Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

Contents lists available at ScienceDirect

Chemical Engineering & Processing: Process Intensification


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

Transesterification of vegetable oil with methanol using solid base catalyst T


of calcium oxide under ultrasonication

Yoshihiro Kojimaa, , Shinya Takaib
a
Institute of Materials and Systems for Sustainability, Nagoya University, Furo-cho, Chikusa-ku, Nagoya, Aichi, 464-8603, Japan
b
Department of Chemical Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya, Aichi, 464-8603, Japan

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

Keywords: Transesterification of vegetable oil with methanol using calcium oxide as a solid base catalyst was carried out
Transesterification under ultrasonication and mechanical stirring conditions, and the transesterification rate was compared. In
Biodiesel addition, the effects of ultrasonic power, catalyst mass and molar ratio of methanol to oil on methanolysis rate of
Calcium oxide oil under ultrasonication were investigated. The biodiesel yield reached 88% at about 2.5 and 5 hs under ul-
Ultrasonication
trasonication and mechanical stirring, respectively, thus demonstrating that ultrasonication was more effective
Methanolysis rate
method for the acceleration of transesterification rate. The increase in both the ultrasonic power per unit volume
of mixture solution and the catalyst mass resulted in an increase in the methanolysis rate. It was found that the
larger amount of methanol in an excess over the stoichiometric amount to oil played more significant role in
enhancement of production rate of biodiesel in early reaction period limited by mass transfer of oil, but was less
favorable for the enhancement of methanolysis rate in the reaction period limited by chemical reaction.

1. Introduction purification of biodiesel product by applying the heterogeneous base


catalyst such as calcium oxide which can be recovered from mixture
Recently, biofuels have widely been used around the world, due to solution. However, the reaction rate of transesterification by hetero-
the increase in energy demand and concerns about energy and en- geneous catalyst is relatively slower compared to that by homogeneous
vironmental issues such as the depletion of fossil fuels and greenhouse catalyst [5]. In order to understand characteristics of transesterification
warming. Especially, fatty acid methyl esters (FAMEs), also called in the presence of heterogeneous catalysts, improve the transester-
“biodiesel fuel”, have become an alternative fuel for diesel engines, ification reaction rate and explore the reaction mechanism, many re-
because they have similar fuel properties to fossil fuel and can be used searchers have studied the transesterification reaction of vegetable oil
without any modification of the engines [1]. Biodiesel is most com- to biodiesel for many types of heterogeneous catalysts such as MgO,
monly produced by transesterification reactions of vegetable oils with CaO, BaO, PbO, MnO2, K2CO3/c-Al2O3, KF/c-Al2O3 and CaO/SiO2
mono-alkyl alcohol such as methanol. The formula scheme for the [2,5,10–18]. Singh and Fernando investigated the transesterification
production of FAMEs out of triglyceride is as shown below [2,3] kinetics of heterogeneous catalysts of MgO, CaO, BaO, PbO and MnO2,
Scheme1, with soybean oil and reported that heterogeneous catalysts require
In order to shift the equilibrium to the right of reaction shown in much higher temperatures and pressures to achieve acceptable con-
Scheme 1, alcohol is commonly added in an excess over the stoichio- version levels compared to homogeneous catalysts [2]. For transester-
metric amount. For the production process of FAMEs, homogeneous ification of soybean oil using CaO catalyst, Kouzu et al. reported that
catalyst such as sodium hydroxide (NaOH) or potassium hydroxide CaO catalyst prepared via calcination in a helium gas flow more sig-
(KOH) is commonly used [4,5]. High conversion of triglyceride can be nificantly had activity compared with that in an ambient air [13].
achieved in comparatively short reaction times under mild conditions Granados et al. investigated the influence of H2O and CO2 on the de-
in the presence of homogeneous catalyst. On the contrary, its dis- terioration of the catalytic performance by contact with room air and
advantage is that homogenous catalysts can not be reused. Moreover, reported that CO2 is the main deactivating agent [10]. Liu et al. re-
several washing steps are usually required to remove the catalyst re- ported that water content in methanol resulted in a significant increase
sidues from the ester products, giving rise to increased costs of biodiesel in biodiesel yield in transesterification reaction process in the presence
production [5–9]. Hence, an attempt has been made to simplify of CaO [5]. Kouzu et al. investigated an active phase of CaO by


Corresponding author.
E-mail address: ykojima@imass.nagoya-u.ac.jp (Y. Kojima).

https://doi.org/10.1016/j.cep.2018.12.007
Received 14 March 2018; Received in revised form 15 December 2018; Accepted 16 December 2018
Available online 19 December 2018
0255-2701/ © 2018 Elsevier B.V. All rights reserved.
Y. Kojima, S. Takai Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

synthesis of biodiesel fuel using a solid catalyst such as CaO, because it


appeared that physical/chemical and heating effects induced by ultra-
sound were capable of improving transesterification reaction rate of
vegetable oil with mono-alkyl alcohol in the presence of solid catalyst.
High power ultrasound technique has been applied for the preparation
of solution, the synthesis and destruction of chemical species, the im-
provement and acceleration of chemical reaction rate and the pre-
paration of materials [19–25]. During ultrasound propagation in a li-
quid, the liquid flow which is termed acoustic streaming would
Scheme 1. Transesterification of triglyceride oil to biodiesel.
generate in the same direction as the ultrasound wave propagation. In
addition, high power ultrasonic irradiation to a solution causes the
characterizing the catalyst collected after achieving the conversion of formation, growth and collapse of cavitation bubbles in the solution.
soybean oil into its methyl ester at reflux of methanol in a glass batch The implosive collapse of the cavitation bubbles during the adiabatic
reactor and indicated that CaO combined with the by-produced glycerol compression creates a local field with an enormous temperature of ~
to form calcium diglyceroxide as a major component of the collected 5000 K and pressures of ~100 MPa, also called “hotspot”. In/around
catalyst [16,17]. They also investigated the characteristics of transes- the hotspot, molecules solvent and dissolved gas species are decom-
terification using calcium diglyceroxide [16]. In the methanolysis posed pyrolytically to produce free-radicals. The chemical reactions
process of sunflower oil using CaO calcined at various temperatures, proceed by contact the radicals with molecule and/or the other radi-
Veljković et al. studied the kinetics of methanolysis under mechanical cals; chemical effect of ultrasound arises. The compression, expansion
stirring condition [14]. They analyzed the kinetics of methanolysis of and collapse of the cavitation bubbles also cause the formation of mi-
sunflower oil on the following assumptions: “the overall methanolysis crostreaming, high-speed microjets and generation of the shockwave in
rate in the initial reaction period is limited by the triglyceride mass the solution [26–29]. The physical effects due to acoustic streaming,
transfer rate towards the active sites of catalyst surface; in the later high-speed microjets, microstreaming and shockwave enable us to
reaction period, the overall methanolysis rate is limited by the rate of agitate solution effectively. They also play an important role in not only
the chemical reaction between a methoxide ion and triglyceride; and disruption of immiscible liquid layers but also acceleration of mass
the conversion of triglyceride follows the pseudo-first order reaction transfer at the liquid-solid or/and liquid-liquid interface regions. Sev-
kinetics. According to the assumptions, the rates of triglyceride mass eral researchers have investigated ultrasound application for synthesis
transfer and triglyceride reaction have about the same value: of biodiesel fuel in the presence of homogeneous catalysts such as
dcA m NaOH and KOH, and have reported that the ultrasonication is useful for
(−rA ) = − = k s,A⋅θ⋅am⋅(cA − cA,s)⋅ cat = k mt,A⋅(cA − cA,s) = k⋅cA,s,
dt V the enhancement of production rate of biodiesel [30–38].
(1) In this study, an attempt was made to synthesize biodiesel fuel via
the transesterification reactions of soybean oil with methanol using CaO
where (-rA) [mol・dm−3・min-1] is the rate of triglyceride consumption, solid catalyst under ultrasonic irradiation. The pseudo-first-order re-
cA [mol・dm−3] is the triglyceride concentration in the liquid phase; t
action kinetics model was applied to the methanolysis of soybean oil to
[min] is the reaction time; ks,A [m・min-1] is the triglyceride mass determine the respective methanolysis rate constants, km and kc, in a
transfer coefficient; θ [-] is the fraction of the available active specific
triglyceride mass transfer controlled regime and a subsequent chemical
catalyst surface; am [m2・g-1] is the active specific catalyst surface; cA,s reaction controlled regime. We compared the difference in the metha-
[mol・dm−3] is the concentration of triglyceride adsorbed on the cata-
nolysis rates of oil between ultrasonic treatment and mechanical stir-
lyst surface per liquid phase volume; mcat [g] is the catalyst mass; V ring treatment. In addition, the effects of ultrasonic power, catalyst
[m3] is the reaction mixture volume; kmt,A [min-1] is the volumetric
mass and molar ratio of methanol to oil on the methanolysis rate of oil
triglyceride mass transfer coefficient; and k [min-1] is the pseudo-first were investigated under ultrasonication.
order reaction rate constant. By rearranging Eq. (1), the following
equation can be obtained:
2. Experimental
dcA k⋅k mt,A
− = ⋅cA = kapp⋅cA,
dt k mt,A + k (2) 2.1. Reagents
−1
where kapp [min ] is the apparent rate constant which is related to the
mass transfer, as well as the chemical reaction rate. When kmt,A is sig- Soybean oil, methanol (CH3OH), calcium carbonate (CaCO3), hy-
nificantly smaller than k in the initial phase, then drogen chloride (1 mol・l−1 HCl aq) and heptane (CH3(CH2)5CH3, 99+
%) were purchased from Wako Chemicals. Co. Soybean oil used in this
− ln(1 − xA ) = kapp t + C = k mt,A t + C , (3) work had a saponification value of 188–195 and an iodine value of 123-
142. Density (specific gravity) of soybean oil was 0.925 g・cm-3 (298 K).
where xA [-] is the conversion of triglyceride and C [-] is the integration
Methyl heptadecanoate (CH3(CH2)15COOCH3, 97+%) was obtained
constant. As the reaction proceeds, the amount of methanol adsorbed
from Tokyo Chemical Co.
on the active catalyst surface decreases, while the fraction of the
available active catalyst surface for triglyceride adsorption and the
volumetric triglyceride mass transfer coefficient increase. When kmt,A is 2.2. Preparation of calcium oxide catalyst
significantly larger than k in the initial phase, then
− ln(1 − xA ) = kapp t + C = kt + C , To prepare calcium oxide (CaO) catalyst, CaCO3 was calcinated in a
(4)
tubular furnace under helium atmosphere. During the calcination
and the chemical reaction between triglyceride and methanol controls treatment, the temperature rose at the rate of 6 K・min−1 from room
the overall methanolysis rate. Thus, Veljković et al. applied the pseudo- temperature and was kept at 1273 K for 5 h, followed by cooling down
first-order reaction kinetics model to the methanolysis, and determined at the rate of 6 K・min−1 until room temperature. The results of XRD
the methanolysis rate constants, km (= kapp = kmt,A) and kc (= kapp = analysis in Fig. 1 reveal the complete conversion of CaCO3 to CaO via
k), in a triglyceride mass transfer controlled regime and a subsequent the calcination treatment. The specific surface area and mean particle
chemical reaction controlled regime, respectively. diameter of the CaO obtained after the calcination of CaCO3 was about
Our research interest was focused on ultrasonically assisted 7.98 m2・g−1 and 18.7 μm, respectively.

102
Y. Kojima, S. Takai Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

The revolution rate of rotor blades was 400 rpm.


During the reaction, the samples (0.5 mL) were removed from the
reaction mixture at 15 or 30 min time interval, immediately quenched
by adding 0.2 mL HCl solution (0.1 mol・L−1) to neutralize the CaO, and
then centrifuged (2500 rpm, 10 min). 250 mg of products mixture was
withdrawn from the upper ester layer of the centrifuged mixture and
dissolved in a 5 mL heptane solution including the 50 mg methyl hep-
tadecanoate. The prepared solution was used to determine the amount
of methyl ester formed in the transesterification reaction by gas chro-
matographic analysis.
The molar ratio of methanol to soybean oil was in the range of
Fig. 1. XRD analysis of materials prepared by calcination of CaCO3. 6:1–18:1. The volume of soybean oil was 70 mL except when the effect
of molar ratio of methanol to soybean oil under the same mixture vo-
lume on the methanolysis rate of oil was investigated. The mass ratios
2.3. Experimental apparatus and operating conditions
of CaO catalyst to soybean oil were 0.5, 1.0 and 1.5 wt%.
Fig. 2 shows a schematic diagram of the experimental apparatus and
a photograph of reaction vessel. A reaction vessel consists of a stainless- 2.4. Measurement
steel cylindrical vessel (Honda Electronics) and a glass cap equipped
with a reflux. The stainless-steel cylindrical vessel has a double tube The amount of methyl ester formed in the transesterification reac-
structure (inner diameter, φ50 mm and height, 115 mm). The bottom of tion was measured by GC-FID (GC-4000, GL Sciences Inc.) equipped
the vessel played a role of vibrating plate to which a Langevin type with a column (Inart-cap 225, GL Sciences Inc.). Oven temperature,
transducer with a diameter of 45 mm is attached. A mixture of me- injector temperature and detector temperature were set at 488, 523 and
thanol, soybean oil and CaO catalyst was situated in the vessel equipped 523 K, respectively.
with a thermocouple (Copper and constantan, Omega Engineering Inc.) According to the European regulated procedure EN14103, the yield
which is connected with a temperature recorder (NR-500, NR-TH08, (y [wt%]) of the FAME product was determined as follows:
Keyence Corp.) and a personal computer. Temperature of the mixture
(∑ Ai − AMH ) ⎤ CMH VMH × 100
was maintained at 333 K during reaction using a circulating water bath. y [wt%] = ⎡ ,

⎣ AMH ⎥
⎦ W (6)
During transesterification reactions, either the mixture in a reaction
vessel was irradiated with ultrasound or it was agitated by mechanical
where ∑ Ai is the total peak area of FAME mixture, AMH the peak area
stirring. Ultrasound (about 43 kHz) was emitted into the mixture from
of methyl heptadecanoate which is the internal standard, CMH the
the transducer using a power amplifier (Honda Electronics L-400BM-L)
concentration of the methyl heptadecanoate in heptane (10 g・L−1),
and a function generator (NF W1946B). Ultrasonic power emitted was
VMH the volume of the methyl heptadecanoate solution (5 mL) and W
measured by calorimetry as recommended elsewhere [39]. The ultra-
the weight of the products mixture (250 mg). In this study, the con-
sonic power PUS was calculated using the following equation:
version of triglyceride, x[-], was calculated on the assumption that it
ΔT was equal to the y/100.
PUS = MCP The X-ray diffraction analysis (Rint-2500TTR, Rigaku) was carried
Δt (5)
out to identify the CaO produced by calcination of CaCO3. The specific
where M, CP and ΔT/Δt denote liquid mass, specific heat of liquid at surface area of the CaO catalyst was measured by a surface area mea-
constant pressure and rate of temperature rise, respectively. For a tra- suring instrument (BELSORP-mini, BEL Japan Inc.). The laser diffrac-
ditional stirring process, a stirring device and rotor blades were used. tion particle size distribution analyzer (LA-920, HORIBA) was

Fig. 2. Schematic diagram of the experimental apparatus and photograph of reaction vessel.

103
Y. Kojima, S. Takai Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

Table 1
The first-order reaction rate constants under ultrasonic and mechanical stirring
conditions. Oil content: 70 mL; CaO/oil mass ratio: 1 wt%; methanol/oil molar
ratio: 6/1; ultrasonic power: 5.0 W; revolution rate of stirring: 400 rpm.
method km×103[min−1] kc×103[min−1]

Mechanical stirring 2.1 ( ± 0.2) 13 ( ± 2)


Ultrasonication 0.5 ( ± 0.2) 30 ( ± 2)

soybean oil triglyceride via its transesterification. In both cases of ul-


trasonication and mechanical stirring, the results in Fig. 4 reveal that
the dependence of –ln(1-x) on reaction time is linear in each regime
controlled by mass transfer or chemical reaction, indicating that the
pseudo-first-order reaction kinetics model can probably be applied to
the methanolysis of soybean oil under ultrasonication condition as well
Fig. 3. Time-dependent change in biodiesel yield under ultrasonic and me- as mechanical stirring condition. According to the pseudo-first-order
chanical stirring conditions. Oil content: 70 mL; CaO/oil mass ratio: 1 wt%; reaction kinetics model, the reaction rate constants, km and kc, were
methanol/oil molar ratio: 6/1; ultrasonic power: 5.0 W; revolution rate of estimated from the respective slopes in the initial and latter regions in
stirring: 400 rpm.
Fig. 4. The rate constants estimated from the slopes are listed in Table 1.
The reaction rate constant, km, under mechanical stirring condition was
employed to measure the mean particle diameter of the CaO catalyst. about 4.2 times as large as that under ultrasonication condition, sug-
gesting that the stirring treatment at 400 rpm can disturb methanol
3. Results and discussion including solid catalysts effectively in an oil layer and subsequently
contribute to the formation of methanol/oil emulsion. On the contrary,
3.1. Ultrasonic treatment and stirring treatment ultrasonication at 5.0 W is not probably intensive enough to cause the
fast acoustic streaming and formation/collapse of cavitation bubbles
Fig. 3 shows the typical time-change of yield of FAMEs produced via which are responsible for uniformly forming the methanol/oil emulsion
transesterification reactions of soybean oil with methanol by ultra- and dispersing CaO catalysts over the mixture solution. In addition,
sonication and mechanical stirring methods at a 6:1 molar ratio of when ultrasound propagates in the solution including soybean oil of the
methanol to oil. Until early about 75 min, the reaction rate under ul- high viscosity, it appears that it is not easy to cause the formation/
trasonication condition was slower than that under mechanical stirring collapse of cavitation bubbles in the solution compared to liquids of the
condition. However, after about 75 min, the reaction rate was ac- low viscosity such as water and methanol, because cavitation threshold
celerated under ultrasonication condition compared to mechanical increases with an increase in viscosity of liquid generally. As shown in
stirring condition. The FAMEs yield reached 88% at about 2.5 h and 5 h Table 1, however, the reaction rate constant, kc, obtained by ultra-
for ultrasonication and mechanical stirring, respectively; thus the re- sonication treatment was about 2.3 times as large as that by mechanical
sults demonstrate that enhancement in the overall average transester- stirring treatment. As the methanolysis reactions progress, the amount
ification reaction rate is achieved under ultrasonication condition. of formed FAMEs increases and to the contrary the amount of soybean
Veljković et al. studied the kinetics of methanolysis of sunflower oil oil decreases; the changes in the amounts of reactants and products
in the presence of CaO catalyst under mechanical stirring condition with reaction time seem to cause a decrease in the viscosity of the
[14]. They applied the pseudo-first-order reaction kinetics model to the mixture solution, due to mutual interaction among methanol, FAMEs,
methanolysis, and determined the methanolysis rate constants, km and oil and others. Cavitation threshold gradually decreases with the de-
kc, in a triglyceride mass transfer controlled regime and a subsequent crease in viscosity of the mixture solution and is eventually equal to the
chemical reaction controlled regime, respectively. Fig. 4 shows the acoustic pressure equivalent to the applied ultrasound power; then the
plots of [–ln(1-x)] against reaction time under ultrasonic and me- physical effects such as microstreaming and microjet resulting from the
chanical stirring treatments, where x denotes the conversion of the formation/collapse of the cavitation bubbles are immediately in-
tensified in the mixture solution, probably contributing to evolution of
emulsification, uniformly effective dispersion of CaO catalysts in the
solution, efficient cleaning of the solid catalyst surface and/or promo-
tion of mass transfer of the solution toward the solid catalyst surface. In
addition, implosive collapse of cavitation bubbles may offer the local
hot spot with enormous temperature in the mixture solution. It is
supposed that especially cavitation effects induced under ultrasonic
irradiation condition cause considerable acceleration of the reaction
rate in the chemical reaction controlled regime as well as the transition
regime which follows the initial triglyceride mass transfer controlled
region. We conclude that ultrasound is a promising technique for en-
hancement of the overall average reaction rate of biodiesel production
in the presence of inhomogeneous catalyst.

3.2. Effect of ultrasonic power

Fig. 4. Plots of [–ln(1-x)] against reaction time under ultrasonic and mechan- The effect of ultrasonic power on the reaction rate constants, km and
ical stirring conditions. Oil content: 70 mL; CaO/oil mass ratio: 1 wt%; me- kc was investigated. The results in Table 2 demonstrate that an increase
thanol/oil molar ratio: 6/1; ultrasonic power: 5.0 W; revolution rate of stirring: of the ultrasonic power resulted in acceleration of the transesterifica-
400 rpm. tion reaction rates. The km values were about 0.6 × 10–3, 0.7 × 10–3

104
Y. Kojima, S. Takai Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

Table 2 Table 4
Effect of ultrasonic power on the first-order reaction rate constants. Oil content: Effect of molar ratio of methanol to oil on the first-order reaction rate constants
70 mL; CaO/oil mass ratio: 1 wt%; methanol/oil molar ratio: 12/1. under the same oil content condition. Oil content: 70 mL; CaO/oil mass ratio:
1 wt%; ultrasonic power: 5.0 W.
power km×103[min−1] kc×103[min−1]
molar ratio km×103[min−1] kc×103[min−1]
1.2W 0.6 ( ± 0.1) 19 ( ± 1)
3.0W 0.7 ( ± 0.2) 26 ( ± 2) oil:methanol = 1:6 0.5 ( ± 0.2) 30 ( ± 2)
5.0W 1.1 ( ± 0.2) 28 ( ± 2) oil:methanol = 1:12 1.1 ( ± 0.2) 28 ( ± 2)
oil:methanol = 1:18 1.6 ( ± 0.5) 21 ( ± 2)

and 1.1 × 10–3 min–1 at 1.2. 3.0 and 5.0 W, respectively, indicating that
the higher ultrasonic power was favarable for mass transfer. In general, 3.4. Effect of molar ratio of methanol to oil
it is believed that an increase in the ultrasonic power leads to in-
tensification of acoustic streaming and cavitation effects, which may be 3.4.1. Condition of constant amount of oil
responsible for the effective mixing of methanol with soybean oil in a The dependence of the reaction rate constants, km and kc, on the
mass transfer controlled regime. As shown in Table 2, the kc also in- molar ratio of methanol to soybean oil was investigated. The results are
creased from 19 × 10−3 to 28 × 10−3 min-1 for an increase in the ul- summarized in Table 4. The experiments were performed under the
trasonic power from 1.2 to 5.0 W. In the mixture solution, implosive constant conditions of the soybean oil content (70 mL) and ultrasonic
collapse of cavitation bubbles can offer the local hot spot with en- power. Hence the total volumes of sample mixture were about 84, 105
ormous temperature, which is probably higher under higher ultrasonic and 121 mL for the 6:1, 12:1 and 18:1 molar ratios of methanol to oil,
power condition. Additionally, it is considered that the implosive col- respectively, thus meaning that the ultrasonic energy per unit mixture
lapse of cavitation bubbles causes locally high speed flow such as mi- solution volume decreased with an increase in the molar ratio of me-
crojet, which can contribute to the enhancement in collision of me- thanol to oil. The ultrasonic energy per unit mixture volume at the 18:1
thanol alkoxide molecule with soybean oil molecule. molar ratio of methanol to oil was about 1.15 and 1.39 times as low as
those at the molar ratios of 12:1 and 6:1, respectively. However, as can
be seen in Table 4, the rate constant, km, increased from
3.3. Effect of CaO catalyst mass 0.5 × 10−3 min-1 to 1.6 × 10−3 min-1 for an increase in the molar ratio
from 6:1 to 18:1. When the oil content in mixture is relatively smaller,
Table 3 shows the effect of catalyst mass on the transesterification the viscosity of the mixture is probably lower. The acoustic streaming
reaction rates of oil to biodiesel product under ultrasonication. The accelerates more smoothly in the mixture under the lower viscosity
results demonstrate that both km and kc depended on catalyst mass. As condition and thus can disturb the interface of oil/methanol layer,
catalyst mass was incresed from 0.5 to 1.5 wt%, the km increased from probably leading to intensification of emulsification which can con-
0.7 × 10−3 to 1.4 × 10−3 min-1, indicating that the larger catalyst tribute to enhancement of the oil mass transfer. On the other hand, as
mass leads to the enhancement of transesterification reaction rate in a seen in Table 4, for 6:1, 12:1 and 18:1 molar ratios of methanol to oil,
triglyceride mass transfer controlled regime. The pseudo-first-order the kc values were 30 × 10–3, 28 × 10–3 and 21 × 10–3 min–1, respec-
reaction kinetics model proposed by Veljković et al. suggests that the km tively, suggesting that the higher rate constant is obtained at the lower
increases proportionally with catalyst mass [14]. The results in Table 3 molar ratio of methanol to oil. As the molar ratio increases, the ultra-
were similar to the suggestion. Veljković et al. reported that kc did not sonic energy per unit mixture volume decreases, resulting in a decrease
depend on the catalyst mass, while our results demonstrate that the kc in the number of local hot spots with enormous temperature which are
slightly increased from 26 × 10−3 to 30 × 10−3 min-1 as catalyst mass formed by implosive collapse of cavitation bubbles. Additionally, an
increased from 0.5 to 1.5 wt%. This may be because an increase in the increase in the amount of methanol in an excess over the stoichiometric
solid catalyst particle content causes the intensification of cavitation amount to oil may give rise to lowering of collision frequency between
effects. Tuziuti et al. studied about the enhancement of chemical re- soybean oil molecules and molecules of calcium alkoxide.
action with a photocatalyst of titanium dioxide by ultrasonic irradiation
and reported that the effect of titanium dioxide particle addition on the
enhancement of chemical reaction is due to the increase in the cavita- 3.4.2. Condition of constant volume of methaol/oil mixture
tion bubbles [40]. Based on the results in Table 3, it is entirely fair to In the Section 3.4.1, we reported the effect of the molar ratio of
say that the overall average rate of biodiesel production in the presence methanol to soybean oil on the reaction rate under the constant
of CaO solid catalyst under ultrasonication increases with the catalyst amounts of soybean oil and ultrasonic power. In this case, un-
mass. Liu et al. reported that the biodiesel yield was significantly im- fortunately, the ultrasonic power per unit mixture volume was different
proved with the increase of CaO solid catalyst mass in the experiments with each molar ratio condition. Hence, to maintain the ultrasonic
of transesterification reactions of soybean oil under magnetic stirring power per unit mixture volume, the dependence of reaction rates on the
condition [5]. Their results are similar to those obtained in our study. molar ratio of methanol to soybean oil was investigated under the
constant conditions of the methanol/oil mixture volume and ultrasonic
power. As shown in Table 5, the mass transfer controlled rate constant,
km, increased from 0.5 × 10−3 min-1 to 1.3 × 10−3 min-1 at the molar
ratio of methanol to soybean oil of 6:1-18:1 and the dependence of km

Table 3 Table 5
Effect of CaO catalyst mass on the first-order reaction rate constants. Oil con- Effect of molar ratio of methanol to soybean oil on the first-order reaction rate
tent: 70 mL; CaO/oil mass ratio: 1 wt%; methanol/oil molar ratio: 12/1; ul- constants under the same mixture volume condition. Methanol/oil mixtute
trasonic power: 5.0 W. amount: 105 mL; CaO/oil mass ratio: 1 wt%; ultrasonic power: 5.0 W.
CaO catalyst mass km×103[min−1] kc×103[min−1] molar ratio km×103[min−1] kc×103[min−1]

0.5 wt% 0.7 ( ± 0.2) 26 ( ± 2) oil:methanol = 1:6 0.5 ( ± 0.2) 28 ( ± 2)


1.0 wt% 1.1 ( ± 0.2) 28 ( ± 2) oil:methanol = 1:12 1.1 ( ± 0.2) 28 ( ± 2)
1.5 wt% 1.4 ( ± 0.2) 30 ( ± 2) oil:methanol = 1:18 1.3 ( ± 0.1) 24 ( ± 1)

105
Y. Kojima, S. Takai Chemical Engineering & Processing: Process Intensification 136 (2019) 101–106

on the molar ratio was similar to that seen in Table 4. On the other catalyst, J. Japan Inst. Energy 85 (2006) 135–141.
hand, the chemical reaction controlled rate constant, kc at the higher [14] V.B. Veljković, O.S. Stamenković, Z. Todorović, M.L. Lazić, D.U. Skala, Kinetics of
sunflower oil methanolysis catalyzed by calcium oxide, Fuel 88 (2009) 1554–1562.
molar ratio of 18:1, was slightly lower, perhaps because of ineffective [15] S. Yamanaka, M. Kouse, K. Kadota, A. Shimosaka, Y. Shirakawa, J. Hidaka, Catalysis
impingement of the soybean oil molecules on the molecules of calcium by CaO/SiO2 composite particle for biodiesel production, Kagaku Kougaku
alkoxide due to the excess amount of methanol. However, the rate Ronbunshu 33 (3) (2007) 483–489.
[16] M. Kouzu, T. Kasuno, M. Tajika, S. Yamanaka, J. Hidaka, Active phase of calcium
constant, kc, at the molar ratio of 12:1 was almost the same as that at oxide used as solid base catalyst for transesterification of soybean oil with refluxing
the 6:1 molar ratio. The higher methanol content in the mixture may methanol, Appl. Catal. A Gen. 334 (2008) 357–365.
have a negative effect on impingement of the soybean oil molecules on [17] M. Kouzu, S. Yamanaka, J. Hidaka, M. Tsunomori, Heterogeneous catalysis of cal-
cium oxide used for transesterificationof soybean oil with refluxing methanol, Appl.
the molecules of calcium alkoxide as described above; but it is con- Catal. A Gen. 355 (2009) 94–99.
sidered that several cavitation effects which are responsible for the [18] M. Kouzu, T. Kasuno, M. Tajika, Y. Sugimoto, S. Yamanaka, J. Hidaka, Calcium
acceleration of the methanolysis rate are intensified at the 12:1 molar oxide as a solid base catalyst for transesterificationof soybean oil and its application
to biodiesel production, Fuel 87 (2008) 2798–2806.
ratio compared to the 6:1 molar ratio, because the viscosity of mixture
[19] L.M. Cubillana-Aguilera, M. Franco-Romano, M.L.A. Gil, I. Naranjo-Rodríguez,
under the higher molar ratio condition seems to be lower. J.L. Hidalgo-Hidalgo de Cisneros, J.M. Palacios-Santander, New, fast and green
The comparison of the reaction times required to recach the bio- procedure for the synthesis of gold nanoparticles based on sonocatalysis, Ultrason.
diesel yield of about 85% showed no significant differences among the Sonochem. 18 (2011) 789–794.
[20] Y. Kojima, H. Imazu, K. Nishida, Physical and chemical characteristics of ultra-
molar ratio conditions. As a result, it was found that the amount of sonically-prepared water-in-diesel fuel: effects of ultrasonic horn position and water
biodiesel product per units reaction time and cosumption energy was content, Ultrason. Sonochem. 21 (2) (2014) 722–728.
higher at the lower moalr ratio in the moalr ratio range of 6:1 to 18:1 [21] Y. Hirai, K. Nakabayashi, M. Kojima, M. Atobe, Size-controlled spherical polymer
nanoparticles: synthesis with tandem acoustic emulsification followed by soap-free
under ultrasonication; thus it seems reasonable to conclude that ultra- emulsion polymerization and one-step fabrication of colloidal crystal films of var-
sound is a promising method for improvement in biodiesel productivity ious colors, Ultrason. Sonochem. 21 (6) (2014) 1921–1927.
via methanolysis in the presence of solid base catalysts. [22] Y. Kojima, T. Fujita, E.P. Ona, H. Matsuda, S. Koda, N. Tanahashi, Y. Asakura,
Effects of dissolved gas species on ultrasonic degradation of (4-chloro-2-methyl-
phenoxy) acetic acid (MCPA) in aqueous solution, Ultrason. Sonochem. 12 (5)
4. Conclusions (2005) 359–365.
[23] A. Khataee, A. Karimi, S. Arefi-Oskoui, R.D.C. Soltani, Y. Hanifehpour, B. Soltani,
S.W. Joo, Sonochemical synthesis of Pr-doped ZnO nanoparticles for sonocatalytic
In a transesterification reaction of soybean oil with methanol using degradation of Acid Red 17, Ultrason. Sonochem. 22 (2015) 371–381.
heterogeneous catalyst of CaO, ultrasonication was more effective for [24] E. Ruiz, H. Rodríguez, J. Coro, V. Niebla, A. Rodríguez, R. Martínez-Alvarez,
the improvement of the production rate of biodiesel compared to tra- H. Novoa de Armas, M. Suárez, N. Martín, Efficient sonochemical synthesis of alkyl
4-aryl-6-chloro-5-formyl-2-methyl-1,4-dihydropyridine-3-carboxylate derivatives,
ditional mechanical stirring. The transesterification reaction rate of oil
Ultrason. Sonochem. 19 (2012) 221–226.
with methanol increased with the ultrasonic power and catalyst mass. [25] T.J. Mason, Industrial sonochemistry - potential and practicality, Ultrasonics 30
The higher molar ratio of methanol to oil allowed the enhancement of (1992) 192–196.
production rate of biodiesel in an oil mass transfer controlled regime. [26] W. Lauterborn, C. Ohl, Cavitation bubble dynamics, Ultrason. Sonochem. 4 (1997)
65–75.
On the contrary, however, the lower molar ratio of methanol to oil was [27] A. Tezel, S. Mitragotri, Interactions of inertial cavitation bubbles with stratum
found more favorable for the enhancement of methanolysis rate in a corneum lipid bilayers during low-frequency sonophoresis, Biophys. J. 85 (2003)
chemical reaction controlled regime. 3502–3512.
[28] L. Wolloch, J. Kost, The importance of microjet vs shock wave formation in sono-
phoresis, J. Control. Release 148 (2010) 204–211.
References [29] J. Wu, G. Du, Streaming generated by a bubble in an ultrasound field, J. Acoust.
Soc. Am. 101 (1997) 1899–1907.
[30] H.D. Hanh, N.T. Dong, K. Okitsu, Y. Maeda, R. Nishimura, Effects of molar ratio,
[1] M. Mittelbach, Process technologies for biodiesel production, in: W. Soetaert,
catalyst concentration and temperature on transesterification of triolein with
E.J. Vandamme (Eds.), Biofuel, John Wiley & Sons, Ltd., 2009, pp. 79–93.
ethanol under ultrasonic irradiation, J. Jpn. Petrol. Inst. 50 (4) (2007) 195–199.
[2] A. Singh, S.D. Fernando, Reaction kinetics of soybean oil transesterification using
[31] H.D. Hanh, N.T. Dong, C. Stavarache, K. Okitsu, Y. Maeda, R. Nishimura,
heterogeneous metal oxide catalysts, Chem. Eng. Technol. 30 (12) (2007)
Methanolysis of triolein by low frequency ultrasonic irradiation, Energy Convers.
1716–1720.
Manage. 49 (2008) 276–280.
[3] R. Abd Rabu, I. Janajreh, D. Honnery, Transesterification of waste cooking oil:
[32] H.D. Hanh, N.T. Dong, K. Okitsu, R. Nishimura, Y. Maeda, Biodiesel production
process optimization and conversion rate evaluation, Energy Convers. Manage. 65
through transesterification of triolein with various alcohols in an ultrasonic field,
(2013) 764–769.
Renew. Energy 34 (2009) 766–768.
[4] G. Vicente, M. Martinez, J. Aracil, Integrated biodiesel production: a comparison of
[33] H.D. Hanh, N.T. Dong, K. Okitsu, R. Nishimura, Y. Maeda, Biodiesel production by
different homogeneous catalysts systems, Bioresour. Technol. 92 (2004) 297–305.
esterification of oleic acid with short-chain alcohols under ultrasonic irradiation
[5] X. Liu, H. He, Y. Wang, S. Zhu, X. Piao, Transesterification of soybean oil to bio-
condition, Renew. Energy 34 (2009) 780–783.
diesel using CaO as a solid base catalyst, Fuel 87 (2008) 216–221.
[34] C. Stavarache, M. Vinatoru, R. Nishimura, Y. Maeda, Conversion of vegetable oil to
[6] T. Ebiura, T. Echizen, A. Ishikawa, K. Murai, T. Baba, Selective transesterification
biodiesel using ultrasonic irradiation, Chem. Lett. 32 (8) (2003) 716–717.
of triolein with methanol to methyl oleate and glycerol using alumina loaded with
[35] C. Stavarache, M. Vinatoru, R. Nishimura, Y. Maeda, Fatty acids methyl esters from
alkali metal salt as a solid-base catalyst, Appl. Catal. A Gen. 283 (2005) 111–115.
vegetable oil by means of ultrasonic energy, Ultrason. Sonochem. 12 (5) (2005)
[7] H. Fukuda, A. Kondo, H. Noda, Biodiesel fuel production by transesterification of
367–372.
oils, J. Biosci. Bioeng. 92 (2001) 405–416.
[36] C. Stavarache, M. Vinatoru, Y. Maeda, Ultrasonic versus silent methylation of ve-
[8] J.K. Hak, S.K. Bo, J.K. Min, M.P. Young, Transesterification of vegetable oil to
getable oils, Ultrason. Sonochem. 13 (2006) 401–407.
biodiesel using heterogeneous base catalysts, Catal. Today 93 (2004) 315–320.
[37] N.N. Mahamuni, Y.G. Adewuyi, Application of taguchi method to investigate the
[9] M.P. Dorado, E. Ballesteros, F.J. Lopez, M. Mittelbach, Optimization of alkali-cat-
effects of process parameters on the transesterification of soybean oil using high
alyzed transesterification of brassica carinata oil for biodiesel production, Energy
frequency ultrasound, Energy Fuels 24 (2010) 2120–2126.
Fuels 18 (2004) 77–83.
[38] S.M. Joshi, P.R. Gogate, S.S. Kumar, Intensification of esterification of karanja oil
[10] M.L. Granados, M.D.Z. Poves, D.M. Alonso, R. Mariscal, F.C. Galisteo, R. Moreno-
for production of biodiesel using ultrasound assisted approach with optimization
Tost, J. Santamaría, J.L.G. Fierro, Biodiesel from sunflower oil by using activated
using response surface methodology, Chem. Eng. Process. Process Intensif. 124
calcium oxide, Appl. Catal. B Environ. 73 (2007) 317–326.
(2018) 186–198.
[11] S. Gryglewicz, Rapeseed oil methyl esterd preparation using heterogeneous cata-
[39] S. Koda, T. Kimura, T. Kondo, H. Mitome, A standard method to calibrate sono-
lysts, Bioresour. Technol. 70 (1999) 249–253.
chemical efficiency of an individual reaction system, Ultrason. Sonochem. 10
[12] Z. Huaping, W.U. Zongbin, C. Yuanxiong, Z. Ping, D. Shijie, L. Xiaohua,
(2003) 149–156.
M. Zongqiang, Preparation of biodiesel catalyzed by solid super base of calcium
[40] T. Tuziuti, K. Yasui, Y. Iida, H. Taoda, S. Koda, Effect of particle addition on so-
oxide and its refining process, Chinese J. Catal. 27 (5) (2006) 391–396.
nochemical reaction, Ultrasonics 42 (1-9) (2004) 597–601.
[13] M. Kouse, M. Umemoto, T. Kasuno, M. Tajika, Y. Aihara, Y. Sugimoto, J. Hidaka,
Biodiesel production from soybean oil using calcium oxide as s heterogeneous

106

You might also like