Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Nb2 O5 -catalyzed kinetics of fatty acids


esterification for reactive distillation
process simulation

Mauro Banchero ∗ , Giuseppe Gozzelino


Dipartimento di Scienza Applicata e Tecnologia, Politecnico di Torino, Torino, Corso Duca degli Abruzzi, 24,
10129 Torino, Italy

a r t i c l e i n f o a b s t r a c t

Article history: On the basis of kinetic data obtained for the first time in the presence of pelletized niobium
Received 6 March 2015 oxide as a catalyst, the esterification with methanol of single fatty acids or their mixtures has
Received in revised form 27 May been simulated in a reactive distillation unit, which is a promising intensification technique
2015 in the frame of biodiesel synthesis. The kinetic behaviour of the reversible esterification of
Accepted 31 May 2015 capric, lauric, myristic, palmitic, stearic and oleic acid at different temperatures was fitted
Available online 6 June 2015 through a pseudo-homogeneous reversible model suitable for the process design. The sim-
ulations showed that the reactivity of the acids and the acid composition of the feed affect
Keywords: both the column temperature profile and conversion, and that the benefit of a methanol
Fatty acid esterification excess is limited by the temperature decrease in the reactive section. The simulations also
Reactive distillation provided some preliminary design guidelines about the role of the main parameters, such
Pseudo-homogeneous reversible as top reflux ratio, column pressure, feed temperature and number of stages, on the process
kinetics performance.
Process intensification © 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Process simulation

1. Introduction and its esterification reaction with a fatty acid is slow and
limited by chemical equilibrium.
The need of alternatives to fossil fuels and of sustainable The limitations of the esterification process can be over-
technologies for energy production have raised the research come by different process intensification strategies (Gole and
interests towards biodiesel, which is a fuel derived from the Gogate, 2012; Maddikeri et al., 2012; Mazubert et al., 2013).
triglycerides contained in renewable sources such as edible, Integrated reactive separation technologies such as reactive
non-edible or waste oils, fatty acid cuts and algae (Aransiola distillation, reactive absorption, reactive extraction and mem-
et al., 2014; Gole and Gogate, 2012; Maddikeri et al., 2012; Ullah brane reactors are typical examples (Kiss and Bildea, 2012).
et al., 2015). The integration of the separation and reaction in a single oper-
The biodiesel is a mixture of organic esters that can be ation unit overcomes equilibrium limitations and provides
obtained from the triglycerides through two different routes: benefits such as energy savings by heat-integration systems.
the transesterification and the hydroesterification process. As Reactive distillation (RD) is one of the most used technolo-
far as the hydroesterification process is concerned, the trigly- gies, as reported by numerous studies (Kiss and Bildea, 2012).
cerides are hydrolyzed to free fatty acids and then esterified However, ideal RD systems assume that the reactants are the
with an alcohol (Maddikeri et al., 2012; Mazubert et al., 2013). compounds with intermediate volatilities and the products
Methanol is the most used alcohol in the biodiesel production are the heaviest and the lightest components. This does not


Corresponding author. Tel.: +39 0110904703.
E-mail address: mauro.banchero@polito.it (M. Banchero).
http://dx.doi.org/10.1016/j.cherd.2015.05.043
0263-8762/© 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301 293

always happen in practical applications and the investigation 2. Materials and methods
of the RD esterification of long-chain carboxylic acids with
methanol is particularly interesting because it is one of the 2.1. Materials
scarcely-investigated cases with the most unfavourable rel-
ative volatilities, where the reactants are the heaviest and All reactants were supplied by Sigma-Aldrich (USA) in the
lightest components (Chen et al., 2013; Tung and Yu, 2007). In 95–99.8% purity range and were used as received. The niobium
addition, the long-chain acids and their methyl esters exhibit (V) oxide that was used as the catalyst was kindly provided by
a lower difference in vapour pressures with respect to short- CBMM (Brazil). The catalyst (bulk density = 1370 kg m−3 , bed
chain systems so enhancing the complexity of their separation porosity = 0.70) was provided in the form of cylindrical pel-
in the RD column. lets (diameter = 5 mm and height = 4 mm) and was activated at
Previous process design studies report the results about 200 ◦ C for 1 h before its use (de Araújo Gonçalves et al., 2011).
the esterification of a single fatty acid (Banchero et al., The acid site density of the activated catalyst was evaluated
2014; Dimian et al., 2009; Kiss, 2011; Kiss et al., 2008, 2006; according to the method developed by Lopez et al. (2007) and
Machado et al., 2011; Steinigeweg and Gmehling, 2003) and resulted equal to 420 ␮mol g−1 .
the employed kinetic models are simplified since the esterifi-
cation is considered as a one-way irreversible reaction (Dimian 2.2. Experimental apparatus and procedures
et al., 2009; Kiss, 2011; Kiss et al., 2008, 2006; Machado et al.,
2011). The few reports about the RD simulation of the esterifi- The apparatus employed for the experiments was composed
cation of fatty acid mixtures or hydrolyzed oils (Cossio-Vargas by a stainless steel batch reactor (volume 300 ml) equipped
et al., 2012, 2011; Machado et al., 2013) hypothesise homoge- with a stationary annular basket-container for the catalyst
neous catalyzed systems where the equilibrium is attained on and a magnetically driven stirrer that provides fluid circula-
each reactive tray (Cossio-Vargas et al., 2012, 2011) or make use tion (Fig. 1). The vessel was equipped with an electrical heating
of models where the hydrolyzed oil is considered as a single mantle, a sampling line, pressure and temperature gauges and
component (Machado et al., 2013). a methanol-feeding device constituted by a 100-ml stainless
Reliable simulations or process design studies related to steel container connected to the reactor and to a nitrogen
the esterification of hydrolyzed oils either in conventional cylinder.
or unconventional layouts would require the knowledge of
appropriate reversible kinetic models for all the different fatty
acids that compose the mixture under the same working con-
ditions and over the same catalyst. Unfortunately the kinetic
studies available in the literature generally involve pure or oil-
diluted single fatty acids over different catalysts (Ilgen, 2014;
Omota et al., 2003; Patel and Brahmkhatri, 2013; Rattanaphra
et al., 2011; Steinigeweg and Gmehling, 2003; Tesser et al.,
2005). Recently, a kinetic study of the esterification of fatty
acids with methanol over powdered niobium oxide has been
performed (de Araújo Gonçalves et al., 2011) but the results
cannot be used in RD simulation because only the param-
eters referring to a one-way irreversible reaction have been
reported.
In the present work, experiments in the presence of pel-
lets of niobium oxide are reported and a pseudo-homogeneous
reversible kinetic model of the esterification of six fatty acids
(capric, lauric, myristic, palmitic, stearic and oleic acid) with
methanol is obtained. The acids were selected as the most
representative fatty acids in the composition of triglycerides
of different biodiesel feedstock (Lin et al., 2011). With respect
to previous literature works the kinetic model accounts for
both the direct and reverse reactions and, for the first time,
was obtained at the same working conditions (90–170 ◦ C) and
over the same catalyst in its pelletized form for all the tested
acids.
In the second part of this work the obtained kinetic
model was used to simulate a conventional RD unit as a
case-study. Simulations were performed either with single
fatty acids or mixtures that represent the hydrolysis prod-
uct of palm-oil-like or coconut-oil-like feedstock. Aim of
the simulations is to point out the different reactivity of
the single acids and how the composition of acid mixtures Fig. 1 – Experimental batch reactor. (1) Reactor; (2) electrical
from hydrolyzed oils affects the performance of the column. heating mantle; (3) annular stationary basket container; (4)
The effect of the most significant design parameters such magnetically driven stirring device; (5) temperature
as top reflux ratio (RR), column pressure, feed temperature indicator; (6) pressure indicator; (7) stainless steel container
and number of stages over the column performance is also for methanol addition; (8) nitrogen supply; (9) sampling
reported. line.
294 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301

In a standard experimental procedure the reactor was filled where XA is the conversion of the fatty acid, CA0 is the initial
in with the fatty acid and the catalyst and pre-heated before concentration of both reactants and Keq = k1 /k2 is the equilib-
the addition of methanol, which was introduced using a nitro- rium constant.
gen overpressure. The pre-heating temperature was set in The solution of Eq. (2) leads to the following equation:
order to reach the scheduled reaction temperature after the
 2a X − 2 − a   −2 + a 
addition of methanol. The stirring rate of the mixer was 1 A 2 2
In Y = In = a3 k1 CA0 t (3)
kept constant at 700 rpm for all the tests. This stirring rate 2a1 XA − 2 + a2 −2 − a2
was selected after performing preliminary experiments to
investigate the influence of mass transfer limitations on the where a1 = (1 − 1/Keq ), a2 = 2(1 − a1 )1/2 and a3 = (a2· mcat )/V. The
measured kinetics. For all the investigated systems exper- plot of ln Y versus t gives a straight line through the origin
iments at different stirring rates showed that above the whose slope allows k1 and k2 to be determined, provided
threshold value of 500 rpm the reaction did not evidence any that Keq is available. The equilibrium constant Keq can be
increase in velocity. The rotation speed of 700 rpm guarantees estimated from the experimental data. When the reaction
that external diffusive phenomena can be neglected and the reaches the thermodynamic equilibrium the plot of XA versus
reactive system can be considered in kinetic regime. Exper- time achieves an asymptotic value, XAeq , that allows the
iments were conducted in the 90–170 ◦ C thermal range with equilibrium constant Keq to be calculated from the following
stoichiometric amounts of reactants and a catalyst content equation:
equal to 5% of the acid amount. Samples of the reaction mix-
XA
2
ture were withdrawn at different times and the conversion Keq = (4)
2
was evaluated through the variation of the acid fraction of the (1 − XAeq )
withdrawn samples after removing the volatile components
(water and methanol) by heating each sample at 120 ◦ C for 1 h that is obtained assuming dXA /dt = 0 in Eq. (2).
in a ventilated oven. The acid fraction was measured accord-
ing to standard acid–base titration procedures recommended 2.4. Simulation set-up
in the literature (de Araújo Gonçalves et al., 2011; Rattanaphra
et al., 2011; Tesser et al., 2005). All the RD simulations were carried out through the RADFRAC
steady-state model implemented in ASPEN Plus chemical
process simulator (V 7.3). RADFRAC is a rigorous physical equi-
2.3. Kinetic model librium stage model able to solve the set of equations that
describe each equilibrium stage: material balances, enthalpy
The experimental tests in the batch reactor were modelled balances and phase equilibrium relations between the vapour
with the following pseudo-homogeneous second-order equi- and the liquid leaving the stage. The phase equilibrium was
librium kinetics (Rattanaphra et al., 2011): described assuming the vapour phase as ideal and the liquid
phases as real, on the basis of the UNIFAC-Dortmund model
as it was suggested by many literature works (Kiss, 2011; Kiss
dCA
V = rA V = −k1 mcat CA CS + k2 mcat CE CW (1) and Bildea, 2012; Machado et al., 2013, 2011; Steinigeweg and
dt
Gmehling, 2003). The pseudo-homogeneous kinetic model, rA ,
reported in Eq. (1), which accounts for the chemical reaction
where rA is the reaction rate, V is the volume of the pseudo- rate in the liquid phase, was included in the set of equations
homogeneous reacting system, mcat is the catalyst weight, that described the reactive section of the RD column. The
CA , CB , CE , CW are the molar concentrations of fatty acid, kinetic constants are reported in Table 1. The main param-
methanol, fatty acid methyl ester and water, respectively, k1 eters of the simulated RD column are reported in Table 2; they
and k2 are the forward and backward rate constants referred were defined with reference to literature works (Dimian et al.,
to the catalyst loading. The effect of temperature on the reac- 2009; Kiss, 2011), which investigated the esterification of dode-
tion rates was considered through the conventional Arrhenius canoic acid with methanol, and to the result of the parameter
equation with k1 = k1∞ ·exp(−E1 /RT) and k2 = k2∞ ·exp(−E2 /RT) design discussed in Section 3.4.
where k1∞ and k2∞ are the pre-exponential factors and E1 The reported results refer to the simulation of a RD unit fed
and E2 the activation energies. Since CA = CB = CA0 (1 − XA ) and with pure fatty acids or mixtures of fatty acids similar to those
CE = CW = CA0 XA , the following equation becomes: that can obtained from the hydrolysis of hypothetical palm-
oil-like (mixture A) or coconut-oil-like (mixture B) feedstock.
  The compositions of the mixtures used in the simulations are
dXA 1 2
V = k1 mcat CA0 (1 − XA )2 − X (2) reported in Table 3. They were obtained averaging the exper-
dt Keq A
imental data of real palm oil and coconut oil compositions

Table 1 – Kinetic and thermodynamic parameters for the esterification of fatty acids with methanol.
Fatty acid k1∞ (m6 kmol−1 s−1 kg−1 ) E1 (kJ kmol−1 ) k2∞ (m6 kmol−1 s−1 kg−1 ) E2 (kJ kmol−1 ) Keq 1 (−)

Capric (C10:0) 48.0 60.5 × 103 31.7 60.1 × 103 1.33


Lauric (C12:0) 16.8 55.7 × 103 19.7 57.0 × 103 1.26
Myristic (C14:0) 5.14 53.2 × 103 11.0 54.2 × 103 0.63
Palmitic (C16:0) 3.84 52.3 × 103 6.59 52.0 × 103 0.53
Stearic (C18:0) 3.61 52.2 × 103 4.72 50.7 × 103 0.48
Oleic (C18:1) 499 66.6 × 103 187 66.0 × 103 2.33

1
Calculated at T = 125 ◦ C.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301 295

Table 2 – Parameters for the RD simulation with


stoichiometric reactant ratio.
Parameter Value

Total number of stages 25


Number of reactive stages 20 (from 3 to 22)
Catalyst bulk density 1370 kg m−3
Catalyst porosity 0.70
Liquid holdup per stage 0.1 m3
Fatty acid feed on stage 3 at 160 ◦ C 5.83 kmol h−1
Methanol feed on stage 22 at 104.5 ◦ C 5.89 kmol h−1
Pressure 4 atm
Top reflux ratio (mass ratio) 0.10 kg kg−1
Bottoms rate 5.89 kmol h−1
Type of condenser (stage 1) Total
Type of reboiler (stage 25) Kettle

reported in the literature (Aranda et al., 2009; Jang and Tan,


2012; Lin et al., 2011; Nakpong and Wootthikanokkhan, 2010).
Simulations with pure acids as well as those with mixtures A
and B were conducted assuming the acidic feed on stage 3 as
reported in Table 2.
Fig. 2 – Conversion profiles versus time for the
3. Results and discussion esterification of different acids with methanol at 110 ◦ C.

3.1. Experimental evaluation of the reaction rate trend can be related to the mechanism of acid-catalyzed ester-
constants ification that involves the nucleophilic attack of the alcohol
to the catalyst-protonated carboxylic moiety of the fatty acid.
The regression of the experimental conversion profiles versus The reactivity should be influenced both by the steric effect of
time with the kinetic model described in Section 2.3 provided the chain length of the alkyl group and by the polarity of the
the kinetic constants for all the investigated systems. A selec- fatty acid molecule. It has been reported that a decrease in
tion of the experimental results is reported in Figs. 2 and 3: chain length or the presence of unsaturated bonds lead to an
the different kinetic behaviour of the fatty acids in presence increase in the polarity of the fatty acid molecule resulting in
of Nb2 O5 at 110 ◦ C is compared in Fig. 2 whereas the results higher reactivity (de Araújo Gonçalves et al., 2011). The slight
obtained for the esterification of oleic acid are reported in difference between the activation energies of the forward and
Fig. 3 as a plot example. All the evaluated kinetic parameters reverse kinetic constants (Table 1) suggests an almost ather-
are summarized in Table 1, which also reports the equilib- mal character of the esterification reaction and supports the
rium constants. As shown in Fig. 3a, the conversion increases observed negligible role of temperature on the equilibrium
versus time until a steady state condition is achieved, which is conversion.
the conversion at equilibrium. The equilibrium constant can
then be determined through Eq. (4). Fig. 3a also shows that 3.2. RD simulation for the esterification of a single
the equilibrium conversion is almost independent of temper- fatty acid
ature. The same result was obtained for all the tested systems
and evidences that temperature cannot be used as a process The pseudo-homogeneous kinetic model described in previ-
parameter to enhance the equilibrium conversion of the ester- ous sections was used to simulate a conventional RD unit as a
ification reaction. case-study. The performance of the column operating on each
The data reported in Table 1 show that reactivity depends single fatty acid fed at a fixed molar flow rate at stoichiomet-
on the chain length of the fatty acid (from C10 to C18) except ric conditions and with the parameters reported in Table 2
for oleic acid that exhibits an anomalous high reactivity. This was investigated. The conversion results as well the complete
mass and energy balances of the process are summarized in
Table 4. The table also reports the maximum equilibrium con-
Table 3 – Composition of the mixtures of fatty acids (% version that can be obtained in a conventional continuous
w/w).
reactor in thermodynamic equilibrium conditions.
Fatty acid Mixture A (palm Mixture B The conversion results show the same trend as the kinetic
oil)1 (%) (coconut oil)2 (%)
parameters. All the investigated systems show a conversion in
Capric (C10:0) – 6.8 the RD column that is much higher than the equilibrium con-
Lauric (C12:0) 0.6 48.4 version. Conversions up to 94–96% can be obtained for capric,
Myristic (C14:0) 0.6 19.1 lauric and oleic acid, which exhibit the highest reactivity.
Palmitic (C16:0) 44.8 10.9 The temperature and the reaction rate profiles along the RD
Stearic (C18:0) 7.6 4.4
column are reported in Fig. 4 while Fig. 5 shows the vapour and
Oleic (C18:1) 46.4 10.4
liquid compositions. Both figures report the results obtained
1
Mixture of fatty acids obtained from the hydrolysis of hypothetical for the esterification of palmitic acid but similar results were
palm-oil-like feedstock. obtained for the other investigated systems.
The column is operated in the 124–394 ◦ C temperature
2
Mixture of fatty acids obtained from the hydrolysis of hypothetical
coconut-oil-like feedstock. range but the reaction is conducted in nearly isothermal
296 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301

70 5
4.5
60
4 90°C
50 3.5 110°C
conversion ( %)

3 130°C
40 90°C
2.5 (b) 150°C
110°C
30
2 170 °C
130°C
20 (a) 150°C 1.5

170 ° 1
10
0.5
0 0
0 50 0 1000 1500 0 50 100 150 200
t (min)
t (min)

Fig. 3 – Conversion profiles (a) and plot of ln Y (b) versus time at different temperatures for the esterification of oleic acid
with methanol.
conditions with a temperature value in the reactive section ester reaches its maximum of concentration. The composition
close to 160 ◦ C (Fig. 4a), which is the temperature of the fatty profiles in the liquid phase give reason of the reaction rate pro-
acid feed. The same temperature trend in the reactive sec- file reported in Fig. 4b: a maximum is displayed at the top stage
tion was found for all the systems except for the esterification of the reactive section where the forward reaction definitely
of oleic acid where a temperature value of approximately prevails on the backward one due to the high concentration
130 ◦ C was achieved. This can be related to the different phys- of the fatty acid. In lower stages the forward reaction rate
ical properties of this unsaturated fatty acid. Even though its is reduced and is partially balanced by the backward reac-
molecular weight is close to that of the “long-chain” stearic tion until a flat pattern is displayed. Even though the ester
acid, which is solid at room temperature, its physical proper- concentration at the bottom stages of the reactive section is
ties are similar to lighter fatty acids that are liquid at room maximum the backward reaction is inhibited by the nearly
temperature. complete depletion of water in this zone. This confirms that
In the liquid phase (Fig. 5a) of the reactive section the con- integrating separation and reaction in the reactive section
centration of methanol increases from the top to the bottom results in shifting the chemical equilibrium to the right.
while that of the fatty acid shows the opposite trend. The
top stages of the reactive section operate with a large excess 3.3. RD simulation for the esterification of fatty acid
of fatty acid and the depletion of acid along the column is mixtures
compensated by the increase in the methanol content. As far
as the products are concerned it can be observed that while The RD column described in Table 2 was simulated assuming
the ester concentration increases from the top to the bottom fatty acid mixtures as the feed on stage 3. The composition
of the column, water is nearly completely absent where the of the fatty acid mixtures is reported in Table 3. Conversion

Table 4 – Conversion, mass balance, product purity, top/bottom temperature and heat requirement for the esterification
of the fatty acids in the RD column.
Capric (C10:0) Lauric (C12:0) Myristic (C14:0) Palmitic (C16:0) Stearic (C18:0) Oleic (C18:1)

Equilibrium conversion1 54% 53% 44% 42% 41% 61%


Conversion in the RD 96% 95% 83% 80% 77% 94%
column
Fatty acid feed (kg h−1 ) 1004 1168 1332 1495 1659 1647
Methanol feed (kg h−1 ) 189 189 189 189 189 189
Top stream mass flow 111 111 120 122 125 110
rate (kg h−1 )
Bottom stream mass 1082 1246 1401 1562 1723 1726
flow rate (kg h−1 )
Product purity in 96.2% 95.2% 82.4% 80.7% 77.4% 94.5%
bottom stream (ester
content) (% w/w)
Top condenser 138 136 126 124 123 136
temperature (◦ C)
Condenser heat duty 69 69 68 68 68 69
(kW)
Bottom reboiler 290 330 364 394 422 411
temperature (◦ C)
Reboiler heat duty (kW) 166 227 314 359 434 488

1
Maximum equilibrium conversion attainable in a isothermal continuous reactor working at the average temperature of the RD reactive section.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301 297

Temperature (°C) Reacon rate (kmol/h)


100 200 300 400 0 0.2 0.4 0.6
0 0
(a) (b)

5 5

10 10
Stage Stage
reacve secon
reacve secon

15 15

20 20

25 25

Fig. 4 – Temperature (a) and reaction rate (b) profiles along the RD unit for the esterification of palmitic acid with methanol
(P = 4 atm, RR = 0.1, reactive stages from 3 to 22, temperature of the acid feed = 160 ◦ C, temperature of the methanol
feed = 104.5 ◦ C).

results, mass and energy balances are reported in Table 5. Sim- operate at temperatures 13–14 ◦ C lower than those achieved
ilar conversion results in the RD column were obtained for the when the esterification of mixture B was conducted. In fact,
two mixtures: 87% for mixture A and 91% for mixture B. In both the results reported in Section 3.2 pointed out that when the
cases the conversion is higher than the maximum conversion esterification of oleic acid was investigated a lower tempera-
that can be obtained in a conventional continuous reactor in ture of the reactive section was observed.
thermodynamic equilibrium. The slightly higher conversion It can be concluded that the composition profile of the
obtained for mixture B with respect to mixture A can be related hydrolyzed oil can influence the performance of the RD
to the different composition of the two mixtures. While mix- column in two ways. From one hand higher amount of short-
ture A (hydrolyzed palm oil) is richer in long-chain fatty acids, chain or unsaturated fatty acid components involves higher
mixture B (hydrolyzed coconut oil) is richer in lighter compo- reaction conversion rates due to the more reactive charac-
nents. However, mixture A is quite rich in oleic acid, which is ter of these compounds. On the other hand the presence
one of the most reactive fatty acid due to the presence of the of unsaturated fatty acids, such as oleic acid, with physical
unsaturated bond. properties quite different from those of the homologous sat-
The temperature of the reactive section of the RD column urated compounds may mitigate the conversion performance
is another important aspect to be considered. When mixture due to a reduction in the temperature level of the reactive
A was fed to the RD column the reactive section was found to section.

mole fracon mole fracon


0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0

5 5

reacve secon

10 Methanol 10
Stage Stage
Water
Acid
15 15 Methanol
Ester
Water
reacve secon Acid
20 20 Ester

(a) (b)
25 25

Fig. 5 – Liquid (a) and vapour (b) molar composition profiles along the RD unit for the esterification of palmitic acid with
methanol (P = 4 atm, RR = 0.1, reactive stages from 3 to 22, temperature of the acid feed = 160 ◦ C, temperature of the methanol
feed = 104.5 ◦ C).
298 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301

increased. For example, when the RD column was operated


Table 5 – Conversion, mass balance, product purity,
top/bottom temperature and heat requirement for the under an acid to methanol molar ratio equal to 1:3 conversions
esterification of the mixtures of fatty acids in the RD for mixtures A and B increased up to 96 and 98%, respectively,
column. and the ester content in the bottom streams could overcome
Mixture A Mixture B the above-cited limit for vehicle use of biodiesel. It must be
(palm oil)1 (coconut oil)2 pointed out, however, that increasing the methanol flow rate,
which is the most volatile compound, also results in a tem-
Equilibrium conversion3 50% 51%
perature decrease of the reactive section. In the above-cited
Conversion in the RD 87% 91%
column
example, where the RD column is operated under an acid
Fatty acid feed (kg h−1 ) 1570 1325 to methanol molar ratio equal to 1:3, the temperature of the
Methanol feed (kg h−1 ) 189 189 reactive section is approximately 10 ◦ C lower than when the
Top stream mass flow rate 116 114 column is operated under stoichiometric conditions, which
(kg h−1 ) partially limits the benefit of operating under an excess of one
Bottom stream mass flow 1643 1399
reactant.
rate (kg h−1 )
Product purity in bottom 87.8% 91.4%
stream (ester content) 3.4. The role of the main design parameters on the RD
(% w/w) performance
Top condenser 129 131
temperature (◦ C)
Condenser heat duty (kW) 69 69
The above-reported simulation results were obtained after
Bottom reboiler 402 353 investigating the effect of the critical design parameters over
temperature (◦ C) the RD column performance: top reflux ratio, column pressure,
Reboiler heat duty (kW) 422 290 temperature of the feed streams and number of stages. These
parameters, which influenced the esterification behaviour of
1
Mixture of fatty acids obtained from the hydrolysis of hypothetical
all the investigated systems in a similar way, are briefly ana-
palm-oil-like feedstock.
2
Mixture of fatty acids obtained from the hydrolysis of hypothetical
lyzed hereinafter.
coconut-oil-like feedstock.
3
Maximum equilibrium conversion attainable in a isothermal con-
3.4.1. The top reflux ratio
tinuous reactor working at the average temperature of the RD
Fig. 6 shows the role of the top reflux ratio (RR) on the perfor-
reactive section.
mance of the RD unit. Higher conversions and lower energy
demand can be achieved operating at low RR values (Fig. 6a).
Another important aspect to be considered in the biodiesel In the RD configuration, the reflux acts like an internal recy-
synthesis from hydrolyzed oils is related to the product purity. cle of a methanol–water mixture whose composition depends
Table 5 reports that the ester content in the bottom stream is on the adopted RR. Higher RR values result in higher amounts
lower than 96.5%, which is recommended for vehicle use of of liquid flowing through the column (Fig. 6b), which involves
biodiesel (Knothe, 2006). This would require a further increase higher heat duties both at the reboiler and condenser. Other
in the number of reactive stages or the addition of a purifi- consequences of the increased amount of liquid across the
cation unit of the bottom product. Alternatively, reactant column are connected with the decrease in the average tem-
conversion and ester content in the bottom stream could be perature in the reactive zone and the increase in water amount
increased operating under an excess of methanol to achieve a in the liquid (Fig. 6b). Both phenomena contribute to the
further shift-to-the-right of the chemical equilibrium. For this decrease in conversion shown in Fig. 6a. The temperature
reason, some simulation tests were performed at higher acid reduction decreases the overall reaction rate while the higher
to methanol molar ratios. As it could be expected, conversion amount of water results in shifting the chemical equilibrium

80 500 10
160
9
70
140 8
400
60
conversion 120 7
Temperature (°C)
Conversion (%)

Flow rate (kmol/h)


Heat duty (kW)

50 reboiler heat duty 300 6


100
condenser heat duty temperature
40 5
80 total liquid flow rate
200 4
30 60 liquid water flow rate
(a) (b)
3
20 40
100 2
10 20 1
0 0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Reflux rao Reflux rao

Fig. 6 – The effect of the RR on the performance of the RD unit for the esterification of stearic acid with methanol (P = 4 atm,
reactive stages from 3 to 22, temperature of the acid feed = 160 ◦ C, temperature of the methanol feed = 104.5 ◦ C). (a)
Conversion and heat duties versus RR. (b) Average values of the temperature and liquid flow rates in the reactive section
plotted versus RR.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301 299

100 500 180 1


90 160 0.9
80 400 140 0.8
conversion
70 0.7

Temperature (°C)
reboiler heat duty
Conversion (%)

120

Heat duty (kW)

Mole fracon (-)


60 condenser heat duty 300 temperature 0.6
100 methanol content in the liquid phase
50 0.5
80
40 200 0.4
60
30 (a) (b) 0.3
20 100 40 0.2
10 20 0.1
0 0 0 0
1 3 5 7 9 1 3 5 7 9
Pressure (atm) Pressure (atm)

Fig. 7 – The effect of the pressure on the performance of the RD unit for the esterification of lauric acid with methanol
(RR = 0.1, reactive stages from 3 to 22, temperature of the acid feed = 160 ◦ C, temperature of the methanol feed = 104.5 ◦ C). (a)
Conversion and heat duties versus pressure. (b) Average values of the temperature and liquid methanol mole fraction in the
reactive section plotted versus pressure.

towards the reactants (Steinigeweg and Gmehling, 2003). An temperature profile of the reactive section of the column
optimized value of 0.1 for RR was selected, as reported in (Fig. 8b). It can be concluded that higher temperatures of the
Table 2, since lower values did not result in a further increase acid feed are beneficial to the reactant conversion and the tem-
in the reactant conversion. perature selection can be a useful tool to control the thermal
profile of the reactive section of the RD column. As far as the
3.4.2. The column pressure simulations reported in this work are concerned it was con-
The effect of pressure on the performance of the RD col- sidered appropriate not to exceed the temperature of 170 ◦ C
umn is shown in Fig. 7. The reactant conversion increases in the reactive section, which was the maximum temperature
because the temperature of the reactive section is higher investigated in the kinetic experiments, and a value of 160 ◦ C
when the column is operated at higher pressures (Fig. 7b). for the acid feed was selected, as reported in Table 2.
This is the consequence of the increased boiling point of The role of the temperature of the alcohol feed on the RD
the components that results in higher temperatures of the performance was also investigated but no significant results in
top condenser and bottom reboiler as well as higher energy terms of reactant conversion were obtained. The simulations
demand at the reboiler (Fig. 7a). Furthermore the pressure showed that the temperature of the alcohol feed only affects
increase results in higher methanol concentration in the liq- the temperature profile of the stripping section (i.e. the section
uid phase (Fig. 7b). This last point is particularly relevant below the reactive zone). The liquid methanol stream (Table 2)
for this RD system where reactant methanol is the most was, then, fed on stage 22 at its boiling point, which is equal
volatile component and the reaction takes place in the liquid to 104.5 ◦ C at 4 atm, as suggested by previous works (Dimian
phase. Operating at higher pressure limits the drawback of the et al., 2009; Kiss, 2011).
unfavourable boiling point ranking of reactants and products.
The higher methanol concentration in the liquid enhances the
forward reaction resulting in increased conversion. An opti- 3.4.4. The number of stages
mized value of 4 atm for the column pressure was selected, Fig. 9 reports the reactant conversion of the RD column versus
as reported in Table 2, since higher values would require too the number of stages in the reactive section. Data are reported
high energy demand without resulting in significant increase for the esterification of capric acid at two different pressures.
in the reactant conversion. Furthermore the selected pressure In both cases the RD performance is significantly higher than
could guarantee to operate at temperature values in the bot- the equilibrium conversion, which is unaffected by the change
tom reboiler (Tables 4 and 5) which are not so much higher in the working pressure since the reaction occurs in the
than the corresponding normal boiling points of the fatty acids liquid phase. Furthermore, the chemical equilibrium is also
and their methyl esters (270–375 ◦ C) so preventing the risk of unaffected by temperature as explained in Section 3.1. High
significant thermal decomposition of the product. conversion can be obtained operating at higher pressure of
the column with lower number of stages since the reaction
3.4.3. The temperature of the acid and alcohol feed can be conducted at higher temperature levels. A value of
streams 20 reactive stages was selected, as reported in Table 2, since
The role of the temperature of the acid feed on the reac- at 4 atm a higher number of stages did not result in signifi-
tant conversion and energy demand of the RD column is cant improvement in the conversion. It must be pointed out,
displayed in Fig. 8. The higher energy demand connected with however, that the most appropriate selection of the working
the temperature increase of the acid feed, which involves its pressure and the consequent number of reactive stages should
pre-heating, is compensated by a reduction of the heat duty at be done after an economical optimization of the process. In
the reboiler as shown in Fig. 8a. As a consequence, a change fact, if the column is operated at lower pressures a higher
in this parameter doesn’t significantly influence the overall number of reactive stages is needed whereas if it is run at
energy demand. On the other hand the reactant conversion higher pressures higher energy demand and higher reboiler
is increased because the temperature of the feed affects the temperatures are required as discussed in Section 3.4.2.
300 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301

Temperature (°C)
0 100 200 300 400
0

100 500
90
5
80 400 reacve
secon
70
Conversion (%)

Heat duty (kW)


60 300 10
TACID =120 °C
Stage
50 TACID =160 °C
40 conversion 200 15 TACID =220 °C
reboiler heat duty
30
condenser heat duty
(a) (b)
20 100
20
10
0 0
120 140 160 180 200 220 240
TACID (°C) 25

Fig. 8 – The effect of the temperature of the acid feed on the performance of the RD unit for the esterification of myristic acid
with methanol (P = 4 atm, RR = 0.1, reactive stages from 3 to 22, temperature of the methanol feed = 104.5 ◦ C). (a) Conversion
and heat duties versus the temperature of the acid feed. (b) Temperature profiles along the column at different temperatures
of the acid feed.

The role of the number of stages in the rectifying and strip- The scarce influence of the number of stages in the rec-
ping sections of the RD column (i.e. the sections above and tifying and stripping sections of the column is connected
below the reactive zone) was also investigated. It was found with this particular system where the reactants are the heavi-
that increasing the number of stages in these two sections did est and the lightest component. Increasing the number of
not affect the performance of the RD unit. This is evidenced by stages in these two zones would not result in increasing
the composition profiles reported in Fig. 5. The liquid phase in the purity of the products in the top and bottom streams
the rectifying section is almost composed by methanol and since they are the components with intermediate volatili-
water and no further stages are required to improve their ties. It can be concluded that the RD column of this specific
separation from the “heavy” components. A similar situation process mainly consists in the reactive section. This signifi-
occurs in the stripping section as far as the high-boiling com- cantly differs from the ideal RD systems where the rectifying
ponents are concerned. It was, then, decided to work with the and stripping sections are intended to increase the recovery
minimum number of stages in these two sections as reported efficiency of products, which are the lightest and heaviest
in Table 2. compound.

4. Conclusions
100
90 A pseudo-homogeneous reversible kinetic model for the ester-
ification of six different fatty acids with methanol has been
80
experimentally obtained in the presence of pellets of niobium
70 oxide for the first time. The obtained model, which can be used
Conversion (%)

to design conventional or unconventional esterification units


60
in the frame of the biodiesel synthesis, was used to design a
50 RD column through a commercial software as a case-study.
40 equilibrium conversion Results pointed out the different reactivity of the sin-
gle acids while simulations with different hypothetical
30 hydrolyzed-oil feedstock of industrial interest showed how
P=1 atm the oil composition affects the reaction rate and temperature
20
P=4 atm profile of the column. Simulation tests at different reactant
10 molar ratios pointed out that the benefit of methanol excess
0 is partially limited by the temperature decrease in the reactive
10 15 20 25 30 35 40 section.
Number of reacve stages The simulations also provided some preliminary design
guidelines for this specific RD process with unfavourable boil-
Fig. 9 – The effect of the number of reactive stages on the ing point ranking of reactants and products:
performance of the RD unit for the esterification of capric
acid with methanol at different pressures (RR = 0.1,
temperature of the methanol feed = 104.5 ◦ C, temperature of • higher conversions and lower energy consumption are
the acid feed = 160 ◦ C). achieved operating at low reflux ratios;
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 292–301 301

• operating at higher pressure limits the drawback of the Kiss, A.A., 2011. Heat-integrated reactive distillation process for
unfavorable boiling point ranking of reactants and products; synthesis of fatty esters. Fuel Process. Technol. 92, 1288–1296.
• the temperature profile in the reactive section, which Kiss, A.A., Bildea, C.S., 2012. A review of biodiesel production by
integrated reactive separation technologies. J. Chem. Technol.
operates at isothermal conditions, is controlled by the tem-
Biotechnol. 87, 861–879.
perature of the acid feed; Kiss, A.A., Omota, F., Dimian, A.C., Rothenberg, G., 2008. Biodiesel
• the temperature of the alcohol feed does not affect the col- by catalytic reactive distillation powered by metal oxides.
umn performance; Energy Fuels 22, 598–604.
• the number of stages in the stripping and rectifying sections Kiss, A.A., Omota, F., Dimian, A.C., Rothenberg, G., 2006. The
should be minimum since they do not influence neither the heterogeneous advantage: biodiesel by catalytic reactive
conversion performance nor product purity. distillation. Top. Catal. 40, 141–150.
Knothe, G., 2006. Analyzing biodiesel: standards and other
methods. J. Am. Oil Chem. Soc. 83, 823–833.
References Lin, L., Cunshan, Z., Vittayapadung, S., Xiangqian, S., Mingdong,
D., 2011. Opportunities and challenges for biodiesel fuel. Appl.
Aranda, D.A.G., de Araújo Gonçalves, J., Peres, J.S., Ramos, A.L.D., Energy 88, 1020–1031.
de Melo, C.A.R., Antunes, O.A.C., Furtado, N.C., Taft, C.A., 2009. Lopez, D.E., Goodwin Jr., G.G., Bruce, D.A., 2007.
The use of acids, niobium oxide and zeolite catalysts for Transesterification of triacetin with methanol on nafion acid
esterification reaction. J. Phys. Org. Chem. 22, resins. J. Catal. 245, 381–391.
709–711. Machado, G.D., Aranda, D.A.G., Castier, M., Cabral, V.F.,
Aransiola, E.F., Ojumu, T.V., Oyekola, O.O., Madzimbamuto, T.F., Cardozo-Filho, L., 2011. Computer simulation of fatty acid
Ikhu-Omoregbe, D.I.O., 2014. A review of current technology esterification in reactive distillation columns. Ind. Eng. Chem.
for biodiesel production: state of the art. Biomass Bioenergy Res. 50, 10176–10184.
61, 276–297. Machado, G.D., Pellegrini Pessoa, F.L., Castier, M., Aranda, D.A.G.,
Banchero, M., Kusumaningtytas, R.D., Gozzelino, G., 2014. Cabral, V.F., Cardozo-Filho, L., 2013. Biodiesel production by
Reactive distillation in the intensification of oleic acid esterification of hydrolyzed soybean oil with ethanol in
esterification with methanol—a simulation case-study. J. Ind. reactive distillation columns: simulation studies. Ind. Eng.
Eng. Chem. 20, 4242–4249. Chem. Res. 52, 9461–9469.
Chen, H., Huang, K., Liu, W., Zhang, L., Wang, S., Wang, S.J., 2013. Maddikeri, G.L., Pandit, A.B., Gogate, P.R., 2012. Intensification
Enhancing mass and energy integration by external recycle in approaches for biodiesel synthesis from waste cooking oil: a
reactive distillation columns. AICHE J. 59, 2015–2032. review. Ind. Eng. Chem. Res. 51, 14610–14628.
Cossio-Vargas, E., Barroso-Muňoz, F.O., Hernandez, S., Mazubert, A., Poux, M., Aubin, J., 2013. Intensified processes for
Segovia-Hernandez, J.G., Cano-Rodriguez, M.I., 2012. FAME production from waste cooking oil: a technological
Thermally coupled distillation sequences: steady state review. Chem. Eng. J. 233, 201–223.
simulation of the esterification of fatty organic acids. Chem. Nakpong, P., Wootthikanokkhan, S., 2010. High free fatty acid
Eng. Process. 62, 176–182. coconut oil as a potential feedstock for biodiesel production in
Cossio-Vargas, E., Hernandez, S., Segovia-Hernandez, J.G., Thailand. Renewable Energy 35, 1682–1687.
Cano-Rodriguez, M.I., 2011. Simulation study of the Omota, F., Dimian, A.C., Bliek, A., 2003. Fatty acid esterification by
production of biodiesel using feedstock mixtures of fatty acids reactive distillation: Part 2—Kinetics based design for
in complex reactive distillation columns. Energy 36, sulphated zirconia catalyst. Chem. Eng. Sci. 58, 3175–3185.
6289–6297. Patel, A., Brahmkhatri, V., 2013. Kinetic study of oleic acid
de Araújo Gonçalves, J., Ramos, A.L.D., Rocha, L.L.L., Domingos, esterification over 12-tungstophosphoric acid catalyst
A.K., Monteiro, R.S., Peres, J.S., Furtado, N.C., Taft, C.A., anchored to different mesoporous silica supports. Fuel
Aranda, D.A.G., 2011. Niobium oxide solid catalyst: Process. Technol. 113, 141–149.
esterification of fatty acids and modelling for biodiesel Rattanaphra, D., Harvey, A.P., Thanapimmetha, A., Srinophakun,
production. J. Phys. Org. Chem. 24, 54–64. P., 2011. Kinetic of myristic acid esterification with methanol
Dimian, A.C., Bildea, C.S., Omota, F., Kiss, A.A., 2009. Innovative in the presence of triglycerides over sulphated zirconia.
process for fatty acid esters by dual reactive distillation. Renewable Energy 36, 2679–2686.
Comput. Chem. Eng. 33, 743–750. Steinigeweg, S., Gmehling, J., 2003. Esterification of a fatty acid by
Gole, V.L., Gogate, P.R., 2012. A review on intensification of reactive distillation. Ind. Eng. Chem. Res. 42, 3612–3619.
synthesis of biodiesel from sustainable feed stock using Tesser, R., Di Serio, M., Guida, M., Nastasi, M., Santacesaria, E.,
sonochemical reactors. Chem. Eng. Process. 53, 1–9. 2005. Kinetics of oleic acid esterification with methanol in the
Ilgen, O., 2014. Investigation of reaction parameters, kinetics and presence of triglycerides. Ind. Eng. Chem. Res. 44, 7978–7982.
mechanism of oleic acid esterification with methanol by Tung, S.T., Yu, C.C., 2007. Effects of relative volatility ranking to
using Amberlyst 46 as a catalyst. Fuel Process. Technol. 124, the design of reactive distillation. AICHE J. 53, 1278–1297.
134–139. Ullah, K., Ahmad, M., Sofia Sharma, V.K., Lu, P.M., Harvey, A.,
Jang, J.J., Tan, C.S., 2012. Biodiesel production from coconut oil in Zafar, M., Sultana, S., 2015. Assessing the potential of algal
supercritical methanol in the presence of cosolvent. J. Taiwan biomass opportunities for bioenergy industry: a review. Fuel
Inst. Chem. Eng. 43, 102–107. 143, 414–423.

You might also like