Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Fuel 370 (2024) 131829

Contents lists available at ScienceDirect

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

Review article

Jatropha curcas oil a potential feedstock for biodiesel production: A


critical review
Joseph V.L. Ruatpuia a, Gopinath Halder b, Michael Vanlalchhandama c, Fanai Lalsangpuii d,
Rajender Boddula e, *, Noora Al-Qahtani e, f, *, Subramaniapillai Niju g, Thangavel Mathimani h, i,
Samuel Lalthazuala Rokhum a, *
a
Department of Chemistry, National Institute of Technology Silchar, Silchar 788010, Assam, India
b
Department of Chemical Engineering, National Institute of Technology Durgapur, Durgapur, India
c
Department of Biotechnology, School of Life Sciences, Mizoram University, Tanhril, Mizoram 796001, India
d
Department of Botany, Mizoram University, Tanhril, Mizoram 796001, India
e
Center for Advanced Materials (CAM), Qatar University, Doha 2713, Qatar
f
Central Laboratories Unit (CLU), Qatar University, Doha 2713, Qatar
g
Department of Biotechnology PSG College of Technology, Coimbatore 641004, Tamil Nadu, India
h
Institute of Research and Development, Duy Tan University, Da Nang, Viet Nam
i
School of Engineering and Technology, Duy Tan University, Da Nang, Viet Nam

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

Keywords: Biodiesel is a low-emissions diesel substitute fuel made from renewable resources. The excessive reliance on
Biodiesel production edible oils for biodiesel production raises concerns about food-versus-fuel issues including starvation in poor
Jatropha curcas oil nations. In contrast, non-edible plant oils are extensively utilized in developing countries due to the high cost of
Renewable feedstock
edible oils, also driven by their high demand as a food source. In the past few years, research efforts have been
Physicochemical properties
Life cycle assessment
directed towards identifying cost-effective and readily available feedstocks for biodiesel production, given that
75 % of the overall production expenses depend upon the choice of feedstock. Therefore, this review aims to
provide an up-to-date overview of advancements and developments associated with converting non-edible
Jatropha curcas oil (JCO) into biodiesel. A brief comparison of physiochemical properties across different
geographical regions is also provided. The study delves into the efficiency, combustion, and emission properties
of diesel engines fueled by Jatropha biodiesel. To visualize the evolving landscape of this research, a total of 400
publications from the Scopus Collection databases, spanning the periods of 2002–2022 and 2012–2022, were
analyzed using VOSviewer. Additionally, life cycle assessment (LCA) facilitates an understanding of the practical
implications and benefits of employing Jatropha biodiesel as a fuel source. Life cycle cost analysis (LCCA), was
conducted to study the economic viability of the JCO biodiesel. A brief conclusion about the potential JCO
feedstock in biodiesel production and their future perspective is presented.

diesel engines as a direct replacement for diesel fuel is becoming more


1. Introduction and more prevalent worldwide [3,4]. In several European countries, a
blend known as B20, consisting of 20 % biodiesel and 80 % conventional
Growing environmental concerns, diminishing fossil fuel reserves, diesel, is widely used. Generally, all diesel equipment can effectively
increasing energy consumption, and the global depletion of fossil fuels handle biodiesel blends up to B20, and most storage and distribution
are expected to increase by 28 % from 2015 to 2040, all of which systems are also well-suited for these blends [5]. So, no significant en­
collectively drive the hunt for bio-based fuel substitutes [1]. Biodiesel gine modifications are needed for these lower-level mixtures. Further­
has garnered significant attention as a potential substitute due to its more, numerous engines, often designed with minimal or no
renewable nature, eco-friendliness, and the possibility of curbing adjustments, are capable of running higher blends, including pure bio­
resource depletion during production [2]. The use of biodiesel fuels in diesel (B100) [6].

* Corresponding authors at: Center for Advanced materials (CAM), Qatar University, Doha 2713, Qatar (N.A.-Q.), Department of Chemistry, National Institute of
Technology Silchar, India (S.L.R.).
E-mail addresses: researchraaj@gmail.com (R. Boddula), noora.alqahtani@qu.edu.qa (N. Al-Qahtani), rokhum@che.nits.ac.in (S.L. Rokhum).

https://doi.org/10.1016/j.fuel.2024.131829
Received 29 January 2024; Received in revised form 22 April 2024; Accepted 30 April 2024
Available online 4 May 2024
0016-2361/© 2024 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Abbreviations LCCA Life cycle cost analysis


MOFs Metal-organic frameworks
BSFC Brake specific fuel consumption MTOR Methanol to oil molar ratio
BTE Break thermal efficiency MUSFA Monounsaturated fatty acids
CN Cetane number NOx Nitrogen oxides
FAMEs Fatty acid methyl esters PM Particulate matter
FDCA-Hf 2,5-furandicarboxylic acid with hafnium PUSFA Polyunsaturated fatty acids
FPP Filter plugging point SAFACAM Sulfonic acid functionalized carbonaceous material
GHGs Greenhouse gases SFA Saturated fatty acids
ID Ignition delay UHC Unburned hydrocarbons
JCO Jatropha curcas oil USD United States dollar
LCA Life cycle assessment

To ensure the long-term sustainability of bio-based energy produc­ growing region in the world is displayed in Fig. 1.
tion, it is important to transform conventional fuels into more valuable Hence, the main objective of the present review is to provide an
and contemporary biofuels like bioethanol, biodiesel, and biogas, overview of ongoing advancements in the research and development of
instead of simply burning them directly [7]. Moreover, the production biodiesel production from Jatropha curcas oil (JCO). This review also
and utilization of biofuels sourced from locally abundant resources not includes a brief analysis of the physiochemical characteristics of JCO
only enhances energy security but also fosters agricultural development and biodiesel originating from various geographical regions, as well as a
and creates employment opportunities boosting the economy. In gen­ comparison of the effectiveness, combustion, and emission characteris­
eral, there are four primary techniques for producing biodiesel: direct tics of diesel engines fueled by Jatropha biodiesel. Pertinent data was
use of vegetable oil, micro-emulsion, thermal cracking, and trans­ gathered from both published and unpublished sources, and the insights
esterification [8]. However, the majority of diesel engines are not obtained from this study can serve as valuable resources for the
compatible with direct vegetable oil use due to its high viscosity, which advancement of sustainable Jatropha biodiesel production. In addition
can lead to engine damage and excessive emissions [9]. Because of their to visualizing and mapping the developments in JCO biodiesel produc­
low cetane numbers (CN) and energy content, biodiesel produced tion, a comprehensive bibliometric analysis was conducted. VOSviewer
through micro-emulsion and thermal cracking frequently fails to fully was utilized to analyze 400 publications sourced from the Scopus
burn. Transesterification, on the other hand, is widely employed due to Collection databases, covering the timeframes of 2002–2022 and
its ease in converting vegetable oil into biodiesel, rendering it the most 2012–2022. The analysis revealed a notable trend over the past decade,
widely adopted method for biodiesel production [10]. Acids, bases, and indicating an increasing utilization of Jatropha curcas oil as a biodiesel
enzymes are the major types of transesterification catalysts used in the feedstock. Furthermore, a study was conducted on the life cycle
manufacture of biodiesel [11,12]. Homogeneous catalysts are not ideal assessment (LCA) and life cycle cost analysis (LCCA) of JCO biodiesel to
for this process due to several constraints and drawbacks, including non- comprehend the practical implications, economic viability and benefits
recyclability and the creation of significant volumes of wastewater of using Jatropha biodiesel as a fuel source.
during catalyst separation and washing [13,14]. In the meanwhile, it is
anticipated that heterogeneous (solid) catalysts will eventually replace 2. Techniques for Jatropha oil extraction and yield globally
homogeneous catalysts due to the ease of catalyst recovery and ad­
vancements in product purification [15,16]. Heterogeneous catalysts Methods for extracting oil play a crucial role in the process of bio­
are less problematic to dispose of, easily recyclable, and reusable diesel production. The Fig. 2 showcases a variety of oil extraction
[17,18]. As a result, it has been claimed that using heterogeneous techniques employed in Jatropha Curcas Oil production. The discussion
transesterification catalysts can lower the cost of producing biodiesel below covers different extraction methods, including solvent extraction,
[19,20]. supercritical fluid extraction, mechanical screw press, and traditional
According to research, feedstock alone accounts for more than 75 % extraction, detailing their processes, advantages, and drawbacks.
of biodiesel production costs [21]. In this regard researchers are moti­ Table 1 displays the oil content of Jatropha seeds from various culti­
vated to explore second-generation feedstocks because first-generation vated regions and methods of oil extraction.
feedstocks are commonly used for food purposes [22]. Second- Solvent extraction stands as the predominantly preferred technique
generation biofuels are those produced from non-food sources. Within employed for extracting oils due to its high yield and quality, utilizes
second-generation feedstocks one potential source of feedstock for bio­ polar solvents to extract oils efficiently and is currently utilized for in­
diesel is Jatropha a non-edible oil produced from plants containing seeds dustrial purposes. On average, the solvent extraction process can recover
[23]. Under ideal environmental conditions it demonstrates greater crop 90–98 % of the available oil [26]. It was believed that utilizing solvent
performances and yield, Jatropha has drawn attention for its ease of extraction could prove to be economically viable for producing biodiesel
adaptation in marginal and non-agricultural areas with tropical and starting from a daily output of 50 tons and beyond [27]. These tech­
subtropical temperatures. Jatropha is a resilient plant that is a good niques yield higher quantities of oil with lower water content compared
source of vegetable oil feedstock for biodiesel because of its quick to traditional and mechanical extraction methods.
growth, simplicity in reproduction, and needs for minimal agitation In traditional extraction, seeds are first removed from the shell,
[24]. Jatropha has been suggested as a potential energy crop because it dried, and roasted. The traditional extraction approach from Tanzania
produces high-quality oil and can be processed more affordably than found that estimated yield 22 % yielded 67 % of the accessible oil from
other feedstocks. Cultivating Jatropha on degraded land can contribute Table 1. Despite its low operational cost and suitability for semi-skilled
to improving the structure and quality of the ecosystem. Without much operators, traditional extraction yields less oil with higher water content
modification, a diesel engine can successfully run on biodiesel made and volatile components compared to other methods [28]. The tradi­
from Jatropha [25]. Furthermore, there is potential for broader utili­ tional methods have been used for many years but are less efficient in oil
zation of Jatropha and biodiesel in generating heat, light, and elec­ recovery compared to newer techniques.
tricity. According to Plantation International (PI) largest Jatropha plant- The mechanical press method stands out as the most conventional

2
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Fig. 1. Largest Jatropha curcas plant growing region in the world, where purple color represent the region producing more than 200,000 ha, gold color represent the
region producing between 20,000 and 200,000 ha, blue color represent the region producing between 5000 and 20,000 ha and red color represent the region
producing 1–5000 ha. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

of available oil, respectively [27]. However this methods yield lower oil
recovery rates with resulting oil containing impurities such as water,
metals, and dust, all compromising the quality of biodiesel production.
Mechanical extractions exhibit superior efficiency compared to tradi­
tional methods.
While in supercritical CO2 extraction, where CO2 acts as a solvent to
remove oils from seeds, eliminating the need for degumming and
dehydration processes. This method reduces time and energy re­
quirements compared to traditional solvent extraction. It’s also envi­
ronmentally friendly due to minimal solvent usage and oil purification.
From Table 1, a study conducted in China, shows that supercritical fluid
extraction yielded a high oil content ranges from (40.28 %) at 90 ◦ C and
Fig. 2. Different methods of oil extraction techniques employed for jatropha 0.5 MPa, outperforming traditional and mechanical extraction. How­
curcas oil extraction. ever, the primary constraints associated with this technique include the
elevated expenses involved in the procedure, the necessity for special­
ized equipment, and the prerequisite for initial drying of raw materials
Table 1
The differences in oil content of Jatropha seeds grown in different regions and to decrease moisture levels to below 20 % [29].
different oil extracted techniques.
3. JCO biodiesel a potential source of sustainable energy
Growing Method of oil extraction Estimated oil yield References
regions (wt.%)
3.1. Major feedstocks used for biodiesel production
Brazil Solvent extraction 31.6 [30]
China Solvent extraction 38.90 [31]
Ghana Solvent extraction 43.2–48.7 [32] The extensive range of available feedstocks is a major benefit for
Nigeria Solvent extraction 47.25 [33] manufacturing biodiesel as a substitute fuel [15,41]. According to
Mexico Solvent extraction 57.4 [34] research, feedstock alone accounts for more than 75 % of biodiesel
India Solvent extraction 13.7–54.4 [35]
production costs as shown in Fig. 3 [8,42]. Therefore, choosing the
Tanzania Traditional extraction 22.02 [28]
Tanzania Mechanical extraction 26.15 [28]
appropriate feedstock is essential to guaranteeing cheap biodiesel pro­
India Mechanical extraction 25–30 [36] duction costs. The availability of feedstock for biodiesel by nation wise
China Supercritical CO2 40.28 [37] from the literature [15] is provided in Table 2. To date, a handful of
extraction more than 350 oil-producing plants known to exist, edible oil such as
Indonesia Supercritical CO2 43.51 [38]
soybean, palm, sunflower, safflower, cottonseed, rapeseed, and peanut
extraction
Spain Supercritical CO2 32.2 [39] oils are seen to be potential replacements for fossil fuels [43]. However,
extraction as concerns mount regarding competition for food, land, and water re­
India Supercritical CO2 60 [40] sources in the production of first-generation bioenergy crops, attention
extraction
has shifted towards second-generation feedstocks sourced from non-
edible oil plants like Jatropha, mahua, and castor. These alternatives
approach for extracting oil compared to other methods. This technique have garnered significant interest due to their adaptability, even
involves employing either a manual ram press or an engine-driven screw thriving in adverse climatic and soil conditions prevalent in arid and
press. Furthermore, the presses utilized for mechanical oil extraction are semi-arid regions. Moreover, they possess a remarkably high oil content
primarily tailored for specific plants and are not efficient for a wide ranging between 63.2 % and 66.4 %, surpassing that of soybean (18.6
range of feedstocks. While the manual ram press technique, capable of %), linseed (33.3 %), and palm kernel (44.6 %). Table 3 provides a
recovering 60–65 % and engine-driven screw press recovering 75–80 % comparison of oil content, seed yield, and oil yield from various second-

3
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Table 3
Comparison of oil content, seed yield and yield of oil from various second
generation feedstocks.
Sl. Types of Oil content Seed yield (Mg/ha Oil yield (Mg/ha
No. Oil (wt.%) Year− 1) Year− 1)

1. Caster 45–60 2.5 0.5–1.0


2. Jatropha 40–60 2.0 2.0–3.0
3. Linseed 35–45 1.0 0.5–1.0
4. Mahua 35–40 2.0 1.0–4.0
5. Karanja 30–40 0.6 2.0–4.0

biodiesel. This can be accomplished through four primary methods:


blending or directly using vegetable oil, thermal cracking, micro-
emulsification, and transesterification [44]. Among these techniques,
transesterification has emerged as the most promising and is currently
employed for large-scale biodiesel production. A catalyst such as an
acid, an alkali, or an enzyme, was used to transesterify edible and non-
edible oil with methanol, which resulted in the production of fatty acid
methyl ester (FAMEs) and crude glycerol as a byproduct [45]. The molar
ratio of alcohol to oil, the type of catalyst, the reaction temperature,
stirring speed, and the purity of the reactants are the variables that
determine the transesterification reaction for effective biodiesel pro­
Fig. 3. General cost breakdown in biodiesel production. duction [46]. The result of transesterifying feedstock is let to settle,
forming two layers: a top layer made of biodiesel and a bottom layer
made of crude glycerol. Either gravity separation or centrifugation is
Table 2 used to separate crude glycerol. However, trace amounts of catalyst,
Several countries with various types of feedstocks used for biodiesel production. glycerol, methanol, and other impurities were present in the top layer of
Sl. Country Feedstock biodiesel. To get the desired product biodiesel processes like washing,
No. drying, and impurity removal must be carried out after the trans­
1. UK Rapeseed/waste cooking oil esterification reaction [15]. The general flowchart of JCO biodiesel
2. France Rapeseed/sunflower production using the transesterification process is shown in Fig. 4.
3. Peru Palm/Jatropha For the typical homogenous biodiesel synthesis techniques, the alkali
4. Argentina Soybeans
must be separated and recovered this is a huge challenge and necessi­
5. Italy Rapeseed/sunflower
6. Greece Cottonseed tates dosing similar doses of acid, i.e., HCl, to neutralize the employed
7. Spain Linseed oil/sunflower catalyst [47]. The formation of a persistent emulsion as a result of the
8. India Jatropha/Pongamia pinnata (Karanja)/soybean/rapeseed/ aqueous quenching makes it difficult to separate methyl ester and
sunflower glycerol. Additionally, wastewater from the process necessitates extra
9. Turkey Sunflower/rapeseed
10. Brazil Soybeans/palm oil/castor/cotton oil
treatment [48]. The benefit of solid material catalysts is they may be
11. Sweden Rapeseed easily collected for reuse once the reaction is finished, which eliminates
12. Norway Animal fats the processing costs related to homogeneous catalysts. An eco-friendly
13. Australia Jatropha/Pongamia/waste cooking oil/animal tallow method of producing biodiesel is using heterogeneous catalyst. More
14. Indonesia Palm oil/Jatropha/coconut
importantly, the heterogeneous based catalyzed technique also elimi­
15. Japan Waste cooking oil
16. Malaysia Palm oil nates soap generation in homogeneous alkali transesterification, the
17. Philippines Coconut/Jatropha oil soap that was produced interferes with the separation of the methyl ester
18. Singapore Palm oil layer from the glycerol layer and is surface active. Various types of
19. Pakistan Jatropha oil Palm/Jatropha/coconut oil heterogeneous catalysts were shown in Table. 4 below.
20. Iran Palm/Jatropha/castor/algae oil
21 Bangladesh Rubber seed/Pongamia pinnata oil
Khan et al. [61] investigated and compared the transesterification of
22. Ghana Palm oil JCO using both homogeneous (potassium hydroxide; KOH) and het­
23. New Waste cooking oil/tallow erogeneous (calcium oxide; CaO) catalysis. The highest yield of 96 %
Zealand was achieved with CaO nanoparticles within 5.5 h of reaction time,
24. Mali Jatropha oil
utilizing 4 wt% catalyst and an 18:1 methanol-to-oil molar ratio. While
25. Germany Rapeseed
26. Ireland Frying oil/animal fat the optimal conditions for KOH included a methanol to oil molar ratio of
27. Cuba Jatropha curcas/Moringa/neem oil 9:1, a catalyst concentration of 5 wt%, and a reaction duration of 3.5 h,
28. Mexico Animal fat/waste oil resulting in a 92 % yield. Both CaO and KOH facilitated high yields of
29. USA Soybeans/waste oil/peanut FAME during JCO transesterification, however, CaO achieving a higher
30. Zimbabwe Jatropha oil Kenya Castor oil
yield (96 %) compared to KOH (92 %). The smaller crystalline size of
31. Canada Rapeseed/animal fat/soybean oil
32. China Jatropha/waste cooking oil/rapeseed oil CaO, as determined by XRD analysis shown in Fig. 5, provided a larger
surface area, enhancing its catalytic activity. Conversely, the reduced
yield of KOH was attributed to its heightened basicity, which promoted
generation feedstocks, indicating JCO as a standout performer with unwanted hydrolysis and saponification reactions.
superior overall yield potential for biodiesel production. A strong, magnetically retrievable Fe3O4@SiO2-SO3H core@shell
nanoparticulate acid catalyst was explored by Changmai et al. [49] and
3.2. Transesterification process for JCO biodiesel production using was effectively synthesized by a step-by-step coating, co-precipitation,
various catalysts and functionalization process. Using the ideal reaction conditions of a
9:1 methanol: oil molar ratio, an 8 wt% catalyst loading, an ambient
There are various technologies available to enhance the quality of temperature of 80 ◦ C, and a reaction time of 3.5 h, the catalyst displayed

4
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Fig. 4. General flowchart of biodiesel production from JCO using the Transesterification process.

Table 4
Different types of heterogeneous catalysts used in JCO biodiesel production.
Sl.No. Catalyst Methanol to oil molar ratio Catalyst loading Reaction time Reaction temperature Yield References
(wt.%) (h) (◦ C) (%)

1. Fe3O4@SiO2-SO3H 9:1 8 3.5 80 98.1 [49]


2. Waste oyster shells 9:1 5 3 65 91.1 [50]
3. Waste eggshell of Gallus domesticus 6:1 2 2 60 98 [51]
4. CaO-La2O3 25:1 3 3 160 98.7 [52]
5. SAFACAM 20:1 9 1 120 98.7 [53]
6. MOF-5 36:1 0.75 9.59 145 93.3 [54]
7. zirconium-based MOFs 25:1 4 4 69.85 97.57 [55]
8. FDCA-Hf 5:1 1.0 1 90 98.0 [56]
9. Zn8@Fe-C400 40:1 7 4 160 100 [57]
10. Immobilized Enterobacter aerogenes lipase 4:1 NA 48 55 94.0 [58]
11. Lipase of Thermomyces lanuginosus 1.5:1 0.5 24 35 80.7 [59]
12. Immobilized Candida cylindracea lipase NA 8.0 24 40 78.0 [60]

Fig. 5. XRD analysis of (a) KOH and (b) CaO. Reproduced with permission of Elsevier from [61].

98.1 % conversion of JCO to biodiesel. In another work, Amesho et al. nanocatalyst from waste domesticated gallus eggshell. After allowing
[50] created a CaO-based heterogeneous catalyst from recycled oyster the reaction to run at 90 ◦ C for 120 min, a high FAME yield (98 %) was
shells, subsequently employed as a proficient and sustainable catalyst achieved.
source in the transesterification reaction of JCO, it proved to be an In the meantime, Lee et al. [52] developed a straightforward
effective and beneficial renewable resource. As a result of the break­ approach for producing biodiesel from non-edible Jatropha oil with a
down, adsorption, and precipitation processes aided by betaine high free fatty acid content, employing a bifunctional acid-base catalyst
amphoteric surfactants, Teo et al. [51], created a novel super basic CaO-La2O3 It was found that the integrated metal–metal oxide between

5
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Ca and La enhanced the catalytic activity due to well dispersion of CaO esterification of Jatropha oils with high acidic values into biodiesel,
on the composite surface and thus, increased the surface acidic and basic achieving yields of up to 98 %. Furthermore, the FDCA-Hf hybrid, this
sites as compared to that of bulk CaO and La2O3 metal oxide. Further, sturdy metal–organic framework, demonstrated high stability and could
Ruatpuia et al. [53] study the application of sulfonic acid functionalized be reused without any loss of activity.
carbonaceous material (SAFACAM) as a catalyst for biodiesel production In another study, Wang et al. [57] synthesized the magnetic acid-
from inedible feedstock JCO. A better biodiesel 98.7 ± 0.6 % conversion base amphoteric nanoparticles that enabled simultaneous esterifica­
was attained under optimal conditions, such as 9 wt% 20:1 MTOR and tion and transesterification of biodiesel production from Jatropha oils
kept the reaction at 120 ◦ C for 50 min. The JCO conversion in the fifth with high acid value at low temperatures. A biodiesel yield of 100 %
recycling was found to be 83.0 ± 0.8 %. A readily available natural from Jatropha was obtained in 4 h at 160 ◦ C, using a methanol/oil molar
biomass derivative (glucose) along with JCO has the capacity to lower ratio of 40:1 and a catalyst dosage of 7 wt%. Furthermore, the catalyst
the overall cost of biodiesel production by converting low-cost raw demonstrated reusability, maintaining biodiesel yields of over 94.3 %
materials. Fig. 6 displayed the general flowchart for the synthesis of even after 10 cycles, with an acid value of 6.3 mg KOH/g.
biomass-derived sulfonic acid-functionalized carbonaceous materials. The employment of an enzyme (lipase) as a biocatalyst for the pro­
Porous materials known as metal–organic frameworks (MOFs) are duction of biodiesel have been studies by Kumari et al. [58] where 94 %
formed by connecting molecular building components, including yield of biodiesel was obtained using Immobilized E. aerogenes lipase to
organic linkers and metal ions/inorganic clusters, through robust coor­ catalyze the transesterification reaction The process involved an oil:
dination interactions [63]. Due to their remarkable tunability, porosity, methanol molar ratio of 1:4, 50 U of immobilized lipase per gram of oil,
and crystallinity, MOFs have demonstrated significant potential in and a t-butanol:oil volume ratio of 0.8:1 at 55 ◦ C over a 48-hour reaction
various sectors, including separation, storage, delivery, and catalysis period. Remarkably, the lipase retained its activity even after under­
[64,65]. Fig. 7 shows different types of MOF used in biodiesel going seven cycles of repeated use, exhibiting negligible loss.
production.
Youssef et al. [54] employ MOF-5 (Metal–organic framework MOF- 4. Properties of synthesized biodiesel and Jatropha oil
5) as a novel heterogeneous acid catalyst to carry out the simulta­
neous esterification/transesterification of two non-edible oils: waste 4.1. Jatropha oil and biodiesel physicochemical properties collected from
cooking oil and Jatropha curcas oil. Furthermore, it was confirmed that different region
the biodiesel produced complies with the ASTM D6751 quality standard
for certain biofuel characteristics. In another study, Dai et al. [55] The assessment of various physicochemical characteristics of Jatro­
created zirconium-based MOFs, employing the impregnation method to pha curcas oil indicates its potential as a viable raw material for biodiesel
synthesize the catalyst utilized in the preparation of biodiesel from JCO. production. Brief physicochemical properties are shown in Fig. 8 below.
Under the most favorable reaction conditions, the reaction temperature According on measurement temperatures and growing places, JCO
reached 343 K, the catalyst dosage about oil was 4 wt%, the methanol to estimated density ranges from 913 to 940 kg/m3 as shown in Table 5.
oil ratio was 25:1, and a reaction time of 4 h resulted in an impressive When compared to the comparable biodiesel produced by the trans­
average conversion rate of 97.57 % for JCO. According to Li et al. [56] esterification of JCO, Jatropha oil has a substantially higher density
the robust acid-base catalytic sites of 2,5-furandicarboxylic acid with 838–883 kg/m3 estimated from different tropical and subtropical re­
hafnium (FDCA-Hf) enabled the efficient simultaneous (trans) gions, along with their respective region provided in Table 6. Addi­
tionally, studies revealed that the composition of fatty acids in the
feedstock affects both the oil density and biodiesel properties [66,67]. In
comparison to biodiesel made from unsaturated fatty acids, saturated
fatty acid biodiesel has a lower density [66].
The term “viscosity” refers to the amount of internal friction that
occurs when one part of a fluid moves over another, obstructing the flow
of the fluid. The estimated viscosity of JCO, which is quite viscous,
varied from 17 to 46.3 mm2s− 1 based on the growth conditions and agro
ecological conditions at the time of measurement (Table 6). As the
measurement temperature increased, both Jatropha oil and biodiesel
viscosity decreased, as seen in Tables 5 and 6. From the result, during
assessing the samples Jatropha oil viscosity from different regions,
consideration should be given to the measuring temperature. Re­
searchers generally agree that the estimated viscosities were still too
varied from what gasoline and diesel had in common 2.26 mm2s− 1 [68].
Vegetable oils’ enormous molecular mass and complex chemical
makeup contribute to their high viscosity. Additionally, the fatty acid
composition of the feedstock is related to the viscosity of oil.
The fuel’s flashpoint for biodiesel is the temperature at which it ig­
nites upon contact with a flame. A lower flashpoint can lead to fuel filter
and line blockages, making its importance. The flashpoints varied for
Jatropha oil 190.5–240 ◦ C and JCO biodiesel 135–170 ◦ C as shown in
Tables 5 and 6. The outcome suggests that the transesterification and
purification have lower the Jatropha oil flashpoint when it converted in
biodiesel. Yet, the ignition points of Jatropha oil and biodiesel far
exceeded that of conventional petrol-diesel, at 71.5 ◦ C. Moreover,
Jatropha oil and biodiesel’s flashpoint values varied significantly among
diverse cultivation areas.
Fig. 6. General flowchart for the synthesis of biomass-derived sulfonic acid- The cloud point, pour point, and filter plugging point (FPP) are the
functionalized carbonaceous materials. Reproduced with permission of Elsev­ three key factors that govern the use of biodiesel at a lower temperature.
ier from [62]. Pour point is the lowest temperature at which fuel can flow freely or the

6
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Fig. 7. Several metal organic framework broadly classified acidic, basic, enzymatic and bifunctional catalysts for biodiesel production. Reproduced with permission
of Elsevier from [63].

Fig. 8. Jatropha curcas oil and JCO biodiesel physicochemical properties studied.

Table 5
Jatropha oil physiochemical property originated from different regions.
Country Density (kg Viscosity (mm2 Flashpoint Pour point Cloud point Acid value (mg Calorific value (MJ Cetane References
m− 3) s− 1) (◦ C) (◦ C) g− 1) g− 1) number

Brazil 921 − − 2 − 8.4 40.3 − [30]


Ghana − 17.03 − − − 3.5 − − [32]
China 915.11 28.44 190.5 − − 12.8 38.9 − [66]
Cape-Verd 926.02 46.32 240 1.0 6.0 2.2 38.9 − [69]
Malaysia 9401 24.54 225 4.0 − 28.0 38.7 − [70]
India 926.02 33.54 228 − − 9.0 16.8 37.6 56.1 [71]
Thailand 918.04 − − − − 1.0 39.8 47.7 [72]
South- 918.04 − − − − 4.6 39.4 42.5 [72]
Africa
Borneo 914.04 − − − − 4.6 40.0 56.0 [72]
Philippines 913.04 − − − − 0.9 38.9 33.0 [72]
1, 2, 3
and 4 represent measurements taken at 15, 20, 30, and 40 ◦ C, respectively.

temperature at which wax can gel the gasoline, whereas the temperature biodiesel’s reported cold flow characteristics, however, are all margin­
at which the cloud of wax in the fuel initially becomes apparent after ally higher than those of gasoline and diesel [66,69].
cooling is known as the cloud point [67,73]. The biodiesel cold flow The calorific value of a fuel, often known as its heating value, is a key
characteristics is influenced by the level of saturation. Higher numbers sign of its energy content. JCO biodiesel has a calorific value that is quite
linked to specific cold flow parameters indicate more saturated bio­ similar to that of gasoline and diesel [69]. In their study, Ong et al.
diesel. Unsaturated fatty acid-dominated biodiesel performs well in (2013) investigated the calorific value of biodiesel and Jatropha oil,
engines at lower temperatures than fuels mostly made of saturated fatty finding them to be 38.96 MJ kg− 1 and 40.42 MJ kg− 1, respectively.
acids. Furthermore, the ignition quality is compromised by the reduced Furthermore, it was found that the calorific value of Jatropha oil man­
CN found in the fatty acid with a larger unsaturation percentage. Table 6 ufactured in Brazil amounted to 40.31 MJ kg− 1 [30]. The data presented
shows the typical pour and cloud characteristics of biodiesel produce in Table 6 indicates that the calorific value of Jatropha oil differs across
from JCO studied over several geographic locations. Jatropha several tropical regions, ranging from 37.6 to 40.3 MJ kg− 1. In another

7
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Table 6
Jatropha biodiesel physiochemical properties originated from different geographical locations.
Country Density (kg Viscosity (mm2 Flashpoint Pour point Cloud point Acid value (mg Calorific value (MJ Cetane References
m− 3) s− 1 ) (◦ C) (◦ C) (◦ C) g− 1) g− 1) number

India 880.22 7.322 170.0 − 1 0.04 40.0 50.6 [31]


China 838.81 3.914 161.8 2.0 3 0.24 40.4 50.0 [66]
Malaysia 880.01 4.84 135.0 2.0 0.40 39.2 [70]
India 875.02 13.882 168.0 − 6.0 1 0.37 39.9 [80]
India 883.02 4.844 162.3 − 3.0 2.3 0.45 38.4 [81]
Nigeria 867.31 2.314 142.0 − 10 2 0.59 [82]

Physiochemical Standards of conventional diesel, ASTM D 6751-06and EN 14214 of JCO


Diesel 814–840 2.5–5.7 50–98 − 20 to 5 − 10 to − 5 42–45.9 45–55 [83]
ASTM D 860 to 9001 1.9 to 6.0 130 − 15 to 10 − 3 to 12 47 [83]
6751–06
EN 14,214 860 to 9001 3.5–5–04 >101 − − 0.5 − 51 [68]
1, 2, 3, and 4
Measurements taken at 15, 20, 30, and 40 ◦ C, respectively.

study, Aransiola et al. [74] when compared to gasoline and diesel, (C20:0), linolenic acid (C18:3), and erucic acid (C22:1) were identified.
biodiesel calorific values were between 9 and 12 % lower. JCO and Sinha et al. [84] provided evidence that as the seeds aged, there was a
biodiesel have a lower heating value than petrol and diesel, which re­ significant increase in both the oil production and the levels of oleic and
sults in higher fuel consumption for equivalent tasks. linoleic acid. Furthermore, with the aging of the seeds, there was a
In comparison to petrol diesel, biodiesel often has a greater CN [75]. decrease in the levels of palmitic and linolenic acids. Moreover, the fatty
An important factor in determining the quality of the fuel is the CN, acid composition of JCO could be influenced by factors such as local
which represents how well the biodiesel ignites [76]. Biodiesel’s ignition environment, agronomic techniques, genotype, climate, and other con­
delay (ID) is influenced by its CN [77]. ID is the period of time between ditions. Table 7 reveals notable discrepancies in the fatty acid compo­
the beginning of injection and the onset of combustion. Shorter IDs sition of Jatropha oil/biodiesel among various regions.
correspond to greater CNs, and vice versa. Since JCO biodiesel contains
a lot of oxygen, it has a higher CN. Consequently, it has a shorter ID time 5. Engine performances and emission of Jatropha biodiesel
than a petrol-diesel engine [67]. Due to its complete combustion, a high
CN leads to an uncomplicated cold temperature start, reduced exhaust Numerous types of biodiesel, such as Karanja, Polanga, Mahua,
emissions, decreased noise, and minimal white smoke [78]. The cetane Rubber seed, Cotton seed, Jojoba oil, Tobacco oil, Neem, Linseed oil,
number of JCO biodiesel was measured at 50. Furthermore, Tables 5 and and Jatropha, have been thoroughly studied as potential fuel sources
6 present the CN values of biodiesel and Jatropha oil obtained from over the years [90]. In contrast to other biodiesel varieties, Jatropha
several studies, revealing significant variations depending on the biodiesel presents notable benefits as it boasts a lower viscosity, making
geographical locations of cultivation [69]. Higher CN values and fewer it easier to utilize and transport. Moreover, it offers a high Cetane
carbon chain branches result in shorter ID and longer carbon chains, number and calorific value, both crucial factors for ensuring efficient
lower CN levels are caused by shorter, more branched carbon chains combustion within engines [83]. The test results of Jatropha biodiesel in
[73,79]. various engine models are shown in Table 8. Overall, the performance of
Jatropha biodiesel was found to be satisfactory, with no notable tech­
nical problems. Jatropha biodiesel presents a combination of high break
4.2. Fatty acid composition of Jatropha curcas oil thermal efficiency (BTE) and reduced fuel usage. According to the study
conducted by Sahoo et al, [91] utilizing a 20 % blend of Jatropha bio­
The physicochemical properties of fats and oils are greatly influenced diesel could result in a brake power that is 0.09 %–2.64 % higher than
by the types of fatty acids that make up triglycerides. The oil’s fatty acid that of diesel fuel. Brake specific fuel consumptions (BSFC) for all the
composition, alongside factors such as chain length and degree of biodiesel blends with diesel increases with blends and decreases with
unsaturation, plays a vital role in revealing the physical properties of speed. It is also observed that brake specific energy consumption (BSEC)
both the oil and biodiesel. The saturation level of fatty acids can increases with blends and decreases with speed. The maximum increase
generally be categorized into saturated fatty acids Cn:0 (SFA), mono­ in BSEC for B100 at 1200 rpm was 20.21 % which is higher than that of
unsaturated fatty acids Cn:1 (MUSFA), and polyunsaturated fatty acids conventional diesel. It was also found that the Smoke emission reduces
Cn: ≥2 (PUSFA). The increased methyl ester levels of saturated or with blends and engine speeds. The use of B20, B50, B100, caused a
polyunsaturated fatty acids have a negative impact on the various reduction in smoke in the range of 28.57 %, 40.9 %, 64.28 % with
physical and chemical properties. As a result, oils that are rich in respect to diesel at a rated engine speeds of 2200 rpm, respectively. It is
MUSFA, are recommended to use feedstock with reduced levels of observed that the minimum reduction in smoke of at least 1.29 % for
PUSFA and controlled SFA acids for the production of biodiesel. Bio­ B20 for lower engine speeds (1200–1600 rpm) and 15.84 % for higher
diesel derived from oilseeds containing elevated levels of SFA exhibited engine speeds (1800–2200 rpm). The Smoke emission maximum
enhanced oxidation stability and heating value. In contrast, biodiesels reduction (above 60 %) is seen in case of B100 at engine speed
obtained from oils containing higher levels of MUSFA and PUSFA, this (1200–1600 rpm). However, as the biodiesel content in the fuel blend
biodiesel displayed unfavorable cold flow characteristics because it so­ increases, the engine power decreases [92]. Similarly, Huang et al, [93]
lidified at higher temperatures. According to Islam et al. [72] the ma­ study the experimental performances and emissions of a diesel engine
jority of jatropha oil (75 %) consists of unsaturated fatty acids, with using two different biodiesels derived from Chinese pistache oil and
approximately 76–78 % of oleic and linoleic acids, the major constitu­ jatropha oil compared with pure diesel and found that at higher engine
ents of the fatty acid composition in JCO were palmitic acid (C16:0) speeds with greater biodiesel content in the fuel blends, Jatropha bio­
having 0.14 %–20 %, stearic acid (C18:0) having 1.48 % 17 %, oleic acid diesel leads to an approximate 6.8 % increase in brake-specific fuel
(C18:1) ranging from 12.8 % to 48 %, and linoleic acid, which makes up consumption (BSFC). However, the improvement in BTE is only
28.7 %–46.7 % of its composition. Additionally, minor components such marginally increased by 0.1–6.7 %. Smoke emissions from the engine
as myristic acid (C14:0), palmitoleic acid (C16:1), arachidic acid

8
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Table 7
Jatropha oil/biodiesel fatty acid composition collected from different regions.
Country Myristic acid Palmitic acid Palmitoleic acid Stearic acid Oleic acid Linoleic acid Linolenic acid Other References
(14:0) (16:0) (16:1) (18:0) (18:1) (18:2) (18:3)

Malaysia 0.1 13 0.7 5.8 43.5 35.4 0.3 0.2 [66]


Indonesian 0–0.1 14.1–15.3 0–1.3 3.7–9.8 34.3–44.8 29.0–43.2 0–0.3 0–0.5 [85]
Brazil 0.14 6.1 21.8 46.4 [30]
China 13.23 0.85 5.4 40.62 36.99 0.22 [31]
Nigeria 20.3 7.4 41.8 32.6 2.72 [33]
Nicaragua 0.1 13.6 0.8 7.4 34.6 42.2 0.2 0.6 [86]
Cape Verde 0.1 15.1 0.9 7.1 43.7 31.4 0.2 0.5 [87]
India 11.88 7.8 42.72 30.08 0.28 4.7 [79]
Thailand 15.2 0.7 6.8 43.6 32.2 0.2 0.3 [88]
Congo- 15.6 1.27 5.8 40.1 37.6 [89]
Brazzaville

Table 8
Jatropha biodiesel engine performance and results at different test conditions.
Types of Engine Test Condition Result References

BTE BSFC Engine power

3-Cylinder, 4-stroke diesel Varied speeds (1200, 1800, 20 % blend of biodiesel The best BSFC improvement is The maximum increase in [91]
engine, WC, D: 3440 cc, 2200 rpm) and different increased the engine power observed with B20. Later power is observed for B50
CR: 18:1, RS: 2200 rpm blends (B20, B50, B100) by 0.09–2.64 % increases with blends at 2000 and 2100 rpm.
decreases BSFC
1-Cylinder, 4-stroke diesel Varied speeds (1500, 2000 An increase of 0.2–3.5 % and An increase of 9.3 % and 6.8 % Comparable to conventional [93]
engine, D: 815 cc, RP: 8.82 rpm) 0.1–6.7 % were found for BSFC were found for 1500 and pure diesel
kW, CR: 17:1, RS: 2000 1500 and 2000 rpm, 2000 rpm respectively
rpm respectively
4-Cylinder, 4-stroke diesel Varied speeds at full load Lower than diesel fuel Comparatively similar to fossil Greater than diesel fuel [83]
engine, TC, D: 1609 cc, condition diesel operation
RP:84.5 kW, CR: 18.5:1,
RS: 3800 rpm
1-Cylinder, 4 Stroke diesel Varied blends (B20, B40, B50, Greater than diesel fuel at 20 % blend offers lower than [83]
engine, CR: 16.5:1, RP: B60, B80, B100) at different about 20–80 % blends that of diesel fuel, while other
5HP, RS: 1500 rpm loads (25 %, 50 %, 75 %, 100 blends performed similarly
%)
1–Cylinder, 4-stroke diesel Various blends (B10, B20, B50, Decreased with increasing Less than diesel fuel Greater than diesel fuel [94]
engine, WC, D: 1007 cc, B100) and varied speeds percentage of biodiesel in the
CR: 18.5:1, RS:2400 rpm (1000–2400 rpm) fuel
1-Cylinder, 4-stroke diesel Varied speeds (1800, 2500, Less than diesel fuel Greater than diesel fuel [95]
engine, CR: 18:1, D: 395 cc, 3200 rpm)
RP: 5.59 kW, RS: 3600 rpm

fuelled by the biodiesels are lowered significantly than that fuelled by oxides of nitrogen (NOx) emission is increased from 6 % to 26 % for the
diesel. At the engine speed of 1500 rpm, the reduction of smoke emis­ 10 % and 20 % biodiesel blends, this can be attributed to the lean air/
sions are between 8 % and 35 % and At the engine speed of 2000 rpm, fuel ratio, as biodiesel contains 12 % more oxygen in its molecular
the reduction of smoke emissions are 12 %–57 %. structure. In addition, the higher viscosity of the blend of fuel leads to a
In another work Prabhu et al, [96] study the total emissions for bigger droplet size and shorter ignition delay. In summary, the use of
jatropha oil, palm oil, soybean oil, and petroleum fuels and have been biodiesel and their blend results in a reduction in Smoke emission, hy­
figured as 18.5, 34.1, 52.1, and 87.1 g/MJ of CO2-equivalent, respec­ drocarbon, carbon monoxide, and particulate matter. Fig. 9 displayed
tively and found that total emissions from Jatropha biodiesel are the schematic diagram of test set up for a three cylinder DI diesel engine.
minimum. Also reveal that the following exhaust emissions are
decreased by using jatropha biodiesel such as, hydro carbon (HC) by
44–68 %, carbon monoxide (CO) by 48–72 %, Carbon dioxide (CO2) by 5.1. Effects of impurities present in biodiesel
52–78 %, and smoke by 49–73 %. According to research conducted by
Xue et al. [97] suggests that pure biodiesel leads to substantial re­ Insufficient separation and purification of biodiesel lead to a greater
ductions in various emissions such as PM, CO, and CO2, while NOx levels presence of impurities that have a negative impact on the performance
are found to increase compared to those of diesel fuel. According to their of engines. While a certain level of impurities is allowed in commercially
findings, the average reduction in PM ranges from 50 to 72.73 %, CO produced biodiesel, it must still adhere to international standards and
from 50 to 73 %, HC from 45 to 67 %, and CO2 from 50 to 80 %. On the specifications. According to reports, most of the impurities found in
other hand, the use of pure biodiesel results in an increase in NOx levels biodiesel remain consistent regardless of the feedstock, alcohol, catalyst,
by 5.58–25.97 %. In a similar vein, Mofijur et al. [98] examined the or operational conditions used during the production process [99].
engine performance and emission traits of no blend (conventional Table 9 provides a summary of how impurities affect biodiesel, requiring
diesel), as well as biodiesel blends with petroleum diesel comprising 10 the use of separation and purification methods that are appropriate for
% biodiesel and 20 % biodiesel. They observed a reduction in hydro­ raw biodiesel meeting the required standards.
carbon (HC) emissions by 3.84 % and 10.25 %, and carbon monoxide
(CO) emissions decreased by 16 % and 25 % for the 10 % and 20 % 6. Life cycle assessment of biodiesel production
blends, respectively, compared to the no blend scenario. However,
Life cycle assessment (LCA) is critical for assessing the environmental

9
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Fig. 9. Schematic diagram of test set up for a three cylinder DI diesel engine. Reproduced with permission of Elsevier from [91].

Table 9
Effects of impurities present in product biodiesel and its drawbacks
[100,99,101,102].
Impurities Drawbacks

Water Decreases the intensity of combustion heat, system corrosion, fuel


pump malfunction, ice crystal formation, bacterial growth, and
piston pitting.
Glycerol Decantation, fuel tank bottom deposits, injector fouling, storage
problem, settling problems, and severity of engine durability
problems
Soap/catalyst Pose corrosion problems, damage injectors, plugging of filters, and
weakening of engine
Glycerides Turbidity, higher viscosity, and deposition at piston, valves, and
nozzles
Methanol Lower the flash points, lower viscosity, and causes corrosion
Free fatty Less oxidation stability and corrosion of vital engine components
acids

impact of synthesized biodiesel. The LCA process begins with resource


extraction and ends with final product formation and wastewater
disposal [103]. The LCA is a standardized ISO approach designed to
identify the transfer of pollution between different stages or types of
environmental impact. This strategy adheres to the so-called “from
cradle to grave” philosophy and takes into account a product’s complete
life cycle [104]. A key method in the production of biodiesel is the Fig. 10. Different stages in Life cycle assessment of biodiesel production.
systematic analysis of the distribution of energy as well as a mass that
reveals the most efficient processes within the production sequence and
artificial ponds [105]. Comparing the outcomes of different life cycle
points out opportunities for enhancing biodiesel production.
assessments is challenging because of the diverse range of functional
units assessed and the lack of standardization.
6.1. Step-wise life cycle assessment for biodiesel production
6.1.1. Agricultural cultivation
The following consecutive steps make up the life cycle assessment of There are various smaller stages, within the agricultural cultivation
biodiesel produced from first- and second-generation sources as shown phase, including: (i) manufacturing agricultural inputs such as pesti­
in Fig. 10: (i) agricultural cultivation, (ii) transportation, (iii) biodiesel cides, seeds and fertilizers, etc., (ii) producing capital items, such as farm
production, and (iv) combustion [103]. In the case of third-generation buildings and equipment, and (iii) conducting agricultural tasks such as
biodiesels, the farming phase encompasses the cultivation of algae in

10
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

planting, applying nutrients, harvesting, and clearing land. The lifespan outputs, such as wastes, emissions, and end products, for the production
of a plant plays a significant role in the LCA of both first- and second- of biodiesel [114,115]. Inventory can stem from either literature anal­
generation biodiesel. In the case of perennial plants, the cultivation ysis or process simulation. Commercial and public databases like Euro­
stages mentioned earlier occur periodically. Nonetheless, it is essential pean Life Cycle Database, Ecoinvent, US Life Cycle Inventory Database,
to also take into account additional stages such as pre-nursery, nursery, and the Biograce offer a streamlined data collection process, contingent
and immature plantation phases [106]. Additionally, the use of agri­ upon maintaining requisite data quality [116] which help the researcher
cultural feedstocks is linked to the occupation of land and changes in significantly.
land use, which can significantly contribute to the overall greenhouse
gas emissions throughout the lifecycle of biodiesel. The application of 6.3. Life cycle impact assessment (LCIA) of JCO biodiesel production
fertilizers to enhance the productivity of energy crops leads to an
elevation in the production of nitrous oxide, while deforestation reduces Life cycle impact assessment (LCIA) assesses potential environmental
the absorption of CO2. Therefore, the environmental consequences of impacts across different environmental categories. During this stage of
the agricultural phase will be affected by factors such as deforestation, the LCA, specific environmental impacts are quantified using the results
degradation of meadows, the tillage system, and the specific conditions of the LCI flows. The resultant data can be acquired through LCA soft­
of the locations including air, temperature, and soil organic carbon ware like Umberto, Gabi, Simapro, etc. In LCIA, two primary approaches
content [107]. are employed: problem-oriented methods, often referred to as mid-
points, and damage-oriented methods, also known as end points. Nazir
6.1.2. Transportation and oil extraction stages and Setyaningsih [115] when comparing the LCIA of biodiesel produc­
The following stages make up the life cycle assessment for the tion from palm oil and JCO using Simapro 7 and ECO indicator 99, found
transportation step: (i) moving agricultural products to the oil extraction that the production of biodiesel from palm oil had a greater overall
facility, (ii) moving the oil to the biodiesel manufacturing facility, and environmental impact and GHG emissions than that of biodiesel from
(iii) moving the biodiesel to the final user. The oil extraction process is jatropha oil. The production of palm oil biodiesel results in a 66 %
further categorized into three stages: (a) upstream activities involved in greater environmental burden compared to jatropha biodiesel produc­
producing input materials such as chemicals, fuels, and electricity, (b) tion. It is also found that when compared to other life cycle stages,
the production of capital goods, and (c) the operations of oil extraction cultivation has the greatest environmental impact.
plants, such as extraction of crude oil, treatment of wastewater. Several
state-of-the-art reviews have examined the various aspects of the bio­ 6.4. Life cycle interpretation
diesel life cycle, revealing that accidents pose risks at every stage
including equipment malfunction, operator error, uncontrollable Life cycle interpretation is this stage where the results obtained are
external circumstances, ignition, and other things that can all contribute summarized and analyzed. It involves formulating sets of recommen­
to these accidents [108]. Unfortunately, such occurrences can result in dations, identifying limitations, and drawing conclusions from the entire
accidents that cause harm, deaths, and oil and fuel spills. These negative LCA analysis [116]. Interpretation can occur either through literature
outcomes can significantly impact the soil and water environment analysis or by employing software-guided techniques in process
[109]. simulation.

6.1.3. Biodiesel production stage 6.5. Life cycle cost analysis (LCCA) and economic viability of JCO
Various steps make up the biodiesel production stage, including: (i) biodiesel production
obtaining input materials (such as chemicals and electricity), (ii)
manufacturing capital goods, and (iii) performing various operations at The Life Cycle Cost Analysis (LCCA) is incorporated into a life cycle
the biodiesel production plant (such as manufacturing, refinement, assessment (LCA) framework to holistically assess the energy, environ­
treatment of wastewater, etc.). The specific technique utilized to pro­ mental, and economic consequences of producing Jatropha biodiesel.
duce biodiesel has a considerable impact on the total life cycle assess­ The LCCA comprises expenses for feedstock (including feedstock oil,
ment. Once the transesterification process is complete, resulting the methanol, and catalyst) as well as additional expenditures like power
formation of biodiesel esters glycerol and catalyst components are usage, transportation, and labor [117]. The feedstock’s cost-
frequently left behind. Additionally, the process of transesterification effectiveness and effective performance significantly influence the po­
requires substantial energy input for heating and stirring, as well as the tential for the economical and productive commercialization of the
utilization of catalysts. However, despite these difficulties, using enzy­ trans-esterification process, aimed at generating substantial quantities of
matic catalysts or adding CO2 to supercritical lipid extraction and biodiesel [118]. As per the information from GlobalPetrolPrice.com, the
transesterification processes could significantly reduce the environ­ current worldwide average cost for diesel is 1.27 United State Dollars
mental effects of biodiesel synthesis, as previously discussed [110]. (USD) per liter. Table 10 presents different production costs for JCO
biodiesel per liter in USD.
6.1.4. Combustion stage According to Liu et al. [119] the study conducted in China utilizing
The combustion phase involves analyzing the emissions produced by LCCA reveals the Jatropha biodiesel cost to be 796.32 USD/ton, which is
both stationary and mobile engines using biodiesel. Typically, when merely 0.7 USD/L (since JCO biodiesel density is 0.88 g cm− 3 so 1 ton =
biodiesel is burned, it results in a reduced release of carbon monoxide 1136 L). The finding shows that Jatropha oil accounted for only 44.37 %
(CO), greenhouse gases (GHGs), particulate matter (PM), and unburned
hydrocarbons (UHC) compared to regular diesel fuel [111]. However, it Table 10
does tend to generate higher levels of nitrogen oxide (NOx) emissions in Several studies of LCCA JCO biodiesel production.
comparison to conventional diesel fuel. Nevertheless, numerous recent
Sl. Authors Method of JCO biodiesel Cost References
studies have revealed that when biodiesel is combusted under optimized No. production per Litre in USD
circumstances NOx emissions are reduced [112,113].
1. Liu et al. Transesterification 0.7 [119]
2. Kumar et al. Transesterification 0.58 [120]
6.2. Life cycle inventory (LCI) of JCO biodiesel production 3. Tewari Transesterification 0.56 [121]
4. Singh et al. Transesterification 0.74 [122]
The life cycle inventory (LCI) is a technique step that consists in­ 5. Yusuf and Supercritical 0.78 [123]
Kamarudin methanol
ventorying the inputs, such as energy, water, and raw materials, and the

11
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

of the overall cost of producing biodiesel, which is significantly lower conversion. Using this process, an initial capital investment of USD 7.4
compared to other edible feedstocks like soybean oil and palm oil. The million was established, along with a minimum bio-methanol produc­
energy required to produce 1 ton of Jatropha biodiesel amounts to tion price of 0.91 USD/L. Accordingly, the biodiesel production plants
17566.16 MJ, with fertilizer usage accounting for 78.14 % of the total decrease the cost of biodiesel production by 26.36 %. In the meantime,
energy and methanol production using up 18.65 %. Singh et al. [122] reported that when produced on a large scale, Jatro­
In another study, Kumar et al. [120] found that the price of Jatropha pha biodiesel becomes economically viable due to the revenue gener­
biodiesel in India stood at 0.58 USD per liter, Yet, incorporating 5 % ated by byproducts like seed cake and glycerol in the biodiesel
biodiesel, priced at Rs. 48, with petro-diesel, which is currently sold at production process.
around Rs. 41 per liter, will result in an approximate rise of Rs. 0.40 per
liter in the overall cost. If the blending is increased to 20 %, the cost will 7. Bibliometric mapping of JCO biodiesel production
elevate by Rs. 1.6 per liter.
In a similar vein, Tewari [121] examine the life cycle cost analysis of Bibliometric analyses is a widely recognized and rigorous approach
JCO biodiesel in India and ascertain that the overall cost of biodiesel for scrutinizing extensive volumes of scientific information. They facil­
amounts to 0.56 USD/L. Notably, the specific production cost is itate the visualization of the progression of a particular research topic
contingent upon the production scale. Government policies, and in­ and concurrently showing significant insights into the swiftly evolving
centives for producers, processors, and consumers play a role in deter­ domains within that field [127]. A detailed bibliometric mapping was
mining the outcome in certain instances. conducted to provide a deeper grasp of the developments in the field of
Meanwhile, Singh et al. [122] have documented that the production Jatropha curcas biodiesel production. Utilizing VOS viewer software, a
of one liter of Jatropha biodiesel carries a net cost estimated at 0.74 knowledge map was generated, incorporating data from a total of 400
USD, which is lower in comparison to the existing diesel cost of 1.27 USD publications sourced from the Scopus Collection databases spanning the
per liter. years 2012–2022. The aim was to scrutinize and visually represent the
The LCCA findings from Table 10 and Table 11 reveal the JCO bio­ developments in this field, with a focus on biodiesel keywords. Through
diesel Cost per Litre in USD and Cost per Tons in USD and show sig­ these mapping techniques, researchers can glean valuable insights into
nificant potential in Jatropha oil-based biodiesel production [124]. It is the current pulse of research development and anticipate future trends.
crucial to note that in Table 10 while various ways have been used to To comprehensively capture the diverse timelines associated with
produce biodiesel from Jatropha oil, the bulk of these technologies have various generations of biofuels. Fig. 11a illustrates the bibliometric
only been evaluated on a small scale, either in laboratory settings or mapping derived from the time range 2002–2022. The analysis reveals
pilot projects [68]. According to studies, economic concerns account for that biodiesel derived from feedstocks such as, oleic acid, waste cooking
24 % of the numerous factors affecting the cultivation of Jatropha [125]. oil, palm oil, soybean oil, and microalgae dominates as feedstocks for
However, in Table 11, a comparison was made between B100 JCO biodiesel production, whereas jatropha curcas oil is notably absent in
biodiesel and other B100 second-generation biodiesel feedstocks. The the image. However, over the last decade 2012–2022, as depicted in
analysis revealed that the price of biodiesel derived from JCO was the Fig. 11b, there is a significant surge in the popularity of jatropha as a
most economical among the other feedstocks, have lower price per tons feedstock for biodiesel production. This rise in prominence could be
showing high potential for a feedstock for biodiesel production in large attributed to improved engine performance and favorable outcomes
scale. This affordability is expected to incentivize consumers to adopt from life cycle cost analysis, highlighting the growing preference for
biodiesel for road use. Consequently, it is anticipated that this will drive jatropha oil as a viable feedstock.
up the demand for biodiesel in vehicles, thereby bolstering the economy
in rural areas. 8. Conclusions and future prospect
The synthesized biodiesel must not only meet performance and
emissions standards but also should compete economically with fossil This review has provided a comprehensive overview of the potential
petroluem diesel. Yusuf and Kamarudin [123] conducted a study of Jatropha curcas oil as a biodiesel feedstock. By analyzing current
comparing the economic feasibility of producing biodiesel from JCO, progress and attributes associated with JCO biodiesel production, it has
canola oil, and waste frying oil. Through techno-economic analysis, they filled a gap in the literature regarding the utilization of non-edible
determined that the total production costs were USD 31.20 million, USD feedstocks for sustainable fuel generation. The exploration of Jatropha
50.9 million, and USD 35.51 million per year, respectively. The pro­ curcas oil represents a promising move towards eco-friendly fuel pro­
duction of biodiesel from JCO was found to yield higher glycerol credits duction, highlighted by its increasing trend as a biodiesel feedstock over
owing to the purity of the glycerol, thereby generating increased reve­ the past decade. Key findings indicate that the solvent extraction process
nue. This study demonstrates that biodiesel production from JCO is the yields the highest oil recovery 90–98 % of the available oil from jatropha
most cost-effective option, offering purities comparable to those re­ curcas oil making it a favorable extraction technique. Physiochemical
ported in other studies. In another study, Makepa et al. [126] found that properties of JCO and JCO biodiesel vary depending on geographical
enhancing the economic efficiency of biodiesel production is achievable regions and extraction methods, Jatropha biodiesel exhibits a similar
through a reduction in the methanol costs needed for the trans­ density and lower calorific value, while the viscosity of both JCO and
esterification process. A thermodynamic model is developed to predict JCO biodiesel is quite variable, ranging from 17 to 46.3 mm2s− 1 and
bio-methanol production by simulating steam gasification of Jatropha 2.31 to 13.88 mm2s− 1, respectively, depending on growth and agro-
curcas L. (Euphorbiaceae) seedcake and Fischer-Tropsch synthesis of the ecological conditions at the time of measurement. Among the various
syngas, employing Aspen Plus. The integration of gasification with methods used in biodiesel production, transesterification process is the
Fischer-Tropsch synthesis offers an alternative pathway for biomass most suitable method. More than 95 % biodiesel yield is obtain using
5:1–40:1 methanol to oil molar ratio, 1–9 catalyst loading, 60–160 ◦ C
temperature and 1–10 h reaction time. Blending JCO biodiesel with
Table 11 conventional diesel shows potential for enhancing power and reducing
Comparison of B100 price from different feedstock [70].
smoke emissions, although challenges remain in addressing NOx emis­
Sl. No. Types of feedstock B100 biodiesel Cost per Tons in USD sions during combustion. Moreover, life cycle cost analysis suggests that
1. Rapeseed oil 940–965 large-scale production of B100 biodiesel from JCO could be economi­
2. Soybean oil 800–805 cally viable, indicating significant potential for commercialization.
3. Palm oil 720–750 Overall, JCO offers a promising alternative feedstock for biodiesel pro­
4. Jatropha oil 400–500
duction, capable of reducing production costs and environmental

12
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

Fig. 11. The bibliometric mapping of biofuel production using VOSviewer incorporating data from a total of 400 publications sourced from the Scopus Collection
databases spanning the years 2002–2022 (a) last 20 years bibliometric mapping, (b) last 10 years bibliometric mapping.

impact. future research, enhancing the emission levels of NOx during the com­
The future studies should take into account the abundant availability bustion of biodiesel is necessary due to the higher potency of nitrous
of non-edible feedstocks. Although Jatropha Curcas Oil has shown oxide (N2O) as a greenhouse gas compared to CO2. However design of
promising results in biodiesel production and has the potential to reduce low cost catalyst also play a crucial role for setting up large-scale bio­
overall production costs, a thorough investigation into its commercial diesel production utilizing a low cost inedible feedstock.
availability for large-scale biodiesel production is warranted. Scaling up
biodiesel production should be prioritized in future outlook to facilitate
a more comprehensive evaluation of life cycle cost analysis. Also in

13
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

CRediT authorship contribution statement [14] Tang ZE, Lim S, Pang YL, Ong HC, Lee KT. Synthesis of biomass as heterogeneous
catalyst for application in biodiesel production: state of the art and fundamental
review. Renew Sustain Energy Rev 2018;92:235–53. https://doi.org/10.1016/j.
Joseph V.L. Ruatpuia: Writing – original draft, Validation, Data rser.2018.04.056.
curation, Conceptualization. Gopinath Halder: Writing – review & [15] Changmai B, Vanlalveni C, Ingle AP, Bhagat R, Rokhum L. Widely used catalysts
editing. Michael Vanlalchhandama: Validation, Data curation. Fanai in biodiesel production: a review. RSC Adv 2020;10:41625–79. https://doi.org/
10.1039/d0ra07931f.
Lalsangpuii: Formal analysis, Data curation. Rajender Boddula: [16] Chakraborty R, Bepari S, Banerjee A. Application of calcined waste fish (Labeo
Writing – review & editing. Noora Al-Qahtani: Writing – review & rohita) scale as low-cost heterogeneous catalyst for biodiesel synthesis. Bioresour
editing. Subramaniapillai Niju: Validation, Data curation. Thangavel Technol 2011;102:3610–8. https://doi.org/10.1016/j.biortech.2010.10.123.
[17] Laskar IB, Rajkumari K, Gupta R, Chatterjee S, Paul B, Rokhum L. Waste snail
Mathimani: Writing – review & editing. Samuel Lalthazuala Rokhum: shell derived heterogeneous catalyst for biodiesel production by the
Supervision, Investigation, Conceptualization. transesterification of soybean oil. RSC Adv 2018;8:20131–42. https://doi.org/
10.1039/c8ra02397b.
[18] Kafuku G, Lee KT, Mbarawa M. The use of sulfated tin oxide as solid superacid
catalyst for heterogeneous transesterification of Jatropha curcas oil. Chem Pap
Declaration of competing interest
2010;64:734–40. https://doi.org/10.2478/s11696-010-0063-1.
[19] Helwani Z, Aziz N, Bakar MZA, Mukhtar H, Kim J, Othman MR. Conversion of
The authors declare that they have no known competing financial Jatropha curcas oil into biodiesel using re-crystallized hydrotalcite. Energy
Convers Manag 2013;73:128–34. https://doi.org/10.1016/j.
interests or personal relationships that could have appeared to influence
enconman.2013.04.004.
the work reported in this paper. [20] Kiss AA, Omota F, Dimian AC, Rothenberg G. The heterogeneous advantage:
biodiesel by catalytic reactive distillation. Top Catal 2006;40:141–50. https://
Data availability doi.org/10.1007/s11244-006-0116-4.
[21] Ermias N, Berta A, Amente C, Setarge Y. Biodiesel production in Ethiopia: current
status and future prospects. Sci Afr 2023;19:e01531. https://doi.org/10.1016/j.
Data will be made available on request. sciaf.2022.e01531.
[22] Hoekman SK, Broch A, Robbins C, Ceniceros E, Natarajan M. Review of biodiesel
composition, properties, and specifications. Renew Sustain Energy Rev 2012;16:
Acknowledgement 143–69. https://doi.org/10.1016/j.rser.2011.07.143.
[23] Selvaraj R, Praveenkumar R, Moorthy IG. A comprehensive review of biodiesel
production methods from various feedstocks. Biofuels 2019;10:325–33. https://
This work was supported by Qatar University through a National
doi.org/10.1080/17597269.2016.1204584.
Capacity Building Program Grant (NCBP), [QUCP-CAM-22/24-463]. [24] Zhang F, Tian XF, Fang Z, Shah M, Wang YT, Jiang W, et al. Catalytic production
Statements made herein are solely the responsibility of the authors. of Jatropha biodiesel and hydrogen with magnetic carbonaceous acid and base
Open Access funding provided by the Qatar National Library. synthesized from Jatropha hulls. Energy Convers Manag 2017;142:107–16.
https://doi.org/10.1016/j.enconman.2017.03.026.
[25] Kumar V, Kumar A. Assessment and usability of Jatropha biodiesel blend with
References phenolic antioxidant to control NOx emissions of an unmodified diesel engine.
Environ Sci Pollut Res 2023;30:108051–66. https://doi.org/10.1007/s11356-
023-29995-4.
[1] Elsayed M, Eraky M, Osman AI, Wang J, Farghali M, Rashwan AK, et al.
[26] He Y, Peng T, Guo Y, Li S, Guo Y, Tang L, et al. Nontoxic oil preparation from
Sustainable valorization of waste glycerol into bioethanol and biodiesel through
Jatropha curcas L. seeds by an optimized methanol/n-hexane sequential
biocircular approaches: a review. Environ Chem Lett 2023. https://doi.org/
extraction method. Ind Crops Prod 2017;97:308–15. https://doi.org/10.1016/j.
10.1007/s10311-023-01671-6.
indcrop.2016.12.034.
[2] Sahu G, Gupta NK, Kotha A, Saha S, Datta S, Chavan P, et al. A review on
[27] Bhuiya MMK, Rasul MG, Khan MMK, Ashwath N, Azad AK. Prospects of 2nd
biodiesel production through heterogeneous catalysis route. ChemBioEng Rev
generation biodiesel as a sustainable fuel — Part: 1 selection of feedstocks, oil
2018;5:231–52. https://doi.org/10.1002/cben.201700014.
extraction techniques and conversion technologies. Renew Sustain Energy Rev
[3] Venu H, Appavu P. Al2O3 nano additives blended Polanga biodiesel as a potential
2016;55:1109–28. https://doi.org/10.1016/j.rser.2015.04.163.
alternative fuel for existing unmodified DI diesel engine. Fuel 2020;279:118518.
[28] Aboubakar X, Goudoum A, Bébé Y, Mbofung CMF. Optimization of Jatropha
https://doi.org/10.1016/j.fuel.2020.118518.
curcas pure vegetable oil production parameters for cooking energy. South Afr J
[4] Ruatpuia JVL, Halder G, Shi D, Halder S, Lalthazuala S. Comparative life cycle
Chem Eng 2017;24:196–212. https://doi.org/10.1016/j.sajce.2017.09.002.
cost analysis of bio-valorized magnetite nanocatalyst for biodiesel production:
[29] Rubio-rodríguez N, De DSM, Beltrán S, Jaime I, Sanz MT, Rovira J. Supercritical
modeling, optimization, kinetics and thermodynamic study. Bioresour Technol
fluid extraction of fish oil from fish by-products: a comparison with other
2024;393:130160. https://doi.org/10.1016/j.biortech.2023.130160.
extraction methods. J Food Eng 2012;109:238–48. https://doi.org/10.1016/j.
[5] Doan BQ, Nguyen XP, Pham VV, Dong TMH, Pham MT, Le TS. Performance and
jfoodeng.2011.10.011.
emission characteristics of diesel engine using ether additives: a review. Int J
[30] de Oliveira JS, Leite PM, de Souza LB, Mello VM, Silva EC, Rubim JC, et al.
Renew Energy Dev 2022;11:255–74. https://doi.org/10.14710/
Characteristics and composition of Jatropha gossypiifoliaand Jatropha curcas L.
ijred.2022.42522.
oils and application for biodiesel production. Biomass Bioenergy 2009;33:
[6] Balat M. Potential alternatives to edible oils for biodiesel production - a review of
449–53. https://doi.org/10.1016/j.biombioe.2008.08.006.
current work. Energy Convers Manag 2011;52:1479–92. https://doi.org/
[31] Wang R, Hanna MA, Zhou WW, Bhadury PS, Chen Q, Song BA, et al. Production
10.1016/j.enconman.2010.10.011.
and selected fuel properties of biodiesel from promising non-edible oils:
[7] Chen C, Qu S, Guo M, Lu J, Yi W, Liu R, et al. Waste limescale derived recyclable
Euphorbia lathyris L., Sapium sebiferum L. and Jatropha curcas L. Bioresour
catalyst and soybean dregs oil for biodiesel production: analysis and
Technol 2011;102:1194–9. https://doi.org/10.1016/j.biortech.2010.09.066.
optimization. Process Saf Environ Prot 2021;149:465–75. https://doi.org/
[32] Agyemang NS, Antwi K. Variations in oil content and biodiesel yield of Jatropha
10.1016/j.psep.2020.11.022.
curcas from different agro-ecological zones of Ghana. Int J Renew Sustain Energy
[8] Koh MY, Tinia TI. A review of biodiesel production from Jatropha curcas L. oil.
2014;3:76–81. https://doi.org/10.11648/j.ijrse.20140304.11.
Renew Sustain Energy Rev 2011;15:2240–51. https://doi.org/10.1016/j.
[33] Akintayo ET. Characteristics and composition of Parkia biglobbossa and Jatropha
rser.2011.02.013.
curcas oils and cakes. Bioresour Technol 2004;92:307–10. https://doi.org/
[9] Singh SP, Singh D. Biodiesel production through the use of different sources and
10.1016/S0960-8524(03)00197-4.
characterization of oils and their esters as the substitute of diesel: a review.
[34] Makkar HPS, Kumar V, Oyeleye OO, Saikia A, Angulo-Escalante MA, Becker K.
Renew Sustain Energy Rev 2010;14:200–16. https://doi.org/10.1016/j.
Jatropha platyphylla, a new non-toxic Jatropha species: physical properties and
rser.2009.07.017.
chemical constituents including toxic and antinutritional factors of seeds. Food
[10] Axelsson L, Franzén M, Ostwald M, Berndes G, Lakshmi G, Ravindranath NH.
Chem 2011;125:63–71. https://doi.org/10.1016/j.foodchem.2010.08.037.
Perspective: Jatropha cultivation in southern India: assessing farmers’
[35] Kumar R, Das N. Survey and selection of Jatropha curcas L. germplasm:
experiences. Biofuels Bioprod Biorefin 2012;6:246–56. https://doi.org/10.1002/
assessment of genetic variability and divergence studies on the seed traits and oil
bbb.
content. Ind Crops Prod 2018;118:125–30. https://doi.org/10.1016/j.
[11] Abdullah SHYS, Hanapi NHM, Azid A, Umar R, Juahir H, Khatoon H, et al.
indcrop.2018.03.032.
A review of biomass-derived heterogeneous catalyst for a sustainable biodiesel
[36] Raheman H, Mondal S. Biogas production potential of jatropha seed cake.
production. Renew Sustain Energy Rev 2017;70:1040–51. https://doi.org/
Biomass Bioenergy 2012;37:25–30. https://doi.org/10.1016/j.
10.1016/j.rser.2016.12.008.
biombioe.2011.12.042.
[12] Buasri A, Chaiyut N, Loryuenyong V, Wongweang C, Khamsrisuk S. Application of
[37] Liu J, Chen P, He J, Deng L, Wang L, Lei J, et al. Extraction of oil from Jatropha
eggshell wastes as a heterogeneous catalyst for biodiesel production. Sustain
curcas seeds by subcritical fluid extraction. Ind Crops Prod 2014;62:235–41.
Energy 2013;1:7–13. https://doi.org/10.12691/rse-1-2-1.
https://doi.org/10.1016/j.indcrop.2014.08.039.
[13] Boey PL, Maniam GP, Hamid SA. Performance of calcium oxide as a
heterogeneous catalyst in biodiesel production: a review. Chem Eng J 2011;168:
15–22. https://doi.org/10.1016/j.cej.2011.01.009.

14
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

[38] Chen WH, Chen CH, Chang CMJ, Chiu YH, Hsiang D. Supercritical carbon dioxide [63] Gouda SP, Dhakshinamoorthy A, Rokhum SL. Metal-organic framework as a
extraction of triglycerides from Jatropha curcas L. seeds. J Supercrit Fluids 2009; heterogeneous catalyst for biodiesel production: a review. Chem Eng J Adv 2022;
51:174–80. https://doi.org/10.1016/j.supflu.2009.08.010. 12:100415. https://doi.org/10.1016/j.ceja.2022.100415.
[39] Fernández CM, Fiori L, Ramos MJ, Pérez Á, Rodríguez JF. Supercritical extraction [64] Mahmoud E. Recent advances in the design of metal–organic frameworks for
and fractionation of Jatropha curcas L. oil for biodiesel production. J Supercrit methane storage and delivery. J Porous Mater 2021;28:213–30. https://doi.org/
Fluids 2015;97:100–6. https://doi.org/10.1016/j.supflu.2014.11.010. 10.1007/S10934-020-00984-Z/FIGURES/6.
[40] Mouahid A, Bouanga H, Crampon C, Badens E. Supercritical CO2 extraction of oil [65] Hao M, Qiu M, Yang H, Hu B, Wang X. Recent advances on preparation and
from Jatropha curcas: an experimental and modelling study. J Supercrit Fluids environmental applications of MOF-derived carbons in catalysis. Sci Total
2018:2–11. https://doi.org/10.1016/j.supflu.2017.11.014. Environ 2021;760:143333. https://doi.org/10.1016/J.
[41] Atadashi IM, Aroua MK, Aziz AA. High quality biodiesel and its diesel engine SCITOTENV.2020.143333.
application: a review. Renew Sustain Energy Rev 2010;14:1999–2008. https:// [66] Ong HC, Silitonga AS, Masjuki HH, Mahlia TMI, Chong WT, Boosroh MH.
doi.org/10.1016/j.rser.2010.03.020. Production and comparative fuel properties of biodiesel from non-edible oils:
[42] Lim S, Teong LK. Recent trends, opportunities and challenges of biodiesel in Jatropha curcas, Sterculia foetida and Ceiba pentandra. Energy Convers Manag
Malaysia: an overview. Renew Sustain Energy Rev 2010;14:938–54. https://doi. 2013;73:245–55. https://doi.org/10.1016/j.enconman.2013.04.011.
org/10.1016/j.rser.2009.10.027. [67] Canakci M, Sanli H. Biodiesel production from various feedstocks and their effects
[43] Abdulla R, Chan ES, Ravindra P. Biodiesel production from Jatropha curcas: a on the fuel properties. J Ind Microbiol Biotechnol 2008;35:431–41. https://doi.
critical review. Crit Rev Biotechnol 2011;31:53–64. https://doi.org/10.3109/ org/10.1007/s10295-008-0337-6.
07388551.2010.487185. [68] Ewunie GA, Morken J, Lekang OI, Yigezu ZD. Factors affecting the potential of
[44] Amani A, Rahmati S, Fakhlaei R, Barati B, Wang S, Doherty W, et al. Emerging Jatropha curcas for sustainable biodiesel production: a critical review. Renew
technologies for biodiesel production: processes, challenges, and opportunities. Sustain Energy Rev 2021;137:110500. https://doi.org/10.1016/j.
Biomass Bioenergy 2022;163:106521. https://doi.org/10.1016/j. rser.2020.110500.
biombioe.2022.106521. [69] Ramesh A, Palanichamy K, Tamizhdurai P, Umasankar S, Sureshkumar K,
[45] Das A, Li H, Kataki R, Agrawal PS, Moyon NS, Gurunathan B, et al. Terminalia Shanthi K. Sulphated Zr–Al2O3 catalysts through jatropha oil to green-diesel
arjuna bark – a highly efficient renewable heterogeneous base catalyst for production. Mater Lett 2019;238:62–5. https://doi.org/10.1016/j.
biodiesel production. Renew Energy 2023;212:185–96. https://doi.org/10.1016/ matlet.2018.11.158.
j.renene.2023.05.066. [70] Hamzah NHC, Khairuddin N, Siddique BM, Hassan MA. Potential of Jatropha
[46] Saikia K, Moyon NS, Chai F. Process optimization and kinetic studies of Musa curcas L. as biodiesel feedstock in Malaysia: a concise review. Processes 2020;8:
glauca catalyzed biodiesel production 2023;36. 1–11. https://doi.org/10.3390/pr8070786.
[47] Welter RA, Santana HS, Gaziola L, Torre D, Barnes MC. Biodiesel production by [71] Sparsh S, Siva C, Sahu R, Shyam R, Kaur J, Sahu M, et al. Future prospects of
heterogeneous catalysis and eco-friendly routes. ChemBioEng Rev 2023:86–111. biodiesel production from jatropha in India. Mater Today Proc 2022;63:A22–6.
https://doi.org/10.1002/cben.202200062. https://doi.org/10.1016/j.matpr.2022.07.273.
[48] Das S, Anal JMH, Kalita P, Saikia L, Rokhum SL. Process optimization of biodiesel [72] Islam AKMA, Yaakob Z, Anuar N, Primandari SRP, Osman M. Physiochemical
production using waste snail shell as a highly active nanocatalyst 2023;2023. properties of Jatropha curcas seed oil from different origins and candidate plus
[49] Changmai B, Wheatley AEH, Rano R, Halder G, Selvaraj M, Rashid U, et al. plants (CPPs). JAOCS, J Am Oil Chem Soc 2012;89:293–300. https://doi.org/
A magnetically separable acid-functionalized nanocatalyst for biodiesel 10.1007/s11746-011-1908-7.
production. Fuel 2021;305:121576. https://doi.org/10.1016/j. [73] Demirbas A, Karslioglu S. Biodiesel production facilities from vegetable oils and
fuel.2021.121576. animal fats. Energy Sources, Part A Recover Util Environ Eff 2007;29:133–41.
[50] Amesho KTT, Lin YC, Chen CE, Cheng PC, Shangdiar S. Kinetics studies of https://doi.org/10.1080/009083190951320.
sustainable biodiesel synthesis from Jatropha curcas oil by exploiting bio-waste [74] Aransiola EF, Ojumu TV, Oyekola OO, Madzimbamuto TF. A review of current
derived CaO-based heterogeneous catalyst via microwave heating system as a technology for biodiesel production: state of the art. Biomass Bioenergy 2013:
green chemistry technique. Fuel 2022;323:123876. https://doi.org/10.1016/j. 1–22. https://doi.org/10.1016/j.biombioe.2013.11.014.
fuel.2022.123876. [75] Kumar R, Dixit AK. Combustion and emission characteristics of variable
[51] Teo SH, Islam A, Masoumi HRF, Taufiq-Yap YH, Janaun J, Chan ES, et al. compression ignition engine fueled with Jatropha curcas ethyl ester blends at
Effective synthesis of biodiesel from Jatropha curcas oil using betaine assisted different compression ratio. J Renew Energy 2014;2014:1–12. https://doi.org/
nanoparticle heterogeneous catalyst from eggshell of Gallus domesticus 2017;vol. 10.1155/2014/872923.
111. https://doi.org/10.1016/j.renene.2017.04.039. [76] Van Gerpen J. Biodiesel processing and production. Fuel Process Technol 2005;
[52] Lee HV, Juan JC, Taufiq-Yap YH. Preparation and application of binary acid-base 86:1097–107. https://doi.org/10.1016/j.fuproc.2004.11.005.
CaO-La2O3 catalyst for biodiesel production. Renew Energy 2015;74:124–32. [77] Avhad MR, Sánchez M, Bouaid A, Martínez M, Aracil J, Marchetti JM. Modeling
https://doi.org/10.1016/j.renene.2014.07.017. chemical kinetics of avocado oil ethanolysis catalyzed by solid glycerol-enriched
[53] Ruatpuia JVL, Changmai B, Pathak A, Alghamdi LA, Kress T, Halder G, et al. calcium oxide. Energy Convers Manag 2016;126:1168–77. https://doi.org/
Green biodiesel production from Jatropha curcas oil using a carbon-based solid 10.1016/j.enconman.2016.07.060.
acid catalyst: a process optimization study. Renew Energy 2023;206:597–608. [78] Bello EI, Mogaji TS, Agge M. The effects of transesterification on selected fuel
https://doi.org/10.1016/j.renene.2023.02.041. properties of three vegetable oils. Mech Eng 2011;3:218–25.
[54] Ben YC, Yam AC, Rivera AZJM. Simultaneous esterification/transesterification of [79] Mazumdar P, Borugadda VB, Goud VV, Sahoo L. Physico-chemical characteristics
waste cooking oil and Jatropha curcas oil with MOF - 5 as a heterogeneous acid of Jatropha curcas L. of North East India for exploration of biodiesel. Biomass
catalyst. Int J Environ Sci Technol 2021;18:3313–26. https://doi.org/10.1007/ Bioenergy 2012;46:546–54. https://doi.org/10.1016/j.biombioe.2012.07.005.
s13762-020-03088-y. [80] Tambunan AH, Situmorang JP, Silip JJ, Joelianingsih A, Araki T. Yield and
[55] Dai Q, Yang Z, Li J, Cao Y, Tang H, Wei X. Zirconium-based MOFs-loaded ionic physicochemical properties of mechanically extracted crude Jatropha curcas L
liquid-catalyzed preparation of biodiesel from Jatropha oil. Renew Energy 2021; oil. Biomass Bioenergy 2012;43:12–7. https://doi.org/10.1016/j.
163:1588–94. https://doi.org/10.1016/j.renene.2020.09.122. biombioe.2012.04.004.
[56] Li H, Yang T, Fang Z. Biomass-derived mesoporous Hf-containing hybrid for [81] Sharma Dugala N, Singh Goindi G, Sharma A. Evaluation of physicochemical
efficient Meerwein-Ponndorf-Verley reduction at low temperatures. Appl Catal B characteristics of Mahua (Madhuca indica) and Jatropha (Jatropha curcas) dual
Environ 2018;227:79–89. https://doi.org/10.1016/j.apcatb.2018.01.017. biodiesel blends with diesel. J King Saud Univ - Eng Sci 2021;33:424–36. https://
[57] Wang YT, Fang Z, Yang XX, Yang YT, Luo J, Xu K, et al. One-step production of doi.org/10.1016/j.jksues.2020.05.006.
biodiesel from Jatropha oils with high acid value at low temperature by magnetic [82] Aderibigbe FA, Mustapha SI, Adewoye TL, Mohammed IA, Gbadegesin AB,
acid-base amphoteric nanoparticles. Chem Eng J 2018;348:929–39. https://doi. Niyi FE, et al. Qualitative role of heterogeneous catalysts in biodiesel production
org/10.1016/j.cej.2018.05.039. from Jatropha curcas oil. Biofuel Res J 2020;7:1159–69. https://doi.org/
[58] Kumari A, Mahapatra P, Garlapati VK, Banerjee R. Enzymatic transesterification 10.18331/BRJ2020.7.2.4.
of Jatropha oil. Biotechnol Biofuels 2009;2:1–7. https://doi.org/10.1186/1754- [83] Thapa S, Indrawan N, Bhoi PR. An overview on fuel properties and prospects of
6834-2-1. Jatropha biodiesel as fuel for engines. Environ Technol Innov 2018;9:210–9.
[59] Bueso F, Moreno L, Cedeño M, Manzanarez K. Lipase-catalyzed biodiesel https://doi.org/10.1016/j.eti.2017.12.003.
production and quality with Jatropha curcas oil: exploring its potential for [84] Sinha P, Islam A, Negi MS, Tripathi SB. Changes in oil content and fatty acid
Central America. J Biol Eng 2015;9:1–7. https://doi.org/10.1186/s13036-015- composition in Jatropha curcas during seed development. Ind Crop Prod 2015;77:
0009-9. 508–10. https://doi.org/10.1016/j.indcrop.2015.09.025.
[60] Kabbashi NA, Mohammed NI, Alam MZ, Mirghani MES. Hydrolysis of Jatropha [85] Berchmans HJ, Hirata S. Biodiesel production from crude Jatropha curcas L. seed
curcas oil for biodiesel synthesis using immobilized Candida cylindracea lipase. oil with a high content of free fatty acids. Bioresour Technol 2008;99:1716–21.
J Mol Catal B Enzym 2015;116:95–100. https://doi.org/10.1016/j. https://doi.org/10.1016/j.biortech.2007.03.051.
molcatb.2015.03.009. [86] Foidl N, Foidl G, Sanchez M, Mittelbach M, Hackel S. Jatropha curcas L. as a
[61] Khan K, Ul-Haq N, Rahman WU, Ali M, Rashid U, Ul-Haq A, et al. Comprehensive source for the production of biofuel in Nicaragua. Bioresour Technol 1996;58:
comparison of hetero-homogeneous catalysts for fatty acid methyl ester 77–82. https://doi.org/10.1016/S0960-8524(96)00111-3.
production from non-edible jatropha curcas oil. Catalysts 2021;11. https://doi. [87] Sarin R, Sharma M, Sinharay S, Malhotra RK. Jatropha-Palm biodiesel blends: an
org/10.3390/catal11121420. optimum mix for Asia. Fuel 2007;86:1365–71. https://doi.org/10.1016/j.
[62] Ao S, Changmai B, Vanlalveni C, Van LM, Wheatley AEH, Lalthazuala S. Biomass fuel.2006.11.040.
waste-derived catalysts for biodiesel production: recent advances and key [88] Supamathanon N, Wittayakun J, Prayoonpokarach S. Properties of Jatropha seed
challenges. Renew Energy 2024;223:120031. https://doi.org/10.1016/j. oil from Northeastern Thailand and its transesterification catalyzed by potassium
renene.2024.120031.

15
J.V.L. Ruatpuia et al. Fuel 370 (2024) 131829

supported on NaY zeolite. J Ind Eng Chem 2011;17:182–5. https://doi.org/ Northwestern US. J Clean Prod 2017. https://doi.org/10.1016/j.
10.1016/j.jiec.2011.02.004. jclepro.2017.11.172.
[89] Kumar A, Sharma S. An evaluation of multipurpose oil seed crop for industrial [108] Rivera SS, Olivares RDC, Baziuk PA. Assessment of biofuel accident risk: a
uses (Jatropha curcas L.): a review. Ind Crops Prod 2008;28:1–10. https://doi. preliminary study 2015;II:1–6.
org/10.1016/j.indcrop.2008.01.001. [109] Gupta PK, Fate ASM. Transport, and bioremediation of biodiesel and blended
[90] Ashraful AM, Masjuki HH, Kalam MA, Fattah IMR, Imtenan S, Shahir SA, et al. biodiesel in subsurface environment: a review. J Environ Eng 2020;146:1–10.
Production and comparison of fuel properties, engine performance, and emission https://doi.org/10.1061/(ASCE)EE.1943-7870.0001619.
characteristics of biodiesel from various non-edible vegetable oils: a review. [110] Taher H, Giwa A, Abusabiekeh H, Al-zuhair S. Biodiesel production from
Energy Convers Manag 2014;80:202–28. https://doi.org/10.1016/j. Nannochloropsis gaditana using supercritical CO 2 for lipid extraction and
enconman.2014.01.037. immobilized lipase transesteri fi cation: economic and environmental impact
[91] Sahoo PK, Das LM, Babu MKG, Arora P, Singh VP, Kumar NR, et al. Comparative assessments. Fuel Process Technol 2020;198:106249. https://doi.org/10.1016/j.
evaluation of performance and emission characteristics of jatropha, karanja and fuproc.2019.106249.
polanga based biodiesel as fuel in a tractor engine. Fuel 2009;88:1698–707. [111] Nabi N, Zare A, Hossain FM, Ristovski ZD, Brown RJ. Reductions in diesel
https://doi.org/10.1016/j.fuel.2009.02.015. emissions including PM and PN emissions with diesel-biodiesel blends. J Clean
[92] Singh D, Sharma D, Soni SL, Inda CS, Sharma S, Sharma PK, et al. Prod 2017;166:860–8. https://doi.org/10.1016/j.jclepro.2017.08.096.
A comprehensive review of physicochemical properties, production process, [112] Gharehghani A, Asiaei S, Khalife E, Najafi B, Asiaei S, Khalife E, et al.
performance and emissions characteristics of 2nd generation biodiesel feedstock: Simultaneous reduction of CO and NOx emissions as well as fuel consumption by
Jatropha curcas. Fuel 2021;285:119110. https://doi.org/10.1016/j. using water and nano particles in diesel – biodiesel blend. J Clean Prod 2018.
fuel.2020.119110. https://doi.org/10.1016/j.jclepro.2018.10.338.
[93] Huang J, Wang Y, Qin J, Roskilly AP. Comparative study of performance and [113] Chen H, Xie B, Ma J, Chen Y. NOx emission of biodiesel compared to diesel:
emissions of a diesel engine using Chinese pistache and jatropha biodiesel. Fuel higher or lower ? Appl Therm Eng 2018;137:584–93. https://doi.org/10.1016/j.
Process Technol 2010;91:1761–7. https://doi.org/10.1016/j. applthermaleng.2018.04.022.
fuproc.2010.07.017. [114] Gnansounou E. Crops and starch-based. 2020. doi: 10.1016/B978-0-444-63585-
[94] Reksowardojo IK, Lubis IH, A WMS, Brodjonegoro TP, Soerawidjaja TH, 3.00004-8.
Arismunandar W. Performance and exhaust gas emissions of using biodiesel fuel [115] Nazir N, Setyaningsih D. Life cycle assessment of biodiesel production from palm
from physic nut (Jatropha curcas L.) oil on a direct injection diesel engine (DI) oil and jatropha Oil in Indonesia 2010:1–6.
2018:1232–6. [116] Kaewcharoensombat U, Prommetta K, Srinophakun T. Journal of the Taiwan
[95] Ganapathy T, Gakkhar RP, Murugesan K. Influence of injection timing on Institute of Chemical Engineers Life cycle assessment of biodiesel production from
performance, combustion and emission characteristics of Jatropha biodiesel jatropha. J Taiwan Inst Chem Eng 2011;42:454–62. https://doi.org/10.1016/j.
engine. Appl Energy 2011;88:4376–86. https://doi.org/10.1016/j. jtice.2010.09.008.
apenergy.2011.05.016. [117] Yang J, Cong W, Zhu Z, Miao Z, Wang Y-T, Nelles M, et al. Microwave-assisted
[96] Prabhu A, Venkata Ramanan M, Jayaprabakar J. Production, properties and one-step production of biodiesel from waste cooking oil by magnetic bifunctional
engine characteristics of Jatropha biodiesel–a review. Int J Ambient Energy 2021; SrO–ZnO/MOF catalyst. J Clean Prod 2023;395:136182. https://doi.org/
42:1810–4. https://doi.org/10.1080/01430750.2018.1557548. 10.1016/j.jclepro.2023.136182.
[97] Xue J, Grift TE, Hansen AC. Effect of biodiesel on engine performances and [118] Ruatpuia JVL, Halder G, Mohan S, Gurunathan B, Li H, Chai F, et al. Microwave-
emissions. Renew Sustain Energy Rev 2011;15:1098–116. https://doi.org/ assisted biodiesel production using ZIF-8 MOF-derived nanocatalyst: a process
10.1016/j.rser.2010.11.016. optimization, kinetics, thermodynamics and life cycle cost analysis. Energy
[98] Mo M, Masjuki HH, Kalam MA, Atabani AE. Evaluation of biodiesel blending, Convers Manag 2023;292:117418. https://doi.org/10.1016/j.
engine performance and emissions characteristics of Jatropha curcas methyl enconman.2023.117418.
ester: Malaysian perspective 2013;55. doi: 10.1016/j.energy.2013.02.059. [119] Liu Y, Zhu Z, Zhang R, Zhao X. Life cycle assessment and life cycle cost analysis of
[99] Suthar K, Dwivedi A, Joshipura M. A review on separation and purification Jatropha biodiesel production in China. Biomass Convers Biorefinery 2022.
techniques for biodiesel production with special emphasis on Jatropha oil as a https://doi.org/10.1007/s13399-022-03614-7.
feedstock. Asia-Pacific J Chem Eng 2019;14:1–19. https://doi.org/10.1002/ [120] Kumar S, Chaube A, Kumar S. Sustainability issues for promotion of Jatropha
apj.2361. biodiesel in Indian scenario: a review. Renew Sustain Energy Rev 2012;16:
[100] Atabani AE, Silitonga AS, Badruddin IA, Mahlia TMI, Masjuki HH, Mekhilef S. 1089–98. https://doi.org/10.1016/j.rser.2011.11.014.
A comprehensive review on biodiesel as an alternative energy resource and its [121] Tewari DN. Report of the committee on development of biofuel. Planning
characteristics. Renew Sustain Energy Rev 2012;16:2070–93. https://doi.org/ Commission, Government of India; 2003. http://planningcommission.nic.in/
10.1016/j.rser.2012.01.003. reports/genrep/cmtt_bio.pdf.
[101] Wakil MA, Kalam MA, Masjuki HH, Atabani AE, Rizwanul Fattah IM. Influence of [122] Singh K, Singh S, Kumar A, Kharkwal H, Avikal S. Performance, energy, emission
biodiesel blending on physicochemical properties and importance of and cost analysis of Jatropha (Jatropha curcas) oil as a biofuel for compression
mathematical model for predicting the properties of biodiesel blend. Energy ignition engine. Mater Today Proc 2021;43:348–54. https://doi.org/10.1016/j.
Convers Manag 2015;94:51–67. https://doi.org/10.1016/j. matpr.2020.11.675.
enconman.2015.01.043. [123] Yusuf NNAN, Kamarudin SK. Techno-economic analysis of biodiesel production
[102] Schobing J, Tschamber V, Brillard A, Leyssens G. Impact of biodiesel impurities from Jatropha curcas via a supercritical methanol process. Energy Convers Manag
on carbon oxidation in passive regeneration conditions: influence of the alkali 2013;75:710–7. https://doi.org/10.1016/j.enconman.2013.08.017.
metals. Appl Catal B Environ 2018;226:596–607. https://doi.org/10.1016/j. [124] Dawa A, Klemola T, Saloniemi I, Niemelä P, Vuorisalo T. Energy for Sustainable
apcatb.2017.12.011. Development Factors affecting genetic and seed yield variability of Jatropha
[103] Pikula K, Zakharenko A, Stratidakis A, Nosyrev A, Mezhuev Y, Tsatsakis A. Green curcas (L.) across the globe: a review. Energy Sustain Dev 2017. https://doi.org/
Chemistry Letters and Reviews The advances and limitations in biodiesel 10.1016/j.esd.2017.09.002.
production: feedstocks, oil extraction methods, production, and environmental [125] Raj N, Neupane P, Bahadur B, Quiroz-arita C, Manandhar S, Bradley TH.
life cycle assessment 2020. doi: 10.1080/17518253.2020.1829099. Stochastic economic and environmental footprints of biodiesel production from
[104] Saranya G, Ramachandra TV. Life cycle assessment of biodiesel from estuarine Jatropha curcas Linnaeus in the different federal states of Nepal. Renew Sustain
microalgae. Energy Convers Manag X 2020;8:100065. https://doi.org/10.1016/j. Energy Rev 2020;120:109619. https://doi.org/10.1016/j.rser.2019.109619.
ecmx.2020.100065. [126] Makepa DC, Fumhirwa DV, Tambula S, Fumhirwa DV, Tambula S. Performance
[105] Campbell PK, Beer T, Batten D. Life cycle assessment of biodiesel production from analysis, techno-economic and life cycle assessment of Jatropha curcas L.
microalgae in ponds. Bioresour Technol 2011;102:50–6. https://doi.org/ (Euphorbiaceae) seedcake gasification and Fischer- Tropsch integrated process for
10.1016/j.biortech.2010.06.048. bio-methanol production integrated process for bio-methanol production.
[106] Ali M, Akram A, Ghobadian B, Ra S, Heijungs R, Tabatabaei M. Environmental Biofuels 2024;15:57–66. https://doi.org/10.1080/17597269.2023.2216957.
impact assessment of olive pomace oil biodiesel production and consumption: a [127] Sales MB, Borges PT, Nazareno M, Filho R, Lizandra R, Castro AP, et al.
comparative lifecycle assessment 2016;106. doi: 10.1016/j.energy.2016.03.010. Sustainable feedstocks and challenges in biodiesel production: an advanced
[107] Mohammad S, Tabatabaie H, Tahami H, Murthy GS. A regional life cycle bibliometric analysis 2022.
assessment and economic analysis of camelina biodiesel production in the Pacific

16

You might also like