizopus Enzymatic Y > 82.2 The results suggested that He et al. (2016)
arrhizus lipase Transesterification loofah sponge is a potential
fungi carrier for an
immobilized whole-cell
biocatalyst.
Olive Oil Methanol- Nanoparticulate Rhizopus Enzymatic Y = 98 Novel lipase formulation for Sharma et al.
Ethanol arrhizus lipase Transesterification efficient catalysis in lipid-to- (2018)
biodiesel conversion
Soapstock oil Methanol Y = 95.2
(continued on next page)

27
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Table 5 (continued )
Feedstock Alcohol Catalyst Technology Results Problem solved Ref

lipase B from Candida Esterification followed by Developed process enables the Su and Wei
antarctica immobilized on a transesterification simple, efficient, and green (2014)
macroporous acrylic resin production of biodiesel from
soapstock oil, providing with a
potential industrial
application.

Note: NA = not available; LCA = Life cycle assessment; X = Conversion; Y = Yield; BBP = biodiesel breakeven price; NPV = net present value.

short-alcohols chain using base catalysts such as NaOH and KOH In Brazil, there is a large production of vegetable grains, which permit
(Quayson et al., 2020a). Meanwhile, National Biofuel Policy (Reno­ the attainment of huge amounts of vegetable oils. Part of this oil is used
vaBio) was employed by 2017 under Brazilian Ministry of Mines and to produce fuel. In this sense, in 2005 biodiesel was compulsorily
Energy (Law 13.576/2017) to reach the biofuel production goals and introduced in energy matrix, being added to diesel oil at 13% content by
to satisfy the COP21 Paris endorsement (Lima et al., 2020). Expan­ 2021. The main Brazilian agricultural crop is soybeans (Vieira et al.,
sion of biodiesel production implemented by Brazilian government is 2021) and consequently the main raw material for biodiesel production
attributed to economic growth, clean environment, and sustainable is the soybean oil (69%), followed by bovine fat (17%), cottonseed oil
biofuels to mitigate carbon emissions (Klein et al., 2019). The Bra­ (1.7%) and other raw materials (Cremonez et al., 2015; da Silva César
zilian biodiesel production was 4.6 million m3 which 71.4% coming et al., 2019). It should be pointed out that soybean oil is the most pro­
from soybean oil and 9% from beef tallow followed palm oil, cot­ duced in the world (El-Hamidi and Zaher, 2018). In the United States the
tonseed oil and waste oils (Ramos et al., 2022). Utilization of soy­ use of soy oil as raw material for biodiesel and the application of
bean oil in biodiesel production is caused by massive feedstock with enzymatic transesterification using methanol are the best actual tech­
the production capacity of 128 million tons by 2020 (Ramos et al., nology for biodiesel production.
2022). It is noteworthy in Table 5 that each country has potential to produce
• United States. USA is known well as the second highest biodiesel its own biodiesel, for internal consumption and exportation. It is shown
producer in the world after Indonesia with its production of 6.5 that feedstocks can be related to internal production of lipases and
billion liters in 2019 (Babadi et al., 2022). Renewable Fuel Standard supports, such as porous solids like zeolites in Indonesia and activated
programme was created to reduce greenhouse gas emissions and carbon from babassu coconut in Brazil. Regarding the most used tech­
expands the renewable oil sectors while reducing reliance on im­ nologies shown in Table 5, in Fig. 9 the most promise technologies for
ported oil (OECD/FAO, 2019). The decrees were offered to raise a the four country cases are shown. Countries with limited vegetable oils
concern towards the environmental issue, especially global warming resources must have developed the local biodiesel production in order to
that is caused by the greenhouse gas emissions. Approximately, 5.93 minimize the biodiesel import and supply (Rahman et al., 2021) because
billion liters of biodiesel were produced by United States of Americ the high biodiesel production cost is mostly (70–85%) caused by
(USA) in 2016 (Chen et al., 2018) and over 1 billion gallons would be expensive raw material price (Sitepu et al., 2020). Meanwhile, 15–30%
produced in 2025 (Babadi et al., 2022). The common feedstock of high total process budget is attributed to high cost of catalyst, energy
employed in USA is soybean oil (55%) thus the soybean-based bio­ consumption, catalyst treatment, and biodiesel purification thus explo­
diesel production was capable of 76% reduction of carbon emission ration of the most effective and efficient biodiesel techbology has been
in USA (Chen et al., 2018). Besides, the biodiesel in USA can be also always conducted (Babadi et al., 2022).
produced from recycled oil (13%) followed by corn oil (12%), animal Commonly, the biodiesel processes consist of four primary technol­
fat (10%), and other edible and non-edible oils (10%). Among ogies which are (i) direct use of mixed oils, (ii) micro-emulsion, (iii)
available biodiesel technologies, the commercial soybean biodiesel pyrolysis, and (iv) transesterification. These technologies are basically
production is conducted by transesterification technology. The USA applied in the biodiesel production aiming to mitigate oil viscosity,
produced biodiesel demonstrates that 95% of biodiesel (B100) improve stability in oxidation process and increase volatility thus the
products fulfill ASTM D6751 standard (Alleman et al., 2013). The obtained biodiesel can be implemented in the engine diesel (Sundus
biodiesel transesterification technology using high FFAs contained et al., 2017). Among those technologies, the transesterification tech­
oils is more synergistic and effective if compared to the oils with low nology is convinced as the most prospective and potential process to use
FFAs contents (Chen et al., 2018). in the biodiesel production due to low energy consumption and efficient
viscosity alleviation (Silitonga et al., 2020). Commercially, the world
Biodiesel production in these countries relies on abundant feedstock. biodiesel is produced over the transesterification technology using high
In percentage, each country contributes woth 8%, 2%, 10% and 17% of FFAs containing vegetable oils and alcohols (Mahdi et al., 2021). The
the re global production, respectively (Lv et al., 2021). Technologies produced biodiesel can be blended by petroleum-based diesel fuel to
applied is counted on the process suitability for the available feedstock appropriately use in diesel engines (Pérez et al., 2018) and should meet
and need as well. In Table 5, technologies with the best applied pro­ ASTM D6751 standard (Joshi et al., 2010) by reducing the biodiesel
cesses are listed for each of these countries. It can be verified in Table 5 viscosity (Ramalingam et al., 2020).
that, for instance, Indonesia prefers to use crude palm oil as the feed­ In the first transesterification step, triglycerides are converted into
stock to produce biodiesel, because it has large palm plantation in each diglycerides and further converted to monoglycerides which are subse­
province thus the low palm oil price offers much profits for Indonesia quently converted to glycerol (Algoufi et al., 2017). The trans­
sustainable biofuel and bioenergy (Simaremare et al., 2021). The esterification process can be conducted either with catalysts or without
transesterification technology is chosen because it is very simple and catalysts by employing primary or secondary monohydric aliphatic al­
cost-effective technology (Bhatia et al., 2021). cohols (Mueanmas et al., 2019). On the other hand, the trans­
In Malaysia, palm oil feedstock and transesterification reaction are esterification process can be proceeded over enzymes as its catalyst
implemented in technologies that have high TRL, providing high bio­ (Bashir et al., 2022) because enzymes are convinced as the renewable
diesel yield and low process price (Alkabbashi et al., 2009). The process and recyclable biocatalysts thus it is capable of reduction of the biodiesel
turned economically feasible with the inclusion of significant recycle process cost (binti Ramlee et al., 2022). To improve the catalyst per­
streams and additional separation equipment (Waudby and Zein, 2021). formance and transesterification efficiency, the catalyst and biocatalysts

28
H.I. Mahdi et al.
29

Fig. 9. Most used technologies for biodiesel using nanobiocatalysts in Indonesia, Malaysia, Brazil and USA.

Chemosphere 319 (2023) 138003


H.I. Mahdi et al. Chemosphere 319 (2023) 138003

in nano particle size called nanocatalysts and nanobiocatalysts have cycle of biodiesel. Furthermore, the implementation of nanocatalyst or
been explored nowadays. This phenomenon is attributed that the nanobiocatalyst is suitable for developing country because it can reduce
nanocatalysts and nanobiocatalysts can significantly enhance the overall costs of the reaction process thanks to its performance for
Brønsted-Lewis acidity, enlarged active sites, and easier accessibility of repeated reaction cycle. Moreover, agroindustrial waste produced from
reactant and product (Azelee et al., 2022). Consequently, the process developed country can be used to synthesized a biodegradable carrier
results in high outcomes and long-term stability with high reusability in that are green and environmental friendly. This strategies can reduce
order to reduce the process cost. waste and cost associated with catalyst disposal. In addition, the support
These improvements are proposed to be implemented in these can be reused by detaching the enzymes from the support and reload
countries, in order to biodiesel production through nanobiocatalysts and with fresh and active enzymes. In this context, the cost for development
internal feedstocks be possible for small and big farmers, as example of of carrier can be minimized. On top of that, the produced biodiesel has
China, that produced biodiesel in this way since 2005. Therefore, shown comparable chemical and physical properties with diesel fuels.
regarding nanobiocatalysts and nanocatalysts, this review provide However, there are still some existing problems especially in develop­
enough information to put these technologies in a new level of readiness ment of catalyst. Thus, more efforts should be put on the development of
and commercialization. Also, points where literature should be novel catalysts with the aim to achieve practical application of nano­
completed demonstrate that research needs a narrow direction catalysts and nanobiocatalysts.
regarding each country or community, where biodiesel production from
nanocatalysts could be a reality in the near future. Authors contribution statement

8. Conclusions and future perspective of nanocatalyst and Hilman Ibnu Mahdi: Writing - review & editing; Conceptualization;
nanobiocatalysts Data curation; Formal Analysis; Hilman Ibnu Mahdi, Rangabha­
shiyam Selvasembian, Lucas Meili and Gayathri Rangasamy:
The production of alternative fuel poses many challenges in term of Conceptualization; Validation; Supervision. Nurfadhila Nasya Ramlee,
catalyst synthesis, proses development and commercialization process. Faisal Amir, José Leandro da Silva Duarte, Leonardo Hadlich de
Hence, development of a selective, thermostability, and high recycla­ Oliveira, Yu-ShenCheng, Nur Izyan Wan Azelee: Data curation;
bility catalyst is crucial for biodiesel industry. The conventional method Formal Analysis; Visualization.
of biodiesel production using homogeneous catalysts requires high en­
ergy consumption for purification and separation of products, diffi­ Declaration of competing interest
culties in catalyst removal, non-recyclability of the catalysts, problems
of corrosion in the reactor and soap formation. Meanwhile, other pros­ The authors declare that they have no known competing financial
pect catalysts, called heterogeneous solid acid and base catalysts, have interests or personal relationships that could have appeared to influence
several deficiencies including soap formation during the reaction, bad the work reported in this paper.
stability in storage as a result of the existence of water and CO2 contents,
and the requirement of acid pretreatments for low-quality reactants. Data availability
The nano-sized heterogeneous catalysts (nanocatalysts and nano­
biocatalysts) are seen to be a better solution due to its larger surface area Data will be made available on request.
and more spacious active sites leading to increased biodiesel yield. There
are several factors that can enhance the catalytic performance and sta­ Acknowledgements
bility including suitable operating conditions, synthesis methods, and
types of metal oxides impregnated with the nanocatalysts. However, the Authors NIWA and NNR greatly appreciate the Ministry of Higher
nanocatalysts have poor activity for reactants with high FFAs and water Education Malaysia under the Fundamental Research Grant Scheme
contents causing soap formation, whilst needs to be carried out at high (FRGS/1/2020/TK0/UTM/02/8) and Universiti Teknologi Malaysia
reaction temperature, longer reaction time, high methanol to oil ratio (UTM) for the support. Authors JLSD and LM thank to Coordenação de
and susceptible to catalyst deactivation after repeated cycle. Aperfeiçoamento de Pessoal de Nível Superior (CAPES, Brazil), Conselho
Alternatively, lipases have a great prospect for the biodiesel pro­ Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Brazil)
duction because those can produce biodiesel even at high FFAs and and Fundação de Amparo à Pesquisa do Estado de Alagoas (FAPEAL,
water containing feedstocks. The enzymes-based biodiesel production Brazil). Author HIM really appreciate PT Global Amines Indonesia and
takes place at mild reaction temperature and does not contribute toward PT Wilmar Nabati Indonesia (WINA-biodiesel production plant) for
environmental pollution. Yet, the enzymes can be easily denatured and supporting and some advice to perfectly enhance the paper quality in
have a very short lifetime. To overcome these drawbacks, immobilized order to support Indonesia government in massive biodiesel production.
lipases in a nano-sized particle called immobilized nanobiocatalysts play
a significant role in the biodiesel production. Immobilized nano­ References
biocatalyst can be used at high enzyme loadings, reusable for several
cycles, and have better protection from enzyme denaturation. Further Abas, N.A., Yusoff, R., Aroua, M.K., Aziz, H.A., Idris, Z., 2020. Production of palm-based
improvement can be studied regarding the separation unit of nano­ glycol ester over solid acid catalysed esterification of lauric acid via microwave
heating. Chem. Eng. J. 382, 122975.
catalysts because conventional separation method involves high oper­ Abdullah, R., Abas, R., Ayatollah, K., 2007. Impact of palm oil-based biodiesel demand
ational cost and leads to catalyst loss. In addition, integration of genetic on palm oil price. Oil Palm Indus. Econ. J. 7, 19–27.
engineering and protein engineering towards producing solvent- Abrahamsson, J., Andreasson, E., Hansson, N., Sandström, D., Wennberg, E.,
Maréchal, M., Martinelli, A., 2015. A Raman spectroscopic approach to investigate
tolerance lipase will solve the problem regarding lipase deactivation the production of biodiesel from soybean oil using 1-alkyl-3-methylimidazolium
during repeated reaction cycle. ionic liquids with intermediate chain length. Appl. Energy 154, 763–770.
Among the latest technologies implemented nowadays, trans­ Adnan, M., Li, K., Xu, L., Yan, Y., 2018. X-shaped ZIF-8 for immobilization rhizomucor
miehei lipase via encapsulation and its application toward biodiesel production.
esterification technology is widely used in the industries in Indonesia,
Catalysts 8, 96.
Malaysia, Brazil and USA. It is the best solution to solve the problem of Adsul, M., Sandhu, S.K., Singhania, R.R., Gupta, R., Puri, S.K., Mathur, A., 2020.
high viscosity along with high efficiency and good cost-saving, high Designing a cellulolytic enzyme cocktail for the efficient and economical conversion
reusability of catalysts, resulting in high conversion with low production of lignocellulosic biomass to biofuels. Enzym. Microb. Technol. 133, 109442.
Afifah, A., Syahrullail, S., Azlee, N.I.W., Rohah, A.M. Synthesis and tribological studies of
cost. The nanocatalyst-assisted process is suitable to be used in devel­ epoxidized palm stearin methyl ester as a green lubricant. J. Clean. Prod. 280,
oping country attributed to its great potential in producing repeated 124320.

30
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Afifah, A., Syahrullail, S., Azlee, N.I.W., Sidik, N.A.C., Yahya, W., Abd Rahim, E., 2019. Ardizzone, S., Bianchi, C., Cappelletti, G., Porta, F., 2004. Liquid-phase catalytic activity
Biolubricant production from palm stearin through enzymatic transesterification of sulfated zirconia from sol–gel precursors: the role of the surface features. J. Catal.
method. Biochem. Eng. J. 148, 178–184. 227, 470–478.
Agarwal, D., Sinha, S., Agarwal, A.K., 2006. Experimental investigation of control of NOx Arezki, R., Jakab, Z., Laxton, D., Matsumoto, A., Nurbekyan, A., Wang, H., Yao, J., 2017.
emissions in biodiesel-fueled compression ignition engine. Renew. Energy 31, Oil prices and the global economy. In: Department, R. (Ed.), International Monetary
2356–2369. Fund.
Agarwal, M., Chauhan, G., Chaurasia, S., Singh, K., 2012. Study of catalytic behavior of Arif, M., Bai, Y., Usman, M., Jalalah, M., Harraz, F.A., Al-Assiri, M., Li, X., Salama, E.-S.,
KOH as homogeneous and heterogeneous catalyst for biodiesel production. Zhang, C., 2020. Highest accumulated microalgal lipids (polar and non-polar) for
J. Taiwan Inst. Chem. Eng. 43, 89–94. biodiesel production with advanced wastewater treatment: role of lipidomics.
Aghababaie, M., Beheshti, M., Razmjou, A., Bordbar, A.-K., 2020. Enzymatic biodiesel Bioresour. Technol. 298, 122299.
production from crude Eruca sativa oil using Candida rugosa lipase in a solvent-free Arumugamurthy, S.S., Sivanandi, P., Pandian, S., Choksi, H., Subramanian, D., 2019.
system using response surface methodology. Biofuels 11, 93–99. Conversion of a low value industrial waste into biodiesel using a catalyst derived
Akhabue, C.E., Osa-Benedict, E.O., Oyedoh, E.A., Otoikhian, S.K., 2020. Development of from brewery waste: an activation and deactivation kinetic study. Waste Manag.
a bio-based bifunctional catalyst for simultaneous esterification and 100, 318–326.
transesterification of neem seed oil: modeling and optimization studies. Renew. Ashjari, M., Garmroodi, M., Asl, F.A., Emampour, M., Yousefi, M., Lish, M.P., Habibi, Z.,
Energy 152, 724–735. Mohammadi, M., 2020. Application of multi-component reaction for covalent
Akia, M., Yazdani, F., Motaee, E., Han, D., Arandiyan, H., 2014a. A review on conversion immobilization of two lipases on aldehyde-functionalized magnetic nanoparticles;
of biomass to biofuel by nanocatalysts. Biofuel Res. J. 16–25. production of biodiesel from waste cooking oil. Process Biochem. 90, 156–167.
Akia, M., Yazdani, F., Motaee, E., Han, D., Arandiyan, H., 2014b. A review on conversion Ashok, A., Kennedy, L.J., Vijaya, J.J., 2019. Structural, optical and magnetic properties
of biomass to biofuel by nanocatalysts. Biofuel Res. J. 1, 16–25. of Zn1-xMnxFe2O4 (0≤ x≤ 0.5) spinel nano particles for transesterification of used
Akinfalabi, S.-I., Rashid, U., Ngamcharussrivichai, C., Nehdi, I.A., 2020a. Synthesis of cooking oil. J. Alloys Compd. 780, 816–828.
reusable biobased nano-catalyst from waste sugarcane bagasse for biodiesel Atabani, A.E., Silitonga, A.S., Badruddin, I.A., Mahlia, T., Masjuki, H., Mekhilef, S., 2012.
production. Environ. Technol. Innov. 18. A comprehensive review on biodiesel as an alternative energy resource and its
Akinfalabi, S.-I., Rashid, U., Ngamcharussrivichai, C., Nehdi, I.A., 2020b. Synthesis of characteristics. Renew. Sustain. Energy Rev. 16, 2070–2093.
reusable biobased nano-catalyst from waste sugarcane bagasse for biodiesel Atabani, A., Silitonga, A., Ong, H., Mahlia, T., Masjuki, H., Badruddin, I.A., Fayaz, H.,
production. Environ. Technol. Innov., 100788 2013. Non-edible vegetable oils: a critical evaluation of oil extraction, fatty acid
Akoh, C.C., Chang, S.-W., Lee, G.-C., Shaw, J.-F., 2007. Enzymatic approach to biodiesel compositions, biodiesel production, characteristics, engine performance and
production. J. Agric. Food Chem. 55, 8995–9005. emissions production. Renew. Sustain. Energy Rev. 18, 211–245.
Al-Jaberi, S.H.H., Rashid, U., Al-Doghachi, F.A.J., Abdulkareem-Alsultan, G., Taufiq- Atadashi, I., Aroua, M.K., Aziz, A.A., Sulaiman, N., 2012. The effects of water on
Yap, Y.H., 2017. Synthesis of MnO-NiO-SO 4 − 2/ZrO 2 solid acid catalyst for methyl biodiesel production and refining technologies: a review. Renew. Sustain. Energy
ester production from palm fatty acid distillate. Energy Convers. Manag. 139, Rev. 16, 3456–3470.
166–174. Atadashi, I., Aroua, M., Aziz, A.A., Sulaiman, N., 2013. The effects of catalysts in
Al-Zuhair, S., Ling, F.W., Jun, L.S., 2007. Proposed kinetic mechanism of the production biodiesel production: a review. J. Ind. Eng. Chem. 19, 14–26.
of biodiesel from palm oil using lipase. Process Biochem. 42, 951–960. Atapour, M., Kariminia, H.-R., Moslehabadi, P.M., 2014. Optimization of biodiesel
Alaei, S., Haghighi, M., Toghiani, J., Vahid, B.R., 2018. Magnetic and reusable MgO/ production by alkali-catalyzed transesterification of used frying oil. Process Saf.
MgFe2O4 nanocatalyst for biodiesel production from sunflower oil: influence of fuel Environ. Protect. 92, 179–185.
ratio in combustion synthesis on catalytic properties and performance. Ind. Crop. Awais, M., Musmar, S.e.A., Kabir, F., Batool, I., Rasheed, M.A., Jamil, F., Khan, S.U.,
Prod. 117, 322–332. Tlili, I., 2020. Biodiesel production from Melia azedarach and Ricinus communis oil
Alaei, S., Haghighi, M., Rahmanivahid, B., Shokrani, R., Naghavi, H., 2020. Conventional by transesterification process. Catalysts 10, 427.
vs. hybrid methods for dispersion of MgO over magnetic Mg–Fe mixed oxides Azelee, N.I.W., da Silva Santos, D.H., Meili, L., Mahdi, H.I., 2022. Commercial green
nanocatalyst in biofuel production from vegetable oil. Renew. Energy 154, diesel production under hydroprocessing technology using solid-based
1188–1203. heterogeneous catalysts. Green Diesel: Alternative Biodiesel Petrodiesel. Springer
Algoufi, Y., Akpan, U., Kabir, G., Asif, M., Hameed, B., 2017. Upgrading of glycerol from 149–204.
biodiesel synthesis with dimethyl carbonate on reusable Sr–Al mixed oxide catalysts. Babadi, A.A., Rahmati, S., Fakhlaei, R., Barati, B., Wang, S., Doherty, W., Ostrikov, K.,
Energy Convers. Manag. 138, 183–189. 2022. Emerging technologies for biodiesel production: processes, challenges, and
Ali, S.A., Suboyin, A., Haj, H.B., 2018. Unconventional and conventional oil production opportunities. Biomass Bioenergy 163, 106521.
impacts on oil price: lessons learnt with glance to the future. J. Global Econ. 6, 1–9. Babaki, M., Yousefi, M., Habibi, Z., Mohammadi, M., Brask, J., 2015. Effect of water,
Alkabbashi, A.N., Alam, M.Z., Mirghani, M.E.S., Al-Fusaiel, A.M.A., 2009. Biodiesel organic solvent and adsorbent contents on production of biodiesel fuel from canola
production from crude palm oil by transesterification process. J. Appl. Sci. 9, oil catalyzed by various lipases immobilized on epoxy-functionalized silica as low
3166–3170. cost biocatalyst. J. Mol. Catal. B Enzym. 120, 93–99.
Alleman, T.L., Fouts, L., Chupka, G., 2013. Quality Parameters and Chemical Analysis for Babaki, M., Yousefi, M., Habibi, Z., Mohammadi, M., Yousefi, P., Mohammadi, J.,
Biodiesel Produced in the United States in 2011. National Renewable Energy Lab. Brask, J., 2016. Enzymatic production of biodiesel using lipases immobilized on
(NREL), Golden, CO (United States). silica nanoparticles as highly reusable biocatalysts: effect of water, t-butanol and
Ambat, I., Srivastava, V., Sillanpää, M., 2018. Recent advancement in biodiesel blue silica gel contents. Renew. Energy 91, 196–206.
production methodologies using various feedstock: a review. Renew. Sustain. Energy Badgujar, V.C., Badgujar, K.C., Yeole, P.M., Bhanage, B.M., 2019. Enhanced biocatalytic
Rev. 90, 356–369. activity of immobilized steapsin lipase in supercritical carbon dioxide for production
Amini, Z., Ilham, Z., Ong, H.C., Mazaheri, H., Chen, W.-H., 2017. State of the art and of biodiesel using waste cooking oil. Bioproc. Biosyst. Eng. 42, 47–61.
prospective of lipase-catalyzed transesterification reaction for biodiesel production. Badoei-dalfard, A., Malekabadi, S., Karami, Z., Sargazi, G., 2019. Magnetic cross-linked
Energy Convers. Manag. 141, 339–353. enzyme aggregates of Km12 lipase: a stable nanobiocatalyst for biodiesel synthesis
Amoah, J., Ho, S.-H., Hama, S., Yoshida, A., Nakanishi, A., Hasunuma, T., Ogino, C., from waste cooking oil. Renew. Energy 141, 874–882.
Kondo, A., 2016. Lipase cocktail for efficient conversion of oils containing Badoei-Dalfard, A., Shahba, A., Zaare, F., Sargazi, G., Seyedalipour, B., Karami, Z., 2021.
phospholipids to biodiesel. Bioresour. Technol. 211, 224–230. Lipase immobilization on a novel class of Zr-MOF/electrospun nanofibrous
Andrade, M.F., Parussulo, A.L., Netto, C.G., Andrade, L.H., Toma, H.E., 2016. Lipase polymers: biochemical characterization and efficient biodiesel production. Int. J.
immobilized on polydopamine-coated magnetite nanoparticles for biodiesel Biol. Macromol. 192, 1292–1303.
production from soybean oil. Biofuel Res. J. 3, 403–409. Ballotin, F.C., da Silva, M.J., Lago, R.M., de Carvalho Teixeira, A.P., 2020a. Solid acid
Andrade, T.A., Errico, M., Christensen, K.V., 2017. Influence of the reaction conditions catalysts based on sulfonated carbon nanostructures embedded in an amorphous
on the enzyme catalyzed transesterification of castor oil: a possible step in biodiesel matrix produced from bio-oil: esterification of oleic acid with methanol. J. Environ.
production. Bioresour. Technol. 243, 366–374. Chem. Eng., 103674
Andrade, T.A., Martin, M., Errico, M., Christensen, K.V., 2019. Biodiesel production Ballotin, F.C., da Silva, M.J., Lago, R.M., de Carvalho Teixeira, A.P., 2020b. Solid acid
catalyzed by liquid and immobilized enzymes: optimization and economic analysis. catalysts based on sulfonated carbon nanostructures embedded in an amorphous
Chem. Eng. Res. Des. 141, 1–14. matrix produced from bio-oil: esterification of oleic acid with methanol. J. Environ.
Andrade, T.A., Errico, M., Christensen, K.V., 2020. Implementation of Biodiesel Chem. Eng. 8, 103674.
Production Process Using Enzyme-Catalyzed Routes. Process Systems Engineering Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., Pandey, R., 2010.
for Biofuels Development, pp. 191–220. Commercializing lignocellulosic bioethanol: technology bottlenecks and possible
Andreo-Martínez, P., Ortiz-Martínez, V.M., García-Martínez, N., de los Ríos, A.P., remedies. Biofuels, Bioproducts and Biorefining: Innov. Sustain. Econ. 4, 77–93.
Hernández-Fernández, F.J., Quesada-Medina, J., 2020. Production of biodiesel Barbosa, O., Ortiz, C., Torres, R., Fernandez-Lafuente, R., 2011. Effect of the
under supercritical conditions: state of the art and bibliometric analysis. Appl. immobilization protocol on the properties of lipase B from Candida Antarctica in
Energy 264, 114753. organic media: enantiospecifc production of atenolol acetate. J. Mol. Catal. B Enzym.
Araki, C.A., Marcucci, S.M.P., da Silva, L.S., Maeda, C.H., Arroyo, P.A., Zanin, G.M., 71, 124–132.
2018. Effects of a combination of lipases immobilised on desilicated and thiol- Barbosa, O., Torres, R., Ortiz, C., Berenguer-Murcia, A.n., Rodrigues, R.C., Fernandez-
modified ZSM-5 for the synthesis of ethyl esters from macauba pulp oil in a solvent- Lafuente, R., 2013. Heterofunctional supports in enzyme immobilization: from
free system. Appl. Catal. Gen. 562, 241–249. traditional immobilization protocols to opportunities in tuning enzyme properties.
Aransiola, E.F., Ojumu, T.V., Oyekola, O., Madzimbamuto, T., Ikhu-Omoregbe, D., 2014. Biomacromolecules 14, 2433–2462.
A review of current technology for biodiesel production: state of the art. Biomass Barbosa, O., Ortiz, C., Berenguer-Murcia, Á., Torres, R., Rodrigues, R.C., Fernandez-
Bioenergy 61, 276–297. Lafuente, R., 2015. Strategies for the one-step immobilization–purification of
enzymes as industrial biocatalysts. Biotechnol. Adv. 33, 435–456.

31
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Baroutian, S., Aroua, M.K., Raman, A.A., Sulaiman, N.M.N., 2008. Density of palm oil- United States with induced land use change impacts. Bioresour. Technol. 251,
based methyl ester. J. Chem. Eng. Data 53, 877–880. 249–258.
Barros, S., 2019. Brazil Biofuels Annual 2019. USDA. Cheng, J., Qiu, Y., Huang, R., Yang, W., Zhou, J., Cen, K., 2016. Biodiesel production
Bartha-Vári, J.-H., Moisă, M.E., Bencze, L.C., Irimie, F.-D., Paizs, C., Toșa, M.I., 2020. from wet microalgae by using graphene oxide as solid acid catalyst. Bioresour.
Efficient biodiesel production catalyzed by nanobioconjugate of lipase from Technol. 221, 344–349.
Pseudomonas fluorescens. Molecules 25, 651. Cheng, J., Qiu, Y., Zhang, J., Huang, R., Yang, W., Fan, Z., 2017. Conversion of lipids
Bashir, M.A., Wu, S., Zhu, J., Krosuri, A., Khan, M.U., Aka, R.J.N., 2022. Recent from wet microalgae into biodiesel using sulfonated graphene oxide catalysts.
development of advanced processing technologies for biodiesel production: a critical Bioresour. Technol. 244, 569–574.
review. Fuel Process. Technol. 227, 107120. Cherian, E., Yazhini, D., Victor, M., Baskar, G., 2019. Production of biodiesel from pork
Baskar, G., Soumiya, S., Aiswarya, R., 2016. Biodiesel production from pongamia oil fat using alumina-doped calcium oxide nanocomposite as heterogeneous catalyst.
using magnetic composite of zinc oxide nanocatalyst. Int. J. Mod. Sci. Technol. 1, Energy Sources, Part A Recovery, Util. Environ. Eff. 1–10.
129–137. Chhetri, A., Watts, K., Islam, M., 2008. Waste cooking oil as an alternate feedstock for
Bastos, R.R.C., da Luz Corrêa, A.P., da Luz, P.T.S., da Rocha Filho, G.N., Zamian, J.R., da biodiesel production. Energies 1, 3–18.
Conceição, L.R.V., 2020. Optimization of biodiesel production using sulfonated Ching-Velasquez, J., Fernández-Lafuente, R., Rodrigues, R.C., Plata, V., Rosales-
carbon-based catalyst from an amazon agro-industrial waste. Energy Convers. Quintero, A., Torrestiana-Sánchez, B., Tacias-Pascacio, V.G., 2020. Production and
Manag. 205, 112457. Characterization of Biodiesel from Oil of Fish Waste by Enzymatic Catalysis.
Beck, Á., Pölczmann, G., Eller, Z., Hancsók, J., 2014. Investigation of the effect of Renewable Energy.
detergent–dispersant additives on the oxidation stability of biodiesel, diesel fuel and Chingakham, C., David, A., Sajith, V., 2019. Fe3O4 nanoparticles impregnated eggshell
their blends. Biomass Bioenergy 66, 328–336. as a novel catalyst for enhanced biodiesel production. Chin. J. Chem. Eng. 27,
Bencze, L.C., Bartha-Vari, J.H., Katona, G., Tosa, M.I., Paizs, C., Irimie, F.D., 2016. 2835–2843.
Nanobioconjugates of Candida Antarctica lipase B and single-walled carbon Chouhan, A.S., Sarma, A., 2011. Modern heterogeneous catalysts for biodiesel
nanotubes in biodiesel production. Bioresour. Technol. 200, 853–860. production: a comprehensive review. Renew. Sustain. Energy Rev. 15, 4378–4399.
Benjumea, P., Agudelo, J., Agudelo, A., 2008. Basic properties of palm oil Chua, S.Y., Goh, C.M.H., Tan, Y.H., Mubarak, N.M., Kansedo, J., Khalid, M.,
biodiesel–diesel blends. Fuel 87, 2069–2075. Walvekar, R., Abdullah, E., 2020. Biodiesel synthesis using natural solid catalyst
Bernal, C., Illanes, A., Wilson, L., 2014. Heterofunctional hydrophilic–hydrophobic derived from biomass waste—a review. J. Ind. Eng. Chem. 81, 41–60.
porous silica as support for multipoint covalent immobilization of lipases: Chung, K.-H., Park, B.-G., 2009. Esterification of oleic acid in soybean oil on zeolite
application to lactulose palmitate synthesis. Langmuir 30, 3557–3566. catalysts with different acidity. J. Ind. Eng. Chem. 15, 388–392.
Bhatia, S.K., Bhatia, R.K., Jeon, J.-M., Pugazhendhi, A., Awasthi, M.K., Kumar, D., Cirujano, F.G., Dhakshinamoorthy, A., 2021. Engineering of active sites in metal–organic
Kumar, G., Yoon, J.-J., Yang, Y.-H., 2021. An overview on advancements in biobased frameworks for biodiesel production. Adv. Sustain. Syst. 5.
transesterification methods for biodiesel production: oil resources, extraction, Coggon, R., Vasudevan, P.T., Sanchez, F., 2007. Enzymatic transesterification of olive oil
biocatalysts, and process intensification technologies. Fuel 285, 119117. and its precursors. Biocatal. Biotransform. 25, 135–143.
binti Ramlee, N.N., Mahdi, H.I., Azelee, N.I.W., 2022. Biodiesel Production Using Costa, M.W., Oliveira, A.A., 2022. Social life cycle assessment of feedstocks for biodiesel
Enzymatic Catalyst. Biofuels and Bioenergy. Elsevier, pp. 133–169. production in Brazil. Renew. Sustain. Energy Rev. 159, 112166.
Booramurthy, V.K., Kasimani, R., Subramanian, D., Pandian, S., 2020. Production of Cremonez, P.A., Feroldi, M., Nadaleti, W.C., de Rossi, E., Feiden, A., de Camargo, M.P.,
biodiesel from tannery waste using a stable and recyclable nano-catalyst: an Cremonez, F.E., Klajn, F.F., 2015. Biodiesel production in Brazil: current scenario
optimization and kinetic study. Fuel 260, 116373. and perspectives. Renew. Sustain. Energy Rev. 42, 415–428.
Borah, M.J., Devi, A., Borah, R., Deka, D., 2019. Synthesis and application of Co doped da Conceição, L.R.V., Carneiro, L.M., Giordani, D.S., de Castro, H.F., 2017. Synthesis of
ZnO as heterogeneous nanocatalyst for biodiesel production from non-edible oil. biodiesel from macaw palm oil using mesoporous solid catalyst comprising 12-
Renew. Energy 133, 512–519. molybdophosphoric acid and niobia. Renew. Energy 113, 119–128.
Borges, M.E., Díaz, L., 2012. Recent developments on heterogeneous catalysts for da Silva César, A., Conejero, M.A., Ribeiro, E.C.B., Batalha, M.O., 2019. Competitiveness
biodiesel production by oil esterification and transesterification reactions: a review. analysis of “social soybeans” in biodiesel production in Brazil. Renew. Energy 133,
Renew. Sustain. Energy Rev. 16, 2839–2849. 1147–1157.
Bose, S., Armstrong, D.W., Petrich, J.W., 2010. Enzyme-catalyzed hydrolysis of cellulose Dahdah, E., Estephane, J., Haydar, R., Youssef, Y., El Khoury, B., Gennequin, C.,
in ionic liquids: a green approach toward the production of biofuels. J. Phys. Chem. Aboukaïs, A., Abi-Aad, E., Aouad, S., 2020. Biodiesel production from refined
B 114, 8221–8227. sunflower oil over Ca–Mg–Al catalysts: effect of the composition and the thermal
Boviatsi, E., Papadaki, A., Efthymiou, M.N., Nychas, G.J.E., Papanikolaou, S., da Silva, J. treatment. Renew. Energy 146, 1242–1248.
A., Freire, D.M., Koutinas, A., 2020. Valorisation of sugarcane molasses for the Dai, Y.-M., Wu, J.-S., Chen, C.-C., Chen, K.-T., 2015. Evaluating the optimum operating
production of microbial lipids via fermentation of two Rhodosporidium strains for parameters on transesterification reaction for biodiesel production over a LiAlO2
enzymatic synthesis of polyol esters. J. Chem. Technol. Biotechnol. 95, 402–407. catalyst. Chem. Eng. J. 280, 370–376.
Boxi, S.S., Paria, S., 2014. Effect of silver doping on TiO 2, CdS, and ZnS nanoparticles for Damanik, N., Ong, H.C., Chong, W.T., Silitonga, A.S., 2017. Biodiesel production from
the photocatalytic degradation of metronidazole under visible light. RSC Adv. 4, Calophyllum inophyllum− palm mixed oil. Energy Sources, Part A Recovery, Util.
37752–37760. Environ. Eff. 39, 1283–1289.
Busto, M., Meza, V., Ortega, N., Perez-Mateos, M., 2007. Immobilization of naringinase De, A., Boxi, S.S., 2020. Application of Cu impregnated TiO2 as a heterogeneous
from Aspergillus Niger CECT 2088 in poly (vinyl alcohol) cryogels for the debittering nanocatalyst for the production of biodiesel from palm oil. Fuel 265, 117019.
of juices. Food Chem. 104, 1177–1182. de Castro, S.M., de Castro, A.M., 2012. Assessment of the Brazilian potential for the
Cabral, N.M., Lorenti, J.P., Plass, W., Gallo, J.M.R., 2020. Solid acid resin Amberlyst 45 production of enzymes for biofuels from agroindustrial materials. Biomass Conv.
as a catalyst for the transesterification of vegetable oil. Front. Chem. 8, 305. Biorefinery 2, 87–107.
Can, Ö., 2014. Combustion characteristics, performance and exhaust emissions of a De Diego, T., Manjón, A., Lozano, P., Iborra, J.L., 2011. A recyclable enzymatic biodiesel
diesel engine fueled with a waste cooking oil biodiesel mixture. Energy Convers. production process in ionic liquids. Bioresour. Technol. 102, 6336–6339.
Manag. 87, 676–686. Dechakhumwat, S., Hongmanorom, P., Thunyaratchatanon, C., Smith, S.M.,
Canakci, M., 2007. The potential of restaurant waste lipids as biodiesel feedstocks. Boonyuen, S., Luengnaruemitchai, A., 2020. Catalytic activity of heterogeneous acid
Bioresour. Technol. 98, 183–190. catalysts derived from corncob in the esterification of oleic acid with methanol.
Canakci, M., Van Gerpen, J., 2001. Biodiesel production from oils and fats with high free Renew. Energy 148, 897–906.
fatty acids. Trans. ASAE 44, 1429. Dehghani, S., Haghighi, M., 2020. Sono-enhanced dispersion of CaO over Zr-Doped
Cantone, S., Ferrario, V., Corici, L., Ebert, C., Fattor, D., Spizzo, P., Gardossi, L., 2013. MCM-41 bifunctional nanocatalyst with various Si/Zr ratios for conversion of waste
Efficient immobilisation of industrial biocatalysts: criteria and constraints for the cooking oil to biodiesel. Renew. Energy 153, 801–812.
selection of organic polymeric carriers and immobilisation methods. Chem. Soc. Rev. Deive, F.J., Rodríguez, A., 2020. Ionic Liquids for Enzyme-Catalyzed Production of
42, 6262–6276. Biodiesel. Green Sustainable Process for Chemical and Environmental Engineering
Cao, L., 2006. Carrier-bound Immobilized Enzymes: Principles, Application and Design. and Science. Elsevier, pp. 31–47.
John Wiley & Sons. Dewi, Citra, Slamet, A.S., 2019. Novel approach of esterification process using
Chang, M.Y., Chan, E.-S., Song, C.P., Biodiesel production catalysed by low-cost liquid heterogeneous catalyst in biodiesel synthesis from waste cooking oil. IOP Conf. Ser.
enzyme Eversa® Transform 2.0: effect of free fatty acid content on lipase methanol Mater. Sci. Eng. 509.
tolerance and kinetic model. Fuel 283, 119266. Di Serio, M., Tesser, R., Pengmei, L., Santacesaria, E., 2008. Heterogeneous catalysts for
Chattopadhyay, S., Sen, R., 2020. Materials and methods for biodiesel production. biodiesel production. Energy Fuel. 22, 207–217.
Sustain. Agric. Rev. 39, 179–204. Ding, C., Liang, J., Zhou, Z., Li, Y., Peng, W., Zhang, G., Zhang, F., Fan, X., 2019.
Chen, X., Du, W., Liu, D., 2008. Response surface optimization of biocatalytic biodiesel Photothermal enhanced enzymatic activity of lipase covalently immobilized on
production with acid oil. Biochem. Eng. J. 40, 423–429. functionalized Ti3C2TX nanosheets. Chem. Eng. J. 378, 122205.
Chen, Y., Xiao, B., Chang, J., Fu, Y., Lv, P., Wang, X., 2009. Synthesis of biodiesel from Dizge, N., Keskinler, B., 2008. Enzymatic production of biodiesel from canola oil using
waste cooking oil using immobilized lipase in fixed bed reactor. Energy Convers. immobilized lipase. Biomass Bioenergy 32, 1274–1278.
Manag. 50, 668–673. dos Santos, T.C., Santos, E.C., Dias, J.P., Barreto, J., Stavale, F.L., Ronconi, C.M., 2019.
Chen, G., Liu, J., Qi, Y., Yao, J., Yan, B., 2016. Biodiesel production using magnetic Reduced graphene oxide as an excellent platform to produce a stable Brønsted acid
whole-cell biocatalysts by immobilization of Pseudomonas mendocina on Fe3O4- catalyst for biodiesel production. Fuel 256, 115793.
chitosan microspheres. Biochem. Eng. J. 113, 86–92. Du, L., Li, Z., Ding, S., Chen, C., Qu, S., Yi, W., Lu, J., Ding, J., 2019. Synthesis and
Chen, R., Qin, Z., Han, J., Wang, M., Taheripour, F., Tyner, W., O’Connor, D., Duffield, J., characterization of carbon-based MgO catalysts for biodiesel production from castor
2018. Life cycle energy and greenhouse gas emission effects of biodiesel in the oil. Fuel 258, 116122.

32
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Dumri, K., Hung Anh, D., 2014. Immobilization of lipase on silver nanoparticles via Guan, F., Peng, P., Wang, G., Yin, T., Peng, Q., Huang, J., Guan, G., Li, Y., 2010.
adhesive polydopamine for biodiesel production. Enzym. Res. 2014, 389739. Combination of two lipases more efficiently catalyzes methanolysis of soybean oil for
El-Hamidi, M., Zaher, F.A., 2018. Production of vegetable oils in the world and in Egypt: biodiesel production in aqueous medium. Process Biochem. 45, 1677–1682.
an overview. Bull. Natl. Res. Cent. 42, 1–9. Guldhe, A., Singh, B., Rawat, I., Permaul, K., Bux, F., 2015. Biocatalytic conversion of
Encinar, J.M., Gonzalez, J.F., Rodríguez-Reinares, A., 2005. Biodiesel from used frying lipids from microalgae Scenedesmus obliquus to biodiesel using Pseudomonas
oil. Variables affecting the yields and characteristics of the biodiesel. Ind. Eng. Chem. fluorescens lipase. Fuel 147, 117–124.
Res. 44, 5491–5499. Guo, F., Fang, Z., Xu, C.C., Smith Jr., R.L., 2012. Solid acid mediated hydrolysis of
Esmaeili, H., Foroutan, R., 2018. Optimization of biodiesel production from goat tallow biomass for producing biofuels. Prog. Energy Combust. Sci. 38, 672–690.
using alkaline catalysts and combining them with diesel. Chem. Chem. Technol. 1 Guo, J., Sun, S., Liu, J., 2020. Conversion of waste frying palm oil into biodiesel using
(12), 120–126. free lipase A from Candida Antarctica as a novel catalyst. Fuel 267, 117323.
Fan, Y., Ke, C., Su, F., Li, K., Yan, Y., 2017. Various types of lipases immobilized on Gupta, J., Agarwal, M., Dalai, A., 2020. An overview on the recent advancements of
dendrimer-functionalized magnetic nanocomposite and application in biodiesel sustainable heterogeneous catalysts and prominent continuous reactor for biodiesel
preparation. Energy Fuel. 31, 4372–4381. production. J. Ind. Eng. Chem.
Farid, M.A.A., Roslan, A.M., Hassan, M.A., Hasan, M.Y., Othman, M.R., Shirai, Y., 2020. Gurunathan, B., Ravi, A., 2015a. Biodiesel production from waste cooking oil using
Net energy and techno-economic assessment of biodiesel production from waste copper doped zinc oxide nanocomposite as heterogeneous catalyst. Bioresour.
cooking oil using a semi-industrial plant: a Malaysia perspective. Sustain. Energy Technol. 188, 124–127.
Technol. Assessments 39, 100700. Gurunathan, B., Ravi, A., 2015b. Process optimization and kinetics of biodiesel
Fatima, Z., 2021. Use of Nanomaterials for the Immobilization of Industrially Important production from neem oil using copper doped zinc oxide heterogeneous
Enzymes. nanocatalyst. Bioresour. Technol. 190, 424–428.
Fatima, A., Mumtaz, M.W., Mukhtar, H., Akram, S., Touqeer, T., Rashid, U., Ul Gwenzi, W., Simbanegavi, T.T., Mahdi, H.I., Azelee, N.I.W., Muisa-Zikali, N.,
Mustafa, M.R., Nehdi, I.A., Saiman, M.I., 2020. Synthesis of lipase-immobilized Rangabhashiyam, S., 2022. Occurrence and ecological health risks of microplastics.
CeO2 nanorods as heterogeneous nano-biocatalyst for optimized biodiesel Emerg. Contam. Terr.-Aquat.-Atmos. Continuum 243–270.
production from eruca sativa seed oil. Catalysts 10, 231. Halim, S.F.A., Kamaruddin, A.H., 2008. Catalytic studies of lipase on FAME production
Ferrero, G.O., Rojas, H.J., Argaraña, C.E., Eimer, G.A., 2016. Towards sustainable biofuel from waste cooking palm oil in a tert-butanol system. Process Biochem. 43,
production: design of a new biocatalyst to biodiesel synthesis from waste oil and 1436–1439.
commercial ethanol. J. Clean. Prod. 139, 495–503. Halim, R., Gladman, B., Danquah, M.K., Webley, P.A., 2011. Oil extraction from
Ferrero, G.O., Faba, E.M.S., Rickert, A.A., Eimer, G.A., 2020. Alternatives to rethink microalgae for biodiesel production. Bioresour. Technol. 102, 178–185.
tomorrow: biodiesel production from residual and non-edible oils using biocatalyst Hama, S., Tamalampudi, S., Yoshida, A., Tamadani, N., Kuratani, N., Noda, H.,
technology. Renew. Energy 150, 128–135. Fukuda, H., Kondo, A., 2011. Process engineering and optimization of glycerol
Feyzi, M., Shahbazi, E., 2015. Catalytic performance and characterization of Cs–Ca/ separation in a packed-bed reactor for enzymatic biodiesel production. Bioresour.
SiO2–TiO2 nanocatalysts for biodiesel production. J. Mol. Catal. Chem. 404, Technol. 102, 10419–10424.
131–138. Hama, S., Noda, H., Kondo, A., 2018. How lipase technology contributes to evolution of
Feyzi, M., Nourozi, L., Zakarianezhad, M., 2014. Preparation and characterization of biodiesel production using multiple feedstocks. Curr. Opin. Biotechnol. 50, 57–64.
magnetic CsH2PW12O40/Fe–SiO2 nanocatalysts for biodiesel production. Mater. Hanefeld, U., Gardossi, L., Magner, E., 2009. Understanding enzyme immobilisation.
Res. Bull. 60, 412–420. Chem. Soc. Rev. 38, 453–468.
Feyzi, M., Hosseini, N., Yaghobi, N., Ezzati, R., 2017. Preparation, characterization, Hanif, M., Shamsuddin, A., Nomanbhay, S., Fazril, I., Kusumo, F., Akhiar, A., 2019. Energ
kinetic and thermodynamic studies of MgO-La2O3 nanocatalysts for biodiesel y saving potential using elite jatropha curcas hybrid for biodiesel production i n
production from sunflower oil. Chem. Phys. Lett. 677, 19–29. Malaysia. Int. J. Recent Technol. Eng. 8, 6281–6287.
Firdaus, M.Y., Guo, Z., Fedosov, S.N., 2016. Development of kinetic model for biodiesel Hannon, M., Gimpel, J., Tran, M., Rasala, B., Mayfield, S., 2010. Biofuels from algae:
production using liquid lipase as a biocatalyst, esterification step. Biochem. Eng. J. challenges and potential. Biofuels 1, 763–784.
105, 52–61. Harahap, F., Silveira, S., Khatiwada, D., 2019. Cost competitiveness of palm oil biodiesel
Fjerbaek, L., Christensen, K.V., Norddahl, B., 2009. A review of the current state of production in Indonesia. Energy 170, 62–72.
biodiesel production using enzymatic transesterification. Biotechnol. Bioeng. 102, Harreh, D., Saleh, A., Reddy, A., Hamdan, S., 2018. An experimental investigation of
1298–1315. karanja biodiesel production in Sarawak, Malaysia. J. Eng. 2018.
Foroutan, R., Mohammadi, R., Esmaeili, H., Bektashi, F.M., Tamjidi, S., 2020. Hartmann, M., Kostrov, X., 2013. Immobilization of enzymes on porous silicas–benefits
Transesterification of waste edible oils to biodiesel using calcium oxide@ and challenges. Chem. Soc. Rev. 42, 6277–6289.
magnesium oxide nanocatalyst. Waste Manag. 105, 373–383. Hashmi, S., Gohar, S., Mahmood, T., Nawaz, U., Farooqi, H., 2016. Biodiesel production
Galadima, A., Muraza, O., 2014. Biodiesel production from algae by using heterogeneous by using CaO-Al2O3 Nano catalyst. Int. J. Eng. Res. Sci. 2, 43–49.
catalysts: a critical review. Energy 78, 72–83. Hasni, K., Ilham, Z., Dharma, S., Varman, M., 2017. Optimization of biodiesel production
Gao, J., Yin, L., Feng, K., Zhou, L., Ma, L., He, Y., Wang, L., Jiang, Y., 2016. Lipase from Brucea javanica seeds oil as novel non-edible feedstock using response surface
immobilization through the combination of bioimprinting and cross-linked protein- methodology. Energy Convers. Manag. 149, 392–400.
coated microcrystal technology for biodiesel production. Ind. Eng. Chem. Res. 55, He, Q., Shi, H., Gu, H., Naka, G., Ding, H., Li, X., Zhang, Y., Hu, B., Wang, F., 2016.
11037–11043. Immobilization of Rhizopus oryzae ly6 onto loofah sponge as a whole-cell
Garcia-Galan, C., Berengue Murcia, Á., Fernandez Lafuente, R., Rodrigues, R.C., 2011. biocatalyst for biodiesel production. Bioresources 11, 850–860.
Potential of different enzyme immobilization strategies to improve enzyme He, Y., Wang, X., Zhang, Y., Guo, Z., Jiang, Y., Chen, F., 2019. Enzymatic ethanolysis
performance. Adv. Synth. Catal. 353, 2885–2904. subjected to Schizochytrium biomass: sequential processing for DHA enrichment and
Gardy, J., Hassanpour, A., Lai, X., Ahmed, M.H., 2016. Synthesis of Ti (SO4) O solid acid biodiesel production. Energy Convers. Manag. 184, 159–171.
nano-catalyst and its application for biodiesel production from used cooking oil. He, Y., Zhang, B., Guo, S., Guo, Z., Chen, B., Wang, M., 2020. Sustainable biodiesel
Appl. Catal. Gen. 527, 81–95. production from the green microalgae Nannochloropsis: novel integrated processes
Gardy, J., Osatiashtiani, A., Céspedes, O., Hassanpour, A., Lai, X., Lee, A.F., Wilson, K., from cultivation to enzyme-assisted extraction and ethanolysis of lipids. Energy
Rehan, M., 2018. A magnetically separable SO4/Fe-Al-TiO2 solid acid catalyst for Convers. Manag. 209, 112618.
biodiesel production from waste cooking oil. Appl. Catal. B Environ. 234, 268–278. Helwani, Z., Aziz, N., Kim, J., Othman, M., 2016. Improving the yield of Jatropha
Gardy, J., Nourafkan, E., Osatiashtiani, A., Lee, A.F., Wilson, K., Hassanpour, A., Lai, X., curcas’s FAME through sol–gel derived meso-porous hydrotalcites. Renew. Energy
2019. A core-shell SO4/Mg-Al-Fe3O4 catalyst for biodiesel production. Appl. Catal. 86, 68–74.
B Environ. 259, 118093. Hernández-Hipólito, P., Juárez-Flores, N., Martínez-Klimova, E., Gómez-Cortés, A.,
Gawande, M.B., Pandey, R.K., Jayaram, R.V., 2012. Role of mixed metal oxides in Bokhimi, X., Escobar-Alarcón, L., Klimova, T.E., 2015. Novel heterogeneous basic
catalysis science—versatile applications in organic synthesis. Catal. Sci. Technol. 2, catalysts for biodiesel production: sodium titanate nanotubes doped with potassium.
1113–1125. Catal. Today 250, 187–196.
Ghalandari, A., Taghizadeh, M., Rahmani, M., 2019. Statistical optimization of the Hidayat, A., Rochmadi Wijaya, K., Budiman, A., 2015. Esterification of free fatty acid on
biodiesel production process using a magnetic core-mesoporous shell KOH/Fe3O4@ palm fatty acid distillate using activated carbon catalysts generated from coconut
γ-Al2O3 nanocatalyst. Chem. Eng. Technol. 42, 89–99. shell. Procedia Chem. 16, 365–371.
Ghaly, A., Dave, D., Brooks, M., Budge, S., 2010. Production of biodiesel by enzymatic Hirkude, J., Padalkar, A.S., 2014. Experimental investigation of the effect of compression
transesterification. Am. J. Biochem. Biotechnol. 6, 54–76. ratio on performance and emissions of CI engine operated with waste fried oil methyl
Girelli, A.M., Astolfi, M.L., Scuto, F.R., 2020. Agro-industrial wastes as potential carriers ester blend. Fuel Process. Technol. 128, 367–375.
for enzyme immobilization: a review. Chemosphere 244, 125368. Hossain, S.Z., Razzak, S.A., Al Shater, A.F., Moniruzzaman, M., Hossain, M.M., 2020.
Gohain, M., Laskar, K., Paul, A.K., Daimary, N., Maharana, M., Goswami, I.K., Recent Advances in Enzymatic Conversion of Microalgal Lipids into Biodiesel.
Hazarika, A., Bora, U., Deka, D., 2020. Carica papaya stem: a source of versatile Energy & Fuels.
heterogeneous catalyst for biodiesel production and C–C bond formation. Renew. Hosseini, S.A., 2022. Nanocatalysts for biodiesel production. Arab. J. Chem., 104152
Energy 147, 541–555. Hu, S., Guan, Y., Wang, Y., Han, H., 2011. Nano-magnetic catalyst KF/CaO–Fe3O4 for
Gomes Filho, J., Peiter, A., Pimentel, W., Soletti, J., Carvalho, S., Meili, L., 2015. biodiesel production. Appl. Energy 88, 2685–2690.
Biodiesel production from Sterculia striata oil by ethyl transesterification method. Huber, M.L., Lemmon, E.W., Kazakov, A., Ott, L.S., Bruno, T.J., 2009. Model for the
Ind. Crop. Prod. 74, 767–772. thermodynamic properties of a biodiesel fuel. Energy Fuel. 23, 3790–3797.
Gomiero, T., 2015. Are biofuels an effective and viable energy strategy for industrialized Ibrahim, M.L., Khalil, N.N.A.N.A., Islam, A., Rashid, U., Ibrahim, S.F., Mashuri, S.I.S.,
societies? A reasoned overview of potentials and limits. Sustainability 7, 8491–8521. Taufiq-Yap, Y.H., 2020. Preparation of Na2O supported CNTs nanocatalyst for
efficient biodiesel production from waste-oil. Energy Convers. Manag. 205, 112445.

33
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Idris, N.A., Lau, H.L.N., Wafti, N.S.A., Mustaffa, N.K., Loh, S.K., 2020. Glycerolysis of Knothe, G., 2010. Biodiesel and renewable diesel: a comparison. Prog. Energy Combust.
palm fatty acid distillate (PFAD) as biodiesel feedstock using heterogeneous catalyst. Sci. 36, 364–373.
Waste Biomass Valor. 12, 735–744. Knothe, G., Razon, L.F., 2017. Biodiesel fuels. Prog. Energy Combust. Sci. 58, 36–59.
IEA, 2016. World Energy Outlook Special Report. International Energy Agency, France. Konwar, L.J., Boro, J., Deka, D., 2014. Review on latest developments in biodiesel
Illanes, A., Fajardo, A., 2001. Kinetically controlled synthesis of ampicillin with production using carbon-based catalysts. Renew. Sustain. Energy Rev. 29, 546–564.
immobilized penicillin acylase in the presence of organic cosolvents. J. Mol. Catal. B Köse, Ö., Tüter, M., Aksoy, H.A., 2002. Immobilized Candida Antarctica lipase-catalyzed
Enzym. 11, 587–595. alcoholysis of cotton seed oil in a solvent-free medium. Bioresour. Technol. 83,
Ingle, A.P., Chandel, A.K., Philippini, R., Martiniano, S.E., da Silva, S.S., 2020. Advances 125–129.
in nanocatalysts mediated biodiesel production: a critical appraisal. Symmetry 12, Krishnamurthy, K., Sridhara, S., Kumar, C.A., 2020. Optimization and kinetic study of
256. biodiesel production from Hydnocarpus wightiana oil and dairy waste scum using
Iqbal, M.M.A., Toemen, S., Razak, F.I.A., Rosid, S.J.M., Azelee, N.I.W., 2020. Catalytic snail shell CaO nano catalyst. Renew. Energy 146, 280–296.
methanation over nanoparticle heterostructure of Ru/Fe/Ce/γ-Al2O3 catalyst: Kulkarni, R.M., Britto, P.J., Narula, A., Saqline, S., Anand, D., Bhagyalakshmi, C.,
performance and characterisation. Renew. Energy 162, 513–524. Herle, R.N., 2020. Kinetic studies on the synthesis of fuel additives from glycerol
Islam, A., Taufiq-Yap, Y.H., Chu, C.-M., Chan, E.-S., Ravindra, P., 2012. Synthesis and using CeO2–ZrO2 metal oxide catalyst. Biofuel Res. J. 7, 1100–1108.
characterization of millimetric gamma alumina spherical particles by oil drop Kumar, D., Ali, A., 2013. Transesterification of low-quality triglycerides over a Zn/CaO
granulation method. J. Porous Mater. 19, 807–817. heterogeneous catalyst: kinetics and reusability studies. Energy Fuel. 27,
Jambulingam, R., Shalma, M., Shankar, V., 2019. Biodiesel production using lipase 3758–3768.
immobilised functionalized magnetic nanocatalyst from oleaginous fungal lipid. Kumar, B.R., Saravanan, S., 2016. Use of higher alcohol biofuels in diesel engines: a
J. Clean. Prod. 215, 245–258. review. Renew. Sustain. Energy Rev. 60, 84–115.
Jayaraman, J., Alagu, K., Appavu, P., Joy, N., Jayaram, P., Mariadoss, A., 2020. Kumar, V., Thakur, I.S., 2020. Biodiesel Production from Transesterification of Serratia
Enzymatic production of biodiesel using lipase catalyst and testing of an unmodified Sp. ISTD04 Lipids Using Immobilised Lipase on Biocomposite Materials of
compression ignition engine using its blends with diesel. Renew. Energy 145, Biomineralized Products of Carbon Dioxide Sequestrating Bacterium. Bioresource
399–407. Technology, 123193.
Jegannathan, K.R., Abang, S., Poncelet, D., Chan, E.S., Ravindra, P., 2008. Production of Kumar, A., Kumar, K., Kaushik, N., Sharma, S., Mishra, S., 2010. Renewable energy in
biodiesel using immobilized lipase—a critical review. Crit. Rev. Biotechnol. 28, India: current status and future potentials. Renew. Sustain. Energy Rev. 14,
253–264. 2434–2442.
Jiang, C., Jia, J., Zhai, S., 2014. Mechanistic understanding of toxicity from Kumari, V., Shah, S., Gupta, M.N., 2007. Preparation of biodiesel by lipase-catalyzed
nanocatalysts. Int. J. Mol. Sci. 15, 13967–13992. transesterification of high free fatty acid containing oil from Madhuca indica. Energy
Johari, A., Nyakuma, B.B., Nor, S.H.M., Mat, R., Hashim, H., Ahmad, A., Zakaria, Z.Y., Fuel. 21, 368–372.
Abdullah, T.A.T., 2015. The challenges and prospects of palm oil based biodiesel in Kumari, A., Mahapatra, P., Garlapati, V.K., Banerjee, R., 2009. Enzymatic
Malaysia. Energy 81, 255–261. transesterification of Jatropha oil. Biotechnol. Biofuels 2, 1.
Joshi, H., Moser, B.R., Toler, J., Walker, T., 2010. Preparation and fuel properties of Lai, J.-Q., Hu, Z.-L., Wang, P.-W., Yang, Z., 2012. Enzymatic production of microalgal
mixtures of soybean oil methyl and ethyl esters. Biomass Bioenergy 34, 14–20. biodiesel in ionic liquid [BMIm][PF6]. Fuel 95, 329–333.
Jumiarni, D., Anggraini, N., 2021. Characterization of microalgae from lowlands in Lakshmi, S.B.A.V.S., Pillai, N.S., Mohamed, M.S.B.K., Narayanan, A., 2020. Biodiesel
South Sumatera (Indonesia) as a potential source for biodiesel production. J. Phys.: production from rubber seed oil using calcined eggshells impregnated with Al 2 O 3
Conf. Ser. IOP Publ., 012002 as heterogeneous catalyst: a comparative study of RSM and ANN optimization. Braz.
Kalavathy, G., Baskar, G., 2019. Synergism of clay with zinc oxide as nanocatalyst for J. Chem. Eng. 1–18.
production of biodiesel from marine Ulva lactuca. Bioresour. Technol. 281, 234–238. Lam, M.K., Lee, K.T., Mohamed, A.R., 2009. Life cycle assessment for the production of
Karmakar, B., Samanta, S., Halder, G., 2020. Delonix regia heterogeneous catalyzed two- biodiesel: a case study in Malaysia for palm oil versus jatropha oil. Biofuels,
step biodiesel production from Pongamia pinnata oil using methanol and 2- Bioproducts and Biorefining: Innov. Sustain. Econ. 3, 601–612.
propanol. J. Clean. Prod. 255, 120313. Lam, M.K., Lee, K.T., Mohamed, A.R., 2010. Homogeneous, heterogeneous and
Karp, S.G., Medina, J.D., Letti, L.A., Woiciechowski, A.L., de Carvalho, J.C., Schmitt, C. enzymatic catalysis for transesterification of high free fatty acid oil (waste cooking
C., de Oliveira Penha, R., Kumlehn, G.S., Soccol, C.R., 2021. Bioeconomy and oil) to biodiesel: a review. Biotechnol. Adv. 28, 500–518.
biofuels: the case of sugarcane ethanol in Brazil. Biofuels, Bioproducts and Laskar, I.B., Rokhum, L., Gupta, R., Chatterjee, S., 2020. Zinc oxide supported silver
Biorefining 15, 899–912. nanoparticles as a heterogeneous catalyst for production of biodiesel from palm oil.
Katchalski-Katzir, E., 1993. Immobilized enzymes—learning from past successes and Environ. Prog. Sustain. Energy 39, e13369.
failures. Trends Biotechnol. 11, 471–478. Law, S.Q., Halim, R., Scales, P.J., Martin, G.J., 2018. Conversion and recovery of
Kaur, M., Ali, A., 2011. Lithium ion impregnated calcium oxide as nano catalyst for the saponifiable lipids from microalgae using a nonpolar solvent via lipase-assisted
biodiesel production from karanja and jatropha oils. Renew. Energy 36, 2866–2871. extraction. Bioresour. Technol. 260, 338–347.
Kaur, N., Ali, A., 2015. Biodiesel production via ethanolysis of jatropha oil using Lee, T.S., Vaghjiani, J.D., Lye, G.J., Turner, M.K., 2000. A systematic approach to the
molybdenum impregnated calcium oxide as solid catalyst. RSC Adv. 5, large-scale production of protein crystals. Enzym. Microb. Technol. 26, 582–592.
13285–13295. Lee, O.K., Kim, Y.H., Na, J.-G., Oh, Y.-K., Lee, E.Y., 2013. Highly efficient extraction and
Kaur, M., Malhotra, R., Ali, A., 2018. Tungsten supported Ti/SiO2 nanoflowers as lipase-catalyzed transesterification of triglycerides from Chlorella sp. KR-1 for
reusable heterogeneous catalyst for biodiesel production. Renew. Energy 116, production of biodiesel. Bioresour. Technol. 147, 240–245.
109–119. Lee, J.-C., Lee, B., Ok, Y.S., Lim, H., 2020. Preliminary Techno-Economic Analysis of
Kawakami, K., Oda, Y., Takahashi, R., 2011. Application of a Burkholderia cepacialipase- Biodiesel Production over Solid-Biochar. Bioresource technology, 123086.
immobilized silica monolith to batch and continuous biodiesel production with a Leung, D.Y., Wu, X., Leung, M., 2010. A review on biodiesel production using catalyzed
stoichiometric mixture of methanol and crude Jatropha oil. Biotechnol. Biofuels 4, transesterification. Appl. Energy 87, 1083–1095.
42. Li, Q., Yan, Y., 2010. Production of biodiesel catalyzed by immobilized Pseudomonas
Kazemifard, S., Nayebzadeh, H., Saghatoleslami, N., Safakish, E., 2018. Assessment the cepacia lipase from Sapium sebiferum oil in micro-aqueous phase. Appl. Energy 87,
activity of magnetic KOH/Fe 3 O 4@ Al 2 O 3 core–shell nanocatalyst in 3148–3154.
transesterification reaction: effect of Fe/Al ratio on structural and performance. Li, X., Xu, H., Wu, Q., 2007. Large-scale biodiesel production from microalga Chlorella
Environ. Sci. Pollut. Control Ser. 25, 32811–32821. protothecoides through heterotrophic cultivation in bioreactors. Biotechnol. Bioeng.
Keihani, M., Esmaeili, H., Rouhi, P., 2018. Biodiesel production from chicken fat using 98, 764–771.
nano-calcium oxide catalyst and improving the fuel properties via blending with Li, N.-W., Zong, M.-H., Wu, H., 2009. Highly efficient transformation of waste oil to
diesel. Phys. Chem. Res. 6, 521–529. biodiesel by immobilized lipase from Penicillium expansum. Process Biochem. 44,
Khan, I., Saeed, K., Khan, I., 2019. Nanoparticles: properties, applications and toxicities. 685–688.
Arab. J. Chem. 12, 908–931. Li, S.-F., Fan, Y.-H., Hu, J.-F., Huang, Y.-S., Wu, W.-T., 2011a. Immobilization of
Khatiwada, D., Palmén, C., Silveira, S., 2021. Evaluating the palm oil demand in Pseudomonas cepacia lipase onto the electrospun PAN nanofibrous membranes for
Indonesia: production trends, yields, and emerging issues. Biofuels 12, 135–147. transesterification reaction. J. Mol. Catal. B Enzym. 73, 98–103.
Kilpin, K.J., Dyson, P.J., 2013. Enzyme inhibition by metal complexes: concepts, Li, S.-F., Fan, Y.-H., Hu, R.-F., Wu, W.-T., 2011b. Pseudomonas cepacia lipase
strategies and applications. Chem. Sci. 4. immobilized onto the electrospun PAN nanofibrous membranes for biodiesel
Kim, H.-J., Kang, B.-S., Kim, M.-J., Park, Y.M., Kim, D.-K., Lee, J.-S., Lee, K.-Y., 2004. production from soybean oil. J. Mol. Catal. B Enzym. 72, 40–45.
Transesterification of vegetable oil to biodiesel using heterogeneous base catalyst. Li, H., Lv, P., Wang, Z., Miao, C., Yuan, Z., 2020a. Biodiesel Continuous Esterification
Catal. Today 93, 315–320. Process Experimental Study and Equipment Design. Biomass Conversion and
Kim, J., Grate, J.W., Wang, P., 2008. Nanobiocatalysis and its Potential Applications. Cell Biorefinery, pp. 1–8.
Press, pp. 639–646. Li, J., Zhang, J., Shen, S., Zhang, B., William, W.Y., 2020b. Magnetic responsive
Kim, M., DiMaggio, C., Yan, S., Salley, S., Ng, K., 2010. The synergistic effect of alcohol Thermomyces lanuginosus lipase for biodiesel synthesis. Mater. Today Commun. 24,
mixtures on transesterification of soybean oil using homogeneous and heterogeneous 101197.
catalysts. Appl. Catal. Gen. 378, 134–143. Lim, S., Yap, C.Y., Pang, Y.L., Wong, K.H., 2019. Biodiesel synthesis from oil palm empty
Klein, B.C., Chagas, M.F., Watanabe, M.D.B., Bonomi, A., Maciel Filho, R., 2019. Low fruit bunch biochar derived heterogeneous solid catalyst using 4-benzenediazonium
carbon biofuels and the New Brazilian National Biofuel Policy (RenovaBio): a case sulfonate. J. Hazard Mater., 121532
study for sugarcane mills and integrated sugarcane-microalgae biorefineries. Renew. Lima, L.N., Oliveira, G.C., Rojas, M.J., Castro, H.F., Da Rós, P.C., Mendes, A.A.,
Sustain. Energy Rev. 115, 109365. Giordano, R.L., Tardioli, P.W., 2015. Immobilization of Pseudomonas fluorescens
Knezevic, Z., Siler-Marinkovic, S., Mojovic, L., 2004. Immobilized lipases as practical lipase on hydrophobic supports and application in biodiesel synthesis by
catalysts. Acta Period. Technol. 151–164.

34
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

transesterification of vegetable oils in solvent-free systems. J. Ind. Microbiol. Mateo, C., Palomo, J.M., Fernandez-Lorente, G., Guisan, J.M., Fernandez-Lafuente, R.,
Biotechnol. 42, 523–535. 2007. Improvement of enzyme activity, stability and selectivity via immobilization
Lima, M., Mendes, L., Mothé, G., Linhares, F., de Castro, M., Da Silva, M., Sthel, M., 2020. techniques. Enzym. Microb. Technol. 40, 1451–1463.
Renewable energy in reducing greenhouse gas emissions: reaching the goals of the Math, M., Irfan, G., 2007. Optimization of Restaurant Waste Oil Methyl Ester Yield.
Paris agreement in Brazil. Environ. Dev. 33, 100504. Maulidiyah, Nurdin, M., Fatma, F., Natsir, M., Wibowo, D., 2017. Characterization of
Ling, J.S.J., Tan, Y.H., Mubarak, N.M., Kansedo, J., Saptoro, A., Nolasco-Hipolito, C., methyl ester compound of biodiesel from industrial liquid waste of crude palm oil
2019. A review of heterogeneous calcium oxide based catalyst from waste for processing. Anal. Chem. Res. 12, 1–9.
biodiesel synthesis. SN Appl. Sci. 1. Mazumdar, P., Dasari, S.R., Borugadda, V.B., Srivasatava, G., Sahoo, L., Goud, V.V.,
Liu, Y., Wu, H., Yan, Y., Dong, L., Zhu, M., Liang, P., 2013. Lipase-catalyzed 2013. Biodiesel production from high free fatty acids content Jatropha curcas L. oil
transesterification for biodiesel production in ionic liquid [Emim] Tfo. Int. J. Green using dual step process. Biomass Conv. Biorefinery 3, 361–369.
Energy 10, 63–71. Mehra, S., 2012. Physiochemical properties characteristics of Jatropha Curcas biodiesel
Long, Y.-D., Fang, Z., Su, T.-C., Yang, Q., 2014. Co-production of biodiesel and hydrogen and its blends with tetrachloroethylene addition. Int. J. Chem. Res. 30–33.
from rapeseed and Jatropha oils with sodium silicate and Ni catalysts. Appl. Energy Mehrasbi, M.R., Mohammadi, J., Peyda, M., Mohammadi, M., 2017. Covalent
113, 1819–1825. immobilization of Candida Antarctica lipase on core-shell magnetic nanoparticles for
López, E.N., Medina, A.R., Moreno, P.A.G., Cerdán, L.E., Valverde, L.M., Grima, E.M., production of biodiesel from waste cooking oil. Renew. Energy 101, 593–602.
2016. Biodiesel production from Nannochloropsis gaditana lipids through Melani, N.B., Tambourgi, E.B., Silveira, E., 2020. Lipases: from production to
transesterification catalyzed by Rhizopus oryzae lipase. Bioresour. Technol. 203, applications. Separ. Purif. Rev. 49, 143–158.
236–244. Mguni, L.L., Meijboom, R., Jalama, K., 2012. Biodiesel Production over Nano-MgO
Lotero, E., Liu, Y., Lopez, D.E., Suwannakarn, K., Bruce, D.A., Goodwin, J.G., 2005. Supported on Titania.
Synthesis of biodiesel via acid catalysis. Ind. Eng. Chem. Res. 44, 5353–5363. Mihajlovski, K., Buntić, A., Milić, M., Rajilić-Stojanović, M., Dimitrijević-Branković, S.,
Lotero, E., Goodwin Jr., J.G., Bruce, D.A., Suwannakarn, K., Liu, Y., Lopez, D.E., 2006. 2020. From agricultural waste to biofuel: enzymatic potential of a bacterial isolate
The catalysis of biodiesel synthesis. Catalysis 19, 41–83. streptomyces fulvissimus CKS7 for bioethanol production. Waste Biomass Valor.
Lozano, P., Bernal, J.M., Piamtongkam, R., Fetzer, D., Vaultier, M., 2010. One-phase 1–10.
ionic liquid reaction medium for biocatalytic production of biodiesel. ChemSusChem Modi, M.K., Reddy, J., Rao, B., Prasad, R., 2007. Lipase-mediated conversion of
3, 1359–1363. vegetable oils into biodiesel using ethyl acetate as acyl acceptor. Bioresour. Technol.
Lu, J., Chen, Y., Wang, F., Tan, T., 2009. Effect of water on methanolysis of glycerol 98, 1260–1264.
trioleate catalyzed by immobilized lipase Candida sp. 99–125 in organic solvent Mohd Noor, N.A., Nurdin, S., Yaakob, Z., Mahmud, M.S., 2019. Intensification of
system. J. Mol. Catal. B Enzym. 56, 122–125. biodiesel synthesis reactor using biphasic homogenous catalytic reaction: parametric
Luz Martinez, S., Romero, R., López, J.C., Romero, A., Sanchez Mendieta, V., study. Ind. Eng. Chem. Res. 58, 9173–9178.
Natividad, R., 2011. Preparation and characterization of CaO nanoparticles/NaX Mohebbi, S., Rostamizadeh, M., Kahforoushan, D., 2020. Effect of molybdenum promoter
zeolite catalysts for the transesterification of sunflower oil. Ind. Eng. Chem. Res. 50, on performance of high silica MoO3/B-ZSM-5 nanocatalyst in biodiesel production.
2665–2670. Fuel 266, 117063.
Lv, L., Dai, L., Du, W., Liu, D., 2021. Progress in enzymatic biodiesel production and Muanruksa, P., Kaewkannetra, P., 2020. Combination of fatty acids extraction and
commercialization. Processes 9, 355. enzymatic esterification for biodiesel production using sludge palm oil as a low-cost
Ma, Y., Wang, Q., Zheng, L., Gao, Z., Wang, Q., Ma, Y., 2016. Mixed methanol/ethanol on substrate. Renew. Energy 146, 901–906.
transesterification of waste cooking oil using Mg/Al hydrotalcite catalyst. Energy Muanruksa, P., Dujjanutat, P., Kaewkannetra, P., 2020. Entrapping immobilisation of
107, 523–531. lipase on biocomposite hydrogels toward for biodiesel production from waste frying
Madhuvilakku, R., Piraman, S., 2013. Biodiesel synthesis by TiO2–ZnO mixed oxide acid oil. Catalysts 10, 834.
nanocatalyst catalyzed palm oil transesterification process. Bioresour. Technol. 150, Mueanmas, C., Nikhom, R., Petchkaew, A., Iewkittayakorn, J., Prasertsit, K., 2019.
55–59. Extraction and esterification of waste coffee grounds oil as non-edible feedstock for
Mahajan, S., Konar, S.K., Boocock, D.G.b., 2006. Determining the acid number of biodiesel production. Renew. Energy 133, 1414–1425.
biodiesel. J. Am. Oil Chem. Soc. 83, 567–570. Mulalee, S., 2020. Continuous and circulated batch processes for esterification of free
Mahdavi, M., Abedini, E., hosein Darabi, A., 2015. Biodiesel synthesis from oleic acid by fatty acids by novozym 435 in expanded bed reactor: a case study of palm fatty acid
nano-catalyst (ZrO2/Al2O3) under high voltage conditions. RSC Adv. 5, distillate. J. Oil Palm Res.
55027–55032. Musa, I.A., 2016. The effects of alcohol to oil molar ratios and the type of alcohol on
Mahdi, H.I., Muraza, O., 2016. Conversion of isobutylene to octane-booster compounds biodiesel production using transesterification process. Egypt. J. Petrol. 25, 21–31.
after methyl tert-butyl ether phaseout: the role of heterogeneous catalysis. Ind. Eng. Naderi, F., Nayebzadeh, H., 2019. Performance and stability assessment of Mg-Al-Fe
Chem. Res. 55, 11193–11210. nanocatalyst in the transesterification of sunflower oil: effect of Al/Fe molar ratio.
Mahdi, H.I., Muraza, O., 2019. An exciting opportunity for zeolite adsorbent design in Ind. Crop. Prod. 141, 111814.
separation of C4 olefins through adsorptive separation. Separ. Purif. Technol. 221, Najeeb, J., Akram, S., Mumtaz, M.W., Danish, M., Irfan, A., Touqeer, T., Rashid, U.,
126–151. Ghani, W.A.W.A.K., Choong, T.S.Y., 2021. Nanobiocatalysts for biodiesel synthesis
Mahdi, H.I., Irawan, E., Nuryoto, N., Jayanudin, J., Sulistyo, H., Sediawan, W.B., through transesterification—a review. Catalysts 11, 171.
Muraza, O., 2016. Glycerol carbonate production from biodiesel waste over modified Naveenkumar, R., Baskar, G., 2019. Biodiesel production from Calophyllum inophyllum
natural clinoptilolite. Waste Biomass Valor. 7, 1349–1356. oil using zinc doped calcium oxide (Plaster of Paris) nanocatalyst. Bioresour.
Mahdi, H.I., Bazargan, A., McKay, G., Azelee, N.I.W., Meili, L., 2021. Catalytic Technol. 280, 493–496.
deoxygenation of palm oil and its residue in green diesel production: a current Nayak, P.K., Dash, U., Rayaguru, K., Krishnan, K.R., 2016. Physio-chemical changes
technological review. Chem. Eng. Res. Des. 174, 158–187. during repeated frying of cooked oil: a Review. J. Food Biochem. 40, 371–390.
Mahmoud, H.R., 2019. Bismuth silicate (Bi4Si3O12 and Bi2SiO5) prepared by ultrasonic- Nayak, S.N., Bhasin, C.P., Nayak, M.G., 2019. A review on microwave-assisted
assisted hydrothermal method as novel catalysts for biodiesel production via oleic transesterification processes using various catalytic and non-catalytic systems.
acid esterification with methanol. Fuel 256, 115979. Renew. Energy 143, 1366–1387.
Mak, K.-H., Yu, C.-Y., Kuan, I.-C., Lee, S.-L., 2009. Immobilization of Pseudomonas Nayebzadeh, H., Saghatoleslami, N., Tabasizadeh, M., 2016. Optimization of the activity
cepacia lipase onto magnetic nanoparticles for biodiesel production. Sci. Technol. of KOH/calcium aluminate nanocatalyst for biodiesel production using response
Vision 5, 19–23. surface methodology. J. Taiwan Inst. Chem. Eng. 68, 379–386.
Malakar, B., Das, D., Mohanty, K., 2020. Optimization of glucose yield from potato and Nematian, T., Salehi, Z., Shakeri, A., 2020a. Conversion of bio-oil extracted from
sweet lime peel waste through different pre-treatment techniques along with enzyme Chlorella vulgaris micro algae to biodiesel via modified superparamagnetic nano-
assisted hydrolysis towards liquid biofuel. Renew. Energy 145, 2723–2732. biocatalyst. Renew. Energy 146, 1796–1804.
Mansir, N., Taufiq-Yap, Y.H., Rashid, U., Lokman, I.M., 2017. Investigation of Nematian, T., Shakeri, A., Salehi, Z., Saboury, A.A., 2020b. Lipase immobilized on
heterogeneous solid acid catalyst performance on low grade feedstocks for biodiesel functionalized superparamagnetic few-layer graphene oxide as an efficient
production: a review. Energy Convers. Manag. 141, 171–182. nanobiocatalyst for biodiesel production from Chlorella vulgaris bio-oil. Biotechnol.
Mansir, N., Teo, S.H., Rashid, U., Taufiq-Yap, Y.H., 2018. Efficient waste Gallus Biofuels 13, 57.
domesticus shell derived calcium-based catalyst for biodiesel production. Fuel 211, Nezhad, M.K., Aghaei, H., 2020. Tosylated Cloisite as a New Heterofunctional Carrier for
67–75. Covalent Immobilization of Lipase and its Utilization for Production of Biodiesel
Mardhiah, H.H., Ong, H.C., Masjuki, H., Lim, S., Pang, Y.L., 2017. Investigation of from Waste Frying Oil. Renewable Energy.
carbon-based solid acid catalyst from Jatropha curcas biomass in biodiesel Ngoie, W.I., Oyekola, O.O., Ikhu-Omoregbe, D., Welz, P.J., 2020. Valorisation of edible
production. Energy Convers. Manag. 144, 10–17. oil wastewater sludge: bioethanol and biodiesel production. Waste Biomass Valor.
Marín-Suárez, M., Méndez-Mateos, D., Guadix, A., Guadix, E.M., 2019. Reuse of 11, 2431–2440.
immobilized lipases in the transesterification of waste fish oil for the production of Nielsen, P., Rancke Madsen, A., Holm, H., Burton, R., 2016. Production of biodiesel using
biodiesel. Renew. Energy 140, 1–8. liquid lipase formulations. J. Am. Oil Chem. Soc. 93, 905–910.
Maruyama, T., Nakajima, M., Uchikawa, S., Nabetani, H., Furusaki, S., Seki, M., 2000. Nikhom, R., Tongurai, C., 2014. Production development of ethyl ester biodiesel from
Oil-water interfacial activation of lipase for interesterification of triglyceride and palm oil using a continuous deglycerolisation process. Fuel 117, 926–931.
fatty acid. J. Am. Oil Chem. Soc. 77, 1121. Nogueira Jr., C.A., Feitosa, F.X., Fernandes, F.A., Santiago, R.S., de Sant’Ana, H.B., 2010.
Matassoli, A.L., Corrêa, I.N., Portilho, M.F., Veloso, C.O., Langone, M.A., 2009. Densities and viscosities of binary mixtures of babassu biodiesel+ cotton seed or
Enzymatic synthesis of biodiesel via alcoholysis of palm oil. Appl. Biochem. soybean biodiesel at different temperatures. J. Chem. Eng. Data 55, 5305–5310.
Biotechnol. 155, 44–52. Noureddini, H., Gao, X., Philkana, R., 2005. Immobilized Pseudomonas cepacia lipase for
biodiesel fuel production from soybean oil. Bioresour. Technol. 96, 769–777.

35
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Nusterer, E., Blöchl, P., Schwarz, K., 1996. Interaction of water and methanol with a Ramlee, N.N., Mahdi, H.I., Azelee, N.I.W., 2022. Adsorption immobilization:
zeolite at high coverages. Chem. Phys. Lett. 253, 448–455. mechanisms and interactions between support and lipase. In: Rahman, R.A.,
OECD/FAO, 2016. Biofuels OECD/FAO-Agricultural Outlook 2016-2025 (Paris). Shaarani, S.M. (Eds.), Enzyme Immobilization for Bioprocessing. PENERBIT UTM
OECD/FAO, 2019. Biofuels. OECD-FAO Agricultural Outlook 2019-2028. OECD-FAO. PRESS, UTM, Johor, pp. 39–69.
Osorio-González, C.S., Gómez-Falcon, N., Sandoval-Salas, F., Saini, R., Brar, S.K., Ramos, M.D.N., Milessi, T.S., Candido, R.G., Mendes, A.A., Aguiar, A., 2022. Enzymatic
Ramírez, A.A., 2020. Production of biodiesel from Castor oil: a review. Energies 13, catalysis as a tool in biofuels production in Brazil: current status and perspectives.
2467. Energy Sustain. Dev. 68, 103–119.
Outili, N., Kerras, H., Nekkab, C., Merouani, R., Meniai, A.H., 2020. Biodiesel production Rana, Q.U.A., Irfan, M., Ahmed, S., Hasan, F., Shah, A.A., Khan, S., Rehman, F.U.,
optimization from waste cooking oil using green chemistry metrics. Renew. Energy Khan, H., Ju, M., Li, W., 2019. Bio-catalytic transesterification of mustard oil for
145, 2575–2586. biodiesel production. Biofuels 1–8.
Pan, H., Li, H., Liu, X.-F., Zhang, H., Yang, K.-L., Huang, S., Yang, S., 2016. Mesoporous Ranganathan, S.V., Narasimhan, S.L., Muthukumar, K., 2008. An overview of enzymatic
polymeric solid acid as efficient catalyst for (trans) esterification of crude Jatropha production of biodiesel. Bioresour. Technol. 99, 3975–3981.
curcas oil. Fuel Process. Technol. 150, 50–57. Rashtizadeh, E., Farzaneh, F., Talebpour, Z., 2014. Synthesis and characterization of
Park, Y.-M., Lee, D.-W., Kim, D.-K., Lee, J.-S., Lee, K.-Y., 2008. The heterogeneous Sr3Al2O6 nanocomposite as catalyst for biodiesel production. Bioresour. Technol.
catalyst system for the continuous conversion of free fatty acids in used vegetable 154, 32–37.
oils for the production of biodiesel. Catal. Today 131, 238–243. Rathod, V.K., Pandit, A.B., 2010. Enzymatic hydrolysis of oil in a spray column. J. Mol.
Pastore, C., Lopez, A., Mascolo, G., 2014. Efficient conversion of brown grease produced Catal. B Enzym. 67, 1–9.
by municipal wastewater treatment plant into biofuel using aluminium chloride Rattanaphra, D., Harvey, A., Srinophakun, P., 2010. Simultaneous conversion of
hexahydrate under very mild conditions. Bioresour. Technol. 155, 91–97. triglyceride/free fatty acid mixtures into biodiesel using sulfated zirconia. Top.
Pasupulety, N., Gunda, K., Liu, Y., Rempel, G.L., Ng, F.T., 2013. Production of biodiesel Catal. 53, 773–782.
from soybean oil on CaO/Al2O3 solid base catalysts. Appl. Catal. Gen. 452, 189–202. Ravi, A., Gurunathan, B., Rajendiran, N., Varjani, S., Gnansounou, E., Pandey, A.,
Patchimpet, J., Simpson, B.K., Sangkharak, K., Klomklao, S., 2020. Optimization of You, S., Raman, J.K., Ramanujam, P., 2020. Contemporary Approaches towards
process variables for the production of biodiesel by transesterification of used Augmentation of Distinctive Heterogeneous Catalyst for Sustainable Biodiesel
cooking oil using lipase from Nile tilapia viscera. Renew. Energy 153, 861–869. Production. Environmental Technology & Innovation, 100906.
Paul Abishek, M., Patel, J., Prem Rajan, A., 2014. Algae oil: a sustainable renewable fuel Rebelo, L.P., Netto, C.G., Toma, H.E., Andrade, L.H., 2010. Enzymatic kinetic resolution
of future. Biotechnol. Res. Int. 2014. of (RS)-1-(Phenyl) ethanols by Burkholderia cepacia lipase immobilized on magnetic
Peer, M.S., Kasimani, R., Rajamohan, S., Ramakrishnan, P., 2017. Experimental nanoparticles. J. Braz. Chem. Soc. 21, 1537–1542.
evaluation on oxidation stability of biodiesel/diesel blends with alcohol addition by Rial, R.C., de Freitas, O.N., Nazário, C.E.D., Viana, L.H., 2020. Biodiesel from soybean oil
rancimat instrument and FTIR spectroscopy. J. Mech. Sci. Technol. 31, 455–463. using Porcine pancreas lipase immobilized on a new support: p-nitrobenzyl cellulose
Peiter, A.S., Lins, P.V., Meili, L., Soletti, J.I., Carvalho, S.H., Pimentel, W.R., xanthate. Renew. Energy 149, 970–979.
Meneghetti, S.M., 2020. Stirring and mixing in ethylic biodiesel production. J. King Rios, N.S., Neto, D.M.A., Dos Santos, J.C.S., Fechine, P.B.A., Fernández-Lafuente, R.,
Saud Univ. Sci. 32, 54–59. Gonçalves, L.R.B., 2019. Comparison of the immobilization of lipase from
Perera, F., 2017. Pollution from fossil-fuel combustion is the leading environmental Pseudomonas fluorescens on divinylsulfone or p-benzoquinone activated support.
threat to global pediatric health and equity: solutions exist. Int. J. Environ. Res. Publ. Int. J. Biol. Macromol. 134, 936–945.
Health 15. Rocha, T.G., Pedro, H.d.L., de Souza, M.C., Monteiro, R.R., dos Santos, J.C., 2020. Lipase
Pérez, A., Ramos, R., Montero, G., García, C., Coronado, M., Campbell, H., Delgado, R., cocktail for optimized biodiesel production of free fatty acids from residual chicken
Suástegui, A., 2018. Development and implementation of virtual instrumentation for oil. Catal. Lett. 1–12.
the measurement of operating parameters of an engine using diesel-biodiesel Rodrigues, R.C., Pessela, B.C., Volpato, G., Fernandez-Lafuente, R., Guisan, J.M.,
mixtures. Biofuels-Chall. Opportun. IntechOpen. Ayub, M.A., 2010. Two step ethanolysis: a simple and efficient way to improve the
Pesaresi, L., Brown, D., Lee, A.F., Montero, J., Williams, H., Wilson, K., 2009. Cs-doped enzymatic biodiesel synthesis catalyzed by an immobilized–stabilized lipase from
H4SiW12O40 catalysts for biodiesel applications. Appl. Catal. Gen. 360, 50–58. Thermomyces lanuginosus. Process Biochem. 45, 1268–1273.
Pinheiro, B.B., Rios, N.S., Aguado, E.R., Fernandez-Lafuente, R., Freire, T.M., Fechine, P. Rodrigues, R.C., Ortiz, C., Berenguer-Murcia, Á., Torres, R., Fernández-Lafuente, R.,
B., dos Santos, J.C., Gonçalves, L.R., 2019. Chitosan activated with divinyl sulfone: a 2013. Modifying enzyme activity and selectivity by immobilization. Chem. Soc. Rev.
new heterofunctional support for enzyme immobilization. Application in the 42, 6290–6307.
immobilization of lipase B from Candida Antarctica. Int. J. Biol. Macromol. 130, Romdhane, I.B.-B., Romdhane, Z.B., Bouzid, M., Gargouri, A., Belghith, H., 2013.
798–809. Application of a chitosan-immobilized Talaromyces thermophilus lipase to a batch
Pollardo, A.A., Lee, H.-s., Lee, D., Kim, S., Kim, J., 2018. Solvent effect on the enzymatic biodiesel production from waste frying oils. Appl. Biochem. Biotechnol. 171,
production of biodiesel from waste animal fat. J. Clean. Prod. 185, 382–388. 1986–2002.
Pratas, M.J., Freitas, S.V., Oliveira, M.B., Monteiro, S.C., Lima, Á.S., Coutinho, J.A., Royon, D., Daz, M., Ellenrieder, G., Locatelli, S., 2007. Enzymatic production of biodiesel
2011. Biodiesel density: experimental measurements and prediction models. Energy from cotton seed oil using t-butanol as a solvent. Bioresour. Technol. 98, 648–653.
Fuel. 25, 2333–2340. Rueda, N., Dos Santos, C.S., Rodriguez, M.D., Albuquerque, T.L., Barbosa, O., Torres, R.,
Prinsen, P., Luque, R., 2019. Chapter 1. Introduction to Nanocatalysts. Nanoparticle Ortiz, C., Fernandez-Lafuente, R., 2016. Reversible immobilization of lipases on
Design and Characterization for Catalytic Applications in Sustainable Chemistry, octyl-glutamic agarose beads: a mixed adsorption that reinforces enzyme
pp. 1–36. immobilization. J. Mol. Catal. B Enzym. 128, 10–18.
Putrasari, Y., Praptijanto, A., Santoso, W.B., Lim, O., 2016. Resources, policy, and Safakish, E., Nayebzadeh, H., Saghatoleslami, N., Kazemifard, S., 2020. Comprehensive
research activities of biofuel in Indonesia: a review. Energy Rep. 2, 237–245. assessment of the preparation conditions of a separable magnetic nanocatalyst for
Qiu, F., Li, Y., Yang, D., Li, X., Sun, P., 2011. Heterogeneous solid base nanocatalyst: biodiesel production from algae. Algal Res. 49, 101949.
preparation, characterization and application in biodiesel production. Bioresour. Saha, B.C., Woodward, J., 1997. Fuels and Chemicals from Biomass.
Technol. 102, 4150–4156. Sahani, S., Roy, T., Sharma, Y.C., 2019. Clean and efficient production of biodiesel using
Quah, R.V., Tan, Y.H., Mubarak, N.M., Kansedo, J., Khalid, M., Abdullah, E.C., barium cerate as a heterogeneous catalyst for the biodiesel production; kinetics and
Abdullah, M.O., 2020. Magnetic biochar derived from waste palm kernel shell for thermodynamic study. J. Clean. Prod. 237, 117699.
biodiesel production via sulfonation. Waste Manag. 118, 626–636. Sahota, L., Tiwari, G., 2017. Analytical characteristic equation of nanofluid loaded active
Quayson, E., Amoah, J., Hama, S., Kondo, A., Ogino, C., 2020a. Immobilized lipases for double slope solar still coupled with helically coiled heat exchanger. Energy
biodiesel production: current and future greening opportunities. Renew. Sustain. Convers. Manag. 135, 308–326.
Energy Rev. 134, 110355. Sai, B.A., Subramaniapillai, N., Mohamed, M.S.B.K., Narayanan, A., 2020. Optimization
Quayson, E., Amoah, J., Rachmadona, N., Hama, S., Yoshida, A., Kondo, A., Ogino, C., of continuous biodiesel production from rubber seed oil (RSO) using calcined
2020b. Biodiesel-mediated biodiesel production: a recombinant Fusarium eggshells as heterogeneous catalyst. J. Environ. Chem. Eng. 8, 103603.
heterosporum lipase-catalyzed transesterification of crude plant oils. Fuel Process. Salis, A., Bhattacharyya, M.S., Monduzzi, M., Solinas, V., 2009. Role of the support
Technol. 199, 106278. surface on the loading and the activity of Pseudomonas fluorescens lipase used for
Rad, A.S., Nia, M.H., Ardestani, F., Nayebzadeh, H., 2018. Esterification of waste chicken biodiesel synthesis. J. Mol. Catal. B Enzym. 57, 262–269.
fat: sulfonated MWCNT toward biodiesel production. Waste Biomass Valor. 9, Sani, Y.M., Daud, W.M.A.W., Aziz, A.A., 2014. Activity of solid acid catalysts for
591–599. biodiesel production: a critical review. Appl. Catal. Gen. 470, 140–161.
Rahman, A., Dargusch, P., Wadley, D., 2021. The political economy of oil supply in Santaraite, M., Sendzikiene, E., Makareviciene, V., Kazancev, K., 2020. Biodiesel
Indonesia and the implications for renewable energy development. Renew. Sustain. production by lipase-catalyzed in situ transesterification of rapeseed oil containing a
Energy Rev. 144, 111027. high free fatty acid content with ethanol in diesel fuel media. Energies 13, 2588.
Rahman Talukder, M.M., Min Puah, S., Chuan Wu, J., Jae Won, C., Chow, Y., 2006. Santos, J.C.S.d., Barbosa, O., Ortiz, C., Berenguer-Murcia, A., Rodrigues, R.C., Fernández
Lipase-catalyzed methanolysis of palm oil in presence and absence of organic solvent Lafuente, R., 2015. Importance of the Support Properties for Immobilization or
for production of biodiesel. Biocatal. Biotransform. 24, 257–262. Purification of Enzymes.
Rahmanulloh, A., 2019. Indonesia Biofuels Annual Report 2019. USDA. Santos, S., Puna, J., Gomes, J., 2020. A review on bio-based catalysts (immobilized
Raita, M., Champreda, V., Laosiripojana, N., 2010. Biocatalytic ethanolysis of palm oil enzymes) used for biodiesel production. Energies 13, 3013.
for biodiesel production using microcrystalline lipase in tert-butanol system. Process Saoud, K., 2018. Nanocatalyst for biofuel production: a review. Green Nanotechnol.
Biochem. 45, 829–834. Biofuel Prod. 39–62.
Ramalingam, S., Murugesan, E., Ganesan, P., Rajendiran, S., 2020. Characteristics Sarvestani, N.S., Abbaspour-Fard, M.H., Tabasizadeh, M., Nayebzadeh, H., Van, T.C.,
analysis of julifora biodiesel derived from different production methods. Fuel 280, Jafari, M., Ristovski, Z., Brown, R.J., 2020. Synthesize of magnetite Mg-Fe mixed
118579. metal oxide nanocatalyst by urea-nitrate combustion method with optimal fuel ratio
for reduction of emissions in diesel engines. J. Alloys Compd., 155627

36
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Schmid, R.D., Verger, R., 1998. Lipases: interfacial enzymes with attractive applications. Sitepu, E.K., Heimann, K., Raston, C.L., Zhang, W., 2020. Critical evaluation of process
Angew. Chem. Int. Ed. 37, 1608–1633. parameters for direct biodiesel production from diverse feedstock. Renew. Sustain.
Seffati, K., Esmaeili, H., Honarvar, B., Esfandiari, N., 2020. AC/CuFe2O4@ CaO as a Energy Rev. 123, 109762.
novel nanocatalyst to produce biodiesel from chicken fat. Renew. Energy 147, Siva, S., Marimuthu, C., 2015. Production of biodiesel by transesterification of algae oil
25–34. with an assistance of nano-CaO catalyst derived from egg shell. Int. J. ChemTech
Selvakumar, P., Sivashanmugam, P., 2017. Optimization of lipase production from Res. 7, 2112–2116.
organic solid waste by anaerobic digestion and its application in biodiesel Stamenković, O.S., Veličković, A.V., Veljković, V.B., 2011. The production of biodiesel
production. Fuel Process. Technol. 165, 1–8. from vegetable oils by ethanolysis: current state and perspectives. Fuel 90,
Shah, S., Gupta, M.N., 2007. Lipase catalyzed preparation of biodiesel from Jatropha oil 3141–3155.
in a solvent free system. Process Biochem. 42, 409–414. Stattman, S.L., Mol, A.P., 2014. Social sustainability of Brazilian biodiesel: the role of
Shah, S., Sharma, A., Gupta, M.N., 2006. Preparation of cross-linked enzyme aggregates agricultural cooperatives. Geoforum 54, 282–294.
by using bovine serum albumin as a proteic feeder. Anal. Biochem. 351, 207–213. Straathof, A.J., Panke, S., Schmid, A., 2002. The production of fine chemicals by
Shah, K.A., Parikh, J.K., Maheria, K.C., 2015. Use of sulfonic acid-functionalized silica as biotransformations. Curr. Opin. Biotechnol. 13, 548–556.
catalyst for esterification of free fatty acids (FFA) in acid oil for biodiesel production: Su, E., Wei, D., 2014. Improvement in biodiesel production from soapstock oil by one-
an optimization study. Res. Chem. Intermed. 41, 1035–1051. stage lipase catalyzed methanolysis. Energy Convers. Manag. 88, 60–65.
Shahbazi, M.R., Khoshandam, B., Nasiri, M., Ghazvini, M., 2012. Biodiesel production Subileau, M., Jan, A.H., Drone, J., Rutyna, C., Perrier, V., Dubreucq, E., 2017. What
via alkali-catalyzed transesterification of Malaysian RBD palm oil – characterization, makes a lipase a valuable acyltransferase in water abundant medium? Catal. Sci.
kinetics model. J. Taiwan Inst. Chem. Eng. 43, 504–510. Technol. 7, 2566–2578.
Shahid, E., Jamal, Y., 2008. A reviews as vehicle fuel. Renew. Sustain. Energy Rev. 12, Sugiarto, B., Suryantoro, M., Yubaidah, S., Attharik, M., 2018. The effect of antioxidant
2484. additives on the growth of deposits on the use of biodiesel fuel (B100) at certain
Shalihah, R., Widiarti, N., Suprapto, S., Yanuar, E., Prasetyoko, D., 2020. Modified ZnO temperatures. In: IOP Conference Series: Earth and Environmental Science. IOP
to Indonesia Limestone as heterogeneous catalysts for biodiesel production from Publishing, 012075.
Reutealis trisperma Oil. Malays. J. Fund. Appl. Sci. 16, 649–653. Sukarno, I., Matsumoto, H., Susanti, L., 2016. Transportation energy consumption and
Shameer, P.M., Ramesh, K., 2017. Green technology and performance consequences of emissions-a view from city of Indonesia. Fut. Cities Environ. 2, 1–11.
an eco-friendly substance on a 4-stroke diesel engine at standard injection timing Sulaiman, N.F., Bakar, W.A.W.A., Toemen, S., Kamal, N.M., Nadarajan, R., 2019. In
and compression ratio. J. Mech. Sci. Technol. 31, 1497–1507. depth investigation of bi-functional, Cu/Zn/γ-Al2O3 catalyst in biodiesel production
Shao, G.N., Sheikh, R., Hilonga, A., Lee, J.E., Park, Y.-H., Kim, H.T., 2013. Biodiesel from low-grade cooking oil: optimization using response surface methodology.
production by sulfated mesoporous titania–silica catalysts synthesized by the sol–gel Renew. Energy 135, 408–416.
process from less expensive precursors. Chem. Eng. J. 215, 600–607. Sulihatimarsyila, A.N., Lau, H.L., Nabilah, K., Azreena, I.N., 2019. Refining process for
Sharma, S., Kumar, H., Poonia, M., Jethoo, A., 2012. Combustion and performance production of refined palm-pressed fibre oil. Ind. Crop. Prod. 129, 488–494.
studies of biodiesel fuelled diesel engine: a review. Int. J. Mech. Eng. Appl. Res. 3, Sundus, F., Fazal, M., Masjuki, H., 2017. Tribology with biodiesel: a study on enhancing
128–132. biodiesel stability and its fuel properties. Renew. Sustain. Energy Rev. 70, 399–412.
Sharma, R.K., Saxena, M., O’Neill, C.A., Ramos, H.A., Griebenow, K., 2018. Synthesis of Sutrisno, B., Muhammad, A., Genta, Z., Hidayat, A., 2021. Indonesia natural zeolite as
Rhizopus arrhizus lipase nanoparticles for biodiesel production. ACS Omega 3, heterogeneous catalyst for biodiesel production. Key Eng. Mater. Trans. Tech. Publ.
18203–18213. 91–95.
Shaw, J.-F., Chang, S.-W., Lin, S.-C., Wu, T.-T., Ju, H.-Y., Akoh, C., Chang, R.-H., Taher, H., Nashef, E., Anvar, N., Al-Zuhair, S., 2019. Enzymatic production of biodiesel
Shieh, C.-J., 2008. Continuous enzymatic synthesis of biodiesel with Novozym 435. from waste oil in ionic liquid medium. Biofuels 10, 463–472.
Energy Fuel. 22, 840–844. Taher, H., Giwa, A., Abusabiekeh, H., Al-Zuhair, S., 2020. Biodiesel production from
Sheikh, R., Choi, M.-S., Im, J.-S., Park, Y.-H., 2013. Study on the solid acid catalysts in Nannochloropsis gaditana using supercritical CO2 for lipid extraction and
biodiesel production from high acid value oil. J. Ind. Eng. Chem. 19, 1413–1419. immobilized lipase transesterification: economic and environmental impact
Sheldon, R., Pelt, S., 2013. Van Enzyme immobilisation in biocatalysis: why, what and assessments. Fuel Process. Technol. 198, 106249.
how. Chem. Soc. Rev. 42, 6223–6235. Tahvildari, K., Anaraki, Y.N., Fazaeli, R., Mirpanji, S., Delrish, E., 2015. The study of CaO
Shimada, Y., Watanabe, Y., Sugihara, A., Tominaga, Y., 2002. Enzymatic alcoholysis for and MgO heterogenic nano-catalyst coupling on transesterification reaction efficacy
biodiesel fuel production and application of the reaction to oil processing. J. Mol. in the production of biodiesel from recycled cooking oil. J. Environ. Health Sci. Eng.
Catal. B Enzym. 17, 133–142. 13, 73.
Shrivastava, P., Verma, T.N., Samuel, O.D., Pugazhendhi, A., 2020. An experimental Talha, N.S., Sulaiman, S., 2016. Overview of catalyst in biodiesel production. ARPN J.
investigation on engine characteristics, cost and energy analysis of CI engine fuelled Eng. Appl. Sci. 11.
with Roselle, Karanja biodiesel and its blends. Fuel 275, 117891. Tan, T., Lu, J., Nie, K., Deng, L., Wang, F., 2010. Biodiesel production with immobilized
Silalahi, F.T.R., Simatupang, T.M., Siallagan, M.P., 2020. Biodiesel produced from palm lipase: a review. Biotechnol. Adv. 28, 628–634.
oil in Indonesia: current status and opportunities. AIMS Energy 8, 81–101. Tan, Y.H., Abdullah, M.O., Nolasco-Hipolito, C., Taufiq-Yap, Y.H., 2015. Waste ostrich-
Silitonga, A.S., Shamsuddin, A.H., Mahlia, T.M.I., Milano, J., Kusumo, F., Siswantoro, J., and chicken-eggshells as heterogeneous base catalyst for biodiesel production from
Dharma, S., Sebayang, A.H., Masjuki, H.H., Ong, H.C., 2020. Biodiesel synthesis used cooking oil: catalyst characterization and biodiesel yield performance. Appl.
from Ceiba pentandra oil by microwave irradiation-assisted transesterification: ELM Energy 160, 58–70.
modeling and optimization. Renew. Energy 146, 1278–1291. Tang, Y., Gu, X., Chen, G., 2013. 99% yield biodiesel production from rapeseed oil using
Sim, J.H., Harun@ Kamaruddin, A., Bhatia, S., 2009. Effect of mass transfer and enzyme benzyl bromide–CaO catalyst. Environ. Chem. Lett. 11, 203–208.
loading on the biodiesel yield and reaction rate in the enzymatic transesterification Tariq, M., Ali, S., Khalid, N., 2012. Activity of homogeneous and heterogeneous catalysts,
of crude palm oil. Energy Fuel. 23, 4651–4658. spectroscopic and chromatographic characterization of biodiesel: a review. Renew.
Sim, J.H., Kamaruddin, A.H., Bhatia, S., 2010a. Biodiesel (FAME) productivity, catalytic Sustain. Energy Rev. 16, 6303–6316.
efficiency and thermal stability of lipozyme TL IM for crude palm oil Tat, M.E., Van Gerpen, J.H., 2000. The specific gravity of biodiesel and its blends with
transesterification with methanol. J. Am. Oil Chem. Soc. 87, 1027–1034. diesel fuel. J. Am. Oil Chem. Soc. 77, 115–119.
Sim, J.H., Kamaruddin, A.H., Bhatia, S., 2010b. The feasibility study of crude palm oil Tavares, G.R., Gonçalves, J.E., dos Santos, W.D., da Silva, C., 2017. Enzymatic
transesterification at 30 C operation. Bioresour. Technol. 101, 8948–8954. interesterification of crambe oil assisted by ultrasound. Ind. Crop. Prod. 97,
Simaremare, A., Cahyo, N., Indrawan, H., 2021. Techno–economic study of utilizing CPO 218–223.
as fuel replacement for existing diesel power plant. In: IOP Conference Series: Teo, S.H., Goto, M., Taufiq-Yap, Y.H., 2015. Biodiesel production from Jatropha curcas L.
Materials Science and Engineering. IOP Publishing, 042037. oil with Ca and La mixed oxide catalyst in near supercritical methanol conditions.
Simsek, S., 2020. Effects of biodiesel obtained from Canola, sefflower oils and waste oils J. Supercrit. Fluids 104, 243–250.
on the engine performance and exhaust emissions. Fuel 265, 117026. Teo, S.H., Islam, A., Taufiq-Yap, Y.H., 2016. Algae derived biodiesel using nanocatalytic
Singh, S., Singh, D., 2010. Biodiesel production through the use of different sources and transesterification process. Chem. Eng. Res. Des. 111, 362–370.
characterization of oils and their esters as the substitute of diesel: a review. Renew. Thangaraj, B., Piraman, S., 2016. Heteropoly acid coated ZnO nanocatalyst for Madhuca
Sustain. Energy Rev. 14, 200–216. indica biodiesel synthesis. Biofuels 7, 13–20.
Singh, D., Singal, S., Garg, M., Maiti, P., Mishra, S., Ghosh, P.K., 2015. Transient Thangaraj, B., Jia, Z., Dai, L., Liu, D., Du, W., 2020. Lipase NS81006 immobilized on
performance and emission characteristics of a heavy-duty diesel engine fuelled with functionalized ferric-silica magnetic nanoparticles for biodiesel production. Biofuels
microalga Chlorella variabilis and Jatropha curcas biodiesels. Energy Convers. 11, 811–819.
Manag. 106, 892–900. Thliveros, P., Kiran, E.U., Webb, C., 2014. Microbial biodiesel production by direct
Singh, D., Sharma, D., Soni, S., Sharma, S., Kumari, D., 2019. Chemical compositions, methanolysis of oleaginous biomass. Bioresour. Technol. 157, 181–187.
properties, and standards for different generation biodiesels: a review. Fuel 253, Thoai, D.N., Tongurai, C., Prasertsit, K., Kumar, A., 2017. A novel two-step
60–71. transesterification process catalyzed by homogeneous base catalyst in the first step
Singh, D., Sharma, D., Soni, S., Sharma, S., Sharma, P.K., Jhalani, A., 2020. A review on and heterogeneous acid catalyst in the second step. Fuel Process. Technol. 168,
feedstocks, production processes, and yield for different generations of biodiesel. 97–104.
Fuel 262, 116553. Thoai, D.N., Tongurai, C., Prasertsit, K., Kumar, A., 2019. Review on biodiesel
Siregar, K., Sofyan, H., Nasution, I., Ichwana, R., Syafriandi, S., Sofiah, I., Miharza, T., production by two-step catalytic conversion. Biocatal. Agric. Biotechnol. 18,
2021. Study of life cycle assessment in biodiesel production from crude palm oil and 101023.
its benefits for the sustainability of oil palm industry in Aceh province Indonesia. In: Tian, X., Dai, L., Liu, M., Liu, D., Du, W., Wu, H., 2016. Lipase-catalyzed methanolysis of
IOP Conference Series: Earth and Environmental Science. IOP Publishing, 012017. microalgae oil for biodiesel production and PUFAs concentration. Catal. Commun.
84, 44–47.

37
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Toldrá-Reig, F., Mora, L., Toldrá, F., 2020. Developments in the use of lipase Wilkanowicz, S.I., Hollingsworth, N.R., Saud, K., Kadiyala, U., Larson, R.G., 2020.
transesterification for biodiesel production from animal fat waste. Appl. Sci. 10, Immobilization of calcium oxide onto polyacrylonitrile (PAN) fibers as a
5085. heterogeneous catalyst for biodiesel production. Fuel Process. Technol. 197, 106214.
Tran, D.-T., Yeh, K.-L., Chen, C.-L., Chang, J.-S., 2012. Enzymatic transesterification of Woodford, J.J., Parlett, C.M., Dacquin, J.P., Cibin, G., Dent, A., Montero, J., Wilson, K.,
microalgal oil from Chlorella vulgaris ESP-31 for biodiesel synthesis using Lee, A.F., 2014. Identifying the active phase in Cs-promoted MgO nanocatalysts for
immobilized Burkholderia lipase. Bioresour. Technol. 108, 119–127. triglyceride transesterification. J. Chem. Technol. Biotechnol. 89, 73–80.
Tran, D.-T., Chen, C.-L., Chang, J.-S., 2016. Continuous biodiesel conversion via Xie, W., Ma, N., 2009. Immobilized lipase on Fe3O4 nanoparticles as biocatalyst for
enzymatic transesterification catalyzed by immobilized Burkholderia lipase in a biodiesel production. Energy Fuel. 23, 1347–1353.
packed-bed bioreactor. Appl. Energy 168, 340–350. Xie, W., Wang, H., 2020. Immobilized polymeric sulfonated ionic liquid on core-shell
Uzun, B.B., Kılıç, M., Özbay, N., Pütün, A.E., Pütün, E., 2012. Biodiesel production from structured Fe3O4/SiO2 composites: a magnetically recyclable catalyst for
waste frying oils: optimization of reaction parameters and determination of fuel simultaneous transesterification and esterifications of low-cost oils to biodiesel.
properties. Energy 44, 347–351. Renew. Energy 145, 1709–1719.
Vahid, B.R., Haghighi, M., 2017. Biodiesel production from sunflower oil over MgO/ Xing-cai, L., Jian-Guang, Y., Wu-Gao, Z., Zhen, H., 2004. Effect of cetane number
MgAl2O4 nanocatalyst: effect of fuel type on catalyst nanostructure and improver on heat release rate and emissions of high speed diesel engine fueled with
performance. Energy Convers. Manag. 134, 290–300. ethanol–diesel blend fuel. Fuel 83, 2013–2020.
Vahid, B.R., Saghatoleslami, N., Nayebzadeh, H., Toghiani, J., 2018. Effect of alumina Xiong, W., Li, X., Xiang, J., Wu, Q., 2008. High-density fermentation of microalga
loading on the properties and activity of SO42− /ZrO2 for biodiesel production: Chlorella protothecoides in bioreactor for microbio-diesel production. Appl.
process optimization via response surface methodology. J. Taiwan Inst. Chem. Eng. Microbiol. Biotechnol. 78, 29–36.
83, 115–123. Xu, J., Zhang, R., Han, Z., Wang, Z., Wang, F., Deng, L., Nie, K., 2020. The highly-stable
Van Gerpen, J., 1996. Cetane Number Testing of Biodiesel. Proceedings, Third Liquid immobilization of enzymes on a waste mycelium carrier. J. Environ. Manag. 271,
Fuel Conference: Liquid Fuel and Industrial Products from Renewable Resources. 111032.
American Society of Agricultural Engineers St. Joseph, MI, pp. 197–206. Yaakob, Z., Narayanan, B.N., Padikkaparambil, S., 2014. A review on the oxidation
Vardast, N., Haghighi, M., Dehghani, S., 2019. Sono-dispersion of calcium over Al-MCM- stability of biodiesel. Renew. Sustain. Energy Rev. 35, 136–153.
41used as a nanocatalyst for biodiesel production from sunflower oil: influence of Yagiz, F., Kazan, D., Akin, A.N., 2007. Biodiesel production from waste oils by using
ultrasound irradiation and calcium content on catalytic properties and performance. lipase immobilized on hydrotalcite and zeolites. Chem. Eng. J. 134, 262–267.
Renew. Energy 132, 979–988. Yan, S., Salley, S.O., Ng, K.S., 2009. Simultaneous transesterification and esterification of
Vasić, K., Hojnik Podrepšek, G., Knez, Ž., Leitgeb, M., 2020. Biodiesel production using unrefined or waste oils over ZnO-La2O3 catalysts. Appl. Catal. Gen. 353, 203–212.
solid acid catalysts based on metal oxides. Catalysts 10, 237. Yan, S., Salley, S.O., Ng, K.S., 2012. Methods and catalysts for making biodiesel from the
Veeramalai, S., Ramlee, N.N., Mahdi, H.I., Manas, N.H.A., Ramli, A.N.M., Illias, R.M., transesterification and esterification of unrefined oils. Google Patents.
Azelee, N.I.W., 2022. Development of organic porous material from pineapple waste Yang, G., Tian-Wei, T., Kai-Li, N., Fang, W., 2006. Immobilization of lipase on
as a support for enzyme and dye adsorption. Ind. Crop. Prod. 181, 114823. macroporous resin and its application in synthesis of biodiesel in low aqueous media.
Veljković, V., Lakićević, S., Stamenković, O., Todorović, Z., Lazić, M., 2006. Biodiesel Chin. J. Biotechnol. 22, 114–118.
production from tobacco (Nicotiana tabacum L.) seed oil with a high content of free Yee, K.F., Tan, K.T., Abdullah, A.Z., Lee, K.T., 2009. Life cycle assessment of palm
fatty acids. Fuel 85, 2671–2675. biodiesel: revealing facts and benefits for sustainability. Appl. Energy 86,
Venkat Reddy, C.R., Oshel, R., Verkade, J.G., 2006. Room-temperature conversion of S189–S196.
soybean oil and poultry fat to biodiesel catalyzed by nanocrystalline calcium oxides. Yee, K.F., Wu, J.C., Lee, K.T., 2011. A green catalyst for biodiesel production from
Energy Fuel. 20, 1310–1314. jatropha oil: optimization study. Biomass Bioenergy 35, 1739–1746.
Veny, H., Baroutian, S., Aroua, M.K., Hasan, M., Raman, A.A., Sulaiman, N.M.N., 2009. Yesilyurt, M.K., Arslan, M., Eryilmaz, T., 2019. Application of response surface
Density of Jatropha curcas seed oil and its methyl esters: measurement and methodology for the optimization of biodiesel production from yellow mustard
estimations. Int. J. Thermophys. 30, 529–541. (Sinapis alba L.) seed oil. Int. J. Green Energy 16, 60–71.
Verma, M.L., Barrow, C.J., Puri, M., 2013. Nanobiotechnology as a novel paradigm for Yoosuk, B., Udomsap, P., Puttasawat, B., Krasae, P., 2010. Modification of calcite by
enzyme immobilisation and stabilisation with potential applications in biodiesel hydration–dehydration method for heterogeneous biodiesel production process: the
production. Appl. Microbiol. Biotechnol. 97, 23–39. effects of water on properties and activity. Chem. Eng. J. 162, 135–141.
Verma, M.L., Puri, M., Barrow, C.J., 2016a. Recent trends in nanomaterials immobilised Yousefi, S., Haghighi, M., Vahid, B.R., 2018. Facile and efficient microwave combustion
enzymes for biofuel production. Crit. Rev. Biotechnol. 36, 108–119. fabrication of Mg-spinel as support for MgO nanocatalyst used in biodiesel
Verma, P., Sharma, M., Dwivedi, G., 2016b. Impact of alcohol on biodiesel production production from sunflower oil: fuel type approach. Chem. Eng. Res. Des. 138,
and properties. Renew. Sustain. Energy Rev. 56, 319–333. 506–518.
Verma, M.L., Abraham, R.E., Puri, M., 2020. Nanobiocatalyst Designing Strategies and Yu, C.-Y., Huang, L.-Y., Kuan, I., Lee, S.-L., 2013. Optimized production of biodiesel from
Their Applications in Food Industry. Biomass, Biofuels, Biochemicals. Elsevier, waste cooking oil by lipase immobilized on magnetic nanoparticles. Int. J. Mol. Sci.
pp. 171–189. 14, 24074–24086.
Verziu, M., Cojocaru, B., Hu, J., Richards, R., Ciuculescu, C., Filip, P., Parvulescu, V.I., Yücel, Y., Demir, C., Dizge, N., Keskinler, B., 2011. Lipase immobilization and production
2008. Sunflower and rapeseed oil transesterification to biodiesel over different of fatty acid methyl esters from canola oil using immobilized lipase. Biomass
nanocrystalline MgO catalysts. Green Chem. 10, 373–381. Bioenergy 35, 1496–1501.
Vieira, B., Nadaleti, W.C., Sarto, E., 2021. The effect of the addition of castor oil to Yusuff, A.S., Owolabi, J.O., 2019. Synthesis and characterization of alumina supported
residual soybean oil to obtain biodiesel in Brazil: energy matrix diversification. coconut chaff catalyst for biodiesel production from waste frying oil. South Afr. J.
Renew. Energy 165, 657–667. Chem. Eng. 30, 42–49.
Vieitez, I., Callejas, N., Irigaray, B., Pinchak, Y., Merlinski, N., Jachmanián, I., Yusuff, A.S., Adeniyi, O.D., Azeez, S.O., Olutoye, M.A., Akpan, U.G., 2019. Synthesis and
Grompone, M.A., 2014. Acid value, polar compounds and polymers as determinants characterization of anthill-eggshell-Ni-Co mixed oxides composite catalyst for
of the efficient conversion of waste frying oils to biodiesel. J. Am. Oil Chem. Soc. 91, biodiesel production from waste frying oil. Biofuels, Bioproducts and Biorefining 13,
655–664. 37–47.
Vijayalakshmi, S., Anand, M., Ranjitha, J., 2020. Microalgae-based biofuel production Zafiropoulos, N.A., Ngo, H.L., Foglia, T.A., Samulski, E.T., Lin, W., 2007. Catalytic
using low-cost nanobiocatalysts. Microalgae Cultiv. Biofuels Prod. Elsevier 251–263. synthesis of biodiesel from high free fatty acid-containing feedstocks. Chem.
Vyas, A.P., Verma, J.L., Subrahmanyam, N., 2010. A review on FAME production Commun. 3670–3672.
processes. Fuel 89, 1–9. Zahan, K., Kano, M., 2018. Biodiesel production from palm oil, its by-products, and mill
Wahab, A.G., 2019. Biofuels Annual. USDA, Kuala Lumpur. effluent: a review. Energies 11.
Wang, X., Liu, X., Zhao, C., Ding, Y., Xu, P., 2011. Biodiesel production in packed-bed Zdarta, J., Meyer, A., Jesionowski, T., Pinelo, M., 2018. A general overview of support
reactors using lipase–nanoparticle biocomposite. Bioresour. Technol. 102, materials for enzyme immobilization: characteristics, properties, practical utility.
6352–6355. Catalysts 8.
Wang, Y.-T., Yang, X.-X., Xu, J., Wang, H.-L., Wang, Z.-B., Zhang, L., Wang, S.-L., Zhang, K.-P., Lai, J.-Q., Huang, Z.-L., Yang, Z., 2011. Penicillium expansum lipase-
Liang, J.-L., 2019. Biodiesel production from esterification of oleic acid by a catalyzed production of biodiesel in ionic liquids. Bioresour. Technol. 102,
sulfonated magnetic solid acid catalyst. Renew. Energy 139, 688–695. 2767–2772.
Waudby, H., Zein, S.H., 2021. A circular economy approach for industrial scale biodiesel Zhang, H., Liu, T., Zhu, Y., Hong, L., Li, T., Wang, X., Fu, Y., 2020a. Lipases immobilized
production from palm oil mill effluent using microwave heating: design, simulation, on the modified polyporous magnetic cellulose support as an efficient and recyclable
techno-economic analysis and location comparison. Process Saf. Environ. Protect. catalyst for biodiesel production from Yellow horn seed oil. Renew. Energy 145,
148, 1006–1018. 1246–1254.
Wen, L., Wang, Y., Lu, D., Hu, S., Han, H., 2010a. Preparation of KF/CaO nanocatalyst Zhang, Q., Lei, D., Luo, Q., Wang, J., Deng, T., Zhang, Y., Ma, P., 2020b. Efficient
and its application in biodiesel production from Chinese tallow seed oil. Fuel 89, biodiesel production from oleic acid using metal–organic framework encapsulated
2267–2271. Zr-doped polyoxometalate nano-hybrids. RSC Adv. 10, 8766–8772.
Wen, Z., Yu, X., Tu, S.-T., Yan, J., Dahlquist, E., 2010b. Biodiesel production from waste Zhang, R., Zhu, F., Dong, Y., Wu, X., Sun, Y., Zhang, D., Zhang, T., Han, M., 2020c.
cooking oil catalyzed by TiO2–MgO mixed oxides. Bioresour. Technol. 101, Function promotion of SO42− /Al2O3–SnO2 catalyst for biodiesel production from
9570–9576. sewage sludge. Renew. Energy 147, 275–283.
Wicaksono, W.P., Jati, S.A., Yanti, I., Jiwanti, P.K., 2021. Co-solvent free electrochemical Zhang, Y., Niu, S., Lu, C., Gong, Z., Hu, X., 2020d. Catalytic performance of NaAlO2/
synthesis of biodiesel using graphite electrode and waste concrete heterogeneous γ-Al2O3 as heterogeneous nanocatalyst for biodiesel production: optimization using
catalyst: optimization of biodiesel yield. Bull. Chem. React. Eng. Catal. 19. response surface methodology. Energy Convers. Manag. 203, 112263.

38
H.I. Mahdi et al. Chemosphere 319 (2023) 138003

Zhong, L., Feng, Y., Wang, G., Wang, Z., Bilal, M., Lv, H., Jia, S., Cui, J., 2020. Production Zhou, X., Liu, J., Huang, T., Bian, H., Wang, R., Sha, J., Dai, H., 2020. Near-complete
and use of immobilized lipases in/on nanomaterials: a review from the waste to enzymatic hydrolysis efficiency of Miscanthus using hydrotropic fractionation at
biodiesel production. Int. J. Biol. Macromol. atmospheric pressure. Ind. Crop. Prod. 149, 112365.
Zhou, W., Zhuang, W., Ge, L., Wang, Z., Wu, J., Niu, H., Liu, D., Zhu, C., Chen, Y., Zhou, Y., Searle, S., Kristiana, T., 2021. Opportunities for Waste Fats and Oils as
Ying, H., 2019. Surface functionalization of graphene oxide by amino acids for Feedstocks for Biodiesel and Renewable Diesel in indonesia.
Thermomyces lanuginosus lipase adsorption. J. Colloid Interface Sci. 546, 211–220.

39

You might also like