Sunteți pe pagina 1din 10

For reprint orders, please contact reprints@future-science.

com

Review

Application of ultrasonication in transesterification


processes for biodiesel production
Biofuels (2012) 3(4), 479488

Brian He* & Jon H Van Gerpen


In biodiesel production, adequate mixing is required to create sufficient contact between the vegetable oil
or animal fat and alcohol, especially at the beginning of the reaction. Application of ultrasonication provides
sufficient mixing and energy so that the transesterification can proceed at a faster rate due to two effects.
First, ultrasonic cavitation and microbubble formation, which are caused by the ultrasonic energy introduced
by the sonotrode, greatly improve the interfacial contact between the immiscible methanol and plant oil/
animal fat mixture, thus increasing the reaction rate. Second, the formation and bursting of microbubbles
caused by ultrasonic cavitation intensifies the local energy transfer and energizes the reactant molecules,
thus enhancing the overall reaction rate. The other possible beneficial aspect of ultrasonication may be
ultrasonic energy-induced free radical formation, which initiates chain reactions, as has been observed in
other organic systems, although it is not fully understood in transesterification yet.

Biodiesel production involves the transesterification of


plant oils or animal fats with an alcohol to produce alkyl
esters. This reaction is reversible and proceeds stepwise.
To enhance the completeness of this reaction, it is common to add 1.5- to two-times the stoichiometric amount
of alcohol. With methanol as the alcohol, this exceeds
the solubility of methanol in the oil or fat, so droplets
are formed from the methanol that exceeds the solubility limit [1] . These droplets represent a polar phase
that captures most of the dissociated sodium methoxide catalyst. Although the solubility of oil or fat in the
methanol phase is low (12%), the reaction occurs as
the triglycerides diffuse through the droplet interface
where they come into contact with methanol and catalyst. The reaction rate is widely considered to be limited by this diffusion process [1] . Later in the reaction,
glycerol accumulates in this polar phase, maintaining
the isolation of the catalyst and, thus, hindering the
diffusion processes, which, in turn, limits the reaction
rate. In conventional processes, this problem of limited
methanol solubility in vegetable oils or animal fat is
typically overcome by intensifying mechanical mixing

or simply allowing extended time to let the product


esters build up, which helps by improving the methanol
solubility in oils or fats. As an attempt to address this
issue in an alternative way, application of ultrasonication
in this system is studied by many researchers.
Ultrasonic excitation helps increase the liquidliquid
interfacial area through emulsification, which is important for the formation of vapor bubbles and cavitation bubbles in viscous liquids, such as plant oils and
animal fats. Vapor bubbles within the liquid, such as
methanol bubbles generated mechanically or ultrasonically in liquid oils or fats, oscillate and move with the
steady currents in the bulk liquid caused by the highfrequency acoustic oscillations or acoustic streaming.
This acoustic streaming may also be caused by differences in the ultrasound velocity, which is a property
of the materials, traveling in alcohol (e.g., 1182m/s
in ethanol [2]) and in vegetable oils (e.g., 1430m/s in
soybean oil [3]). Simultaneously, the cavitation bubbling action pushes the liquid towards the interfacial
surfaces of the vapor bubbles, where it interacts with the
bubbles. This phenomenon enhances the mass transfer

Department of Biological & Agricultural Engineering, University of Idaho, PO Box442060, Moscow, ID 83844-2060, USA
*Author for correspondence: Tel.: +1 208 885 7435; Fax: +1 208 885 8923; E-mail: bhe@uidaho.edu

future science group

10.4155/BFS.12.35 2012 Future Science Ltd

ISSN 1759-7269

479

Review He & Van Gerpen

Key terms
Ultrasonication: In the scope of
discussion in this article, ultrasonication
is an application of ultrasound to
transmit energy into a fluid (typically a
liquid) in the form of high frequency
(>20kHz) ultrasound wave oscillation for
the purpose of, for example, material
structure destruction, intensive mixing
and/or chemical reaction induction.
Ultrasonication is typically generated
through an ultrasound transducer or
sonotrode.
Sonotrode: Generator or transducer of
ultrasonic waves. The commonly seen
sonotrodes include the ones that are
based on magnetostrictive and
piezoelectric properties of special
materials. The shape of sonotrodes can
be circular, polygonal, planar and
tubular for different applications.
Sonochemical reactors: Chemical
reactors for the study of sonochemistry
of various systems. The design of
sonochemical reactors can be batch or
continuous, tubular or stirred tanks,
single pass or circulated, depending on
the application requirements.

across the interfaces of the bubbles


and, thus, accelerates the chem
ical reaction rates under diffusionlimited conditions such as the early
stage of transesterification of oils
and fats in biodiesel production.
Meanwhile, ultrasonic cavitation
bubbles are much smaller in size
and the total number per volume is
much more stable than those generated mechanically [4] . The very large
number of fine cavitation bubbles
greatly increases the interfacial area
available for chemical reactions.
Ultrasonic cavitation is also claimed
to produce a microenvironment of
very high temperature and pressure [5,6] , which may create highly
active reaction intermediates and
lead to faster transesterification
reaction rates.

Sonotrodes
The ultrasonic generators or transSonochemistry: Study of chemical
ducers are also commonly referred
changes, in a reactive substance or
to as sonotrodes. Based on the
mixture of substances, which are
mechanism of how the ultrasound
induced by the application of
high-frequency sounds, typically
is generated, Mason has classified
ultrasound, of high power (or energy) to
the ultrasonic sonotrodes into three
the system.
types [7] . The first type of sonotrode
is the liquid whistle, in which a high
pressure liquid jet, realized by a high-pressure pump and
a converging orifice, causes a thin metal blade to vibrate
at high frequency, thus creating intensive local eddies
and violent liquid flows. This type of sonotrode is very
useful for applications of liquid mixing and homogenization. The second type is the magnetostrictive sonotrode,
which uses materials of construction that have the property of magnetostriction; that is, when placed in a magnetic field, these materials change their size depending
on the application of a magnetic field. Such materials
include nickel alloys, cobalt/iron (Co/Fe) alloys and
aluminum/iron/chromium (Al/Fe/Cr) alloys. A drawback of this type of sonotrode is that its upper frequency
is limited to approximately 70kHz. The third type is
the piezoelectric sonotrode, which is the most commonly used due to its wide ultrasonic frequency range.
The principle of the piezoelectric sonotrode is the energy
conversion between electrical pulses and mechanical
vibrations through a polarized material (e.g., a piezoelectric ceramic) with two electrodes attached to it, in
which the electrical energy and acoustic energy are transformed back and forth [8] . Piezoelectric sonotrodes can
be constructed in different power sizes and frequency

480

Biofuels (2012) 3(4)

ranges, and are used in situations that require powerful


ultrasonication.
Ultrasonic sonotrodes for sonochemical applications are constructed in the forms of ultrasonic baths,
ultrasonic horns or cup horns, planar transducers and
vibrating/cavitating tubes [9] . The means of applying
sonication includes both direct and indirect techniques;
sonochemical reactors can, for example, be designed
with circular and polygonal shapes, planar and tubular sonotrodes, for batch and continuous systems [9] .
The scale-up of sonochemical reactors, especially the
design of sonotrodes for large reactors, however, has
many engineering challenges to be considered and
overcome [7,10] .
Ultrasonic phenomena in reactive solutions
Ultrasound refers to frequencies above the human hearing range, which is approximately 20kHz [8] . Although
ultrasound can be extended to a much wider frequency
range, the low frequencies (~2060kHz) are most commonly used [4] . Due to its unique properties, ultrasound
has found many applications in medical and industrial
fields. The industrial and experimental applications
include enhancing chemical reactions, crystallization,
cell disruption and solids dispersion in liquidliquid or
liquidsolid systems. Ultrasonication is also a useful
tool for liquid mixing or emulsification. Most industrial
ultrasonic cleaning devices are operated at 40kHz. If
the purpose of ultrasonication is to damage, such as
the devices used in laboratories for biological cell disruption, lower frequencies (e.g., 20kHz) are typically
used. When applied in homogeneous media, ultrasound
transfers a substantial amount of energy through rapid
but short displacement movements. In addition to the
application time, the major factor that decides the total
energy output, expressed by the transducer displacement, velocity and acceleration, is the power of the
ultrasonic transducer or sonotrode [4] .
The effect of ultrasonic waves on liquids has been
discussed by Suslick [11] . When ultrasonic waves are
passed through a mixture of immiscible liquids, such
as vegetable oil and methanol, acoustic pressure causes
acoustic cavitation and fine emulsions in the liquids.
The result of such ultrasonic cavitation and emulsion
formation is extremely fine cavities or microbubbles,
with diameters in the range of micrometers. These
microbubbles provide tremendously large interfacial
areas. To achieve such an effect, however, a high enough
energy density of ultrasound input (i.e., the ultrasonic
energy per unit surface area to which it is applied) is
a necessity.
Cavitation can also occur from various causes, such
as a turbulent fluid flow pattern (i.e., hydrodynamic
cavitation). However, the sizes of the cavities created by

future science group

Application of ultrasonication in transesterification processes for biodiesel production Review

ultrasonication are much smaller compared with those


by turbulent fluid flow patterns. In ultrasonic cavitation, the ultrasound is composed of waves of expansion
(negative pressure) and compression (positive pressure)
[4] . These expansion and compression waves cause preexisting bubbles to grow and then collapse. When the
microbubbles undergo implosive collapse, smaller cavities are formed, which then serve as the nucleation sites
for the next cycle. The resonance size of these microbubbles, which is approximately 170m at 20kHz and
3.3m at 1000kHz [11] , before implosive collapse, is
dependent on the ultrasonic frequency and the acoustic power. Other parameters that also affect ultrasonic
cavitation and the sizes of the microbubbles include the
presence of dissolved gases, the liquid working medium,
and the ambient temperature and pressure [9] .
The formation of cavities and their dynamics of
growth and collapse depend on the system under
consideration. The mechanism of cavity formation
in liquidliquid systems is quite different from those
on liquidsolid interfaces. The implosive collapse of
microbubbles during ultrasonic cavitation is associated
with high local energy exchanges. The predicted locally
effective temperatures are thousands of degrees. Suslick
and co-workers have claimed, based on their experimental determination, that the effective ultrasonic
cavitation temperature in the gas phase of the cavities
was as high as 5000K, and in the liquid phase it was
1900K [5,6] , although this high temperature may not
considerably increase the bulk liquid temperature due
to the low energy density carried by these microbubbles.
The internal pressures of the collapsing microbubbles
were theoretically predicted to be astonishingly high;
as much as approximately 20003000atm for organic
solvents [12] . Such high energy generation and exchange
would be expected to have a tremendous effect on the
system, both physically (e.g., local effective temperature spikes) and chemically (e.g., enhanced chemical
reactions). The high energy generated by ultrasonication at the microlevel could energize organic chemical
substances to form free radicals. These highly active
radicals react with other chemicals and initiate chain
reactions in many organic systems (discussed in the
next section). The study of ultrasound-induced chemical reactions and related chemistry via radical formation
and other pathways has become a scientific field, known
as sonochemistry [9,13] .
It has been observed that the constructive effect of
ultrasonication on chemical reactions does not exist proportionally for all ultrasonic power ranges; the effect
increases as the ultrasonic power increases over a certain
range. When the ultrasonic power is further increased
above this level, the positive effect of ultrasonication
decreases suddenly. Such a phenomenon is known as

future science group

decoupling [7,14] . Therefore, the optimal level of ultrasonic power application is dependent on the chemical
system of interest.
Ultrasonication techniques have been used in many
fields, including the control of airborne contamination,
remediation of contaminated soils, removal of biological contamination from water and treatment of sewage
sludge [15] . Applications have been expanded to many
other industrial fields; new developments and improvements are being made constantly. The complete understanding of the sonochemistry and the design of sonochemical reactors, however, as stated by Thompson and
Doraiswamy, requires the mutual and continued effort
of both scientists and engineers [9] .
In summary, the mechanisms of ultrasonication
that possibly affect chemical reactions are threefold:
the ultrasonic oscillating pressures cause rapid movement of liquid clusters at the molecular level, which
thus greatly increases mixing of the microenvironment;
ultrasonication creates acoustic cavitation and leads to
the formation of microbubbles in liquids the creation
and collapse of the microbubbles generates high pressure
and temperature locally at the molecular cluster levels, which can provide the energy needed for chemical
reactions; and the combined effect of the previous two
phenomena may pass tremendous energy to molecules
and energize them into free radicals, which lead to chain
reactions and subsequent products, although this last
point is not yet fully understood.
Effect of ultrasonication on chemical reactions
Early research has shown that application of ultrasonication in aqueous systems causes water molecules to
dissociate and form H2O2 and H2 [16] . Since the 1970s,
research has been conducted on ultrasonication-induced
free radical reactions in organic liquids [17,18] . The radical formation is fundamentally dependent on the effective ultrasonic cavitation temperature and the internal
pressures at the moment of cavity collapse, which, in
turn, depend on the properties of the given chemical
compounds.
Research has shown that ultrasonication of alkanes
can yield various decomposed products, from hydrogen to short-chain hydrocarbons, a very similar phenomenon to that of thermochemical decomposition
or pyrolysis of organic compounds. In work by Suslick
etal., various decomposed products were obtained when
saturated n-decane (C10H22) was exposed to ultrasonic
irradiations of approximately 100W/cm2 at 20kHz [19] .
These products included H2, CH4, acetylene (C2H2),
ethylene (C2H4) and propylene (C3H6). Suslick etal.
suggested that the mechanism of n-decane decomposition by ultrasonic irradiation at room temperature is
similar to that of alkane pyrolysis by the Rice radical

www.future-science.com

481

Review He & Van Gerpen

chain mechanism [19] . The Rice radical chain mech


anism is a theory that describes the chain reactions
occurring in organic compound thermal decomp
osition. It was initially proposed by Rice [2026] , and
was further developed by other researchers [27] . It is now
widely accepted in the field of organic chemistry.
As early as 1942, Miyagawa reported that application
of ultrasonication increased the reaction rates of acetate hydrolysis and fat saponification [28] . In the 1990s,
intensive studies were conducted by many researchers
on ultrasonication application in organic reaction systems. For example, Lin and Yen applied the ultrasonication technique in their research on upgrading asphalt
products and demonstrated that the improved effectiveness of ultrasonication on their system was based on free
radical involved reactions [29] . Another example is the
esterification of carboxylic acids, in which ultrasonication greatly enhanced the reactions and provided a
convenient method for esterification with homogenous
sulfuric acid catalyst [30] . In the review by Thompson
and Doraiswamy, more than 30 research reports of
sonochemical processes in organic synthesis, organometallic chemistry, polymerization and biotechnology have been summarized [9] . Based on the reaction
mechanism, the effectiveness of ultrasonication was
mainly due to the ultrasound initiating the reactions,
enhancing the rate of reaction and altering the reaction pathway. However, there were also reports that
showed the ultrasonication does not have any effect on
some chemical reactions, such as those following ionic
reaction mechanisms [31,32] .
In recent years, ultrasonication has been used to
enhance the transesterification of plant oils and animal
fats for biodiesel production. In conventional basecatalyzed transesterification of plant oils or animal
fats with a short-chain alcohol, typically methanol, the
responsible catalytic functional species is the ionized
methoxide (also named methylate) group, or CH3O-,
which is formed by reacting methanol with a strong base
(alternatively, it can be obtained by reacting metallic
potassium or sodium with anhydrous methanol). If the
base is potassium hydroxide, the formation of potassium
methoxide is:
CH3 OH + KOH D CH3 OK + H 2 O
Equation 1

Potassium methoxide partially dissociates into


methoxide ion in polar solvents:
CH3 OK D CH3 O- + K+
Potassium
methoxide

Methoxide
ion

Potassium
ion

Equation 2

482

Biofuels (2012) 3(4)

Then, to the best of our knowledge, the mechanism


for transesterification of triglycerides can be described
as follows. The methoxide ion attacks the carboxyl
groups at the CO double bonds of the triglycerides
to form transesterified products (i.e., alkyl esters) via
a couple of unstable intermediate complex species, and
regenerates the methoxide to start the next round of
reaction (Figure1) .
Similar processes will happen to diglycerides to yield
monoglycerides, and monoglycerides to glycerol in the
transesterification of plant oils and animal fats for
biodiesel production.
As mentioned earlier, because methanol and plant
oils or animal fats are immiscible, it is generally
accepted that the transesterification of plant oils or
animal fats with methanol is a mass transfer-limited
reaction. The improvement of the reaction through
intensive mechanical agitation supports this assumption. Therefore, it is reasonable to expect that application of ultrasonication will introduce large quantities of microbubbles via acoustic cavitation and, thus,
greatly increase the interfacial area between the plant
oils or animal fats and methanol, which leads to a
faster transesterification reaction [33] . According to the
report by Wu etal., ultrasonication produced average
droplet sizes that were 42% smaller than those generated using standard impellers in a methanol/soybean
oil model system [34] . This enhancement effect was
also observed in transesterification with other alcohols
including ethanol, propanol, butanol, hexanol, octanol and decanol, although limited effect was observed
on secondary alcohols [35,36] . Efforts have been made
to explain the actual mechanism using an ultrasonic
cavitation model that combines the mass transfer via
diffusion and the dynamics of bubble formation. The
conclusion was that the fundamental physical effect
behind the ultrasonication is the radial motion of the
ultrasonic cavitation bubbles [37] . The ultrasonication
not only provides the energy to generate the cavitation
bubbles but also the required activation energy for initiating the transesterification [38] . However, this effect
was not significant unless the sonotrode was placed at
or near the interface of the immiscible liquidliquid
mixture [3941] . It was also shown that when combined
with microwave radiation, the effect of ultrasonicationinduced mixing is more significant [42] . The total reaction time (1min ultrasonic mixing and 2min closed
microwave irradiation) was much shorter than has been
reported for complete transesterification (conversion
rate of 98.5% on a molar basis).
To further explore the mechanism that supports the
ultrasonication effect on improved transesterification,
Kalva etal. developed a different model and designed
a series of experiments to verify their theory [43] . The

future science group

Application of ultrasonication in transesterification processes for biodiesel production Review

bubble-dynamics model includes the considerations of


bubble dynamics, mass and heat transfer across the bubbles, overall energy balance and the velocity field generated by the radial motion of the microbubbles. After
simulation via the model and analysis of the experimental results, Kalva etal. concluded that the interfacial area
is increased by the fine emulsion formed between the oil
and methanol, and that it is this physical mechanism,
not any chemical effect, which is responsible for the
beneficial action of ultrasonication. The authors also
concluded that the production of radical species does
not play any role in the induction or acceleration of
the reaction by participating species [43] . The reason
was said to be that the moderate intensity of ultrasonic
cavitation in methanol does not generate temperatures
and pressures high enough to warrant production of
radicals. This conclusion is quite contradictory to those
made by previous research groups in other chemical systems [19,28,29,44,45] , especially on the local temperature,
which was claimed to be very high under ultrasonic
cavitation [5,44] .
In a conventional, mechanically agitated system,
the rate-limiting step is said to be the transesterification of di- to mono-glycerides [46,47] . According to
Stavarache etal., the limiting step in transesterification
shifts from the transesterification of monoglycerides
to esters in ultrasonicated systems [48] . The evidence
is that monoglycerides were detected in higher concentration than diglycerides, indicating a slow transesterification of monoglycerides to esters. Additionally, the
transesterification occurs first with saturated fatty acids
mostly at the beginning of the reaction, while the transesterification of unsaturated fatty acids takes place at a
slower pace [48] .
Ultrasonication-assisted transesterification for
biodiesel production
The application of ultrasonication to enhance the transesterification of vegetable oil to biodiesel was reported as
early as in 2003 [49] . Maeda etal. were granted patents
on their ultrasonication biodiesel production process
after they communicated their research in 2003 [101103] ,
and afterwards continued looking into this field [5053] .
The findings of reduced reaction time, low catalyst
application rate and less excess methanol requirement
make this unique technology very attractive. Since then,
researchers have extensively explored the advantages of
ultrasonication and its feasibility for advancing biodiesel
production technology. More than 70 research reports
have been published from 2003 todate.
Research has shown that the application of ultrasonication is effective and can be applied in the production of biodiesel from various feedstocks of virgin oils
[52,5459] . Ultrasonic processing provides similar yields of

future science group

O
CH2

R2OOC CH

CH2

CH2

CH3 O

CH2

R2OOC CH
CH2
(II)

COOR3

Alkyl ester
(biodiesel)

(II)
O
+
+ H O CH3
COOR3

CH2 O
R2OOC CH
CH2

Methanol

C R1
O
CH3

COOR3

(I)
CH2

R2OOC CH

CH3

CH2

(I)
CH2

R1

R1

CH3
COOR3

Methoxide
O

C
O

CH2

R2OOC CH

R2OOC CH

COOR3

Triglyceride

CH2

R1

C
+

Di-glyceride

+ CH3
COOR3

Methoxide

Figure1. Mechanism for transesterification of triglycerides catalyzed by


methoxide.

biodiesel with a much shorter reaction time as compared


with the conventional procedure, and it can achieve far
more complete reactions if the same reaction time is
employed. Ultrasonication shows similar enhancement
effects on the conversion of waste vegetable oils and
yellow greases [6062] , and other high fatty acids feedstocks [6365] to biodiesel, as well as with ethanol as the
transesterifying alcohol [50,66] .
Although the reported ultrasonic effects vary widely,
the major benefit observed from ultrasonication is the
much shorter reaction time. Once applied, the ultrasonication process can reduce the transesterification
process to tens of minutes [38,45,6770] and even as low as
1min [71] . It is concluded by Stavarache etal. [48] that
the transesterification takes place in the first 310min
when ultrasonication is used with a base catalyst. The
much shorter reaction time was also observed when a
heterogeneous catalyst was used [68] . Meanwhile, the
application rate of catalyst can be reduced by approximately two- to three-times due to the use of ultrasonication [45] . In an investigation of ultrasonication applying
modes, Chand etal. studied the effect of pulse mode
(5son/25soff ) versus continuous mode (for 15s)
ultrasonication. The experiments with ultrasonic treatment resulted in a 96% conversion to biodiesel in less
than 90s in the pulse mode. In the continuous mode
treatment, a biodiesel yield of 86% was achieved in
15s of ultrasonication, which is approximately equivalent to that obtained from the control experiment (i.e.,
without ultrasonication) after 3045min of reaction

www.future-science.com

483

Review He & Van Gerpen

(8386%) [72] . Others have concluded that continuous ultrasonication provides better emulsification,
and pulse ultrasonication was found to be too mild
to deliver adequate mixing to the immiscible mixture
[62,70] . The advantage of reduced reaction time was also
observed in a continuous flow system, where a reaction
time of 20min was considered adequate despite the
relatively lower conversion rates of approximately 90%
with 7.8 and 19l/h throughputs [73] .
Teixeira etal. conducted a quick comparison of conventional and ultrasound-assisted transesterification
of beef tallow in a laboratory set-up. Under similar
experimental conditions (i.e., a working sample of
200g beef tallow with 6:1 methanol-to-tallow molar
ratio, 0.5%wt potassium hydroxide catalyst and a reaction temperature of 60C), the ultrasound-assisted
process resulted in a comparable conversion rate and
similar fuel quality to that of a conventional process
[71] . However, the reaction time was reduced to 70s
under an ultrasonication power input of 400 W at
24kHz (which equals ~14J per gram of animal fat)
as compared with 1h for the conventional process.
The authors have claimed that this fast reaction could
be due to the high-speed mixing and mass transfer
between the less-miscible methanol and triglycerides
under sonication, as well as to the formation of a microemulsion resulting from the collapse of ultrasonic cavitation bubbles and ultrasonic jets; this claim is consistent with other reports. Most researchers believe that
the effect of ultrasonication on enhancing transesterification lies mainly in intensifying the mixing of the
immiscible methanol and triglyceride phases, especially
at the beginning of the reaction [33,37,42] . The mixing
enhancement is largely due to the collapse of ultrasonic
cavitation bubbles and the reduced droplet sizes of low
boiling temperature methanol in less-miscible triglycerides [34] . Another observation was that the ultrasonication does not affect the ester profiles as compared
with those from conventional, non-ultrasonication
procedures if potassium is used as the catalyst, indicating that ultrasonication does not decompose fatty
acids to other chemicals (e.g.,radicals) as observed in
other organic systems [19] . On the other hand, results
of improved fatty acid separation were observed under
ultrasonication when sodium hydroxide was used as
the catalyst [74] .
It was demonstrated that ultrasonication also
enhances heterogeneous catalyst activities in trans
esterification of plant oils [37,68,7577] . With heterogeneous catalysts, nanoemulsions (a micelle size of 5.1nm
was observed) can be created by using ultrasonication
in a transesterification system [78] , which would greatly
increase the surface areas of heterogeneous catalysts for
active catalytic sites and enhance the reaction. When

484

Biofuels (2012) 3(4)

calcium oxide was used as a catalyst, the conversion of


palm oil was only 5.5% after 60min reaction under
mechanical agitation. When ultrasonication was
applied (100W at 20kHz per 250ml reacting mixture, which was approximately equivalent to 1820kJ
per l of reacting mixture), the conversion increased to
75% in the same time frame [37] ; this observation agrees
with a previous study for the same catalyst [79] . The
catalysts after ultrasonication maintain their physical
and chemical properties, and can be reused without
significant deactivation [76] . Other oxide catalysts of
alkaline earth metals, such as barium oxide and strontium oxide, demonstrated similar increases in activity
under ultrasonication [76] .
Ultrasonication also showed effectiveness for transesterification catalyzed by enzymes. When used under
two different ultrasonication conditions, Novozym
435 (Candida antarctica lipase B immobilized on polyacrylic resin) showed significantly enhanced activity.
Under optimized conditions (50% of ultrasonic power
or approximately equivalent to 7200kJ energy input
per liter of reacting mixture, 50rpm vibration, 6%
Novozym 435 and 40C), 96% yield of biodiesel from
soy bean oil was achieved in 4h and the enzyme showed
no obvious activity loss [80] . Kumar etal. investigated
ultrasonication-assisted, lipase-catalyzed transesterification of jatropha oil and showed similar effectiveness
[69] . With a lower methanol-to-oil molar ratio of 4:1,
it took 30min to achieve a conversion rate of 85%
(with ultrasonication of 100W at 24kHz per 25ml
reacting mixture, which was equivalent to ~9100kJ
energy input per l of reacting mixture). It was seen
that the reaction time was dramatically shorter from
that used in conventional enzymatic transesterification
processes (usually several hours), however, with a very
high energy cost. Evidentially, the locally effective high
temperatures predicted by Suslick and co-workers [5,6]
do not show a significant effect on enzyme deactivation
as observed by Kumar etal. [69] .
It seems apparent that the higher the power input,
the faster the reaction rate of transesterification.
However, this claim is valid only to a certain point if
ultrasonication is used. Kumar etal. reported that when
a 200W/24kHz ultrasonication transducer was used
in a 25-ml reaction device, the enhancement effect of
ultrasonication levels were at 50% of the maximum
amplitude, or 100W per 25ml of reacting mixture
[68] . The explanation was that when a large amount of
ultrasonic power enters a system, a much larger quantity of ultrasonic cavitation bubbles is generated in the
solution. Excessive bubbles likely merge and form larger
and more stable bubbles and, thus, create a barrier to
acoustic energy transfer [81] . This is the phenomenon
referred to earlier as decoupling [7,14] .

future science group

Application of ultrasonication in transesterification processes for biodiesel production Review

It is generally understood that the ultrasonic frequency affects the size of the cavitation bubbles. The
higher the frequency, the finer the cavitation bubbles,
thus the larger the surface area between the alcohol and
the oils [9,82] . This leads to a greater enhancement of
the transesterification rate. Mahamuni and Adewuyi
studied the application of high-frequency ultrasonication in biodiesel production [82,83] . Among the three
frequencies of 581, 611 and 1300kHz, 611kHz was
claimed to perform well at all three different power
levels (58, 91 and 183W per 450ml reacting mixture), although a maximum biodiesel yield close to
95% was achieved at 581 kHz [83] . It was also found
in their optimization study on the effects of ultrasound
frequency and power combinations that, to reach the
same level conversion of 82%, the optimum combination of frequency and power was 611kHz and 139W.
This was the lowest energy input of the nine cases
studied, corresponding to 273J/g of ultrasonic energy
(~240kJ/l biodiesel) as compared with much higher
energy input under the other operating conditions [82] .
From an engineering perspective, the ultrasonic
power consumption is a key piece of data for the
ultrasound-assisted transesterification of plant oils
and animal fats. Unfortunately, most studies do not
report the ultrasonic power input when reporting
research on transesterification. The level of ultrasonication varies widely and is frequently not presented in
a quantitative manner. In a recently reported study,
the ultrasonic power input into the reaction system
was included in the account. When the sonotrode
power was set to 450W and the reaction time was
30min, the ultrasonic energy densities were 6337.8,
5959.7 and 4926.5 kJ/l for the reaction systems of
canola, soybean and corn oil, respectively, due to the
different capabilities of absorbing ultrasonic energy
by these oils. Consequentially, a high transesterification rate or biodiesel yield was achieved within 30min
[84] . In another study, Cintas etal. showed a power
consumption analysis on a bench-scale continuousflow reactor (3.3l/h) with a power input of 600W at
21.5kHz [41] . To make a complete transesteri fication
in a two-stage reaction, the corresponding energy
consumption was reported to be 0.28kWh per liter
of biodiesel (~1000kJ/l) with 35min ultrasonication
after 30 min conventional heating and mechanical
stirring at 45C [41] .
To gather the information needed for efficient
process design and operation of sonochemical reactors for biodiesel production, the practical aspects of
frequency and intensity of sonication, initial radius
and gas content of the cavities, and the temperature
effect on the cavitation phenomena have been studied
[85] . Process analysis and theoretical simulation with

future science group

numerical solutions of the dynamic models were conducted. The authors reported that to maximize the
ultrasonication effect, a combination of lower frequencies at higher intensity of irradiation was preferred over
higher frequencies and lower intensities of irradiation. In addition, sonochemical reactors with triple
frequency showed considerably higher overall cavitational activities than that of single or dual frequency
at equivalent power dissipation levels. The authors
commented that higher intensity of ultrasonication is
preferred but an optimum intensity may be observed
due to the mechanism of the chemical reaction propagation similar to the concept of the activation energy
in chemical reactions such as pyrolysis [85] . This comment addresses the question about whether there is an
optimum ultrasonication condition as raised by many
researchers. Combined with the physical configuration
and mechanical design [9] , researchers would have a
better understanding of the engineering aspects of
sonochemical reactors for biodiesel production.
Ultrasonication applications in biodiesel production
have attracted attention from both the research community and the industry. However, the sonochemistry
and the design of sonochemical reactors for transesterification of plant oils and animal fats for biodiesel production, as with sonochemistry applications in other
fields, are far from being completely understood. Thus,
they require the continued effort of scientists and engineers from the research and industry communities to
explore this novel technical area.
Additionally, cost analysis is an important aspect
in evaluating ultrasonication applications in biodiesel
production and is of interest to many readers. Cost
analysis could not be included in this short review due
to the fact that most of studies currently available in
the literature are small-scale laboratory explorations
and the results from these studies are either not representative or not directly transferable into commercialscale including process scale-up and control schemes.
Cost analysis, however, should be reviewed in the
future based on additional studies with scaled-up processes to benefit the potential users of this technology
in this field.
Future perspective
It is likely that the application of ultrasound will be
further developed into the commercial production of
biodiesel in the next 510years. As the industry consolidates and new sources of feedstock, such as algae oil
and jatropha oil, become available, plant capacities will
increase to improve economies of scale. Installation
of ultrasonication reactors can provide a good quality
reaction in a shorter time, thus allowing more throughput in the plant. The combination of heterogeneous

www.future-science.com

485

Review He & Van Gerpen

catalysis with ultrasound will provide a robust process


to satisfy increasingly challenging quality specifications. Even small producers will likely find ultrasound
to be an excellent way to enhance productivity.
On the other hand, research will be needed to optimize reactor and sonotrode design to reduce energy
consumption and, thus, reduce the overall life cycle
energy requirement of the biodiesel production process.

Financial & competing interests disclosure


The authors have no relevant affiliations or financial involvement
with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the
manuscript. This includes employment, consultancies, honoraria,
stock ownership or options, expert testimony, grants or patents
received or pending, or royalties. No writing assistance was utilized
in the production of this manuscript.

Executive summary
Ultrasonication-enhanced mixing
When ultrasonication is applied in transesterification of plant oils and animal fats, ultrasonic cavitation and microbubble formation greatly
improve the interfacial contact between the immiscible methanol and plant oil/animal mixture, thus increasing the reaction rate.
The formation and bursting of the microbubbles caused by ultrasonic cavitation intensifies the local energy transfer and energizes the
reactant molecules, thus enhancing the overall reaction rate.
Ultrasonication-induced free radicals
There may be ultrasonic energy-induced free radical formation in the ultrasound-assisted transesterification system, which initiates chain
reactions as has been observed in other organic systems, although it is not yet fully understood for transesterification.
Ultrasonication in biodiesel production
Ultrasonication in transesterification of plant oils and animal fats for biodiesel production has been extensively researched. The most
significant effect of ultrasonication is the greatly shortened reaction time.
There are large variations in reported values for power consumption of ultrasonication, which, along with sonochemical reactor scale-up
and other engineering aspects, requires further investigation.
The application of ultrasonication and other novel technologies, such as heterogeneous catalysts, could lead to better efficiency for
commercial biodiesel production.

References

Papers of special note have been highlighted as:


of interest
of considerable interest

Cheeke J. Fundamentals and Applications of


Ultrasonic Waves. CRC Press, NY, USA
(2002).

Thompson LH, Doraiswamy LK.


Sonochemistry: science and engineering.
Indust. Eng. Chem. Res. 38(4), 12151249A
(1999).

15

Mason TJ. Sonochemistry and the


environment providing a green link
between chemistry, physics and engineering.
Ultrason. Sonochem. 14(4), 476483 (2007).

Very thorough review that describes the


fundamental principles and applications
aspects of ultrasonication; one of the best
articles of its nature.

16

Del Duca M, Yeager E, Davies M, Hovorka F.


Isotopic techniques in the study of the
sonochemical formation of hydrogen
peroxide. J.Acoust. Soc. Am. 30(4), 301307
(1958).

17

Donaldson DJ, Farrington MD, Kruus P.


Cavitation-induced polymerization of
nitrobenzene. J. Phys. Chem. 83(24),
31303135 (1979).

18

Diedrich GK, Kruus P, Rachlis LM.


Cavitation-induced reactions in pure
substituted benzens. Can. J. Chem. 50(11),
17431750 (1972).

19

Suslick KS, Gawienowski JJ, Schubert PF,


Wang HH. Alkane sonochemistry. J.Phys.
Chem. 87(13), 22992301 (1983).

n n

Van Gerpen JH. Biodiesel processing and


production. Fuel Process. Technol. 86(10),
10971107 (2005).

Darrigo G, Paparelli A. Sound-propagation in


water-methanol mixtures at low temperatures.
1. Ultrasonic velocity. J.Chem. Phys. 88(1),
405415 (1988).

Wong SH, Watkins RD, Kupnik M, Pauly


KB, Khuri-Yakub BT. Feasibility of
MR-temperature mapping of ultrasonic
heating from a CMUT. IEEE Trans. Ultrason.
Ferroelectr. Freq. Control 55(4), 811818
(2008).

Shol A. Industrial applications of ultrasound.


In: Ultrasound: Its Chemical, Physical, and
Biological Effects. Suslick K (Ed.). VCH
Publishers, Inc., NY, USA, 97122 (1988).

Flint EB, Suslick KS. The temperature of


cavitation. Science 253(5026), 13971399
(1991).

Suslick KS, Hammerton DA, Cline RE. The


sonochemical hot-spot. J.Am. Chem. Soc.
108(18), 56415642 (1986).

486

Mason TJ. Large scale sonochemical


processing: aspiration and actuality. Ultrason.
Sonochem. 7(4), 145149 (2000).

Biofuels (2012) 3(4)

n n

10

n n

11

Gogate PR, Pandit AB. Sonochemical


reactors: scale up aspects. Ultrason. Sonochem.
11(34), 105117 (2004).
Valuable article on engineering aspects of
ultrasonication application.
Suslick K. Homogeneous sonochemistry. In:
Ultrasound: Its Chemical, Physical, and
Biological Effects. Suslick K (Ed.). VCH
Publishers, Inc, NY, USA, 123163 (1988).

12 Scrivastava S, Berkowitz N. The relationship

between internal pressure and ultrasonic


velocity. Can. J.Chem. 41, 17871793 (1962).
13 Suslick KS. Sonochemistry. Science

comparative study of polymer degradation


and iodine oxidation. J.Phys. Chem. 94(12),
51695172 (1990).

20 Rice FO. The decomposition of organic

compounds into free radicals. Trans. Faraday


Soc. 30, 152169 (1934).

247(4949), 14391445 (1990).


n n

14

Important article that hypothesized the high


temperature spots created by ultrasonication
in liquid.
Henglein A, Gutierrez M. Chemical effect of
continuous and pulsed ultrasound a

21

Rice FO, Dooley MD. The thermal


decomposition of organic compounds from
the standpoint of free radicals. XII. The
decomposition of methane. J.Am. Chem. Soc.
56(7), 27472749 (1934).

future science group

Application of ultrasonication in transesterification processes for biodiesel production Review


22 Rice FO, Glasebrook AL. The thermal

35

decomposition of organic compounds from


the standpoint of free radicals. XI. The
methylene radical. J.Am. Chem. Soc. 56(7),
23812383 (1934).
23 Rice FO, Herzfeld KF. The thermal

R, Maeda Y. Biodiesel production by


esterification of oleic acid with short-chain
alcohols under ultrasonic irradiation
condition. Renew. Energy 34(3), 780783
(2009).

24 Rice FO, Johnston WR. The thermal

26 Rice FO. The thermal decomposition of

Abdullah AZ. Ultrasonic-assisted biodiesel


production process from palm oil using
alkaline earth metal oxides as the
heterogeneous catalysts. Fuel 89(8),
18181825 (2010).
Base-catalyzed fast transesterification of
soybean oil using ultrasonication. Energy
Fuels 21(2), 11611164 (2007).

40 Monnier H, Wilhelm AM, Delmas H. The

Einhorn C, Einhorn J, Dickens MJ, Luche JL.


Organic sonochemistry some illustrative
examples of a new fundamental approach.
Tetra. Lett. 31(29), 41294130 (1990).

G. A new pilot flow reactor for high-intensity


ultrasound irradiation. Application to the
synthesis of biodiesel. Ultrason. Sonochem.
17(6), 985989 (2010).
42 Hsiao M-C, Lin C-C, Chang Y-H, Chen

L-C. Ultrasonic mixing and closed microwave


irradiation-assisted transesterification of
soybean oil. Fuel 89(12), 36183622 (2010).
43 Kalva A, Sivasankar T, Moholkar VS.

Physical mechanism of ultrasound-assisted


synthesis of biodiesel. Indust. Eng. Chem. Res.
48(1), 534544 (2009).
44 Adewuyi YG. Sonochemistry: environmental

science and engineering applications. Indust.


Eng. Chem. Res. 40(22), 46814715 (2001).

32 Luche JL, Einhorn C, Einhorn J,

Sinisterragago JV. Organic sonochemistry a


new interpretation and its consequences.
Tetra. Lett. 31(29), 41254128 (1990).
33 Colucci JA, Borrero EE, Alape F. Biodiesel

from an alkaline transesterification reaction


of soybean oil using ultrasonic mixing. J. Am.
Oil Chem. Soc. 82(7), 525530 (2005).
34 Wu P, Yang Y, Colucci JA, Grulke EA. Effect

of ultrasonication on droplet size in biodiesel


mixtures. J. Am. Oil Chem. Soc. 84(9),
877884 (2007).
n

In-depth study on ultrasonication-induced


mixing in micro level.

future science group

51

Hanh HD, Dong NT, Okitsu K, Maeda Y,


Nishimura R. Test temperature dependence of
transesterification of triolein under lowfrequency ultrasonic irradiation condition.
Jpn J. Appl. Phys. Pt 1 46(7B), 47714774
(2007).

52

Hanh HD, Dong NT, Starvarache C, Okitsu


K, Maeda Y, Nishimura R. Methanolysis of
triolein by low frequency ultrasonic
irradiation. Energy Conver. Manage. 49(2),
276280 (2008).

53 Thanh LT, Okitsu K, Sadanaga Y, Takenaka

N, Maeda Y, Bandow H. Ultrasound-assisted


production of biodiesel fuel from vegetable
oils in a small scale circulation process.
Bioresour. Technol. 101(2), 639645 (2010).

41 Cintas P, Mantegna S, Gaudino EC, Cravotto

30 Khurana JM, Sahoo PK, Maikap GC.

31

Nishimura R. Effects of molar ratio, catalyst


concentration and temperature on
transesterification of triolein with ethanol
under ultrasonic irradiation. J. Jpn Petrol. Inst.
50(4), 195199 (2007).

influence of ultrasound on micromixing in a


semi-batch reactor. Chem. Eng. Sci.
54(1314), 29532961 (1999).

29 Lin JR, Yen TF. An upgrading process

Sonochemical esterification of carboxylic


acids in presence of sulphuric acid. Synth.
Commun. 20(15), 22672271 (1990).

50 Hanh HD, Dong NT, Okitsu K, Maeda Y,

Influence of ultrasound on mixing on the


molecular scale for water and viscous liquids.
Ultrason. Sonochem. 6(12), 6774 (1999).

acetate hydrolysis and fat saponification.


J.Soc. Org. Syn. Chem. (Japan) 7, 167 (1942).
through cavitation and surfactant. Energy
Fuels 7(1), 111118 (1993).

Maeda Y. Conversion of vegetable oil to


biodiesel using ultrasonic irradiation. Chem.
Lett. 32(8), 716717 (2003).

39 Monnier H, Wilhelm AM, Delmas H.

radicalradical reactions disproportionation


vs combination. Chem. Rev. 73(5), 441464
(1973).
28 Miyagawa I. Application of ultrasonication in

49 Stavarache C, Vinatoru M, Nishimura R,

38 Singh AK, Fernando SD, Hernandez R.

organic compounds from the standpoint of


free radicals. I. Saturated hydrocarbons.
J.Am. Chem. Soc. 53(2), 19591972 (1931).
27 Gibian MJ, Corley RC. Organic

of ultrasonically assisted transesterification of


various vegetable oils with methanol.
Ultrason. Sonochem. 14(3), 380386 (2007).

37 Mootabadi H, Salamatinia B, Bhatia S,

25 Rice FO. The thermal decomposition of

organic compounds from the standpoint of


free radicals. III. The calculation of the
products formed from paraffin
hydrocarbons. J.Am. Chem. Soc. 55,
30353040 (1933).

48 Stavarache C, Vinatoru M, Maeda Y. Aspects

36 Hanh HD, Dong NT, Okitsu K, Nishimura

decomposition of organic compounds from


the standpoint of free radicals. VI. The
mechanism of some chain reactions. J.Am.
Chem. Soc. 56, 284289 (1934).
decomposition of organic compounds from
the standpoint of free radicals. V. The
strength of bonds in organic molecules. J.Am.
Chem. Soc. 56, 214219 (1934).

Hanh HD, Dong NT, Okitsu K, Nishimura


R, Maeda Y. Biodiesel production through
transesterification of triolein with various
alcohols in an ultrasonic field. Renew. Energy
34(3), 766768 (2009).

45

Stavarache C, Vinatoru M, Nishimura R,


Maeda Y. Fatty acids methyl esters from
vegetable oil by means of ultrasonic energy.
Ultrason. Sonochem. 12(5), 367372 (2005).

46 Freedman B, Butterfield RO, Pryde EH.

Transesterification kinetics of soybean oil.


J.Am. Oil Chem. Soc. 63(10), 13751380
(1986).
47 Komers K, Stloukal R, Machek J, Skopal F.

Biodiesel from rapeseed oil, methanol and


KOH. 3. Analysis of composition of actual
reaction mixture. Eur. J. Lipid Sci. Technol.
103(6), 363371 (2001).

54 Sarma AK, Sarmah JK, Barbora L etal.

Recent inventions in biodiesel production and


processing a review. Recent Pat. Eng. 2(1),
4758 (2008).
55

Van Manh D, Chen Y-H, Chang C-C, Chang


M-C, Chang C-Y. Biodiesel production from
Tung oil and blended oil via ultrasonic
transesterification process. J. Taiwan Inst.
Chem. Eng. 42(4), 640644 (2011).

56 Yang QL, Zhu F, Sun J, Qin S. Research of

synthesizing biodiesel from Suaeda salsa by


ultrasonic-assisted transesterification. In: 2nd
International Conference on Manufacturing
Science and Engineering, ICMSE 2011, April
911 2011. Trans Tech Publications, Guilin,
China, 39253931 (2011).
57 Yustianingsih L, Zullaikah S, Ju YH.

Ultrasound assisted in situ production of


biodiesel from rice bran. J.Energy Inst. 82(3),
133137 (2009).
58 Fan X, Wang X, Chen F. Ultrasonically

assisted production of biodiesel from crude


cottonseed oil. Int. J. Green Energy 7(2),
117127 (2010).
59 Georgogianni KG, Kontominas MG,

Pomonis PJ, Avlonitis D, Gergis V.


Conventional and in situ transesterification of
sunflower seed oil for the production of

www.future-science.com

487

Review He & Van Gerpen


biodiesel. Fuel Process. Technol. 89(5),
503509 (2008).

69 Kumar G, Kumar D, Poonam, Johari R,

Singh CP. Enzymatic transesterification of


Jatropha curcas oil assisted by ultrasonication.
Ultrason. Sonochem. 18(5), 923927 (2011).

60 Babajide O, Petrik L, Amigun B, Ameer F.

Low-cost feedstock conversion to biodiesel via


ultrasound technology. Energies 3(10),
16911703 (2010).
61

Georgogianni KG, Kontominas MG, Tegou


E, Avlonitis D, Gergis V. Biodiesel
production: reaction and process parameters
of alkali-catalyzed transesterification of waste
frying oils. Energy Fuels 21(5), 30233027
(2007).

62 Hingu SM, Gogate PR, Rathod VK.

Ultrasound-assisted synthesis of biodiesel


from palm fatty acid distillate. Indust. Eng.
Chem. Res. 48(17), 79237927 (2009).
64 Santos FFP, Malveira JQ, Cruz MGA,

Fernandes FAN. Production of biodiesel by


ultrasound assisted esterification of
Oreochromis niloticus oil. Fuel 89(2), 275279
(2009).
65

Worapun I, Pianthong K, Thaiyasuit P,


Thinvongpituk C. Effect on the use of
ultrasonic cavitation for biodiesel production
from crude Jatropha curcas L. seed oil with a
high content of free fatty acid. In: 4th
International Conference on Experimental
Mechanics, November 1820 2009. SPIE,
Singapore (2010).

66 Rodrigues S, Mazzone LCA, Santos FFP,

Cruz MGA, Fernandes FAN. Optimization of


the production of ethyl esters by ultrasound
assisted reaction of soybean oil and ethanol.
Braz. J. Chem. Eng. 26(2), 361366 (2009).
67 Kumar D, Kumar G, Poonam, Singh CP.

Fast, easy ethanolysis of coconut oil for


biodiesel production assisted by
ultrasonication. Ultrason. Sonochem. 17(3),
555559 (2010).
68 Kumar D, Kumar G, Poonam, Singh CP.

Ultrasonic-assisted transesterification of
Jatropha curcus oil using solid catalyst, Na/
SiO2. Ultrason. Sonochem. 17(5), 839844
(2010).

488

Biofuels (2012) 3(4)

79 Arzamendi G, Arguinarena E, Campo I,

Zabala S, Gandia LM. Alkaline and


alkaline-earth metals compounds as catalysts
for the methanolysis of sunflower oil. Catalysis
Today 133, 305313 (2008).

One of a few examples of ultrasonicationassisted biodiesel preparation via enzymatic


processes.

70 Ji JB, Wang JL, Li YC, Yu YL, Xu ZC.

80 Yu D, Tian L, Wu H etal. Ultrasonic

irradiation with vibration for biodiesel


production from soybean oil by Novozym
435. Process Biochem. 45(4), 519525 (2010).

Preparation of biodiesel with the help of


ultrasonic and hydrodynamic cavitation.
Ultrasonics 44(Suppl. 1), e411e414 (2006).
71 Teixeira LSG, Assis JCR, Mendonca DR etal.

Comparison between conventional and


ultrasonic preparation of beef tallow biodiesel.
Fuel Process. Technol. 90(9), 11641166
(2009).

Synthesis of biodiesel from waste cooking oil


using sonochemical reactors. Ultrason.
Sonochem. 17(5), 827832 (2010).
63 Deshmane VG, Gogate PR, Pandit AB.

biodiesel production process. Chem. Eng.


Technol. 30(11), 14811487 (2007).

81

82 Mahamuni NN, Adewuyi YG. Optimization

of the synthesis of biodiesel via ultrasoundenhanced base-catalyzed transesterification of


soybean oil using a multifrequency ultrasonic
reactor. Energy Fuels 23, 27572766 (2009).

72 Chand P, Chintareddy VR, Verkade JG,

Grewell D. Enhancing biodiesel production


from soybean oil using ultrasonics. Energy
Fuels 24(3), 20102015 (2010).
73 Stavarache C, Vinatoru M, Maeda Y, Bandow

83 Mahamuni NN, Adewuyi YG. Application of

taguchi method to investigate the effects of


process parameters on the transesterification
of soybean oil using high frequency
ultrasound. Energy Fuels 24(3), 21202126
(2010).

H. Ultrasonically driven continuous process


for vegetable oil transesterification. Ultrason.
Sonochem. 14(4), 413417 (2007).
74

Stavarache C, Vinatoru M, Maeda Y.


Ultrasonic versus silent methylation of
vegetable oils. Ultrason. Sonochem. 13(5),
401407 (2006).

84 Lee SB, Lee JD, Hong IK. Ultrasonic energy

effect on vegetable oil based biodiesel


synthetic process. J.Indust. Eng. Chem. 17(1),
138143 (2011).

75 Georgogianni KG, Katsoulidis AP, Pomonis

PJ, Kontominas MG. Transesterification of


soybean frying oil to biodiesel using
heterogeneous catalysts. Fuel Process. Technol.
90(5), 671676 (2009).

85 Prabhu AV, Gogate PR, Pandit AB.

Optimization of multiple-frequency
sonochemical reactors. Chem. Eng. Sci.
59(2223), 49914998 (2004).

76 Verziu M, Coman SM, Richards R,

Parvulescu VI. Transesterification of Vegetable


Oils Over CaO Catalysts. Elsevier, Amsterdam,
The Netherlands, 6470 (2011).
77 Georgogianni KG, Katsoulidis AK, Pomonis

PJ, Manos G, Kontominas MG. Trans


esterification of rapeseed oil for the
production of biodiesel using homogeneous
and heterogeneous catalysis. Fuel Process.
Technol. 90(78), 10161022 (2009).
78 Ye X, Fernando S, Wilson W, Singh A.

Canals A, Hernandez MD. Ultrasoundassisted method for determination of chemical


oxygen demand. Analyt. Bioanalyt. Chem.
374(6), 11321140 (2002).

Patents
101 Maeda Y, Vinatoru M, Stavarach CE, Iwai K,

Oshige H: US6884900 (2005).


102 Maeda Y, Vinatoru M, Stavarach CE, Iwai K,

Oshige H: EP1411042 (2004).


103 Maeda Y, Vinatoru M, Stavarach CE, Iwai K,

Oshige H: JP156022A (2004).


n

Early research on ultrasonication application


in biodiesel production.

Application of amphiphilic catalysts


utrasonication, and nanoemulsions for

future science group

S-ar putea să vă placă și