You are on page 1of 26

Accepted Manuscript

Title: CFD modeling of multiphase flow in reactive distillation


column

Authors: Maryam Mazarei Sotoodeh, Mortaza Zivdar, Rahbar


Rahimi

PII: S0255-2701(18)30165-X
DOI: https://doi.org/10.1016/j.cep.2018.04.034
Reference: CEP 7276

To appear in: Chemical Engineering and Processing

Received date: 11-2-2018


Revised date: 23-4-2018
Accepted date: 29-4-2018

Please cite this article as: Sotoodeh MM, Zivdar M, Rahimi R, CFD modeling of
multiphase flow in reactive distillation column, Chemical Engineering and Processing
- Process Intensification (2010), https://doi.org/10.1016/j.cep.2018.04.034

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
CFD modeling of multiphase flow in reactive
distillation column
Maryam Mazarei Sotoodeha, Mortaza Zivdara,1,Rahbar Rahimia

a
Department of Chemical Engineering, University of Sistan and Baluchestan, Zahedan, P.O.
Box 98164-161, Iran

T
IP
1
Mortaza Zivdar, Tel: +98 915 341 4268. E-mail: mzivdar@eng.usb.ac.ir

R
Graphical Abstract

SC
U
N
A
M
D
TE
EP
CC

Highlights:

 Mass transfer and chemical reaction in reactive distillation column were studied.
A

 Katapak SP-11&12 were used to investigate the effect of internals.


 Penetration theory of Higbie was applied to the mass transfer coefficient of liquid.

1
Abstract
This work presents a 3D VOF two-phase model to investigate the hydrodynamics, chemical reaction heat, and
multicomponent mass transfer of 1, 1 diethoxy butane (DEB) production in a reactive distillation column. Two
types of internal, katapak SP-11 and 12, were studied. Penetration theory of Higbie was applied to the mass
transfer coefficient of liquid. Mass fraction of each component in the gas phase at the outlet of stages was shown
and compared to the experimental data. Results showed that the butanal of the liquid phase enters to the gas
phase with the rising of gas along the column, which some amounts of ethanol, water, and DEB in the gas phase
is absorbed in the liquid. Moreover, the mass transfer in open channels is more than closed channels. The
comparison shows that they are in satisfactory agreement, with relative errors lower than 10%. Furthermore,
according to CFD results, the conversions are decreased due to the generation of products. As the reaction is
exothermic, the rate of rising in temperature decreases as the conversion rate decreases. The mean relative errors
for prediction of mole fraction and temperature were 10% and 3%, respectively.

T
Keywords: Reactive distillation, Multicomponent mass transfer, Reaction heat, katapak SP,
Computational fluid dynamics.

IP
1. Nomenclature

R
A pre-exponential factor

SC
a
mass transfer area, m2

a Permeability, m2

C Concentration, mol/m3 U
N
C2 Resistance coefficient, m-1
A
D Diffusion coefficient, m2/s
M

dh Hydraulic diameter, m
D

dp Diameter of catalyst particle, mm


TE

Ea Activation energy, J/mol

Fs Gas capacity factor, m/s(kg/m3)0.5


EP


F Additional forces in the Navier–Stokes equations, N
CC

g Gravitational acceleration, m/s2

j Diffusion flux, mol/m2s


A

k Thermal conductivity, W/m K

K Mass transfer coefficient, m/s

P Pressure, Pa

R Universal gas constant, J/mol.K

2
r Reaction rate, mol/m3.s

T Temperature, K

te Exposure time, s

tR Thickness catalyst filled pockets, mm

v Velocity, m/s

T
w Catalyst loading, Kg cat/m3

IP
x Mole fraction of each component in liquid phase

R
Greek letters

SC
∆Hr Enthalpy of reaction, kJ/mol

α Overall corrugation angle


U
N
αe Corrugation angle
A
Γ Cross sectional fraction
M

ε Porosity, dimensionless

εp Void fraction
D

Λ Volume fraction catalyst


TE

μ Dynamic viscosity (Pa.s)


EP

ρ Density (kg/m3)

σ Surface tension (N/m)


CC

X Channel ratio

δ Film thickness, m
A

ω Mass fraction

Subscripts

CB Catalyst bag

3
D Distillation section

i Component

j Stage in the reactive distillation column

G Gas phase

hG hydraulic Gas

T
L Liquid phase

IP
e effective

R
q Phase

SC
2. Introduction
Reactive distillation processes are used in several areas of chemical engineering. Due to

U
occurring chemical reaction and distillation in one unit, the capital and operational cost of
reactive distillation systems are reduced compared with conventional reactor-distillation
N
column [1, 2]. These processes also offer many other advantages over traditional methods
such as simplification or elimination of the separation system (capital savings), improving
A
selectivity and reducing by-product formation, catalyst requirement reduction significantly,
avoidance of azeotropes, heat integration benefits, and preventing hot spot formation.
M

Therefore, in many cases, the uses of these systems are preferred. Despite all the benefits of
reactive separation, there are some limitations, including volatility constraints, the
requirement of residence time and catalysts lifetime, scale up to large flows, process
D

conditions mismatch, and the liquid phase reaction [1, 3-5].


Recently, production of biodiesel additives increased significantly. 1,1 diethoxy butane is one
TE

of the acetals that can be used as oxygenated diesel additives. Acetals can be produced from
an acid-catalysed reaction between alcohols and aldehydes [6-11]. Due to high conversion,
low energy consumption and reduction of capital costs, the reactive distillation process is one
EP

of the most preferred technologies in the production of acetals. Also, this process is applied
for the limited equilibrium reactions such as reversible liquid phase reactions. Because of
removing the products continuously, selectivity and conversion are increased [12-16].
CC

Numerous experimental works were accomplished on the various types of reactive separation
processes. Some of these works were focused on the synthesis of materials (e.g., acetal, ) by
reactive separation processes [17-23]. The main scope of others was on the hydrodynamics
behavior and effect of internals on the process [24-27]. According to the cost of
A

hydrodynamic experiments and various internal equipment (including various packing and
catalysts), the use of CFD simulation has recently become the most useful way to study on
various types of packing and synthesis in reactive separation processes [28-30].
Van Baten et al. [31] studied the hydrodynamics of a reactive distillation sieve tray column.
Catalyst containing wire-gauze envelopes were disposed along the liquid flow direction. A 3D
two-phase model was developed to determine liquid clear height on the tray as a function of
tray geometry and operating conditions. A 3D steady state one phase model was presented by

4
Kloker et al. [2] to obtain the influence of different catalytic internals on the reactive
distillation of n-hexyl acetate from hexanol and acetic acid.
Egorov et al. [32] have proposed a new modelling methodology for reactive separation which
exploits a combination of modern CFD facilities and the rate-based process simulation
approach. Hydrodynamics and mass transfer correlations were obtained by using CFD
simulations. Zivdar et al. [33] investigated the dry pressure drop within the catalyst packed
channels of katapak-S by using CFD simulations. Also, they presented the gas flow path line
in the packing sheet and elbows.
Physical and reactive mass transfer absorption in gas-liquid flow was simulated by Haroun et
al. [34]. Haroun et al. [35] simulated reactive mass transfer at the liquid interface in a two-
phase flow. 2-D simulations were performed to investigate interfacial Mass transfer and liquid

T
hold up by Haroun et al. [36]. In all three papers, they were using VOF model to capture the
gas-liquid interface motion in structured packing. Also, they found that Higbie theory well

IP
predicts the liquid side mass transfer.
A 3D two-phase CFD model was established to study the separation performance of

R
structured catalytic packing by Dai et al. [37]. Two types of structured catalytic packings (i.e.
BH-1 and BH-2 types) were used in simulations. Liu et al. [38] analysed the multi-scale

SC
structure of a reactive distillation column by Aspen Plus with Fluent software. The reactive
distillation column is divided into four scales: column scale, tray scale, fluid mechanical scale
and molecule scale. Tray efficiency was calculated using CFD simulations. Ding et al. [39]
presented a 3D model for simulating the winpak-based modular catalytic structured packing
U
by using CFD. Mazarei et al. [40] investigated the pressure drop and flow pattern in the
modular catalytic structured packings. Also, they illustrated the effect of geometry on the
N
hydrodynamics and characterisation of flow in the katapak SP modules.
It appears from above review that there is lack of knowledge in reactive distillation processes.
A
As experimental setups are so expensive and measuring parameters locally (i.e., mass
fraction, temperature, pressure, velocity, etc.) are not possible, CFD simulation is the best
M

approach to obtained accurate data in whole RD systems. Furthermore, it is simplified to


evaluate the various internals effects. At previous work, a 3D VOF model that was developed
to evaluate dry and wet pressure drop of katapak SP-11 & 12 [40] is validated by the Behrens
D

[41] experimental data. Also, the catalyst bags of katapak SP were simulated as porous media
to show the liquid velocity distribution in these bags. The goal of this work is to model the
TE

reaction, heat and multicomponent mass transfer that occurred in packings of RD columns.
For this purpose, at first RD column base on Agirre [9] experimental work was simulated by
commercial software. Secondly, a 3D two-phase VOF model was carried out. Finally, the
EP

results of multicomponent mass transfer in the separation section and the reaction and heat
that appeared in the catalyst bags were compared with commercial software results. Fig. 1
represents the flowchart, describing the methodology of this work.
CC

3. Experimental description
Agirre [9] studied on the production of 1,1 diethoxy butane in a packed bed column that
used katapak SP-11 in the reactive section and Amberlyst resins as a catalyst.
A

DEB is obtained from ethanol and butanal in an exothermic reversible reaction. The
stoichiometry of the reaction between them is as follow [7, 9]:

2 Ethanol+1 Butanal↔1 DEB+ 1 Water

Moreover, they reported kinetic parameters of the reaction. The experiments were
performed in a batch stirred tank reactor with Amberlyst resins as a catalyst. Due to
experiments results, the reaction follows an elemental rate. Table 1 shows the Arrhenius

5
parameters. The activation energy and the pre-exponential factor obtained in the kinetic study
were implemented in the model.

Reactive distillation experiments were carried out in a continuous packed catalytic distillation
column. The inner diameter and the height of column are 0.05 and 0.75 m, respectively. The
column included three sections: rectification, reactive and stripping section with a total
condenser and reboiler. The height of sections was 0.15 m, 0.3 m, and 0.3 m, respectively.
The reactive part is filled with three katapak SP-11 modules. Multiknit structured packing is
used in the rectification and stripping sections.
The ethanol and butanal, as reactants, were fed to the column in the stoichiometric ratio at
the upper and the lower part of the catalytic section, respectively. The column was thermally

T
isolated and operated at atmospheric pressure. Ethanol and butanal have entered the column at
the temperature of 343 K and 338 K, respectively.

IP
4. Model development method

R
4.1. Simulations and assumptions
Several assumptions have been used in the simulations due to complicated geometry, lack

SC
of appropriate understanding of multiscale phenomena in katapak SP, and time-consuming
simulations. Although, the hydraulic parameters such as dry and wet pressure drop have
shown a periodic or "quasi-steady state" manner, their changes were considered negligible

U
during simulation time. These parameters have some effect on the gas-liquid contact time and
mass transfer rate between two-phases. As regards that these changes are negligible; the
N
steady state assumption is acceptable. Also, due to complicated geometry of catalyst particles
in the reactive section, the catalyst bags were assumed as a porous medium [32].
A
In the first part, hydrodynamic simulations, the continuity and momentum equations solved
numerically for each phase. For single-phase flow, the air is used as the gas phase while in the
M

two-phase analysis, air and water assumed as the gas and liquid phases, respectively.
Simulations were done at atmospheric pressure and ambient temperature. Therefore, one can
neglect the use of energy balance equations. As the liquid and gas are assumed to be
D

Newtonian, isothermal, and incompressible, their physical properties are kept constant.
Moreover, mass transfer and reactions, as well as capillary rise, were neglected in this part of
TE

the simulations.
In the second part, based on the Agirre [9] experiments, gas and liquid phases are the
mixtures of four components. Both phases are consist of ethanol and butanal as reactants and
EP

DEB and water as products of the reaction. As noted in the references [42-44], in two-phase
systems, we can consider that all resistance to mass transfer concentrated in one phase.
Therefore, in the present work, it is assumed that the dominant resistance to the mass transfer
is in the liquid phase.
CC

As like as previous work [40], Simulations were done at the liquid load point for both
geometries. All the walls were assumed impermeable, so liquid cannot move from catalyst
bags to corrugated sheets and vice versa.
A

4.2. Governing equations


4.2.1. Hydrodynamic model
The VOF model is well suited for tracking the interface between the two immiscible
phases, which is validated by a variety of researchers for gas-liquid CFD simulations of
structured packing [34, 35, 45-49]. The hydrodynamic model, that was presented in our
previous work [40], is used in this work. So, a summary of model is presented in this section.
A detailed description of the model can be found in Mazarei et al. [40].

6
The VOF model accounts for the effect of surface tension along the interface between the
phases. The continuum surface force (CSF) model was applied as surface tension model [34-
36, 45-47]. In the CSF model that was proposed by Brackbill et al. [50], a surface force is
formulated to model numerically surface tension effects at fluid interfaces having finite
thickness [50]. The BSL model was used to solve the turbulent viscosity in the simulations
[46, 47]. The diameter of catalyst particles is 1mm, and  CB , the void fraction of catalyst bags
is 0.37.

4.2.2. Mass transfer equations


To predict the multicomponent mass transfer in a system with C component, C-1 mass
transfer equations must be solved in each phase. The mass transfer equation for phase q is

T
expressed as:

IP
 
 
.  q q v i, q  . q ji ,q   q Si ,q (1)

R
where i , q is the mass fraction of component i in phase q, Si , q is the source of species i in

SC
phase q due to interfacial mass transfer and ji ,q is the diffusion flux of component i in phase
q. Mass transfer between phases for each component must satisfy the balance condition, so
Si , L   Si ,G The source term for liquid phase can be described by the following equation:
.
U
N
Si  Ki ae  ( xi  x*i ) (2)
A
M

where K i is the mass transfer coefficient, ae referred to the effective mass transfer area, xi is
the mole fraction of component i, and x *i is the equilibrium mole fraction of component i in
D

the liquid phase. In the above equation, Onda relation is used as the effective mass transfer
area [41, 51].
TE

As mentioned in section 3.1., mass transfer resistance in the gas phase is neglected.
Therefore, the mass transfer coefficient is equal to liquid mas transfer coefficient. It is found
that penetration theory of Higbie well predicts the liquid side mass transfer [34-36, 41, 45,
EP

49]. The average mass transfer coefficient is K L  2 DL te [43]. DL is the liquid side
diffusion coefficient and te is the exposure time. Exposure time is defined with the
characteristic length and effective liquid velocity. Behrens [41] investigated about the
CC

hydrodynamics and mass transfer in catalytic structured packings and presented a model
based on the katapak SP geometry. In this model, the mass transfer coefficient in the laminar
film is as follow:
DL Le
A

KL  2 (3)
0.9d hG

The diffusivity coefficient of each component in the quadratic liquid is computed by Cullinan
equation, in the simulations. This equation is an extension of Vignes’ equation to
multicomponent systems [52].
In Eq. 8, the gas hydraulic diameter is defined as [41]:

7
bh  2s 2
d hG  bh
0.5 (4)
 bh  2s   bh  2s   2 
2
bh  2s
     
 2h   b   2h

where b is the corrugation base width, h is the corrugation height, s is the corrugation side
length and δ is the liquid film thickness in the channels. The film thickness is defined as [41,
53]:

T
3 L L ,D
 3 (5)

IP
a p  L g sin( L )

where μL is the dynamic viscosity, vL,D the liquid velocity in the distillation section that takes

R
the excess liquid above the catalytic load point into account and is based on the volume flow
liquid in this part of the packing, ap the geometric installed area, ρL the liquid density, g the

SC
gravitational acceleration and aL the effective liquid flow angle. aL is the flow angle along
which the liquid descends through the channels and is defined as [41] :



cos90  ac 


U
N
  (6)
a L  a tan
  b  
 sin 90  ac  cos a tan   
A
   2h   
M

where ac is the corrugation inclination angle to the horizontal, i.e. 41º.


The effective liquid velocity in Eq. 3 is defined as [41]:
D

 L,D
 Le  (7)
 D hL sin( a L )
TE

where ɛD is the porosity of the distillation section in the katapak SP which is equal to that of
the mellapak plus and hL is the dynamic liquid hold-up in the channels.
EP

4.2.3. Energy equation


Energy equation is solved in the following form:
CC

 (8)
.( hv )  .(kT )  S h
A

where h is the mass averaged enthalpy and T is the shared temperature field. k is the volume
averaged thermal conductivity. Sh includes the heat of chemical reaction and any other
volumetric heat sources. The source term is obtained by the following equation:

S h  r * H r (9)

8
where r and H r are rate and enthalpy of reaction, respectively. These parameters are
illustrated in the next section.

4.2.4. Reaction equations


The global reaction can be symbolized as:
k
2A  B  C  D
k4

A reaction mechanism was proposed by Agirre [9]

T
d [C ] (10)
 wkC A2 C B  wk 4CC CD

IP
dt

R
where w is catalyst loading (Kg cat/m3), Cj is the concentration of “j” (mol/m3), k=k3K and

SC
C r r r r
K  HA . The relationship among all the component rates is A  B  C  D . Based on
C AC B a b c d
this equation, rate laws for all the components are:

rA  2 wkC A2 C B  2wk 4 CC C D U (11)


N
A
rB   wkC A2 C B  wk 4 CC C D (12)
M

rC  wkC A2 C B  wk 4 C C C D (13)
D

rD  wkC A2 C B  wk 4 C C C D (14)
TE

Also, kinetic constants follow Arrhenius’ correlation:


EP

  Ea  (15)
Ln(k )  Ln( A)   
 RT 
CC

where Ea is the activation energy (J/mol), R is the universal gas constant (J/(mol K)), T is the
temperature in Kelvin, A is the pre-exponential factor and k is the kinetic constant.
A

Agirre [9] presented a temperature dependent equation for enthalpy of reaction, so the
enthalpy changes locally in the computational domains:

H r (T j )  31098  (73.069)  (T j  298.15) (16)

4.3. Computational domains and boundary conditions


9
In this study, the results of simulations were validated by Behrens [41] and Agirre [9]
experimental works. Both of them have been used katapak SP in the reactive distillation
columns.
Various types of the internals are used in reactive distillation columns. The internals can
affect the transfer phenomena which occur in the distillation columns [41, 54, 55]. Katapak
SP is a type of modular catalytic structured packings which is used in reactive separation
columns. This kind of packing is composed of two distinct parts: separation and reactive
sections. The katapak SP module is made of alternating vertical layers of structured packing
sheets and catalyst bags. The catalyst bags are made of wire gauze envelopes containing
Amberlyst catalyst particles. Mellapak Plus is used as a separation layer. The ratio of
separation to reactive layers can be varied to declare flexibility between separation and

T
reaction processes. The suffix 11 and 12 denote an alternating arrangement as one corrugated

IP
sheet with one catalyst bag and two corrugated sheets with one catalyst bag, respectively[41,
56].

R
Computational domains contained a piece of katapak SP that consists of three sections: one
separation section in the middle and catalyst bag on each side as a reactive section. Fig. 2

SC
shows the schematic of katapak SP-11 and 12. The contact of gas and liquid happen in
structured packing sheet, and the separation process is done in this layer. Only the liquid
phase flows through the reactive section while the reaction is carried out in contact with the
catalyst particles.

U
In katapak SP-11, one sheet of Mellapak Plus is drawing as a separation layer. In this
situation, all of the channels of corrugated sheet are in the vicinity of catalyst bags. These
N
channels called closed channels. In katapak SP-12, two sheets of Mellapak Plus are drawing
between catalyst bags as a separation layer. In this case, two types of channels are created,
A
closed channels and open channels. Open channels are the crisscrossing channels in the
middle of two sheets. The vapour is not able to flow through the catalyst bags, because of
M

their compact structure, so the cross-sectional area available for the vapour is reduced.
There are three useful ratios in the structure of katapak SP which known as channel ratio,
cross-sectional ratio and volume fraction of the catalyst. These parameters differentiate both
D

katapak SP-11 and katapak SP-12 and affect the pressure drop of modules and consequently
the mass transfer and reaction performance of packing. Some characteristics of geometries
TE

and ratios are listed in Table 2.

Grid independency study was done for computational domains. Several sizes of mesh (i.e.,
EP

0.2, 0.3, 0.35, 0.4 and 4.5 mm) were tested to make sure of the numerical accuracy of
simulations. Eventually, the size of 0.35mm and 0.3 mm were chosen for separation and
reactive sections based on the simulation results. The computational time required for each
CC

two-phase simulation is more than one week on a core i7 CPU running on an eight-core 2.67
GHz with 18.0 GB of RAM.
A

On the bottom of domains, static pressure was selected for catalyst bags and structured
packing sheets as the boundary condition. At the top, liquid inlet velocity and gas outlet
velocity were specified for both domains. Also, in the simulations of mass transfer and
reaction, the mass fraction of each component is defined. No slip wall boundary condition
was selected for the liquid flow while free slip wall boundary condition was used for the gas
phase.

5. Results and discussion


10
5.1. Hydrodynamics
The hydrodynamic behavior of flow affects the mass transfer and reaction that occurred in
the column. In our previous study [40], a 3D steady state VOF simulations were done for
investigating dry gas and two-phase flow in katapak SP-11 and 12. To making sure of
hydrodynamics model that is used in the simulations, the results were compared with the
experimental results of Behrens [41]. Then, this model is used in the simulation of DEB
production. In this section, a summary of hydrodynamic results is reviewed. The complete
illustration of hydrodynamic simulations and results are presented in Mazarei et al. [40].
5.1.1. Dry pressure drop
The dry pressure drop of one phase flow for both geometries is presented and compared

T
with experimental results. Fig. 3 compares simulation and experimental dry pressure drop for
all Fs in katapak SP-11 and 12. The solid line in the figure denotes that simulated pressure

IP
drop is equal to experimental one and dotted lines over and under the solid line represents
15% deviations. It can be seen that the trends of simulated pressure drop vs. Fs exhibit the

R
same behaviour with experimental data. In the other hands, by increasing the Fs, the pressure
drop is increased and vice versa. Although, the simulated pressure drops are under and

SC
overpredict in all Fs, for katapak SP-11 and 12, respectively.
Since the cross-sectional fraction available for gas flow in the katapak SP-12 is larger than
katapak SP-11 (i.e. Г11: Г12 = 0.4: 0.52), the pressure drop of this packing is lower than the

U
katapak SP-11. The same trend was observed by Behrens [41].
N
5.1.2. Two-phase pressure drop
The wet pressure drop within a specific range of Fs (≈ 0.55 – 1.2 (m/s (kg/m3)0.5) is
A
obtained. All the simulations were done at the liquid loading point. As expected, the pressure
drop increases with increasing Fs. The discrepancy between the simulation results and
M

experimental data for Katapak SP-11 and 12 are shown in a parity plot in Fig.4. As explained
in section 4.1.1, the solid line in the figure denotes the simulation pressure drop is equal to
experimental data and dotted lines over and under the solid line represents 9% deviations. The
D

simulation results were found in good agreement with the experimental data showing about
7% and 11% deviation for katapak SP-11 and 12, respectively.
TE

Separation section in katapak SP-11 only includes closed channels. So, the pressure and
velocity profiles are the same in all channels. However, katapak SP-12 contains both open and
closed channels. Therefore, this difference led to different pressure drop profile in the
EP

channels of packing layer, which is shown in Fig. 5. It is clear from that, the pressure drop in
the closed channels is approximately 70% higher than open channels; while the pressure drop
of separation section in katapak SP-12 is between these two types of channel. As mentioned
CC

in section 3.3, the differences in the pressure drop profiles occur because of the difference in
the channels nature.
Since the boundaries around the catalyst bag have been defined as impermeable walls,
A

liquid velocity near the wall boundaries is zero while it is uniform inside the bag. As can be
seen from Fig. 6, the last result is reached by assuming the catalyst bag as homogeneous and
isotropic porous medium.

5.2. Multicomponent mass transfer


Agirre [9] presented the mass fraction of each component, mass flow of each phase, and
temperature only at the distillate and bottom of the column. Since the data between stages are
needed to validate the CFD results, the column was simulated by commercial software. Based

11
on the Agirre [9], the NRTL model was chosen to predict the thermodynamic equilibrium of
the mixture. This equation is implemented in the simulations. The simulated column is as like
as experimental column that illustrated in section 3 and consisting of three reactive stages.
Consequently, mass fraction, temperature, and liquid and gas mass flows of each stage were
obtained to validate the CFD results.
Mass fraction of each component at the inlet and outlet of each stage in the separation
section were shown in Fig. 7. According to experimental data, with the ascent of gas along the
column, the butanal of the liquid phase enters the gas phase. In return, little amount of the
ethanol, water and, DEB in the gas phase are absorbed in the liquid. These trends observed
clearly in CFD results, too. Fig. 8 a, b and c show the mass fraction of components at the first,
second and third reactive stage, respectively. At the lower reactive stage, the third stage, the

T
amount of butanal is at its lowest level in the gas phase. Whatever, the gas moves upward in
the column, the butanal in the gas phase increases. This process is reversed for ethanol. It can

IP
be observed that the results obtained from CFD simulations have good agreement with
experimental data and relative errors are lower than 10%. Matching the results of simulations

R
and experimental data can confirm the assumptions and closure models used.
As explained in Section 4.1.2, two-phase hydrodynamics in katapak SP-12, the pressure

SC
drop in open and closed channels are not equal. Also, according to Behrens [41], the HETP of
the katapak SP-12 is less than the HETP of the katapak SP-11. The differences in pressure
drop and HETP are due to the nature of the open and closed channels and cross sectional
fraction in each module. Fig. 7. d shows the difference in the amount of mass transfer in open
U
and closed channels for ethanol and butanal. As can be seen, the mass transfer in open
channels is more than closed ones.
N
A
5.3. Chemical reaction
Fig. 8 represents the mole fraction of each component at the inlet and outlet of reactive
M

stages. Considering the mole fraction of reactants in all stage are higher than products,
reaction moves in the forward direction. Simultaneously the mole fraction of reactants
decreased, the mole fraction of products increased. The results of CFD simulations were
D

found in acceptable deviations with commercial software data showing about 10% deviations.
Fig. 8 d shows the conversion of ethanol and butanal. As it is clear, the conversion of
TE

reaction decreased during the generation of products. So, the conversion of the first stage is
higher than the second and third stages.
Fig. 9 represents the temperature of the liquid at the inlet and outlet of each stage.
EP

Comparison between CFD simulation results and commercial software data shows good
agreement about 3% deviations. As shown in Fig. 8 d and Fig. 9, since the reaction is
exothermic, the rate of rising in temperature decreases as the conversion rate decreases.
CC

Given the compatibility of the results of simulations with the experimental results, it is
clear that the assumptions considered, including considering catalytic bags as a porous
medium, are acceptable.
A

Conclusions
In the present study, a 3-D two-phase model for the study of hydrodynamic, chemical reaction
heat and multicomponent mass transfer of 1,1 diethoxy butane production in the reactive
distillation process was investigated. In the reactive distillation column, katapak SP-11 and 12
were used to understand more about the reactive separation internals.

12
Penetration theory was used to prediction the liquid mass transfer coefficient in the
multicomponent mixture. Mass fraction of each component at the inlet and outlet of each
stage in the separation section were compared to the experimental data and commercial
software results. According to experimental data, with the ascent of gas along the column, the
butanal of the liquid phase enters the gas phase. This process is reversed for ethanol, water,
and DEB in the gas phase. It can be observed that CFD results show good agreement with
experimental data and relative errors are lower than 10%. Matching the results of simulations
and experimental data can confirm the assumptions and closure models used. Due to
differences in the nature of the open and closed channels and cross sectional fraction of
modules, the mas transfer of the katapak SP-12 is more than that in the katapak SP-11.
The mole fraction of each component and temperature at the inlet and outlet of catalyst

T
bags of all reactive stages were illustrated. As the results show, the conversion is decreased
due to the generation of products and increases the mole fraction of DEB and water in the

IP
liquid. Also, considering that the reaction is exothermic, the rate of rising in temperature
decreases as the conversion rate decreases. The results show that the mean relative errors for

R
prediction of mole fraction and temperature were 10% and 3%, respectively.
Finally, it can be concluded that despite all the limitations of the CFD, the use of this

SC
knowledge can help to understand better the phenomena occurring in the process of reactive
distillation.

U
N
A
M
D
TE
EP
CC
A

13
References
[1] R. Taylor, R. Krishna, modeling reactive distillation, Chem. Eng. Sci. 55 (2000) 5183-
5229. https://doi.org/10.1016/S0009-2509(00)00120-2
[2] M. Klöker, E.Y. Kenig, A. Górak, On the development of new column internals for
reactive separations via integration of CFD and process simulation, Catal. Today 79-80
(2003) 479–485. https://doi.org/10.1016/S0920-5861(03)00068-3
[3] A. Tuchlenski, A. Beckmann, D. Reusch, R. Dussel, U. Weidlich, R. Janowsky, Reactive
distillation - industrial applications, process design & scale-up, Chem. Eng. Sci. 56 (2001)
387-394. https://doi.org/10.1016/S0009-2509(00)00240-2
[4] C. Noeres, E.Y. Kenig, A. Gorak, Modelling of reactive separation process: reactive
absorption and reactive distillation, Chem. Eng. Process. 42 (2003) 157-178.

T
https://doi.org/10.1016/S0255-2701(02)00086-7

IP
[5] B. YANG, J. WU, G. ZHAO, H. WANG, S. LU, Multiplicity Analysis in Reactive
Distillation Column Using ASPEN PLUS, Chin. J. Chem. Eng. 14 (2006) 301—308.
https://doi.org/10.1016/S1004-9541(06)60075-X

R
[6] V.M.T.M. Silva, A.E. Rodrigues, Synthesis of diethylacetal: thermodynamic and kinetic
studies, Chem. Eng. Sci. 56 (2001) 1255-1263. https://doi.org/10.1016/S0009-

SC
2509(00)00347-X
[7] M. Rahaman, N.S. Graca, C.S.M. Pereira, A.E. Rodrigues, Thermodynamic and kinetic
studies for synthesis of the acetal (1,1-diethoxybutane) catalyzed by Amberlyst 47 ion-

U
exchange resin, Chem Eng J. 264 (2015) 258-267. https://doi.org/10.1016/j.cej.2014.11.077
[8] I. Agirre, V.L. Barrio, B. Gueme, J.F. Cambra, P.L. Arias, Catalytic reactive distillation
N
process development for 1,1 diethoxybutane production from renewable sources, Bioresour.
Technol. 102 (2011) 1289-1297. https://doi.org/10.1016/j.biortech.2010.08.064
A
[9] I. Agirre, Innovative reaction systems for acetal (1,1 diethoxy butane) production from
renewable sources, Chemical and Environmental Engineering Department, Engineering
M

Faculty of Bilbao, Bilbao, Spain, 2010,


[10] I. Agirre, M.B. Guemez, A. Motelica, H.M. Veen, J.F. Venteb, L.A. Pedro, A techno-
economic comparison of various process options for the production of 1,1-diethoxy butane, J.
D

Chem. Technol. Biotechnol. 87 (2012) 943–954. https://doi.org/10.1002/jctb.3704


[11] M.R. Capeletti, L. Balzano, G. Puente, M. Laborde, U. Sedran, Synthesis of acetal (1,1-
TE

diethoxyethane) from ethanol and acetaldehyde over acidic catalysts, Appl. Catal., A: General
198 (2000) L1–L4. https://doi.org/10.1016/S0926-860X(99)00502-5
[12] B. Bessling, J. Löning, A. Ohligschläger, G. Schembecker, K. Sundmacher,
EP

Investigations on the Synthesis of Methyl Acetate in a Heterogeneous Reactive Distillation


Process, Chem. Eng. Technol. 21 (1998) 393-400. https://doi.org/10.1002/(SICI)1521-
4125(199805)21:5<393::AID-CEAT393>3.0.CO;2-9
[13] N. Calvar, B. Gonzalez, A. Dominguez, Esterification of acetic acid with ethanol:
CC

Reaction kinetics and operation in a packed bed reactive distillation column, Chem. Eng.
Process. 46 (2007) 1317–1323. https://doi.org/10.1016/j.cep.2006.10.007
[14] F. Forner, M. Brehelin, D. Rouzineau, M. Meyer, J.U. Repke, Startup of a reactive
A

distillation process with a decanter, Chem. Eng. Process. 47 (2008) 1976–1985.


https://doi.org/10.1016/j.cep.2007.09.005
[15] I.K. Lai, Y.C. Liu, C. Yu, Production of high-purity ethyl acetate using reactive
distillation: Experimental and start-up procedure, Chem. Eng. Process. 47 (2008) 1831–1843.
https://doi.org/10.1016/j.cep.2007.10.008
[16] K. Prasertsit, C. Mueanmas, C. Tongurai, Transesterification of palm oil with methanol
in a reactive distillation column, Chem. Eng. Process. 70 (2013) 21-26.
https://doi.org/10.1016/j.cep.2013.05.011

14
[17] J.C. Gonzalez, H. Subawalla, J.R. Fair, Preparation of tert-Amyl Alcohol in a Reactive
Distillation Column.2. Experimental Demonstration and Simulation of Column
Characteristics, Ind. Eng. Chem. Res 36 (1997) 3845-3853. https://doi.org/10.1021/ie960808l
[18] M. Schmitt, H. Hasse, K. Althaus, H. Schoenmakers, L. Götze, P. Moritz, Synthesis of n-
hexyl acetate by reactive distillation, Chem. Eng. Process. 43 (2004) 397–409.
https://doi.org/10.1016/S0255-2701(03)00124-7
[19] S. Bhatia, A.L. Ahmad, A.R. Mohamed, S.Y. Chin, Production of isopropyl palmitate in
a catalytic distillation column: Experimental studies, Chem. Eng. Sci. 61 (2006) 7436–7447.
https://doi.org/10.1016/j.ces.2006.08.039
[20] M. Brehelin, F. Forner, D. Rouzineau, J.U. Repke, X. Meyer, M. Meyer, G. Wozny,
Production of n-propyl acetate by reactive distillation Experimental and Theoretical Study,

T
Chem. Eng. Res. Des. 85 (2007) 109–117. https://doi.org/10.1205/cherd06112
[21] J. Holtbruegge, S. Heile, P. Lutze, A. Górak, Synthesis of dimethyl carbonate and

IP
propylene glycol in a pilot-scale reactive distillation column: Experimental investigation,
modeling and process analysis, Chem Eng J. 234 (2013) 448–463.

R
https://doi.org/10.1016/j.cej.2013.08.054
[22] M.F. Fernandez, B. Barroso, X. Meyer, M. Meyer, M. Le Lann, G.C. Le Roux, M.

SC
Brehelin, Experiments and dynamic modeling of a reactive distillation column for the
production of ethyl acetate by considering the heterogeneous catalyst pilot complexities,
Chem. Eng. Res. Des. 91 (2013) 2309-2322. https://doi.org/10.1016/j.cherd.2013.05.013
[23] C. Buchaly, P. Kreis, A. Górak, Hybrid Separation Processes – Combination of Reactive
U
Distillation with Membrane Separation Chem. Eng. Process. 46 (2007) 790-799.
https://doi.org/10.1016/j.cep.2007.05.023
N
[24] M. Schmitt, C. Scala, P. Moritz, H. Hasse, n-Hexyl acetate pilot plant reactive distillation
with modified internals, Chem. Eng. Process. 44 (2005) 677–685.
A
https://doi.org/10.1016/j.cep.2003.07.005
[25] E. Harbou, M. Schmitt, S. Parada, C. Grossmann, H. Hasse, Study of heterogeneously
M

catalysed reactive distillation using the D+R tray- a novel type of laboratory equipment,
Chem. Eng. Res. Des. 89 (2011) 1271-1280. https://doi.org/10.1016/j.cherd.2011.01.011
[26] A. Kołodziej, M. Jaroszy´nski, I. Bylica, Mass transfer and hydraulics for KATAPAK-S,
D

Chem. Eng. Process. 43 (2004) 457–464. https://doi.org/10.1016/S0255-2701(03)00134-X


[27] P. Moritz, H. Hasse, Fluid Dynamics in Reactive Distillation Packing Katapak-S, Chem.
TE

Eng. Sci. 54 (1999) 1367-1374. https://doi.org/10.1016/S0009-2509(99)00078-0


[28] L. Spiegel, W. Meier, Distillation Columns with Structured Packings in the Next Decade,
Chem. Eng. Res. Des. 81 (2003) 39-47. https://doi.org/10.1205/026387603321158177
EP

[29] J.M. van Baten, R. Krishna, Gas and liquid phase mass transfer within KATAPAK-S
structures studied using CFD simulations, Chem. Eng. Sci. 57 (2002) 1531 – 1536.
https://doi.org/10.1016/S0009-2509(02)00026-X
CC

[30] C. von Scala, M. Wehrli, G. Gaiser, Heat transfer measurements and simulation of
KATAPAK-M catalyst supports Chem. Eng. Sci. 54 (1999) 1375-1381.
https://doi.org/10.1016/S0009-2509(99)00077-9
[31] J.M. van Baten, J. Ellenberger, R. Krishna, Hydrodynamics of reactive distillation tray
A

column with catalyst containing envelopes: experiments vs. CFD simulations, Catal. Today
66 (2001) 233–240. https://doi.org/10.1016/S0920-5861(00)00625-8
[32] Y. Egorov, F. Menter, M. Kloker, E.Y. Kenig, On the combination of CFD and rate-
based modeling in the simulation of reactive separation processes, Chem. Eng. Process. 44
(2005) 631–644. https://doi.org/10.1016/j.cep.2003.10.011
[33] M. Zivdar, R. Rahimi, M. Nasr, M. Haghshenasfard, CFD Simulations of pressure drop
in KATAPAK-S Structured Packing, Iran. J. Chem. Chem. Eng 2 (2005) 64-71

15
[34] Y. Haroun, D. Legendre, L. Raynal, Direct numerical simulation of reactive absorption in
gas-liquid flow on structured packing using interface capturing method, Chem. Eng. Sci. 65
(2010) 351-356. https://doi.org/10.1016/j.ces.2009.07.018
[35] Y. Haroun, D. Legendre, L. Raynal, Volume of fluid method for interfacial reactive mass
transfer: application to stable liquid film, Chem. Eng. Sci. 65 (2010) 2896-2909.
https://doi.org/10.1016/j.ces.2010.01.012
[36] Y. Haroun, L. Raynal, D. Legendre, Mass transfer and liquid hold up determination in
structured packing by CFD, Chem. Eng. Sci. 75 (2012) 342-348.
https://doi.org/10.1016/j.ces.2012.03.011
[37] C. Dai, Z. Lei, Q. Li, B. Chen, Pressure drop and mass transfer study in structured
catalytic packings, Sep. Purif. Technol. 98 (2012) 78–87.

T
https://doi.org/10.1016/j.seppur.2012.06.035
[38] J. Liu, B. Yang, S. Lu, C. Yi, Multi-scale study of reactive distillation, Chem Eng J. 225

IP
(2013) 280-291. https://doi.org/10.1016/j.cej.2013.03.046
[39] H. Ding, J. Li, X. W., C. Liu, CFD simulation and optimization of Winpak-based

R
modular catalytic structured packing, Ind. Eng. Chem. Res. 54 (2015) 2391-2403.
https://doi.org/10.1021/ie503998v

SC
[40] M. Mazarei Sotoodeh, M. Zivdar, R. Rahimi, CFD simulation of dry and wet pressure
drops and flow pattern in catalytic structured packings, JCHPE 51 (2017) 27-37.
https://doi.org/10.22059/jchpe.2017.62163
[41] M. Behrens, Hydrodynamics and Mass Transfer of Modular Catalytic Structured
Packing, Technische universiteit Delft, 2006,
U
[42] N. Padoin, A.T.O. DalToe, L.P. Rangel, K. Ropelato, C. Soares, Heat and Mass transfer
N
modeling for multicomponent multiphase flow with CFD, Int. J. Heat Mass Transfer 73
(2014) 239-249. https://doi.org/10.1016/j.ijheatmasstransfer.2014.01.075
A
[43] R. Taylor, R. Krishna, Multicomponent mass transfer, John wiley and Sons Inc., New
York, United states, 1993
M

[44] V. Alopaeus, calculation of multicomponent mass transfer between dispersed and


continuous phases, Helsinki University of Technology, 2001,
[45] J.B. Haelssig, A.Y. Tremblay, J. Thibault, S.G. Etemad, Direct numerical simulation of
D

interphase heat and mass transfer in multicomponent vapour-liquid flows, Int. J. Heat Mass
Transfer 53 (2010) 3947-3960. https://doi.org/10.1016/j.ijheatmasstransfer.2010.05.013
TE

[46] S. Shojaee, S.H. Hosseini, A. Rafati, G. Ahmadi, Prediction of the effective area in
structured packings by computational fluid dynamics, Ind. Eng. Chem. Res 50 (2011) 10833-
10842. http://doi.org/10.1021/ie200088d
EP

[47] S.H. Hosseini, S. S., G. Ahmadi, M. Zivdar, Computational fluid dynamics studies of dry
and wet pressure drops in structured packings, J Ind Eng Chem. 18 (2012) 1465-1473.
https://doi.org/10.1016/j.jiec.2012.02.012
CC

[48] T. Zarei, E. Abedini, R. Rahimi, J. Khorshidi, Computational fluid dynamics on the


hydrodynamic characteristics of the conical cap tray, Korean J. Chem. Eng. 34 (2017) 969-
976. https://doi.org/10.1007/s11814-017-0004-6
[49] J. Chen, C. Liu, X. Yuan, G. Yu, CFD simulation of flow and mass transfer in structured
A

packing distillation columns, Chin. J. Chem. Eng. 17 (2009) 381-388.


https://doi.org/10.1016/S1004-9541(08)60220-7
[50] J.U. Brackbill, D.B. Kothe, C. Zemach, A Continuum Method for Modeling Surface
Tension, 100 (1992) 335-354. https://doi.org/10.1016/0021-9991(92)90240-Y
[51] J.F. Rejl, V. Linek, T. Moucha, E. Prokopova, L. Valenz, F. Hovorka, Vapour and liquid
side volumetric mass transfer coefficients measured in distillation column. Comparison with
data calculated from absorption correlations, Chem. Eng. Sci. 61 (2006) 6069-6108.
https://doi.org/10.1016/j.ces.2006.06.003

16
[52] Perry's Chemical Engineering Handbook, Seventh Edition ed., McGraw Hill
[53] Z. Olujic, M. Behrens, L. Spiegel, Experimental characterization and modeling of the
performance of a large-specific-area high-capacity structured packing, Ind. Eng. Chem. Res
46 (2007) 883-893. https://doi.org/10.1021/ie051146f
[54] R. Rahimi, M. Mazarei Sotoodeh, E. Bahramifar, The effect of tray geometry on the
sieve tray efficiency, Chem. Eng. Sci. 76 (2012) 90-98.
https://doi.org/10.1016/j.ces.2012.01.006
[55] Z. Olujic, M. Jodecke, A. Shilkin, G. Schuch, B. Kaibel, Equipment improvement trends
in distillation, Chem. Eng. Process. 48 (2009) 1089–1104.
https://doi.org/10.1016/j.cep.2009.03.004
[56] E. Brunazzi, Mass Transfer Study on Catalytic Structured Packings for Reactive

T
Separations, Chem Eng Trans. 43 (2015) 1015-1020. https://doi.org/10.3303/CET1543170

R IP
SC
U
N
A
M
D
TE
EP
CC
A

17
T
R IP
SC
Figure 1. Methodology flowchart

U
N
A
M
D
TE
EP
CC

Figure 2. Simple schematic of computational domains, Left: geometry of katapak SP-12. Right:
geometry of katapak SP-11 [41]
A

18
13
12 katapak SP-11
11 katapak SP-12
Simulation pressure drop (mbar/m)

10
9 +15%
8
7 -15%
6
5
4
3

T
2
1

IP
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13

R
Experimental pressure drop (mbar/m)
Figure 3. Simulation and Experimental pressure drop for all Fs in katapak SP-11 and 12.

SC
6

5
katapak SP-11
U
Simulation pressure drop

N
katapak SP-12
+9 %
4
A
(mbar/m)

-9 %
3
M

2
D

1
TE

0
0 1 2 3 4 5 6
Experimental pressure drop (mbar/m)
EP

Figure 4. Katapak SP-11 and 12 wet pressure drop vs. Fs.


CC
A

19
3.5
3
Pressure drop (mbar/m)

2.5
2
1.5
closed channels
1
open channels
0.5 separation layer in MCSP-12

T
0 0.2 0.4 0.6 0.8 1
h/H

IP
Figure 5. Pressure drop profile along the closed channels, open channels, and katapak SP-12
separation sections.

R
SC
U
N
A
M
D
TE
EP

Figure 6. Liquid velocity distribution in the center of catalyst bag


CC
A

20
0.6 inlet 0.6 inlet
out-Column simulation out-Column simulation
0.5 out-CFD 0.5 out-CFD

Mass fraction
0.4
Mass fraction

0.4

0.3 0.3

0.2 0.2

0.1 0.1

T
0 0
DEB Ethanol Butanal Water DEB Ethanol Butanal Water

IP
a: First reactive stage b: Second reactive stage

R
inlet
0.7 inlet 0.6

SC
out-Column simulation out-closed channels
0.6 out-CFD 0.5 out-open channels

0.5
Mass fraction

0.4
Mass fraction

0.4
0.3 U
N
0.3
0.2
A
0.2
0.1
0.1
M

0 0
DEB Ethanol Butanal Water DEB Ethanol Butanal Water
D

c: Third reactive stage d: First reactive stage in katapak SP-12


TE

Figure 7. The mole fraction of each component at the inlet and outlet of separation section,
compared with commercial software results in a, b, c: katapak SP-11, d: first reactive stage in
katapak SP-12.
EP
CC
A

21
0.6 inlet 0.6 inlet
out-Column simulation out-Column simulation
0.5 0.5 out-CFD
out-CFD
mole fraction

0.4

mole fraction
0.4
0.3 0.3
0.2 0.2
0.1 0.1

T
0 0
DEB ETHANOL N-BUT-01 WATER DEB ETHANOL N-BUT-01 WATER

IP
a: First reactive stage b: Second reactive stage

R
0.6 inlet
30

SC
out-Column simulation
0.5 out-CFD 25 Ethanol Butanal

0.4 20
U
mole fraction

Conversion

N
0.3 15

0.2 10
A

0.1 5
M

0 0
DEB ETHANOL N-BUT-01 WATER first stage second stage third stage
D

c: Third reactive stage d: Conversion of Ethanol and Butanal


TE

Figure 8 . a, b, c: The mole fraction of each component at the inlet and outlet of stages. d:
Conversion of ethanol and butanal.
EP
CC
A

22
inlet
100 out-Column simulation
90 out-CFD
80
Temperature (˚C)

70
60
50
40
30
20

T
10
0

IP
first stage second stage third stage

Figure 9. Temperature at the inlet and outlet of each stage, compared with commercial

R
software data

SC
U
N
A
M
D
TE
EP
CC
A

23
Table 1. Arrhenius correlation’s parameters for the acetalization reaction [9]
Forward reaction Reverse reaction
Ea (J/mol) 35,505 59,752

A 1.08

  
m3
3

 1.06E+5

  
m3
2


 mol 2 s kgcat   mol s kgcat 
   

T
R IP
SC
U
N
A
M
D
TE
EP
CC
A

24
Table 2: Geometrical characteristics of packings [41]

Property Symbol katapak SP-11 katapak SP-12

Cross sectional fraction Γ 0.40 0.52


cross sectional area for separation section
Г=
total cross sectional area

Volume fraction catalyst Λ 0.46 0.34


volume occupied by the catalyst
Λ=
volume occupied by katapak SP module

T
Channel ratio X 0 0.5

IP
number of open channels per packing layer
X=
total number of channels per packing layer

R
Void fraction εp 0.55 0.7

SC
Thickness catalyst filled pockets tR 13.6 mm 13 mm

Height packing element hpe 200 mm

Corrugation height H
U
6.5 mm
N
Corrugation base width B 9.85 mm
A
Corrugation angle αe or c 41°
M

Overall corrugation angle α 45°

Catalyst diameter dp 1 mm
D
TE
EP
CC
A

25

You might also like