Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Renewable and Sustainable Energy Reviews 116 (2019) 109399

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


journal homepage: www.elsevier.com/locate/rser

Biodiesel process intensification through catalytic enhancement and T


emerging reactor designs: A critical review
Kang Yao Wonga,b, Ng Jo-Hana,b,c,∗, Cheng Tung Chongd, Su Shiung Lame, Wen Tong Chongf
a
Energy Technology Research Group, University of Southampton, Southampton, SO17 1BJ, United Kingdom
b
Faculty of Engineering and Physical Sciences, University of Southampton Malaysia (UoSM), 79200, Iskandar Puteri, Johor, Malaysia
c
UTM Centre for Low Carbon Transport in Cooperation with Imperial College London, Universiti Teknologi Malaysia, Skudai, Johor, Malaysia
d
China-UK Low Carbon College, Shanghai Jiao Tong University, Lingang, Shanghai, 201306, China
e
Pyrolysis Technology Research Group, Institute of Tropical Aquaculture and Fisheries Research, Faculty of Ocean Engineering Technology and Informatics, Universiti
Malaysia Terengganu, 21030, Kuala Nerus, Terengganu, Malaysia
f
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, Kuala Lumpur, 50603, Malaysia

A R T I C LE I N FO A B S T R A C T

Keywords: In recent years, biodiesel has proven to become an acceptable alternative source of energy, due to its attractive
Biodiesel properties such as reduced emissions, renewability, and energy sustainability. In this review, the processing
Transesterification methodologies for biodiesel are extensively evaluated, particularly the mechanisms and theories of the in-
Reactor tensification approaches. Recent techniques implemented to produce biodiesel vary but are aimed at addressing
Intensification
production challenges such as the natural mass transfer limitation of reactants, feasibility in upscaling, and ease
Process integration
of downstream processing. Key trends within the context of biodiesel processing technologies include, transitions
from batch-type to continuous reactors for transesterification, edible to non-edible biodiesel feedstocks,
homogeneous to heterogeneous catalytic transesterification, and base-type to acid catalyst driven transester-
ification. Overarching insights indicate that the focus of interests are converging towards modularity and greater
sustainability, while maintaining feasibility in commercialisation. Techno-economic analysis revealed that there
are Southeast Asian nations that have immense potential for biodiesel production, and are able to almost replace
their diesel consumption completely. Different biodiesel standards have localised context to facilitate the pre-
dominant feedstock used in the production, storage and usage. Therefore, the selection of biodiesel feedstock has
to ensure lowered emissions when replacing diesel to mitigate potential health hazards.

1. Introduction such as, premature mortality from ambient air pollution [4], labour
capacity, vector-borne diseases, and impact on food security due to
Conventional energy are usually sourced from the combustion of rising global temperature [5]. Under these treaties and agreements,
fossil fuels such as coal, crude oil, and natural gas which are both non- countries are required to meet their emissions target through national
renewable and non-sustainable. By estimation, the total world petro- measures to deal with mitigation of greenhouse gases. For example,
leum and other liquids reserves in 2018 are consumed at a rate of 36.5 recently the Paris Agreement rulebook has been finalised during the
billion barrels globally [1], and are accounted for 79.5% of the world's COP24 (Conference of Parties) organised by UNFCCC (United Nations
total energy consumption, where 29% of the total energy consumption Framework Convention on Climate Change) in Katowice, Poland, such
are used in transportation sector [2]. World energy consumption are that countries set out to keep global warming to well below 2° Celsius
projected to increase from 575 quadrillion Btu to 736 quadrillion Btu above pre-industrial levels by 2100; and to limit the increase to 1.5°
between 2015 and 2040, representing a total energy consumption rise Celsius [6].
of 28%. Most of the increases are expected to be contributed by the non- Biodiesel has always been one of the major alternative energy
OECD countries with a share of 41% in contrast to a 9% increment for sources to replace fossil fuel, due to its attractive properties such as
OECD countries [3]. Mitigation obligations such as the Kyoto Protocol reduced emissions, renewability, and sustainability. Additionally, bio-
in 1997 and the Paris Agreement in 2015, have fostered the growth of diesel can also be produced from renewable sources that are obtained
renewable energy usage due to the pressing climate-related hazards naturally, such as plant oil, animal fat, and microorganism, which


Corresponding author. Energy Technology Research Group, University of Southampton, Southampton, SO17 1BJ, United Kingdom.
E-mail address: J.Ng@soton.ac.uk (J.-H. Ng).

https://doi.org/10.1016/j.rser.2019.109399
Received 4 March 2019; Received in revised form 10 September 2019; Accepted 12 September 2019
Available online 28 September 2019
1364-0321/ © 2019 Elsevier Ltd. All rights reserved.
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Nomenclature XRD X-ray diffraction

Abbreviation Description Symbols Description Units

BSFC Brake specific fuel consumption Ca Cavitation number


BTE Brake thermal efficiency ΔP Pressure drop across cavitation device Pa
Btu British thermal unit p Local pressure Pa
CSTR Continuous-stirred tank reactor PD Power dissipated into system J/s
EGA-MS Mass spectrometry pV Vaporising pressure Pa
FAME Fatty acid methyl ester v Velocity of fluid flow m/s
FFA Free fatty acid V0 Volumetric flow rate through the cavitation device m3/s
LCA Life cycle assessment α Ratio of hole diameter on orifice to pipe diameter
PTFE Polytetrafluoroethylene β Ratio of total flow area on orifice plate to the cross-sec-
RBD Refined bleached deodorised tional flow area of pipe
RIMM Rectangular interdigital micromixer β0 Ratio of total flow area on orifice plate to the cross-sec-
RSM Response surface methodology tional flow area of pipe
SCMH Simultaneous cooling and microwave heating Ɛ′ Dielectric constant
SIMM Slit interdigital micromixer Ɛ’’ Dielectric loss
UHC Unburned hydrocarbon ρ Density kg/m3
WCO Waste cooking oil

makes it a promising fuel from a sustainability perspective. Biodiesel manufacturer's guideline. Additionally, blends which consist of
crops can also act as carbon sequestering plants, which can contribute 6%–20% biodiesel are popular choices as they represent good balance
to a net carbon emission reduction in the life-cycle of biodiesel pro- of fuel performance and cost, which complies to the prescribed stan-
duction. However, the land devoted for biofuel feedstock also restrict dards as specified by ASTM D7467 [10,11]. Blending allows biodiesel to
the availability of food producing crops, hence priority should be be combusted in an ICE with minimal or without modification, thus,
considered to use marginal soils for sustainable biodiesel production reducing the consumption and dependency on fossil oil. Studies have
[7]. Biodiesel is proven to be more environmentally friendly, due to the also shown signs of reduction in tail pipe emissions such as carbon
reduced emissions during combustion and superior biodegradability monoxides (CO), unburned hydrocarbon (UHC), and soot via usage of
when compared with fossil fuel [8,9]. biodiesel-diesel blends [12,13]. However, nitrogen oxides (NOx) emis-
Generally, biodiesel are volumetrically blended with commercial sions for biodiesel-fuelled combustions are showing mixed results, with
fossil diesel up to 20% and used in the existing internal combustion higher saturation fatty acid methyl ester (FAME) showing predisposi-
engines (ICE) without modifications, depending on the engine tion for lower NOx emissions.

Fig. 1. Crude oil prices, biodiesel research articles and patents for years 2000–2018.

2
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Currently, a collection of methodologies such as continuous stirred- December 2007 and June 2009, causing the crude oil prices to collapse.
tank, microtubular mixer, microwave technique, and cavitation in- From it, crude oil prices have fallen more than 36% over a year, fol-
tensification has been reported to enhance biodiesel production lowed by the declination of biodiesel-related patents by 10% and 16%
through transesterification of vegetable feedstocks [14]. The process is in 2010 and 2012, respectively. Similarly, biodiesel-related research fell
usually performed in the presence of a catalyst, either in homogeneous by 16% in 2012, which corresponds to a lag of 3 years from the effect of
or heterogeneous forms to improve the rate of reactions [15]. These the Great Recession. Thus, the series of cascading effects from the
catalysts can cause huge environment impacts as they are often in- global economic downturn show that scientific studies are less re-
organic bases that require downstream post-processing. To date, the sponsive to the fluctuation of crude oil economies in terms of time, as
number of catalysts discovered for biodiesel transesterification has in- compared to patent work. Another noticeable trend between the years
creased substantially since the classification of homogeneous, hetero- 2014–2016 was observed, when the average crude oil prices fell by
geneous, and enzymatic catalysts which draw attention away from approximately 50% of its original price at 2014, due to a combination
physical intensification methodologies. The main challenges in bio- of factors. This is due to the increase of supply from new sources, such
diesel transesterification remains with the mass transfer limit due to the as shale oil and shale gas, in addition to oil producers maintaining the
reactants being immiscible by nature [16]. Consequently, transester- level of crude oil production. At the same time, global demand of en-
ification process is also a reversible process which obeys the Le Châ- ergy from emerging economies showed a decrease, which led to a
telier's principle of chemical equilibrium therefore limiting the bio- global overproduction of crude oil. Hence, biodiesel-related technolo-
diesel yield. gies have plummeted immediately in tandem with crude oil prices,
Biodiesel reactors have also been tailored specifically to tackle the while the numbers of research articles remain at a steady output.
challenges mentioned above. The amounts of reactor design and bio- Current research approaches towards biodiesel production still heavily
diesel-related research have shown positive correlations with respect to relies on the usage of catalyst and type of biodiesel reactor. According
the fluctuation of crude oil prices as shown in Fig. 1. The patent search to Chuah et al. [14], new and cleaner biodiesel intensification tech-
in Fig. 1 for biodiesel reactor designs was based on the World In- nologies are in demand, which should be more environmental friendly
tellectual Property Organization (WIPO) [17], whereas the published from an operating condition perspective while maintaining high yield.
scientific articles related to biodiesel reactors are derived from the Thus, the purpose of this article is to critically review the devel-
Scopus database [18], this is also represented in Table 1. opment of current biodiesel production technologies and also to study
The number of scientific literatures covering the period of January the trend of emerging techniques for biodiesel intensification such as
2000 to December 2018 for this particular topic was 2325, while the continuous stirred tank, microchannel, cavitation, microwave, and co-
total number of patents filed from the year 2000–2018 was 2454. The solvent while assessing the novelty and maturity of these technologies.
WTI and Brent crude oil price data was compiled from the U.S. Energy The techno-economical perspective of biodiesel production on a global
Information Administration [19], while the Dubai crude oil price for the level is also addressed to highlight several high potential biodiesel
same period was source from World Bank Open Data [20]. producing country. Besides, a comprehensive analysis of biodiesel
From Fig. 1, it is observable that the average WTI, Brent and Dubai standards variations is also discussed to elucidate the differences of
crude oil prices directly influence both the patents filed and scientific specifications from the major biodiesel producing countries.
articles published, as the latter two clearly lag behind the fluctuation
trends of crude oil prices. The grey shaded area represents the Great
Recession which has led to a diminishing demand for energy between

Table 1
Sum of biodiesel reactor patent from the year 2000–2018, sourced from WIPO [17].
Sum of Patent Year 20
Countries
00 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 Grand Total

Australia 2 1 1 2 4 1 1 6 7 10 8 8 7 3 1 4 2 3 2 73
Bulgaria 1 1 1 3
Canada 1 1 1 4 5 14 14 11 9 13 6 2 3 3 2 5 94
China 1 2 2 10 58 68 84 69 76 69 32 99 66 86 80 2 804
Denmark 1 1 1 1 3 2 2 1 12
Egypt 1 1 2
France 1 1 1 3
Germany 1 3 2 2 1 1 10
India 3 2 7 13 16 9 16 16 13 12 6 6 7 8 134
Indonesia 1 1 1 1 4
Israel 1 1 1 1 4
Japan 1 2 4 9 8 7 5 5 4 1 1 2 49
Malaysia 1 1 3 1 7 9 7 9 9 7 5 3 4 3 4 73
Mexico 1 2 2 1 7 10 9 13 6 5 5 2 1 6 6 76
Portugal 1 2 1 1 5
Republic of Korea 9 9 5 14 13 4 7 18 11 10 4 2 5 111
Romania 1 2 2 5
Russian Federation 2 3 2 2 1 1 1 1 4 4 21
Singapore 1 1 1 2 2 1 8
South Africa 1 1 2 1 5
Spain 1 1 2 4 3 1 2 14
Tunisia 1 1
United Kingdom 1 1 2 2 3 1 1 1 1 13
United States 1 1 1 5 4 13 18 26 26 47 50 48 29 27 43 21 19 9 6 394
Patent Officesa 1 2 6 3 11 20 46 67 65 65 66 47 36 39 19 18 11 14 536
Grand Total 4 3 6 15 19 43 72 183 232 285 256 267 209 160 227 142 148 133 50 2454

a
Collection of patents from European Patent Office and Patent Cooperation Treaty.

3
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

2. Transesterification for biodiesel production Another study from Hoang and Le [22] compared the performance
between preheated and unheated straight Jatropha oil in a diesel engine
Transesterification is a popular approach for biodiesel production using various spraying cone angle and penetration length. The results
because it is straightforward as compared to the other production from 300 h of endurance test reviewed that carbon deposits in the in-
methods such as microemulsion of oil, pyrolysis, and fractional jector hole using SVO were considerably higher than preheated SVO,
cracking [8]. Rudolf Diesel, the pioneer of diesel engine conducted hence leading to an increase in BSFC and reduction in BTE.
experiments with straight vegetable peanut oil in diesel engine in the The aim of transesterification is to reduce the overall viscosity,
late 1890s. However, straight vegetable oil (SVO) in diesel engine is without significantly trading off the calorific value of the resultant fuel
deemed to be too viscous for combustion, which can cause huge var- [23]. Transesterification uses the triglycerides in oil or fat to react with
iation in the ignition delay, causing inconsistency during the ignition alcohols such as methanol or ethanol in the presence of a catalyst. The
cycle that leads to damage in engine. According to a recent SVO review conversion consists of three consecutive reversible reactions in series,
from Che Mat et al. [21], the direct use of SVO can lead to an increase in each further breaking down the glycerides into simpler form, forming a
brake specific fuel consumption (BSFC) and reduces brake thermal ef- methyl ester group. Thus, at the end of the transesterification, three
ficiency (BTE), when compared with diesel. The higher viscosity and molecules of methyl esters and one molecule of glycerol are produced.
density of SVO lowers the atomisation rate, which later leads to a lower In general, glycerol is produced as a by-product, however glycerol can
combustion efficiency and increases tailpipe emissions such as CO. be refined to a higher purity due to its commercial value in

Fig. 2. Conventional biodiesel production. Adapted from Refs. [26,27] with permission from Elsevier. Licence Number: 4541360938195, 4541361018788.

4
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

pharmaceutical, food, and cosmetic industries. In addition, value-added homogeneous base catalyst in transesterification seems to be sig-
fuel additives such as acetins can be produced through the valorising nificant. However, this method is still preferred on an industrial scale
process of glycerol esterification with acetic acid and acetic anhydride production, because of its consistency in production and economic
[24,25]. viability.
The summary of the chemical reaction process is shown in Reaction Table 3 summarises the homogeneous base type catalyst used in
1 [8], and the overview of transesterification production chain is shown transesterification from the year of 2000–2018. In a study by Darnoko
in Fig. 2. and Cheryan [42], transesterification of refined, bleached, and deo-
dorised (RBD) palm oil is carried out with KOH and methanol: oil ratio
CH2 − OOC − R1 CH2 − OH of 6:1 in a continuous-stirred tank reactor (CSTR), at a temperature
R1 − COO − R 4
60 °C for 90 min. The FAME concentration (wt. %) varies from 80% to
CH − OOC − R2 3R 4 OH catalyst R2 − COO − R 4 CH − OH
+ ↔ + 90% when KOH catalyst loading from 0.5 to 1.2 wt % is used, with the
Alcohol maximum FAME concentration at 1.0 wt % catalyst. In another report
R3 − COO − R 4
CH2 − OOC − R3 CH2 − OH by Dias et al. [43], transesterification of soybean, sunflower, and waste
Ester
Triglycerides Glycerol cooking oil (WCO) were studied using bases such as NaOH, KOH, and
(1) CH3ONa. However, biodiesel yield suffered a decrease from 99.4% to
below 96.5% when 1.2 wt % of catalytic loading was used for all the
mentioned catalysts, as compared to 0.8 wt %. The observations are
3. Biodiesel catalytic enhancement potentially due to the competing reaction of saponification producing
soap as unwanted by-product. On the other hand, the catalyst type
3.1. Homogeneous base transesterification causes insignificant differences for biodiesel yield when the optimised
loading was used.
FFA + OH− (RO−) → Soap + H2 O (ROHL) (2)
3.2. Homogeneous acid transesterification/esterification
Biodiesel can be produced with the aid of a base-type catalyst. Base-
catalysed transesterification methods are widely used as they are rela- The main chemical pathway of biodiesel production with base type
tively cheap, highly reactive, and effective at mild reacting conditions catalyst is through transesterification of triglycerides in vegetable oil.
[28]. However, presence of water content and free fatty acid (FFA) in However, acid type catalyst can convert feedstocks to biodiesel via es-
the feedstocks will encourage other chemical pathways such as ester- terification of free fatty acids and transesterification of triglycerides
ification and saponification [29]. Base type catalysts promote the sa- concurrently. Reactions 1 and 3 show examples of esterification che-
ponification process in biodiesel feedstock with the presence of FFA mical processes of FFA-triglycerides-methanol to produce water, ester
[28,30,31]. Reaction 2 shows the chemical pathway of saponification, and glycerol.
where FFA reacts with hydroxide ion to form soap, leading to reduction
R − COOH + R 4 OH catalyst H2 O R − COOH2 CR 4
in catalyst activity. ↔ +
FFA Alcohol Water Ester (3)
Thus, transesterification with base catalyst are prone to soap for-
mation for either high FFA content, or high catalyst loading, which The acid catalysed biodiesel production avoids the saponification
ultimately leads to a decrease in biodiesel yield. Therefore, the pre- process which leads to soap formation and complicated downstream
requisite of using base-catalysed transesterification is to ensure bio- processing, hence increasing the final yield of biodiesel and reducing
diesel feedstocks have a FFA level of less than 1 wt%, through refine- production cost [51]. This process allows cheaper feedstocks such as
ment or pre-treatment resulting in production cost increase [32,33]. waste cooking oil, which is high in FFA content to be used for biodiesel
Table 2 tabulates the base catalyst transesterification of other conversion without pre-treatment. Additionally, feedstocks from edible
published work with FFA content reported for the different feedstocks, oil such as palm, sunflower, and soybean; feedstocks from non-edible
organised by year. In recent years, the transition from the use of first oils such as Jatropha curcas, and Pongamia pinnata (Karanja) which are
generation feedstocks to second generation feedstocks due to their non- high in FFA contents (> 2%) in nature will also benefit from using acid
competing nature with food crop, also led to complications in biodiesel catalyst [52,53].
production as they are naturally high in FFA content. Thus, biodiesel Table 4 summarises the homogeneous acid type catalyst in biodiesel
and glycerine will become increasingly difficult to separate due to the transesterification. Farag et al. [54] reported using a model oil which
formation of soap as a by-product [34]. consist of 50% sunflower oil and 50% soybean oil to simulate the direct
Besides, an additional separation process is required to remove the esterification of FFA. The effect of different acid catalyst type was found
homogeneous base from the end product, usually through neutralisa- to be significant, with the percentage of FFA conversion in the order of
tion of acids where trace amount of water is produced. This causes H2SO4 > HCl > AlCl3 > FeSO4 > SnCl·H2O. In the optimum condi-
problems to the purification process. The drawbacks of using a tions, a 96.6% FFA conversion can be achieved at 6:1 methanol molar

Table 2
FFA content in different feedstocks for homogeneous base type transesterification. Adapted from Ref. [28] with permission from Elsevier. Licence Number:
4541360793123.
Feedstocks Feedstock Generation FFA content (wt. %) Reference Year

Soybean First <1 Freedman et al. [35] 1984


Edible oil and fats First < 0.05 Ma and Hanna [36] 1999
Review of edible oil and animal fat First <1 Canakci and Van Gerpen [37] 2001
Review < 0.5 Zhang et al. [38] 2003
Rubber seed oil Second 2 Ramadhas et al. [32] 2005
Jatropha oil Second <1 Kumar Tiwari et al. [39] 2007
Polanga oil Second ≤2 Sahoo et al. [33] 2007
Karanja oil Second ≤3 Sharma and Singh [23,40] 2008
Review ≤2 Lam et al. [28] 2010
Polanga oil Second ≤2 Ong et al. [41] 2014

5
K.Y. Wong, et al.

Table 3
Summary of the homogeneous base catalysts used for biodiesel transesterification (2000–2018).
Feedstock Reactor parameters Oil quantity Catalyst Loading (wt. %) Alcohol Alcohol: Oil molar Temperature (°C) Time (min) Yield (%) Ref. Year
ratio

Palm oil Magnetic stirrer 500 g KOH 1.0 Methanol 6:1 60 90 97.0 [42] 2000
Palm oil Static mixer at 350 rpm – CH3ONa 0.125a Methanol 10:1 70 1 99.0 [44] 2004
Palm oil Static mixer at 300 rpm – NaOH 0.2a Methanol 6:1 70 1 99.0 [44] 2004
Rapeseed oil Static mixer at 600 rpm 500 g CH3OK 1.0 Methanol 6:1 65 120 96.0 [45] 2008
Rapeseed oil Static mixer at 600 rpm 500 g CH3ONa 1.0 Methanol 6:1 65 120 73.0 [45] 2008
Soybean oil Magnetic stirrer 200 g CH3ONa 0.4 Methanol 6:1 60 60 93.0 [43] 2008
Sunflower oil Magnetic stirrer 200 g CH3ONa 0.4 Methanol 6:1 60 60 95.0 [43] 2008
Waste cooking oil Magnetic stirrer 200 g CH3ONa 0.4 Methanol 6:1 60 60 92.0 [43] 2008
Rapeseed oil Static mixer at 600 rpm 500 g KOH 1.0 Methanol 6:1 65 120 96.0 [45] 2008
Soybean oil Magnetic stirrer KOH 0.4 Methanol 6:1 60 60 94.0 [43] 2008
Sunflower oil Magnetic stirrer 200 g KOH 1.0 Methanol 6:1 60 60 93.0 [43] 2008

6
Waste cooking oil Magnetic stirrer 200 g KOH 0.8 Methanol 6:1 60 60 90.0 [43] 2008
Rapeseed oil Static mixer at 600 rpm 500 g NaOH 1.0 Methanol 6:1 65 120 82.0 [45] 2008
Soybean oil Magnetic stirrer 200 g NaOH 0.4 Methanol 6:1 60 60 97.0 [43] 2008
Sunflower oil Magnetic stirrer 200 g NaOH 0.4 Methanol 6:1 60 60 96.0 [43] 2008
Waste cooking oil Magnetic stirrer 200 g NaOH 1.0 Methanol 6:1 60 60 90.0 [43] 2008
Duck tallow Static stirrer at 600 rpm – KOH 1.0 Methanol 6:1 65 180 97.0 [46] 2009
Palm oil Constant total volume, with magnetic stirrer 303 ml KOH 8.5 Dimethyl Carbonate 9:1 75 480 96.2 [47] 2010
(DMC)
Waste cooking oil Magnetic stirrer at 600 rpm – NaOH 0.5 Methanol 7.5:1 50 30 96.0 [48] 2012
Fish oil Magnetic stirrer at 600 rpm 50 g KOH 0.5 Methanol 6:1 32 60 96.0 [49] 2013
Rapeseed oil Preheated oil with magnetic starrier at 600 rpm 6g KOH 1.0 Methanol 6:1 60 60 91.0 [30] 2014
Palm oil Magnetic stirring at 600 rpm – CH3ONa 0.32 Methanol 5.48:1 55 40 85.0 [50] 2017
Waste rapeseed oil 5.5% FFA and 3 wt % Water content with static mixer at 40 ml CH3ONa 3.0 Methanol 18:1 60 5 97.0 [31] 2018
600 rpm

a
Catalyst loading in mol/kg.
Renewable and Sustainable Energy Reviews 116 (2019) 109399
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

ratio, temperature of 60 °C, H2SO4 catalytic loading at 2.5% w/w,

2008

2008

2010

2010

2010

2010

2013
Year
agitation speed at 300 rpm, and reaction time of 60 min. The optimised
conditions were replicated on home and restaurant cooking oil to
[55]

[55]

[54]

[54]

[54]

[56]

[57]
Ref

compare the FFA conversion.


Aranda et al. [55] compared the use of four different homogeneous
type acid catalysts, with methanesulfonic (CH4O3S) and sulfuric acid,
FFA Conversion

(H2SO4) outperforming the rest in the esterification of palm oil by ap-


proximately 60% in yield difference. Although a 3:1 methanol molar
ratio was implemented, a conversion higher than 90% was able to be
90.0

90.0

96.6

87.9

88.0

95.0

80.6
(%)

achieved, due to higher FFA proportion in palm oil, and esterification


being the primary chemical reaction. Non-catalytic conditions were also
compared, where a small amount of H2SO4 catalyst at 0.01 wt %, was
(min)
Time

180

180

180

312
able to promote from 30% to more than 70%.
60

60

90
Temperature (°C)

3.3. Heterogeneous base transesterification

Heterogeneous catalysts are viable candidates to replace homo-


geneous-type catalysts, as they are potentially more sustainable than
130

130

60

60

60

65

87

the aforementioned homogeneous catalysts [58]. However, the post-


processing of homogeneous catalyst transesterification/esterification
Alcohol: Oil molar

involves considerable amount of energy during the purification stage,


owing to the need for catalyst removal [59]. Nonetheless, the disposal
of wastewater generated have augmented the overall processing ex-
penses [60]. Additionally, the catalysts often diffuses into the glycerol
ratio

15:1
3:1

3:1

6:1

6:1

6:1

9:1

layer, reducing the quality of glycerol and reusability of catalyst.


As a result, heterogeneous catalysts are developed to take advantage
Methanol

Methanol

Methanol

Methanol

Methanol

Methanol

Methanol
Alcohol

of the ease of separation and reusability. The base solid catalysts are
generally derived from the alkaline earth metals. The catalyst are
oxides of elements such as Mg, Ca, Sr, and Ba which exhibit strong basic
properties at their active sites, when used as a catalyst in biodiesel
Loading (wt.

production. For example, the CaO alkali solid catalyst can be found in
natural occurring resources, such as limestones, egg shells, and oyster
0.063a
0.1

1.0

2.5

2.5

2.5

0.5

shells which makes it a primary candidate in heterogeneous catalysed


%)

transesterification due to its cost effectiveness [61,62].


Table 5 summarises studies on heterogeneous basic type catalyst for
1-butyl-3-methylimidazolium hydrogen sulfate

biodiesel transesterification. Granados et al. [63], studied the dete-


rioration of catalytic performance via air contact and reusability of
Summary of the homogeneous acid catalysts used for biodiesel transesterification (2008–2013).

catalyst, through characterisation technique such as X-ray diffraction


(XRD), and evolved gas analysis by mass spectrometry (EGA-MS). The
CaO catalyst was reported to be sensitive to H2O and CO2 in the at-
mosphere, and results indicate that fresh solid CaO are susceptible to
hydration to form Ca(OH)2 and carbonation to form CaCO3, reducing
the catalytic activity of CaO in transesterification. As a result, evacua-
tion of hydrated and carbonated CaO at elevated temperature is able to
(BMIM HSO4)

reactivate the catalyst, hence improving the biodiesel yield in the


transesterification process. The CaO catalyst treated at 700 °C is able to
Catalyst

CH4O3S
H2SO4

H2SO4

H2SO4

achieve a yield of 94% for biodiesel, and retains an 81% conversion


AlCl3
HCl

after eight consecutive uses. However, it was discovered that active


species from solid catalyst has a tendency to leach into the reactants,
Oil Quantity

resulting in a performance decline after several transesterification cy-


15.8 ml

cles.
307 g

307 g

100 g

100 g

100 g

Another improved version of the CaO catalyst was done by co-pre-


cipitation method of a lanthanum-based calcium oxide (CaO–La2O3)


Constant stirring rate

[72], resulting in a mixed solid oxide which was demonstrated to be


Reactor parameters

feasible for solid base transesterification, using crude jatropha oil as a


Catalyst loading in terms of mol.
Static mixer at

Static mixer at

Static mixer at

Static mixer at

Static mixer at

Static mixer at

feedstock. Catalytic performance of the Ca–La catalyst was measured


through studying the reaction conditions, which enable the results to
500 rpm

500 rpm

300 rpm

300 rpm

300 rpm

360 rpm

characterise the physico-chemical properties of the catalyst. Results of


the transesterification activity using different Ca:La atomic ratios of the
CaO–La2O3 catalyst showed that pure LaO2O3 catalyst has zero effect on
Palm/rubber seed oil
Sunflower/soybean

Sunflower/soybean

Sunflower/soybean

transesterification. However, further increment of CaO loading from


2 wt % to 4 wt % showed that the CaO active phase enhances FAME
yield from 52% to 82% due to an increase in basicity, which is directly
Oleic acid
Feedstock

Palm oil

Palm oil

correlated to catalyst activity. Subsequently, the reusability and re-


Table 4

oil

oil

oil

generation of catalyst can be done by a simpler alkali solution washing,


a

without significant leaching of catalyst, thus, retaining majority of its

7
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

catalytic performance.

2003
2003

2007
2007
2008
2011
2011
2011
2014

2018

2019

2019
Year
Nanomaterials have also gained traction due to its unique properties
for enhanced biodiesel production. Metal oxide nano-catalyst such as
[64]
[64]

[63]
[63]
[65]
[66]
[66]
[67]
[68]

[69]

[70]

[71]
Ref

titanium dioxide, calcium oxide, magnesium oxide, and strontium oxide


are able to be recovered easily from the reactants and be reused, for
Yield %

> 90.0

98.03
82.8
98.2

60.0

93.0
97.0
99.0
81.0
94.7

98.0

98.0
more economical biodiesel production by exploiting their magnetic
properties [73]. Besides, some nano-particles are also biocompatible,
which makes them possible to extract microalgae oil without destroying
(min)

the cells due to their highly selective mesoporous structure that absorbs
Time

0.67
0.67
120
120

120

480

180
90
90

81

60
lipid [74].
2
Temperature (°C)

3.4. Heterogeneous acid transesterification

Despite all of the advantages of using homogeneous-type catalysts in


73.6
190
190

190

120
transesterification, it is complicated and costly to separate and recover
60
60

60
60

60

60

the liquid catalysts when they form a single phase with the end pro-
Alcohol: Oil molar

ducts. On the contrary, solid acid catalysts are superior, as compared to


homogeneous catalysts and base type catalysts. This is because they are
easier to be separated from the end products, which leads to better
reusability and recovery of the solid catalyst, hence effectively lowering
3.7:1
3.7:1
ratio

20:3
20:3

13:1
13:1
12:1

15:1

10:1
6:1

8:1

9:1

the processing cost [75]. However, heterogeneous acid transesterifica-


tion do suffer from poorer catalytic performance, hence higher tem-
Methanol
Methanol

Methanol
Methanol
Methanol
Methanol
Methanol
Methanol
Methanol

Methanol

Methanol

Methanol
Alcohol

perature and pressure reaction condition are often required [76].


There are several prerequisites for the selection of a strong hetero-
geneous acid catalyst. These are properties that favour the esterification
Loading (wt.

process, as tabulated in Table 6. Examples of typical heterogeneous acid


catalyst are shown in Fig. 3.
1.84
1.84

0.75

Table 7 summarises the experimental studies using heterogeneous


0.8

5.5

2.5

6.0

2.5
%)

3
3

1
1

acid in the transesterification process. Sulfated metal oxide is one of the


most popular solid acid catalysts, as it is classified as a superacid which
He gas at 900 °C

heavily favours the transesterification process. Superacids of such are


550 °C for 5 h
PDMS-PEO,
Calcination

more environmentally friendly, when used for esterification of biodiesel


700 °C -
Method

and can be obtained easily [77]. Zirconium oxide (ZrO2) as a support


500 °C

500 °C

700 °C

for catalytic scaffolding, has also been studied under the sulfated metal

oxide group. Positive influence was observed from its acidic catalytic
Sodium modified fluorapatite

behaviour, with simultaneous oxidising and reducing properties on the


Magnetic bamboo charcoal
Summary of the heterogeneous basic catalysts used for biodiesel transesterification (2003–2018).

surfaces [78]. Zirconium oxides are generally impregnated with sulfuric


Activated carbon/CaO

acid to be converted into sulfated zirconia, SO42−/ZrO2. Alternatively,


ZnO nanoparticles

the surface acidity can also be modified through the introduction of


(K/BC-Fe2O3)

other metal oxides such as sulfate or tungstate, to provide greater cat-


(Na/FAP)

alytic activity.
Catalyst

CaCO3
t-MgO

Patel et al. [79] investigated the synthesis of sulfated zirconia and


MgO

CaO
CaO

BaO
SrO
SrO

the biodiesel conversion of oleic acid for the esterification of FFA.


Catalytic loading of sulfated zirconia has been observed to have a po-
Oil Quantity

sitive correlation with biodiesel yield. For the concentration of 1.0 N


100 ml

SO42−/ZrO2 (0.1 g), an initial yield of 18% was achieved. Using the
50 g
50 g

15 g
15 g

same operation condition, further improvement on the catalyst was



done by using a second stage calcination process at 600 °C for an ex-


tended 4 h, which improved the esterification yield to 40%. However,
Microwave (2.45 GHz) and Ultrasound

the reaction rates of the esterification process is considered to be re-


irradiations (25 kHz) (100W/100W)

latively poor. The esterification process was optimised through para-


metric study of reaction time and catalyst loading, which leads to an
Continuous stirring at 700 rpm
Magnetic stirring at 400 rpm

Magnetic stirrer at 600 rpm

optimised operating condition of 0.5 g catalytic loading at 12 h reaction


Static mixer at max speed
Static mixer at 1000 rpm
Static mixer at 1000 rpm

time, achieving 90% yield. The catalyst is then regenerated to de-


termine its reusability in the esterification process. A reduction of 3%
Reactor parameters

Microwave 1100 W
Microwave 1100 W

yield for one-cycle esterification on oleic acid was observed. In all,


Not specified
Not specified

catalytic activity for heterogeneous acid catalysts are still relatively


Static mixer

weaker when compared with its homogeneous counterparts.


Doyle et al. [84] proposed the synthesis of zeolite catalysts with
kaolin clay, which can be used in the esterification process. This is a
favourable catalyst candidate because of its low cost, and also structural
Waste cooking oil

Waste cooking oil

Mesua ferrea oil

advantage of having both mesoporous [91] and microporous [92] fra-


Sunflower oil
Sunflower oil

Rapeseed oil
Soybean oil
Soybean oil

Soybean oil

meworks. Zeolite Y was prepared as a scaffold, and then impregnated


Canola oil
Canola oil
Feedstock

Palm oil

with ammonium nitrate to form HY zeolite to be used as a catalyst for


Table 5

esterification. A maximum yield of 85% was achieved by using an op-


erating condition of 1 h reaction time, 6:1 M ratio of ethanol/oleic acid,

8
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Table 6
Criteria, advantage, and disadvantage of an ideal solid acid catalyst.
Criteria of an ideal solid acid catalyst Advantages of an ideal solid acid catalyst Disadvantages of an ideal solid acid catalyst

• Strong Brønsted acid group that are highly


polarised hydroxyl group, at the surface of
• Insensitive to free fatty acid and moisture which
enables the use of a lower grade oil.
• Catalytic activity of solid base is greater than solid acid,
hence operating conditions such as reaction
catalyst. temperature and pressure for using a solid acid catalyst
are usually higher.
• Strong Lewis acid group which are actively
promoting transesterification, provided that the
• transesterification
One-step process where esterification and
can simultaneously takes place.
• The tendency of acid leaching into the end product,
particularly the sulfate group, H2SO4, into reactant,
active sites are not poisoned by water. promoting homogenous catalytic activity that interferes
with the measurement.
• Unique
texture.
interconnected large pore structure and • Eliminates the need of washing for the biodiesel,
while reducing potential corrosion issue downstream
of production.
• A hydrophobic catalytic surface. • Easier separation of catalyst with the end product,
hence more efficient for the recovery process of
catalyst that also allows for extensive reusability.
• For the continuous flow-type reactor, the end
products separation and purification costs can be
minimised.
• Avoid
sites.
deactivation, poisoning and leaching of acid

and 70 °C reaction temperature. The reusability study of the catalyst that are proven to be effective in biodiesel production, due to their
showed a decrease in conversion from 85% to 77% after a single cycle. strong cross-linked structures which exhibit acidic active sites that are
The catalytic performance reduction is speculated to be related to the used in esterification and transesterification reactions [93,94]. Ad-
adsorption of water at the active site of the catalyst. The esterification ditionally, these resins swell under the presence of polar components,
process using the HY zeolite produced from kaolin clay was then leading to a change in the magnitude of cross-links that will determine
compared against a commercially sourced HY zeolite catalyst. The HY the rigidity of the resins [85]. High swelling capacity is also correlated
kaolin zeolite was able to perform better at the first hour of the reac- to lower crosslinking [90,91], which indicates an improvement in the
tion, but the conversion reduces to the same as compared to the com- surface area and pore diameter of the resins. This will result in a higher
mercial sample after 2 h. The unique low Si/Al ratio of HY kaolin density of acid sites for the resin to facilitate the esterification process.
zeolite catalyst contributed to a higher catalytic activity at an earlier
stage of free fatty acid esterification, due to the higher density of acid
sites and greater Brønsted acid strength.
Ion-exchange resin catalysts are also heterogeneous acid catalysts

Fig. 3. Typical heterogeneous acid catalyst families for biodiesel esterification and transesterification.

9
K.Y. Wong, et al.

Table 7
Summary of the heterogeneous acid catalysts used for biodiesel transesterification (2008–2018).
Feedstock Reactor parameters Oil quantity Catalyst Catalyst preparation Loading Alcohol Alcohol: Oil Temperature (°C) Time (hr) Yield % Ref Year
(ml) (wt. %) molar ratio

Waste cooking oil High temperature and – Sulfated zirconia Zirconium hydroxide powder with H2SO4, dried at 3.0 Methanol 9:1 120 4 93.6 [78] 2008
pressure in an (SO42−/ZrO2) 110 °C then calcined at 600 °C.
autoclave
Oleic acid Batch reactor with 100 Sulfated zirconia Hydrated zirconia with aqueous ammonia solution 0.5a Methanol 40:1 60 12 90.0 [79] 2013
magnetic stirrer (SO42−/ZrO2) dried at 100 °C, then calcined at 600 °C for 4 h.
Waste cooking oil Stainless steel batch 300 Sulfated tin oxide Amorphous tin oxide in H2SO4 solution for 6 h, 3.0 Methanol 15:1 150 3 92.3 [80] 2009
reactor with magnetic SO42−/SnO2) then calcined at 300 °C for 2 h.
stirrer
Soybean oil Sealed capillary tubes 0.3 ETS-10 zeolite Calcined at 500 °C in flowing air at a heating rate of 0.03a Methanol 6:1 120 24 94.6 [81] 2004
in preheated furnace 4 °C/min and held for 10 h.
Sunflower oil Mechanical stirrer 100 Fly-ash Na-X zeolite Ion exchange with potassium acetate, dried at 3.0 Methanol 6:1 65 8 83.5 [82] 2012
110 °C for 2 h, then calcined at 500 °C for 2 h.
RBD Palm oil Continuous stirring at 125 Natural zeolite Zeolite impregnation with KOH for 24 h, dried at 3.0 Methanol 7:1 60 4 95.01 [83] 2013
500 rpm 110 °C for 24 h, calcined at 450 °C for 4 h.
Oleic acid Batch reactor with 50 HY kaolin zeolite NaY zeolite in ammonium nitrate at 100 °C for 4 h, 5.0 Ethanol 6:1 70 1 85.0 [84] 2016
continuous stirring dried at 100 °C for 6 h, stirred in oxalic acid for 8 h,

10
dried again at 100 °C, then calcined at 550 °C for
5 h.
Waste cooking oil Magnetic stirrer at – Amberlyst-15 (ion-exchange Wet form resins dried in oven for 12 h at 110 °C. 2.0 Methanol 20b 60 3 45.7%c [85] 2008
1500 rpm resins) Amberlyst-15 with high crosslinking, catalyst
surface area of 53 m2/g, 33% porosity, average
pore diameter of 30 nm.
Rapeseed oil/ Sealed pressure – ST-DVB (cation-exchange Resin ball swelled in dichloroethane at 60 °C, then 10.0 Methanol 15:1 100 – 97.8%c [86] 2015
oleic acid mix endurance tubes with resin) mixed with H2SO4. Catalyst surface area of
stirring 185 m2/g, average pore diameter of 9.7 nm.
Waste cooking oil Mechanical stirrer SO4/Fe–Al–TiO2 Stepwise deposition of alumina and iron oxides on 3.0 Methanol 10:1 90 2.5 96.0 [87] 2018
commercial TiO2 NPs.
Olive oil Mechanical stirrer at Activated bentonite Bentonite with size of about 130 μm activated with 5.0 Methanol 15:1 62 27 67.0 [88] 2018
450 rpm 10 N H2SO4 for 2 h.
Palmitic acid Magnetic stirring 1.0 g Zirconia-supported Sol-gel method followed by hydrothermal 30.0 Methanol 15.0 ml 60 6 > 90.0 [89] 2018
tungstophosphoric treatment.
heteropolyacid
Oleic acid – – F−-SO42-/MWCNTs 1.0 g MWCNTs and 2 g NaF was dissolved first in 0.9 Methanol 12:1 65 6 90.0 [90] 2019
100 mL 0.1 M H2SO4 under constant stirring

a
Catalyst loading in grams.
b
Percentage of volume (vol. %).
c
Percentage in terms of oil conversion.
Renewable and Sustainable Energy Reviews 116 (2019) 109399
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

4. Biodiesel reactors and technology advancements

2000

2008

2009

2017
Year
4.1. Continuous stirred tank reactor (CSTR)

[95]

[96]

[97]

[98]
Ref
Continuous stirred tank reactor (CSTR) is an extension of the batch-

Yield (%)
type reactors, which has a similar fundamental agitation mechanism.

97.3

97.6

83.7

81.0
Batch processing is one of the most common modes of biodiesel pro-
duction, which enables biodiesel to be produced in a single batch over a
period of time [61–63,71]. However, through the use of CSTR, con-

Reaction time
tinuous production of biodiesel can be achievable using a simple setup.
Nevertheless, large scale production of biodiesel that utilises batch

(min)
processing technique requires large reactor volumes, which increases

275
60

20
capital investment, making it less economically viable. Furthermore,

6
the batch-to-batch variation is also a common issue in biodiesel batch

Temperature (°C)
processing, since the quality of biodiesel tends to differ slightly for
every processing batches. The CSTR technology is straightforward to
scale up as compared to other types of reactors. All variation of CSTR
shares similar concepts of utilising physical disruption through me-

60

60

60

30
chanical stirring. The mechanical stirrer is physically attached to a
motor externally to provide adequate mixing and energy input to the

Alcohol: Oil molar


system. The reactants and products are simultaneously added and re-
moved, to maintain a steady-state continuous system in a CSTR. Table 8
summarises the continuous stirred tank system, with varying designs

ratio
and operating conditions.

c
6:1

6:1

3:1

0.6
Typically, CSTR involves more than a single stage, where the con-
figuration of the system consists of a reactor and a phase separator as
shown in Fig. 4. In the beginning stage of the reactor, alcohol and tri-

Methanol-buffer
glycerides are allowed to be partially converted into biodiesel. This is
then followed by a glycerol removal step for the reacted mixture in the
Methanol

Methanol

Methanol

solution
Alcohol

second stage of CSTR, which is used to promote the forward reaction of


transesterification through shifting of the chemical equilibrium. The
single-stage reactor is able to ensure the complete conversion of oil
Catalytic Loading (wt.

through the highly efficient production and separation system.


Darnoko and Cheryan [95] produced biodiesel through a two-stage
CSTR system with RBD palm oil. The second stage of the prototype
developed is a separator which is used for glycerol removal. The op-
eration of the CSTR requires the batch mode to be enabled over a re-
1.7a

sidence time of 40 min during the initialisation. This is then followed by


%)

40
1

activating the feed pumps to deliver the feedstocks and methanol at a


consistent rate. Another feed pump is then used to move the mixture to
the separator, where glycerol is later removed through phase separa-
Immobilized lipase

tion. During the initialisation stage, 82.7% biodiesel yield was


achieved. However, the final yield suffered a significant drop, only
achieving a 58.8% yield after a steady-state of five volume replacement
Catalyst

enzyme
H2SO4
NaOH

cycles in the continuous mode. Optimisation was also performed by


KOH
Summary of continuous stirred tank reactor for biodiesel production.

changing the residence time during batch mode, and increasing the
number of volume replacements to allow better stability in the con-
CSTR with 6-stages at 300 rpm and production rate of

Single stage CSTR at 800 rpm, oil feed at 1.8 mL/min


CSTR with 4-stages (303–706 rpm) and production
CSTR batch processing, using 500 g oil with 20 vol

tinuous transesterification system. A maximum yield of 97.3% is


achieved in steady-state, after using the optimal conditions of 20 vol
Catalyst loading in percentage of volume (vol. %).

replacement at a residence time of 60 min, 6:1 methanol to oil molar


ratio, using a 1 wt % catalyst loading, and 60 °C reaction temperature.
Prateepchaikul et al. [97] used a CSTR system with a series of four
Alcohol feed at 0.6 ml/min continuously.

CSTR reactors stacked together as shown in Fig. 5. The stacked CSTR


has a smaller reactor footprint as compared to the CSTR in multiple
stages. They are also relatively easier to be integrated into the feed
pump system, since pressure drop is reduced as compared to a multiple
Reactor parameters

stage CSTR. Raw, un-degummed, mixed crude palm oil which contains
Catalyst loading in grams.
rate of 8.3 L/h.

8–12% FFA was used as feedstock for the acid-catalysed esterification


replacement.

process. As significant amount of water is produced as a by-product


17.3 L/h

which interferes with the conversion of biodiesel, gravity separation is


required to remove the water content. The esterified oil and acidified
water obtained from the gravity separator are found to be at 83.7 and
Sunflower oil

16.3 vol %, respectively. A centrifugal separator can be employed to


Feedstock

Palm oil

Palm oil

Palm oil

significantly improve the separation process.


Table 8

b
a

11
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 4. Schematic drawing of a single reactor CSTR. Adapted from Ref. [95].

operated under milder conditions, since they have superior rates of


reaction, which reduce the energy requirement for the transesterifica-
tion process.
Generally, microchannel reactors have relatively smaller footprints,
as compared to other types of biodiesel reactors, which gives rise to its
advantage of lower capital, reduced pumping cost for biodiesel pro-
duction [101]. Microchannel reactors utilise microfluidic devices, such
as a micromixer for its agitation process. Since most of the chemical
reaction for microreactors involve mixing of two or more fluids, the use
of a micromixer can enhance the mixing efficiency at microscale [102].
Micromixer can be either active or static type mixer of various geo-
metries. Active mixing actuators are usually powered externally by
electric or magnetic excitation, whereby static mixers involve passive
mixing through the use of pressure differences. Both type of mixers
utilise different mixing principles, defined by their operations at dif-
ferent ranges of production capacities and mixing speeds [103]. Two
types of intensification can occur in the micromixer [103], (i) hetero-
geneous mixing created by convection, (ii) homogeneous mixing at the
molecular level which is driven by diffusion of adjacent domains. The
common types of micromixer are namely the T-junction, J-junction, Y-
Fig. 5. Schematic of CSTR impellers in series. Reprinted and adapted from Ref. junction, and cross-junction as shown in Fig. 6. These mixers are used to
[97] with permission from Elsevier. Licence Number: 4541360505466. direct the flow of reactants into the main reacting channel, where
mixing will occur through the aforementioned intensification methods.
4.2. Microchannel reactor A summary of microreactor research using various feedstock accom-
panied with their performances is tabulated in Table 9.
Microchannel reactors or sometimes as known as microreactors, are Sun et al. [105] studied four different types of static micromixers,
designed to overcome the issues of low contact surface-to-volume ratio namely, a T-junction, J-junction, rectangular interdigital micromixer
of reactants [99]. Microreactors are intensively researched and can be (RIMM), and a slit interdigital micromixer (SIMM). Interdigital type
categorised as, microtubular reactors, zig-zag microchannel reactors, micromixers have alternating feed channels to periodically create liquid
etched microchannel reactors, and microfluidic laminated lab chip. lamellae, which are used to study the hydrodynamics for mixing two
Microchannel reactors utilise a meso or micro scale geometry for their different liquid streams [111]. In their work, the outlet of the micro-
reacting channels, which can have different hydraulic diameters in the mixer are connected to a transparent polytetrafluoroethylene (PTFE)
range of 10–1000 μm. The narrow channel enables a high surface-to- tube of 1.2 mm internal diameter, to observe the formation of oil dro-
volume ratio transesterification process, which effectively decreases the plets and flow patterns. Stainless steel Dixon rings, are filter rings with
diffusion length of reactants, hence increasing the mass transfer to massive surface-to-volume ratio were packed in the reacting channel to
promote reaction rates. Microreactors typically have approximately a avoid the phase separation during the transesterification process, as
10000–50000 m2/m3 surface-to-volume ratio, which are quoted to be shown in Fig. 7. Overall, the multilamination micromixer, RIMM and
10–50 and 100–500 times greater when compared with conventional SIMM are superior micromixers which can achieve approximately
laboratory vessels of 1000 m2/m3 and production vessel of 100 m2/m3, double in the biodiesel yield when compared with than the T- and J-
respectively [100]. Thus, microchannel reactors are often selectively junction passive micromixers. The result shows significant

12
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 6. Micromixers with (a) T-junction, (b) J-junction, (c) Y-junction, (d) cross-junction.

improvement to biodiesel yield with the use of the interdigital micro- Many researchers have reported a decrease in reaction rates when
mixers, mainly because of the enhanced contact area between the me- higher operating temperatures of 70 °C and above for biodiesel pro-
thanol and oil. The optimised operating condition at operating condi- duction were implemented. This is because the reaction temperature
tions of 8:1 methanol to oil molar ratio, with a residence time of 17 s, used is above the methanol boiling point of 64.7 °C. This consequently
under reaction temperature of 70 °C, and a biodiesel throughput of results in creating a gas-liquid mixture within the reacting channel,
10 mL/min combined with the Dixon rings-packed PTFE tube, allows a forming a layer of vapour that insulates the oil-alcohol interface. Thus,
biodiesel yield of 99.5% to be achieved. the specific interfacial surface reduces and therefore biodiesel yield
The larger interfacial surface area allows the feedstock oil and al- decreases. The results are in accordance with from past research
cohol, which are both immiscible in nature to mix homogeneously, as [104,105,108,109]. However, Sun et al. [105] observed the pressure
oil droplets are formed in the stream of alcohol shown in Fig. 8. The drop in the reaction system, suggesting a change in the methanol
degree of mixing and size of droplets formed are highly dependent on boiling point as calculated by the Antoine equation. Therefore, this
the angle of impact from the micromixer. For instance, T-junction, J- allows a higher reacting temperature to be used without completely
junction, and Y-junction micromixer have an angle of impact of 180°, boiling the methanol. This also suggest that the induced pressure drop
90°, and less than 90°, respectively. Another important factor which can have a significant effect on improving the rates of reaction, for the
controls the droplet size in microreactors is the residence time, which is limited operating temperature of microreactors which tends to be lower
defined as the average length of time that the reactants spend in the than the boiling point of alcohol used.
microchannel, from the beginning of the micromixer to the output. The Further intensification can be done by packing the microchannel
residence time is usually controlled by varying the channel length and with device that can either improve the passive mixing or increase the
also combined flow rates of oil and alcohol. As the channel length in- surface-to-volume ratio to allow greater mass transfer. Aghel et al.
creases, the mixture flow will suffer from a pressure drop, hence ef- [107] used a wire coil insert to change the geometric properties of the
fectively decreasing the flow rates and mixing potential. reactor, which performs the same role of a micromixer. This is done to
A study conducted by Kaewchada et al. [108] shows the effects of provide more convectional mixing throughout the reacting channel as
micromixer types and residence time on the biodiesel yield. Two types compared to using micromixers. Wire coils of different geometries were
of static micromixers were tested namely, the T- and J-junction, where used, with parameters such as wire coil length and wire pitch length, as
the T-junction outperforms the J type for the range of 5–20 s. The flow shown in Fig. 10. Response surface methodology (RSM) is used to op-
rate is observed to decrease when channel length is kept constant, timise the independent operating conditions for the transesterification
which resulted in pressure drop across the reaction channel, leading to process. The largest improvement was observed when the methanol to
a significant drop in biodiesel yield from 97.14 to 91.43%. On the other oil molar ratio was increased from 6 to 9, where the wire coil micro-
hand, prolonged residence time can cause the droplet size to increase, reactor is capable of achieving a 94% biodiesel yield, as compared to a
therefore effectively allowing less interfacial area for the oil and me- biodiesel yield of less than 90% without the wire coil. However, when
thanol during mixing. Fig. 9 shows the oil droplet size for two types of excessive methanol to oil molar ratio of 12:1 is used, the yield decreases
micromixer at various residence times. The optimal biodiesel yield of due to the reversible reaction of transesterification. Overall, the wire
97.14% is achieved with catalyst loading of 1 wt %, operating at re- coil insert demonstrated an improvement over non-wire coil, where the
action temperature of 60 °C, using methanol to oil molar ratio of 6:1, use of a long wire coil (30 cm) with a shorter pitch length (0.5 mm)
and a residence time of 5s. improved the biodiesel yield by 6%.

13
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

4.3. Cavitation reactor

2009
2010
2013
2014
2016
2017
2019
Year

[104] The discovery of the cavitation phenomena can be traced back to


[105]
[106]
[107]
[108]
[109]
[110]
1754, first mentioned by Leonhard Euler in the theory of turbines.
Ref

However, the use of cavitation only started to gain attention after ap-
Yield (%)

proximately a century later. Cavitation occurs in flowing liquid systems


91.76
95.41

98.26
97.3
94.8

97.5
99.0
when the liquid experiences a sudden change of pressure to below its
vapour pressure, where bubble cavities are formed and quickly im-
Residence time (s)

plodes as it reaches a high pressure area [113]. The collapsing cavities


releases enormous amount of energy, causing high local temperature
and local pressure of few thousand Kelvin and atmospheres, respec-
tively [114]. Cavitation are used not just in application such as che-
180
534
120

mical reactor design, but also found in submerged rotating propeller of


18
44
5
5

ships, control valves, pumps, which causes significant wear damages.


Temperature (°C)

There are four different types of cavitations, namely, acoustic, hydro-


dynamic, optical and particle [115]. Fig. 11 summarises each of the
categories and the cause of the phenomena. Sonification and hydro-
dynamic cavitations are preferred for chemical reactor application due
62.4
56
70
65
65
60
65

to its favourable intensification effect [116].


Alcohol: Oil molar ratio

4.3.1. Sonification
Sonochemical reactors can be classified into indirect and direct
cavitation. The former involves generating and applying ultrasound
waves to reactants through a vessel or flask, which are immersed in a
bath of liquid. This type of indirect irradiation is often utilised for
9.4:1
24:1
6:1
8:1
9:1
6:1
9:1

cleaning rather than promoting chemical reaction. For the latter, an


ultrasonic probe, usually known as an ultrasound horn is used for
Methanol
Methanol
Methanol
Methanol
Methanol
Methanol
Methanol
Alcohol

generating the oscillation from an ultrasonic transducer. Fig. 12 shows


a typical setup for ultrasonic chemical reactor using direct cavitation
with an immersed horn.
Loading (wt. %)

Typically, the low and high frequency ranges are defined as


30 mg powder

20–100 kHz and 2–10 MHz, respectively [118]. The high frequency
ultrasound is usually coupled with low power application, useful for
1.16

medical scanning and analytical chemistry. On the contrary, low fre-


1.2
1.0

1.3
1.2
1

quency operation is generally used with higher power output, which is


Catalyst

suitable for ultrasonic cleaning bath and process intensification of


H2SO4
NaOH
KOH

KOH
KOH

KOH
CaO

chemical reactions, especially for increasing the mass transfer across


non-miscible reactants [118,119]. The use of high frequency in ultra-
Flow rate (ml/hr)

sonic intensification for biodiesel production is generally not suitable,


as the energy released from the imploding bubbles are weaker than the
impingement of the reactants [120]. However, recent development of
piezo-electric based ultrasonic reactor has shown progress in over-
175.2

5000
16.1
150

175

coming the problem with high frequency, low power operation [121].

Table 10 summarises the sonication reactor studies that are used for
Internal diameter (mm)

biodiesel transesterification process.


Summary of microchannel reactors for biodiesel transesterification.

Mahamuni and Adewuyi [124] have studied the use of a multi-


frequency ultrasonic reactor, which operates at frequencies of 323 kHz,
581 kHz, 611 kHz, and 1300 kHz, and power input settings of
13–223 W. At the optimum frequency, an increase in input power
0.240a

0.667a
0.600
0.508
0.508
0.900

generally raises the final biodiesel yield, albeit with diminishing im-
0.8

provements. Increased frequency reduces the critical size of the cav-


Rectangular interdigital micromixer

ities, hence also decreases the imploding interval of the cavities [124].
Additionally, the use of higher frequency also promotes smaller cavi-
tation bubbles with increased transport activities. Similarly, as cavita-
tion bubbles reduces in size, the energy released locally during cavities
implosion are further reduced, limiting significant reactions to some
extent. Therefore, there is an optimum point between operating fre-
quency of ultrasound and rate of bubble cavities collapsing, for max-
Mixer type

T-junction

T-junction
T-junction
T-junction
T-junction
T-junction

imum reaction rates. It was also proposed that the 611 kHz and 139 W
Hydraulic diameter.

frequency-power combination form the optimum parameters for max-


imum possible micro-level mixing. As such, increasing operating fre-
quency has positive effects on reaction kinetics as it leads to a decrease
Cottonseed oil
Soybean oil

Soybean oil

in overall volume and implosion time of bubbles.


Feedstock

Pork lard

Palm oil

In a study from Mostafaei et al. [128], probe diameter, ultrasonic


Table 9

WCO

WCO

amplitude, vibration pulse, and irradiation distance were extensively


a

studied. Flow patterns of ultrasound in the reactor can be associated

14
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 7. Micromixer reactor setup with delay loop packed with Dixon ring. Reprinted and adapted with permission from Refs. [105,112]. Copyright (2009) American
Chemical Society.

with probe diameters, where larger probe leads to higher reaction rate. an optimised operating condition of 50 ml/min flow rate, irradiation
Whereas smaller probe can have higher intensity concentrated at the tip distance of 75 mm, probe diameter of 28 mm, vibration pulse at 62%,
area, resulting in an inhomogeneous ultrasonic field. The amplitude and and ultrasonic amplitude of 56%. The biodiesel yield converges to
pulse have similar roles in affecting the energy input to the reaction 91.12% experimentally, which concurs with the model prediction of
medium. An increase in amplitude is followed by increasing ultrasonic 91.6%. Fig. 14 shows the schematic of an ultrasonic-assisted reactor for
power, then reaction enhancement. However, overcommitting with continuous biodiesel production. In the reported experiment, the opti-
high power induces secondary reaction of biodiesel, further cracking it mised flow rate is relatively high when compared with the reactor's
into fractions of shorter-chained organic compound, and oxidising the working volume of 350 ml, hence the accumulation of glycerol phase is
biodiesel into ketones and aldehydes. Fig. 13 summarises the effects of negligible.
increasing frequency and power operation in ultrasonic chemical re- Studies by Aghbashlo et al. developed and evaluated the potential of
actor. using a low power, and high frequency piezo-ultrasonic reactor
Ultrasonic-induced cavitation can be also integrated into a con- [25,121,131]. The novel ultrasonic reactor was able to achieve a
tinuous transesterification system to efficiently produce biodiesel. The 97.12% yield conversion using an operating condition of 6:1 methanol-
implementation of continuously-synthesised biodiesel can achieve to-oil molar ratio, 60 °C reaction temperature, and 10 min ultra-
greater energy efficiency and reduce processing cost for the transes- sonification time, with a specific energy consumption of 378 kJ/kg.
terification process, hence it is more economically feasible for larger Additionally, exergoeconomic and exergoenvironmental analyses were
scale production. Mostafaei et al. [128] utilises waste cooking oil to implemented to optimise the biodiesel production from a thermo-
produce biodiesel continuously with an ultrasonic reactor. The waste dynamic, economic, and environmental perspective. The optimised
cooking oil is first pretreated as the excess FFA and water content in the operating condition yields a cost and environmental impact per unit
oil can reduce the efficacy of transesterification process. Micro-emul- exergy of 2.23 USD/GJ and 1.32 mPts/GJ, respectively.
sion can be achieved by using the ultrasonic to cause micro-streaming,
which is a rapid oscillatory of fluid motion to provide intense micro- 4.3.2. Hydrodynamic
level mixing in the reactants for enhanced biodiesel conversion. Re- According to Bernoulli's effect, a an incompressible fluid undergoes
gression model was used to optimise the continuous reactor, leading to velocity variation in a constriction, the pressure will also change

Fig. 8. Formation of biodiesel in a T-junction mixer of microchannel reactors.

15
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

cavitation reactor has similar underlying principle of operation when


compared with its sonificator reactors counterpart, where both reactor
utilises the formation, growth, and implosion of cavity bubbles to
achieve high intensity of micro-emulsion and high level of micro-
mixing in multiphase reactants [134]. However, hydrodynamic reactors
are often packaged as a static mixing reactor. Mixture of reactants are
allowed to flow through a series of orifices, valves, or throttles in an
open or closed-loop system to achieve uniform mixing of contents
through the manipulation of pressure difference. Table 11 shows the
cavitation reactors that are used for biodiesel production.
Cavitation number (Ca) is a dimensionless number used in hydraulic
devices and flow systems. It characterises the potential of flow cavita-
tion by drawing relationship between the difference of a local absolute
pressure from the vapour pressure and kinetic energy per volume, as
shown in equation below [142].

p − pv
Ca = 1
2
ρv 2 (1a)

where ρ is the density of the fluid, p is the local pressure, pv is the


vaporising pressure, and v is the velocity of the fluid flow.
In a research from Maddikeri et al. [138], biodiesel intensification is
investigated by manipulating cavitation number using inlet pressure
from 1 to 5 bar for the hydrodynamic cavitation reactor. The effect of
inlet pressure on cavitation number using three cavitation devices of
varying geometries is shown in Fig. 15. It is noticeable that the decrease
in cavitation number is associated with the increase in device inlet
Fig. 9. Effect of residence time on pressure for T- and J-junction: (a) 5 s, T-
pressure, which is likely due to an increase in velocity at the throat of
junction; (b) 10 s, T-junction; (c) 20 s, T-junction; (d) 5 s, J-junction; (e) 10 s, J-
junction; (f) 20 s, J-junction (10 mg/g KOH, methanol‐to‐oil molar ratio 6:1,
venture leading to a higher cavitation activity.
and 60 °C). Reprinted from Ref. [108] with permission from Canadian Journal There are considerable differences in the upstream pressure from 1
of Chemical Engineering. Licence Number: 4541360123331. to 2 bar in terms of the changes in cavitation number. However, beyond
3 bar the improvement is not significant. The observation can be jus-
tified by choked cavitation, due to mass transfer resistance being in-
conversely to obey the energy conservation principle of flowing fluids
significant at a higher flow rate, therefore the inherent reaction kinetics
[132]. The Bernoulli's effect allows the fluid in the chemical reactor to
becomes rate limiting [100]. Overall, slit venturi has the lowest cavi-
achieve vapour pressure as it accelerates through these constrictions,
tation numbers as compared to orifice and circular venture, as it has the
creating a local boiling effect which aids in agitating the mixture as a
highest volumetric flow rate resulting in a greater mass transfer across
whole, often known as hydrodynamic cavitation [133]. Hydrodynamic
the constriction. As such, the frequency of cavitation increases. The

Fig. 10. Schematic of microreactor with wire coil of 1 mm pitch length. Reprinted from Ref. [107] with permission from Elsevier. Licence Number: 4541351481631.

16
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 11. Cavitation classifications.

power dissipation from hydrodynamic cavitation can be measured ac- 12:1 alcohol to oil molar ratio, and an inlet pressure of 3 bar.
cording to equation (2) using pressure drop and volumetric flow [138]. Cavitational yield is a generalised parameter to enable comparison
to be made between cavitational reactors [135], and equated by,
PD = ΔP × V0 (2a)
Amount of methyl ester (g )
where PD is the power dissipated into the system in J/s, ΔP is the Cavitational yield =
Energy sup plied (J ) (3a)
pressure drop across the cavitation device in Pa, and V0 is the volu-
metric flow rate through the cavitation device in m3/s. The equation above provides a measure of efficiency based of the
Slit venturi which has the lowest cavitation number is able to pro- energy supplied from the reactor and the amount of biodiesel produced.
duce the highest power, at approximately 1.1 times and 1.5 times This can also be easily extrapolated to other type of biodiesel reactors to
higher than those of circular venturi and orifice, respectively. The re- allow a cross-reactor comparison regardless of reactor category.
sults for biodiesel yield are also dependent on the power input of the Furthermore, Ghayal et al. [137] studied the relationships between
devices, with slit venturi, circular venturi, and orifice having yields of operating pressure of a hydrodynamic cavitation reactor and flow
89%, 82%, and 64%, respectively. The operating condition of the hy- geometries such as hole size, number of holes, and total perimeter of
drodynamic cavitation reactor utilises a catalytic loading of 1 wt %, a holes on a orifice plate. This allows the characterisation of the

Fig. 12. Schematic diagram of ultrasonic system setup used for the preparation of emulsion fuel and cracking of diesel oil. Reprinted from Ref. [117] with permission
from Elsevier. Licence Number: 4541351383318.

17
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

cavitation activities in biodiesel transesterification using a hydro-

2006

2007

2009
2010
2010

2013

2015

2016

2017
2017
Year
dynamic reactor. The results show similar trends to that of Maddikeri
et al., where the improvements in cavitation and rate of reaction were
[122]

[123]

[124]
[125]
[126]

[127]

[128]

[129]

[130]
[121]
Ref.

observed when upstream pressure increases. However, increasing


pumping liquid flow rate had diminishing returns in terms of im-
Power density

274.809 kJ/L
provement. This work extensively correlates the cavitation effects to

124.3 kJ/L

378 kJ/kg
556 kJ/L

3
flow geometries of orifice plates, such as the ratio of total perimeter of

2 W/cm
900 J/g

holes to total flow area on plate, α, the ratio of hole diameter on orifice
– to pipe diameter, β, and the ratio of total flow area on orifice plate to


the cross-sectional flow area of pipe, β0. Higher biodiesel yield is
Power (W)

achieved with increasing values of α, β, β0 for orifice plates, which are

102.8
2000
150

600

139
100

200
physically represented by greater number of holes and smaller hole size

80

40
31
of the orifice. These flow geometry properties contribute to more cavity
generating spots, allowing greater shear layer area to exist, hence mass
Frequency

transfer is further enhanced through better emulsification of multiphase


(kHz)

1700
reactants.
19.7

611
45

20
20

25

24

20

20 Chuah et al. [139] utilised an optimised orifice plate to demonstrate


large scale biodiesel transesterification via a hydrodynamic cavitation
(wt. %)

86.61

97.12
system with a 50 L capacity. The reactor is assisted by a double dia-
b
Yield

96.0

99.0
95.2
99.0

91.6

96.1

99.0
100

phragm pump in their pilot biodiesel plant as shown in Fig. 16. Waste
cooking oil is used as the feedstock of choice because of its low cost and
Reaction time

abundance in availability. The optimised operating parameters of the


biodiesel plant was able to produce biodiesel of more than 96.5% in
(min)

240

180

15 min of reaction time. Thus, the biodiesel produced complies with the
15

20

30
60
50

90
10

international biodiesel standards of American standards, ASTM D 6751


Temperature (°C)

and British standards, EN 14214. Effects of oil to methanol molar ratio


were investigated from 1:4 to 1:7, where only 1:6, and 1:7 resulted in
achieving a high biodiesel yield. Molar ratio with lower level of me-
thanol is insufficient when diffusion of methanol into glycerol layer
45

48

45
65
35

60

45

40

60

takes place. In contrast, high methanol to oil molar ratio level also

decreases conversion slightly, as the dilution of oil with methanol oc-


Alcohol to oil

curs. The case with catalyst loading of 1.0 wt % performs better as


molar ratio

compared to 1.25 wt %, where the latter was observed to promote sa-


ponification reaction to form soap, contributing to biodiesel conversion
6:1

6:1

6:1
9:1
5:1

6:1

6:1

9:1

5:1
6:1

inefficiency. Besides, operating temperatures of 50 °C and 65 °C are


optimal for biodiesel transesterification in a hydrodynamic cavitation
reactor, since higher temperature promotes solubility between oil and
carbonate
Methanol

Methanol

Methanol
Methanol
Methanol

Methanol

Methanol
Methanol
Dimethyl
Alcohol

Acetate
Methyl

methanol, increasing the interfacial contact surface of two reactants, for


better mass transfer [143].
However, there are limited amount of studies in continuous bio-
diesel production using a hydrodynamic cavitation reactor, with an
Loading
(wt. %)

open-loop flow system. Most of the setup is based on a closed-loop


0.23
1.0

0.5
3.0
0.7

1.0

3.0

1.0

recirculation of reactants via a pump, through constriction devices to


10

induce cavitation over certain amount of time. The biggest issue of


Lipase enzyme

Lipase enzyme
Novazym 435

hydrodynamic cavitation is the static mixing of reactants, which re-


quires a significant amount of pressure differences, usually above 1 bar
Catalyst

upstream pressure for the inlet cavitation device [137,138,140]. This


KOH

KOH

KOH

KOH

KOH

KOH
SrO

allows the fluid to reach a desired velocity change, hence achieving


Summary of sonification reactor for biodiesel production.
Oil quantity

vaporising pressure for the cavitation process to take place. Therefore,


as the volumetric flow increases, closed-loop systems are favourable as
450 mla
250 mla
1000 g

they take up less space. Continuous hydrodynamic reactor for biodiesel



transesterification is viable by implementing a series of constriction


Continuous flow reactor (constant

devices to increase the cavitation zone. Subsequently, the down-scaling


Ultrasonic water bath (50% duty

Continuous flow reactor (50 ml/


Continuous flow reactor at 8 L/

of reactor's total capacity is recommended to compensate for the long


Submerged ultrasonic horn at

Piezoelectric-based ultrasonic
Ultrasonic horn submerged

flow pathway of an open-loop system.


cycle) with static mixer at

Probe diameter at 3.5 cm


residence time) at 6 vol

(2 cm), 60% duty cycle


Reacting parameters

Conversion of feedstock oil.

4.4. Microwave reactor


Volume of total reactants.
replacement

Microwaves are in fact not micro in size, instead, they have wave-
100 rpm

lengths ranging from millimetre to meter scale. The common fre-


reactor
min)
2 cm

quencies of a microwave are between 300 MHz (100 cm) and 300 GHz
min

(0.1 cm) [144]. In the electromagnetic spectrum, microwave is situated


Soybean oil

Soybean oil

between infrared and radio wavelength, hence it is a considerably large


Canola oil

Canola oil
Feedstock
Table 10

Palm oil

Palm oil

wave as compared to other type of electromagnetic radiation.


WCO

WCO

WCO

WCO

The main application of microwaves is the generation of heat, often


b
a

in household microwave oven operating at 2.45 GHz specifically to

18
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 13. Effects of increasing frequency and power operation in ultrasonic cavitation reactor.

avoid interference with telecommunications. Fundamentally, the mi- have a penetration depth of 0.76 cm in methanol at 20 °C, and 1.4 cm at
crowave produced by a magnetron, is carrying both electric and elec- 60 °C. Therefore, external agitation or stirring is still required for uni-
tromagnetic fields that are directly perpendicular, and travelling at the form heating of the bulk fluid. Thus, the scaling ability of microwave-
speed of light. The fields interact with dielectrics such as polar mole- assisted transesterification is fairly poor as compared to other proces-
cules and ions, to continuously realign them along the field's direction sing techniques.
of propagation. Heat energy is then generated from the rapid oscillation Nevertheless, many research have claimed that high biodiesel yields
of rotating polar molecules with its surrounding molecule via frictional can be obtained in a short amount of time, through the use of micro-
forces. Large ions can also contribute to the heating of liquid media wave irradiation. Lin et al. [154] reported that microwave heating
when the electric field changes in direction, as a result of kinetic energy system was used to improve the palm biodiesel's yield. The microwave
dissipation in the form of heat [145]. Microwave irradiation heating is reactor can simultaneously reduce both reaction time and energy con-
more efficient as compared to conventional conduction and convection sumption of biodiesel production, as compared to conventional heating
within material as it does not depend on the material properties, such as system. When the process is optimised, a 99.5% biodiesel yield was
thermal conductivity, specific heat capacity, and density of material achievable in 3 min through microwave irradiation, combined with
[146]. mechanical stirring from a magnetic stirrer. This represents up to a total
Biodiesel transesterification reaction involves reactions such as li- of 30 times reduction in reaction time and 22 times reduction in total
pids and alcohol which are good microwave absorbers as they contain energy consumption, when compared with conventional heating. Be-
high level of dielectrics [145]. Thus, microwave-assisted transester- sides, using a higher methanol to oil molar ratio decreases the biodiesel
ification allows reaction intensification using alternative heating yield, which is due to the dilution of heat dissipation from the micro-
method, rather than the less efficient conduction from water bath or waves, since methanol is a better microwave absorber as compared to
heating element. Table 12 tabulates the microwave reactors that are oil. The optimised operating parameter runs on 750 W microwave
used for biodiesel production from different studies. power output, with 6:1 methanol to oil molar ratio, using a Ch3ONa
Although microwave can heat up reactant rapidly, it is only re- catalyst loading of 0.75 wt%.
stricted to a small volume of fluid due to the limited penetrating depth The dielectric constant, Ɛ′ characterises how well a dielectric com-
of microwave. According to Grant and Halstead [160], penetrating pound absorbs microwaves and stores energy, while dielectric loss Ɛ’’,
depth is a function of reactant's temperature. For example, microwaves represents the ability of compound to convert absorbed energy into heat

Fig. 14. Schematic diagram of experimental setup for ultrasonic assisted continuous biodiesel production. Reprinted from Ref. [128] with permission from Elsevier.
Licence Number: 4541351298136.

19
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

2006

2008

2010
2013
2014

2015

2016

2019
Year

[122]

[135]

[136]
[137]
[138]

[139]

[140]

[141]
Ref.
Cavitational yield
(mg/J)

2.58a

0.17

1.28
1.22

1.25

0.91


Yield (wt.

b
97.0

92.0

80.0
95.0
89.2

98.1

97.0

99.0
%)
Reaction time

Fig. 15. Cavitation number produced by different inlet pressure of different


(min)

cavitation devices. Reprinted from Ref. [138] with permission from Elsevier.
60

90

30
60
60

15

18

Licence Number: 4540140273191.


Temperature (°C)

[155]. When microwave heating system is used in biodiesel transes-


terification, the selection of alcohol is important, as it holds a sig-
< 57.1c
45–55

nificant volume proportion of the mixture. However, the selection of oil


45

28

60

60

55

35

is less significant since the dielectric constant is between the value of 3


and 5 [161]. Alcohol with high dielectric constant and dielectric loss is
favourable since it is a better microwave absorber. In theory, methanol
Alcohol to oil
molar ratio

will outperform the other primary alcohols as a solvent for microwave


irradiation as the dielectric constants for methanol, ethanol, and pro-
4.5:1

6.8:1
10:1

12:1

panol are, 32.7, 24.5, and 17.9, respectively [161]. Yuan et al. [150]
6:1

6:1

6:1

6:1

reported that the effect of methanol is greater as compared to ethanol


for the transesterification process, which suggest that more energy is
Methanol

Methanol

Methanol
Methanol

Methanol

Methanol

Methanol

transferred into the oil-methanol system. Other researcher such as


Alcohol

acetate
Methyl

Barnard et al. [147], Lertsathapornsuk et al. [148], and Leadbeater


et al. [149], have used methanol, ethanol, and butanol, to obtain bio-
diesel yield of 99.0%, 97.0%, and 93.0%, respectively.
Loading (wt.

Calculated values, where soybean methyl ester molecular weight is assumed as 292.2 g/mol.

Microwave heating has good synergy with heterogeneous catalyst,


especially those with high microwave absorption characteristic. Since
catalytic effect takes place on the active site of solid catalyst, reactants
1.0

1.0

1.0
1.0
1.0

1.0

1.0

1.0
%)

adsorb on the external and internal surface which will be acting as


microwave hot spots, promoting rate of reaction for biodiesel transes-
terification [150]. Yuan et al. [150] used activated carbon as hetero-
methoxide
Potassium

geneous catalyst carrier, as it has a high volume to surface area and


Catalyst

H2SO4

NaOH

NaOH

strong microwave radiation absorbing characteristics. The highly polar


KOH

KOH

KOH

KOH
Summary of hydrodynamic cavitation reactor for biodiesel production.

sulfuric acid (H2SO4) is also used as a loading catalyst combined with


activated carbon (H2SO4/C). This combination further enhances the
Operating pressure

microwave absorption properties. The experiment compares the effect


between homogeneous (H2SO4) and heterogeneous (H2SO4/C) type of
catalyst, under the heating influence of microwave irradiation. Sulfuric
acid concentration is controlled to be identical for both cases, with the
(MPa)

0.7

0.3
0.3

0.2

0.3

0.7

H2SO4/C reaching reaction equilibrium faster, hence a biodiesel yield


of more than 90% is achieved as compared to 40% when H2SO4 is used


Plate with 21 hole at 1 mm

Plate with 21 hole at 1 mm


Single orifice with 1000 g

as catalyst for the same reaction time.


100 hole at 0.3 mm each

Below boiling point of methyl acetate.

In the recent years, microwave heating has capture more attention


Single 10 mm orifice
Reacting parameters

Single orifice plate

in 3rd generation biodiesel feedstock processing. The high cost of


Single slit venturi

diameter each

diameter each

downstream treatments which involves harvesting, drying, extraction,


Orifice plate

Yield in percentage of mol.

separation, and transesterification of 3rd generation feedstocks such as


microalgae is not economically favourable for upscale production
[162]. Thus, microwave technologies are used in attempts to (i) provide
oil

efficient heating, (ii) reduce number of intermediate process by com-


Sunflower and palm

bining stages, (iii) provide better extraction of lipids from feedstocks.


Rubber seed oil

Loong et al. [162] developed a one-step transesterification using si-


Soybean oil

Thumba oil

multaneous cooling and microwave heating (SCMH), where cell dis-


Feedstock
Table 11

ruption and transesterification of wet microalgae (Nannochloropsis sp.)


oil

WCO
WCO

WCO

WCO

was performed simultaneously in a single step. The one-step SCMH was


b
a

20
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 16. Schematic of a 50 L capacity hydrodynamic cavitation reactor for biodiesel production, and hydrodynamic cavitation devices (a) slit venturi, (b) circular
venturi, and (c) orifice plate. Reprinted and adapted from Refs. [138,139] with permission from Elsevier. Licence Number: 4540140273191, 4540140126930.

able to achieve a 75.0% biodiesel production level, as compared to 4.5. Cosolvent-assisted transesterification
water bath heating of 13.46%, and one-step microwave heating of
15.39%. The cooling stage promotes the microwave absorption prop- Cosolvent is usually a secondary solvent used in a reaction that does
erties of molecules, hence improving the penetration of microwave ir- not react with the primary solvent and be involved in the primary re-
radiation, subsequently promoting the microalgae lipid extraction and action pathway. Instead, the use of a cosolvent is to amplify the mag-
transesterification yield. nitude of effectiveness in trace quantities, prior to the primary solvent.
Continuous flow microwave reactor has also been explored for In the context of biodiesel transesterification, the cosolvent used for the
biodiesel production, as reported by Choedkiatsakul et al. [156]. process should be inert towards the chemistry of transesterification, and
Commercially available microwave system can help to streamline the does not deactivates the catalyst [9].
reactor setup and design. However, the size of these systems still re- Biodiesel transesterification uses oil, which mainly consists of tri-
stricts the biodiesel production to be performed within laboratory scale. glycerides that are nonpolar and hydrophobic. The oil is used to react
Fig. 17 shows the configuration of the continuous biodiesel production with alcohol which is polar and hydrophilic, thus, using the analogy of
microwave system. The use of thin coiled tubing in the microwave ‘like dissolve like’, oil and alcohol is immiscible in nature. Therefore,
system is advantageous to efficiently transfer energy. Since there is no when oil and alcohol are mixed together, the transesterification process
mechanical stirrer in a continuous microwave system, homogeneous only occurs on the contact surfaces of oil and alcohol, resulting in the
heating can only be done via full penetration of microwave to reactants. formation of a two-phase system.
The microwave system is able to achieve a biodiesel yield of 99.4%, in The cosolvent used for biodiesel production should be soluble in
105 s of residence time. The operating conditions entail a methanol to both fluids, acting as an interface solvent between oil and alcohol. This
oil ratio of 12:1, a NaOH catalyst loading of 1.0 wt %, under a micro- encourages the formation of a single-phase system. Biodiesel transes-
wave power output of 400 W at 70 °C reacting temperature. The mea- terification has been proven to achieve higher yield by using cosolvent
sured energy consumption using the system is 0.1167 kWh/L, which is such as Acetone (ACT), Tetrahydrofuran (THF), diethyl ether (DEE),
half of the requirement for conventional heating process. and FAME [163–165]. The solubility of some solvents and organic
solvents are compared with transesterification compound, shown in
Table 13, while biodiesel transesterification yield through different

21
K.Y. Wong, et al.

Table 12
Summary of microwave reactor for biodiesel production and their respective operating conditions.
Feedstock Reacting parameters Catalyst Loading Alcohol Alcohol to oil Temperature (°C) Reaction Power Yield Yield Efficiency Ref. Year
(wt. %) molar ratio time (min) Output (W) (wt. %)

WCO Continuous microwave reactor at 7.2 L/ KOH 1.0 Methanol 6:1 50 4.5 1600 94.6 26 kJ/L [147] 2007
min
WCO Continuous microwave reactor mixed via NaOH 3.0 Methanol 12:1 78 0.5 800 97.0 74.8 Wh/L [148] 2008
T-junction 0.222 kWh/L
Soybean oil Batch microwave reactor with magnetic KOH 1.0 1-Butanol 6:1 120 1.0 1600 93.0 – [149] 2008
stirrer using 1L of oil
Castor oil Microwave reactor with static mixing at H2SO4/C 5.0 Methanol 12:1 65 60 200 94.0 – [150] 2009
600 rpm using 20 g of oil
Yellow horn seed oil Microwave batch reactor with 10 g of oil H2SO4 1.0 Methanol 20:1 90 30 500 95.2 – [151] 2010
Yellow horn seed oil Microwave batch reactor with 10 g of oil Heteropolyacid 1.0 Methanol 12:1 60 10 500 96.2 – [151] 2010

22
(Cs2·5H0·5PW12O40)
Dry algae oil Dry algae biomass of 2 g, simultaneous KOH 2.0 Methanol 12:1 60–64 4–5 400 80.13 – [152] 2011
(Nannochloropsis) lipid extraction and transesterification via
microwave irradiation
Jatropha curcas oil Continuous recycling flow in microwave CaO 3.17 Methanol 8.42:1 60 67.9 800 97.1 – [153] 2013
reactor at 8 ml/min
Palm oil – CH3ONa 0.75 Methanol 6:1 – 3 750 99.5 – [154] 2014
WCO Microwave and static mixer at 300 rpm KOH 1.0 Methanol 6:1 60 8 300 91.3 – [155] 2014
with total reactant volume of 80 mL
Palm oil Reactant premixed by static stirrer, then NaOH 1.0 Methanol 12:1 70 1.75 400 99.4 0.1167 kWh/L [156] 2015
feed into microwave reactor continuously
Kapok seed oil Microwave and static stirrer with 50 g of KOH 2.15 Methanol 9.85:1 57.09 3.29 3000 98.9 – [157] 2015
oil
Palm oil Pressurised microwave reactor at 7 bar NaOH 1.0 Methanol 6:1 – 2.3 780 97.82 – [158] 2018
Ceiba pentandra oil Microwave and static stirrer at 800 rpm KOH 0.84 Methanol 6:1 100 6.46 850 96.19 – [159] 2019
Renewable and Sustainable Energy Reviews 116 (2019) 109399
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 17. Schematic of continuous flow microwave reactor for biodiesel production. Adapted from Ref. [156] with permission from Elsevier. Licence Number:
4540140009135.

Table 13 alcohol and base type catalyst loading are shown in Table 14.
Dissolution of reactants and products in organic solvents at 20 °C. Reproduced Table 15 summarises the cosolvent-assisted biodiesel transester-
from Ref. [166] with permission from The Royal Society of Chemistry. Licence ification studies. Alhassan et al. [165] demonstrated the production of
Number: 4540131081234. biodiesel from cotton seed oil by using commercially available cosol-
Solvents Organic Solvents vent such as, DEE, Dichlorobenzene (CBN), and ACT, to produce a co-
solvent system with mixtures of methanol at different levels of con-
Compounds H2O CH3OH ACT THF IPA AN DEE FAME centration. Under the reaction temperature of 60 °C, methanol to oil
Triglycerides X X ⌾ ⌾ ⌾ X ⌾ ⌾
FAME X X ⌾ ⌾ ⌾ ⌾ ⌾ ⌾
molar ratio of 6:1, and catalyst concentration of 0.75%, the biodiesel
Glycerol ⌾ ⌾ X X ⌾ X X X yield achieved exceeded 90% conversion in the first 10 min with 10 vol
CH3OH ⌾ ⌾ ⌾ ⌾ ⌾ ⌾ ⌾ X % cosolvent, as compared to non-cosolvent transesterification observed
Oleic acida X ⌾ ⌾ ⌾ ⌾ X ⌾ ⌾ at below 70% yield. The experimental results suggest that the reaction
Stearic acidb X X X ⌾ X X X X
proceeds as soon as phase equilibrium between solvent and oil estab-
Soap ⌾ ⌾ ⌾ ⌾ ⌾ Δ Δ Δ
lished in the first few minutes. All of the different cosolvent systems
⌾: Miscible. enable at least 90% biodiesel yield performance in the first 10 min
Δ: Partially miscible. except for DEE which resulted lower yield. The findings also agree to
X: Immiscible. the results reported by Luu et al. [163], where non-solvent system has a
Organic solvent used: Methanol (CH3OH), Acetone (ACT), Tetrahydrofuran slower reaction rate. This is reflected by a maximum difference of 15%
(THF), Isopropyl alcohol (IPA), Acetonitrile (AN), Diethyl ether (DEE), Fatty biodiesel yield at the 100th minute point of the transesterification
acid methyl ester (FAME). process (< 100 min), as compared to cosolvent used such as Diethyl
a
Melting points of oleic acid is 16.3 °C. Ketone, Ethyl Methyl Ketone, AN, and Ethyl Acetate.
b
Melting points of stearic acid is 69.9 °C.
Furthermore, the separation time of glycerol from biodiesel

Table 14
Biodiesel transesterification yield through different alcohol and base type catalyst loading. Reprinted and adapted from Ref. [167] with permission from Elsevier.
Licence Number: 4540100574669.
Catalyst Catalyst Loading (%) Biodiesel Yield (%)

Methanol Ethanol Isopropanol Butanol

Time (min) Time (min) Time (min) Time (min)

10 20 40 60 10 20 60 10 20 60 20 40 > 60
NaOH 0.5 – – – 80 – 79 – 78 – – – – 83
1.0 91 – – – – 40 – 47 – – – 90 –
1.5 35 – – – 47 – – 79 – – 89 – –
KOH 0.5 – – – 86 – 55 – 76 – – – – 91
1.0 – – 85 – – 57 – 60 – – – 91 –
1.5 – 83 – – 44 – – 29 – – 97 – –

23
K.Y. Wong, et al.

Table 15
Summary of cosolvent-assisted biodiesel production process.
Feedstock Reacting parameters Cosolvent Cosolvent Catalyst Catalyst Alcohol Alcohol to Temperature (°C) Reaction Yield (wt. Ref. Year
loading (wt. %) loading (wt. oil molar time (min) %)
%) ratio

Soybean oil Tetrahydrofuran to methanol volume Tetrahydrofuran 38.4 NaOH/ 1.0 Methanol 6:1 20 10 87b [168] 1996
ratio of 1.6
a
Corn oil Cosolvent pressurised at 500 kPa, Dimethyl ether 0.57 p-toluenesulfonic 4.0 Methanol 6:1 80 120 97.1 [169] 2009
reactor in oil bath and shaker with acid
2.6 Hz shaking speed.
WCO, Canola, Acetone premixed with 40 g of oil by Acetone 25 KOH 1.0 Methanol 4.5:1 25 30 98.1c, [164] 2013
Catfish, Jatropha magnetic stirrer until homogeneous 98.3c,
curcas oil before transesterification 97.0c,
97.9c
Jatropha curcas oil Acetone premixed with 30 g of oil Acetone 20 KOH 1.0 Methanol 6:1 40 30 99.0 [163] 2014

24
Cotton seed oil Magnetic stirring at 350 rpm Dichlorobenzene and 10d KOH 0.75 Methanol 6:1 55 10 96.85 [165] 2014
Acetone
Microalgae oil Wet algae of 0.75 g reacted in sealed Ethyl acetate 6.67e H2SO4 4.06 M Ethyl acetate 6.67f 114 120 97.8 [170] 2017
tube with no agitation.
Linseed oil Linseed in-situ transesterification at Tetrahydrofuran 30 KOH 6.8 Methanol 10:1g 40 90 93.15 [171] 2018
700 rpm agitation rate
Sunflower oil Magnetic stirrer at 900 rpm Crude biodiesel 10 CaO 0.602 mol Methanol 6:1 50 300 99.9 [172] 2019

a
DME to methanol ratio.
b
Estimation from results.
c
Yield are according in the order of feedstocks; Catfish oil's reaction temperature at 35 °C.
d
Cosolvent loading in percentage of volume (vol. %).
e
Cosolvent loading in ml/g.
f
6.67 ml Ethyl acetate per gram of dried algae.
g
Alcohol to oil molar ratio in percentage of weight (wt. %).
Renewable and Sustainable Energy Reviews 116 (2019) 109399
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Table 16 production is also extensively studied by Thanh et al. [164]. It was


Specific density, relative viscosity, and yield of FAME with cosolvent via mixing observed that after a 60-min transesterification, the glycerol separation
ratio of 2:1 ratio wt. % at 15 °C. Reproduced from Ref. [166] with permission times require 37, 30, and 50 min using acetone at 20, 25, 30 wt % of oil,
from The Royal Society of Chemistry. Licence Number: 4540131081234. respectively. However, without cosolvent the phase separation requires
Mixture Density Relative Biodiesel yieldb a significant longer time of 4 h. The addition of cosolvent allows to
(g.cm-3) viscositya (%) increase (i) the difference in specific gravity and (ii) difference in re-
lative viscosity between the two phases [166]. As a result, glycerol can
Water 1.0000 1.000 –
FAME + isopropanol 0.8456 4.130 99
be separated from the reacted system and precipitated within minute
FAME + tetrahydrofuran 0.8797 2.630 93 scale, because of the large differences in specific gravity and similar
FAME + acetonitrile 0.8311 2.020 90 viscosities between the two components. For the case of using acetone
FAME + acetone 0.8395 1.830 99 concentration at 30 wt %, the increase in phase separation time can be
FAME + dimethyl ether 0.8202 1.690 93
pointed towards the relationship between the relative concentrations of
FAME 0.8797 0.885 88
glycerol and methanol to the overall mixture. When the acetone con-
a
Relative viscosity to water. centration increases, the relative proportion of excess methanol-gly-
b
Based on reaction condition of: molar ratio methanol to canola oil, 4.5:1; cerol decreases in the overall system, hence they will be less likely to
cosolvent to oil, 25 wt%.; KOH catalyst loading, 0.5 wt%; Temperature, 20 °C; merge and form a single phase that separates from the biodiesel layer.
Reaction time, 60 min. Yield is an estimation from data points extracted from The specific gravities and viscosities of FAME and potential cosol-
graph [166]. vent for transesterification are tabulated in Table 16, with the FAME-
isopropanol system separating the fastest and dimethyl ether the
slowest for glycerol.

5. Trends and perspectives

5.1. Techno-economic analysis

The current biodiesel production volume as shown in Fig. 18 illus-


trates that there are only a few countries that contribute more than a
billion litres of biodiesel volume production annually, such as United
States of America (USA), Brazil, Argentina, Germany, France, Spain,
Poland, Indonesia, Thailand and China. The cumulative production of
the aforementioned countries are 25.6 billion litres, equivalent to
74.8% of the total world production.
The versatility of biodiesel feedstock allows many other countries
across the continent to yield great potential for larger volume of bio-
diesel production. Thus, a data-driven predictive model was used to
estimate the maximum biodiesel production volume of 155 countries as
shown in Fig. 19 [173]. The model is supported with primary sources of
data obtained from the Food and Agriculture Organization of the United
Nations Statistics Division [174], Index Mundi [175] and US Energy
Information Administration [176]. In addition to the previously men-
Fig. 18. Worldwide leading biodiesel production in 2017, by continent (in tioned country, eight other countries such as, Malaysia, Ukraine, Ca-
billion litres). nada, Netherlands, Russia, Belgium, Philippines, and Colombia have
immense potential for biodiesel production volume exceeding a billion
litres annually. The Southeast Asian nations such as Malaysia and In-
donesia have the capacity to produce up to 16.3 and 30.3 billion litres

Fig. 19. Maximum biodiesel production volume (in billion litres).

25
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Fig. 20. Maximum biodiesel blend levels achievable for each country.

of biodiesel per year, due to the excess availability of palm feedstock. predominant feedstock used in the production, storage, and usage of the
Fig. 20 shows the maximum biodiesel blend levels that is achievable biofuel for the countries. For example, the nature of Malaysia's tropical
for the 155 countries, with respect to their level of diesel consumption. climate requires a higher average storage temperature, hence a more
It is indicated that biodiesel can potentially replace all the diesel fuel stringent requirement for fuel flash point, as shown in Table 17. This
consumption (B100) for only one country, namely, Malaysia, provided helped in allaying fears in Malaysia when the country implemented a
that the economic climate is viable for biodiesel production. There are mandatory blending of 10% biodiesel for fossil diesel in December 2018
other countries situated between the red and black lines in Fig. 20 that [177], to exploit the potential that the country has to reduce un-
have the potential to implement B10 to B100, while countries suitable exported crude palm oil and also improve energy security [177].
for B1 to B10 blends are in between the black and blue lines, and In recent years, stringent regulations on emissions have been an
countries not having potential above B1 are below the blue line. important driver to encourage environmentally friendly fuel in internal
There are 25, 41, and 23 countries that are able to volumetrically combustion engines. The emissions from diesel engines such as CO,
replace 10–100%, 1–10% and below 1% biodiesel blend in diesel, re- NOx, UHC and soot are often correlated with health hazards [4,5].
spectively. The large volume of diesel consumption in countries such as Nevertheless, biodiesel are often labelled as fuel with lower emissions,
China, Indonesia, and Japan have made it less feasible for their hence the emissions of 14 biodiesel feedstocks were studied and com-
economies to implement a higher blending ratio of biodiesel solely with pared against diesel emissions in Table 18.
the locally available feedstock. Clearly, biodiesel fuel has potential in reducing the dominant
emissions from combustion engines when compared with traditional
5.2. Biodiesel policy, standards and emissions analysis diesel fuel. From Table 18, coconut biodiesel has the most significant
reduction in emissions as it is the only feedstock that achieved all four
Biodiesel made from vegetable oil or animal fat has long-chain alkyl reduction in emissions, and highest reduction in smoke opacity, which
from either methyl, ethyl, or propyl esters. Often, waste cooking oil is is often associated with the particulate matter concentration. Although
post-processed to be used as a feedstock for biodiesel production which most of the biodiesel from different feedstock has a reduction in
are derived from both vegetable and animal fat, hence the variance in emissions level, feedstocks such as olives and rice bran might not be
the feedstock quality can be quite significant. As such, the quality of good candidates for biodiesel production due to its poorer performance
biodiesel has to be controlled during production to ensure it conforms in emissions.
to the standards specified by international bodies.
The standards are designed to provide specifications for the physico- 5.3. Practical implications of this study
chemical characteristics of the biodiesel. The two major international
biodiesel standards are the American Standards for Testing Materials The prevailing batch-type reactor is used as a benchmark against
(ASTM D6751) and the European Standards (EN14214). Some biodiesel other types of biodiesel processing reactors. However, the comparison
producing countries adopt the specification as guidelines from ASTM made is often unfair as the biodiesel yield and reaction time are the only
and EN to define their own biodiesel standards. The biodiesel standards factors to be compared. As such, other more important factors such as
mandated in the United States (ASTM), the Europe Union (EN), the processing capacity, energy efficiency, space requirement, en-
Malaysia, Australia, India (IS), Japan (JIS), Brazil (ANP), and South vironmental impact, and scalability of technology have to be taken into
Africa (SANS) are tabulated in Table 17. consideration for a holistic comparison [186]. For microreactor, the
Typically, biodiesel standards have localised context to favour the conventional flow production is inefficient due to the nature of the two

26
K.Y. Wong, et al.

Table 17
Biodiesel standards for various countries [178–185].
Property Units EN 14214:2008 ASTM D6751-12 Malaysia PME Australia Biodiesel India IS 15607:2005 Japan JIS Brazil ANP 42 South Africa SANS
Standards 2014 Standards 2003 K2390:2016 1935

Ester content %(m/m) > 96.5 – > 96.5 > 96.5 > 96.5 > 96.5 – > 96.5
Density at 15°C kg/m³ 860–900 – 860–900 860–890 860–900 860–900 – 860–900
Viscosity at 40°C mm2/s 3.5–5.0 1.9–6.0 3.5–5.0 3.5–5.0 2.5–6.0 2.0–5.0 Reporta 3.5–5.0
Flash point °C > 101 > 93 > 120 > 120 > 120 > 100 > 100 > 120
Cloud point °C – Reporta – – – – – –
Pour point °C – – – – – – – –
Cold filter plugging point °C – – < 15 – – Reporta – (−)4/(+)3b
Sulfur content mg/kg < 10 < 15 ppmc < 10 < 10 < 50 < 10 < 10 < 10
< 0.05%d
Carbon residue (on 10% distillation %(m/m) < 0.3 < 0.050 (100% – < 0.3 < 0.05(100% < 0.3 < 0.05(100% < 0.3
residue) sample) sample) sample)
Acid value mg KOH/g < 0.5 < 0.5 < 0.8 < 0.5 < 0.5 < 0.8 < 0.5
Sulfated ash content %(m/m) < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02
Water content mg/kg < 500 < 500 < 500 < 500 < 500 < 500 < 500 < 500
Total contamination mg/kg < 24 – < 24 < 24 < 24 < 24 Reporta < 24
Cetane number – > 51 > 47 > 51 > 51 > 51 > 51 > 45 > 51
Copper strip corrosion (3 h at 50°C) rating Class 1 < No. 3 Class 1 No.3 or Class 1e No. 1 No. 1 No. 1 No. 1
Oxidation stability, 110°C hr >8 >3 > 10 >6 >6 > 10 >6 >6

27
Iodine value g iodine/ < 120 – < 110 – Reporta Reporta Reporta < 140
100 g
Linolenic acid methyl ester %(m/m) < 12 – < 12 – – < 12 – < 12
Polyunsaturated (≥ 4 double bonds) %(m/m) <1 – <1 – – – – <1
methyl esters
Methanol content %(m/m) < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.5 < 0.2
Monoglycerides content %(m/m) < 0.7 < 0.4 < 0.7 – – < 0.7 < 1.0 < 0.8
Diglycerides content %(m/m) < 0.2 – < 0.2 – – < 0.2 < 0.25 < 0.2
Triglycerides content %(m/m) < 0.2 – < 0.2 – – < 0.2 < 0.25 < 0.2
Free glycerol %(m/m) < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02
Total glycerol %(m/m) < 0.25 < 0.24 < 0.25 < 0.25 < 0.25 < 0.25 < 0.38 < 0.25
Phosphorus content mg/kg <4 < 0.001% wt. <4 < 10 < 10 <4 < 10 < 10
Group I metals (Na + K) mg/kg <5 <5 <5 <5 Reporta <5 < 10 <5
Group II metals (Ca + Mg) mg/kg <5 <5 <5 <5 Reporta <5 Reporta <5
Distillation temperature, 90% °C – < 360 – < 360 – – < 360 –
recovered (T90)

a
Report as measured.
b
For winter/summer.
c
For S15.
d
For S500.
e
For < 10 mg/kg sulfur.
Renewable and Sustainable Energy Reviews 116 (2019) 109399
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Table 18 extensive research work, for scaling-up into commercial applications.


Percentage change in emissions of NO, CO, UHC, and smoke opacity of different In-depth understanding and analysis between batch-type and con-
biodiesel as compared to fossil diesel. tinuous-type transesterification are vital for designing hybrid reactors
Feedstock CO (%) NO (%) UHC (%) Smoke opacity (%) that integrates more than one type of technologies, such as microwave
and mechanical stirring, which can lead to more sustainable biodiesel
Coconuts −0.90 −0.45 −22.77 −51.98 production processes. Additionally, environmental sustainability tools
Cottonseed 9.50 7.30 5.71 −37.98
should be considered in the biofuel industry moving forward, such as
Groundnut 8.92 6.36 8.61 −21.16
Kapok 7.38 4.91 3.38 −34.82 life cycle assessment (LCA), energy and exergy analyses of production
Maize 9.07 7.81 5.42 −23.55 methodologies to evaluate their economic values and environmental
Olives 7.31 3.92 15.02 11.98 impact.
Palm 2.00 −1.39 −4.60 −35.91
Palm kernel 3.52 5.24 −12.49 −31.44
Rapeseed 9.12 6.84 4.35 −39.85
Acknowledgement
Rice bran 7.77 5.49 13.33 12.22
Safflower 9.00 8.25 3.69 −24.37 The authors would like to acknowledge the support provided by the
Sesame 9.30 8.92 3.30 −35.70 Malaysian Ministry of Education (MOE) under the Fundamental
Soybean 9.36 8.74 −0.76 −45.14
Research Grant Scheme (FRGS) through project FRGS/1/2016/TK07/
Sunflower 9.89 8.18 6.17 −31.60
USMC/02/2, and also thank Universiti Malaysia Terengganu under
Golden Goose Research Grant Scheme (GGRG)(Vot 55191) for sup-
phase behaviour between oil and alcohol, which reduces their mis- porting Dr Lam to perform this joint initiative with University of
cibility during the initialisation of transesterification reaction. Both the Southampton Malaysia.
droplet and pulsating approaches in microreactor have proven to in-
crease yield due to the enhanced surface to volume ratio. Currently, Appendix A. Supplementary data
homogeneous catalyst are favoured in hydrodynamic reactor due to the
ease of implementation. The development of heterogeneous-type cata- Supplementary data to this article can be found online at https://
lyst and cosolvent-assisted transesterification can address the issue of doi.org/10.1016/j.rser.2019.109399.
biodiesel feedstocks with high FFA, while simultaneously reducing the
amount of alcohol and catalyst requirements. Besides, the introduction References
of cosolvent can improve the miscibility between the oil and methanol
during transesterification, as it acts as an interface for the polar and [1] U.S.. Energy information adminisatration, short-term energy outlook (STEO).
https://www.eia.gov/outlooks/steo/pdf/steo_full.pdf; 2018, Accessed date: 14
non-polar phases. However, separation of cosolvent will incur addi- January 2019.
tional processing cost, which is not economical. Cosolvent has also [2] Renewable energy policy network for the 21st century, renewables global status
showed potential in terms of enhancing phase separation between report. www.ren21.net; 2018, Accessed date: 12 February 2019.
[3] U.S. Energy Information Administration. International energy outlook. www.eia.
biodiesel and glycerol, as it increases the difference in specific gravity gov/ieo; 2017, Accessed date: 12 February 2019.
and relative viscosity between the two phases. Microwave reactor has [4] Watts N, Amann M, Ayeb-Karlsson S, Belesova K, Bouley T, Boykoff M, Byass P, Cai
proven to be more effective when coupled with another intensification W, Campbell-Lendrum D, Chambers J, Cox PM, Daly M, Dasandi N, Davies M,
Depledge M, Depoux A, Dominguez-Salas P, Drummond P, Ekins P, Flahault A,
method. Nonetheless, commercial microwave reactors have a fixed
Frumkin H, Georgeson L, Ghanei M, Grace D, Graham H, Grojsman R, Haines A,
operation frequency of 2.45 GHz, which may not be efficient for large Hamilton I, Hartinger S, Johnson A, Kelman I, Kiesewetter G, Kniveton D, Liang L,
volume biodiesel processing due to a low penetration depth. Therefore, Lott M, Lowe R, Mace G, Odhiambo Sewe M, Maslin M, Mikhaylov S, Milner J,
microwave-assisted biodiesel reactor are confined to small batch pro- Latifi AM, Moradi-Lakeh M, Morrissey K, Murray K, Neville T, Nilsson M,
Oreszczyn T, Owfi F, Pencheon D, Pye S, Rabbaniha M, Robinson E, Rocklöv J,
cessing, which has scalability issue and slow biodiesel output. Thus, Schütte S, Shumake-Guillemot J, Steinbach R, Tabatabaei M, Wheeler N, Wilkinson
continuous reactor is introduced to achieve a higher biodiesel P, Gong P, Montgomery H, Costello A. The Lancet Countdown on health and cli-
throughput, with low residence time to compensate for the limited mate change: from 25 years of inaction to a global transformation for public
health. Lancet 2018;391:581–630. https://doi.org/10.1016/S0140-6736(17)
microwave reactor footprint. 32464-9.
Beside the challenges in the technological aspect of biodiesel pro- [5] Watts N, Amann M, Arnell N, Ayeb-Karlsson S, Belesova K, Berry H, Bouley T,
duction, the economic climate is also a significant factor in current Boykoff M, Byass P, Cai W, Campbell-Lendrum D, Chambers J, Daly M, Dasandi N,
Davies M, Depoux A, Dominguez-Salas P, Drummond P, Ebi KL, Ekins P, Montoya
biofuel production industry. As shown from the techno-economic ana- LF, Fischer H, Georgeson L, Grace D, Graham H, Hamilton I, Hartinger S, Hess J,
lysis, some countries such as Indonesia and Malaysia have huge po- Kelman I, Kiesewetter G, Kjellstrom T, Kniveton D, Lemke B, Liang L, Lott M, Lowe
tentials for biodiesel production and consumption, but there is no in- R, Sewe MO, Martinez-Urtaza J, Maslin M, McAllister L, Mikhaylov SJ, Milner J,
Moradi-Lakeh M, Morrissey K, Murray K, Nilsson M, Neville T, Oreszczyn T, Owfi
centive for replacing traditional diesel at a higher blend level due to the F, Pearman O, Pencheon D, Pye S, Rabbaniha M, Robinson E, Rocklöv J, Saxer O,
industry being financially not viable [187]. Selectivity in feedstock can Schütte S, Semenza JC, Shumake-Guillemot J, Steinbach R, Tabatabaei M, Tomei J,
improve the efficacy of biodiesel as a driver to reduce greenhouse gases, Trinanes J, Wheeler N, Wilkinson P, Gong P, Montgomery H, Costello A. The 2018
report of the Lancet Countdown on health and climate change: shaping the health
since the emissions of different biodiesel can vary with a significant
of nations for centuries to come. Lancet 2018;392:2479–514. https://doi.org/10.
margin. 1016/S0140-6736(18)32594-7.
[6] Katowice climate change conference. UNFCCC; 2018https://unfccc.int/katowice,
Accessed date: 14 January 2019 – december(n.d.).
[7] Hugh M. Preventing the progression of climate change: one drug or polypill?
6. Conclusion and future prospects
Biofuel Res. J. 2017;13:536. https://doi.org/10.18331/BRJ2017.4.1.2.
[8] Demirbas A. Progress and recent trends in biodiesel fuels. Energy Convers Manag
In recent years, biodiesel has gained significant traction aided by 2009;50:14–34. https://doi.org/10.1016/j.enconman.2008.09.001.
the advancement in production technologies. However, the number of [9] Banković-Ilić IB, Stamenković OS, Veljković VB. Biodiesel production from non-
edible plant oils. Renew Sustain Energy Rev 2012;16:3621–47. https://doi.org/10.
biodiesel-related research is still susceptible to fluctuations, due to the 1016/j.rser.2012.03.002.
volatility of crude oil prices which directly impacts the development of [10] Jeon J, Park S. Effect of injection pressure on soot formation/oxidation char-
biodiesel technologies. Also, biodiesel production methodologies still acteristics using a two-color photometric method in a compression-ignition engine
fueled with biodiesel blend (B20). Appl Therm Eng 2018;131:284–94. https://doi.
face challenges from different aspects, such as mass transfer limitation, org/10.1016/J.APPLTHERMALENG.2017.12.005.
long residence time, scalability of technology, and expensive in- [11] ASTM International. ASTM D7467-18b, standard specification for diesel fuel oil,
strumentations. Apart from developing novel technologies in biodiesel biodiesel blend (B6 to B20). 2018. https://doi.org/10.1520/D7467-18B. West
Conshohocken, PA.
production, optimisation of the existing technology still requires

28
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

[12] Ng JH, Ng HK, Gan S. Advances in biodiesel fuel for application in compression (Jatropha curcas) with high free fatty acids: an optimized process. Biomass
ignition engines. Clean Technol Environ Policy 2010;12:459–93. https://doi.org/ Bioenergy 2007;31:569–75. https://doi.org/10.1016/J.BIOMBIOE.2007.03.003.
10.1007/s10098-009-0268-6. [40] Sharma YC, Singh B. Development of biodiesel: current scenario. Renew Sustain
[13] Ng J-HH, Ng HK, Gan S. Characterisation of engine-out responses from a light-duty Energy Rev 2009;13:1646–51. https://doi.org/10.1016/j.rser.2008.08.009.
diesel engine fuelled with palm methyl ester (PME). Appl Energy 2012;90:58–67. [41] Ong HC, Masjuki HH, Mahlia TMI, Silitonga AS, Chong WT, Leong KY.
https://doi.org/10.1016/j.apenergy.2011.01.028. Optimization of biodiesel production and engine performance from high free fatty
[14] Chuah LF, Klemeš JJ, Yusup S, Bokhari A, Akbar MM, Rí J, Kleme S C J, Yusup S, acid Calophyllum inophyllum oil in CI diesel engine. Energy Convers Manag
Bokhari A, Akbar MM, Klemeš JJ, Yusup S, Bokhari A, Akbar MM. A review of 2014;81:30–40. https://doi.org/10.1016/J.ENCONMAN.2014.01.065.
cleaner intensification technologies in biodiesel production. J Clean Prod [42] Darnoko D, Cheryan M. Kinetics of palm oil transesterification in a batch reactor. J
2016;146:181–93. https://doi.org/10.1016/j.jclepro.2016.05.017. Am Oil Chem Soc 2000;77:1263–7. https://doi.org/10.1007/s11746-000-0198-y.
[15] Shan R, Lu L, Shi Y, Yuan H, Shi J. Catalysts from renewable resources for biodiesel [43] Dias JM, Alvim-Ferraz MCMM, Almeida MF. Comparison of the performance of
production. Energy Convers Manag 2018;178:277–89. https://doi.org/10.1016/J. different homogeneous alkali catalysts during transesterification of waste and
ENCONMAN.2018.10.032. virgin oils and evaluation of biodiesel quality. Fuel 2008;87:3572–8. https://doi.
[16] Noureddini H, Zhu D. Kinetics of transesterification of soybean oil. J Am Oil Chem org/10.1016/j.fuel.2008.06.014.
Soc 1997;74:1457–63. https://doi.org/10.1007/s11746-997-0254-2. [44] Foon CS, May CY, Ngan MA, Hock CC, Sit Foon C, Yuen May C, Ngan MA, Cheng
[17] World Intellectual Property Organization. WIPO - search international and na- Hock C. Kinetics study on transesterification of palm oil. J Oil Palm Res
tional patent collections (n.d.). https://patentscope.wipo.int/search/en/search.jsf, 2004;16:19–29http://palmoilis.mpob.gov.my/publications/joprv16n2-cheng.pdf,
Accessed date: 17 January 2019. Accessed date: 26 February 2018.
[18] SCOPUS. Scopus preview - Scopus - welcome to Scopus (n.d.). https://www. [45] Rashid U, Anwar F. Production of biodiesel through optimized alkaline-catalyzed
scopus.com/results/handle.uri, Accessed date: 18 January 2019. transesterification of rapeseed oil. Fuel 2008;87:265–73. https://doi.org/10.1016/
[19] U.S. Energy Information Administration. Spot prices for crude oil and petroleum j.fuel.2007.05.003.
products (n.d.). https://www.eia.gov/dnav/pet/pet_pri_spt_s1_a.htm, Accessed [46] Chung KH, Kim J, Lee KY. Biodiesel production by transesterification of duck
date: 18 January 2019. tallow with methanol on alkali catalysts. Biomass Bioenergy 2009;33:155–8.
[20] The World Bank. World Bank open data | data (n.d.). https://data.worldbank.org/, https://doi.org/10.1016/j.biombioe.2008.04.014.
Accessed date: 18 January 2019. [47] Zhang L, Sheng B, Xin Z, Liu Q, Sun S. Kinetics of transesterification of palm oil
[21] Che Mat S, Idroas MY, Hamid MF, Zainal ZA. Performance and emissions of and dimethyl carbonate for biodiesel production at the catalysis of heterogeneous
straight vegetable oils and its blends as a fuel in diesel engine: a review. Renew base catalyst. Bioresour Technol 2010;101:8144–50. https://doi.org/10.1016/j.
Sustain Energy Rev 2018;82:808–23. https://doi.org/10.1016/J.RSER.2017.09. biortech.2010.05.069.
080. [48] Uzun BB, Kiliç M, Özbay N, Pütün AE, Pütün E. Biodiesel production from waste
[22] Hoang AT, Le AT. Trilateral correlation of spray characteristics, combustion frying oils: optimization of reaction parameters and determination of fuel prop-
parameters, and deposit formation in the injector hole of a diesel engine running erties. Energy 2012;44:347–51. https://doi.org/10.1016/j.energy.2012.06.024.
on preheated Jatropha oil and fossil diesel fuel. Biofuel Res J 2019;21:909–19. [49] Fadhil AB, Ali LH. Alkaline-catalyzed transesterification of Silurus triostegus
https://doi.org/10.18331/BRJ2019.6.1.2. Heckel fish oil: optimization of transesterification parameters. Renew Energy
[23] Sharma YC, Singh B. Development of biodiesel from karanja, a tree found in rural 2013;60:481–8. https://doi.org/10.1016/j.renene.2013.04.018.
India. Fuel 2008;87:1740–2. https://doi.org/10.1016/j.fuel.2007.08.001. [50] Thoai DN, Tongurai C, Prasertsit K, Kumar A. A novel two-step transesterification
[24] Aghbashlo M, Tabatabaei M, Rastegari H, Ghaziaskar HS. Exergy-based sustain- process catalyzed by homogeneous base catalyst in the first step and hetero-
ability analysis of acetins synthesis through continuous esterification of glycerol in geneous acid catalyst in the second step. Fuel Process Technol 2017;168:97–104.
acetic acid using Amberlyst®36 as catalyst. J Clean Prod 2018;183:1265–75. https://doi.org/10.1016/J.FUPROC.2017.08.014.
https://doi.org/10.1016/J.JCLEPRO.2018.02.218. [51] Gebremariam SN, Marchetti JM. Biodiesel production through sulfuric acid cata-
[25] Aghbashlo M, Tabatabaei M, Jazini H, Ghaziaskar HS. Exergoeconomic and ex- lyzed transesterification of acidic oil: techno economic feasibility of different
ergoenvironmental co-optimization of continuous fuel additives (acetins) synthesis process alternatives. Energy Convers Manag 2018;174:639–48. https://doi.org/
from glycerol esterification with acetic acid using Amberlyst 36 catalyst. Energy 10.1016/J.ENCONMAN.2018.08.078.
Convers Manag 2018;165:183–94. https://doi.org/10.1016/J.ENCONMAN.2018. [52] Wang Y-T, Fang Z, Yang X-X, Yang Y-T, Luo J, Xu K, Bao G-R. One-step production
03.054. of biodiesel from Jatropha oils with high acid value at low temperature by mag-
[26] Sharma YC, Singh B, Upadhyay SN. Advancements in development and char- netic acid-base amphoteric nanoparticles. Chem Eng J 2018;348:929–39. https://
acterization of biodiesel: a review. Fuel 2008;87:2355–73. https://doi.org/10. doi.org/10.1016/J.CEJ.2018.05.039.
1016/j.fuel.2008.01.014. [53] Thiruvengadaravi KV, Nandagopal J, Baskaralingam P, Sathya Selva Bala V,
[27] Leung DYC, Wu X, Leung MKH. A review on biodiesel production using catalyzed Sivanesan S. Acid-catalyzed esterification of karanja (Pongamia pinnata) oil with
transesterification. Appl Energy 2010;87:1083–95. https://doi.org/10.1016/j. high free fatty acids for biodiesel production. Fuel 2012;98:1–4. https://doi.org/
apenergy.2009.10.006. 10.1016/J.FUEL.2012.02.047.
[28] Kee Lam M, Teong Lee K, Rahman Mohamed A, Lam MK, Lee KT, Mohamed AR. [54] Farag HA, El-maghraby A, Taha NA. Optimization of factors affecting esterification
Homogeneous, heterogeneous and enzymatic catalysis for transesterification of of mixed oil with high percentage of free fatty acid. Fuel Process Technol
high free fatty acid oil (waste cooking oil) to biodiesel: a review. Biotechnol Adv 2011;92:507–10. https://doi.org/10.1016/j.fuproc.2010.11.004.
2010;28:500–18. https://doi.org/10.1016/j.biotechadv.2010.03.002. [55] Aranda DAG, Santos RTP, Tapanes NCO, Ramos ALD, Antunes OAC. Acid-cata-
[29] Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic lyzed homogeneous esterification reaction for biodiesel production from palm
catalysis for transesterification of high free fatty acid oil (waste cooking oil) to fatty acids. Catal Lett 2008;122:20–5. https://doi.org/10.1007/s10562-007-
biodiesel: a review. Biotechnol Adv 2010;28:500–18. https://doi.org/10.1016/J. 9318-z.
BIOTECHADV.2010.03.002. [56] Khan MA, Yusup S, Ahmad MM. Acid esterification of a high free fatty acid crude
[30] Eze VC, Phan AN, Harvey AP. A more robust model of the biodiesel reaction, al- palm oil and crude rubber seed oil blend: optimization and parametric analysis.
lowing identification of process conditions for significantly enhanced rate and Biomass Bioenergy 2010;34:1751–6. https://doi.org/10.1016/j.biombioe.2010.
water tolerance. Bioresour Technol 2014;156:222–31. https://doi.org/10.1016/j. 07.006.
biortech.2014.01.028. [57] Mohammad Fauzi AH, Saidina Amin NA. Optimization of oleic acid esterification
[31] Eze VC, Phan AN, Harvey AP. Intensified one-step biodiesel production from high catalyzed by ionic liquid for green biodiesel synthesis. Energy Convers Manag
water and free fatty acid waste cooking oils. Fuel 2018;220:567–74. https://doi. 2013;76:818–27. https://doi.org/10.1016/j.enconman.2013.08.029.
org/10.1016/J.FUEL.2018.02.050. [58] Jacobson K, Gopinath R, Meher LC, Dalai AK. Solid acid catalyzed biodiesel pro-
[32] Ramadhas ASS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA duction from waste cooking oil. Appl Catal B Environ 2008;85:86–91. https://doi.
rubber seed oil. Fuel 2005;84:335–40. https://doi.org/10.1016/j.fuel.2004.09. org/10.1016/J.APCATB.2008.07.005.
016. [59] Mansir N, Taufiq-Yap YH, Rashid U, Lokman IM. Investigation of heterogeneous
[33] Sahoo PK, Das LM, Babu MKG, Naik SN. Biodiesel development from high acid solid acid catalyst performance on low grade feedstocks for biodiesel production: a
value polanga seed oil and performance evaluation in a CI engine. Fuel review. Energy Convers Manag 2016;141:171–82. https://doi.org/10.1016/j.
2007;86:448–54. https://doi.org/10.1016/J.FUEL.2006.07.025. enconman.2016.07.037.
[34] Gui MM, Lee KT, Bhatia S. Feasibility of edible oil vs. non-edible oil vs. waste [60] Avhad MR, Marchetti JM. A review on recent advancement in catalytic materials
edible oil as biodiesel feedstock. Energy 2008;33:1646–53. https://doi.org/10. for biodiesel production. Renew Sustain Energy Rev 2015;50:696–718. https://
1016/j.energy.2008.06.002. doi.org/10.1016/j.rser.2015.05.038.
[35] Freedman B, Butterfield RO, Pryde EH. Transesterification kinetics of soybean oil. [61] Kouzu M, Hidaka JS. Transesterification of vegetable oil into biodiesel catalyzed
J Am Oil Chem Soc 1986;63:1375–80. https://doi.org/10.1007/BF02679606. by CaO: a review. Fuel 2012;93:1–12. https://doi.org/10.1016/j.fuel.2011.09.
[36] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol 015.
1999;70:1–15. https://doi.org/10.1016/S0960-8524(99)00025-5. [62] Kouzu M, ya Yamanaka S, suke Hidaka J, Tsunomori M. Heterogeneous catalysis of
[37] Canakci M, Van Gerpen J. Biodiesel production from oils and fats with high free calcium oxide used for transesterification of soybean oil with refluxing methanol.
fatty acids. Trans ASAE 2001;44:1429–36http://biodieseleducation.org/ Appl Catal A Gen 2009;355:94–9. https://doi.org/10.1016/j.apcata.2008.12.003.
Literature/Journal/2001_Canakci_Biodiesel_production.pdf, Accessed date: 7 May [63] Granados ML, Poves MDZ, Alonso DM, Mariscal R, Galisteo FC, Moreno-Tost R,
2018. Santamaría J, Fierro JLG. Biodiesel from sunflower oil by using activated calcium
[38] Zhang Y, Dubé M, McLean D, Kates M. Biodiesel production from waste cooking oxide. Appl Catal B Environ 2007;73:317–26. https://doi.org/10.1016/j.apcatb.
oil: 1. Process design and technological assessment. Bioresour Technol 2006.12.017.
2003;89:1–16. https://doi.org/10.1016/S0960-8524(03)00040-3. [64] Jeon H, Kim DJ, Kim SJ, Kim JH. Synthesis of mesoporous MgO catalyst templated
[39] Kumar Tiwari A, Kumar A, Raheman H. Biodiesel production from jatropha oil by a PDMS-PEO comb-like copolymer for biodiesel production. Fuel Process

29
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

Technol 2013;116:325–31. https://doi.org/10.1016/j.fuproc.2013.07.013. [90] Shu Q, Zou W, He J, Lesmana H, Zhang C, Zou L, Wang Y. Preparation of the F−-
[65] Kouzu M, Kasuno T, Tajika M, Sugimoto Y, Yamanaka S, Hidaka J. Calcium oxide SO42-/MWCNTs catalyst and kinetic studies of the biodiesel production via es-
as a solid base catalyst for transesterification of soybean oil and its application to terification reaction of oleic acid and methanol. Renew Energy 2019;135:836–45.
biodiesel production. Fuel 2008;87:2798–806. https://doi.org/10.1016/j.fuel. https://doi.org/10.1016/J.RENENE.2018.12.067.
2007.10.019. [91] Liu Y, Pinnavaia TJ. Assembly of wormhole aluminosilicate mesostructures from
[66] Koberg M, Abu-Much R, Gedanken A. Optimization of bio-diesel production from zeolite seeds. J Mater Chem 2004;14:1099–103. https://doi.org/10.1039/
soybean and wastes of cooked oil: combining dielectric microwave irradiation and b315193j.
a SrO catalyst. Bioresour Technol 2011;102:1073–8. https://doi.org/10.1016/j. [92] Belviso C, Cavalcante F, Lettino A, Fiore S. A and X-type zeolites synthesised from
biortech.2010.08.055. kaolinite at low temperature. Appl Clay Sci 2013;80–81:162–8. https://doi.org/
[67] Wan Z, Hameed BHH. Transesterification of palm oil to methyl ester on activated 10.1016/j.clay.2013.02.003.
carbon supported calcium oxide catalyst. Bioresour Technol 2011;102:2659–64. [93] Tesser R, Casale L, Verde D, Di Serio M, Santacesaria E. Kinetics and modeling of
https://doi.org/10.1016/j.biortech.2010.10.119. fatty acids esterification on acid exchange resins. Chem Eng J 2010;157:539–50.
[68] Martinez-Guerra E, Gude VG. Transesterification of used vegetable oil catalyzed by https://doi.org/10.1016/j.cej.2009.12.050.
barium oxide under simultaneous microwave and ultrasound irradiations. Energy [94] Honkela ML, Root A, Lindblad M, Krause AOI. Comparison of ion-exchange resin
Convers Manag 2014;88:633–40. https://doi.org/10.1016/j.enconman.2014.08. catalysts in the dimerisation of isobutene. Appl Catal A Gen 2005;295:216–23.
060. https://doi.org/10.1016/j.apcata.2005.08.023.
[69] Liu K, Wang R, Yu M. An efficient, recoverable solid base catalyst of magnetic [95] Darnoko D, Cheryan M. Continuous production of palm methyl esters. J Am Oil
bamboo charcoal: preparation, characterization, and performance in biodiesel Chem Soc 2000;77:1269–72. https://doi.org/10.1007/s11746-000-0199-x.
production. Renew Energy 2018;127:531–8. https://doi.org/10.1016/J.RENENE. [96] Leevijit T, Tongurai C, Prateepchaikul G, Wisutmethangoon W. Performance test of
2018.04.092. a 6-stage continuous reactor for palm methyl ester production. Bioresour Technol
[70] Essamlali Y, Amadine O, Fihri A, Zahouily M. Sodium modified fluorapatite as a 2008;99:214–21. https://doi.org/10.1016/j.biortech.2006.11.052.
sustainable solid bi-functional catalyst for biodiesel production from rapeseed oil. [97] Prateepchaikul G, Somnuk K, Allen M. Design and testing of continuous acid-
Renew Energy 2019;133:1295–307. https://doi.org/10.1016/J.RENENE.2018.08. catalyzed esterification reactor for high free fatty acid mixed crude palm oil. Fuel
103. Process Technol 2009;90:784–9. https://doi.org/10.1016/j.fuproc.2009.03.008.
[71] Borah MJ, Devi A, Borah R, Deka D. Synthesis and application of Co doped ZnO as [98] Ilmi M, Kloekhorst A, Winkelman JGM, Euverink GJW, Hidayat C, Heeres HJ.
heterogeneous nanocatalyst for biodiesel production from non-edible oil. Renew Process intensification of catalytic liquid-liquid solid processes: continuous bio-
Energy 2019;133:512–9. https://doi.org/10.1016/J.RENENE.2018.10.069. diesel production using an immobilized lipase in a centrifugal contactor separator.
[72] Taufiq-Yap YH, Teo SH, Rashid U, Islam A, Hussien MZ, Lee KT. Transesterification Chem Eng J 2017;321:76–85. https://doi.org/10.1016/J.CEJ.2017.03.070.
of Jatropha curcas crude oil to biodiesel on calcium lanthanum mixed oxide cat- [99] Yasvanthrajan N, Nabera A, Salike S, Valan DT, Sivakumar P, Muthukumar K,
alyst: effect of stoichiometric composition. Energy Convers Manag Arunagiri A. An overview on the process intensification of microchannel reactors
2014;88:1290–6. https://doi.org/10.1016/J.ENCONMAN.2013.12.075. for biodiesel production. Chem Eng Process Process Intensif 2018. https://doi.org/
[73] Nizami A-S, Rehan M. Towards nanotechnology-based biofuel industry. Biofuel 10.1016/J.CEP.2018.12.008.
Res J 2018;18:798–9. https://doi.org/10.18331/BRJ2018.5.2.2. [100] Maddikeri GL, Pandit AB, Gogate PR. Intensification approaches for biodiesel
[74] Akubude VC, Nwaigwe KN, Dintwa E. Production of biodiesel from microalgae via synthesis from waste cooking oil: a review. Ind Eng Chem Res 2012;51:14610–28.
nanocatalyzed transesterification process: a review. Mater Sci Energy Technol https://doi.org/10.1021/ie301675j.
2019;2:216–25. https://doi.org/10.1016/J.MSET.2018.12.006. [101] Madhawan A, Sharma V, Kuila A, Arora A, Das J. Microreactor technology for
[75] Mardhiah HH, Ong HC, Masjuki HH, Lim S, Lee HV. A review on latest develop- biodiesel production: a review. Biomass Convers Bioref 2017;8:485–96. https://
ments and future prospects of heterogeneous catalyst in biodiesel production from doi.org/10.1007/s13399-017-0296-0.
non-edible oils. Renew Sustain Energy Rev 2017;67:1225–36. https://doi.org/10. [102] Santana HS, Silva JL, Taranto OP. Numerical simulation of mixing and reaction of
1016/j.rser.2016.09.036. Jatropha curcas oil and ethanol for synthesis of biodiesel in micromixers. Chem
[76] Singh A, Gaurav K. Advancement in catalysts for transesterification in the pro- Eng Sci 2015;132:159–68. https://doi.org/10.1016/j.ces.2015.04.014.
duction of biodiesel: a review. J Biochem Technol 2018;7:1148–58https://pdfs. [103] Mansur EA, YE M, WANG Y, DAI Y. A state-of-the-art review of mixing in mi-
semanticscholar.org/163a/743c7b1ac31f5f5c594c264f3c69bf889441.pdf, crofluidic mixers. Chin J Chem Eng 2008;16:503–16. https://doi.org/10.1016/
Accessed date: 17 August 2019. S1004-9541(08)60114-7.
[77] Su F, Guo Y. Advancements in solid acid catalysts for biodiesel production. Green [104] Wen Z, Yu X, Tu ST, Yan J, Dahlquist E. Intensification of biodiesel synthesis using
Chem 2014;16:2934–57. https://doi.org/10.1039/C3GC42333F. zigzag micro-channel reactors. Bioresour Technol 2009;100:3054–60. https://doi.
[78] Fu B, Gao L, Niu L, Wei R, Xiao G. Biodiesel from waste cooking oil via hetero- org/10.1016/j.biortech.2009.01.022.
geneous superacid catalyst SO4\ 2-/ZrO2. Energy Fuel 2009;23:569–72. https:// [105] Sun P, Wang B, Yao J, Zhang L, Xu N. Fast synthesis of biodiesel at high
doi.org/10.1021/ef800751z. throughput in microstructured reactors. Ind Eng Chem Res 2010;49:1259–64.
[79] Patel A, Brahmkhatri V, Singh N. Biodiesel production by esterification of free fatty https://doi.org/10.1021/ie901320s.
acid over sulfated zirconia. Renew Energy 2013;51:227–33. https://doi.org/10. [106] Tanawannapong Y, Kaewchada A, Jaree A. Biodiesel production from waste
1016/j.renene.2012.09.040. cooking oil in a microtube reactor. J Ind Eng Chem 2013;19:37–41. https://doi.
[80] Lam MK, Lee KT, Mohamed AR. Sulfated tin oxide as solid superacid catalyst for org/10.1016/j.jiec.2012.07.007.
transesterification of waste cooking oil: an optimization study. Appl Catal B [107] Aghel B, Rahimi M, Sepahvand A, Alitabar M, Ghasempour HR. Using a wire coil
Environ 2009;93:134–9. https://doi.org/10.1016/j.apcatb.2009.09.022. insert for biodiesel production enhancement in a microreactor. Energy Convers
[81] Suppes GJ, Dasari MA, Doskocil EJ, Mankidy PJ, Goff MJ. Transesterification of Manag 2014;84:541–9. https://doi.org/10.1016/j.enconman.2014.05.009.
soybean oil with zeolite and metal catalysts. Appl Catal A Gen 2004;257:213–23. [108] Kaewchada A, Pungchaicharn S, Jaree A. Transesterification of palm oil in a mi-
https://doi.org/10.1016/j.apcata.2003.07.010. crotube reactor. Can J Chem Eng 2016;94:859–64. https://doi.org/10.1002/cjce.
[82] Babajide O, Musyoka N, Petrik L, Ameer F. Novel zeolite Na-X synthesized from fly 22464.
ash as a heterogeneous catalyst in biodiesel production. Catal Today [109] Chueluecha N, Kaewchada A, Jaree A. Biodiesel synthesis using heterogeneous
2012;190:54–60. https://doi.org/10.1016/j.cattod.2012.04.044. catalyst in a packed-microchannel. Energy Convers Manag 2016;141:145–54.
[83] Kusuma RI, Hadinoto JP, Ayucitra A, Soetaredjo FE, Ismadji S. Natural zeolite from https://doi.org/10.1016/j.enconman.2016.07.020.
Pacitan Indonesia, as catalyst support for transesterification of palm oil. Appl Clay [110] Mohadesi M, Aghel B, Maleki M, Ansari A. Production of biodiesel from waste
Sci 2013;74:121–6. https://doi.org/10.1016/j.clay.2012.04.021. cooking oil using a homogeneous catalyst: study of semi-industrial pilot of mi-
[84] Doyle AM, Albayati TM, Abbas AS, Alismaeel ZT. Biodiesel production by ester- croreactor. Renew Energy 2019;136:677–82. https://doi.org/10.1016/J.RENENE.
ification of oleic acid over zeolite Y prepared from kaolin. Renew Energy 2019.01.039.
2016;97:19–23. https://doi.org/10.1016/j.renene.2016.05.067. [111] Hessel V, Hardt S, Löwe H, Schönfeld F. Laminar mixing in different interdigital
[85] Özbay N, Oktar N, Tapan NA. Esterification of free fatty acids in waste cooking oils micromixers: I. Experimental characterization. AIChE J 2003;49:566–77. https://
(WCO): role of ion-exchange resins. Fuel 2008;87:1789–98. https://doi.org/10. doi.org/10.1002/aic.690490304.
1016/j.fuel.2007.12.010. [112] Alford R, Burns M, Burns N. Dixon rings - a revolutionary random column packing.
[86] Fu J, Chen L, Lv P, Yang L, Yuan Z. Free fatty acids esterification for biodiesel Filtration 2011;11:218–23.
production using self-synthesized macroporous cation exchange resin as solid acid [113] Brennen CE, Oxford NY. Cavitation and bubble dynamics. https://authors.library.
catalyst. Fuel 2015;154:1–8. https://doi.org/10.1016/j.fuel.2015.03.048. caltech.edu/25017/5/BUBBOOK.pdf; 1995, Accessed date: 22 January 2019.
[87] Gardy J, Osatiashtiani A, Céspedes O, Hassanpour A, Lai X, Lee AF, Wilson K, [114] Wu P, Bai L, Lin W, Wang X. Mechanism and dynamics of hydrodynamic-acoustic
Rehan M. A magnetically separable SO4/Fe-Al-TiO2 solid acid catalyst for bio- cavitation (HAC). Ultrason Sonochem 2018;49:89–96. https://doi.org/10.1016/J.
diesel production from waste cooking oil. Appl Catal B Environ 2018;234:268–78. ULTSONCH.2018.07.021.
https://doi.org/10.1016/J.APCATB.2018.04.046. [115] Lauterborn W. Lauterborn W, editor. Cavitation and coherent optics BT - cavita-
[88] Rahimzadeh H, Tabatabaei M, Aghbashlo M, Kazemi Shariat Pahani H, Rashidi A, tion and inhomogeneities in underwater acoustics: proceedings of the first inter-
Amir Hossein Goli S, Mostafaei M, Ardjmand M, Nizami A-S. Potential of acid- national conference, göttingen, fed. Rep. Of Germany, Berlin, Heidelberg: Springer
activated bentonite and SO 3 H-functionalized MWCNTs for biodiesel production Berlin Heidelberg; 1980. p. 3–12. https://doi.org/10.1007/978-3-642-51070-0_1.
from residual olive oil under biorefinery Scheme. Front Energy Res 2018;6:137. july 9–11, 1979.
https://doi.org/10.3389/fenrg.2018.00137. Www.Frontiersin.Org. [116] Gogate PR, Tayal RK, Pandit AB. Cavitation: a technology on the horizon. Curr Sci
[89] Alcañiz-Monge J, El Bakkali B, Trautwein G, Reinoso S. Zirconia-supported tung- 2006;91:35–46.
stophosphoric heteropolyacid as heterogeneous acid catalyst for biodiesel pro- [117] Kojima Y, Imazu H, Nishida K. Physical and chemical characteristics of ultra-
duction. Appl Catal B Environ 2018;224:194–203. https://doi.org/10.1016/j. sonically-prepared water-in-diesel fuel: effects of ultrasonic horn position and
apcatb.2017.10.066. water content. Ultrason Sonochem 2014;21:722–8. https://doi.org/10.1016/J.

30
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

ULTSONCH.2013.09.019. [144] Motasemi F, Ani FN. A review on microwave-assisted production of biodiesel.


[118] Tan SX, Lim S, Ong HC, Pang YL. State of the art review on development of ul- Renew Sustain Energy Rev 2012;16:4719–33. https://doi.org/10.1016/j.rser.
trasound-assisted catalytic transesterification process for biodiesel production. 2012.03.069.
Fuel 2019;235:886–907. https://doi.org/10.1016/J.FUEL.2018.08.021. [145] Pozar D. Microwave engineering. fourth ed. 2005. doi:TK7876.P69 2011.
[119] Suslick KS. Sonochemistry. Kirk-othmer encycl. Chem. Technol. John Wiley & [146] Groisman Y, Gedanken A. Continuous flow, circulating microwave system and its
Sons, Inc.; 2000. https://doi.org/10.1002/0471238961.1915141519211912.a01. application in nanoparticle fabrication and biodiesel synthesis. J Phys Chem C
[120] Badday AS, Abdullah AZ, Lee KT, Khayoon MS. Intensification of biodiesel pro- 2008;112:8802–8. https://doi.org/10.1021/jp801409t.
duction via ultrasonic-assisted process: a critical review on fundamentals and re- [147] Barnard TM, Leadbeater NE, Boucher MB, Stencel LM, Wilhite BA. Continuous-
cent development. Renew Sustain Energy Rev 2012;16:4574–87. https://doi.org/ flow preparation of biodiesel using microwave heating. Energy Fuel
10.1016/j.rser.2012.04.057. 2007;21:1777–81. https://doi.org/10.1021/ef0606207.
[121] Aghbashlo M, Tabatabaei M, Hosseinpour S, Hosseini SS, Ghaffari A, Khounani Z, [148] Lertsathapornsuk V, Pairintra R, Aryusuk K, Krisnangkura K. Microwave assisted
Mohammadi P. Development and evaluation of a novel low power, high frequency in continuous biodiesel production from waste frying palm oil and its performance
piezoelectric-based ultrasonic reactor for intensifying the transesterification re- in a 100 kW diesel generator. Fuel Process Technol 2008;89:1330–6. https://doi.
action. Biofuel Res J 2016;12:528–35. https://doi.org/10.18331/BRJ2016.3.4.7. org/10.1016/j.fuproc.2008.05.024.
[122] Ji J, Wang J, Li Y, Yu Y, Xu Z. Preparation of biodiesel with the help of ultrasonic [149] Leadbeater NE, Barnard TM, Stencel LM. Batch and continuous-flow preparation
and hydrodynamic cavitation. Ultrasonics 2006;44:e411–4. https://doi.org/10. of biodiesel derived from butanol and facilitated by microwave heating. Energy
1016/j.ultras.2006.05.020. Fuel 2008;22:2005–8. https://doi.org/10.1021/ef700748t.
[123] Stavarache C, Vinatoru M, Maeda Y, Bandow H. Ultrasonically driven continuous [150] Yuan H, Yang BL, Zhu GL. Synthesis of biodiesel using microwave absorption
process for vegetable oil transesterification. Ultrason Sonochem 2007;14:413–7. catalysts. Energy Fuel 2009;23:548–52. https://doi.org/10.1021/ef800577j.
https://doi.org/10.1016/j.ultsonch.2006.09.014. [151] Zhang S, Zu YG, Fu YJ, Luo M, Zhang DY, Efferth T. Rapid microwave-assisted
[124] Mahamuni NN, Adewuyi YG. Optimization of the synthesis of biodiesel via ul- transesterification of yellow horn oil to biodiesel using a heteropolyacid solid
trasound-enhanced base-catalyzed transesterification of soybean oil using a mul- catalyst. Bioresour Technol 2010;101:931–6. https://doi.org/10.1016/j.biortech.
tifrequency ultrasonic reactor. Energy Fuel 2009;23:2757–66. https://doi.org/10. 2009.08.069.
1021/ef900047j. [152] Patil PD, Gude VG, Mannarswamy A, Cooke P, Munson-McGee S, Nirmalakhandan
[125] Mootabadi H, Salamatinia B, Bhatia S, Abdullah AZ. Ultrasonic-assisted biodiesel N, Lammers P, Deng S. Optimization of microwave-assisted transesterification of
production process from palm oil using alkaline earth metal oxides as the het- dry algal biomass using response surface methodology. Bioresour Technol
erogeneous catalysts. Fuel 2010;89:1818–25. https://doi.org/10.1016/J.FUEL. 2011;102:1399–405. https://doi.org/10.1016/J.BIORTECH.2010.09.046.
2009.12.023. [153] Liao C-C, Chung T-W. Optimization of process conditions using response surface
[126] Thanh LT, Okitsu K, Sadanaga Y, Takenaka N, Maeda Y, Bandow H. Ultrasound- methodology for the microwave-assisted transesterification of Jatropha oil with
assisted production of biodiesel fuel from vegetable oils in a small scale circulation KOH impregnated CaO as catalyst. Chem Eng Res Des 2013;91:2457–64. https://
process. Bioresour Technol 2010;101:639–45. https://doi.org/10.1016/J. doi.org/10.1016/J.CHERD.2013.04.009.
BIORTECH.2009.08.050. [154] Lin Y-C, Hsu K-H, Lin J-F. Rapid palm-biodiesel production assisted by a micro-
[127] Gharat N, Rathod VK. Ultrasound assisted enzyme catalyzed transesterification of wave system and sodium methoxide catalyst. Fuel 2014;115:306–11. https://doi.
waste cooking oil with dimethyl carbonate. Ultrason Sonochem 2013;20:900–5. org/10.1016/j.fuel.2013.07.022.
https://doi.org/10.1016/J.ULTSONCH.2012.10.011. [155] Mazubert A, Taylor C, Aubin J, Poux M. Key role of temperature monitoring in
[128] Mostafaei M, Ghobadian B, Barzegar M, Banakar A. Optimization of ultrasonic interpretation of microwave effect on transesterification and esterification reac-
assisted continuous production of biodiesel using response surface methodology. tions for biodiesel production. Bioresour Technol 2014;161:270–9. https://doi.
Ultrason Sonochem 2015;27:54–61. https://doi.org/10.1016/j.ultsonch.2015.04. org/10.1016/j.biortech.2014.03.011.
036. [156] Choedkiatsakul I, Ngaosuwan K, Assabumrungrat S, Mantegna S, Cravotto G.
[129] Subhedar PB, Gogate PR. Ultrasound assisted intensification of biodiesel produc- Biodiesel production in a novel continuous flow microwave reactor. Renew Energy
tion using enzymatic interesterification. Ultrason Sonochem 2016;29:67–75. 2015;83:25–9. https://doi.org/10.1016/j.renene.2015.04.012.
https://doi.org/10.1016/j.ultsonch.2015.09.006. [157] Bokhari A, Fatt Chuah L, Yusup S, Ahmad J, Rashid Shamsuddin M, Kiat Teng M.
[130] Bhangu SK, Gupta S, Ashokkumar M. Ultrasonic enhancement of lipase-catalysed Microwave-assisted methyl esters synthesis of Kapok (Ceiba pentandra) seed oil:
transesterification for biodiesel synthesis. Ultrason Sonochem 2017;34:305–9. parametric and optimization study. Biofuel Res J 2015;7:281–7. https://doi.org/
https://doi.org/10.1016/J.ULTSONCH.2016.06.005. 10.18331/BRJ2015.2.3.6.
[131] Aghbashlo M, Hosseinpour S, Tabatabaei M, Dadak A. Fuzzy modeling and opti- [158] Mohammad S, Ardebili S, Cravotto G. Flow-mode biodiesel production from palm
mization of the synthesis of biodiesel from waste cooking oil (WCO) by a low oil using a pressurized microwave reactor. Green Process Synth 2019;8:8–14.
power, high frequency piezo-ultrasonic reactor. Energy 2017;132:65–78. https:// https://doi.org/10.1515/gps-2017-0116.
doi.org/10.1016/J.ENERGY.2017.05.041. [159] Silitonga AS, Shamsuddin AH, Mahlia TMI, Milano J, Kusumo F, Siswantoro J,
[132] Batchelor GK. An introduction to fluid dynamics. Cambridge: Cambridge Dharma S, Sebayang AH, Masjuki HH, Ong HC. Biodiesel synthesis from Ceiba
University Press; 2000. https://doi.org/10.1017/CBO9780511800955. pentandra oil by microwave irradiation-assisted transesterification: ELM modeling
[133] Young FR. Cavitation. World Scientific; 1999. and optimization. Renew Energy 2020;146:1278–91. https://doi.org/10.1016/J.
[134] Ramachandran K, Suganya T, Nagendra Gandhi N, Renganathan S. Recent de- RENENE.2019.07.065.
velopments for biodiesel production by ultrasonic assist transesterification using [160] Gabriel C, Gabriel S, Grant EH, Grant EH, Halstead BSJ, Mingos D Michael P.
different heterogeneous catalyst: a review. Renew Sustain Energy Rev Dielectric parameters relevant to microwave dielectric heating. Chem Soc Rev
2013;22:410–8. https://doi.org/10.1016/J.RSER.2013.01.057. 1998;27:213–24. https://doi.org/10.1039/A827213Z.
[135] Kelkar MA, Gogate PR, Pandit AB. Intensification of esterification of acids for [161] Akerlof G. Dielectric constants of some organic solvent-water mixtures at various
synthesis of biodiesel using acoustic and hydrodynamic cavitation. Ultrason temperatures. J Am Chem Soc 1932;54:4125–39. https://doi.org/10.1021/
Sonochem 2008;15:188–94. https://doi.org/10.1016/j.ultsonch.2007.04.003. ja01350a001.
[136] Pal A, Verma A, Kachhwaha SS, Maji S. Biodiesel production through hydro- [162] Loong TC, Idris A, Chee Loong T, Idris A. One step transesterification of biodiesel
dynamic cavitation and performance testing. Renew Energy 2010;35:619–24. production using simultaneous cooling and microwave heating. J Clean Prod
https://doi.org/10.1016/j.renene.2009.08.027. 2016. https://doi.org/10.1016/j.jclepro.2016.03.155.
[137] Ghayal D, Pandit AB, Rathod VK. Optimization of biodiesel production in a hy- [163] Luu PD, Truong HT, Van Luu B, Pham LN, Imamura K, Takenaka N, Maeda Y.
drodynamic cavitation reactor using used frying oil. Ultrason Sonochem Production of biodiesel from Vietnamese Jatropha curcas oil by a co-solvent
2013;20:322–8. https://doi.org/10.1016/j.ultsonch.2012.07.009. method. Bioresour Technol 2014;173:309–16. https://doi.org/10.1016/j.
[138] Maddikeri GL, Gogate PR, Pandit AB. Intensified synthesis of biodiesel using hy- biortech.2014.09.114.
drodynamic cavitation reactors based on the interesterification of waste cooking [164] Thanh LT, Okitsu K, Sadanaga Y, Takenaka N, Maeda Y, Bandow H. A new co-
oil. Fuel 2014;137:285–92. https://doi.org/10.1016/j.fuel.2014.08.013. solvent method for the green production of biodiesel fuel – optimization and
[139] Chuah LF, Yusup S, Abd Aziz AR, Bokhari A, Klemeš JJ, Abdullah MZ. practical application. Fuel 2013;103:742–8. https://doi.org/10.1016/j.fuel.2012.
Intensification of biodiesel synthesis from waste cooking oil (Palm Olein) in a 09.029.
Hydrodynamic Cavitation Reactor: effect of operating parameters on methyl ester [165] Alhassan Y, Kumar N, Bugaje IM, Pali HS, Kathkar P. Co-solvents transesterifica-
conversion. Chem Eng Process Process Intensif 2015;95:235–40. https://doi.org/ tion of cotton seed oil into biodiesel: effects of reaction conditions on quality of
10.1016/j.cep.2015.06.018. fatty acids methyl esters. Energy Convers Manag 2014;84:640–8. https://doi.org/
[140] Bokhari A, Chuah LF, Yusup S, Akbar MM, Kamil RNM. Cleaner production of 10.1016/j.enconman.2014.04.080.
rubber seed oil methyl ester using a hydrodynamic cavitation: optimisation and [166] Maeda Y, Thanh LT, Imamura K, Izutani K, Okitsu K, Van Boi L, Ngoc Lan P, Tuan
parametric study. J Clean Prod 2016;136:31–41. https://doi.org/10.1016/J. NC, Yoo YE, Takenaka N. New technology for the production of biodiesel fuel.
JCLEPRO.2016.04.091. Green Chem 2011;13:1124–8. https://doi.org/10.1039/C1GC15049A.
[141] Bargole S, George S, Kumar Saharan V. Improved rate of transesterification re- [167] Stavarache C, Vinatoru M, Nishimura R, Maeda Y. Fatty acids methyl esters from
action in biodiesel synthesis using hydrodynamic cavitating devices of high throat vegetable oil by means of ultrasonic energy. Ultrason Sonochem 2005;12:367–72.
perimeter to flow area ratios. Chem Eng Process Process Intensif 2019;139:1–13. https://doi.org/10.1016/j.ultsonch.2004.04.001.
https://doi.org/10.1016/J.CEP.2019.03.012. [168] Boocock DGB, Konar SK, Mao V, Sidi H. Fast one-phase oil-rich processes for the
[142] Shah YT, Moholkar VS, Pandit AB. Cavitation reaction engineering. 1999. https:// preparation of vegetable oil methyl esters. Biomass Bioenergy 1996;11:43–50.
doi.org/10.1007/978-1-4615-4787-7. https://doi.org/10.1016/0961-9534(95)00111-5.
[143] Gunvachai K, Hassan MG, Shama G, Hellgardt K. A new solubility model to de- [169] Guan G, Kusakabe K, Sakurai N, Moriyama K. Transesterification of vegetable oil
scribe biodiesel formation kinetics. Process Saf Environ Prot 2007;85:383–9. to biodiesel fuel using acid catalysts in the presence of dimethyl ether. Fuel
https://doi.org/10.1205/PSEP07033. 2009;88:81–6. https://doi.org/10.1016/j.fuel.2008.07.021.

31
K.Y. Wong, et al. Renewable and Sustainable Energy Reviews 116 (2019) 109399

[170] Park J, Kim B, Chang YK, Lee JW. Wet in situ transesterification of microalgae [177] Ministry of Finance Malaysia. Budget 2019. www.treasury.gov.my; 2018,
using ethyl acetate as a co-solvent and reactant. Bioresour Technol Accessed date: 14 February 2019.
2017;230:8–14. https://doi.org/10.1016/J.BIORTECH.2017.01.027. [178] American Society for Testing and Materials. ASTM D6751-02 standard specifica-
[171] Taherkhani M, Sadrameli SM. An improvement and optimization study of bio- tion for biodiesel fuel (B100) blend stock for distillate fuels. 2011.
diesel production from linseed via in-situ transesterification using a co-solvent. [179] European Committee for Standardization. EN 14214 - fatty acid methyl esters
Renew Energy 2018;119:787–94. https://doi.org/10.1016/J.RENENE.2017.10. (FAME) for use in diesel engines and heating applications. 2008.
061. [180] Department of Standards Malaysia. Automotive fuels - palm Methyl Esters (PME)
[172] Todorović ZB, Troter DZ, Đokić-Stojanović DR, Veličković AV, Avramović JM, for diesel engines - requirements and test methods (First revision). 2014.
Stamenković OS, Veselinović LM, Veljković VB. Optimization of CaO-catalyzed [181] Bureau of Indian Standards. IS 15607:2005 bio-diesel (B100) blend stock for diesel
sunflower oil methanolysis with crude biodiesel as a cosolvent. Fuel fuel - specification. 2005.
2019;237:903–10. https://doi.org/10.1016/J.FUEL.2018.10.056. [182] Federal Register of Legislative Inststuments. Fuel Standard ( Biodiesel )
[173] Ng J-H, Teh JX, Wong KY, Wu KH, Chong CT. A techno-economical and auto- Determination 2003 vol. 2009. 2009. p. 1–7.
motive emissions impact study of global biodiesel usage in diesel engines. J Soc [183] Japan Industrial Standards. JIS K2390:2016 Automotive fuels-Fatty acid methyl
Automot Eng Malaysia 2017;1:124–36https://www.researchgate.net/profile/ ester (FAME) as blend stock. 2016. p. 1–8.
Cheng_Tung_Chong/publication/320012083_A_Techno-Economical_and_ [184] AGÊNCIA NACIONAL DO PETRÓLEO. ANP42 - Brazilian biodiesel standard. 2005.
Automotive_Emissions_Impact_Study_of_Global_Biodiesel_Usage_in_Diesel_ [185] South African National Standard. SANS 1935 automotive biodiesel — fatty acid
Engines/links/59c82c72a6fdccc71923fb1f/A-Techno-Economical-and- methyl esters ( FAME ) for diesel engines — requirements and test methods. 2011.
Automotive-Emissions-I, Accessed date: 10 April 2018. [186] Tabatabaei M, Aghbashlo M, Dehhaghi M, Panahi HKS, Mollahosseini A, Hosseini
[174] Food and Agriculture Organization of the United Nations, FAOSTAT. http://www. M, Soufiyan MM. Reactor technologies for biodiesel production and processing: a
fao.org/faostat/en/#home; 2016, Accessed date: 25 November 2016. review. Prog Energy Combust Sci 2019;74:239–303. https://doi.org/10.1016/J.
[175] Mundi Index. Crude oil (petroleum); dated brent - daily price - commodity prices - PECS.2019.06.001.
price charts, Data, and News. https://www.indexmundi.com/commodities/? [187] Raghu Anuradha. Palm stockpiles in world's No. 2 grower seen at 10-month low -
commodity=crude-oil-brent; 2016, Accessed date: 14 August 2019. bloomberg, Bloomberg. https://www.bloomberg.com/news/articles/2019-06-03/
[176] International Energy Agency. Short-term energy outlook. 2016https://www.eia. palm-oil-stockpiles-in-world-s-no-2-grower-seen-at-10-month-low; 2019, Accessed
gov/forecasts/steo/report/global_oil.cfm, Accessed date: 15 November 2016. date: 16 August 2019.

32

You might also like