You are on page 1of 10

Fuel 224 (2018) 489–498

Contents lists available at ScienceDirect

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

Full Length Article

Experimental and kinetic modeling of acid oil (trans)esterification in T


supercritical ethanol

Kallynca Carvalho dos Santos, Fabiane Hamerski, Fernando A. Pedersen Voll, Marcos L. Corazza
Department of Chemical Engineering, Federal University of Paraná, 81531-980 Curitiba, PR, Brazil

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

Keywords: This work reports experimental and kinetic modeling of the coupled esterification and transesterification re-
Biodiesel actions ((trans)esterification) of a model acid oil (a mixture of soybean oil and oleic acid) using ethanol at sub
Supercritical ethanol and supercritical conditions in a continuous tubular reactor. Process variables, namely temperature
Continuous process (220–350 °C), pressure (100–200 bar) and ethanol to acid oil molar ratio (2:1–20:1) as well as the effect of the
Acid oil
initial amount of free fatty acids in the raw material were studied. The best condition for the ethylic supercritical
Kinetics
(trans)esterification of soybean oil with 20 wt% of oleic acid was 280 °C, 100 bar, and ethanol to acid oil molar
ratio of 4:1, where 80% of ester yield was reached in 60 min. It was observed that increasing the free fatty acid
content increased the reaction rate and reaction yield. The proposed autocatalytic kinetic model presented a
good fit in relation to the experimental data.

1. Introduction the reaction is carried out in two steps, ester yields around 90% are
obtained in about 60 min of reaction using homogeneous catalysts for
Biodiesel is considered as one of the main energy source substitutes both reactions (esterification as pre-treatment and transesterification)
for the fossil fuels. In addition to the environmental benefits, one of the [1,9–12].The disadvantages of a homogeneous catalysis process applied
main advantages of biodiesel is the similarity of its properties to the to biodiesel from acid raw material are the large volumes of neu-
petrodiesel, so few modifications are required in the current engine tralization residues, additional products purification steps. The use of a
systems [1]. The conventional method for biodiesel production is the heterogeneous catalyst can overcome these disadvantages, but it needs
transesterification of triacylglycerols with a short chain alcohol, for catalysts capable to deal with such raw materials. The heterogeneous
example, methanol and ethanol, in the presence of alkali catalyst (KOH catalysts are less corrosive, easy to separate and reusable [13]. Many
and NaOH) generating glycerol as a by-product [2–4]. However, alka- acid catalysts can catalyze simultaneous the reactions involved in acid
line transesterification is only suitable for raw materials containing low oils [6,13,14]. From the results presented by Konwar et al. [6], Zhang
free fatty acids (FFA) content (lower than 1 wt%), such as refined ve- et al. [13] and Boz et al. [14] for example, those authors observed the
getable oils [1,5]. This specificity of the raw material makes the bio- versatility of this process, since raw materials with a free fatty acid
diesel production not economically competitive with the petroleum- content within 5–40% were tested. However, a common feature re-
based fuel. garding these heterogeneous catalysts is the long reaction times re-
An alternative to overcome this disadvantage is the use of cheaper quired to obtain high conversions (higher than 8 h).
and low-quality raw materials, such as residual oils, non-edible vege- Transesterification and esterification (biodiesel production) at su-
table oils and animal fats [6]. The use of these oils also can reduce the percritical conditions have been proposed as a potential technology to
wastes with a potential hazard to the environment [7]. These acid oils biodiesel production, in which it is possible to produce alkyl esters with
normally present FFA content above the limit allowed for the transes- high yields in short reaction times. This technology is more tolerant to
terification, what leads to undesirable reactions decreasing the yields of water (moisture) and free fatty acids contents present in the raw ma-
the desired product [8]. terial than other reaction systems [8]. In addition, either transester-
The residual raw material can be processed in a two steps process ification or esterification at supercritical non-catalyzed reaction sys-
(acid esterification as a pretreatment followed by alkaline transester- tems eliminates the need for a catalyst simplifying the purification steps
ification), in a single reaction in the presence of heterogeneous acid [6,15–17].
catalysts or performing the reaction at supercritical conditions. When


Corresponding author at: Department of Chemical Engineering, Federal University of Paraná, PO Box 19011, Polytechnic Center, Curitiba 81531-980, PR, Brazil.
E-mail address: corazza@ufpr.br (M.L. Corazza).

https://doi.org/10.1016/j.fuel.2018.03.102
Received 17 January 2018; Received in revised form 13 March 2018; Accepted 14 March 2018
0016-2361/ © 2018 Elsevier Ltd. All rights reserved.
K.C. dos Santos et al. Fuel 224 (2018) 489–498

1.1. Transesterification reaction at supercritical conditions esterification of a palm fatty acid distillate. They obtained ester yields
above 95% in only 5 min of reaction with 1 wt% of the catalyst. Santos
Saka and Kusdiana [18] were the pioneers of the supercritical et al. [34] evaluated the experimental and kinetic modeling of super-
method aiming the alkyl esters production. They showed that this critical ethyl esterification of fatty acids with different chain lengths.
technology has a reaction rate up to 15 times faster than the alkaline They considered that the reaction is self-catalyzed by the free fatty acids
transesterification, proving the efficiency of this new technology for themselves, and this consideration presented a good prediction of the
producing biodiesel. Supercritical technology is a non-catalytic route, model in relation to the experimental data.
which occurs at high temperatures and pressures (above the critical Some studies show the efficiency of supercritical technology to si-
point of the solvent). Under these conditions, the thermophysical multaneously convert the triglycerides and fatty acids to ester, i.e., in
properties of the alcohol, such as dielectric constant, viscosity and po- the same reactor and condition it is possible to perform the transes-
larity are significantly changed and can be tuned by manipulating the terification and esterification reaction simultaneously, allowing the use
pressure and temperature of the system [17,19]. However, a major of residual raw materials [7,19]. Tsai et al. [22] processed waste
concerning is the thermal degradation of the alkyl esters at high tem- cooking oil in a supercritical reactor and concluded that the presence of
perature, mainly polyunsaturated fatty acids. free fatty acid in the raw material contributed positively to enhance the
Silva et al. [20] studied the thermal degradation of biodiesel pro- ester yield. In that case, the fatty acid has a dual function: reactant of
duced from soybean oil and ethanol. Using thermogravimetric analysis, the esterification reaction and the catalyst of the transesterification
they proved that soybean oil biodiesel is unstable and decompose at reaction. However, the literature is scarce regarding the kinetic mod-
temperatures above 300 °C. As mentioned by Liu et al. [21] and Lin eling involving reactions of acid oils with different free fatty acid
et al. [22], main decomposition reactions consist of isomerization, contents, which is very important in the process modeling, analysis,
polymerization, and pyrolysis. Ethyl fatty acid esters are relatively design and optimization.
stable up to 275 °C, with the pyrolysis effect being pronounced between In the biodiesel production, short chain alcohols such as methanol
300 and 325 °C [21,23]. Other researchers have shown that even more and ethanol are generally used. Most of the works have used methanol
parallel reactions are involved at supercritical conditions, such as gly- due to its low cost and higher reactivity [35,37]. However, methanol is
cerol dehydration and other less favorable reactions (decarboxylation, derived from fossil fuels and its use contradicts the main objective of
methoxylation, and reformation) [24,25] that generates isomers, biodiesel production, which is to produce a sustainable fuel. For bio-
polymers, short chain ester, light hydrocarbons and gases, carboxylic diesel to be considered a fully renewable fuel it is necessary to use a bio-
acids, long chain alkenes, acrolein, acetol and other compounds. Pala- alcohol, such as ethanol. Other advantages of ethanol are its higher
cios-Nereo et al. [26] proposed to proceed the reaction using a heating miscibility with the oil and the ethyl esters present improved cold flow
rate for preventing the thermal decomposition of the reactants and properties, lower greenhouse gases emission, they are more biode-
products. In their study, those authors performed three types of reactor gradable and present higher lubricity. The disadvantage of the ethyl
heating: constant temperature reaction (CTR), gradual heating reaction ester compared to the methyl ester is the higher acid values.
(GHR), both with a reaction time of 75 min, and hybrid reaction (HR) [21,38–40].
combining gradual heating and a phase at constant temperature tota- Glisic and Orlović [16] presented a review considering the costs
lizing reaction time between 65 and 345 min. At CTR they observed a involved in the biodiesel production and for its reduction two factors
decrease in the ester yield in reactions above 320 °C. With GHR, de- must be taken into account, reduction in raw material cost and energy
gradation was not observed, but low yields were obtained. With the HR consumption. The raw material accounts for about 60–80% of the total
method, the authors obtained yields of 99.9% in 345 min at high tem- biodiesel production cost. The supercritical process presents the cap-
peratures. They evaluated that a heating ramp can minimize the de- ability to process low added value raw materials with different content
composition of esters. of free fatty acids and water. According to Glisic and Orlović [16], the
The supercritical transesterification reaction for different oils, such economy in the raw material compensates the energy costs to set the
as sunflower, palm, and soybean oil has been reported by different system under the supercritical conditions of the alcohol. The energy of
authors [17,26–28]. Optimum conditions have been normally reported the system can be optimized from the integration and optimization
to be within 325–350 °C, 160–350 bar and alcohol to oil molar ratio involving cold and hot streams and energy supply to the reboiler in
33:1–43:1, and the ester yields reached were around 80–90%. Tan et al. distillation columns. Glisic and Orlović [16] found that the biodiesel
[29] evaluated the transesterification reaction of palm oil in super- production under supercritical conditions with high molar ratio (42:1)
critical methanol (SCM) and supercritical ethanol (SCE). In their study, requires the same amount of energy as a conventional alkaline trans-
they observed that for the same yield, SCE needs more time than SCM, esterification process.
but the reaction conditions (temperature, pressure and molar ratio) are In general, it can be observed that the literature has pointed out the
milder. In relation to the kinetics of the supercritical transesterification need for high ethanol to oil molar ratios (normally ranged from 25:1 to
reaction, some authors considered the overall reaction and others have 42:1) to obtain high ester yields [22,24,28,35]. The main reason pre-
used all the three reactions, either reversible or irreversible sented is generally the low solubility of triacylglycerols in short chain
[22,28,30–32]. According to these authors, all these kinetic modeling alcohols. In the case of acid raw materials (acid oils), the free fatty acids
approaches present good fit to the experimental data. (FFA) can act as a co-solvent promoting miscibility between the alcohol
and oil. In this case, the FFA might present a similar role of alkyl ester
1.2. Esterification reaction at supercritical conditions in the ternary system short chain alcohol + oil + solute (alkyl ester/
FFA) presented by Dagostin et al. [41], favoring the miscibility of the
The optimal conditions of the supercritical esterification reaction mixture under ambient conditions and reducing the need for the use of
are milder than the supercritical transesterification. As presented in the high molar ratios.
literature, supercritical esterification normally requires molar ration of Aiming to understand and improve the biodiesel production from
alcohol to acid around 6:1, temperatures below 300 °C and pressures low-cost raw material and using optimized amounts of ethanol, this
around 100 bar. Under such conditions, free fatty acid conversions of study is focused on the coupled esterification and transesterification
90% can be obtained in short times (30–50 min) [33,34]. This occurs reactions of a model acid oil (soybean oil doped with oleic acid) in a
because esterification has higher reaction rates compared to transes- continuous tubular reactor at low ethanol to acid oil molar ratio. The
terification [35]. Moreover, these results can be enhanced by the use of influence of different reaction parameters, such as temperature, pres-
catalysts, as shown by Lokman et al. [36]. These authors used carbo- sure, initial free fatty acid content in the oil (Ac0) and ethanol to acid
hydrate derived solid acid catalyst in the supercritical methyl oil molar ratio (Et:AO) on biodiesel yield were studied. A kinetic

490
K.C. dos Santos et al. Fuel 224 (2018) 489–498

modeling able to predict the supercritical (trans)esterification of acid amounts of each reactant, which was stirred using a magnetic stirrer
oil with ethanol is proposed to evaluate the reactions involved in this and thus pumped into the reactor. The molar ratio was calculated ac-
system and its parameters (activation energy and pre-exponential cording to the Eq. (1).
factor) were adjusted.
mol of ethanol
Et: AO =
mol of triacylglycerol + mol of oleic acid (1)
2. Material and methods
Before mixing the acid oil and the ethanol, the reactor was first fed
2.1. Materials with pure ethanol until the system reached the conditions of tempera-
ture (220–300 °C) and pressure (100 bar). Therefore, the reaction
Commercial refined soybean oil (Liza) was purchased in a local mixture was pumped at fixed volumetric flow (5, 2, 1, 0.5 and
market in Curitiba (state of Paraná, Brazil). Ethanol (99.8%, CAS 0.2 cm3.min−1) and samples were collected at the reactor outlet after
number 64-17-5) and oleic acid (90%, CAS number 112-80-1) were reaching twice the estimated residence time for ensuring the system
purchased from Neon and Sigma-Aldrich, respectively. reached the steady-state. For each residence time, samples were col-
Fatty acid profile of soybean oil was performed using a Shimadzu lected for quantifications of water, glycerol, and free fatty acids by ti-
chromatograph (GC 2010 Plus), a capillary column (CP-Wax 58 FFAP tration techniques. Another sample was collected and dried (12 h at
CB, 50 m × 0.25 mm × 0.20 □m), flame ionization detector (FID) and 60 °C and further 2 h at 100 °C) for the quantification of fatty acid ethyl
split injection mode (1:10). The injector and detector temperatures esters and acylglycerols by gas chromatography.
were 250 °C and 280 °C, respectively. The oven temperature was pro-
grammed to increase from 100 °C to 175 °C at a rate of 25 °C.min−1. 2.3. Analytical methods
After that, the temperature was increased to 230 °C at a rate of
4 °C.min−1 and maintained at 230 °C for 15 min. The carrier gas was The water content of the reactants and samples was determined by
Helium at a flow rate of 22.8 cm3.min−1. The sample was prepared volumetric and coulometric titrations using automatic Karl Fischer ti-
according to the official method (Ce 2-66) of American Oil Chemists’ trators, KEM MKA-610 and MKC-610 models, respectively. The FFA
Society (AOCS) to convert triacylglycerol and free fatty acids of soy- quantification was performed by titration according to the method of
bean oil into fatty acid methyl esters (FAMEs). FAMEs were identified American Oil Chemist’s Society (AOCS) official method Ca-5a-40. The
by comparison with retention times of a standard mixture of FAMEs glycerol analysis was performed by the spectrophotometric method, as
(Supelco, MIX FAME 37, St. Louis, MO 63103, USA). Results were ex- proposed by Bondioli and Bella [42].
pressed as a percentage of each individual fatty acid present in the Fatty acid ethyl esters (FAEE), mono, di and triacylglycerols were
sample. quantified by gas chromatography (GC-2010 Plus, Shimadzu
Corporation) equipped with a Select Biodiesel column (Agilent,
2.2. Experimental procedure 15 m × 0.32 mm × 0.10 □m), flame ionization detector (FID) and split
injection mode (1:10). The injector and detector temperatures were
Reactions of acid oil at sub and supercritical ethanol conditions 380 °C and 400 °C, respectively. The initial oven was 50 C and it was
were performed in a bench-scale continuous tubular reactor. The programmed with the following heating ramps 15 °C/min up to 180 °C,
system consists of an HPLC pump (Shimadzu, model LC-20AT), stainless 7 °C/min up to 230 °C, and 10 °C/min up to 380 °C, remaining for 6 min.
steel tubing (316, 1/16″ OD × 0.91, Swagelok) for preheating the Helium was used as the carrier gas at a flow rate of 29.2 cm3.min−1.
mixture, stainless steel tube reactor (316, 1/8″ OD × 0.91, Swagelok) Samples were prepared according to the European standard EN 14105.
with 22 cm3 of capacity, an electrical furnace (Sanchis Fornos Solvent and derivatizing agent used were heptane and N-methyl-N-tri-
Industriais, Brazil) with maximum operating temperature of 400 °C, methylsilyl trifluoroacetamide (MSTFA), respectively. The internal
heat exchanger for cooling the reactor outlet stream. Fig. 1 depicts the standards were lauric acid, methyl palmitate and tricaprin, corre-
schematic diagram of the experimental set up used for the reactions sponding to free fatty acids, fatty acid ethyl ester and mono-di-tria-
evaluated in this work. The pressure in the reactor was controlled by a cylglycerols, respectively.
backpressure valve at the reactor outlet. The temperature and pressure The reaction was evaluated in terms of ester yield ( yieldFAEE ) ac-
were measured by four K-type thermocouples and a pressure transducer cording to the equation:
(Huba Control), respectively, connected in a field logger register
FAEEt
(Novus). yieldFAEE (%) = 100
3TAGI + FFAI (2)
Soybean oil with different initial free fatty acid contents of 5, 20 and
35 wt% of oleic acid (Ac0) and ethanol at different initial ethanol to where TAGI, FFAI refers to the number of moles of triacylglycerols and
acid oil molar ratio (Et:AO) (2:1–20:1) were prepared by mixing known free fatty acids at the inlet of the reactor, respectively, while FAEEt

Fig. 1. Schematic diagram of the experimental apparatus used for esterification and transesterification of acid oil at supercritical conditions.

491
K.C. dos Santos et al. Fuel 224 (2018) 489–498

M Nj
represents the number of moles of fatty acid ethyl esters at the reactor
outlet. OF = ∑∑ (Yijtexp−Yijtcalc )2
j t=1 (14)

2.4. Kinetic modeling In this equation, and Yijtexp Yijtcalc


are the experimental and calculated
content values (wt%) of each component at the experimental point t for
Kinetic of acid oil reactions with ethanol at sub and supercritical each kinetic curve j , M is the number of kinetic curves and Nj is the
conditions of alcohol (Tc = 240.75 °C and Pc = 61.48 bar) were mod- number of experimental data for each kinetic curve. The components i
eled considering the three soybean oil transesterification reactions (Eqs. used in the objective function were DAG, MAG, FFA and FAEE. To
(3)(5)) and the esterification of acid oleic (Eq. (6)). All reactions were evaluate the correlation between the model calculations and the ex-
considered reversible, elementary and self-catalytic by the free fatty perimental data was used the root mean square deviation (rmsd ), Eq.
acid present in the reactant mixture, as shown below: (15), where NOBS is the number of total experimental runs.
k1 M Nj
TAG + ET + FFA ⥫
⥬ FAEE + DAG + FFA
k−1 (3) ∑∑ (Yijtexp−Yijtcalc )2
j t=1
k2
rmsd =
NOBS (15)
DAG + ET + FFA ⥫
⥬ FAEE + MAG + FFA
k−2 (4)
The differential equations that describe the set of reactions in the
k3 tubular reactor according to the proposed mechanism for all compo-
MAG + ET + FFA ⥫
⥬ FAEE + GLY + FFA
k−3 (5) nents are the following:

k4 dCTAG
FFA + ET + FFA ⥫
⥬ FAEE + W + FFA = −r1
dτ (16)
k−4 (6)

where TAG , DAG , MAG , ET , W , FAEE , FFA and GLY represent tria- dCDAG
= r1−r2
dτ (17)
cylglycerol, diacylglycerol, monoacylglycerol, ethanol, water, fatty acid
ethyl ester, free fatty acid and glycerol, respectively, ki and k−i are the dCMAG
forward and backward kinetic rate constants of each reaction i . = r2−r3
dτ (18)
The mass-balance in the continuous tubular reactor for a generic
component A (Eq. (7)) and the reaction rates for each reaction (Eqs. (8) dCGLY
= r3
(11)) in terms of molar concentrations were written as follows: dτ (19)

dCA dCFAEE
= rA = r1 + r2 + r3 + r4
dτ (7) dτ (20)

1 dCET
r1 = k1 ⎛⎜CTAG CET CFFA− CDAG CFAEE CFFA⎞⎟ = −r1−r2−r3−r4
dτ (21)
⎝ keq,1 ⎠ (8)
dCFFA
= −r4
1 dτ (22)
r2 = k2 ⎜⎛CDAG CET CFFA− CMAG CFAEE CFFA⎞⎟
⎝ k eq,2 ⎠ (9) dCW
= r4
dτ (23)
1
r3 = k3 ⎜⎛CMAG CET CFFA− CGLY CFAEE CFFA⎞⎟
keq,3 (10) In this work, the residence time was calculated as proposed by
⎝ ⎠
Santos et al. [34]:
1 VR q̇
r4 = k 4 ⎛⎜CFFA CET CFFA− CW CFAEE CFFA⎞⎟ τ= with υ =
⎝ k eq,4 ⎠ (11) υ ρmix ,R (24)
−3
where Cj are the molar concentration of each component j [mol.cm ], where VR is the reactor volume (22 cm3), υ is the volumetric flow, q̇ is
τ is the residence time [min], ri are the reaction rates of each reaction i the mass flow measured at the end of the reaction of each residence
[mol.min−1.cm−3] and keq,i are the equilibrium constants of each re- time [g.min−1] and ρmix ,R is the density of the initial mixture at reactor
action i , which are described by the ratio between forward and back- temperature and pressure conditions [g.cm−3]. PC-SAFT state equation
ward kinetic rate constants (Eq. (12)). was used to determine the density of the triolein, representing the
soybean oil and oleic acid. PC-SAFT parameters of pure components
ki
keq,i = were taken from the literature [34,44]. The accurate equation of state
k−i (12)
proposed by Dillon and Penoncello [45] was used for the density of
Forward and backward kinetic constants of all self-catalyzed reac- ethanol calculations at each reaction condition. The mixture density
tions were calculated by an Arrhenius-type expression, as indicated in was calculated considering an ideal solution. The kinetic model was
Eq. (13). implemented in Matlab® and the ode23s subroutine was used to solve
the differential equations system (Eqs. (16)(23)).
−E
ki = 10 ki0exp ⎛ ai ⎞
⎝ RT ⎠ (13)
3. Results and discussion
3 −1 2 −1
Pre-exponential factors ki0 [(cm .mol ) .min ], and the activa-
tion energies Eai [kJ.mol−1] of each reaction i were fitted using the The characterization of soybean oil used in the experiments pre-
experimental kinetic data measured. For the kinetic parameters esti- sented the following fatty acid profile: linoleic acid (56.86%), oleic acid
mation, the objective function (OF ) (Eq. (14)) was initially minimized (21.57%), palmitic acid (10.94%), linolenic acid (6.15%), stearic acid
using the stochastic Particle Swarm Optimization (PSO) algorithm [43] (3.31%) and 1.17% others.
and thus the solution was refined using the fminsearch subroutine, Table 1 shows the experiments performed at different conditions of
available in the MATLAB® software. temperature (T), pressure (P), initial free fatty acid content in the

492
K.C. dos Santos et al. Fuel 224 (2018) 489–498

Table 1
Experimental conditions for simultaneous esterification and transesterification of acid soybean oil in ethanol.

Run T [°C] P [bar] Ac0 [wt%] Et:AO Mole fraction** ρmix,R [g.cm−3]

Ethanol Oil Oleic acid

1 283 ± 1 102 ± 3 20 2:1 0.6994 0.1697 0.1308 0.5749


2* 278 ± 1 103 ± 1 22 4:1 0.7909 0.1125 0.0965 0.5407
3 283 ± 2 102 ± 2 20 5:1 0.8238 0.0995 0.0768 0.4804
4 280 ± 1 100 ± 1 22 8:1 0.8909 0.0590 0.0500 0.4157
5 280 ± 1 101 ± 1 20 20:1 0.9524 0.0269 0.0208 0.3326
6* 221 ± 3 100 ± 2 22 4:1 0.7864 0.1151 0.0985 0.7221
7* 249 ± 1 100 ± 2 20 4:1 0.7783 0.1252 0.0965 0.6782
8* 299 ± 1 102 ± 2 20 4:1 0.7787 0.1249 0.0964 0.4453
9 279 ± 1 202 ± 1 20 4:1 0.7912 0.1178 0.0910 0.6832
10 280 ± 1 100 ± 2 5 4:1 0.7909 0.1799 0.0292 0.5445
11 280 ± 1 102 ± 3 35 4:1 0.7908 0.0786 0.1306 0.4996
12 319 ± 2 101 ± 1 22 4:1 0.7902 0.1129 0.0968 0.3764
13 321 ± 1 102 ± 1 22 4:1 0.7896 0.1133 0.0971 0.3773
14 350 ± 1 101 ± 2 22 4:1 0.7897 0.1132 0.0971 0.3323

*
Data set used in the parameters estimation procedure; **Uncertainty in mole fraction u(x) = 0.0003. Legend: Ac0 refers to the initial free fatty acid content in oil; Et:AO is the ethanol to
acid oil molar ratio and ρmix,R is density of the reactant mixture estimated at average conditions in the inlet of the reactor.

reactant oil (A), ethanol to acid oil molar ratio (Et:AO), initial molar compared in terms of magnitude due to the differences in the raw
fraction of components and the density of the mixture at the initial materials, molar ratio and reactor type used in the different works.
conditions (inlet reactor). Temperature and the pressure values are Figs. 2 and 3 show the comparison between experimental and calcu-
presented by the average and standard deviation of collected data along lated values by the kinetic model regarding the components profiles
each experiment. and ester yield at different reaction conditions of runs 2, 6, 7 and 8. For
Reactions at temperatures above 300 °C were not used in the kinetic the parameters estimation, the root mean square deviation (rmsd)
modeling due to the possible occurrence of degradation reactions, as considering the experimental and calculated content values in terms of
suggested in the literature [20,21,23–25]. The kinetic parameters were DAG, MAG, FFA, and FAEE was 3.18 wt%. From the Figs. 2 and 3, and
adjusted using four kinetic set varying the temperature from 220 to the low value of rmsd obtained, it is possible to observe that the kinetic
300 °C (Runs 2, 6, 7 and 8 in Table 1) and fixing the initial oleic acid model proposed was capable to represent the kinetic behavior con-
content in the soybean oil around 20 wt%, pressure of 100 bar, and cerning the esterification and transesterification reactions of acid oil in
ethanol to acid oil molar ratio of 4:1. The pre-exponential factors and ethanol.
the activation energies adjusted are shown in Table 2. The methanol The effect of the temperature on the kinetic and ester yields was
transesterification reaction in the range of 200–320 °C for different oils, evaluated at different temperatures varied from 220 to 300 °C with an
sunflower oil [22] and soybean oil [46], as reported by those authors, ethanol to acid oil ratio of 4:1, 100 bar and initial free fatty acid content
presented activation energy of 85.5 and 92.52 kJ.mol−1, respectively. around 20 wt% (Figs. 2 and 3). The reaction rate increases with in-
While supercritical methanol and ethanol esterification studies [34,47] creasing temperature. The ester yield (Fig. 3) increased from 58% to
obtained activation energy between 72 and 93.4 kJ.mol−1, respec- 82% in the temperature range from 200 to 300 °C. Wang et al. [49]
tively. Farobie et al. [48] studied the transesterification of canola oil in studied the transesterification of crude rapeseed oil in supercritical
supercritical ethanol and considered the three transesterification reac- methanol at condition of 250 °C, 100 bar and methanol to oil molar
tions in the kinetic modeling. The activation energy obtained in our ratio of 36:1 and they obtained 10% yield in ester in 20 min of reaction
work are like the values reported in the literature, at least with same without a catalyst. When comparing this result with the kinetics at
magnitude. The activation energy of the direct and inverse reactions 250 °C (our work), it can be observed an ester yield of 4 times higher
were approximately the same value for the DAG conversion reaction, as and with the use of lower molar ratio. From Fig. 2, it is possible to note
reported by Farobie et al. [48], and in our study this occurred for the that the mass profiles that present the highest temperature effect are the
MAG reaction. However, the work of Farobie et al. [48] presented some diacylglycerol and free fatty acids components. This behavior is justi-
contradictions with the literature, as they can achieve maximum ester fied by the enthalpy change of reaction for each individual reaction
yields in reactions at 400 °C and the literature has mentioned the oc- assumed in the kinetic modeling, since the conversion reaction of DAG
currence of degradation reactions at lower temperatures [20,21,23]. and FFA were the reactions with higher enthalpies, as it can be ob-
The activation energy values obtained in this work are in agreement served by the values presented as supplementary material (Section S2).
with some similar studies presented in the literature. The values were At both two higher temperatures (280 and 300 °C, Fig. 2c and d, re-
spectively), practically all triacylglycerol, diacylglycerol and free fatty
acid were consumed at the end of the reaction and monoacylglycerol is
Table 2 in equilibrium. And, in Fig. 3 it is observed that the ester yield gain is
Kinetic parameters adjusted for the acid soybean oil (trans)esterification in ethanol
small as the temperature increases from 280 to 300 °C. Thus, these
(temperature range: 220–300 °C, P = 100 bar, Ac0 = 20 wt% and Et:AO = 4:1).
conclusions indicate that the optimum operating condition is 280 °C.
i ki0 [(cm3.mol−1)2.min−1] Eai [kJ.mol−1] The kinetic parameters obtained using the set of experimental data
from runs 2, 6, 7 and 8 (Table 1) were used to predict the kinetic re-
1 8.48 94.96 actions at different conditions (Runs 1, 3, 4, 5, 9, 10 and 11 in Table 1)
−1 0.50 62.40
2 8.81 97.50
to access the capability of the mathematical model in predicting the
−2 1.18 30.75 esterification and transesterification of acid oil with ethanol. From
3 8.77 97.09 Figs. 4–9 can be observed that the model was also able to predict well
−3 6.84 80.21 the kinetic profiles and the overall ester yields for different reaction
4 7.27 84.37
conditions than those used in the kinetic parameter estimation.
−4 0.40 20.90
It is worth mentioning that TAG, glycerol and water contents were

493
K.C. dos Santos et al. Fuel 224 (2018) 489–498

Fig. 2. Comparison between experimental data (symbols) and kinetic model (lines) of TAG (□, ), DAG (+, ), MAG (○, ), FFA ( , ), FAEE ( ,
) and GLY ( , ) for the (trans)esterification of acid oil with Et:AO = 4:1 (A) 221 °C, 100 bar, Ac0 = 22 wt% (Run 6 in Table 1), (B) 249 °C, 100 bar, Ac0 = 20 wt%
(Run 7 in Table 1), (C) 278 °C, 103 bar, Ac0 = 22 wt% (Run 2 in Table 1) and (D) 299 °C, 102 bar, Ac0 = 20 wt% (Run 8 in Table 1).

no considered in the kinetic parameters estimation. All results re-


garding these compounds are merely predictions from the mathema-
tical model and its comparison with the experimental values measured
for this compounds. In the kinetic modeling, hydrolysis of TAG, DAG
and MAG were not considered due to the low free water content of
reactants. When evaluating the water content profile along the reac-
tions in the tables presented in the supplementary material (Tables
S1–S14), it is observed that the water amount produced is low and its
content during the reaction did not reduce, i.e., water is not consumed
as a reagent for the hydrolysis reactions. Another hypothesis assumed in
kinetic modeling is that the reaction is catalyzed by the free fatty acid
[34]. All these hypotheses contributed to the good performance and
predictions obtained from the kinetic model proposed.
The literature [22] have mentioned that the fatty acids for this
system have two functions: reactant of the esterification reaction and
the catalyst of the transesterification reaction. This duality of fatty acids
is observed by the results presented in Figs. 2c, 4 and 5. These figures
depict the high influence of the initial free fatty acid content in the
Fig. 3. Comparison between experimental data (symbols) and kinetic model (lines) ester
reactant over the (trans)esterification reactions at conditions around
(FAEE) yield for the (trans)esterification of acid oil with ethanol at different tempera- 280 °C, 100 bar and Et:AO of 4:1. Fig. 5 shows that ester yields of up to
tures: 220 °C (□, ); 250 °C ( , ); 280 °C ( , ) and 300 °C (△, 90% can be achieved with 35 wt% of free fatty acid present in the oil.
) (Run 2, 6, 7 and 8 in Table 1). The increase in the free fatty acid content of the raw material increased
the reaction rate, being able to consume all triacylglycerol up to around

494
K.C. dos Santos et al. Fuel 224 (2018) 489–498

Fig. 4. Comparison between experimental data (symbols) and kinetic model (lines) profile of TAG (□, ), DAG (+, ), MAG (○, ), FFA ( , ), FAEE ( ,
) and GLY ( , ) for the (trans)esterification of acid oil with Et:AO = 4:1 (A) 280 °C, 100 bar, Ac0 = 5 wt% (Run 10 in Table 1) and (B) 280 °C, 102 bar,
Ac0 = 35 wt% (Run 11 in Table 1).

efficiency of the transesterification and esterification reactions carried


out with low proportions of alcohol.
Figs. 2c, 8 and 9 present the effect of pressure on the (trans)ester-
ification reaction of acid oil in ethanol. It can be noted that pressure has
no significant effects on ester yield varying this factor from 100 bar to
200 bar because the compressibility of the system is low due to the high
concentration of free fatty acid and oil in ethanol. In systems with
higher alcohol to oil, e.g. 40:1, the effect of the pressure is much more
pronounced [7,52]. In addition, in the systems with high alcohol con-
centration, the density variation of the mixture is determined by the
properties of the alcohol. Even though using an approximate and
completely predictive approach for the density of the mixture (PC-SAFT
equation of state [34,44] used to estimate the density of pure soybean
oil, represented by the triolein molecule as a model, and pure oleic acid,
and ethanol density was estimated by the accurate equation of state of
Dillon and Penoncello [45] and thus the mixture density was calculated
assuming an ideal solution) the model proposed was capable to predict
the kinetic behavior at 200 bar of pressure (Fig. 8). Fig. 9 indicates that
Fig. 5. Comparison between experimental data (symbols) and kinetic model (lines) ester
the increase in pressure positively influences the initial reaction rate,
(FAEE) yield for the (trans)esterification of acid oil with ethanol at different initial free but the ester yield reached at the end of the reaction is practically the
fatty acid contents (Ac0) in the reactant oil: 5 wt% (□, ); 20 wt% ( , ) same for both pressures. The small gain of the ester yield does not
and 35 wt% ( , ) (Run 2, 10 and 11 in Table 1). compensate for double the pressure.

60 min (Fig. 4). It is also possible to observe that the kinetic model can
predict well the ester yields for systems with a wide range of initial free 3.1. Thermal degradation
fatty acid content.
The ethanol to oil molar ratio is the parameter that presents the A matter of concern in these reactions at supercritical conditions for
greatest differences in the literature, where its optimum conditions vary alkyl ester production is the degradation that both the raw material and
from 12:1 to 45:1 [22,24,28,35,50,51]. The effect of ethanol to acid oil products. As mentioned earlier in this paper, some authors [20,21,23]
molar ratio in the mass profile and ester yield are accessed by com- have presented different results about the ethyl ester yield and de-
paring the results in Figs. 2c, 6 and 7. For the lowest molar ratio tested gradation at temperatures above 300 °C. In general, they agreed that it
(Et:AO = 2:1) the lowest yield was obtained. However, for Et:AO of is unstable between 275 and 325 °C. Polyunsaturated esters, a higher
20:1 the initial reaction rate was the lowest, but at the end of the re- percentage of fatty acids in soybean oil, undergo thermal decomposi-
tion at 300 °C, while monounsaturated FAMEs are relatively stable at
action the ester yield was close to that obtained for molar ratios of
4:1–8:1. From Fig. 7, it is possible to note a similarity in the kinetics of this temperature [53]. However, some studies point out supercritical
transesterification reactions in temperatures above 300 °C and with
the intermediate molar ratios. Therefore, the ethanol to acid oil molar
ratio of 4:1 is found to be an optimized condition to achieve the best high yields. For example, Palacios-Nereo et al. [26] reached 90% of
possible ester yield. In the study of supercritical (trans)esterification ester yield at 320 °C in the soybean oil supercritical transesterification
reaction of soybean fried oil, performed by Gonzalez et al. [7], 75% reaction. At temperatures above 320 °C, they also noticed the de-
ester yield was obtained for a methanol to oil molar ratio of 40:1. In the gradation. Fig. 10 depicts the ester yield profile along the (trans)es-
present study, we have reported ester yield of 80% with ethanol to acid terification reaction of acid oil conducted at different temperatures
oil molar ratio ten times smaller (Fig. 7). This result proves the (220–350 °C) in the presented study. As expected, the temperature has a
positive influence on the yield up to 300 °C, however at temperatures

495
K.C. dos Santos et al. Fuel 224 (2018) 489–498

Fig. 6. Comparison between experimental data (symbols) and kinetic model (lines) profile of TAG (□, ), DAG (+, ), MAG (○, ), FFA ( , ), FAEE ( ,
) and GLY ( , ) for the (trans)esterification of acid oil with ethanol (A) 283 °C, 102 bar, Ac0 = 20 wt%, Et:AO = 2:1 (Run 1 in Table 1), (B) 283 °C, 102 bar,
Ac0 = 20 wt%, Et:AO = 5:1 (Run 3 in Table 1), (C) 280 °C, 100 bar, Ac0 = 22 wt%, Et:AO = 8:1 (Run 4 in Table 1) and (D) 280 °C, 101 bar, Ac0 = 20 wt%, Et:AO = 20:1 (Run 5 in
Table 1).

Fig. 7. Comparison between experimental data (symbols) and kinetic model (lines) ester Fig. 8. Comparison between experimental data (symbols) and kinetic model (lines)
(FAEE) yield for the (trans)esterification of acid oil with ethanol at different ethanol to profile of TAG (□, ), DAG (+, ), MAG (○, ), FFA ( , ),
acid oil molar ratios of 2:1 (□, ); 4:1 ( , ); 5:1 (△, ); 8:1 ( ,
FAEE ( , ) and GLY ( , ) for the (trans)esterification of acid oil
) and 20:1 ( , ) (Run 1–5 in Table 1). with Et:AO = 4:1 at 279 °C, 202 bar, Ac0 = 20 wt% (Run 9 in Table 1).

496
K.C. dos Santos et al. Fuel 224 (2018) 489–498

higher than that the ester degradation is evidenced by the decrease in


the overall ester yield measured, corroborating the results presented in
the literature. And the degradation is much more pronounced at 350 °C
than at 320 °C.
The experimental reproducibility was evaluated for the condition of
320 °C, 100 bar, 20 wt% of oleic acid in the oil (Ac0) and ethanol to acid
oil molar ratio (Et:AO) of 4:1 (Run 12 and 13 in Table 1) and the results
are showed in Fig. 11 as the average and standard deviation for the
ester yield and the residence time (different flow rates into the reactor).
The low deviations confirm the good experimental reproducibility. On
the other hand, the residence time presented a large deviation only for
the last experimental point, where the pump operates at a low feed rate.
Calculation of the reaction time uses the mass flow rate at the reactor
outlet, in which the sample is collected in one minute timed.
When evaluating the experimental results, it was verified that the
best condition for the supercritical esterification/transesterification
reaction of acid oil (20 wt% free fatty acid) in ethanol is the condition
of 280 °C, 100 bar and ethanol to acid oil of 4:1, in which an ester yield
Fig. 9. Comparison between experimental data (symbols) and kinetic model (lines) ester
around 85% was achieved in 60 min of reaction (Fig. 7). This condition
(FAEE) yield for the (trans)esterification of acid oil with ethanol at different pressures: shows the viability of using lower ethanol to the acid oil molar ratios
100 bar ( , ) and 200 bar (×, ) (Runs 2 and 9 in Table 1). those presented in the literature to reach high fatty acid ethyl ester
content from acid oils (waste materials) that can be mandatory to the
economic biodiesel production.

4. Conclusions

The esterification and transesterification reactions of acid oil with


ethanol was investigated in a continuous tubular reactor. The tem-
perature was the parameter that most influenced the reaction yield. The
acid content in the raw material increased the reaction rate varying the
raw material acidity within 5–35 wt% and reaching yields of 50–90% in
about 60 min of reaction time conducted at 280 °C, 100 bar and ethanol
to acid oil molar ratio of 4:1. This work brings two possibilities of
economics in the process of biodiesel production from the use of re-
sidual oils and low amounts of ethanol for the reaction. The kinetic data
were correlated satisfactorily with the model proposed. In addition, the
model was capable to predict satisfactorily different reactions condi-
tions than those used for the parameters adjustment. Applying this
technique (tubular reactor operating at supercritical conditions) for
producing alkyl ester from raw materials with different free fatty acid
content is a promising alternative to reduce costs and intensification of
Fig. 10. Effect of temperature on the (trans)esterification of acid oil with ethanol at
the process.
100 bar, Ac0 = 20 wt% and Et:AO = 4: 220 °C ( ), 250 °C ( ), 280 °C
( ), 300 °C ( ), 320 °C ( ) and 350 °C ( ) (Run 2, 6, 7, 8, mean
Acknowledgments
of 12 and 13 and 14 in Table 1).

The authors want to thank the Brazilian funding agencies (CNPq –


grants 406737/2013-4 and 305393/2016-2, CAPES and Fundação
Araucária) for financial support and scholarships.

Appendix A. Supplementary data

The supplementary material presents the experimental condition for


the (trans)esterification of acid oil (soybean oil with oleic acid) with
ethanol for each experimental condition presented in Table 1 (Table
S1–S14), the mean and standard deviation of the experimental condi-
tions of the runs 12 and 13 in Table 1 (Table S15) and the calculation of
the enthalpy change of reaction involved in kinetic modeling.
Supplementary data associated with this article can be found, in the
online version, at http://dx.doi.org/10.1016/j.fuel.2018.03.102.

References

[1] Alptekin E, Canakci M. Optimization of transesterification for methyl ester pro-


Fig. 11. Reproducibility experimental of (trans)esterification of acid oil with ethanol at duction from chicken fat. Fuel 2011;90:2630–8.
[2] Reyero I, Arzamendi G, Zabala S, Gandía LM. Kinetics of the NaOH-catalyzed
320 °C, 100 bar, A = 22 wt% and Et:AO = 4:1 (Run 12 and 13, Table 1).
transesterification of sunflower oil with ethanol to produce biodiesel. Fuel Process

497
K.C. dos Santos et al. Fuel 224 (2018) 489–498

Technol 2015;129:147–55. Supercrit Fluids 2010;53:82–7.


[3] Likozar B, Levec J. Effect of process conditions on equilibrium, reaction kinetics and [28] Velez A, Soto G, Hegel P, Mabe G, Pereda S. Continuous production of fatty acid
mass transfer for triglyceride transesterification to biodiesel: experimental and ethyl esters from sunflower oil using supercritical ethanol. Fuel 2012;97:703–9.
modeling based on fatty acid composition. Fuel Process Technol 2014;122:30–41. [29] Tan KT, Gui MM, Lee KT, Mohamed AR. An optimized study of methanol and
[4] Kumar R, Tiwari P, Garg S. Alkali transesterification of linseed oil for biodiesel ethanol in supercritical alcohol technology for biodiesel production. J Supercrit
production. Fuel 2013;104:553–60. Fluids 2010;53:82–7.
[5] Cho HJ, Kim SH, Hong SW, Yeo Y-K. A single step non-catalytic esterification of [30] Farobie O, Yanagida T, Matsumura Y. New approach of catalyst-free biodiesel
palm fatty acid distillate (PFAD) for biodiesel production. Fuel 2012;93:373–80. production from canola oil in supercritical tert-butyl methyl ether (MTBE). Fuel
[6] Konwar LJ, Wärnå J, Mäki-Arvela P, Kumar N, Mikkola J-P. Reaction kinetics with 2014;135:172–81.
catalyst deactivation in simultaneous esterification and transesterification of acid [31] Kusdiana D, Saka S. Kinetics of transesterification in rapeseed oil to biodiesel fuel as
oils to biodiesel (FAME) over a mesoporous sulphonated carbon catalyst. Fuel treated in supercritical methanol. Fuel 2001;80:693–8.
2016;166:1–11. [32] Diasakou M, Louloudi A, Papayannakos N. Kinetics of the non-catalytic transes-
[7] Gonzalez SL, Sychoski MM, Navarro-d HJ, Saibene M, Vieitez I, Silva C, et al. terification of soybean oil. Fuel 1998;77:1297–302.
Continuous catalyst-free production of biodiesel through transesterification of [33] Yujaroen D, Goto M, Sasaki M, Shotipruk A. Esterification of palm fatty acid dis-
soybean fried oil in supercritical methanol and ethanol. Energy Fuels tillate (PFAD) in supercritical methanol: effect of hydrolysis on reaction reactivity.
2013;27:5253–9. Fuel 2009;88:2011–6.
[8] Tan KT, Lee KT, Mohamed AR. Effects of free fatty acids, water content and co- [34] Santos PR, Voll FAP, Ramos LP, Corazza ML. Esterification of fatty acids with su-
solvent on biodiesel production by supercritical methanol reaction. J Supercrit percritical ethanol in a continuous tubular reactor. J Supercrit Fluids
Fluids 2010;53:88–91. 2017;126:25–36.
[9] Veljković VB, Lakićević SH, Stamenković OS, Todorović ZB, Lazić ML. Biodiesel [35] Warabi Y, Kusdiana D, Saka S. Reactivity of triglycerides and fatty acids of rapeseed
production from tobacco (Nicotiana tabacum L.) seed oil with a high content of free oil in supercritical alcohols. Bioresour Technol 2004;91:283–7.
fatty acids. Fuel 2006;85:2671–5. [36] Lokman IM, Goto M, Rashid U, Taufiq-Yap YH. Sub- and supercritical esterification
[10] Gupta J, Agarwal M, Dalai AK. Optimization of biodiesel production from mixture of palm fatty acid distillate with carbohydrate-derived solid acid catalyst. Chem Eng
of edible and nonedible vegetable oils. Biocatal Agric Biotechnol 2016;8:112–20. J 2016;284:872–8.
[11] Jain S, Sharma MP. Kinetics of acid base catalyzed transesterification of Jatropha [37] Verma P, Sharma MP. Review of process parameters for biodiesel production from
curcas oil. Bioresour Technol 2010;101:7701–6. different feedstocks. Renewable Sustainable Energy Rev 2016;62:1063–71.
[12] Canakci M, Van Gerpen J. Biodiesel production from oils and fats with high free [38] Sanli H, Canakci M. Effects of different alcohol and catalyst usage on biodiesel
fatty acids. Trans ASAE 2001;44:1429–36. production from different vegetable oils. Energy Fuels 2008;22:2713–9.
[13] Zhang Y, Wong W-T, Yung K-F. One-step production of biodiesel from rice bran oil [39] Knothe G. Dependence of biodiesel fuel properties on the structure of fatty acid
catalyzed by chlorosulfonic acid modified zirconia via simultaneous esterification alkyl esters. Fuel Process Technol 2005;86:1059–70.
and transesterification. Bioresour Technol 2013;147:59–64. [40] Joshi H, Moser BR, Toler J, Walker T. Preparation and fuel properties of mixtures of
[14] Boz N, Degirmenbasi N, Kalyon DM. Esterification and transesterification of waste soybean oil methyl and ethyl esters. Biomass Bioenergy 2009;34:14–20.
cooking oil over Amberlyst 15 and modified Amberlyst 15 catalysts. Appl Catal B [41] Dagostin JLA, Mafra MR, Ramos LP, Corazza ML. Liquid–liquid phase equilibrium
Environ 2015;165:723–30. measurements and modeling for systems involving {soybean oil + ethyl es-
[15] Saka S, Isayama Y, Ilham Z, Jiayu X. New process for catalyst-free biodiesel pro- ters + (ethanol + water)}. Fuel 2015;141:164–72.
duction using subcritical acetic acid and supercritical methanol. Fuel [42] Bondioli P, Bella L. Della. An alternative spectrophotometric method for the de-
2010;89:1442–6. termination of free glycerol in biodiesel. Eur J Lipid Sci Technol 2005;107:153–7.
[16] Glisic SB, Orlović AM. Review of biodiesel synthesis from waste oil under elevated [43] Ferrari JC, Nagatani G, Corazza FC, Oliveira JV, Corazza ML. Application of sto-
pressure and temperature: phase equilibrium, reaction kinetics, process design and chastic algorithms for parameter estimation in the liquid–liquid phase equilibrium
techno-economic study. Renewable Sustainable Energy Rev 2014;31:708–25. modeling. Fluid Phase Equilib 2009;280:110–9.
[17] Gui MM, Lee KT, Bhatia S. Supercritical ethanol technology for the production of [44] Silva C, Soh L, Barberio A, Zimmerman J, Seider WD. Phase equilibria of triolein to
biodiesel: process optimization studies. J Supercrit Fluids 2009;49:286–92. biodiesel reactor systems. Fluid Phase Equilib 2016;409:171–92.
[18] Saka S, Kusdiana D. Biodiesel fuel from rapeseed oil as prepared in supercritical [45] Dillon HE, Penoncello SG. A fundamental equation for calculation of the thermo-
methanol. Fuel 2001;80:225–31. dynamic properties of ethanol. Int J Thermophys 2004;25:321–35.
[19] Patil P, Deng S, Rhodes JI, Lammers PJ. Conversion of waste cooking oil to biodiesel [46] Wang L, Yang J. Transesterification of soybean oil with nano-MgO or not in su-
using ferric sulfate and supercritical methanol processes. Fuel 2009;89:360–4. percritical and subcritical methanol. Fuel 2007;86:328–33.
[20] Silva WC, Castro MPP, Haber Perez V, Machado FA, Mota L, Sthel MS. Thermal [47] Alenezi R, Leeke GA, Winterbottom JM, Santos RCD, Khan AR. Esterification ki-
degradation of ethanolic biodiesel: physicochemical and thermal properties eva- netics of free fatty acids with supercritical methanol for biodiesel production.
luation. Energy 2016;114:1093–9. Energy Convers Manage 2010;51:1055–9.
[21] Liu J, Shen Y, Nan Y, Tavlarides LL. Thermal decomposition of ethanol-based [48] Farobie O, Matsumura Y. A comparative study of biodiesel production using me-
biodiesel: mechanism, kinetics, and effect on viscosity and cold flow property. Fuel thanol, ethanol, and tert-butyl methyl ether (MTBE) under supercritical conditions.
2016;178:23–36. Bioresour Technol 2015;191:306–11.
[22] Tsai Y-T, Lin H-M, Lee M-J. Biodiesel production with continuous supercritical [49] Wang L, He H, Xie Z, Yang J, Zhu S. Transesterification of the crude oil of rapeseed
process: non-catalytic transesterification and esterification with or without carbon with NaOH in supercritical and subcritical methanol. Fuel Process Technol
dioxide. Bioresour Technol 2013;145:362–9. 2007;88:477–81.
[23] Lin R, Zhu Y, Tavlarides LL. Mechanism and kinetics of thermal decomposition of [50] Muppaneni T, Reddy HK, Patil PD, Dailey P, Aday C, Deng S. Ethanolysis of ca-
biodiesel fuel. Fuel 2013;106:593–604. melina oil under supercritical condition with hexane as a co-solvent. Appl Energy
[24] Manuale DL, Torres GC, Vera CR, Yori JC. Study of an energy-integrated biodiesel 2012;94:84–8.
production process using supercritical methanol and a low-cost feedstock. Fuel [51] Sakdasri W, Sawangkeaw R, Ngamprasertsith S. Continuous production of biofuel
Process Technol 2015;140:252–61. from refined and used palm olein oil with supercritical methanol at a low molar
[25] Vieitez I, Da Silva C, Alckmin I, De Castilhos F, Oliveira JV, Grompone MA, et al. ratio. Energy Convers Manage 2015;103:934–42.
Stability of ethyl esters from soybean oil exposed to high temperatures in super- [52] Silva C, Weschenfelder TA, Rovani S, Corazza FC, Corazza ML, Dariva C, et al.
critical ethanol. J Supercrit Fluids 2010;56:265–70. Continuous production of fatty acid ethyl esters from soybean oil in compressed
[26] Palacios-Nereo FJ, Olivares-Carrillo P, De Los Pérez, Ríos A, Quesada-Medina J. ethanol. Ind Eng Chem Res 2007:5304–9.
High-yield non-catalytic supercritical transesterification of soybean oil to biodiesel [53] Ortiz-Martínez VM, Salar-García MJ, Palacios-Nereo FJ, Olivares-Carrillo P,
induced by gradual heating in a batch reactor. J Supercrit Fluids 2016;111:135–42. Quesada-Medina J, De Los Ríos AP, et al. In-depth study of the transesterification
[27] Tat Tan K, Mei Gui M, Teong Lee K, Rahman Mohamed A. An optimized study of reaction of Pongamia pinnata oil for biodiesel production using catalyst-free su-
methanol and ethanol in supercritical alcohol technology for biodiesel production. J percritical methanol process. J Supercrit Fluids 2016;113:23–30.

498

You might also like