Вы находитесь на странице: 1из 14

G Model

CACE-5353; No. of Pages 14 ARTICLE IN PRESS


Computers and Chemical Engineering xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Computers and Chemical Engineering


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

An integrated reactive distillation process for biodiesel production


Eduardo S. Pérez-Cisneros a,∗ , Xenia Mena-Espino a , Verónica Rodríguez-López c ,
Mauricio Sales-Cruz b , Tomás Viveros-García a , Ricardo Lobo-Oehmichen a
a
Universidad Autónoma Metropolitana – Iztapalapa, Departamento de Ingeniería de Procesos e Hidráulica, San Rafael Atlixco No. 186, Colonia Vicentina,
Delegación Iztapalapa, C.P. 09340 México D.F., Mexico
b
Universidad Autónoma Metropolitana – Cuajimalpa, Departamento de Procesos y Tecnología, Avenida Vasco de Quiroga No.4871, Colonia Santa Fe
Cuajimalpa, Delegación Cuajimalpa de Morelos, C.P. 05300 México, D.F., Mexico
c
Universidad Autónoma del Estado de Morelos, Facultad de Farmacia, Av. Universidad 1001 Chamilpa, Cuernavaca C.P. 62209, Morelos, Mexico

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

Article history: An integrated reactive distillation process for biodiesel production is proposed. The reactive separation
Received 1 October 2015 process consists of two coupled reactive distillation columns (RDCs) considering the kinetically controlled
Received in revised form reactions of esterification of the fatty acids (FFA) and the transesterification of glycerides with methanol,
29 December 2015
respectively. The conceptual design of the reactive distillation columns was performed through the con-
Accepted 10 January 2016
Available online xxx
struction of reactive residue curve maps in terms of elements. The design of the esterification reactive
distillation column consisted of one reactive zone loaded with Amberlyst 15 catalyst and for the trans-
esterification reactive column two reactive zones loaded with MgO were used. Intensive simulation of
Keywords:
Reactive distillation the integrated reactive process considering the complex kinetic expressions and the PC-SAFT EOS was
Biodiesel production performed using the computational environment of Aspen Plus. The final integrated RD process was able
Conceptual design to handle more than 1% wt of fatty acid contents in the vegetable oil. However, results showed that the
Reactive residue curve maps amount of fatty acids in the vegetable oil feed plays a key role on the performance (energy cost, catalyst
Element concept load, methanol flow rate) of the integrated esterification–transesterification reactive distillation process.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction substitute fuel. Biodiesel, made from vegetal oils or animal fats, has
a renewable nature and a lower contribution to global warming due
After a century where petroleum has dominated as the preferred to its almost-closed carbon cycle. Analysis has shown that, with the
energy source and as raw material for the production of organic use of biodiesel, CO2 emissions can decrease up to 78%. This biofuel
chemicals, the continuous reduction of oil reservoirs, its increas- is conventionally produced through batch or continuous trans-
ing extraction cost, the emergence of new economies where oil esterification of highly refined vegetable oils with methanol by
demand is increasing exponentially and natural disasters in pro- using homogeneous alkaline catalysts such as sodium or potassium
ducing areas are forcing the oil consumer countries to redefine their hydroxides (Fukuda et al., 2001). Waste cooking oils can also be
energy strategies. In Europe, where oil demand largely exceeds used for biodiesel production, reducing the related production costs
domestic production, the search for new oil substitutes has been and problems concerning the disposal of waste cooking oils that can
in the agenda of both the European Commission and the member cause environmental water contamination. However, the conven-
state governments (Van Gerpen, 2007). Environmental concerns tional technology is not compatible with oils which free fatty acids
are also limiting petroleum based fuels consumption. We have been (FFAs) content exceeds a threshold value of about 1% wt. A possible
witnessing an intensification of air pollution and global warming solution to this drawback could be the development of new tech-
problems due to the emissions of greenhouse gases and other air nologies enabling to employ waste raw materials such as fried oils
contaminants that are produced from the combustion of fossil fuels. or mixture of oils from various sources that cannot be treated in the
As a result of all these issues, studies are nowadays being focused conventional process due to their high content of free fatty acids
on reliable alternatives to conventional diesels. Biodiesel, a blend (Kiss and Bildea, 2012). This perspective discloses the way towards
of fatty acid alkyl esters, is now seen as a sustainable short-term the development of innovative biodiesel production processes such
as those based on heterogeneous two-step reactive-separation pro-
cess, supercritical processes, catalytic hydrodeoxygenation (HDO)
∗ Corresponding author. Tel.: +52 5558044645; fax: +52 5558044900. processes and membrane technology (Yunus khan et al., 2014).
E-mail address: espc@xanum.uam.mx (E.S. Pérez-Cisneros). Recently, Boon-anuwat et al. (2015) presented a study about the

http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
0098-1354/© 2016 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
2 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

design of a continuous reactive distillation process for biodiesel process is used, this is: (a) to perform an analysis of the reactive
production comparing the homogeneous and heterogeneous cat- system in terms of thermodynamic properties, kinetics and catalyst
alyst. They use sodium hydroxide as the homogeneous catalyst available to design a feasible RD process; (b) to perform a con-
and magnesium methoxide as the heterogeneous catalyst for the ceptual design through the computation of reactive residue curve
transesterification of triglyceride with methanol into the reactive maps based on the element concept proposed by Pérez-Cisneros
distillation column. The conclusions from their simulations are et al. (1997) for each occurring chemical reaction; and c) to perform
that the reactive distillation process with the heterogeneous cat- intensive simulation to obtain a feasible and flexible reactive dis-
alyst offers advantages over the conventional process because it tillation column designs and to determine the operating conditions
could eliminate the requirement of post-processing separation and of the integrated process.
purification at cost effective column design and operating con-
ditions. Despite these relevant results, the above work is based 2. The reactive system
only on the transesterification reactions, thus, a pure vegetable
oil is assumed to perform the simulations and a wasted vegetable To describe the reactive system we incorporate the element con-
oil could not be managed. On the other side, Gomez-Castro et al. cept in order to reduce the multicomponent composition space.
(2013) proposed a two-step modified Saka–Dadan reactive distilla- Also, the appropriate thermodynamic model and kinetic expres-
tion process to produce biodiesel with supercritical methanol. They sions to obtain reliable simulation results are considered.
showed that, in terms of production costs and pollutant emission,
the use of reactive distillation for biodiesel production at high pres- 2.1. Chemical reactions and element representation
sure and temperature appears to be an interesting and convenient
alternative. However, it is noticed at this work that the operat- The generic esterification reaction of a carboxylic acid (oleic acid
ing parameters of the reactive distillation column are ambiguously in our case) with methanol, producing methyl oleate (biodiesel) and
extreme for a reactive distillation column setup. Therefore, it is water, is schematically shown below:
clear that additional research and analysis are required to deter-
Amberlyst15
mine if such system could be successfully operated in practice C17 H33 COOH+CH3 OH ←→ C19 H36 O2 + H2 O (1)
and industrial scale. Also, from the experimental point of view, Oleic acid Methanol Methyl Oleate Water

Prasertsit et al. (2013) demonstrated that reactive distillation is a The global transesterification reaction of triolein with methanol
feasible, flexible and reliable alternative to produce biodiesel via to produce three moles of methyl oleate has been selected as a
the transesterification of palm oil with methanol. They used a 4 m model system for the transesterification of a vegetable oil to pro-
long stainless steel column with inside diameter of 1–2 cm, con- duce biodiesel.
nected with 2 L reboiler and found that, for an oil feed flow rate of
Magnesium Oxide
15 ml/min, the optimal condition is at reboiler temperature of 90 ◦ C, C57 H104 O6 +3CH3 OH ←→ 3C19 H36 O2 +C3 H8 O3 (2)
methanol to oil molar ratio of 4.5:1 and the catalyst (KOH) amount Triolein Methanol Methyl Oleate Glycerol

of 1% wt, respect to oil. However, the excess liquid base catalyst with the step by step transesterification reactions as:
should beware for the saponification to form soap in the process.
Magnesium Oxide
Despite the interesting experimental results obtained, the work C57 H104 O6 +CH3 OH ←→ C39 H72 O5 + C19 H36 O2 (3)
deals with a palm oil containing only approximately 0.3–0.6% free Triolein Methanol Diolein Methyl Oleate
fatty acid and less than 1% moisture content, thus, a used vegetable
Magnesium Oxide
oil with higher content of free fatty acid was not considered. C39 H72 O5 + CH3 OH ←→ C21 H40 O4 + C19 H36 O2 (4)
Diolein Methanol Monoolein Methyl Oleate
In the heterogeneous two-step reactive distillation process, the
FFA in the feed is first esterified by acid catalysis to produce one Magnesium Oxide
type of mono-alkyl esters. This esterification step prevents soap C21 H40 O4 +CH3 OH ←→ C3 H8 O3 + C19 H36 O2 (5)
Monoiolein Methanol Glycerol Methyl Oleate
formation during the second reactive separation step, which is the
transesterification of the unreacted triglyceride with alkali or acidic By introducing the element concept proposed by Pérez-Cisneros
catalyst to obtain alkyl esters. This two-step process may be able et al. (1997) the esterification and transesterification reactions
to reduce the overall processing time and it permits processing described in equations (1–5) can be represented in terms of ele-
of feed stocks with high FFA content under normal pressure ments. Table 1 shows the elements definition and the balanced
and temperature operating conditions. For both, the esterifica- reactions in terms of such elements. Also, Table 1 shows the element
tion and transesterification reactions researchers have explored mole fractions which are given in terms of the different species
the use of heterogeneous catalysts. Tesser et al. (2010) studied the participating in each reaction. It should be clear that, while the ele-
transesterification of soybean oil with both basic and acidic het- ments A and C are the same for both reactions, the element B for the
erogeneous catalysts. For esterification reaction the ion exchange esterification reaction means water and for the transesterification
resin Amberlyst 15 proved to be highly effective. On the other reaction means C3 H8 O3 , thus, it is necessary to use a separated ele-
side, Dossin et al. (2006) proposed to use MgO as solid catalyst for ment triangular diagram to represent each reaction in a triangular
the production of biodiesel heterogeneously in a continuous slurry reactive space. Figs. 1 and 2 show the reactive space and the pure
reactor. They showed that the heterogeneous catalysts may allow component location for the esterification and transesterification
good production levels and it presents an interesting alternative reactions, respectively. Note that the triangular element coordinates
to homogeneous catalysts towards a more environment-friendly of the pure components can be straightforward obtained by using
process. Besides, it is generally accepted that a process employ- the element mole fractions definitions and it is done just assign-
ing a solid acid catalyst, if technically feasible, would be preferred ing the unity value to the pure component mole fraction of each
on economic grounds. Thus, in this work, an integrated heteroge- species.
neous two-step reactive distillation process for the production of
biodiesel is investigated. Considering the above, the objective of 2.2. Pure component properties
this work is to develop an integrated reactive distillation process
able to handle a raw material with high FFA content (wasted veg- Due to the great variety of esters species and, in partic-
etable oil) to produce high purity biodiesel. In order to reach this ular, triglycerides, group contribution and corresponding state
objective a systematic procedure for designing of the integrated approaches are the usual methods for the estimation of pure

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx 3

Table 1
Representation of the reactive system in terms of elements.

Reaction Species Elements Reactions in terms of elements Element mole fractions

(1) Oleic Acid A = C18 H32 O WA = (x1 + x3 )/(2x1 + x2 + 2x3 + x4)


Esterification (2) Methanol B = H2 O AB + C = AC + B WB = (x1 + x4 )/(2x1 + x2 + 2x3 + x4 )
(3) Methyl Oleate C = CH3 OH WC = (x2 + x3 )/(2x1 + x2 + 2x3 + x4 )
(4) Water
(1) Triolein A = C18 H32 O
(2) Methanol B = C3 H8 O3 A3 B + C = A2 B + AC WA = (3x1 + x3 + 2x4 + x5 )/(4x1 + x2 + 2x3 + 3x4 + 2x5 + x6 )
Trans esterification (3) Methyl Oleate C = CH3 OH A2 B + C = AB + AC WB = (x1 + x4 + x5 + x6 )/(4x1 + x2 + 2x3 + 3x4 + 2x5 + x6 )
(4) Diolein AB + C = B + AC WC = (x2 + x3 )/(4x1 + x2 + 2x3 + 3x4 + 2x5 + x6 )
(5) Mono Olein
(6) Glycerol

0,0 C= Methanol properties of fatty acids, fatty acid methyl esters, and triglycerides
1,0 using group contribution approaches and applied them to den-
sity and viscosity predictions. Zéberg-Mikkelsen and Stenby (1999)
Pure components developed a group-contribution model for the prediction of melting
0,2
0,8 points and enthalpies of fusion for saturated triglycerides. Ceriani
and Meirelles (2003) presented a group-contribution method for
A=C18H32O the estimation of the vapor pressures and heats of vaporization
0,4
0,6 B=H2O of fatty compounds. Díaz-Tovar et al. (2011) estimated a variety
WC
A
W

C=CH3OH of pure-component properties for edible oils and biodiesel-related


AC=Methyl Oleate
compounds, based on the approaches of Ceriani and Meirelles
0,6
0,4 (2003) and Marrero and Gani (2001) and their respective exten-
sions. In the present work, the pure component properties were
Reactive zone determined using the Marrero and Gani group contribution method
0,8 (2001) incorporated into the Integrated Computer Aided System
0,2
(ICAS) Platform (CAPEC, 2015). Table 2 shows the values of the pure
component properties and the PC-SAFT equation of state parame-
1,0 AB=Oleic Acid
A 0,0 ters used for the phase equilibrium calculations.
B=Water
0,0 0,2 0,4 0,6 0,8 1,0
WB
2.3. Kinetic expressions
Fig. 1. Element representation of the esterification reaction.
The heterogeneous kinetic models employed for the esterifica-
tion reaction and for the transesterification reaction are reported
0.0 C= Methanol
1.0 by Tesser et al. (2010) using Amberlyst 15 as the heterogeneous
Pure components solid catalyst and by Dossin et al. (2006) using MgO as solid cata-
0.2
lyst, respectively. Table 3 shows the complete set of temperature
0.8 dependent parameters and the kinetic expressions used for the
A=C18H32O simulations in this work. It should be noticed in Table 3 that,
0.4 B=C3H8O3 the calculated equilibrium constants keq,1 , keq.2 , and keq.3 , for the
0.6
C=CH3OH glyceride transesterification reactions of Eqs. (3)–(5) appeared to
WC
A
W

AC=Methyl Oleate vary between 0.98 and 1.02 in the temperature range of 313–343 K,
0.6 thus, it is not surprising that these values are close to 1.0 since the
0.4
reactions are almost thermoneutral and the entropy change is very
Reactive zone small (Dossin et al., 2006). For simplicity, all the equilibrium con-
0.8
A3B=Triolein 0.2 stants were taken equal to 1.0 in the simulations and independent
A2B=Diolein of temperature. Also, Dossin et al. (2006) compared the reaction
1.0
A2B AB=Monoolein rates obtained with the kinetic expressions posed in terms of activ-
A 0.0 ities (equations shown in Table 3) with those expressed only in
0.0
A B
0.2 3 0.4 0.6 0.8 1.0 B=Glycerol
WB terms of concentrations and they found negligible differences. Thus,
the kinetic expressions posed in terms of concentrations can also be
Fig. 2. Element representation of the transesterification reactions. used to properly describe the transesterification reactive system.

component properties. Despite the great variety and practical 2.4. Thermodynamic modelling
importance of these substances families, pure component data are
scarce and have been taken into account to a very limited extent for The thermodynamic modelling of methanol-FFA-triglyceride
the development of thermodynamic models intended for a broad mixtures was performed by means of PC-SAFT equation of state.
variety of substances. The critical properties and PC-SAFT parameters were estimated
One of the reasons is that these components are mostly only with the group contribution method of Marrero and Gani (2001)
available as constituents of complex mixtures and are difficult (see Table 2). The PC-SAFT model has shown its capacity to depict
to obtain in pure form. Therefore, specialized approaches have chain molecule phase behaviour providing better predictions for
been developed for the estimation of pure-component proper- all the system investigated and it is recommended as reference EOS
ties of biodiesel-related substances, and little has been published to represent phase equilibria involved in vegetable oils process
on model comparisons. Sales-Cruz et al. (2010) predicted critical (Annesini et al., 2014). Equilibrium calculations were carried out

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
4 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Table 2
Critical properties of the pure components and PC-SAFT equation of state parameters. *Properties determined using the Marrero and Gani (2001) method included in ICAS
Platform (CAPEC, 2015).

Species Critical properties PC-SAFT parameters

MW (gr/mol) Tb (K) Tc (K) Pc (Bar) w m  (A) ε/k (K) εAB (K) AB

Triolein* 885.43 878.74 985.99 2.68 2.549 23.4907 3.925683 264.6772


Methanol 32.042 337.85 512.5 80.84 0.224 1.5255 3.23 188.9 2899.5 0.035176
Methyl Oleate* 296.49 618.93 768.69 11.81 0.967 8.5753 3.850432 256.3849
Diolein* 621 806.33 1886.12 4.44 2.197 16.3139 4.1207 223.8424
Monoolein* 356.55 696.82 846.0 12.13 1.621 15.4767 3.2410 230.1465
Glycerol* 92.095 507.34 701.79 71.85 1.175 10.3570 2.009283 232.4834
Oleic Acid* 282.47 647.04 810.37 13.40 1.150 10.5829 3.522527 244.0751
Water 18.015 373.15 647.13 220.55 0.344 1.0656 3.0007 366.51 2500.7 0.034868

Table 3
Kinetic expressions for esterification and transesterification reactions.

Reaction Catalyst and temperature dependent parameters Kinetic expressions


HMO C R C R
k1 HAO C R −k−1 MO W
AO CR
Esterification Catalyst: Amberlyst 15 r= R R
MEOH
HAO C H C H CR
AO + MO MO + W W
Tesser et al. (2010) r= [mol min−1 gcat −1 ] 1+
CR CR CR
MEOH MEOH MEOH
CiR = conc. in the resin-absorbed liquid phasek1 =
 cm3
  17.46 (kcal/mol)  1 
9.01 exp − T1

grcat min
cm3
  8.78 (kcal/mol)  1
R 273.15
1

k−1 = 5.50 grcat min
exp R 273.15
− T
HAO = 0.46,
HW = 4.74, HMO = 0.13
 1 aME aDG

kMeOH CMeOH − aTG
keq,1
Trans Esterification Catalyst: Magnesium oxide (MgO) rTG = KA aME aDG
1+ +KA CDG +KA CMG +KA CG
Dossin et al. (2006) ri = [mol kgcat −1 s−1 ] keq,1

aTG

1 aME aMG
kMeOH CMeOH −
Ci = [mol m−3 ] keq,2 aDG
rDG = KA aME aMG
1+ +KA CDG +KA CMG +KA CG
ai = activity of species i keq,2 aDG
 
 m3
  −20.1x103 (J /mol)   m3  kMeOH CMeOH − 1 aME aG
keq,3 aMG
kMeOH = 0.148 kgcat s
exp RT
KA = 5.29x10−3 mol
rMG = KA aME aG
1+ +KA CDG +KA CMG +KA CG
keq,3 aMG
keq,1 = keq,2 = keq,3 = 1.0

by means of one of the most up-to-date version of the original Where mi and  i correspond to the number of segments which
EOS which is known as perturbed chain SAFT (PC-SAFT). SAFT is constitute the chain of the ith component and to its diameter,
one of the most important association theory which belongs to the respectively, while xi is the component mole fraction. The dis-
general perturbation category. persion term (disp) in Eq. (6), according to Barker and Henderson
It is based on the Wertheim’s first order perturbation theory (1967), consists of a first and second order contributions
(Wertheim, 1984) and it has been implemented into useful expres-
Adisp A1 A2  ε 
sions in several studies. SAFT EOS considers molecules as chains ij
= + = −2I1 (, m̄) xi xj mi mj ij3 −
composed of freely jointed spherical segments and the basic idea NkT NkT NkT kT
i j
underlying the model is the definition of an interaction potential to

−1 
represent attraction and repulsion forces between segments. The ∂Z hc
molecular model can be refined by assuming the segments to have − m 1 + Z hc +  I2 (, m̄) xi xj mi mj
∂
association sites and partial charges enabling the chains to associate i j
or repulse by hydrogen bonds and dipole interactions. Conse-  ε 2
ij
quently, the residual Helmholtz free energy, is made up by terms × ij3 (8)
stemmed from each of these contributions. The PC-SAFT EOS main kT
innovation relies on considering perturbation respect to spherical where Zhc is the hard-chain contribution expressed in terms of
segments which are connected into rigid chains (hard-chain term) compressibility factor,  is a packing factor and I1 and I2 are inte-
rather than disconnected segments (Gross and Sadowsky, 2001). gral terms. Expressions of these terms depend on the choice of the
Obviously, this means considering interactions between chains interaction potential, which in PC-SAFT model is assumed to be a
instead of single segments and this would allow to capture the modified version of the classical square-well suggested by Chen
behaviour of chain molecule, such as hydrocarbons and polymers, and Kreglewski (1977). The interaction potential is then charac-
more realistically. In this way, the residual Helmholtz free energy terized by the three parameters, mi ,  i and εi , which is the well
can be written as: potential depth. For definition and exhaustive derivation of all the
Ares Ahc Adisp Aassoc terms appearing in Eqs. (6)–(9) it should be reviewed the reference
= + + (6) of Gross and Sadowsky (2001). Mixtures extension provided in Eq.
NkT NkT NkT NkT
(8) requires definition of mixing rules for a pair of unlike segments,
In Eq. (6) the hard-chain (hc) reference term accounts for hard- which according to Berthelot–Lorentz, results may be given as:
sphere (hs) and chain formation contributions
1
  ij = ( + j ), εij = εi εj (1 − kij ) (9)
A hc
=Ahs
xi mi + xi (1 − mi ) ln giihs (ii ) (7) 2 i
i i In this work, binary interaction parameters were not included.

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx 5

1.0 1.0
0.9 Methyl Oleate 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5

X
0.4 0.4
0.3 Methanol 0.3
0.2 0.2
0.1 Oleic Acid 0.1
0.0 0.0
Water
-0.1 -0.1
330 360 390 420 450 480 510 540 570 600
T (K)

Fig. 3. Residue curve map for esterification reaction in terms of elements. Fig. 4. Liquid composition and temperature profile for the residue curve at ˛ = 40.

3. Conceptual design different values of the reaction-separation parameter (˛), which


represents the amount of catalyst loaded into the boiling pot. It
The reactive residue curve maps are a very valuable tool to can be observed in Fig. 3 that there not exist reactive azeotropes
obtain conceptual designs of reactive distillation columns. Such or reactive separation boundaries and, as the amount of catalyst
maps allow us to verify the existence of stable nodes, i.e., reactive loaded increases (˛ = 0.0001–40) the residue curves tend to the AC
azeotropes or reaction-separation boundaries. Also, it is possible to node, which represents the methyl oleate (MO) pure compound.
predict the limits of the operating temperature and pressure of the Also, it can be noted in Fig. 3 that there is a sharp decrease in the
reactive column where a specified conversion could be reached. In methanol composition, even with a very low amount of catalyst
this work, the residue curve maps were calculated for each reaction loaded (˛ = 0.0001), because this compound has the lower boiling
scheme separately following the procedure proposed by Granados- point. However, as the catalyst load increases (˛ = 3, 10, 40), this is,
Aguilar et al. (2008). The heterogeneous kinetic models described in methyl oleate is produced, the slopes of the reactive residue lines
the above section are incorporated into the autonomous differential are diminished pointing to the AC node. From the triangular residue
equations for computation of the reactive residue curves as: curve map shown in Fig. 3, it can be concluded that, even with
 NC    low amounts of catalyst loaded and a rich methanol reacting mix-
dxi  NR
 rj ture, a reactive separation process to produce pure methyl oleate
= xi − yi + ˛
ij −
kj xi (10)
d r0 (biodiesel) is feasible.
j=1 k=1
Fig. 4 shows the corresponding liquid composition and temper-
with the dimensionless reaction-separation parameter ˛ given by ature profile for the residue curve with ˛ = 40. It is interesting to
observe in Fig. 4 that the minimum temperature of the reacting
Mcat r0
˛= (11) mixture is at the initial composition point, this is 337 K, which
V0 is close to the boiling point of pure methanol, while the maxi-
In Eq. (10) NC is the number of components participating in reac- mum temperature is around 585 K, which does not correspond to
tion j, NR is the number of reactions, Mcat is the catalyst mass (kg), the boiling point of methyl oleate, rather such temperature value
V̄0 is the vaporization flow (kmol/h), rj is the intrinsic rate of reac- corresponds to a boiling mixture containing methyl oleate, oleic
tion j (kmol/kg cat. h), r0 is a reference rate of reaction and
ij is the acid, water and insignificant amount of methanol (xOA = 0.3523,
stoichiometric coefficient of compound i in reaction j. The reaction- xMeOH = 1E−08, xMO = 0.96161, xW = 0.00215). It should be clear that
separation parameter ˛ indicates the ratio between the amount of the higher production rate of methyl oleate is between the temper-
catalyst loaded in the distillation vessel (i.e., total wetted reacting atures 338–420 K and after this value the methanol vaporizes and it
area) multiplied by the reference rate of reaction to the vaporiza- is completely consumed at 585 K, remaining in the boiling pot a non
tion flow. In fact, the reaction-separation parameter ˛ indirectly reactive mixture. From Fig. 4 it can be concluded that, at P = 1 atm,
accounts for a liquid residence time. To carry out the integration of the minimum and maximum operating temperatures of a reactive
Eq. (10) it is necessary to calculate: (i) the phase equilibrium and (ii) distillation column considering only oleic acid and methanol as the
the reaction terms. For the computational implementation of the reactants are: 338 K at the top (the lowest temperature) and 590 K
phase equilibrium calculations, the PC-SAFT source code developed at the bottom (the highest temperature). It should be pointed out
by Sales-Cruz (2006) with the required parameters given in Table 2, that the conceptual design of such reactive distillation column does
has been used. For the calculation of the reaction terms the kinetic not consider the presence of triglycerides.
expressions given in Table 3 are evaluated.
3.2. Reactive residue curve maps. Transesterification reaction
3.1. Reactive residue curve maps. Esterification reaction
For the transesterification residue curve map a reactive mix-
For the esterification residue curve map a reactive mixture ture with excess of methanol was used as initial composition
with excess of methanol was used as initial composition point point (xTG = 0.04, xMeOH = 0.95, xMO = 0.004, xDG = 0.002, xMG = 0.002,
(xOA = 0.006, xMeOH = 0.99, xMO = 0.002, xW = 0.002). (OA) represents xG = 0.002). (TG) represents the triolein species, (MeOH) methanol,
the oleic acid, (MeOH) methanol, (MO) methyl oleate and (W) indi- (MO) methyl oleate, (DG) diolein, (MG) monoolein and (G) indicates
cates water compounds. Fig. 3 shows the residue curve map for glycerol compounds. Similar to the esterification reaction, it can be

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
6 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Finally, the minimum and maximum operating temperatures at


P = 1 atm of a conceptual reactive distillation column can be estab-
lished from Fig. 6. The temperature at the top of the column would
correspond to the boiling point of pure methanol (338 K) and the
temperature at the bottom of the column (523 K) would correspond
to a boiling mixture of methyl oleate, triolein, diolein, monoolein
and glycerol (xTG = 0.10802, xMeOH = 1.19E−13, xMO = 0.82872,
xDG = 0.01876, xMG = 0.01875, xG = 0.02574). It should be noted that
the amount of catalyst (˛ = 0.01) is not enough to convert the initial
amount of triolein, thus, the maximum operating temperature to
obtain pure methyl oleate cannot be established.

4. Intensive simulation

All simulations of the reactive distillation columns were carried


out with the steady state model RADFRAC in Aspen Plus environ-
Fig. 5. Residue curve map for transesterification reaction. ment. Stages are numbered from the top to the bottom, with stage
1 as condenser and stage N as reboiler. The chemical reactions take
1.0 1.0 place in the liquid phase and each reactive stage is treated as a per-
fectly mixed continuous stirred tank reactor. Thus, mass and heat
0.9 0.9
transfer issues are neglected. The oil feed to the reactive distillation
0.8 0.8 columns was considered as a mixture of triolein and oleic acid rep-
Methyl Oleate resenting a “simplified used” non-edible vegetal oil. Also, a target
0.7 0.7
Methanol conversion (99.9%) of the fatty acid and the triglyceride was set as a
0.6 0.6
design constraint. Additionally, the PC-SAFT EOS is used to calculate
0.5 0.5 the phase equilibrium and thermodynamic properties.
X

0.4 0.4
0.3
Glycerol
0.3 4.1. Intensive simulation. Esterification RDC
0.2 Triolein 0.2
To determine a feasible structure of the esterification (first step)
0.1 0.1
Monoolein
reactive distillation column, three basic configurations of RDC are
0.0 Diolein 0.0 proposed: (i) a RDC with one reactive zone loaded with solid cata-
-0.1 -0.1 lyst (Amberlyst 15) considering a total condenser; (ii) a RDC with
340 360 380 400 420 440 460 480 500 520 the same reactive zone considering a partial condenser, and (iii) a
T(K) RDC without condenser (no reflux). Fig. 7 shows the basic configu-
rations of the three reactive distillation columns.
Fig. 6. Liquid composition and temperature profile for the residue curve at ␣ = 0.01. Once defined the basic conceptual configurations of the RDCs
and the target conversion of oleic acid (99.9%), a feed flow rate of
observed in Fig. 5 that, as the amount of catalyst loaded increases pure oleic acid (FOA = 100 kmol/h) and an equimolar methanol feed
(˛ = 1.0E−06–0.01) the residue curves tend to the AC node, which (FMeOH = 100 kmol/h) were established as the reactant feed flow
represents the methyl oleate (MO) pure compound. Also, in can be rates. Also, it is considered that 1000 Kg of catalyst is loaded at each
noted in Fig. 5 that there is not any reactive azeotrope and reac- reactive stage. Intensive simulations were performed to determine
tive separation boundary. The trend of the residue curves indicates the minimum number of total stages (reactive and non-reactive) of
that, with a ˛ value greater than 0.01 a complete transformation the RDC and the less energy cost with the three configurations.
of triolein could be reached. However, this conversion would be The intensive simulation was carried out by increasing sequen-
very hard to achieve since the transesterification reactions are tially the number of reactive stages until the target conversion is
a consecutive reactive system. This is, while triolein is reacting achieved. Afterwards, in order to get the desired purity (methanol
with methanol, the production of diolein, monoolein and glyc- at the top and methyl oleate at the bottom), the number of rectifying
erol increases, rendering a mixture containing such compounds. and stripping non-reactive stages were varied. It should be men-
On the other side, it should be mentioned that the values of the tioned that around 300 simulations of each RDC configuration were
reaction-separation parameter (˛), in this case, is three orders of needed to obtain the best structure that clearly fulfil the design con-
magnitude lower than the ˛ values used in the esterification case. straints. Tables 4a–4c summarizes the intensive simulation results
The above indicates that the reaction rates for the transesterifica- for the three RDC configurations.
tion reactions are higher than the esterification one. Therefore, a In Table 4a it can be noted that the run Id: S10-9 shows the
lower amount of catalyst may be expected in the transesterifica- highest conversion of oleic acid, which means that methanol was
tion case. Fig. 6 shows the corresponding liquid composition and completely consumed and the lowest temperature at the top of the
temperature profile for the residue curve with ˛ = 0.01. It can be RDC correspond to the water boiling point. From Table 4b it can be
noted in Fig. 6, in opposition to the esterification reaction, that the observed that the maximum conversion of oleic acid is achieved in
conversion to methyl oleate increases continuously from 340 K to the run Id: S5-14 with 24 total stages and from Table 4c the run
490 K, and as the methanol is consumed roughly completely, vapor- Id: S3-16 presents the maximum conversion with 19 total stages.
ization of some glycerol occurs leading to an increment in the liquid Table 5 shows the final design specifications for the three RDC con-
composition of methyl oleate. It should be pointed out that the figurations after intensive simulation. It can be noted in Table 5 that
final mixture in the boiling pot for the transesterification reaction the best configuration is the RDC without condenser (no reflux),
is not pure methyl oleate, rather a mixture containing some triolein, consisting of 19 total stages (15 reactive stages and 4 separation
diolein and monoolein should be expected. stages).

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx 7

Fig. 7. Basic configurations of RDC for the esterification reaction (first step).

Table 4a
Summarized simulation results for the esterification RDC configuration (a): RDC-with total condenser. Feed ratio 1:1 acid/Methanol (kmol/h). Reflux ratio = 1.0. Catalyst load
of 1000 kg at each reactive stage. OA, oleic Acid; MeOH, methanol.

Input Results

Run Id Stages Feed stage Feed Boilup ratio Oleic Acid Methyl Oleate Temperature (◦ C)
Oleic acid Temperature Conversion Flow rate
(◦ C) (kmol/hr)

Reactive Rec. Strip. Total OA MeOH Top Total Top Bottom

Total From-to

S1-9 11 3–13 2 2 15 2 350 55 1 0.88782 0.222 88.792 87.5 343.4


S2-2 11 3–13 2 2 15 2 350 55 1 0.88782 0.222 88.782 87.5 343.4
S3-5 11 3–13 2 2 15 2 350 55 1.75 0.9175 0.199 91.75 88.4 343.7
S4-2 11 3–13 2 2 15 2 350 55 1.75 0.9175 0.199 91.75 88.4 343.7
S5-5 11 3–13 2 2 15 2 326 55 1.75 0.93503 0.049 93.503 91.8 343.8
S5-16 11 3–13 2 2 15 2 326 60 1.75 0.93553 0.048 93.553 91.8 343.8
S6-8 13 2–14 1 1 15 2 326 60 1.75 0.9502 0.366 95.02 92.7 343.7
S7-1 16 3–18 2 2 20 1 326 60 1.75 0.94341 0.044 94.341 92.3 343.7
S8-5 13 11–23 10 2 25 10 326 60 1.75 0.9939 1.438 99.39 94.5 343.8
S9-5 17 12–28 11 2 30 12 326 55 1.75 0.99738 1.732 99.738 93.7 343.8
S9-9 16 15–30 14 2 32 15 326 55 1.75 0.99802 1.732 99.802 93.7 343.8
S9-12 15 17–32 16 2 34 17 326 55 1.75 0.9982 1.832 99.82 93.7 343.8
S10-1 16 15–30 14 2 32 14 326 55 1.75 0.99146 0.478 99.146 98.9 343.7
S10-9 19 14–32 13 2 34 14 322 55 1.7 0.99929 0.047 99.929 99.8 343.7

Table 4b
Summarized simulation results for the esterification RDC configuration (b): RDC-with partial condenser. Feed ratio 1:1 Oleic acid/Methanol (kmol/h). Reflux ratio = 1. Catalyst
load of 1000 kg at each reactive stage, Oleic acid feed temperature = 326 ◦ C. 2 stripping stages.

Input Results

Run Id Stages Feed stage Reflux ratio Feed Boilup ratio Oleic acid Methyl oleate Temperature (◦ C)
Oleic acid Temp. (◦ C) conversion flow rate
MeOH (kmol/h)

Reactive Rec. Total Top Total Top Bottom

Total From-to

S1-4 11 3–13 2 15 2 0.4 55 1 0.83711 0.295 83.711 99.9 343


S2-8 11 3–13 2 15 2 0.4 55 1.7 0.88631 0.242 88.631 101 343.24
S3-3 11 3–13 2 15 2 0.9 55 2 0.95206 0.059 95.206 98.6 343.3
S4-3 12 4–15 3 17 4 0.9 60 2 0.97145 2.031 97.145 99.4 343.4
S4-6 11 5–16 4 18 5 0.9 60 2 0.97443 2.341 97.443 99.5 343.3
S4-9 11 7–18 6 20 7 0.9 60 2 0.9787 2.788 97.87 99.6 343.3
S4-16 11 7–20 6 22 7 0.9 55 2 0.9789 2.726 97. 89 99.6 343.3
S4-18 11 9–22 8 24 9 0.9 55 2 0.98164 3.015 98.164 99.7 343.3
S5-14 11 9–22 8 24 9 1.2 60 1.75 0.99901 0.015 99.901 99.6 343.3

4.2. Intensive simulation. Transesterification RDC MgO solid catalyst considering one methanol feed; (ii) a RDC with
two reactive zones considering two methanol feeds. Fig. 8 shows
To determine a feasible structure of the transesterification (sec- the basic configurations of the two reactive distillation columns.
ond step) reactive distillation column two basic configurations of The same systematic procedure to determine the best RDC struc-
RDC are proposed: (i) a RDC with one reactive zone loaded with ture of the esterification step was followed to achieve the complete

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
8 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Table 4c
Summarized simulation results for the esterification RDC configuration (c): RDC-without condenser. Feed ratio 1:1 Oleic acid/Methanol (kmol/h). Reflux ratio = 1. Catalyst
load of 1000 kg at each reactive stage, Oleic acid feed temperature = 326 ◦ C. Methanol feed temperature = 55 ◦ C. 2 stripping stages.

Input Results

Run Id Stages Feed Stage Boilup ratio Oleic Acid Methyl Oleate Flow Temperature (◦ C)
Oleic Acid Conversion rate (kmol/hr)

Reactive Rec. Total Top Total Top Bottom

Total From-to

S1-3 11 3–13 2 15 2 1 0.90744 0.001 90.744 147.29 343.7


S2-7 11 3–13 2 15 2 1.6 0.9717 0.001 97.17 154.59 343.62
S3-2 11 4–14 3 16 3 1.9 0.99498 Trace 99.498 100 343.4
S3-8 11 5–15 4 17 4 2 0.99545 <0.001 99.545 104.9 343.4
S3-13 14 3–16 2 18 3 2 0.99885 0.001 99.885 126.2 343.4
S3-16 15 3–17 2 19 3 2 0.99926 <0.001 99.926 99.6098 343.4

Table 5
Intensive simulation. Final design specifications for the three reactive distillation columns after intensive simulation of the esterification step.

Design variable RDC-total condenser RDC-partial condenser RDC-NO condenser

Column pressure (bar) 1 1 1


Total Stages 34 24 19
Rectifying stages 13 8 2
Reactive stages 19 (from 14 to 32) 14 (from 9 to 22) 15 (from 3 to 17)
Stripping stages 2 2 2
Boilup ratio 1 1.75 2
Oil Feed Stage 14 9 3
Methanol feed stage 33 23 18
Oleic acid feed flow (kmol/h) 100 (preheated 322 ◦ C) 100 (preheated 326 ◦ C) 100 (preheated 326 ◦ C)
Methanol fed flow (kmol/h) 100 (preheated 55 ◦ C) 100 (preheated 60 ◦ C) 100 (preheated 55 ◦ C)
Catalyst load (kg cat) per reactive stage 1000 1000 1000
Bottom molar flow of Biodiesel (kmol/h) 99.881 99.848 99.926
Condenser heat duty (kW) −2415.7929 −1836.2888 −2079.0956
Reboiler heat duty (kW) 4568.6906 5047.284 4655.9723
Condenser temperature (◦ C) 99.844 125.980 99.6098
Reboiler temperature (◦ C) 343.733 343.37 343.31

conversion of triolein using, in this case, the two basic configura- 5. The integrated reactive distillation process
tions proposed. Tables 6a and 6b show the summarized intensive
simulation results of the two RDC. It can be noted from those tables With the analysis of the reactive residue curve maps and,
that the best configuration for the esterification reactions is the RDC through intensive simulation of the reactive distillation columns
with two reactive zones and two methanol feeds. Despite the two for each reaction (two-step process) performed in the Aspen Plus
RDC configurations has the same structure (25 total stages and 21 environment, a final integrated reactive distillation process to pro-
reactive stages), the amount of methanol used for complete con- duce biodiesel was obtained. Fig. 9 shows the integrated process
version of triolein is less for the configuration with two methanol considering two reactive distillation columns: (a) a RDC for the
feeds. Table 7 shows the final design specification of the two RDC esterification reaction without condenser and one reactive section
configurations. loaded with a solid catalyst (Amberlyst 15) and, (b) a RDC for the

Fig. 8. Basic configurations of RDC for the transesterification reaction (second step).

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
CACE-5353; No. of Pages 14
G Model
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers

Table 6a
Summarized simulation results for the transesterification RDC configuration (a). RDC-one methanol feed. 3 stripping stages. TO, triolein; DO, diolein; MOO, monoolein; MeOH, methanol.

Run Id Feed ratio TO/MeOH Stages Feed Stage Reflux ratio Feed Triolein Methyl Oleate Flow Temperature (◦ C) Triolein Flor rate Glycerol Flow rate
(kmol/h) temperature Conversion rate (kmol/h) (kmol/h) (kmol/h)
(◦ C)

Reactive Rec. Total TO MeOH TO MeOH Top Total Top Bottom Bottom Top Bottom Top

Total From -to

S1-9 1:6.5 14 4–17 3 20 3 18 1.7 821 51 91.652 4.8E−16 231.19 64.45 326.70 8.35 2.8E−14 57.21 8.9E−11
S2-10 1:6.5 14 4–17 3 20 3 18 1.3 821 51 86.224 3.7E−15 201.90 64.45 180.99 13.77 1.2E−13 42.95 2.7E−10
S3-8 1:6.5 14 4–17 3 20 3 18 1.7 821 55 89.955 8.0E−16 221.62 64.45 170.44 10.04 3.9E−14 52.40 1.1E−10

E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx
S4-8 1:6.5 14 4–17 3 20 3 18 1.7 830 55 89.968 7.9E−16 221.69 64.45 170.38 10.03 3.9E−14 52.44 1.1E−10
S5-6 1:6.5 14 4–17 3 20 3 18 1.7 830 55 91.224 1.7E−17 228.72 64.45 339.78 8.776 2.2E−14 55.91 2.4E−12
S6-1 1:6.5 14 4–17 3 20 2 18 1.7 830 55 90.077 1.4E−14 222.20 64.45 165.28 9.919 3.3E−03 52.51 5.8E−12

ARTICLE IN PRESS
S6-10 1:6.5 16 2–17 1 20 2 18 1.7 830 55 90.241 0.0003 223.13 64.45 164.72 9.754 4.8E−03 52.98 3.8E−04
S7-6 1:15 16 2–17 1 20 2 18 1.7 830 55 98.596 4.5E−05 278.87 64.45 96.02 1.404 2.7E−04 84.08 2.0E−04
S8-5 1:15 16 2–17 1 20 2 18 1.7 830 55 99.561 1.9E−05 289.65 64.45 123.53 0.439 5.1E−05 91.48 1.6E−04
S8-11 1:15 16 2–17 1 20 2 18 3 830 55 99.778 4.2E−06 294.58 64.45 104.49 0.222 3.5E−06 95.37 7.1E−05
S9-5 1:15 21 2–22 1 25 2 23 3 830 55 99.937 1.9E−06 297.16 64.45 329.61 0.06 6.3E−07 97.42 5.8E−05

Table 6b
Summarized simulation results for the transesterification RDC configuration (b). RDC- two methanol feeds. 25 total stages, 3 stripping stages, 1 rectifying stage and 21 reactive stages. TO fed at stage 2 at 830 ◦ C and MeOH at 55 ◦ C.
TO, triolein; DO, diolein; MOO, monoolein; MeOH, methanol.

Run Id Methanol Feed Feed Stage Reflux ratio Triolein Methyl Oleate flow Temperature (◦ C) Triolein flow rate Mono Olein flow Glycerol flow rate
Flow rate (kmol/h) (MeOH) conversion rate (kmol/h) (kmol/h) rate (kmol/h) (kmol/h)

Feed 1 Feed 2 Feed 1 Feed 2 Top Total Top Bottom Bottom Top Bottom Top Bottom Top

S1-3 500 700 6 23 3 99.791 6.5E−06 294.84 64.45 121.89 0.209 9.6E−06 3.87 7.1E−09 95.59 7.9E−05
S1-7 500 700 10 23 3 99.803 6.5E−06 294.90 64.45 121.90 0.197 9.6E−06 3.86 7.1E−09 95.62 7.9E−05
S1-11 500 800 13 23 3 99.803 5.6E−06 295.13 64.45 114.48 0.196 6.7E−06 3.68 5.9E−09 95.82 7.6E−05
S2-2 600 500 6 23 3 99.794 7.8E−06 294.77 64.45 132.83 0.206 1.4E−05 3.94 8.8E−09 95.52 8.3E−05
S2-5 600 500 11 23 3 99.915 7.8E−06 295.56 64.45 133.13 0.085 1.4E−05 3.60 8.8E−09 96.02 8.3E−05
S2-14 1000 500 12 23 3 99.951 4.1E−06 297.28 64.45 104.80 0.049 3.5E−06 2.30 4.1E−09 97.51 7.1E−05
S2-19 1000 500 13 23 3 99.951 4.1E−06 297.28 64.45 104.80 0.049 3.5E−06 2.31 4.1E−09 97.51 7.1E−05
S2-22 1000 500 14 23 3 99.952 4.1E−06 297.28 64.45 104.80 0.048 3.5E−06 2.31 4.1E−09 97.50 7.1E−05
S3-1 600 500 11 23 3.5 99.929 5.3E−06 296.06 64.45 124.24 0.071 7.3E−06 3.2 5.7E−09 96.44 6.7E−05

9
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
10 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Table 7
Intensive simulation. Final design specifications for the two reactive distillation columns after intensive simulation of the transesterification step.

Design variable RDC-one methanol feed RDC-two methanol feeds

Column pressure (bar) 1 1


Total stages 25 25
Rectifying stages 1 1
Reactive stages 21 (from 2 to 22) 21 (from 2 to 22)
Stripping stages 3 3
Reflux ratio 3 3.5
Triolein feed stage 2 2
Methanol feed stage (1) 23 11
Methanol feed stage (2) – 23
Triolein feed flow (kmol/h) 100 (preheated 830 ◦ C) 100 (preheated 830 ◦ C)
Methanol feed flow (1) (kmol/h) 1500 (preheated 55 ◦ C) 600 (preheated 55 ◦ C)
Methanol feed flow (2) (kmol/h) – 500 (preheated 55 ◦ C)
Catalyst load (kg cat) 10,080 (480 per reactive stage) 10,080 (480 per reactive stage)
Bottom molar flow of biodiesel (kmol/h) 297.16 296.06
Condenser heat duty (kW) −47,203.71 −28,006.707
Reboiler heat duty (kW) 60,640.2696 26,853.0692
Condenser temperature (◦ C) 64.45 64.45
Reboiler temperature (◦ C) 329.612 124.24

Fig. 9. Integrated reactive distillation process for biodiesel production.

transesterification reaction with two reactive sections loaded with the temperature profile of the reactive distillation column. Consid-
solid catalyst (MgO) and two methanol feeds. The first reactive ering only the bottom of the column, the temperature goes from
distillation column (esterification step) is fed with a mixture of 349 ◦ C (run Id: EST-1) to 154 ◦ C (run id: EST-13). Therefore, it can be
oleic acid and triolein, modelling a used vegetable oil (86 kmol/h concluded from the simulation results that the energy consumption
of triolein and 14 kmol/h of oleic acid) and 70 kmol/h of methanol. is highly dependent of the triolein-oleic acid mixture concentra-
For the simulation of the integrated reactive process the design tion. Also, the methanol feed flow rate, the boilup ratio and the
goal established were: 99.99% conversion of oleic acid for the catalyst load were manipulated for this RDC. Results of intensive
esterification column and 99.9% conversion of triolein for the trans- simulation (see Table 8) show that the run id: EST-12 has a combi-
esterification column. nation of design and operating variables (800 Kg of catalyst, boilup
As was stated before, the objectives of the present work is to ratio = 0.8 and methanol feed flow rate = 70 kmol/h) which leads to a
develop a reactive distillation process able to deal with mixtures feasible RDC performance in terms of FFA conversion (99.926%) and
representing a used or wasted vegetal oil (triolein and oleic acid), energy consumption. It should be clear that, as the methanol is fed
thus, the effect of the content of FFA in the oil feed, the boil up in an equimolar amount and the target conversion is almost com-
ratio and the methanol feed flow rate, on the performance of the plete, the temperature at the top of the column corresponds to the
esterification RDC, was investigated. For example, first, the reflux water boiling point instead of the boiling point of methanol as was
and boilup ratio were manipulated to achieve a reasonable purity determined with the reactive residue curve map. Also, it should be
of biodiesel. Further, the amount of catalyst was varied until the pointed out that, the final esterification reactive distillation column
conversion of the reactants, the design goals, were reached. Also, design obtained after intensive simulation could manage vegetable
considering the above, different feed compositions were tested to oils with a higher content of free fatty acids than 1% wt.
obtain a robust configuration of the integrated reactive distillation Once the best structure of the RDC for the transesterification
column. Table 8 shows the results simulations obtained with the reaction is found (Section 4.2), intensive simulation of the trases-
RDC configuration without condenser. It can be noted in Table 8 terification RDC in the integrated reactive process is performed to
that the amount of FFA in the oil feed was varied from 0.742 to 0.045 observe the influence of some operating variables. Table 9 shows
weight fraction and how this composition change affects strongly the effect of the reflux ratio, the boilup ratio and the methanol feed

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx 11

Table 8
Intensive Simulation of the integrated reactive distillation process. Effect of the FFA (oleic Acid) content in the oil feed on the esterification reactive distillation column.

Run Id FFA (Oleic Acid) wt fraction in the feed Boilup ratio Methanol feed flow (kmol/h) Biodiesel flow (kmol/h) Temperature (◦ C)

Bottom Top Bottom Top

EST-1 0.742 2 90 89.950 0.0003 349.16 117.75


EST-2 0.561 2 80 79.938 0.0031 355.98 136.78
EST-3 0.427 1.8 70 69.933 0.0093 364.06 150.45
EST-4 0.324 0.8 70 55.414 2.9E−08 197.55 100.02
EST-5 0.242 0.8 130 50.00 3.1E−08 137.70 94.13
EST-6 0.175 0.8 70 40.00 7.3E−08 173.95 99.58
EST-7 0.120 0.8 70 30.00 2.4E−08 165.72 98.68
EST-8 0.074 0.8 70 20.00 8.3E−09 158.64 98.63
EST-9 0.053 (1000 kg cat) 0.8 70 15.00 4.5E−09 155.56 99.26
EST-10 0.053 (900 kg cat) 0.8 70 15.00 4.3E−09 155.56 99.26
EST-11 0.053 (800 kg cat) 0.8 70 14.999 4.1E−09 155.56 99.26
EST-12 0.049 (800 kg cat) 0.8 70 13.999 3.1E−09 154.98 99.98
EST-13 0.045 0.8 70 12.999 1.7E−09 154.41 99.76

Table 9
Intensive simulation of the integrated reactive distillation process. Effect of the boilup ratio, reflux ratio and methanol feed flow rate on the transesterification reactive
distillation column.

Run Id Reflux ratio Boilup ratio Methanol feed flowrate 1 (kmol/h) Biodiesel flow (kmol/h) Temperature (◦ C)

Bottom Top Bottom

TES-1 3 3.5 400 268.38 7.1E−06 331.2


TES-2 3 3.5 500 269.22 5.1E−06 330.7
TES-3 3 3 400 268.49 7.1E−06 330.6
TES-4 3 3 500 269.27 5.1E−06 329.8
TES-5 3 2.5 400 268.56 7.1E−06 329.2
TES-6 3 2.5 500 269.31 5.1E−06 326.9
TES-7 2.5 2.5 400 267.57 1.3E−05 288.5
TES-8 2.5 2.5 500 268.61 9.6E−06 300.0
TES-9 3.5 2.5 400 269.22 4.2E−06 327.3

Table 10
Final design specifications of the integrated reactive distillation process for biodiesel production.

Design variable RDC-NO condenser RDC-two methanol feeds

Column pressure (bar) 1 1


Total stages 19 25
Rectifying stages 2 1
Reactive stages 15 (from 3 to 17) 21 (from 2 to 22)
Stripping stages 2 3
Reflux ratio – 2.5
Boilup ratio 0.8 3
Oil feed stage 3 2
Methanol feed satge (1) 18 11
Methanol feed stage (2) – 23
Feed flow (Kmol/h) 14(Oleic Acid), 86 (Triolein) 85.9999 (Triolein)
Methanol feed flow 1 (kmol/h) 70 (preheated 55 ◦ C) 400 (preheated 55 ◦ C)
Methanol feed flow 2 (kmol/h) – 500 (preheated 55 ◦ C)
Catalyst load (kg cat) 800 (per reactive stage) 10,080 (480 per reactive stage)
Bottom molar flow of Biodiesel (kmol/h) 13.999 268.597
Condenser heat duty (kW) −132.35 −27,521.403
Reboiler heat duty (kW) 117.0387 36,953.4047
Condenser temperature (◦ C) 99.984 64.45
Reboiler temperature (◦ C) 154.98 329.84

Table 11
Stream molar flowrates, temperature and compositions of the integrated reactive distillation process.

Stream 1 2 3 4 5 6 7 8 9 10 11 12 13

Molar flow (kmol/h) 100 70 14.234 155.7695 55.7447 13.9888 86 400 500 700.97 354.766 268.77 85.996
Temperature (◦ C) 326 55 99.984 154.98 64.4962 344.23 830 55 55 64.45 329.24 301.54 301.54
Pressure (atm) 1 1 1 1 1 1 1 1 1 1 1 1 1

Liquid mole fraction


Triolein 0.86 7.35E−09 0.55209 7.03E−13 2.96E−05 0.999999 7.92E−09 0.000158 1.94E−10 0.000651
Methanol 1.0 0.000244 0.35786 0.999376 9.54E−13 6.32E−22 1.0 1.0 0.99999 0.000592 0.000781 1.72E−10
Methyl oleate (biodiesel) 1.16E−09 0.08987 0.000187 0.999916 1.05E−08 1.02E−08 0.757006 0.999218 3.21E−09
Oleic acid 0.14 2.36E−09 4.87E−06 5.17E−10 5.43E−05 6.77E−12
Water 0.999755 0.000179 0.000500 4.23E−21 6.72E−21
Diolein 1.49E−27 0.000592 3.10E−10 0.002442
Monoolein 8.01E−12 0.008005 3.10E−10 0.033024
Glycerol 1.20E−07 0.233647 2.33E−08 0.963882

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
12 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Fig. 10. (a) Temperature column profile of the RDC for the esterification first step.
(b) Temperature column profile for the transesterification second step.
Fig. 11. (a) Methyl Oleate reaction rate column profile for the esterification RDC. (b)
Reaction rate column profiles for the transesterification RDC.

flow rate 1 on the biodiesel production. Results from the intensive


simulation show that the combination of the operating variables the RDC. However, the excess of methanol in the reactive zones pro-
(run id: TES-5. methanol feed 1 = 400 kmol/h, reflux ratio = 3 and motes the reduction of temperature and the isothermal behaviour.
boilup ratio = 2.5) is the best RDC operating option due to the reduc- It should be pointed out that while the triglyceride is consumed
tion of the amount of methanol fed (400 kmol/h) and the energy the diglyceride and monoglyceride are produced, thus, sufficient
consumption (lowest temperature, 329.2 ◦ C), maintaining constant methanol to consume all glycerides must be available at the reac-
the biodiesel flow rate at the bottom of the column. Table 10 shows tive zones. Once the reaction products left the reactive zones, the
the final design specifications of the integrated reactive distilla- temperature at the bottom of the RDC is sharply increased because
tion process. Finally, Table 11 shows the detailed stream molar the vaporization of methanol and condensation of biodiesel, glyc-
flow rates, temperature and composition of the integrated reactive erol and non-converted glycerides.
distillation process. Fig. 11a shows the biodiesel reaction rate profile for the esteri-
Fig. 10a shows the temperature profile for the esterification fication RDC. It can be noted in Fig. 11a that the maximum reaction
RDC. It can be noted in Fig. 10a that the maximum temperature is rate is achieved at the reactive stage 5. In fact, it could be said that
located at the oil feed stage (stage 3). This is a reasonable behaviour biodiesel is mainly produced from the reactive stage 3 to stage 10,
of the RDC since the higher boiling temperature compounds are afterwards a constant reaction rate close to zero is reached. This
introduced at stage 3, this is, Triolein and oleic acid are fed at behaviour is consistent with the decrease of oleic acid concentra-
that stage. On the other side, a deep temperature decrement is tion in the reactive zone (see Fig. 12a).
observed at stage 18 where the fresh methanol is fed. Neverthe- Fig. 11b shows the reaction rate profiles for the transesterifica-
less, the presence of triolein and methyl oleate at the bottom of the tion RDC. It can be noted in Fig. 11b that, while the triolein and
RDC promotes an increment of the temperature at the exit of the methanol are quickly consumed at the reactive stages near to the
column. triolein feed stage, the production of biodiesel, glycerol, diolein and
Fig. 10b shows the temperature profile for the transesterifi- monoolein are promoted at stages 2–5. In fact, the production and
cation RDC. It can be seen in Fig. 10b that the RDC behaves like consumption of the products and reactants of the transesterifica-
an isothermal (low temperature) reactive distillation column. This tion reactions end at reactive stage 16, where all the reaction rates
may be surprising if we consider that there are compounds with reach a zero value. This trend of the reaction rates is consistent with
high boiling points (triolein, diolein, monoolein and glycerol) along the constant liquid profile observed in Fig. 12b.

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx 13

This paper does not consider the aspect of controllability of the


process within the research scope. However, it may be mentioned
in advance that, a control key point, is monitoring temperatures at
the top and bottom of the reactive distillation columns, with their
respective control loops, to ensure the purity of the output product.
Therefore, in a second stage of research, it is considered the study
of controllability through the dynamic simulation of the integrated
process.

6. Conclusions

An integrated heterogeneous two-step reactive distillation pro-


cess for biodiesel production has been developed. The reactive
residue curve maps posed in term of elements disclose the basic
configurations for the reactive distillation columns considering the
esterification and transesterification reactions. Intensive simula-
tion for each RD column was performed and a feasible design
satisfying the design and operating constrains for each RD column
was obtained. This is, a feasible heterogeneous reactive distillation
column capable to transform completely the free fatty acid con-
tained in the vegetable oil (more than 1% wt) was obtained. Also,
for the transesterification RD column, it was found that the con-
figuration with two methanol feeds render a thermically stable
RDC (quasi-isothermal) along the reactive zones. This, theoreti-
cally, would increase the residence time of methanol in such zones,
leading to the complete transformation of the triglycerides and
glycerides present in the reacting mixture. However, results also
show that this RDC is more sensitive to changes in the operat-
ing conditions due to the presence of non-converted glycerides
and glycerol. It can be said that, even with the robust thermo-
dynamic modelling of the phase equilibrium (PC-SAFT EOS), in
order to obtain more reliable simulation results, it is necessary to
overcome the uncertaintities about the predicted thermodynamic
properties and the kinetic expressions used in the transesterifi-
cation reactions. Moreover, in order to determine the flexibility,
and thus the process controllability, it is necessary to perform the
dynamic simulation of the integrated process. Finally, it should be
Fig. 12. (a) Liquid composition column profile of the esterification RDC in the inte- mentioned that the above integrated two-steps heterogeneous RD
grated process. (b) Liquid composition column profile of the transesterification RDC process relies on the strong assumption that the solid catalysts
in the integrated process.
(Amberlyst 15 and MgO) have a severe reaction selectivity and
are highly efficient to convert completely the fatty acid and the
triglycerides present in the vegetable oils. Certainly, this could not
Fig. 12a shows the liquid composition profile along the RDC for be claimed in practical and industrial conditions. In fact, it seems
the esterification step. It can be noted that the exit stream leaving that the traces of non-converted glycerides in biodiesel may lead to
the bottom of the column is a mixture mainly containing triolein, filter plugging problems and many research is now being directed
methanol and methyl oleate. The oleic acid is completely consumed to investigate new alternative processes for producing biodiesel,
and water is vaporized. This bottom stream is further separated in for example, the coupled hydrotreating of vegetable oils and
a conventional distillation column where the triglyceride is sepa- petroleum-diesel.
rated from the remaining alcohol and biodiesel. Finally, in a flash
drum, the pure biodiesel is separated from methanol which may be
Acknowledgements
recycled to the methanol feed of the esterification RDC or send it to
the methanol feed of the transesterification RDC. Fig. 12b shows the
The authors acknowledge the financial support provided by
liquid composition profile along the RDC for the transesterification
CONACyT to Xenia Mena-Espino (Catedra:1665).
reaction. It can be noted that the exit stream leaving the bottom of
the column is a mixture containing mainly methyl oleate and glyc-
erol. The glycerol and the non-converted glycerides in the mixture References
are decanted and separated to produce pure biodiesel. It is clear that
Annesini MC, Gironi F, Guerani W. Phase equilibria of highly asymmetric mixtures
the amount of methanol in the reactive stages of the RDC for the
involved in biodiesel production. Chem Eng Trans 2014;38:67–72.
transesterification step must be high in order to achieve complete Chen SS, Kreglewski A. Applications of the augmented van der Waals theory of fluids.
conversion of all glycerides. I. Pure fluids. Ber Bunsen Phys Chem 1977;81:1048–52.
CAPEC. http://www.capec.kt.dtu.dk/documents/software/tutorials/workshop-
So far, it has been developed a feasible process for the production
capec-2011-propred.pdf; 2015.
of biodiesel from vegetable oils containing a significant fraction of Barker JA, Henderson D. Perturbation theory and equation of state for fluids: a
fatty acids. That is, the integrated reactive distillation process is able successful theory of liquids. J Chem Phys 1967;47:2856–61.
to transform a wasted vegetable oil through two heterogeneous Boon-anuwat N, Kiatkittipong W, Aiouache F, Assabumrungrat S. Process design of
continuous biodiesel production by reactive distillation: comparison between
reaction steps. However, to determine the flexibility, and therefore homogeneous and heterogeneous catalysts. Chem Eng Process: Process Intensif
the process controllability, is required dynamic simulation thereof. 2015;92:33–44.

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008
G Model
CACE-5353; No. of Pages 14 ARTICLE IN PRESS
14 E.S. Pérez-Cisneros et al. / Computers and Chemical Engineering xxx (2016) xxx–xxx

Ceriani R, Meirelles AJA. Predicting vapor–liquid equilibria of fatty systems. Fluid Prasertsit K, Mueanmas C, Tongurai C. Transesterification of palm oil with
Phase Equilib 2003;215:227–36. methanol in a reactive distillation column. Chem Eng Process: Process Intensif
Díaz-Tovar C-A, Gani R, Sarup B. Lipid technology: property prediction and process 2013;70:204–15.
design/analysis in the edible oil and biodiesel industries. Fluid Phase Equilib Sales-Cruz M. Ph.D. Thesis, CAPEC Development of a computer aided modelling
2011;302:284–93. system for bio and chemical process and product design Ph.D. Thesis, CAPEC.
Dossin TF, Reyniers M-F, Berger RJ, Marin GB. Simulation of heterogeneously Department of Chemical Engineering. Technical University of Denmark, March
MgO-catalyzed transesterification for fine-chemical and biodiesel industrial 3; 2006.
production. Appl Catal B: Environ 2006;67:136–48. Sales-Cruz M, Aca-Aca G, Sánchez-Daza O, Loṕez-Arenas T. Predicting critical prop-
Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterification of oils. erties, density and viscosity of fatty acids, triacylglycerols and methyl esters
J Biosci Bioeng 2001;92:405–16. by group contribution methods. In: Pierucci S, Buzzi Ferraris G, editors. 20th
Gomez-Castro FI, Rico-Ramirez V, Segovia-Hernandez JG, Hernandez-Castro S, El- European Symposium on Computer Aided Process Engineering – ESCAPE20.
Halwagi MM. Simulation study on biodiesel production by reactive distillation Amsterdam: Elsevier B.V.; 2010, Poster 101.
with methanol at high pressure and temperature: impact on costs and pollutant Tesser R, Casale L, Verde D, Di Serio M, Santacesaria E. Kinetics and modeling
emissions. Comput Chem Eng 2013;52:204–15. of fatty acids esterification on acid exchange resins. Chem Eng J 2010;157:
Granados-Aguilar AS, Viveros García T, Pérez-Cisneros ES. Thermodynamic analysis 539–50.
of a reactive distillation process for deep hydrodesulfurization of diesel: effect Van Gerpen JH. Biodiesel economics. Montana economics, biological and agricultural
of the solvent and operating conditions. Chem Eng J 2008;143:210–9. engineering. Moscow, ID, USA: University of Idaho; 2007, www.deq.state.mt.us.
Gross J, Sadowsky G. Perturbed-chain SAFT: an equation of state based on a pertur- Wertheim MS. Fluids with highly directional attractive forces. II. Thermodynamic
bation theory for chain molecules. Ind Eng Chem Res 2001;40:1244–60. perturbation theory and integral equations. J Stat Phys 1984;35:35–47.
Kiss AA, Bildea CS. A review of biodiesel production by integrated reactive separation Yunus khan TM, Atabani AE, Irfan Anjum Badruddin, Ahmad Badarudin MS, Khayoon
technologies. J Chem Technol Biotechnol 2012;87:861–79. S, Triwahyono. Recent scenario and technologies to utilize non-edible oils for
Marrero J, Gani R. Group-contribution based estimation of pure component proper- biodiesel production. Renew Sustain Energy Rev 2014;37:840–51.
ties. Fluid Phase Equilib 2001;183–184:183–208. Zéberg-Mikkelsen CK, Stenby EH. Predicting the melting points and the enthalpies
Pérez-Cisneros ES, Gani R, Michelsen ML. Reactive separation systems I. Computa- of fusion of saturated triglycerides by a group contribution method. Fluid Phase
tion of physical and chemical equilibrium. Chem Eng Sci 1997;54:527–43. Equilib 1999;162:7–17.

Please cite this article in press as: Pérez-Cisneros ES, et al. An integrated reactive distillation process for biodiesel production. Computers
and Chemical Engineering (2016), http://dx.doi.org/10.1016/j.compchemeng.2016.01.008

Вам также может понравиться