You are on page 1of 22

Fuel 235 (2019) 886–907

Contents lists available at ScienceDirect

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

Review article

State of the art review on development of ultrasound-assisted catalytic T


transesterification process for biodiesel production
⁎ ⁎
Shiou Xuan Tana, Steven Limb, , Hwai Chyuan Onga, , Yean Ling Pangb
a
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Department of Chemical Engineering, Lee Kong Chian Faculty of Engineering and Science, Universiti Tunku Abdul Rahman, 43000 Selangor, Malaysia

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

Keywords: Excessive utilization of petroleum diesel has led to severe environmental pollution. Biodiesel, which is greener
Ultrasonic and renewable, can be a potential alternative fuel. Biodiesel is produced through transesterification reaction
Reactive extraction between vegetable oil, animal fat or even waste cooking oil (WCO) and alcohol in the presence of catalyst. Under
Transesterification process intensification, ultrasonic irradiation is employed in the transesterification reaction to enhance the
Non-edible
agitation between immiscible reactants. Besides providing intensive mixing, it also offers uniform heating due to
Mechanism
the localized temperature increase and formation of micro jets from the transient collapse of cavitation bubbles,
thus reducing the energy consumption. The focus of this paper is to review the recent research progress on the
ultrasound-assisted catalytic transesterification of non-edible vegetable oils using homogeneous and hetero-
geneous catalysts. The primary factors that affect the operation and efficiency of ultrasound-assisted transes-
terification such as alcohol to oil molar ratio, catalyst loading, reaction time, reaction temperature, energy
consumption, phase separation time, ultrasonic pulse mode and biodiesel conversion or yield have been re-
viewed. The highlights of this review paper are the provisions on the mechanism of ultrasonic reactive extraction
(RE) in the biodiesel production, kinetic study and the existing pilot reactors on the ultrasound-assisted trans-
esterification which are still rarely reviewed in the current literature. Lastly, the challenges and feasibility for
future development in the process intensification of biodiesel production are also addressed.

1. Introduction combustion [4]. Therefore, it tends to produce less smoke and particles
during combustion, resulting in lower emissions of harmful pollutants
Biodiesel is a liquid fuel which comprises of mono alkyl ester of such as particulates, carbon monoxide (CO), unburned hydrocarbons
long-chain fatty acid. In principle, it is a sustainable source of liquid (HCs) and sulphur oxides (SOx) [5,6]. By employing biodiesel compared
transportation fuels as it is synthesized from various types of renewable to diesel, the release of particulates, CO, unburned HCs and SOx would
lipids such as virgin vegetable oils, non-edible vegetable oils, waste be lowered by 40, 44, 68 and 100%, respectively [7]. In contrast, diesel
vegetable oils (WVOs) and animal fats. Biodiesel can be produced from fuel does not contain any oxygen element [8]. The sulphur content in
a variety of sources which include edible oils (e.g. canola, soybean, the diesel fuel contributes to the formation of SOx and sulphuric acid,
sunflower and palm oils), non-edible oils (e.g. Jatropha curcas, which lead to acid rain. The aromatic compounds in the diesel fuel also
Daturametel Linn, Hevea brasiliensis and Calophyllum inophyllum oils), increase particulate emissions and are considered carcinogens. Ben-
waste oils as well as animal fats (e.g. chicken fat, beef tallow and zene, toluene, ethylbenzene, and o-, m-, p-xylene, popularly known as
poultry fat) [1]. The increasing price volatility of fossil fuel has made BTEX compounds, are well documented carcinogenic compounds
biodiesel appears to have significant economic potential as a renewable emitted from the exhaust of a diesel-powered compression ignition (CI)
fuel by lowering the dependency on crude oil foreign imports. Cheaper engine [9]. Furthermore, biodiesel does not require any modifications
and non-edible feedstocks have been employed to produce biodiesel in when it is utilized in diesel engine directly due to its similar properties
order to alleviate the food crisis arising from the utilization of edible to diesel fuel, thus render it a suitable alternative to diesel fuel [10].
feedstocks [2,3]. Transesterification is the most common process to synthesize bio-
Biodiesel has negligible sulphur and aromatics content with about diesel in the presence of alkali, acid or enzyme as catalyst. However,
11% built-in oxygen content, which helps it to undergo complete there are two major challenges encountered in the transesterification


Corresponding authors.
E-mail addresses: stevenlim@utar.edu.my, stevenlim_my2000@hotmail.com (S. Lim), onghc@um.edu.my (H.C. Ong).

https://doi.org/10.1016/j.fuel.2018.08.021
Received 28 March 2018; Received in revised form 12 July 2018; Accepted 4 August 2018
Available online 28 August 2018
0016-2361/ © 2018 Elsevier Ltd. All rights reserved.
S.X. Tan et al. Fuel 235 (2019) 886–907

Nomenclature

R1–COOH free fatty acid


NaOH sodium hydroxide
R1COONa soap
H2O water diglyceride
CH3OH methanol
R1COOCH3 fatty acid methyl ester
k reaction rate constant
X conversion of triglyceride at time,t
t time
triglyceride CTG0 molar concentration of triglyceride at time zero in mol/L

reaction. Firstly, heterogeneous nature of the reaction system poses capital investment and operating costs tremendously [21]. Among the
challenges in mass transfer limitations. Secondly, higher alcohol to oil emerging transesterification techniques, ultrasonication appears to be a
molar ratio is required because of the reversible nature of transester- promising approach as it can enhance mixing, heat and mass transfer
ification reaction. These result in higher operating cost and energy between different phases in alcoholysis process [22]. Consequently, it
consumption which consequently lead to lower biodiesel production can reduce the operating cost and energy consumption as it has lower
efficiency. There are several other transesterification techniques to requirement on alcohol to oil molar ratio, catalyst amount and reaction
produce biodiesel such as non-catalytic supercritical method, biox time as well as no external heating requirement. In addition, phase
process, microwave, membrane technology, reactive distillation and separation of biodiesel from glycerol is simpler and shorter phase se-
ultrasonication. Non-catalytic supercritical method does not require paration time is needed. Its superiority in transesterification reaction
catalyst, thus it does not face the disadvantages encountered in the possesses a huge possibility to be scaled up from laboratory to industrial
catalytic routes such as the necessity of treating the free fatty acids scale. However, it is noteworthy to consider also the problems asso-
(FFAs) and triglycerides (TGs) in different reaction stages, the in- ciated with the utilization of biodiesel in engine. Biodiesel generally has
hibitory effects of water molecules present in the mixture, catalyst de- 14–15% higher fuel consumption than diesel due to its lower calorific
activation, separation of the catalyst from the product mixture, low value (37.9 MJ/kg) than diesel (42.7 MJ/kg) while NO emissions are
glycerol purity, and generation of waste water. However, it is an energy always 15–20% higher than diesel [23]. An overview of the advantages,
intensive and economic infeasible process because it requires high op- disadvantages and recommendations of various emerging biodiesel
erating temperature (553–673 K) and pressure (10–30 MPa) as well as production approaches is tabulated in Table 1.
high methanol-to-triglycerides molar ratio up to 42:1 [11]. For biox This paper reviews in depth the ultrasound-assisted catalytic
process, it employs co-solvents such as tetrahydrofuran (THF) to solu- transesterification for biodiesel synthesis, effect of different parameters
bilize methanol for faster reaction. THF is commonly used due to its towards biodiesel conversion or yield in the ultrasound-assisted
boiling point (339 K) is similar to methanol (337.7 K) which allows the homogeneous and heterogeneous base-catalyzed transesterification as
removal of excess solvents in a single step once the reaction has com- well as comparison between conventional stirring and ultrasonic cavi-
pleted [12,13]. This process is capable to convert oil feedstock con- tation in biodiesel production. In addition, this paper also provides
taining high FFA content (more than 10%) into biodiesel and can be fundamental understanding of ultrasonic mechanism and their effects
conducted under ambient temperature and pressure. However, eco- on transesterification reaction. Most importantly, kinetic study on the
nomical barrier arises due to their close proximity of boiling point be- ultrasound-assisted transesterification, latest pilot reactors for ultra-
tween methanol and co-solvent which makes the product separation to sound-assisted biodiesel production and mechanism of ultrasonic re-
be very challenging [14]. In addition, complete removal of co-solvent active extraction on solid seed for biodiesel production are also re-
from both the glycerol and biodiesel is compulsory due to its hazardous viewed in details.
and toxicity natures [15,16]. Comparing to conventional transester-
ification, microwave method requires about 23 times lower energy
consumption. This is because microwave energy is transmitted directly 2. Introduction of ultrasound
to the reactant, thus eliminating the preheating step. Furthermore,
microwave process has more effective heat transfer system than con- Ultrasound refers to sound wave above human audibility limit
ventional method. Conventional method transfers heat to the reaction which is usually above 20 kHz [29]. It is categorized based on its fre-
via thermal heat reflux whereas microwave transfers energy in the form quency into high-frequency (2–10 MHz) and low-frequency
of electromagnetic wave [7]. Nonetheless, low penetration depth of (20–100 kHz) ultrasound [30]. Application of high-frequency operation
microwave radiation into the absorbing materials limits the scaled-up is more effective for chemical synthesis and wastewater treatment as
to industrial scale [17]. On the other hand, membrane technology in- chemical effects induced are more intensive. Low-frequency operation
tegrates reaction and separation into single step, thus minimizing se- is useful for enhancing the mass transfer across the immiscible reactants
paration costs and recycling requirements. However, membrane tech- by increasing the interfacial surface area between them. Lower fre-
nology still requires further downstream purification since the fatty quency is suggested to be employed for biodiesel synthesis since
acid methyl ester (FAME)-rich phase will still contain methanol, gly- dominant physical effects are required for intensification of transes-
cerol and water [11,18]. For reactive distillation, it is advantageous for terification reaction [22]. High frequency will not benefit biodiesel
esterification process especially when dealing with high FFA feedstocks. synthesis because of the cavitation bubble collapses are weaker than
Compared to the conventional method, it does not require excess al- impingement of the reactants [14]. Diverse forms of energy can be
cohol due to the continuous removal of by-product (water) and is able generated through ultrasonic applicator, which is shown in Fig. 1. An
to reduce the alcohol usage by 66% [11,19]. Even though reactive ultrasonic probe transforms electrical energy input into heat energy and
distillation process has less number of connections between instruments vibrational energy. The vibrational energy is then transformed into
(due to smaller amount of equipment) which reduces the safety issues cavitation energy while some of them is lost through sound reflection
[20], the requirement for reboiler and condenser will increase the [13]. The cavitation energy is further converted into chemical, physical,
and biological effects depending on the application and reaction

887
Table 1
S.X. Tan et al.

Comparison between different biodiesel production approaches.


Methods Description Advantages Disadvantages Recommendations Ref.

Non-catalytic Biodiesel production in which transesterification • Requires no catalyst. • Requires high operating temperature Addition of co-solvents such as carbon dioxide [2,20,24,25]
supercritical occurs by an alcohol at supercritical conditions • Achieves near complete conversion in a (553–673 K) and pressure (10–30 MPa) (CO2), hexane, propane, calcium oxide (CaO) and
method (high temperature and pressure) without catalyst. relatively short time (2–4 min). resulting in much higher heating and subcritical alcohol to reduce the reaction
cooling costs. temperature, pressure and the amount of
• Requires high methanol-to-triglycerides alcohol.*The supercritical condition of CO2 is
molar ratio up to 42:1. 304 K and 7.38 MPa which it is lower than the
supercritical methanol conditions (512 K and
8.09 MPa).

Biox process Non-catalytic biodiesel production with the use of • Short reaction time. • Completely removal of co-solvent is − [14]
co-solvent (usually THF) that generates one phase • Able to handle feeds with high FFA. compulsory due to its hazardous and
oil-rich system. • Recovery of both the excess alcohol and toxicity natures.
the THF co-solvent in a single step. • Residual solvent in the biofuel product
• Absence of catalyst residues in either the can affect the compliance with the
biodiesel phase or the glycerol phase. international standards.

Microwave Continuous oscillation of polar ends of molecules • Short reaction time (10 times shorter • Not suitable to be used in industrial Combination effect of microwave (removal of heat [7,17,26]
or ions by microwave irradiation causes the than conventional methods). scale as high microwave output (power) transfer barrier and ultrasound (removal of mass
collision and friction between the moving • Low oil to methanol ratio. may cause damage to organic molecules transfer barrier) can lead to synergistic effects
molecules and generates localized superheating. • Ease of operation. (TGs) and due to safety aspects. significantly which overcomes the limitation of
• Higher quality and yield product. low penetration depth. However, this technique
• Drastic reduction in quantity of by- leads to high energy consumption for maintaining
products. higher temperature condition which is not
applicable for industrial scale.

888
Reactive distillation Simultaneous chemical reaction and distillation • Simultaneous separation of reactants and • Thermal degradation of the products Reactive absorption could replace reactive [19,21,27]
separation in a single unit. products boosts the conversion and occurs due to a higher temperature distillation as it reduces capital investment and
improves the selectivity by breaking the profile in the column. operating costs due to the absence of the reboiler
reaction equilibrium restrictions. and condenser, higher conversion and selectivity
• Eliminates the need for reheating as no products are recycled in the form of reflux or
because the heat of vaporization boil-up vapours, as well as no occurrence of
provides the heat of reaction in thermal degradation of the products due to a
exothermic reactions. lower temperature profile in the column.

Membrane Combination of reaction and membrane-based • Controlled contact of incompatible • High fabrication cost. − [6,7,11,19,28]
technology separation by taking advantage of the reactants. • Fouling can occur which reduces the
immiscibility of methanol and oil to create a better • Produces FAME-rich products. effectiveness over time.
purification process. • Eliminates undesired side reactions.
Ultrasonication Cavitation bubbles are produced by sound waves • Simple reaction setup. • Necessity of downstream processing − [7,17,22,26]
in a liquid due to pressure variations. • Shorter reaction time. which increases the purification cost.
• Less energy consumption (only one-third • Reaction temperature is slightly higher
to half of the energy that is consumed by for long reactions.
mechanical agitation). • Ultrasonic power must be under control
• Enhanced conversion. due to the soap formation in fast
• Higher oil extraction yield (30–40% reaction.
more when compared to the • FAME yield can be reduced by higher
conventional stirring reactor system). frequencies (40 kHz).
• Higher quality of glycerol production.
Fuel 235 (2019) 886–907
S.X. Tan et al. Fuel 235 (2019) 886–907

Electrical

Heat Vibrational Cavitational

Physical Chemical Biological Sound

Fig. 1. Energy forms, losses and effects in ultrasonic application.

environment [31]. Generally, the beneficial effects of ultrasound on heat and mass transfer in the medium and promoted the desired che-
chemical reactions come from the intense energy produced from the mical reactions [37–39]. Therefore, employment of ultrasound in the
ultrasonic cavitation. Ultrasonic cavitation is the formation, growth and production of biodiesel eliminates the requirement of external agitation
implosive collapse of bubbles (cavities) in a liquid irradiated with ul- and heating because of the formation of micro jets and increment of
trasound. Expansion and compression waves are present during the localized temperature. For the chemical effect of ultrasound, highly
introduction of ultrasound into the liquid, which are resulted from the reactive species such as OH%, HO2% and H% radicals are produced from
negative and positive pressures, respectively. These fluctuations will the dissociation of solvent vapour trapped in the bubble during a
form bubbles which are filled with solvent, solute vapour and dissolved transient implosive collapse of bubbles, which can help to accelerate
gases. These bubbles will then undergo further growth and re- the chemical reaction [33,36].
compression [30]. The bubbles have very short lifetimes which break
down after an approximate period of 400 µs [14]. When the cavitation 3.1. Ultrasound-assisted homogeneous alkaline-catalyzed
bubble collapses, it will produce extremely high local heating and transesterification
elevated pressures as well as enhancing the mass and heat transfer
within the reaction medium owing to the formation of small eddies Alkaline-catalyzed transesterification is commonly applied in the
[30]. production of biodiesel since alkaline catalysts can catalyze transes-
In heterogeneous liquid reaction systems, ultrasound can render terification process faster than acid catalysts [40]. Alkaline metal alk-
physical and chemical effects simultaneously through cavitation bub- oxides and hydroxides are most frequently used alkaline catalysts for
bles. The former emulsifies immiscible reactants by the micro-turbu- ultrasound-assisted transesterification. At low catalyst concentration
lence generated due to the radial motion of cavitation bubbles, in- (0.5 mol%), the former was capable of achieving significantly high
creasing the interfacial area between them and subsequently the yields (> 98%) in short reaction times (30 min). Therefore, alkaline
reaction rate. The latter enhances the chemical reaction due to gen- metal alkoxides were appraised as the most active catalysts [41]. Po-
eration of free radicals during the sudden collapse of cavitation bubbles tassium hydroxide (KOH) and sodium hydroxide (NaOH) are the most
[30]. The beneficial effects of ultrasound on transesterification are ex- commonly used alkaline metal hydroxides. They are normally in-
plained in more details in Section 3. expensive and able to catalyze at low reaction temperatures (from room
temperature to around 333 K) as well as complete reaction in short
reaction time (30 min) [42,43]. On top of that, high conversion rate is
3. Ultrasound-assisted biodiesel productions achieved without intermediate steps [44]. By employing ultrasound in
alkaline-catalyzed transesterification, cavitation events help to reduce
The physical effect of ultrasound on transesterification reaction had mass transfer resistance between immiscible reactants which then ac-
been discussed by quite a number of researchers. In transesterification celerate the chemical reaction. Ea is also reduced which enables the
reaction, ultrasonic irradiation causes bubble cavitations near the phase reaction to be conducted at lower temperature [45]. In addition,
boundary between the alcohol and oil phases, thus producing large homogeneous alkaline-catalyzed transesterification assisted by ultra-
amount of micro bubbles. Some bubbles are stable for the next cycle sound requires lower alcohol to oil molar ratio, catalyst amount, reac-
while the rest will be subjected to violent collapse upon expanding to tion temperature and energy consumption as well as shorter reaction
certain critical size. The asymmetric collapse of the cavitation bubbles time and phase separation time while attaining higher biodiesel yield as
generates micro-turbulence and causes disruption at the phase compared to those without ultrasound. The detailed beneficial effects of
boundary. Micro jets generated from the liquid impingement can reach ultrasound could be found in Section 4.
to a speed up to 200 m/s, creating intimate mixing of the immiscible Nonetheless, alkaline-catalyzed transesterification can only be em-
reactants near the phase boundary and thus induce emulsification. With ployed if the feedstock has FFA level less than 2 wt%, which corre-
the formation of such emulsion, the interfacial region and mass transfer sponds to acid value of 4 mg KOH/g [46]. Oil containing high FFA
between the alcohol and oil phases increases, thus the reaction kinetics content cannot be transesterified directly using alkaline catalyst since
will become faster [14,29,30,32,33]. According to He and Van Gerpen the FFA reacts quickly with the alkaline catalyst through saponification
[34], the formation and bursting of micro bubbles caused by ultrasonic side reaction which is presented in Eq. (1) to produce soap and water.
cavitation intensified the local energy transfer and energized the re- Excessive soap formation tends to gel-up the reaction mixture as soli-
actant molecules, thus enhanced the overall reaction rate. The ultra- dification of soap tends to occur at ambient temperature [47]. The soap
sound not only provides the mechanical energy for mixing but also the formed binds with the catalyst and partially consumes it and hence a
activation energy (Ea) required for transesterification reaction [35]. higher cost will be incurred [44]. Consequently, viscosity of the reac-
Several researchers believed that high local temperature (≥5000 K) tion mixture increases which then interferes with the separation of ester
and pressure (≥1000 atm) [36] would be generated from continuous from glycerol. This will reduce the biodiesel conversion and yield as
compression and rarefaction waves of ultrasonication which then in- well as increases the product separation cost [10,48].
duced intense mixing between the immiscible reactants. This phe-
R1−COOH + NaOH → R1COONa + H2 O (1)
nomena automatically increased the temperature of the reaction
medium with disruption of the microbubbles which thus increased the Furthermore, high percentage of water content in feedstock oil is

889
S.X. Tan et al. Fuel 235 (2019) 886–907

also undesirable. The alkaline homogeneous catalysts must be handled chlorosulfonic acid also counteracted inhibition caused by water
properly as they are highly hygroscopic and tend to absorb water from formed during esterification, which was the primary cause for very slow
air during storage. They can also form water when dissolved in the kinetics of acid catalyzed transesterification. With the usage of chlor-
alcohol reactant [44]. For instance, water molecules can be produced osulfonic acid, 93% of Jatropha biodiesel yield was achieved at op-
from the reaction between the hydroxide and the alcohol even if water- timum conditions of 8.5 wt% catalyst, 20:1 methanol to oil molar ratio,
free alcohol/oil mixture is used [41]. Water can hydrolyze TGs to di- temperature of 331 K and reaction time of 4 h. The Ea for the process
glycerides (DGs) and forms more FFA, with subsequent soap formation (57 kJ/mol) was at least 3 times lower than H2SO4 catalyzed transes-
which is detrimental to the biodiesel yield. The typical hydrolysis re- terification.
action is shown in Eq. (2) [44]. Apart from chlorosulfonic acid, sulfuryl chloride (SO2Cl2) is also
able to catalyze both esterification and transesterification reactions
concurrently. It had been employed in the preparation of biodiesel from
Daturametel Linn oil seeds containing high FFA content using ultrasonic
cleaning bath (40 kHz, 130 W) [53]. Strong acids such as hydrochloric
(2) acid (HCl) and H2SO4 were formed through the vigorous reaction be-
tween water molecules in Daturametel Linn oil and SO2Cl2 under intense
mixing. The maximum biodiesel yield of 95.5% was attained at 8.0 wt%
3.2. Ultrasound-assisted homogeneous acid-catalyzed transesterification SO2Cl2, 7:1 methanol to oil molar ratio and temperature of 318 K with
reaction time of 2 h.
For oil or fat containing high FFA content (> 2 wt%), it was re-
ported that acid-catalyzed transesterification performed better than 3.3. Ultrasound-assisted heterogeneous base-catalyzed transesterification
base-catalyzed transesterification as acid catalysts could catalyze both
esterification and transesterification concurrently. This was because In heterogeneous catalyzed transesterification, the catalyst will be
acid could act as esterification reagent and solvent at the same time. in solid phase in contrary to methanol and oil which are in immiscible
Unlike base-catalyzed transesterification, there is no soap formation in liquid phase. Hence, the overall system is regarded as a triphasic system
the presence of acid catalyst [7,30]. Acid-catalyzed transesterification (liquid/liquid/solid) [54]. By applying ultrasound in the reaction
has higher tolerance towards moisture content of feedstock, which mixture, catalyst size can be reduced to smaller particles, creating new
could otherwise trigger the hydrolysis of esters resulting in regeneration active sites for subsequent reactions. Therefore, the solid catalyst is
of fatty acids [44]. For oil or fat with FFA less than 2 wt%, the kinetic expected to last longer in the ultrasound-assisted process [55]. The
rate of homogeneous acid-catalyzed transesterification was about 4000 effectiveness of heterogeneous catalytic transesterification depends on
times lower than homogeneous base-catalyzed reaction which rendered the activity of the solid catalyst employed [56]. There are several
it unfavourable in commercial applications [49]. Production of FAME concerns for utilizing heterogeneous catalyst in ultrasound-assisted
through the reaction between FFA and methanol is presented in Eq. (3). biodiesel production such as durability, mechanical stability and de-
R−COOH + CH3 OH ↔ R1COOCH3 + H2 O (3) activation of catalyst. These concerns are crucial because the stability of
applying the catalysts might be affected in the long term operation
Sulphuric acid (H2SO4) is generally used as a catalyst in the under vigorous mixing condition [57]. Catalyst reusability is often used
homogeneous acid-catalyzed transesterification because of its acid as the parameter to indicate the catalyst stability. In addition, it is one
strength which is responsible for releasing more H+ species to proto- of the critical criteria in determining the economic feasibility of bio-
nate the carboxylic moiety of the fatty acid [50]. It was found out that diesel production using heterogeneous catalyst [58].
mechanical stirring is not a feasible approach for homogeneous acid- Various solid base catalysts had been reported for biodiesel synth-
catalyzed transesterification as it requires high reaction temperature esis and several selected studies reporting the maximum biodiesel yield
(343–473 K), high pressure (6–10 bar) and long reaction time (> 48 h). from non-edible feedstocks at optimum conditions have been tabulated
Biphasic nature of the reaction system is the main cause for slow ki- in Table 2. It can be observed from Table 2 that maximum biodiesel
netics of homogeneous acid-catalyzed transesterification which affects yields were in the range between 90.0% and 98.5%. In the transester-
the catalyst accessibility to the triglyceride molecules. Although reac- ification of WCO using calcium diglyceroxide (CaDG), 93.5% yield was
tion kinetics could be improved by increasing alcohol to oil molar ratio obtained in the first cycle. Nonetheless, significant decrement in the
(20:1–45:1) and acid concentration (≈5 wt%), it is not practical for yield was noticed in the second and third cycles in which the yields
large scale process as higher operating cost incurred from the excessive dropped to 59% and 50.5%, respectively. This finding might be at-
usage of alcohol and reactor corrosion due to higher acid concentration tributed to better solubility of CaDG in glycerol–methanol mixture [59].
[51]. Therefore, ultrasound is a practical approach to increase the Basically, there are three primary reasons for basic catalyst deactivation
miscibility of the biphasic nature of the reaction system through the [60]. Firstly, absorption of products and by-products on the catalyst
cavitation events. surface reduces the contact between basic sites and reactants. This
Currently, researches on the homogeneous acid-catalyzed transes- claim can be verified through the study done by Deng et al. [60]. The
terification of non-edible oils under ultrasonic irradiation condition biodiesel yield decreased from 95.2% to 80.7% after 1st and 3rd cycle,
were still limited. So far, only Choudhury et al. had conducted trans- respectively. Absorption of viscous liquid (glycerol) on the catalyst
esterification of Jatropha oil using H2SO4 [52] and chlorosulfonic acid surface was found to be the primary suspect of catalyst deactivation.
[33] with the aid of ultrasonication. In their study, biodiesel yield of Therefore, the catalyst was washed with ethanol for removal of viscous
80.12% was successfully obtained at optimum conditions of methanol liquid before subjected to the next cycle. At the 8th cycle reused, bio-
to oil molar ratio of 7, catalyst concentration of 6 wt%, temperature of diesel yield of 89.1% was recorded. Similarly, the presence of calcium
343 K and reaction time of 2 h. Simultaneous esterification and trans- soap and glycerol phase was suspected to cause a decrement in catalytic
esterification with sonication did not succeed due to formation of water activity of the Karabi biodiesel production by Yadav et al. [61]. The
in the esterification step, and sonication alone could not overcome this catalyst maintained its catalytic activity without any considerable
obstacle [52]. In order to overcome the inability of H2SO4 to carry out change up to four reaction cycles (94.1% yield). In the 5th cycle, the
esterification and transesterification simultaneously, Choudhury et al. biodiesel yield decreased to 65.5%. In the transesterification of Jatropha
[33] had utilized chlorosulfonic acid as catalyst instead. Chlorosulfonic oil catalyzed by tripotassium phosphate (K3PO4), its service life only
acid possessed superior ability as it was capable to catalyze both es- could maintain up to three cycles due to the methanol deposition on the
terification and transesterification reactions. Furthermore, catalyst [62]. Secondly, leaching of the active sites into solutions

890
S.X. Tan et al. Fuel 235 (2019) 886–907

contributes to the catalyst deactivation. The biodiesel yields in the

[37]

[60]

[63]

[58]
[59]
[61]
[62]

[64]
Ref.
transesterification of Silybum marianum oil for 1st and 5th cycle were
90.1% and 80.0%, respectively [63]. Leaching of potassium compounds
might be the reason for the drop in the yield. The last reason of catalyst
Catalyst reusability

deactivation can be attributed to the disappearance of active sites due


to the collapse of catalyst structure. This claim was verified through the
X-ray powder diffraction (XRD) characterization conducted by Deng
(times)

et al. [60]. On the other hand, reduction in the biodiesel yield by using
K3PO4 was explained by two possible causes: 1) agglomeration of the


3

4
1
4
3
catalyst in the subsequent cycle of transesterification reaction caused a
Yield (%)

decrement in the number of active sites; 2) Side reaction of K3PO4 with


98.53

95.2

90.1

92.0
93.5
94.1
98.0

95.4
methanol resulted in the formation of dipotassium phosphate (K2HPO4)
and monopotassium phosphate (KH2PO4) during the transesterification
reaction [58].
Reaction time

3.4. Ultrasound-assisted heterogeneous acid-catalyzed transesterification


(min)

120

Ideal solid acid catalyst for transesterification reaction should ex-


15

90

30

90
30

45

30

hibit characteristics such as an interconnected system of large pores, a


Alcohol type; alcohol to oil molar

hydrophobic surface and high concentration of strong acid sites.


However, research on direct usage of solid acid catalyst for transes-
terification reaction has not been widely explored because of its lim-
itations with slow reaction rate and undesirable side reactions.
Moreover, there is still a knowledge gap on the fundamental studies
Methanol; 16:1

12:1
12:1

Methanol; 10:1

dealing with reaction pathways of TGs on solid acid catalysts [56].


Methanol; 9:1

Methanol; 4:1

6:1
9:1

Reports on the role of heterogeneous acid catalyst in ultrasound-as-


Methanol;
Methanol;
Methanol;
Methanol;

sisted process are still relatively rare [65]. The biodiesel yields at op-
ratio

timum conditions by using different solid acid catalysts are tabulated in


Table 3. The highest biodiesel yield (98%) was achieved by using
Temperature (K)

composite of phosphotungstic acid encapsulated in a Fe(III)-based


metal organic framework (PTA-MOF) while the lowest yield (84%) was
achieved by using tungstophosphoric acid/gamma alumina (TPA/Al). It
318

333

323
333
333
323

338

is also noticeable from Table 3 that slightly higher biodiesel yield was

attained by using n-butanol (98%) than ethanol (96%) in the ester-


Catalyst type; loading (wt

ification of oleic acid catalyzed by PTA-MOF. This observation could be


due to higher solubility of n-butanol than ethanol in oleic acid.
KHC4H4O6/TiO2; 5.0

In order to find out the influence of ultrasound on the catalyst


stability (in terms of catalyst leaching), three studies were conducted by
SrO-CaO; 6.0
Na/SiO2; 3.0

Mg/Al; 1.0

K3PO4; 3.0

K3PO4; 1.0

Badday et al. on tungstophosphoric acid/activated carbon (TPA/AC),


CaDG; 1.0
CaO; 5.0

TPA/Al and tungstophosphoric acid/cesium (TPA/Cs) [65–67]. The


results revealed that only 11%, 9.6% and 9.2% of the FAME yield were
%)

achieved under the optimum conditions due to the minimal leaching of


3−
anions, for TPA/AC, TPA/Al and TPA/Cs, respectively. Thus,
Biodiesel yield at optimum conditions by using different solid base catalysts.

PW12 O40
Silybum marianum

all three catalysts had demonstrated good catalyst stability with low
Jatropha oil

Jatropha oil

Jatropha oil

Jatropha oil

leaching behaviour. For sulfonated solid carbon derived from cyclo-


Feedstock

Karabi oil

dextrin (SO3H-CD) without regeneration, the FAME yield decreased in


WCO
WCO

each cycle, from 90.8% (fresh) to approximately 81% (4th cycle). This
oil

observation could be attributed to the deactivation of the active sites


Power (W)

due to their poisoning by other impurities present in the reaction


mixture such as ion-exchange between alkali cations with protons and
200

210

250

375
120

600

210
50

natural adsorption of reactants and products in the system [68]. After


the regeneration process of SO3H-CD with repeating sulfonation, de-
Frequency (kHz)

activation problem was resolved and the catalyst could be reused for 4
cycles without any serious reduction in the FAME yield. For PTA-MOF,
it could be reused for 5 cycles and approximately 5 wt% of PTA was
20–30

leached during the recovery process in both esterification processes


24

40

22
22

20

using ethanol and n-butanol. Electrostatic interaction between PTA


Direct sonication (horn)
Direct sonication (horn)
Direct sonication (horn)
Ultrasonic reactor type

anions and Fe(III) network that was rich in amino groups played a
Indirect sonication

Indirect sonication

Indirect sonication

Indirect sonication

Indirect sonication

crucial part in minimizing leaching from the composite [69].

3.5. Two-step homogeneous and heterogeneous ultrasound-assisted


transesterification
(bath)

(bath)

(bath)

(bath)

(bath)

High initial acid content in non-edible oil will hamper transester-


Table 2

2010

2011

2014

2014
2015
2016
2016

2017
Year

ification approach using alkaline catalysts due to the soap formation.


On the other hand, acid catalyst can catalyze non-edible feedstock

891
S.X. Tan et al.

Table 3
Biodiesel yields at optimum conditions by using different solid acid catalysts.
Year Ultrasonic reactor type Frequency (kHz) Power (W) Amplitude (%) Feedstock Catalyst type; loading (wt%) Temperature (K) Alcohol type; alcohol to oil molar ratio Reaction time (min) Yield (%) Ref.

2013 Direct sonication (probe) 20 400 60 Jatropha oil TPA/AC; 4.23 338 Methanol; 25:1 40 91 [66]
2013 Direct sonication (probe) 20 400 60 Jatropha oil TPA/Al; 4.44 338 Methanol; 19:1 50 84 [65]
2013 Direct sonication (probe) 20 400 60 Jatropha oil TPA/Cs; 2.9 338 Methanol; 25:1 34 90.5 [67]
2015 Direct sonication (probe) 25 − − WCO SO3H-CD; 11.5 390 Methanol; 20:1 8.8 90.8 [68]
2017 Indirect sonication (bath) 37 100 − Oleic acid PTA-MOF; 30.0 Ambient Ethanol; 16:1 15 96 [69]
2017 Indirect sonication (bath) 37 100 − Oleic acid PTA-MOF; 30.0 Ambient n-butanol; 16:1 15 98 [69]

892
Table 4
Optimum conditions for two-step homogeneous and heterogeneous ultrasound-assisted transesterification.
Year Ultrasonic reactor type Frequency (kHz); Power (W) Feedstock Reaction conditions Biodiesel conversion (%) Yield (%) Ref.

2014 Indirect sonication (bath) 35; 35 Jatropha oil 1st step: Methanol to oil (15:1), 5 wt% H2SO4, 338 ± 2 K, 60 min (mechanical stirring) − > 95 [54]
2nd step: Methanol to oil (11:1), 5.5 wt% CaO, 337 K, 120 min (ultrasonication)
2014 Direct sonication (probe) 30; 36 Karanja oil 1st step: Methanol to oil (9:1), 0.5 wt% H2SO4; 303 K, 90 min 84 − [73]
2nd step: Methanol to oil (9:1), 5 wt% Ba(OH)2·8H2O, 303 K, 60 min
2015* Direct sonication (horn) 20; 200 Neem oil 1st step: methanol to oil (0.45 vol%), 0.75 vol% H2SO4, 333 K, 10 min − 98.01 [74]
2nd step: methanol to oil (0.2 vol%), 1.2% KOH, 333 K, 10 min
2015* Direct sonication (horn) 20; 400 Neem oil 1st step: methanol to oil (6:1), 3 wt% H2SO4, 303 K, 30 min 92.72 − [75]
2nd step: methanol to oil (7:1), 1.2% KOH, 323 K, 30 min
*
2015 Direct sonication (horn) 20–30; 1000 Mahua oil 1st step: methanol to oil (0.35 mol%), 1 vol% H2SO4, 323 K, 5 min − 97.4 [71]
2nd step: methanol to oil (1.5 mol%), 0.75% KOH, 323 K, 5 min
2016* Direct sonication (horn) 28; 50 Nerium oleander oil 1st step: methanol to oil (0.4 vol%), 1 vol% H2SO4, 328 K, 10–15 min − 97 [76]
2nd step: methanol to oil (0.2 vol%), 1 v/w% KOH, 328 K, 10–15 min
2016 Direct sonication (probe) 20; 250 Schleichera triguga oil 1st step: Methanol to oil (4:1), 1 vol% H2SO4, 313 K, 20 min − 96.8 [77]
2nd step: Methanol to oil (9:1), 3 wt% Ba(OH)2, 323 K, 80 min
2017 Direct sonication (probe) 30; − Rubber seed oil 1st step: Methanol to oil (15:1), 1.5 wt% H2SO4; 303 K, 120 min − 97 [78]
2nd step: Methanol to oil (8.09:1), 5.38 wt% Ba(OH)2·8H2O, 303 K, 13.2 min

*
Note: denotes for two-step homogeneous ultrasound-assisted transesterification.
Fuel 235 (2019) 886–907
S.X. Tan et al. Fuel 235 (2019) 886–907

containing high acid value but with lower reaction rate. Therefore, a 3.6. Ultrasound-assisted interesterification
two-step process (esterification followed by transesterification) is an
efficient approach to address the undesirable saponification phenom- Interesterification is different from transesterification in term of
enon and slow reaction rate while improving the biodiesel conversion acryl acceptor used and the by-product produced. In biodiesel synthesis
or yield. The feedstock was usually first treated with acid catalysts to via interesterification process catalyzed by base catalyst, alkyl acetate
reduce FFA content to below 2 wt%, followed by transesterification will replace alcohol as acryl acceptor which then produces a short chain
with alkaline catalysts. With the assistance of ultrasound in biodiesel triglyceride instead of glycerol. For instance, triacetin is yielded as by-
production, the processing cost can be greatly reduced due to lower product instead of glycerol when methyl acetate is used during inter-
amount of methanol and operating temperature needed, compared to esterification [79]. This complex reaction is composed of three con-
the conventional two-stage approach [70–72]. Biodiesel conversions or secutive reversible reactions, which are shown in Fig. 2 [25]. In inter-
yields from different feedstocks at optimum conditions by employing esterification reaction catalyzed by acid catalyst, FFA in vegetable oil
two-step homogeneous ultrasound-assisted transesterification are reacts with methyl acetate to form biodiesel and acetic acid [80].
summarized in Table 4. High biodiesel conversion or yield Generally, there are several drawbacks associated with the trans-
(92.72%−98.01%) can be achieved as observed from Table 4. esterification process. For enzyme catalyst, excess utilization of me-
thanol will lead to deactivation of enzyme and the by-product (glycerol)
will have inhibitory effects on the enzymatic activity. These

Fig. 2. Interesterification reaction (Initial 3 equations show the initial steps whereas final equation gives the overall reaction). Adapted with permission from
(Maddikeri GL, Pandit AB, Gogate PR) [25]. Ultrasound assisted interesterification of waste cooking oil and methyl acetate for biodiesel and triacetin production.
Fuel Process Technol 2013;116:241–9. Copyright (2013) Elsevier Ltd.

893
S.X. Tan et al. Fuel 235 (2019) 886–907

unfavourable issues will restrict the application of enzymatic route at the intensification of the transesterification or esterification processes
industrial scale [81]. With the co-production of glycerol, neutralization [84]. This claim had been verified by several researchers through their
is required to prevent biodiesel contamination. Separation and pur- studies such as Kumar et al. (6:1 versus 8:1) [46], Yin et al. (5:1 versus
ification steps are needed for recovery of glycerol. All these steps are 6:1) [49], Mohod et al. (5:1 versus 6:1) [85] and Mohan et al. (4.5:1
energy and resource intensive, decreasing the production efficiency and versus 6:1) [84] in which the values in bracket represent the alcohol to
increasing the production costs. In addition, glycerol is undervalued in oil molar ratios required by ultrasonication and mechanical stirring
the world market due to its excessive supply derived from biodiesel approaches, respectively. Based on their findings, it was revealed that
production [80,82]. With the employment of interesterification instead ultrasonication could achieve 16.67–25% reduction in alcohol to oil
of transesterification, enzyme deactivation could be avoided as triacetin molar ratio as compared to mechanical stirring.
has no adverse effect on the activity of lipase enzyme [81]. Further- Methanol is the most preferable alcohol in transesterification due to
more, production of triacetin is more preferable as it has higher com- its lower cost, industrial availability and superior reactivity as com-
mercial value than glycerol. Triacetin can be used as plasticizer or ge- pared to other alcohols. Its short molecular chain and polar properties
latinizing agent in polymers and explosives and as an additive in serve as its physical and chemical advantages. It can also react with TGs
tobacco, pharmaceutical industries, and cosmetics. In addition, tria- efficiently and dissolve with the basic catalyst easily [2]. Molar ratio of
cetin is soluble in biodiesel and could be added into biodiesel (up to methanol to oil is one of the essential factors affecting the rate of for-
10% by weight) as an additive which could help to improve oxidative mation and yield of FAME. Theoretically, molar ratio of methanol to oil
stability and cold flow properties of biodiesel [25,79]. for the transesterification is 3:1. However, large excess amount of me-
By focusing on the non-edible feedstock, there are only several thanol is required to drive the reversible transesterification reaction
studies conducted on the ultrasound-assisted interesterification. The towards FAME [5]. Excess alcohol in transesterification can enhance
biodiesel yields obtained at optimum reaction conditions using different alcoholysis and ensure complete conversion of oils or fats into esters.
catalysts are shown in Table 5. The superiority of ultrasound as com- Furthermore, better ester conversion can be achieved in a shorter time
pared to conventional approach in interesterification had been proven with a higher alcohol to triglyceride ratio [44]. For heterogeneous
by researchers. For example, Maddikeri et al. [25] observed that higher catalyst, excess alcohol can promote transesterification rate in addition
biodiesel yield was achieved by ultrasound (90%) than conventional to remove product molecules from the catalyst surface and regenerate
stirring (70%) under the same optimum reaction conditions. In the the active sites [86]. The effects of molar ratio of methanol to oil on
study conducted by Subhedar and Gogate [81], they discovered that ultrasound-assisted alkaline transesterification using homogeneous and
ultrasound-assisted interesterification required lower methyl acetate to heterogeneous catalysts are tabulated in Table 6. It is noticed from
oil ratio (9:1 versus 12:1), lower enzyme loading (3% versus 6%), Table 6 that a molar ratio of more than 6:1 was mostly required to
shorter reaction time (3 h versus 24 h) and achieved higher biodiesel achieve optimum biodiesel yield.
yield (96.1% versus 90.1%) as compared to conventional stirring. Based Further increasing the molar ratio beyond the optimum will de-
on the several existing literatures, it is suggested that more studies crease the yield which might be caused by over dilution of biodiesel and
should be conducted on the ultrasound-assisted interesterification with glycerol [38,62,87]. This is because higher dissolution of glycerol and
the utilization of different catalyst types to produce the optimum bio- alcohol in biodiesel will affect the purity of biodiesel if the amount of
diesel yield (chemical and enzyme catalysts). methanol increases in the mixture [87]. Moreover, the equilibrium re-
action might shift towards the reactant side, which reduces the effective
4. Primary factors for ultrasound-assisted transesterification yield [58]. In addition, formation of undesirable suspension resulting
from the unreacted methanol and oil mixture has increased the diffi-
There are several factors that will influence the biodiesel conversion culty of separation and longer duration is required as compared to
or yield in the ultrasound-assisted transesterification such as alcohol to lower methanol ratio [38].
oil molar ratio, catalyst loading, reaction time, reaction temperature, The optimum methanol to oil molar ratio is also affected by the
energy consumption and ultrasonic pulse mode. Although mechanical phase of catalyst. Generally, heterogeneous catalysts require slightly
or magnetic stirring are commonly employed in the production of higher molar ratio than homogeneous catalysts due to slower rate
biodiesel, ultrasonic cavitation appears to be more resource-effective which stemmed from mass transfer limitation [88]. By comparing the
than conventional stirring approach in terms of alcohol to oil ratio, optimum molar ratio required by both homogeneous and hetero-
catalyst loading, reaction time, reaction temperature, energy con- geneous catalysts in Table 6, 16:1 is the highest molar ratio required by
sumption, phase separation time and biodiesel conversion or yield. heterogeneous catalysts whereas 12:1 is the highest molar ratio re-
quired by homogeneous catalysts. This observation tallied with the
4.1. Alcohol to oil molar ratio statement that heterogeneous catalysts required slightly higher molar
ratio than homogeneous catalysts. From Table 6, it is also observed that
A lower alcohol to oil molar ratio will definitely minimize the en- transesterification of WCO had been conducted by several groups of
ergy consumption because alcohol separation using distillation is an researchers by employing homogeneous (KOH and NaOH) and hetero-
energy-intensive operation, which controls the overall economics of geneous (K3PO4 and CaDG) catalysts. The general optimum molar ratios
biodiesel synthesis process [70]. By incorporating ultrasound in the required by each type of catalyst were as followed: KOH and K3PO4
biodiesel synthesis, lower molar ratio will be required as compared to (6:1); NaOH and CaDG (9:1). Similar optimum molar ratio was reported
conventional stirring. This is due to the intense micro-turbulence and using KOH and K3PO4 to achieve maximum biodiesel yield or conver-
microstreaming generated during the cavitation which contributes to sion. This could be attributed to similar catalytic activity of K3PO4 and

Table 5
Optimum reaction conditions for ultrasound-assisted interesterification of non-edible oil.
Feedstock Catalyst Temperature (K) Methyl acetate to oil Catalyst concentration Ultrasonic power Reaction time Biodiesel yield Ref.
molar ratio (wt%) (W) (h) (%)

WCO Potassium methoxide 313 12:1 1 450 0.5 90.0 [25]


WCO Thermomyces lanuginosus 323 9:1 3% (w/v) 80 3 96.1 [81]
(Lipozyme TLIM)
Crambe oil Novozym®435 333 12:1 20 − 6 98.25 [83]

894
S.X. Tan et al. Fuel 235 (2019) 886–907

KOH [62].

[89]
[63]
[58]
[59]

[31]

[31]

[90]
[90]
[77]

[61]
[62]
Ref.

[4]
Yield (%)

4.2. Catalyst loading

88.4
92.0
93.5

96.8

94.1
81

98

98




The reaction of oils and alcohols is a slow process due to the im-
miscibility between oils and alcohols [7]. Hence, catalyst is required for
Biodiesel conversion

transesterification to accelerate the reaction and gives a better con-


version efficiency of oil to biodiesel [91]. Ultrasonication had been
proven by several researchers [46,85,92] that it could help in de-
creasing the catalyst loading required in either ultrasound-assisted
∼80

96.8
(%)

homogeneous or heterogeneous transesterification as compared to

99
99




conventional stirring. In the study conducted by Kumar et al. [46],
mechanical stirring required higher KOH concentration which was 1 wt
(min)
Time

% than ultrasonication to achieve the same conversion of 98.2% in

120
25
90
30

45

80

60
8

1
1
which 0.75 wt% was only needed by ultrasonication. Similar finding
was also reported by Chen et al. [92] in the transesterification of palm
Optimum molar

oil catalyzed by CaO. Lower catalyst amount (8 wt%) was able to


achieve maximum biodiesel yield (92.7%) in the case of ultrasonic
cavitation than magnetic stirring (55% biodiesel yield with 10 wt%
ratio

16:1

12:1

12:1
12:1
6:1

6:1
9:1

9:1

9:1

5:1
7:1
9:1

CaO). Based on their results, it could be deduced that ultrasonication


could achieve 20–25% reduction in catalyst loading. In the esterifica-
Temperature (K)

tion of Karanja oil conducted by Mohod et al. [85], same catalyst


loading (2 wt% H2SO4) was required by ultrasound and mechanical
stirring approaches to reduce the acid value of Karanja oil from
14.15 mg KOH/g oil. Nonetheless, the final acid value achieved by ul-
333
323
333

318

323
323
323

333
323

trasound approach was 50% lower than that of mechanical stirring


Catalyst type; loading

method. Final acid values obtained by ultrasound and mechanical


KHC4H4O6/TiO2; 5.0
Effects of methanol to oil molar ratio on ultrasound-assisted homogeneous and heterogeneous base-catalyzed transesterification.

stirring approaches were 2.7 and 4.1 mg KOH/g oil, respectively.


In homogeneous transesterification, the beneficial effect was due to
Ba(OH)2; 3.0
NaOH; 1.25

NaOH; 1.25
K3PO4; 3.0

K3PO4; 1.0
NaOH; 1.0

NaOH; 1.0
CaDG; 1.0

higher chemical activity in the presence of ultrasound cavitations [46].


KOH; 1.0

CaO; 5.0
KOH; 1

In heterogeneous transesterification, the grinding effect of ultrasound


(wt%)

on the catalyst could help to expose new catalytic active sites and in-
crease the interfacial area between catalyst and reactant [37,77].
Karanj and soybean oil (40:

Therefore, emulsion droplets of alcohol and oil produced could contact


with the catalyst particles easily, thus achieving effective mixing be-
Silybum marianum oil

Schleichera triguga oil

tween liquid reactants and solid catalyst [93]. In addition, micro-


streaming effect of ultrasound could also help to maintain the catalytic
Jatropha oil
60) (w/w)

activity by dislodging the inactive surface formed and attached on the


Karabi oil
Feedstock

denotes for ultrasound-assisted homogeneous alkaline-catalyzed transesterification.

catalyst, thus providing more active sites for the catalytic reaction [94].
WCO

WCO
WCO

WCO

WCO

WCO
WCO

Inadequate catalyst loading can lead to incomplete conversion of TG


into FAME. As the catalyst loading increases, the conversion of TG and
Power (W)

the yield of biodiesel will also increase until an optimum catalyst


loading is attained. Further increment in the catalyst loading will cause
240
250
375
120

150

150

500

500
500
250

600
50

a decrement in the yield and increase the production cost. With ex-
cessive addition of alkaline catalyst beyond the optimum loading, al-
kaline catalyst can cause more TGs to react with the catalyst through
Frequency

saponification [95]. During washing, the soap present in the ester phase
20–30
(kHz)

has the tendency to accumulate at the interfacial region between the


20
40
22
22

20

20

20

20
20
20

20

liquids. Due to its hydrophilic and hydrophobic structure, the soap


molecules are oriented perpendicular to the interfacial region between
Direct sonication (horn) (Pulse

the two immiscible phases: water and FAME. The soap molecules inside
Indirect sonication (bath)

Indirect sonication (bath)


Direct sonication (probe)
Direct sonication (horn)

Direct sonication (horn)


Direct sonication (horn)

Direct sonication (horn)

Direct sonication (horn)

Direct sonication (horn)


Direct sonication (horn)

Direct sonication (horn)

FAME and water molecules will form emulsion, thus hindering the
Ultrasonic reactor type

purification of FAME [96]. In addition, large amount of catalyst in-


(Continuous mode)

creases the solubility of FAME in the glycerol, causing entrainment of


FAME in the glycerol phase after phase separation [97]. Decrement in
the FAME yield with excessive catalyst loading was encountered by
mode)

Hingu et al. [98] and Kesgin et al. [99] in their studies.


In heterogeneous transesterification, the number of catalyst parti-
cles and active sites increase with the catalyst loading and thus in-
4.5:1, 6:1, 9:1,

4.5:1, 6:1, 9:1,


Various molar

6:1, 9:1, 12:1,

creasing the accessibility of TG and alcohol to the catalyst surface.


6:1, 9:1, 12:1

6:1, 9:1, 12:1


3:1, 6:1, 9:1,
4:1, 6:1, 8:1
8:1 to 18:1

9:1 to 17:1

Further increment of catalyst loading beyond the optimum value will


4:1 to 7:1

4:1 to 7:1
4:1 to 7:1

cause either a decrement or negligible effect on the biodiesel conversion


13.5:1

13.5:1
ratio

14:1

12:1

or yield due to the reduction of interspatial space between TG and al-


cohol molecules [77]. Moreover, increasing catalyst beyond the op-
Table 6

2014*

2015*

2015*

2016*
2016*

*
2014
2015
2015

2016

2016

2016
2016
Year

timum loading will cause formation of emulsion layer which makes the
Note:

biodiesel separation to be difficult [62]. In Jatropha biodiesel synthesis

895
S.X. Tan et al. Fuel 235 (2019) 886–907

assisted by ultrasonic irradiation, excessive magnesia-alumina (Mg/Al)

[89]
[63]
[31]

[31]

[58]
[59]

[90]
[90]
[77]
[61]
Ref.

[4]
(> 1.0 wt%) as catalyst was found to cause a decrease in biodiesel yield

Yield (%)
in addition to the formation of emulsion layer [60]. It should be noted
that inadequate catalyst loading will give rise to unsatisfactory bio-

88.4
96.8

92.0
93.5

94.1
81

98


diesel yield and incur higher production costs. Therefore, optimum



amount of catalyst is required to effectively convert the TG into FAME.
It is noticed from Table 7 that the range of catalyst loading to achieve

conversion (%)
maximum yield or conversion was between 0.5 and 1.5 wt% for ultra-

Biodiesel
sound-assisted homogeneous alkaline-catalyzed transesterification
while higher range of catalyst loading (1.0 to 5.0 wt%) was generally

∼85

96.8
100
99



required for ultrasound-assisted heterogeneous base-catalyzed transes-


terification.

(min)
Time

120

100
120
25

40
45
4.3. Reaction time

1
1
Alcohol type; alcohol to oil
It had been documented that shorter reaction time was required by
ultrasonication compared to the conventional stirring. Many re-
searchers had compared the reaction time needed by both approaches

Methanol;16:1

Methanol;12:1
Methanol; 6:1

Methanol; 9:1

Methanol; 9:1

Methanol; 6:1

Methanol; 6:1

Methanol; 6:1
Methanol; 6:1
Methanol; 9:1
Methanol;9:1
in their studies which are shown in Table 8. Ultrasonication can reduce

molar ratio
the reaction time by at least 25% and as high as 95.8% compared to the
conventional stirring. The reactions under ultrasonication are more
reactive compared to mechanical stirring process. This could be at-

Temperature (K)
tributed to the larger interfacial area of reactant droplets due to the
physical effects of cavitation phenomena in terms of intense turbulence
and mixing generated in the reactor which then led to the enhancement
in the biodiesel yield or conversion [8,100,101]. The average interfacial

333

333
333
318

323
323
323
333
area generated by ultrasonication was found to be 67 times higher than


that generated by mechanical agitation. The droplet sizes for ultrasonic
Catalyst type; optimum

KHC4H4O6/TiO2; 5.0
emulsification were in the narrow range of 0.82–44.6 µm. For emulsi-
fication by mechanical agitation, a much wider droplet size distribution
loading (wt%)
Effects of catalyst loading on ultrasound-assisted homogeneous and heterogeneous base-catalyzed transesterification.

was obtained, ranging from 8.1 to 610 µm [102].

Ba(OH)2; 3.0
NaOH; 1.25

NaOH; 0.75
K3PO4; 3.0
NaOH; 0.5

NaOH; 1.5

KOH; 1.25
CaDG; 1.0
Theoretically, the conversion of fatty acid esters will increase with
KOH; 1.0

CaO; 5.0
reaction time [103]. The reaction is slow initially due to the mixing and
dispersion of alcohol into the oil. After some time, the reaction will
proceed rapidly. However, prolonged reaction time will cause a de-
Karanj and soybean oil
Silybum marianum oil

Schleichera triguga oil


crement in the product yield due to the backward reaction of transes-
terification, causing more FFAs to form soaps [44,104]. The transes- (40: 60) (w/w)
terification reaction was found to be dependent on the reaction time in

denotes for ultrasound-assisted homogeneous alkaline-catalyzed transesterification.


the ultrasonic transesterification of Silybum marianum oil catalyzed by
Karabi oil
Feedstock

KHC4H4O6/TiO2 [63]. Low reaction rate in the first 10 min could be due
WCO

WCO

WCO

WCO
WCO

WCO
WCO

to inadequate agitation to enhance proper mixing and dispersion of


methanol and catalyst onto the oil. The maximum biodiesel yield
Power (W)

(90.1%) was attained in 30 min. However, excess reaction time (after


30 min) resulted in slight reduction in biodiesel yield due to reversible
240
250
150

150

375
120
500

500
500
250
50

reaction.
The influence of reaction time (from 0.5 to 2.5 min, with increments
of 0.5 min) on the transesterification of WCO was investigated by
Frequency

Martinez-Guerra and Gude [31] for both pulse and continuous sonica-
20–30
(kHz)

tion conditions. The highest yield of 98% was achieved by pulse soni-
20
40
20

20

22
22
20

20
20
20

cation at 2.5 min and 91% followed by continuous sonication at 1 min.


During pulse sonication, the biodiesel yield increased with reaction
Indirect sonication (bath)

(probe)
Direct sonication (horn)

Direct sonication (horn)

Direct sonication (horn)

Direct sonication (horn)


Direct sonication (horn)
Direct sonication (horn)

(horn)
(horn)

(horn)
Ultrasonic reactor type

time. On the other hand, the biodiesel yield began to drop upon
(Continuous mode)

reaching 1.5 min reaction time for continuous sonication. The drop in
sonication
sonication
sonication
sonication

the biodiesel yield with increasing reaction time by the usage of con-
(Pulse mode)

tinuous sonication had validated the recommendation given by Chand


et al. [105]. Continuous ultrasound irradiation was not recommended
Direct
Direct
Direct
Direct

for long reaction time due to local temperature increase (hot spots)
which could degrade the oil and cause undesired reactions, leading to a
decrement in the biodiesel yield [106].
0.5, 0.75, 1, 1.25

0.5, 0.75, 1, 1.25


Various catalyst
loading (wt%)

0.5, 1.25, 2

0.5, 1.25, 2

0.5, 0.75, 1
0.5 to 1.25
0.5, 1, 1.5

4.4. Reaction temperature


1.0 to 9.0

1.0 to 4.0

1.0 to 4.0
3.0 to 5.0

For reactions with conventional stirring methods, external heating is


paramount and reaction rate will heavily depend on the reaction tem-
Table 7

2014*

2015*

2015*

2016*

2016*
2016*

*
2014

2015
2015

2016
2016
Year

perature with vigorous mechanical mixing [38]. On the other hand,


Note:

ultrasonic mixing eliminates the need of external heating. The

896
S.X. Tan et al. Fuel 235 (2019) 886–907

Table 8
Reaction time needed by employing ultrasonic cavitation and conventional stirring.
Year Feedstock Reaction conditions Ultrasonic cavitation Conventional stirring Reduction Ref.
percentage (%)
Reaction Yield (%) Conversion (%) Reaction Yield (%) Conversion (%)
time (min) time (min)

2010 WCO • Methanol to oil (9:1) 5 − 94.6 120 89.3 95.8 [100]
• 1Temperature:
wt% NaOH (mechanical)
• Methanol to oil343(9:1)K
a

2014 Palm oil • 1 wt% KOH 15 98.5 − 120 89.5 87.5 [8]
• Temperature: 333 K (mechanical)
• Methanol to oil (3:1)
Palm oil • 1 wt% KOH 15 93 − 150 75 90
• Temperature: 333 K (mechanical)
• Methanol to oil ratio (9:1)
2014 Palm oil • 8 wt% CaO 40 70 − 100 70 (magnetic) − 60 [92]
• Temperature: 333 K
• Methanol to oil (8:1)
2014 Silybum • 1 wt% KOH 15 82.46 − 50 82.35 − 70 [107]
marianum oil • Temperature: 333 K (mechanical)
• Methanol to oil (6:1)
2015 Waste fish oil • 1 wt% KOH 30 79.6 − 60 78 − 50 [108]
• Temperature: 328 K (mechanical)
• Methanol to oil (7:1)
2015 Cynara • 1 wt% NaOH 20 97 − 60 95.8 − 40 [109]
cardunculus L. • Temperature: 333 K (mechanical)
seed oil • Methanol to oil (12:1)
2016 Jatropha oil • 1 wt% K PO 45 98 − 60 92* − 25 [62]
• Temperature: 323 K
• Ethanol to oil ratio (6:1)
3 4
b

2016 Citrullus vulgaris • 1.5 wt% crystalline 120 87 − 480 78* − 75 [110]
seed oil • manganese carbonate
Citrullus vulgaris
seed oil
• Isopropanol
(6:1)
to oil ratio 180 86 − 720 74* − 75

• 1.5 wt% crystalline


manganese carbonate
2016 Nerium oleander • 1st step: Methanol to oil
molar ratio (0.4 vol%),
1st step:
10–15
97 − 1st step: 60
2nd step: 60
92 (magnetic) − 75–83 [76]

1 vol% H2SO4, 328 K 2nd step:


• 2nd step: Methanol to oil
molar ratio (0.2 vol%),
10–15

1v/w% KOH, 328 K


2017 Jatropha oil • Methanol to oil (10:1) 30 95.4 − 180 89.3 ≈ 83 [64]
• 6Temperature:
wt% SrO-CaO (magnetic)
• 338 K

Note: * denotes for stirring type is not mentioned in the study; a


denotes for temperature used in conventional stirring only; b
denotes for temperature used in
ultrasonication only.

capability of ultrasound to increase the temperature inherently can be in oil increases and subsequently promotes biodiesel formation [36]. In
proven through the experiments performed by Boffito et al. [111]. In addition, the kinetic rate constant of this endothermic chemical reac-
the ultrasound experiments at 336 K, the thermostat was set at 313 K tion also increases since there is higher energy input for the reaction to
and the ultrasound provided the heat to achieve the desired bath take place [2,36]. Thus, shorter reaction time can also be achieved at
temperature. In fact, the bath temperature reached 336 K within sec- higher temperature. However, it is important to avoid vaporization of
onds of initiating sonication which was a lot faster compared to con- alcohol which will reduce the contact area with the reactant [2]. In
ventional heating. The efficiency of sonication over magnetic stirring at heterogeneous transesterification reaction, the effect of reaction tem-
room temperature had been studied by Ragavan and Roy [112]. perature is very important as the system consists of three phases (oil-
Transesterification of rubber seed oil was completed at 305 K with an alcohol-catalyst). The reaction can be slow due to the presence of dif-
optimum yield of 80.7% which could not be achieved by magnetic fusion barrier among the different phases [63]. The diffusion resistance
stirring method. No visible separation of methyl esters was observed between different phases will be minimized with increasing tempera-
after 15 min of magnetic stirring with methanol/oil/KOH molar ratio ture as the viscosity of the reaction mixture is reduced. Therefore, so-
6:1:0.13 at room temperature. Higher temperature was required by lubility of alcohol in oil phase will increase and the catalyst will have a
magnetic stirring method to initiate the transesterification process. better contact with the reactants [58].
Since ultrasound-assisted transesterification can take place at room Nonetheless, beyond the optimum reaction temperature, the bio-
temperature (303 K), which is much lower than the boiling point of diesel yield will decrease because a higher reaction temperature will
alcohol, it could reduce external heat energy effectively [113]. also accelerate side reactions such as saponification [44,104]. Another
The reaction rate and yield of the biodiesel product are clearly in- reason can be ascribed to the damping effect of the cavitation effects at
fluenced by reaction temperature [2]. At lower reaction temperature, higher operating temperatures [4,98]. At higher reaction temperature,
the oil has higher viscosity and impedes the bubble formation, resulting the equilibrium vapour pressure will increase which leads to easier
in poor mixing between the oil and alcohol-catalyst phases [38,107]. As bubble formation. However, the cavitation bubbles formed contain
reaction temperature rises, the viscosity of oil will decrease and it be- more vapours. The release of solvent vapours interferes with the cavi-
comes susceptible to cavitation. Consequently, the miscibility of alcohol tation effects, causing cushioning of collapse. Therefore, bubbles

897
S.X. Tan et al. Fuel 235 (2019) 886–907

implode with less intensity and reduces the mixing effect of ultrasound 4.6. Phase separation time
on the transesterification reaction, resulting in reduced mass transfer
and biodiesel yield [31,38,59]. Generally, the separation of the reaction mixture containing FAME,
glycerol and excess methanol requires several hours through gravity
settling. The separation time depends on the amount of excess methanol
4.5. Energy consumption used and the temperature of the reaction mixture. Separation will
generally occur faster in the reaction mixture containing less excess of
In terms of energy consumption, mechanical stirring usually re- methanol. This fact could be explained by the differences in the density
quires higher energy consumption than ultrasonication. It was reported of the glycerol phase due to the varying excess volume of methanol.
that ultrasonication only required one-third to a half of the energy that Glycerol has much higher density (1.26 g/cm3) than methanol (0.79 g/
was consumed by mechanical agitation [30]. Yin et al. [49] had re- cm3), and both liquids are polar compounds which will form a uniform
corded the energy consumption for both mechanical stirring and ul- phase at any mixing ratio. Therefore, the glycerol layer with smaller
trasonic probe irradiation methods to reach biodiesel conversion of excess of methanol has a higher density, resulting in a faster phase
95% from sunflower oil. It was observed that mechanical stirring re- separation between the FAME layer and the glycerol layer due to the
quired 0.31 kWh whereas ultrasonic probe irradiation only needed larger difference in the density of both layers. Lower temperature of the
0.18 kWh. Ultrasonication was also discovered to be more energy effi- reaction mixture will contribute to lower dissolution of glycerol and
cient than conventional stirring by other researchers such as Gupta methanol in the FAME layer, thus leading to shorter separation time
et al. [59], Yadav et al. [61] and Brasil et al. [114]. More energy was [5].
consumed by conventional stirring as continuous heating was required As mentioned previously in Sections 4.1 and 4.4, ultrasound ap-
to maintain the temperature of water bath or solution [49,61]. External proach requires lower alcohol to oil molar ratio and lower reaction
heating source is unnecessary in biodiesel synthesis aided by ultrasonic temperature compared to conventional stirring method. Therefore, it
irradiation. This is because cavitation from ultrasound will result in can be assumed that lower excess of alcohol is usually present through
localized increment in temperature at the phase boundary and subse- ultrasound approach. The assumption could be further affirmed since
quently enhanced the reaction. Therefore, less energy is consumed in ultrasonic transesterification requires lower reaction temperature than
the production of biodiesel in the long run [49]. conventional transesterification using either mechanical or magnetic
Delivery of the ultrasonic power to the reaction mixture determines stirring. Hence, the overall temperature of the reaction mixture will be
the degree of cavitation and thus directly affecting the biodiesel yield lower for ultrasonic agitation.
[59]. Optimization of ultrasonic power allows achieving the highest In the transesterification of WVOs carried out by Refaat et al. [115],
biodiesel conversion and yield at the lowest possible power, which re- biodiesel yield of 96.15% was achieved under optimum conditions
duce the production cost of biodiesel. Generally, biodiesel conversion (methanol/oil molar ratio of 6:1, 1 wt% KOH and 338 K) after reaction
increases with an increment in ultrasonic power until optimum ultra- time of 1 h with the aid of mechanical stirring. These optimum condi-
sonic power is reached. When the ultrasonic power is higher, the col- tions were then implemented again using ultrasonication (100 W,
lapse of the cavitation bubbles will be harsher due to higher jet velocity, 20 kHz) in another study by Refaat and El-Sheltawy [116]. It was found
thus enhancing micromixing between the oil and methanol phases at out that the separation time was reduced remarkably from 8 h to 25 min
the phase boundary. Finer emulsion is then formed, increasing mass- in addition to shorter reaction time (reduced from 1 h to 5 min) and
transfer coefficient and biodiesel conversion. On the other hand, bio- higher yields (98–99%) were achieved. According to Kumar et al.
diesel conversion is reduced when further increasing the ultrasonic [117], ultrasonication reduced the separation time from 5 to 10 h to less
power beyond the optimum value. Cushioning effect at higher ultra- than 30 min in the production of biodiesel from Jatropha oil at me-
sonic power reduces the energy transfer into the system, thus providing thanol: ethanol: oil molar ratio of 3:3:1 with 0.75 wt% catalyst. The
lower cavitation activity [36]. Furthermore, huge amount of cavitation ease of separation of glycerol and catalyst was also found to be higher
bubbles will be generated when high ultrasonic power is introduced in the case of ultrasonication compared to mechanical stirring for the
into the reaction mixture. Excessive bubbles will likely merge and form biodiesel synthesis from Nagchampa oil by Gole and Gogate [70], which
larger and more stable (long-lived) bubbles and create a barrier to reduced the purification time and energy.
acoustic energy transfer. This phenomenon of efficiency loss in the
power transfer is referred as decoupling effect [31]. 4.7. Ultrasonic pulse mode

Ultrasound can be applied in pulse mode or continuous mode.

Fig. 3. Wave profiles for (a) continuous sonication (b) pulse sonication. Adapted with permission from (Martinez-Guerra E, Gude VG) [31]. Continuous and pulse
sonication effects on transesterification of used vegetable oil. Energy Convers Manage. 2015;96:268–76. Copyright (2015) Elsevier Ltd.

898
S.X. Tan et al. Fuel 235 (2019) 886–907

Examples of wave patterns for continuous and pulse sonication (5 s on/ detected for further enhancement in the duty cycle to 70% (7 s on/3 s
1 s off) are shown in Fig. 3(a) and (b), respectively. For continuous off) [59]. Therefore, it could be concluded that if the pulse (duty cycle)
sonication (Fig. 3(a)), ultrasound waves are delivered continuously and exceeded the optimum range, the biodiesel conversion or yield would
relaxation intervals are impossible for the liquid phase in the reaction increase with a lower rate. This was because in such condition, the
medium, thus increasing thermal energy of the reaction medium. As a effect of initial vibrational shock applied to the reactants by ultrasound
result, energy input will be partially lost as they are being transformed waves would become identical with uniform waves [87].
into vibrational or cavitation energy. Transesterification reaction may
be energized by the thermal effect of the continuous sonication but
4.8. Biodiesel conversion or yield
prolong exposure is not recommended. This is because unfavourable
emulsification of the reactants may complicate the product separation
In order to compare the biodiesel conversion or yield attained be-
owing to over excitation of the reactants. On the contrary, relaxation
tween conventional stirring and ultrasonication, several researchers
intervals are possible for pulse sonication (Fig. 3(b)), rendering ex-
had conducted the experiments under similar reaction conditions in
citation gap to the reaction mixture to form good emulsions without
which the results are tabulated in Table 9. From Table 9, it can be
increasing thermal energy of the reaction medium significantly.
observed that ultrasonication can increase the biodiesel yield or con-
Nevertheless, longer sonication time is required by pulse sonication
version by 4.0%–27.9% in comparison to conventional stirring. Since
than continuous sonication as lower excitation energy is imparted [31].
the transesterification reaction is mass controlled due to the im-
Chand et al. [105] had proven that pulse mode gave a greater yield of
miscibility of methanol and oil, the micro-turbulence generated due to
esters than continuous mode. In their study, pulse mode (5 s on/25 s
cavitation bubbles can result in larger interfacial area and higher
off) and continuous mode (for 15 s) were compared. 96% biodiesel
temperature and pressure. For conventional stirring, agitation intensity
yield was obtained in less than 90 s in the pulse mode while 86% yield
plays an important role in the transesterification reaction. The mass
was achieved in 15 s for continuous mode. These results are in agree-
transfer of TGs from oil phase towards the methanol–oil interface can
ment with the study carried out by Martinez-Guerra and Gude [31]. In
be the rate limiting step and hence poor mass transfer between the two
addition to lower the yield produced, application of ultrasound in
phases results in slow reaction rate [72]. In short, it can be concluded
continuous mode was also found to cause tip erosion and was less en-
that ultrasonication can achieve higher biodiesel conversion or yield
ergy efficient than pulsed mode [118].
than conventional stirring under comparable reaction conditions.
Pulse is also known as the ratio of ultrasound working time to its
idling time [87]. It could also be expressed in duty cycle, which is the
percentage of ultrasound working time to the total ultrasound working 5. Kinetic study on ultrasound-assisted transesterification
time and idling time. At lower pulse, lower biodiesel conversion is
obtained due to the macro-stirring effect of the ultrasound which is too There are ample of studies which had been published on the ul-
mild to mix the immiscible reactants evenly. By increasing the pulse to trasound-assisted transesterification by utilizing non-edible feedstocks.
a certain extent, the biodiesel conversion increases as better emulsifi- However, the fundamental info such as kinetic studies on ultrasound-
cation of the two immiscible layers is achieved [98]. The effect of using assisted transesterification were rarely investigated [119]. From the
pulsed ultrasound was investigated by Hingu et al. [98] in the biodiesel kinetic study, information on reaction order, reaction rate and Ea can be
synthesis from WCO for 40 min reaction time. For the pulse 2 s on/2 s obtained which is vital for process scale-up.
off (duty cycle of 50%), the extent of conversion was 62% and 5 s on/1 s For the determination of reaction order, different assumptions of the
off (duty cycle of ≈83%) increased the conversion to 65.5%. For pulse reaction order are applied. Pseudo first-order reaction is assumed when
duration of 1 min on/5 s off (≈92%), conversion of 89.5% was acquired the amount of methanol in the reactants is large enough for the me-
and the yield was further increased to 94.5% after dry washing the thanol concentration to remain constant throughout the reaction
esters. In the ultrasonic transesterification of WCO using 1 wt% CaDG as [113,120]. For transesterification of TG due to the presence of excess
a catalyst, methanol to oil molar ratio of 9:1 and 333 K, biodiesel yield methanol, the reaction order can be considered as an irreversible
increased when the duty cycle increased from 30% (3 s on/7 s off) to pseudo second-order reaction [58,59]. Then the reaction rate constant
50% (7 s on/3 s off). However, no significant increment in the yield was (k) can be found using Eq. (4) for first-order reaction and Eq. (5) for
second-order reaction [58,121].

Table 9
Comparison of biodiesel yield or conversion of different feedstocks using ultrasound and conventional stirring approaches.
Year Feedstock Type of ultrasound/stirring Ultrasonic cavitation Conventional stirring Increase percentage Ref.
(%)
Yield (%) Conversion (%) Yield (%) Conversion (%)

2010 WCO • Ultrasonic horn (200 W, 20 kHz) − 89.5 − 57.5 32 [98]


• Mechanical stirring (1000 rpm)
2010 Oreochromis niloticus • Ultrasonic bath (60 W, 40 kHz) 98.2 − ∼85 − 14.2 [50]
oil • Mechanical stirring
2011 Rubber seed oil • Ultrasonic bath (33 ± 3 Hz) 91 − 87 − 4 [112]
• Magnetic stirring (> 300 rpm)
2012 Jatropha oil • Ultrasonic horn (400 W, 40 kHz) − 98 − 79 19 [72]
• Mechanical stirring (600 rpm)
2013 WCO • Stirrer (6-blade turbine, 1000 rpm) 90 − 70 − 20 [25]
2014 WCO • Ultrasonic horn (22 kHz, 375 W) 92 − 59 − 33 [58]
• Overhead stirrer (1000 rpm)
2015 WCO • Ultrasonic horn (120 W, 22 kHz) 93.5 − 65.6 − 27.9 [59]
• Overhead stirrer (800 rpm)
2015 Mahua oil • Ultrasonic horn (1000 W, 20–36 kHz) 97.4 − 93 − 4.4 [71]
• Magnetic stirring (500 rpm)
2016 Schleichera triguga oil • Ultrasonic probe (250 W, 20 kHz) − 96.8 − 63.29 33.51 [77]
• Conventional stirring (did not specify the
condition in the paper)

899
S.X. Tan et al. Fuel 235 (2019) 886–907

k = −ln(1−X )/ t (4) double-step process. Higher Ea of the ultrasonic single step process
could be attributed to catalyst inhibition due to water traces in the
kCTG0 = [X /(1−X )]/ t (5) reaction mixture. On the other hand, absence of water molecules in the
Several researchers had performed kinetic studies on ultrasound- reaction mixture in conventional double-step process contributed to
assisted transesterification in which the rate constant (k), Ea and fre- reduction in Ea. In the transesterification of WCO using CaO (conven-
quency factor (A) for the transesterification at optimum operating tional) and CaDG (ultrasound), CaO (conventional) only required
conditions are tabulated in Table 10. The lower the value of Ea and the around 2/3 of the Ea required by CaDG (ultrasound). The Ea for CaO
higher the value of kinetic constant will indicate a high reaction rate (conventional) and CaDG (ultrasound) were 79 and 119.23 kJ/mol,
[149]. It can be noticed from Table 7 that Ea is the highest during es- respectively. It was estimated that CaDG (ultrasound) should require
terification of Jatropha oil catalyzed by H2SO4 which is 167.419 kJ/ lower Ea than CaO (conventional) due to the presence of ultrasound in
mol. Such high Ea would result in lower reaction kinetics [52]. High enhancing the mass transfer limitation. In addition, catalytic activity of
temperature was required for high Ea in order to increase the solubility CaDG was higher than CaO due to the presence of basic oxygen anion,
of methanol in the ester phase which in turn increased the reaction rate which could easily attract protons from OH group of methanol to form
(reaction kinetics). However, cavitation effects of ultrasound at higher methoxide ions [59]. Therefore, the only possible reason to explain this
temperature would also dampen due to cushioned collapsing of cavities phenomena could be due to the high amount of impurities present in
[94]. As observed from Table 10, the optimum temperature required by the WCO which poisoned the catalyst through absorption [54]. In the
the esterification of Jatropha oil catalyzed by H2SO4 was the highest esterification of Karanja oil conducted by Mohod et al. [85], the kinetic
(343 K). Although bubble formation was easier at higher temperature, rate constants at optimum catalyst loading of 2 wt% H2SO4 were
more vapours were found in the cavitational bubbles. Interference of 0.275 × 10−3 and 0.262 × 10−3 L·mol−1·min−1 using ultrasonication
the solvent vapours with the cavitation effect will affect the mixing and mechanical stirring, respectively. Similarly, in the transesterifica-
effect of the ultrasound on the esterification reaction, resulting in a tion of Silybum marianum oil, Takase et al. [107] discovered that the
decrement in solubility of methanol in ester phase and lower reaction reaction rate constants via ultrasonication with methanol and ethanol
[31,38]. Hence, chlorosulfonic acid was employed instead of H2SO4 in were higher than via mechanical stirring. Higher kinetic rate constant
another study carried out by Choudhury et al. [33]. The Ea for the attained by ultrasonication could be ascribed to better efficacy of ul-
process (57 kJ/mol) was at least 3 times lower than the energy for trasonication than mechanical stirring to generate microemulsion from
H2SO4 catalyzed transesterification. The huge difference of the Ea was the intense mixing of methanol and oil by cavitation effects.
likely due to the difference in the active sites of transesterification re-
action with H2SO4 and chlorosulfonic acid catalyst. 6. Pilot reactors for ultrasound-assisted biodiesel production
It can also be observed from Table 10 that there are three studies
which synthesized WCO using different types of alcohol (methanol and The utilization of ultrasound in biodiesel production has presented a
methyl acetate) and catalysts (potassium methoxide, K3PO4 and CaDG). huge possibility for implementation in pilot or industrial scales.
By rearranging the Ea required by each catalyst in ascending order, However, only a few examples of the relevant case studies were re-
potassium methoxide required the lowest Ea (56.97 kJ/mol), followed ported, which were tabulated in Table 11. From Table 11, it was no-
by K3PO4 (64.24 kJ/mol) and CaDG (119.23 kJ/mol). The optimum ticed that the biodiesel yields were quite high which were mostly above
temperatures to achieve optimum biodiesel yield for potassium meth- 90%. However, lower biodiesel yield was observed when the reactor
oxide, K3PO4 and CaDG were 313 K, 323 K and 333 K, respectively, volume was higher than 6 L. According to the ultrasound-assisted bio-
which affirmed that higher Ea required higher temperature. Lowest Ea diesel synthesis from canola oil conducted by Stavarache et al. [122],
for potassium methoxide could be attributed to its two-phase homo- the biodiesel yield decreased from 95% to 50% when the reactor vo-
geneity as compared to the triphasic system (liquid/liquid/solid) cata- lume increased from 2.62 L to 6.35 L. The reduction in the biodiesel
lyzed by K3PO4 and CaDG. Triphasic system often encountered higher yield was attributed to a decrement in ultrasonic power and bubble
mass transfer limitations which resulted in higher Ea. It was noticed that density, which reduced the contact area between the reactants and
the Ea required by CaDG was almost double compared to the Ea re- induced mass transfer limitation. Another important observation was
quired by both potassium methoxide and K3PO4. The possible reason that when the residence time increased from 20 to 30 min while
could be due to the high impurity of the WCO which hindered the ac- maintaining the reactor volume (6.35 L), the biodiesel yield decreased
tivity of the catalyst [54]. from approximately 90% to 50%. It was suspected that accumulation of
The kinetic behaviours of the ultrasonic single step (chlorosulfonic glycerol at the bottom of the reactor would trap more methanol mo-
acid) and conventional double-step processes (H2SO4 and chlor- lecules into the glycerol layer especially when longer residence time
osulfonic acid) were compared by Choudhury et al. [33]. The Ea for was employed. This could occur since both the glycerol and methanol
ultrasonic single step process was 57.33 kJ/mol which showed a re- were polar compounds; hence they could dissolve with each other at
duction of approximately 40% to 31.29 kJ/mol for the conventional any ratio. Therefore, there was insufficient methanol to promote a

Table 10
Rate constant, activation energy and frequency factor for the transesterification at optimum operating conditions.
Year Feedstock Alcohol Catalyst Reaction Rate constants, k Activation energy, Ea Frequency factor, A Ref.
order (kJ/mol)

2014 Silybum marianum Methanol KOH First 2.3 × 10−2 s−1 (333 K) − − [107]
oil
Ethanol KOH First 7.0 × 10−3 s−1 (353 K) − −
2014 Jatropha oil Methanol Chlorosulfonic acid First 1.33 × 10−2 min−1 (333 K) 57.33 1.57 × 107 min−1 [33]
2014 Jatropha oil Methanol CaO Third 1.46 × 10−3 L2∙mol−2∙s−1 133.5 4.32 × 1017 L2∙mol−2∙s−1 [54]
(338 K)
2015 WCO Methanol K3PO4 Second 0.14 L∙mol−1∙min−1 (323 K) 64.24 2.95 × 109 L∙mol−1∙min−1 [58]
2015 WCO Methanol CaDG Second 0.198 L∙mol−1∙min−1 (333 K) 119.23 9.07 × 1017 L∙mol−1∙min−1 [59]
2016 Schleichera triguga Methanol H2SO4 First 0.236 min−1 (313 K) 21.75 1075.99 min−1 [77]
oil
Methanol Ba(OH)2 Second 0.18 L∙mol−1∙min−1 (323 K) 53.26 6.1 × 107 L∙mol−1∙min−1

900
S.X. Tan et al.

Table 11
Existing pilot reactors for ultrasonic biodiesel production.
Year Process Feedstock Reactor Flowrate Frequency (kHz); Residence time Methanol-to-oil Temperature (K) Catalyst; Yield (%); time (min) Ref.
volume (L) (mL/min) Power (W) (min) molar ratio (mol/ loading (wt%)
mol)

2007 Continuous process Commercial 2.62 − 45; 600 10 6:1 311–313 KOH; − > 90; − [122]
edible oil
20 > 90; −
30 > 90; −
2.62 − 45; 600 10 7.5:1 311–313 KOH > 95; −
20 > 95; −
30 > 95; −
6.35 − 45; 600 20 7.5:1 311–313 KOH less than90; −
30 50; 360
2010 Circulation process Canola 0.8 8000 20; 1000 40 5:1 Room KOH; 0.7 99; 50 [5]
temperature

901
2010 Two-stage continuous process WCO 0.8 1st stage: 1.5 20; 1000 1st stage: 0.53 1st stage: 2.5:1 293–298 KOH; 1st 1st stage: 81; 25 [123]
2nd stage:2.0 2nd stage: 0.40 2nd stage: 1.5:1 stage: 0.7 2nd stage: 99; 20
2nd stage: 0.3
2010 Two-step process (conventional heating under Soybean oil 5 55 21.5; 600 9 6:1 318 KOH; 0.15 − [131]
mechanical stirring for 30 min at 318 K followed
by ultrasonication for 35 min)
2015 Circulation process Soybean oil 5 2000 19; 600 60 6:1 333 MeONa; 0.6 97; 15 [114]
2016 Three-step process (esterification and WCO 6 16 30; − − 8:1 − KOH; − 78; 40 [132]
neutralization in batch mode followed by
transesterification in continuous mode)
2017 Batch mode (longitudinal horn equipped with Karanja oil 4 − 36; 150 − 5:1 Room H2SO4; 2.0 80.9 (in term of acid [85]
two mechanical stirrers) temperature value reduction); 70
Fuel 235 (2019) 886–907
S.X. Tan et al. Fuel 235 (2019) 886–907

constant conversion [122,123]. According to Thanh et al. [123], lower and methanol with the immobilized enzymes. For the case of PBR,
flow rate which corresponded to longer residence time would enhance continuous flow rate allowed the reaction mixture to get as much
the emulsion efficiency of the reactants, resulting to an increment in contact as possible with the immobilized enzymes and to move with the
biodiesel yield. Nonetheless, one important point to note is that bio- generated water from esterification reactions while avoiding the accu-
diesel yield or conversion is affected by the combined effect of flow rate mulation of the traces of glycerol inside the bed. Therefore, biodiesel
or residence time and ultrasonic amplitude in the ultrasonic transes- yield achieved by PBR was twice of that by batch reactor. Since the
terification processes [124]. Employing low ultrasonic amplitude at extent of biodiesel conversion was higher for the case of PBR than batch
higher flow rate decreases the ultrasonic power dissipated into the re- reactor, kinematic viscosity also followed the same trend.
actor, thus lowers the biodiesel yield due to low cavitation and acoustic Furthermore, the reactor characteristics have to be tailored in terms
jets. On the other hand, employing high ultrasonic amplitude at lower of the rheological properties of the reaction mixture (solution or sus-
flow rate increases the ultrasonic power which then promotes the pension, viscosity, surface tension, etc.). This is because the behaviour
backward reaction and therefore decreases biodiesel yield. More studies of micro bubbles will be affected by the rheological properties of re-
should be carried out to investigate the effect of residence time or flow action mixture [128]. For example, the rheological properties of reac-
rate on the biodiesel yield by taking ultrasonic amplitude into con- tion mixture will change when the concentration of alcohol decreases
sideration. while the concentration of glycerol and biodiesel increase at the same
The reason of lacking industrial sonochemical processes might be time. Such changes can easily affect the bubbles dynamics, which
due to the complicated design of efficient cavitation reactors, given the eventually influence the effectiveness of ultrasonic irradiation in the
lack of data linking bubble dynamics with experimental predictions for reaction mixture. When the concentration of alcohol decreases, the
large-scale processes. Biodiesel from Cedrus deodara oil had been syn- viscosity of the reaction media will increase which will impede the
thesized by Mohan et al. [125] using four different ultrasonic reactors growth of the micro bubbles. Its radius size will reduce along with a
which were triple-frequency flow cell, double-frequency flow cell, bath- drop in the bubble temperature and pressure as well as a decrement in
type and horn-type. Energy consumptions by using these four types of oscillatory velocity. This will subsequently result in reduced effect on
ultrasounds were compared and they found that triple-frequency ul- enhancing the quality of the micro emulsion as the reaction proceeds
trasonic reactor was the most energy efficient. The order of energy [128]. Apart from rheological properties of reaction mixture, operating
consumption was triple-frequency flow cell < double-frequency flow conditions of the reaction mixture will also influence the behaviour of
cell < bath-type < horn-type. Their results were in well agreement micro bubbles. According to the study performed by Sajjadi et al. [128],
with the results reported by Gogate et al. [32]. Reaction times to intensity of ultrasound irradiation played the most important role
achieve optimum biodiesel yields for triple-frequency, double-fre- compared to reaction temperature and alcohol concentration. Radius,
quency, bath-type and horn-type ultrasonic reactors were 99% in internal temperature and pressure of bubbles significantly increased
15 min, 99% in 30 min, 99% in 45 min and 98.5% in 50 min, respec- with power amplitude due to stronger expansion and greater energy
tively, at 60% of the maximum rated power conditions. From their accumulation. At the same time, increase in reaction temperature and
studies, it was revealed that triple-frequency ultrasonic reactor was able alcohol concentration made the bubbles characteristics rather mod-
to attain the optimum biodiesel yield under the shortest reaction time erate. It was also found that the growth of bubbles radius was reduced
with the lowest energy consumption. Recently, effects of different ul- by about 4 times as the reaction progressed and the reaction mixture
trasound modes (single-frequency, dual-frequency of sequential mode became more viscous and dense. In order to validate their findings,
(SQM) and simultaneous mode (SMM)) on biodiesel conversion were more studies should be conducted as the available data is limited. In
studied by Yin et al. [126] with the reactants in a counter-current state. addition, the effect of other parameters (catalyst loading, reactor vo-
It was revealed that biodiesel conversion enhanced by single-frequency lume, reactor design and flow rate) on micro bubbles’ behaviour in the
was lower than that enhanced by dual-frequency. This was because ultrasound-assisted biodiesel production at pilot or industrial scale
single-frequency reactor suffered from non-uniform volumetric energy should be examined by employing different catalysts and other non-
dissipation in the reaction medium (due to directional sensitivity of the edible feedstocks.
ultrasound field) that limited the yield of the chemical transformation. Ultrasound can also be synergised in tandem with supercritical
On the other hand, multi-frequency ultrasound had better efficiency method for continuous biodiesel production in pilot reactor. As men-
than the single frequency in terms of higher pressure and temperature tioned in Section 1, high operating temperature, pressure and me-
produced from the implosive collapse of cavitation bubbles which led to thanol-to-triglycerides molar ratio were required by supercritical
higher biodiesel yield. Moreover, multi-frequency ultrasound generated method. All these factors will impair the desired profitability of in-
strong interference pattern between two ultrasound waves. This inter- dustrial production and render supercritical method becomes less
ference reduced the directional sensitivity of single-frequency ultra- competitive with conventional biodiesel production methods [129]. In
sound and aided in achieving more uniform spatial energy dissipation the pilot scale of continuous rapeseed biodiesel production assisted by
in the reaction medium, which resulted in overall volumetric efficiency. supercritical methanol alone, poor mixing of alcohol and oil led to low
For dual-frequency, biodiesel conversion of SMM was higher than that reaction rate and increased reaction time, which contributed to higher
of SQM. It was ascribed to enhanced mechanical disturbance of the energy consumption for the process. With the synergistic effect of ul-
reaction medium by simultaneous dual-frequency irradiation which trasound, reaction rate and biodiesel conversion were increased
generated more cavitation bubbles than sequential dual-frequency ir- through increasing the solubility of oil in alcohol and the area of con-
radiation. Moreover, types of reactors used during ultrasound-assisted tact surface of the immiscible phases (vigorous mixing) [130]. In the
production of biodiesel will also affect the biodiesel yield. Murillo et al. continuous production of biodiesel from rapeseed oil by ultrasound-
[127] had synthesized biodiesel from WVO with the assistance of ul- assisted transesterification in supercritical ethanol conducted by Ma-
trasound and Burkholderia cepacia lipase as catalyst. They evaluated the zanov et al. [129], the effects of ethanol to rapeseed oil molar ratio
kinematic viscosity and the yield of biodiesel produced with the utili- (12:1–20:1), temperature (623 K–653 K), catalyst loading and types
zation of two different reactors which were packed bed reactor (PBR) (ZnO/Al2O3, MgO/Al2O3 and SrO/Al2O3) on fatty acid ethyl ester
and batch reactor. Based on their findings, PBR achieved better bio- (FAAE) yield were investigated. Biodiesel yield of 97.46% was achieved
diesel yield (67.30% versus 32.70%) and lower kinematic viscosity under optimum conditions of 12:1 ethanol to oil molar ratio, SrO/Al2O3
(7.88 mm2∙s−1 versus 25.44 mm2∙s−1 at 293 K) than batch reactor. For (2 wt% SrO) at 623 K. However, there were only few publications re-
the case of batch reactor, discrete quantities of volume were used for porting on the continuous production of biodiesel from edible feedstock
the synthesis of biodiesel. Lack of sufficient flow current resulted in by ultrasound-assisted transesterification combined with supercritical
limited contact of the interfacial surfaces between the particles of WVO method. There was also lack of study pertaining to the case of using

902
S.X. Tan et al. Fuel 235 (2019) 886–907

non-edible feedstock. Therefore, more studies should be carried out to time of 20 min. Kumar et al. [139] attained the maximum biodiesel
investigate the feasibility of continuous biodiesel production assisted by conversion of 92% within 20 min by using methanol to seed ratio (w/w)
ultrasound and supercritical method. of 100:1, > 1− < 2 mm seed size, 1.5 wt% KOH, 50% of ultrasonic
amplitude and 0.3 s cycle.

7. Ultrasonic reactive extraction for biodiesel production


7.1. Mechanism of ultrasonic reactive extraction on oil seed for biodiesel
Reactive extraction (RE) is also known as in situ transesterification production
in which the oil bearing materials such as oil seeds or seed cakes contact
directly with an alcohol solution at ambient temperature and pressure. The suggested ultrasonic extraction mechanism by Jadhav et al.
In this method, alcohol acts as an extracting solvent and an esterifica- [140] is shown in Fig. 4. In the first step of the extraction processes, the
tion reagent, so both oil extraction and transesterification of oil into irregular dry seed particles are surrounded by a solvent layer. Uptake of
biodiesel proceed in one step [30]. In other words, once the oil is ex- the solvent by the irregular dry seed particles causes them to begin
tracted out from seeds, it is subsequently converted to esters [133]. swelling, expanding and acquiring a smoother shape. A dynamic in-
Expensive costs employed for the oil extraction and degumming in the teraction is present at the surface of the seeds in the course of swelling
biodiesel production can be reduced as the isolation and refining of oil and decreases successively upon reaching the maximum limit of swel-
seeds are not needed. In addition, the biodiesel yield could be max- ling. Direct diffusion is then being interfered by the formation of the
imized up to 98% [27,30]. It was reported that RE can reduce biodiesel stagnant layer surrounding the seed particle which possibly causes the
production cost by integrating multiple biodiesel processing stages extraction process to cease. In the second step, solvated compound
which constitute over 70% of the overall cost, even when using refined (seed oil) moves towards the stagnant layer and diffuses into the bulk
oil as feedstock [134]. In order to reduce the alcohol requirement for solvent. The extraction efficiency will be hampered due to the presence
high efficiency of RE, the oil seeds need to be dried prior to the reaction of stagnant layer as a diffusion barrier. Diffusion process will become
taking place [135]. harder if the stagnant layer has high adhesiveness and the extraction
Integration of ultrasonication in the RE process can provide suitable yield will be lower. Moreover, blockage of crevices by the stagnant
pre-treatment condition to change the physical and chemical structure layer could hamper the diffusion of seed oil [140].
of the lignocellulosic seed. Ultrasound used in the process could induce By employing ultrasonication, larger amplitude ultrasonic wave
physical pre-treatment through the formation of cavitation bubbles in passes through the liquid media (solvent). This creates and collapses the
the liquid phase that grow and violently collapse. Furthermore, addi- cavities (bubbles) in a very short time, which contain very high tem-
tion of ultrasonication could induce mechanical treatment for dis- perature and pressure. The asymmetric collapse of these cavities results
rupting biological structure and increase the surface area while en- in the generation of violent shock waves and high-speed jets. These
hancing the saccharification of cellulose for maximum cavitation effect shock waves then strike the surface of the seed, which break the stag-
at 323 K. Moreover, the thermal temperature employed in the process nant layer as well as create cracks, crevices, and microfractures. At the
could deactivate the toxic phorbol ester while producing crude bio- same time, impingement by high-speed jets results in surface peeling,
diesel, glycerol by-product and completely de-oiled seed residues erosion and particle breakdown. Due to these physical effects, the sol-
[136]. vent can easily penetrate the plant’s cell wall, diffuse through the seed,
However, there are only few studies conducted on ultrasound-as- reach the cell tissues, accelerate the extraction of intracellular products
sisted RE directly from non-edible oil seeds. Chadha et al. [137] (i.e. oil bodies) with solvent and permit intracellular products release
achieved 94.1% of biodiesel yield with the assistance of ultrasonic into the extraction environment (solvent) [140,141]. Once the oil
probe (22 kHz; 750 W) under the optimum conditions of methanol to bodies have been extracted, they will be converted to biodiesel si-
seed ratio (w/w) of 10:1, 2–3 mm seed size, 1 wt% NaOH and 80 min multaneously [134].
reaction time. Zahari et al. [138] obtained 62.21% biodiesel yield with The proposed mechanism was confirmed by Jadhav et al. [140]
0.15 N NaOH and < 1 mm seed size. The details of methanol to seed using the field emission scanning electron microscopy (FE-SEM) ana-
ratio, reaction temperature and reaction time in achieving biodiesel lysis of raw and ultrasonically treated date seed. Raw date seed had a
yield of 62.21% were not mentioned in their studies. Koutsouki et al. regular, intact, smooth surface (as shown in Fig. 5(a) and (c)) whereas
[109] synthesized biodiesel from Cynara cardunculus L. seeds with the ultrasonically treated date seed had irregular shape with formation
assistance of ultrasonic horn (24 kHz, 400 W, 80% of ultrasonic am- microfractures and crevices on its surface (as shown in Fig. 5(b) and
plitude and cycle at 0.7). Optimum biodiesel yield of 96% was achieved (d)). The breakdown of seed due to the physical effect of cavitation had
at methanol to oil molar ratio of 550:1, 9.5 wt% NaOH and reaction also been confirmed from the particle size distribution of the fresh and

Fig. 4. Brief overview of ultrasonic extraction mechanism. Adapted with permission from (Jadhav AJ, Holkar CR, Goswami AD, Pandit AB, Pinjari DV) [140].
Acoustic cavitation as a novel approach for extraction of oil from waste date seeds. ACS Sustainable Chemistry & Engineering. 2016;4:4256–63. Copyright (2016)
American Chemical Society.

903
S.X. Tan et al. Fuel 235 (2019) 886–907

Fig. 5. FE-SEM images of raw date seed at (a) 2000 X and (c) 10,000 X and ultrasonically treated date seed at (b) 2000 X and (d) 10,000 X. Adapted with permission
from (Jadhav AJ, Holkar CR, Goswami AD, Pandit AB, Pinjari DV) [140]. Acoustic cavitation as a novel approach for extraction of oil from waste date seeds. ACS
Sustainable Chemistry & Engineering. 2016;4:4256–63. Copyright (2016) American Chemical Society.

ultrasonically treated date seed powder (as shown in Fig. 6). Ultra- feedstocks. This can gain a better insight on the behaviour of the pro-
sonically treated seed had narrower particle size distribution (in the cess under different optimum conditions.
range of 500 nm) as compared to raw seed (in the range of 1600 nm) Ultrasonic RE is a promising approach to synthesize biodiesel from
due to the shrinkage of seed size. seed directly. However, majority of the studies had performed ultra-
sonic RE from liquid oil directly instead of the solid oil-bearing seeds.
8. Challenges and feasibility for future development The optimum parameters obtained such as alcohol to oil ratio, catalyst
loading, ultrasonic power and ultrasonication mode by employing ul-
Based on the review discussed in this work, there are several im- trasonic RE from oil might not be suitable to achieve maximum bio-
minent developments which can be conducted in the future. Since there diesel yield for ultrasonic RE from oil seeds directly. Till now, the sci-
is still a knowledge gap on the fundamental studies dealing with reac- entific data on the ultrasonic RE from seed directly is very limited.
tion pathway of TGs on solid acid catalysts, more in-depth studies Future research could be directed more towards the effect of parameters
should be performed. Besides, more parameter studies can be con- on biodiesel yield or conversion, especially the effect of alcohol to seed
ducted on the two-step heterogeneous ultrasound-assisted transester- ratio instead of alcohol to oil ratio.
ification by utilizing non-edible feedstocks as the literature information The effect of various parameters such as catalyst loading, reactor
is still relatively scarce. For ultrasound-assisted interesterification, volume, reactor design and flow rate on the micro bubble’s behaviour in
parameter studies should be also conducted with different types of ultrasound-assisted biodiesel production at pilot or industrial scale also
catalyst (chemical and enzyme catalysts) and different non-edible requires further investigation. The reason of lacking industrial

904
S.X. Tan et al. Fuel 235 (2019) 886–907

This review has demonstrated that remarkable intensification in terms


of reduction in alcohol to oil ratio (16.67%−25% reduction), catalyst
amount (20%−25% reduction), reaction time (25%−95.8% reduc-
tion), reaction temperature, phase separation time and energy con-
sumption (33%−50% reduction) as well as improved biodiesel con-
version and yield (4%−27.9% increment) can be obtained with the
application of ultrasound in transesterification. Overall, ultrasonication
is advantageous to be used for process intensification in biodiesel pro-
duction and possesses huge potential to prepare biodiesel for industrial
production if the pilot reactor of ultrasound-assisted biodiesel produc-
tion is found to be economically feasible and the biodiesel produced is
of comparable quality and yield to fossil fuels.

Acknowledgements

The authors wish to acknowledge the University of Malaya, Kuala


Lumpur, Malaysia, for funding support of this work under the FRGS-
MRSA (MO014-2016) and University Tunku Abdul Rahman Research
Fund (UTARRF).
Fig. 6. Particle size distribution of fresh date seed powder and ultrasonically
treated date seed powder. Adapted with permission from (Jadhav AJ, Holkar References
CR, Goswami AD, Pandit AB, Pinjari DV) [140]. Acoustic cavitation as a novel
approach for extraction of oil from waste date seeds. ACS Sustainable Chemistry [1] Dharma S, Ong HC, Masjuki H, Sebayang A, Silitonga A. An overview of engine
& Engineering. 2016;4:4256–63. Copyright (2016) American Chemical Society. durability and compatibility using biodiesel–bioethanol–diesel blends in com-
pression-ignition engines. Energy Convers Manage 2016;128:66–81.
[2] Koh MY, Ghazi TIM. A review of biodiesel production from Jatropha curcas L. oil.
sonochemical processes might be due to the complicated design of ef- Renewable Sustainable Energy Rev 2011;15(5):2240–51.
[3] Su C-H. Recoverable and reusable hydrochloric acid used as a homogeneous cat-
ficient cavitation reactors, given the lack of sufficient info linking the alyst for biodiesel production. Appl Energy 2013;104:503–9.
bubble dynamics with experimental predictions for large-scale pro- [4] Parida S, Sahu DK, Misra PK. A rapid ultrasound-assisted production of biodiesel
cesses. Furthermore, the reactor characteristics have to be tailored in from a mixture of Karanj and soybean oil. Energy Sources Part A
2016;38(8):1110–6.
terms of the rheological properties of the reaction mixture (solution or
[5] Thanh LT, Okitsu K, Sadanaga Y, Takenaka N, Maeda Y, Bandow H. Ultrasound-
suspension, viscosity, surface tension, etc.). assisted production of biodiesel fuel from vegetable oils in a small scale circulation
In addition, it has been observed that majority of the studies have process. Bioresour Technol 2010;101(2):639–45.
employed simple designs of sonochemical reactors such as ultrasonic [6] Lee KT, Lim S, Pang YL, Ong HC, Chong WT. Integration of reactive extraction with
supercritical fluids for process intensification of biodiesel production: prospects
horn and ultrasonic bath which may not be effective at commercial and recent advances. Prog Energy Combust Sci 2014;45:54–78.
scale operation [141]. Besides, many experiments were performed in a [7] Talebian-Kiakalaieh A, Amin NAS, Mazaheri H. A review on novel processes of
batch reactor which could not provide a closer basis for the role of biodiesel production from waste cooking oil. Appl Energy 2013;104:683–710.
[8] Manickam S, Arigela VND, Gogate PR. Intensification of synthesis of biodiesel from
ultrasound-assisted reactor in an industrial scale [142]. Therefore, more palm oil using multiple frequency ultrasonic flow cell. Fuel Process Technol
comprehensive theoretical work on the pilot scale with continuous 2014;128:388–93.
processing should be carried out before commercialization of ultra- [9] Singh B, Bux F, Sharma Y. Comparison of homogeneous and heterogeneous cata-
lysis for synthesis of biodiesel from Madhuca indica oil. Chem Ind Chem Eng Q/
sound-assisted biodiesel production plant can be realized. For the CICEQ 2011;17(2):117–24.
combination of ultrasound and supercritical methods for continuous [10] Chuah LF, Bokhari A, Yusup S, Klemeš JJ, Akbar MM, Saminathan S. Optimisation
production of biodiesel in pilot scale, more studies should be performed on pretreatment of kapok seed (Ceiba pentandra) oil via esterification reaction in an
ultrasonic cavitation reactor. Biomass Convers Biorefin 2016:1–9.
for both edible and non-edible feedstocks with varying compositions. [11] Aransiola E, Ojumu T, Oyekola O, Madzimbamuto T, Ikhu-Omoregbe D. A review
Further researches may be directed in terms of the development of of current technology for biodiesel production: state of the art. Biomass Bioenergy
reactor design based on multi-frequency operation [141]. This finding 2014;61:276–97.
[12] Monisha J, Harish A, Sushma R, Krishna Murthy T, Blessy BM, Ananda S. Biodiesel:
would be beneficial in the design of an industrial scale of ultrasound-
a review.
assisted biodiesel production plant. However, no substantial work is [13] Meira M, Quintella CM, Ribeiro EMO, Silva HRG, Guimarães AK. Overview of the
found on energy consumption analysis and comparison among various challenges in the production of biodiesel. Biomass Convers Biorefin
types of ultrasonic reactors for biodiesel production [125]. Hence, more 2014;5(3):321–9.
[14] Badday AS, Abdullah AZ, Lee KT, Khayoon MS. Intensification of biodiesel pro-
studies should be performed to support and verify their claims. Another duction via ultrasonic-assisted process: a critical review on fundamentals and re-
important point is the economic analysis which also needs to be per- cent development. Renewable Sustainable Energy Rev 2012;16(7):4574–87.
formed based on the pilot scale demonstration plant’s performance data [15] Abbasi S, Diwekar UM. Characterization and stochastic modeling of uncertainties
in the biodiesel production. Clean Technol Environ Policy 2014;16(1):79–94.
in order to eliminate the need for unavoidable assumptions in the cost [16] Singh B. Production of Biodiesel from Plant Oils-An Overview. J Biotechnol Bioinf
estimations for the process units [38]. Bioeng 2014;1(2):33–42.
[17] Vyas AP, Verma JL, Subrahmanyam N. A review on FAME production processes.
Fuel 2010;89(1):1–9.
[18] Saleh J, Tremblay AY, Dubé MA. Glycerol removal from biodiesel using membrane
9. Conclusion separation technology. Fuel 2010;89(9):2260–6.
[19] Oh PP, Lau HLN, Chen J, Chong MF, Choo YM. A review on conventional tech-
Biodiesel is a strong candidate to replace fossil fuel as it is en- nologies and emerging process intensification (PI) methods for biodiesel produc-
tion. Renewable Sustainable Energy Rev 2012;16(7):5131–45.
vironmental friendly, economical competitive, technically feasible and [20] Mohapatra S, Das P, Swain D, Satapathy S, Sahu SR. A review on rejuvenated
the sources for biodiesel synthesis are easily available. Ultrasonication techniques in biodiesel production from vegetable oils. Int J Curr Eng Technol
is a promising method to replace mechanical or magnetic stirring in the 2016;6:100–11.
[21] Kiss AA. Novel process for biodiesel by reactive absorption. Sep Purif Technol
biodiesel production process. It is not only capable to eliminate the 2009;69(3):280–7.
mass transfer limitation encountered in the transesterification reaction, [22] Maddikeri GL, Pandit AB, Gogate PR. Intensification approaches for biodiesel
but also eliminates the needs of separate heating and agitation due to synthesis from waste cooking oil: a review. Ind Eng Chem Res
2012;51(45):14610–28.
the localized temperature increment and the formation of micro jets.

905
S.X. Tan et al. Fuel 235 (2019) 886–907

[23] D’Alessandro B, Bidini G, Zampilli M, Laranci P, Bartocci P, Fantozzi F. Straight metal oxides catalyst for ultrasonic-assisted transesterification of Jatropha oil into
and waste vegetable oil in engines: Review and experimental measurement of biodiesel. Aust J Chem 2016.
emissions, fuel consumption and injector fouling on a turbocharged commercial [56] Ramachandran K, Suganya T, Gandhi NN, Renganathan S. Recent developments
engine. Fuel 2016;182:198–209. for biodiesel production by ultrasonic assist transesterification using different
[24] Helwani Z, Othman M, Aziz N, Fernando W, Kim J. Technologies for production of heterogeneous catalyst: a review. Renewable Sustainable Energy Rev
biodiesel focusing on green catalytic techniques: a review. Fuel Process Technol 2013;22:410–8.
2009;90(12):1502–14. [57] Choedkiatsakul I, Ngaosuwan K, Assabumrungrat S. Application of heterogeneous
[25] Maddikeri GL, Pandit AB, Gogate PR. Ultrasound assisted interesterification of catalysts for transesterification of refined palm oil in ultrasound-assisted reactor.
waste cooking oil and methyl acetate for biodiesel and triacetin production. Fuel Fuel Process Technol 2013;111:22–8.
Process Technol 2013;116:241–9. [58] Pukale DD, Maddikeri GL, Gogate PR, Pandit AB, Pratap AP. Ultrasound assisted
[26] Islam A, Taufiq-Yap YH, Chan E-S, Moniruzzaman M, Islam S, Nabi MN. Advances transesterification of waste cooking oil using heterogeneous solid catalyst.
in solid-catalytic and non-catalytic technologies for biodiesel production. Energy Ultrason Sonochem 2015;22:278–86.
Convers Manage 2014;88:1200–18. [59] Gupta AR, Yadav SV, Rathod VK. Enhancement in biodiesel production using
[27] Baskar G, Aiswarya R. Trends in catalytic production of biodiesel from various waste cooking oil and calcium diglyceroxide as a heterogeneous catalyst in pre-
feedstocks. Renewable Sustainable Energy Rev 2016;57:496–504. sence of ultrasound. Fuel 2015;158:800–6.
[28] Kiss AA, Bildea CS. A review of biodiesel production by integrated reactive se- [60] Deng X, Fang Z, Liu Y-H, Yu C-L. Production of biodiesel from Jatropha oil cata-
paration technologies. J Chem Technol Biotechnol 2012;87(7):861–79. lyzed by nanosized solid basic catalyst. Energy 2011;36(2):777–84.
[29] El-Ibiari N, El-Enin SA, Gadalla A, El-Ardi O. El-Diwani G. Biodiesel production [61] Yadav AK, Khan ME, Pal A, Singh B. Ultrasonic-assisted optimization of biodiesel
from castor seeds by reactive extraction conventionally and via ultra-sound using production from Karabi oil using heterogeneous catalyst. Biofuels 2016:1–12.
response surface methodology. IJISET-International Journal of Innovative Science. [62] Jogi R, Murthy YS, Satyanarayana M, Rao TN, Javed S. Biodiesel production from
Eng Technol 2014;1(6).. degummed Jatropha curcas oil using constant-temperature ultrasonic water bath.
[30] Veljković VB, Avramović JM, Stamenković OS. Biodiesel production by ultra- Energy Sources Part A 2016;38(17):2610–6.
sound-assisted transesterification: state of the art and the perspectives. Renewable [63] Takase M, Chen Y, Liu H, Zhao T, Yang L, Wu X. Biodiesel production from non-
Sustainable Energy Rev 2012;16(2):1193–209. edible Silybum marianum oil using heterogeneous solid base catalyst under ultra-
[31] Martinez-Guerra E, Gude VG. Continuous and pulse sonication effects on trans- sonication. Ultrason Sonochem 2014;21(5):1752–62.
esterification of used vegetable oil. Energy Convers Manage 2015;96:268–76. [64] Ali SD, Javed IN, Rana UA, Nazar MF, Ahmed W, Junaid A, et al. Novel SrO-CaO
[32] Gogate PR, Tayal RK, Pandit AB. Cavitation: a technology on the horizon. Curr Sci mixed metal oxides catalyst for ultrasonic-assisted transesterification of Jatropha
2006;91(1):35–46. oil into biodiesel. Aust J Chem 2017;70(3):258.
[33] Choudhury HA, Srivastava P, Moholkar VS. Single-step ultrasonic synthesis of [65] Badday AS, Abdullah AZ, Lee K-T. Ultrasound-assisted transesterification of crude
biodiesel from crude Jatropha curcas oil. AlChE J 2014;60(5):1572–81. Jatropha oil using alumina-supported heteropolyacid catalyst. Appl Energy
[34] He B, Van Gerpen JH. Application of ultrasonication in transesterification pro- 2013;105:380–8.
cesses for biodiesel production. Biofuels 2012;3(4):479–88. [66] Badday AS, Abdullah AZ, Lee K-T. Optimization of biodiesel production process
[35] Hajinezhad A, Abedi S, Ghobadian B, Noorollahi Y. Biodiesel production from from Jatropha oil using supported heteropolyacid catalyst and assisted by ultra-
Norouzak (Salvia lerifolia) seeds as an indigenous source of bio fuel in Iran using sonic energy. Renewable Energy 2013;50:427–32.
ultrasound. Energy Convers Manage 2015;99:132–40. [67] Badday AS, Abdullah AZ, Lee K-T. Ultrasound-assisted transesterification of crude
[36] Mahamuni NN, Adewuyi YG. Optimization of the synthesis of biodiesel via ul- Jatropha oil using cesium doped heteropolyacid catalyst: interactions between
trasound-enhanced base-catalyzed transesterification of soybean oil using a mul- process variables. Energy 2013;60:283–91.
tifrequency ultrasonic reactor. Energy Fuels 2009;23(5):2757–66. [68] Maneechakr P, Samerjit J, Karnjanakom S. Ultrasonic-assisted biodiesel produc-
[37] Kumar D, Kumar G, Singh C. Ultrasonic-assisted transesterification of Jatropha tion from waste cooking oil over novel sulfonic functionalized carbon spheres
curcus oil using solid catalyst, Na/SiO 2. Ultrason Sonochem 2010;17(5):839–44. derived from cyclodextrin via one-step: a way to produce biodiesel at short reac-
[38] Gude VG, Grant GE. Biodiesel from waste cooking oils via direct sonication. Appl tion time. RSC Adv 2015;5(68):55252–61.
Energy 2013;109:135–44. [69] Nikseresht A, Daniyali A, Ali-Mohammadi M, Afzalinia A, Mirzaie A. Ultrasound-
[39] Hong IK, Jeon H, Lee SB. Effect of mixed alcohol reactants on ultrasonic alcoho- assisted biodiesel production by a novel composite of Fe(III)-based MOF and
lysis of canola oil. J Ind Eng Chem 2014;20(3):911–5. phosphotangestic acid as efficient and reusable catalyst. Ultrason Sonochem
[40] Ho WWS, Ng HK, Gan S. Advances in ultrasound-assisted transesterification for 2017;37:203–7.
biodiesel production. Appl Therm Eng 2016;100:553–63. [70] Gole VL, Gogate PR. Intensification of synthesis of biodiesel from nonedible oils
[41] Ejikeme P, Anyaogu I, Ejikeme C, Nwafor N, Egbuonu C, Ukogu K, et al. Catalysis using sonochemical reactors. Ind Eng Chem Res 2012;51(37):11866–74.
in biodiesel production by transesterification processes-an insight. J Chem [71] Bahadur S, Goyal P, Sudhakar K, Prakash Bijarniya J. A comparative study of ul-
2010;7(4):1120–32. trasonic and conventional methods of biodiesel production from mahua oil.
[42] Hara M. Environmentally benign production of biodiesel using heterogeneous Biofuels 2015;6(1–2):107–13.
catalysts. ChemSusChem 2009;2(2):129–35. [72] Worapun I, Pianthong K, Thaiyasuit P. Two-step biodiesel production from crude
[43] Encinar J, Gonzalez J, Rodriguez J, Tejedor A. Biodiesel fuels from vegetable oils: Jatropha curcas L. oil using ultrasonic irradiation assisted. J Oleo Sci
transesterification of Cynara cardunculus L. oils with ethanol. Energy Fuels 2012;61(4):165–72.
2002;16(2):443–50. [73] Saha R, Goud VV. Ultrasound assisted transesterification of high free fatty acids
[44] Leung DY, Wu X, Leung M. A review on biodiesel production using catalyzed karanja oil using heterogeneous base catalysts. Biomass Convers Biorefin
transesterification. Appl Energy 2010;87(4):1083–95. 2014;5(2):195–207.
[45] Na Y, Youyong S, Hua W, Wu Z, Xiaohong H. Experimental study on preparation of [74] Bahadur S, Goyal P, Sudhakar K. Ultrasonic assisted transesterification of neem oil
biodiesel through transesterification with ultrasonic assistant. Materials for for biodiesel production. Energy Sources Part A 2015;37(17):1921–7.
Renewable Energy & Environment (ICMREE), 2011 International Conference on. [75] Prakash Maran J, Priya B. Modeling of ultrasound assisted intensification of bio-
1. IEEE; 2011:284-7. diesel production from neem (Azadirachta indica) oil using response surface
[46] Kumar D, Kumar G, Singh C. Fast, easy ethanolysis of coconut oil for biodiesel methodology and artificial neural network. Fuel 2015;143:262–7.
production assisted by ultrasonication. Ultrason Sonochem 2010;17(3):555–9. [76] Yadav AK, Khan ME, Pal A, Dubey AM. Biodiesel production from Nerium oleander
[47] Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic (Thevetia peruviana) oil through conventional and ultrasonic irradiation methods.
catalysis for transesterification of high free fatty acid oil (waste cooking oil) to Energy Sources Part A 2016;38(23):3447–52.
biodiesel: a review. Biotechnol Adv 2010;28(4):500–18. [77] Sarve AN, Varma MN, Sonawane SS. Ultrasound assisted two-stage biodiesel
[48] Ramadhas AS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA synthesis from non-edible Schleichera triguga oil using heterogeneous catalyst:
rubber seed oil. Fuel 2005;84(4):335–40. Kinetics and thermodynamic analysis. Ultrason Sonochem 2016;29:288–98.
[49] Yin X, Ma H, You Q, Wang Z, Chang J. Comparison of four different enhancing [78] Reshad AS, Panjiara D, Tiwari P, Goud VV. Two-step process for production of
methods for preparing biodiesel through transesterification of sunflower oil. Appl methyl ester from rubber seed oil using barium hydroxide octahydrate catalyst:
Energy 2012;91(1):320–5. process optimization. J Cleaner Prod 2017;142:3490–9.
[50] Santos FF, Malveira JQ, Cruz MG, Fernandes FA. Production of biodiesel by ul- [79] Tian Y, Xiang J, Verni CC, Soh L. Fatty acid methyl ester production via ferric
trasound assisted esterification of Oreochromis niloticus oil. Fuel 2010;89(2):275–9. sulfate catalyzed interesterification. Biomass Bioenergy 2018;115:82–7.
[51] Parkar PA, Choudhary HA, Moholkar VS. Mechanistic and kinetic investigations in [80] dos Santos Ribeiro J, Celante D, Simões SS, Bassaco MM, da Silva C, de Castilhos F.
ultrasound assisted acid catalyzed biodiesel synthesis. Chem Eng J Efficiency of heterogeneous catalysts in interesterification reaction from macaw oil
2012;187:248–60. (Acrocomia aculeata) and methyl acetate. Fuel 2017;200:499–505.
[52] Choudhury HA, Malani RS, Moholkar VS. Acid catalyzed biodiesel synthesis from [81] Subhedar PB, Gogate PR. Ultrasound assisted intensification of biodiesel produc-
Jatropha oil: mechanistic aspects of ultrasonic intensification. Chem Eng J tion using enzymatic interesterification. Ultrason Sonochem 2016;29:67–75.
2013;231:262–72. [82] Tian Y, Xiang J, Verni CC, Soh L, Brahma B, Nath AJ, et al. Fatty acid methyl ester
[53] Mathiarasi R, Partha N. Optimization, kinetics and thermodynamic studies on oil production via ferric sulfate catalyzed interesterification. Biomass Bioenergy
extraction from Daturametel Linn oil seed for biodiesel production. Renewable 2018;115:1–266.
Energy 2016;96:583–90. [83] Tavares GR, Gonçalves JE, dos Santos WD, da Silva C. Enzymatic interesterifica-
[54] Choudhury HA, Goswami PP, Malani RS, Moholkar VS. Ultrasonic biodiesel tion of crambe oil assisted by ultrasound. Ind Crops Prod 2017;97:218–23.
synthesis from crude Jatropha curcas oil with heterogeneous base catalyst: me- [84] Mohan S, Pal A, Singh R. The production of semal oil methyl esters through a
chanistic insight and statistical optimization. Ultrason Sonochem combined process reactor. Energy Sources Part A 2017:1–8.
2014;21(3):1050–64. [85] Mohod AV, Subudhi AS, Gogate PR. Intensification of esterification of non edible
[55] Ali SD, Javed IN, Rana UA, Nazar MF, Ahmed W, Junaid A, et al. SrO-CaO mixed oil as sustainable feedstock using cavitational reactors. Ultrason Sonochem

906
S.X. Tan et al. Fuel 235 (2019) 886–907

2017;36:309–18. [116] Refaat A, El Sheltawy S. Comparing three options for biodiesel production from
[86] Kumar D, Ali A. Transesterification of low-quality triglycerides over a Zn/CaO waste vegetable oil. WIT Trans Ecol Environ 2008;109:133–40.
heterogeneous catalyst: kinetics and reusability studies. Energy Fuels [117] Kumar D, Kumar G, Johari R, Kumar P. Fast, easy ethanomethanolysis of Jatropha
2013;27(7):3758–68. curcus oil for biodiesel production due to the better solubility of oil with ethanol in
[87] Samani BH, Zareiforoush H, Lorigooini Z, Ghobadian B, Rostami S, Fayyazi E. reaction mixture assisted by ultrasonication. Ultrason Sonochem
Ultrasonic-assisted production of biodiesel from Pistacia atlantica Desf. oil. Fuel 2012;19(4):816–22.
2016;168:22–6. [118] Subhedar PB, Gogate PR. Alkaline and ultrasound assisted alkaline pretreatment
[88] Mootabadi H, Salamatinia B, Bhatia S, Abdullah AZ. Ultrasonic-assisted biodiesel for intensification of delignification process from sustainable raw-material.
production process from palm oil using alkaline earth metal oxides as the het- Ultrason Sonochem 2014;21(1):216–25.
erogeneous catalysts. Fuel 2010;89(8):1818–25. [119] Okitsu K, Maeda Y, Bandow H. Ultrasound assisted production of fatty acid methyl
[89] Fayyazi E, Ghobadian B, Najafi G, Hosseinzadeh B. Genetic algorithm approach to esters from transesterification of triglycerides with methanol in the presence of
optimize biodiesel production by ultrasonic system. Chem Prod Process Model KOH catalyst: optimization, mechanism and kinetics. Ultrason Sonochem
2014;9(1):59–70. 2014;21(2):467–71.
[90] Khosravi E, Shariati A, Nikou MRK. Instant biodiesel production from waste [120] Birla A, Singh B, Upadhyay S, Sharma Y. Kinetics studies of synthesis of biodiesel
cooking oil under industrial ultrasonic irradiation. Int J Oil Gas Coal Technol from waste frying oil using a heterogeneous catalyst derived from snail shell.
2016;11(3):308–17. Bioresour Technol 2012;106:95–100.
[91] Wang Y-Y, Đăng TnHp, Chen B-H, Lee,. D-J. Transesterification of triolein to [121] Vyas AP, Verma JL, Subrahmanyam N. Effects of molar ratio, alkali catalyst
biodiesel using sodium-loaded catalysts prepared from zeolites. Ind Eng Chem Res concentration and temperature on transesterification of Jatropha oil with me-
2012;51(30):9959–65. thanol under ultrasonic irradiation. Adv Chem Eng Sci 2011;1(02):45.
[92] Chen G, Shan R, Shi J, Yan B. Ultrasonic-assisted production of biodiesel from [122] Stavarache C, Vinatoru M, Maeda Y, Bandow H. Ultrasonically driven continuous
transesterification of palm oil over ostrich eggshell-derived CaO catalysts. process for vegetable oil transesterification. Ultrason Sonochem
Bioresour Technol 2014;171:428–32. 2007;14(4):413–7.
[93] Zhang F, Fang Z, Wang Y-T. Biodiesel production directly from oils with high acid [123] Thanh LT, Okitsu K, Sadanaga Y, Takenaka N, Maeda Y, Bandow H. A two-step
value by magnetic Na2SiO3@Fe3O4/C catalyst and ultrasound. Fuel continuous ultrasound assisted production of biodiesel fuel from waste cooking
2015;150:370–7. oils: a practical and economical approach to produce high quality biodiesel fuel.
[94] Gaikwad ND, Gogate PR. Synthesis and application of carbon based heterogeneous Bioresour Technol 2010;101(14):5394.
catalysts for ultrasound assisted biodiesel production. Green Process Synth [124] Mostafaei M, Ghobadian B, Barzegar M, Banakar A. Optimization of ultrasonic
2015;4(1):17–30. assisted continuous production of biodiesel using response surface methodology.
[95] Encinar J, Pardal A, Martínez G. Transesterification of rapeseed oil in subcritical Ultrason Sonochem 2015;27:54–61.
methanol conditions. Fuel Process Technol 2012;94(1):40–6. [125] Mohan S, Pal A, Singh R. Biodiesel production from Cedrus deodara oil in different
[96] Encinar J, González J, Pardal A. Transesterification of castor oil under ultrasonic types of ultrasonic reactors and energy analysis. Energy Sources Part A
irradiation conditions. Preliminary results. Fuel Process Technol 2012;103:9–15. 2016;38(24):3709–15.
[97] Noureddini H, Harkey D, Medikonduru V. A continuous process for the conversion [126] Yin X, Zhang X, Wan M, Duan X, You Q, Zhang J, et al. Intensification of biodiesel
of vegetable oils into methyl esters of fatty acids. J Am Oil Chem Soc production using dual-frequency counter-current pulsed ultrasound. Ultrason
1998;75(12):1775–83. Sonochem 2017;37:136–43.
[98] Hingu SM, Gogate PR, Rathod VK. Synthesis of biodiesel from waste cooking oil [127] Murillo G, Sun J, Ali SS, Yan Y, Bartocci P, He Y. Evaluation of the kinematic
using sonochemical reactors. Ultrason Sonochem 2010;17(5):827–32. viscosity in biodiesel production with waste vegetable oil, ultrasonic irradiation
[99] Kesgin C, Yücel S, Özçimen D, Terzioğlu P, Attar A. Transesterification of hazelnut and enzymatic catalysis: a comparative study in two-reactors. Fuel
oil by ultrasonic irradiation. Int J Green Energy 2014;13(3):328–33. 2018;227:448–56.
[100] Sebayang D, Agustian E, Praptijanto A. Transesterification of biodiesel from waste [128] Sajjadi B, Asaithambi P, Abdul Aziz AR, Ibrahim S. Mathematical analysis of the
cooking oil using ultrasonic technique. 2010. effects of operating conditions and rheological behaviour of reaction medium on
[101] Guldhe A, Singh B, Rawat I, Bux F. Synthesis of biodiesel from Scenedesmus sp. by biodiesel synthesis under ultrasound irradiation. Fuel 2016;184:637–47.
microwave and ultrasound assisted in situ transesterification using tungstated [129] Mazanov SV, Gabitova AR, Usmanov RA, Gumerov FM, Labidi S, Amar MB, et al.
zirconia as a solid acid catalyst. Chem Eng Res Des 2014;92(8):1503–11. Continuous production of biodiesel from rapeseed oil by ultrasonic assist trans-
[102] Ramachandran K, Al-Zuhair S, Fong C, Gak C. Kinetic study on hydrolysis of oils esterification in supercritical ethanol. J Supercrit Fluids 2016;118:107–18.
by lipase with ultrasonic emulsification. Biochem Eng J 2006;32(1):19–24. [130] Biktashev SA, Usmanov R, Gabitov R, Gazizov R, Gumerov F, Gabitov F, et al.
[103] Freedman B, Pryde E, Mounts T. Variables affecting the yields of fatty esters from Transesterification of rapeseed and palm oils in supercritical methanol and
transesterified vegetable oils. J Am Oil Chem Soc 1984;61(10):1638–43. ethanol. Biomass Bioenergy 2011;35(7):2999–3011.
[104] Eevera T, Rajendran K, Saradha S. Biodiesel production process optimization and [131] Cintas P, Mantegna S, Gaudino EC, Cravotto G. A new pilot flow reactor for high-
characterization to assess the suitability of the product for varied environmental intensity ultrasound irradiation. Application to the synthesis of biodiesel. Ultrason
conditions. Renewable Energy 2009;34(3):762–5. Sonochem 2010;17(6):985–9.
[105] Chand P, Reddy CV, Verkade JG, Grewell D. Enhancing biodiesel production from [132] Widayat, Satriadi H, Choirudin F, Fitriana A, Kiono BFT. Biodiesel production with
soybean oil using ultrasonics. Am Soc Agric Biol Eng 2008. continuous processing and direct ultrasonic assisted. 2016 International
[106] Salamatinia B, Mootabadi H, Hashemizadeh I, Abdullah AZ. Intensification of Conference on Sustainable Energy Engineering and Application (ICSEEA). IEEE;
biodiesel production from vegetable oils using ultrasonic-assisted process: opti- 2016:122-6.
mization and kinetic. Chem Eng Process Process Intensif 2013;73:135–43. [133] Banković-Ilić IB, Stamenković OS, Veljković VB. Biodiesel production from non-
[107] Takase M, Feng W, Wang W, Gu X, Zhu Y, Li T, et al. Silybum marianum oil as a edible plant oils. Renewable Sustainable Energy Rev 2012;16(6):3621–47.
new potential non-edible feedstock for biodiesel: a comparison of its production [134] Shuit SH, Lee KT, Kamaruddin AH, Yusup S. Reactive extraction and in situ es-
using conventional and ultrasonic assisted method. Fuel Process Technol terification of Jatropha curcas L. seeds for the production of biodiesel. Fuel
2014;123:19–26. 2010;89(2):527–30.
[108] Maghami M, Sadrameli S, Ghobadian B. Production of biodiesel from fishmeal [135] Haas MJ, Scott KM. Moisture removal substantially improves the efficiency of in
plant waste oil using ultrasonic and conventional methods. Appl Therm Eng situ biodiesel production from soybeans. J Am Oil Chem Soc 2007;84(2):197–204.
2015;75:575–9. [136] Zahari MSM, Ismail S, Ibrahim MZ, Lam SS, Mat R. Ultrasonicated Jatropha curcas
[109] Koutsouki A, Tegou E, Kontakos S, Kontominas M, Pomonis P, Manos G. In situ seed residual as potential biofuel feedstock. Jurnal Teknologi 2015;77(1).
transesterification of Cynara cardunculus L. seed oil via direct ultrasonication for [137] Chadha P, Arora AK, Prakash S, Jha MK, Puri SK, Tuli DK, et al. Ultrasonic assisted
the production of biodiesel. Fuel Process Technol 2015;134:122–9. in situ transesterification of Jatropha seed to biodiesel. J Sci Ind Res
[110] Krishnaiah G, Pasnoori S, Santhoshi P, Rajanna K, Rao YR, Patnaik KR. Ultrasonic 2012;71(4):290.
and microwave effects on crystalline Mn (II) carbonate catalyzed biodiesel pro- [138] Zahari MSM, Ismail S, Ibrahim MZ, Lam SS, Mat R. Study of enhanced reactive
duction using watermelon (Citrullus vulgaris) seed oil and alcohol (fibrous flesh) as extraction process using ultrasonication for Jatropha curcas seed. Appl Mech Mater
exclusive green feedstock. Biofuels 2016:1–7. 2014;699.
[111] Boffito D, Galli F, Pirola C, Bianchi C, Patience G. Ultrasonic free fatty acids es- [139] Kumar G. Ultrasonic-assisted reactive-extraction is a fast and easy method for
terification in tobacco and canola oil. Ultrason Sonochem 2014;21(6):1969–75. biodiesel production from Jatropha curcas oilseeds. Ultrason Sonochem
[112] Ragavan SN, Roy DV. Transesterification of rubber seed oil by sonication tech- 2017;37:634–9.
nique for the production of methyl esters. Biomass Convers Biorefin [140] Jadhav AJ, Holkar CR, Goswami AD, Pandit AB, Pinjari DV. Acoustic cavitation as
2011;1(2):105–10. a novel approach for extraction of oil from waste date seeds. ACS Sustainable
[113] Zou H, Lei M. Optimum process and kinetic study of Jatropha curcas oil pre-es- Chem Eng 2016;4(8):4256–63.
terification in ultrasonical field. J Taiwan Inst Chem Eng 2012;43(5):730–5. [141] Shirsath S, Sonawane S, Gogate P. Intensification of extraction of natural products
[114] Brasil AN, Oliveira LS, Franca AS. Circulation flow reactor with ultrasound irra- using ultrasonic irradiations—a review of current status. Chem Eng Process
diation for the transesterification of vegetable oils. Renewable Energy Process Intensif 2012;53:10–23.
2015;83:1059–65. [142] Poosumas J, Ngaosuwan K, Quitain AT, Assabumrungrat S. Role of ultrasonic ir-
[115] Refaat A, Attia N, Sibak HA, El Sheltawy S, ElDiwani G. Production optimization radiation on transesterification of palm oil using calcium oxide as a solid base
and quality assessment of biodiesel from waste vegetable oil. Int J Environ Sci catalyst. Energy Convers Manage 2016;120:62–70.
Technol 2008;5(1):75–82.

907

You might also like