You are on page 1of 9

ASIA-PACIFIC JOURNAL OF CHEMICAL ENGINEERING

Asia-Pac. J. Chem. Eng. 2013; 8: 862–869


Published online 15 April 2013 in Wiley Online Library
(wileyonlinelibrary.com) DOI: 10.1002/apj.1731

Research article
Steady state modeling and simulation of the Oleflex process
for isobutane dehydrogenation considering reaction
network
M. Farsi, A. Jahanmiri* and M. R. Rahimpour
Department of Chemical Engineering, School of Chemical and Petroleum Engineering, Shiraz University, Shiraz, Iran

Received 15 October 2012; Revised 15 February 2013; Accepted 19 February 2013

ABSTRACT: In this study, the performance of moving bed radial flow reactors in Olefex technology to produce isobutene
from isobutane dehydrogenation is studied at steady state condition. The dehydrogenation reactors have been modeled
heterogeneously on the basis of the mass and energy governing laws considering a two-dimensional model. In this system,
isobutane dehydrogenation, hydrogenolysis, propane dehydrogenation, and coke formation reactions occur on the catalyst
surface. The coke deposition on the catalyst surface and the catalyst velocity along the axial direction result to an activity
profile along reactors that has been calculated from proper correlation. To prove the accuracy of the considered mathematical
model and assumptions, simulation results are compared with the plant data at the same process condition. The isobutane
conversion and isobutene selectivity have been obtained at about 38.53% and 90.76%, respectively, which have good
agreement with the plant data. © 2013 Curtin University of Technology and John Wiley & Sons, Ltd.

KEYWORDS: Olefex technology; isobutane dehydrogenation; heterogeneous model; radial flow reactors

INTRODUCTION fixed bed reactors operate adiabatically, and to


regenerate the deactivated catalyst, the reactors are
Currently, because of increasing demand for olefins, changed periodically. In the Olefex technology that
researchers have focused on paraffin dehydrogenation has been commercialized by UOP LLC (Universal Oil
to produce olefins through catalytic processes. Isobutene Products, Illinois, United States), dehydrogenation
as an unsaturated olefin hydrocarbon is a colorless and occurs in the several adiabatic[3]moving bed reactors over
extremely flammable gas at standard pressure and modified Pt-based catalyst. The Olefex reactors
temperature. It is used as a feedstock to produce a variety consist of three moving bed radial-flow reactors, charge
of chemical components such as polybutene, methyl and interstage heaters, and regeneration reactor.
tert-butyl ether and ethyl tertiary butyl ether. In addition, There are few articles in the literature that discuss
alkylation of isobutene by butane produces isooctane as about isobutane dehydrogenation and dehydrogenation
a fuel additive. It is also used as a monomer or copolymer process modeling. Zwahlen and Agnew studied
to produce synthetic rubber and various plastics. isobutane dehydrogenation over a Cr/Al2O3 catalyst
Although isobutene can be separated from refinery in a modified Berty reactor.[4] In this study, the
streams as a side product, the industrial method to produce Langmuir–Hinshelwood–Hougen–Watson model is
isobutene is catalytic dehydrogenation of isobutane over found to provide a good fit to initial rate data. Cortright
Pt-based or Cr-based catalysts. Some commercial et al. presented an accurate rate equation[5]
for isobutane
technologies have been developed for dehydrogenation dehydrogenation over PtSn catalyst. Buzari et al.
of light paraffins such as isobutane, which differ in studied the kinetic of isobutene dehydrogenation over
catalyst, reactor type, and used regeneration system. the commercial Pt–Sn catalyst at a temperature range
The Catofin isobutane dehydrogenation technology of 698–848 K.[6] The results indicated that the
CB&I (Chicago Bridge & Iron Company, The Hague, obtained rate equation based on the Langmuir–
Netherlands), is a cyclic process to produce isobutene Hinshelwood mechanism, considering isobutene
over chromiaalumina catalyst.[1,2] In this process, the adsorption as a rate-limiting step, has very good
agreement with the experimental data. Karinen et al.
investigated isobutane dehydrogenation in a
*Correspondence to: A. Jahanmiri, Department of Chemical
Engineering, School of Chemical and Petroleum Engineering, microreactor over chromia/alumina catalyst.[7] The
Shiraz University, Shiraz, Iran. E-mail: jahanmir@shirazu.ac.ir experimental results showed that the isobutane
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd.
Curtin University is a trademark of Curtin University of Technology
Asia-Pacific Journal of Chemical Engineering MODELING OF ISOBUTANE DEHYDROGENATION 863

conversion and selectivity increases with higher As well as isobutane dehydrogenation, hydrogenolysis,
chromium content. Sahebdelfar et al. modeled the propane dehydrogenation, and coke formation reactions
isobutane dehydrogenation in an adiabatic moving bed take place over the catalyst surface. The high temperature
reactor homogeneously without considering side or residence time causes isobutane cracking to methane
reactions.[8] They neglected the side reactions, whereas and propane (hydrogenolysis reaction) as the main side
the isobutene selectivity in the considered commercial reaction.
process was about 90%. Farsi et al. proposed an axial
flow membrane reactor configuration for isobutane i  C4 H10 þ H2 ↔C3 H8 þ CH4 (2)
dehydrogenation and optimized the process condition to
enhance isobutene production and selectivity.[9] Liang Other side reactions (propane dehydrogenation and
and Hughes studied isobutene synthesis from isobutane coke formation from isobutene reactions) are as follows:
in a membrane reactor over a Pt/Al2O3 catalyst,
experimentally.[10] In addition, they modeled the C3 H8 ↔H2 þ C3 H6 (3)
considered palladium/silver membrane-fixed bed reactor.
In this work, the radial flow reactors in the Olefex i  C4 H8 ! 4 C þ 4 H2 (4)
technology have been modeled heterogeneously
considering the reaction network based on the mass In this work, the kinetic models and reaction rate
and energy conservation laws. In the Olefex technology, parameters have been selected from literature.[12,13]
dehydrogenation occurs in the adiabatic moving bed
reactor over the PtSn/Al2O3 catalyst. The catalyst is
gradually coked as it moves down the reactors. The PROCESS MODELING
outlet catalyst from each reactor is entered to the next
reactor. The outlet catalyst from the last reactor is sent Reaction side
to the regeneration section, and the formed coke on the The Oleflex dehydrogenation process consists of three
catalyst surface is burnt. Then, the regenerated catalyst moving bed radial flow reactors in series wherein the
is sent back to the first reactor. In this system, as well feed stream is entered to the first reactor. The outlet
as the isobutane dehydrogenation reaction, side reactions product from the first reactor is fed to the second
such as hydrogenolysis, propane dehydrogenation, and reactor. Because of coke formation on the catalyst
coke formation reactions occur on the catalyst surface. surface, the outlet catalyst from third reactor is sent to
The coke deposition on the catalyst surface reduces the regeneration section. In the regeneration reactor,
catalyst active site. Thus, coke deposition and axial formed coke is burnt and the regenerated catalyst is
velocity of catalyst along the reactors result to an activity recycled to the first reactor and flows in turn along
profile along the reactors that has been inserted in the reactors. Because the outlet temperature from each
the mathematical model. To prove the accuracy of the reactor drops because of endothermic reactions, the
considered model and assumptions, the results of the interstage heaters are placed between rectors to increase
simulation have been compared with the available the temperature. Fig. 1 shows the schematic diagram of
plant data. the Oleflex process.
Axial velocity of catalyst and coke formation on the
KINETIC MODEL catalyst surface cause an activity, concentration, and
temperature gradient in the axial coordinate. Also, the
In nonoxidative dehydrogenation of isobutane, side radial gradient in the concentration, temperature and
reactions such as cracking, oligomerization, isomerization, activity is due to gas velocity in this direction. Thus,
aromatization, alkylation, and coking occur in the process. a two-dimensional heterogeneous model has been
In the Olefex technology, to increase dehydrogenation developed to simulate the reactors at steady state
reaction rate and minimize the effect of mentioned condition. In the considered mathematical model, the
complex side reactions, Pt–Sn/Al2O3 catalyst is used. In following assumptions are adopted:
this catalytic process dehydrogenation, isobutane cracking
• The gas mixture is in ideal condition (low pressure
and coke formation reactions can be considered.[11]
and high temperature condition).
The isobutane dehydrogenation is a reversible and
• Radial diffusion of mass and energy is negligible.
endothermic reaction. Although increasing the temperature
• The chemical reactions take place on the catalyst
and decreasing the pressure shifts the reaction toward
phase and reactions occurred in the gas phase have
completion, it promotes side reactions and catalyst
been neglected.
deactivation. The kinetic of isobutane dehydrogenation
• The temperature gradient in the solid phase is
over Pt–Sn/Al2O3 catalyst is as follows:
negligible (the Biot number is less that 0.1).
• The feed flows uniformly in the radial direction
i  C4 H10 ↔i  C4 H8 þ H2 (1) of reactors.
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
864 M. FARSI, A. JAHANMIRI AND M. R. RAHIMPOUR Asia-Pacific Journal of Chemical Engineering
Regenerated Catalyst
dCt 1 dPt 1 dT
¼  2 (10)
dr RT dr RT dr

The axial temperature and concentration gradients

Regeneration
Product have been inserted in the governing equations because
of the axial velocity of the catalyst and the occupation
of catalyst voids by gas. Because the catalyst to gas
velocity ratio is very low, moles and energy balance
Feed Deactivated Catalyst equations can be simplified to the following:
Heater
 
g
Figure 1. Schematic diagram of the Oleflex isobutane 1 d ug r Ci;z  
dehydrogenation process. This figure is available in colour  þ av kgi Cis  Cig ¼ 0 (11)
online at www.apjChemEng.com.
r dr
 
Cpg d ug rCt;z Tzg
 þ av hf ðT s  T g Þ ¼ 0 ¼ 0 (12)
r dr
• The catalyst flows uniformly in the axial direction of   X
g
reactors. av kgi Ci;z  Ci;z
s
þ ar;z  ri rB ¼ 0 (13)
• The system is well isolated.
Subject to these assumptions, gas phase component   X
N  
mole and energy balances are expressed by: av hf Tzg  Tzs þ rB ar;z ri ΔHf ;i ¼ 0 (14)
i1
   
d u r C g g
dCi;z 1 d ug r Ct;z XX
1 g i;z
  rB ar;z  ri ¼ 0
  ep ð1  eB Þup (5) r dr
(15)
r dr  dz 
g
av kgi Ci;z  Ci;z
s
¼0
Because of the axial velocity of the catalyst and the
  coke formation on its surface, the coke mass balance
Cpg
d ug rCt;z Tz g
dTz g
  ep ð1  eB Þup Ct;z Cpg (6) should be considered in the model. In this equation, w
r dr dz is the mass fraction of coke to the catalyst.
 g 
av hf Tz  Tz ¼ 0
s
dw
up  ar;z rc ¼ 0 (16)
The component mole and energy balances in the dz
solid phase are expressed by the following:
The pressure drop through the radial direction of the
  X catalytic beds is calculated on the basis of the
g
av kgi Ci;z  Ci;z
s
þ rB ar;z  ri ¼ 0 (7) Tallmadge equation that is usable in laminar and
turbulent regimes.[14]
  dTz s  
 1  ep rB up þ av hf Tzg  Tzs (8)
dz 150 ð1  eÞ2 4:2 ð1  eÞ1:166
X N f ¼ þ (17)
þrB ar;z  ri ðΔHi Þ ¼ 0 Re e3 Re1=6 e3
i1
dP u2 r
¼f (18)
Because of the increasing total mole and reactor area dr Dp
along the radial direction, total mole balance should be
considered in the mathematical model. The total mole The temperature gradient and the axial velocity of
balance is expressed by the following: the catalyst result to an activity profile in the radial
  and the axial direction of the reactors. The catalyst
1 d ug r Ct;z dCt;z activity has been calculated from [15]:
  ep ð1  eB Þup (9)
r dr
XX
dz  
k z
rB ar;z  ri ¼ 0 C1 ¼ Cmax 1  e 1 up ¼ 0 (19)
 
z k2  k z
From the ideal gas condition, total concentration can C2 ¼ Cmax k2  1  e 2 up ¼0 (20)
be expressed by the following: up k1
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
Asia-Pacific Journal of Chemical Engineering MODELING OF ISOBUTANE DEHYDROGENATION 865
ar;z ¼ eða1 C1 þa2 C2 Þ (21) Table 3. Used correlations for physical properties,
mass, and heat transfer coefficient.
where k is a temperature-dependent function, and the Parameter Reference
catalyst activity, a, changes in radial and axial [17]
directions. The parameters of coke deposition equation Gas conductivity
[18]
Mixture heat capacity
have been tabulated in Table 1. The feed specifications, Viscosity of reaction mixtures [18]
the reactor and the catalyst characteristics of the Mass transfer coefficient [19]
[20]
commercial dehydrogenation reactors have been Binary diffusion coefficient
[21]
presented in Table 2.[16] Effective diffusion coefficient in pellet
[22]
Gas–catalyst heat transfer coefficient
Auxiliary equations

There are some parameters in the heterogeneous models


differential equations is not solved analytically and
such as heat and mass transfer coefficients between gas
should be solved numerically. In this system, the reactor
and solid phases that connect the gas and solid phases
length has been discretized in the axial direction, then
governing equations and should be calculated from
the mass and energy governing equations written for
proper correlations. In addition, suitable temperature
each axial segment. Thus, a set of nonlinear ordinary
and component dependent correlations should be chosen
differential equations is obtained in radial direction.
to estimate physical properties of components and mixture
Fourth order Runge–Kutta method has been selected to
such as viscosity, specific heat capacity, heat conductivity,
solve this set of equations in each axial segment. At
and diffusion coefficients along the reactor. The source of
the end of numerical solution, it is possible to plot the
considered correlations to calculate components and
concentration of components and the temperature vs the
mixture physical properties, mass, and heat transfer
reactor length and the radius.
coefficient between phases are summarized in Table 3.

NUMERICAL SOLUTION RESULTS AND DISCUSSION


The mass and energy governing equations create a set of In this section, the process performance to produce
partial differential equations. This set of nonlinear isobutene has been analyzed, and the predicted
component molar flow rate, temperature, and selectivity
Table 1. Parameters of the considered correlation for are presented. To demonstrate the accuracy of the
catalyst activity. considered model and assumptions, the simulation
results have been validated against the available plant
Parameter Value data. The comparison between simulation results and
1
k1 (s ) 7.84  103 plant data for the industrial case is presented in Table 4.
k2 (s1) 1.15  103 It is observed that the predicted isobutane conversion
a1 (kgCat. kgCoke1) 813 and selectivity have good agreement with the observed
a 2 (kgCat. kgCoke1) 289 plant data.
E1 (kJ mol1) 221 Figure 2 shows the isobutene concentration along the
E2 (kJ mol1) 325.8
Cmax (kgCoke. kgCat1) 6.82  104 radial direction of the third reactor. In this reactor,
decreasing catalyst activity in the axial direction due
to coke formation on the catalyst surface and the axial
velocity of the catalyst causes an axial gradient in the
isobutene production. Thus, in the presented results,
Table 2. Feed and product specifications of the the axial average of temperature and mole fractions
commercial dehydrogenation reactors. along the radial direction is presented. The average
Reactor 1 Reactor 2 Reactor 3
component mole fraction in the radial direction is
defined as:
Feed
Temperature ( C) 600 610 604 Table 4. Comparison of simulation results and plant
Flow rate (ton hr1) 106 data for industrial reactor.
Pressure (barg) 1.4 0.9 0.4
Catalyst Absolute Relative
Catalyst loading (ton) 17 18 21 Plant data Model Error (R.E.)
Catalyst density 800
(kg m3) Total conversion 38.8 38.53 0.7%
Catalyst diameter (m) 6  104 Selectivity 90.9 90.76 0.15%

© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
866 M. FARSI, A. JAHANMIRI AND M. R. RAHIMPOUR Asia-Pacific Journal of Chemical Engineering
750

5.55
Isobutene Concentration (mole m )
-3

C4H8

Molar flow rate (mol s )


-1
H2
500 C4H10

4.95

250

4.35
1
0.75 1 0
0 0.5 1 1.5 2 2.5 3
0.5 0.75
0.5 Dimensionless Radius
0.25 0.25
Dimensionless Axis
0 0
Dimensionless Radius (a)
0.14
Figure 2. Isobutene concentration along the radial and
axial direction of the third reactor. This figure is available 0.12
in colour online at www.apjChemEng.com.

C4 H 8 mole fraction
0.08
sum Fi in axial nodes at a specified radial node
yi ðrÞ ¼
sum Ft in axial nodes at a specified radial node
(22) 0.04

Figure 3(a) and (b) shows the isobutene, the isobutane,


and the hydrogen mole flow rates, and the isobutene
mole fraction along the radial direction of reactors, 0
0 0.5 1 1.5 2 2.5 3
respectively. According to this figure, produce isobutene Dimensionless Radius
mole flow in the first, second, and third reactors is about (b)
80.05, 61.75 and 40.6 mol s1. The total produced
isobutene in the process is about 182.4 mol s-1. The Figure 3. (a) Isobutene, isobutane, and hydrogen
isobutane mole flow rate and mole fraction in the first mole flow rate along the radial direction reactors;
reactor has approached to the equilibrium condition (b) Isobutane mole fraction profile along the
radial direction. This figure is available in colour
because of high catalyst activity. Because coke online at www.apjChemEng.com.
deposition on the catalyst surface decreases catalyst
activity, the lower catalyst active sites results to lower
isobutene production and linear concentration profile in
the third reactor. In this system, the higher temperature concentration over the catalyst surface can shift the
at the reactors entrance increases isobutane isobutane dehydrogenation reaction toward completion
dehydrogenation rate in this section. and increases process selectivity.
Figure 4 shows (axial) the average of isobutane Figure 5 (a) and (b) shows the isobutane
conversion in the radial direction of reactors. The total dehydrogenation and the hydrogenolysis reaction rate
isobutane conversion is net of conversion through along the radial direction of reactors, respectively. It is
dehydrogenation and cracking reactions. The isobutane appeared from Figure 5 (a) that the rate of isobutane
conversion in the first, second, and third reactors are dehydrogenation reaction approaches to the equilibrium
about 16.15, 13.36, and 9.02%. The overall isobutane in the first reactor because of higher catalyst activity
conversion has been obtained about 38.53% at the and lower reactor temperature. According to Le
considered operating condition. The isobutane is Châtelier’s principle, when an independent variable of
converted to the isobutene, hydrogen, methane and an equilibrium system such as temperature, pressure, or
propane. Selectivity as main parameter to investigate component concentration changes, the system shifts to
side reactions occurrence, which is defined as the ratio reduce the effect of change and approaches toward a
of produced isobutene per consumed isobutane, has new equilibrium condition. In this system, decreasing
been obtained about 90.76% in this process. The temperature along the radial direction shifts isobutane
reaction kinetics shows that decreasing hydrogen dehydrogenation reaction toward the left side and
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
Asia-Pacific Journal of Chemical Engineering MODELING OF ISOBUTANE DEHYDROGENATION 867
0.4 0.8

Dehydrogenation rate (kmol kgCat s )


-1 -1
0.3 0.6
Isobutane Conversion

0.2 0.4

0.1 0.2

0
0 0 0.5 1 1.5 2 2.5 3
0 0.5 1 1.5 2 2.5 3
Dimensionless Radius Dimensionless Radius
(a)
Figure 4. Isobutane conversion profile. This 0.045
figure is available in colour online at www.
apjChemEng.com.

Cracking rate (kmol kgCat s )


-1 -1
0.03
decreases isobutane production. Also, according to
the stoichiometry Eqns (3–6), increasing hydrogen
content in the reactors by the isobutane and propane
dehydrogenation, and coke formation reactions accel-
erates the hydrogenolysis reaction rate, which results 0.015
to higher isobutane cracking to the propane and the
methane. In addition, propane dehydrogenation decreases
propane concentration in the reaction zone, which results
to higher isobutane cracking. Generally, according to Le 0
0 0.5 1 1.5 2 2.5 3
Châtelier’s principle, decreasing hydrogen concentration
Dimensionless Radius
over the catalyst pellets can shift the equilibrium
limitations toward the lower isobutane cracking. (b)
Figure 6 shows the methane, propane and propene
Figure 5. (a) Isobutane dehydrogenation and (b)
mole fraction in the radial direction of reactors. isobutane cracking reaction rate along the radial
Methane production rate in the first, second, and third direction. This figure is available in colour online
reactors are about 4.25, 7.9 and 6.4 mol s1. The at www.apjChemEng.com.
methane production in the second reactor is higher
compared with the other reactors due to a higher inlet
feed temperature. A part of produced propane in the
reactors is converted to the propene, and results lower temperature at the reactors’ entrance causes lower
propane content compared with the methane. catalyst active sites and decreasing temperature along
Figure 7 shows the radial temperature profile in the gas the radius direction results to increasing catalyst activity.
phase of the reactors. In this process, temperature Also, this figure shows that increasing catalyst residence
decreases because of endothermic reactions. Temperature time in the reactor decreases catalyst activity along
drop in the first reactor is higher compared with the other the axial direction. Produced coke in the catalytic
reactors and conform to isobutane dehydrogenation dehydrocarbon processes is deposited on the catalyst
reaction rate profile. Because reactions take place on the active sites and reduces reaction rate. Coke deposition
catalyst surface, the solid phase temperature will be lower deactivates the catalyst by covering the active sites and
compared with the gas phase temperature and is led to the pore blocking. To reduce coke formation and increasing
heat transfer between solid and gas phases, which results catalyst activity, the hydrogen reach feed is fed to the
to decreasing gas phase temperature. The higher mean hydrocarbon processes. The simulation results show that
temperature in the third reactor results to higher coke the hydrogen mole fraction approaches to 0.56 at the
formation and lower catalyst activity in this reactor. outlet of third reactor, whereas the inlet hydrogen mole
Figure 8(a) and (b) shows the catalyst activity and fraction is about 0.5 in the first reactor. The predicted
coke to catalyst mass fraction profiles along the radial coke to catalyst mass fraction at the outlet of third reactor
direction of reactors. Fig 8(a) shows that higher is obtained at about 1.84%.
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
868 M. FARSI, A. JAHANMIRI AND M. R. RAHIMPOUR Asia-Pacific Journal of Chemical Engineering
0.015 1
Length: 0.5 m
C3H6 Length: 12 m
CH4
C3H8

Catalyst activity
0.01 0.85
Mole fraction

0.005 0.7

0.55
0 0 0.5 1 1.5 2 2.5 3
0 0.5 1 1.5 2 2.5 3
Dimensionless Radius Dimensionless Radius
(a)
Figure 6. Methane, propane, and propene mole 0.03
flow rate along the radial direction of
dehydrogenation reactors. This figure is available
in colour online at www.apjChemEng.com.

Coke fraction (kgCoke kgCat )


-1
0.02

610

590 0.01
Temperature ( C)
o

570
0
0 0.5 1 1.5 2 2.5 3
Dimensionless Radius
550
(b)

Figure 8. (a) Catalyst activity; and (b) coke to


530 catalyst mass fraction profiles along the radial
0 0.5 1 1.5 2 2.5 3 direction of the reactors. This figure is available
Dimensionless Radius in colour online at www.apjChemEng.com.

Figure 7. Radial temperature profiles in the


dehydrogenation reactors. This figure is available
in colour online at www.apjChemEng.com. the third reactor was lower compared with other reactors
due to the higher coke content and lower catalyst activity.
It was explained that hydrogen removal can increase
process selectivity and isobutane conversion.
CONCLUSION

In this study, the isobutane dehydrogenation in moving NOMENCLATURE


bed radial flow reactors were modeled and simulated on
the basis of the mass and energy conservation laws at
steady state condition. To present an exact mathematical a Catalyst activity
model, as well as isobutane dehydrogenation reaction, av Specific surface area of catalyst pellet (m2 m3)
side reactions such as dehydrogenation, hydrogenolysis, Ac Cross section area (m2)
propane dehydrogenation and coke formation reactions Ci Molar concentration of component i (mol m3)
have been considered in the model. It was shown that Cp Specific heat capacity (J mol1 K1))
isobutane conversion is about 38.53% in the outlet of Ct Total mole concentration (mol m3)
the third reactor. In addition, isobutene selectivity was D Diameter (m)
calculated at about 90.76%, which proves the importance f Friction factor
of side reactions. In this system, isobutane conversion in hf Gas-solid heat transfer coefficient (W m2 K1)
© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
Asia-Pacific Journal of Chemical Engineering MODELING OF ISOBUTANE DEHYDROGENATION 869

k Catalyst activity constant (s1) [3] M.M. Bhasin, J.H. Cain, B.V. Vora, T. Imai, R.R. Pujado.
Appl. Catal. A: Gen., 2001; 221, 397–419.
kg Mass transfer coefficient (m s1) [4] A.G. Zwahlen, J.B. Agnew. Ind. Eng. Chem. Res., 1992; 31,
L Reactor length (m) 2088–2093.
P Total pressure (bar) [5] R.D. Cortright, P.E. Levin. A. James. Ind. Eng. Chem. Res.,
1998; 37, 1717–1723.
Re Reynolds number [6] A. Buzari, F. Salehi Rad, M.A. Musavian. Kinetic
r Radial reactor coordinate (m) Investigation and Catalyst Preparation of Isobutane to
ri Rate of reaction (mol kgCat1 s1) Isobutene Dehydrogenation Process, M.Sc Thesis in Tehran
University, Iran, 2006.
T Temperature ( C) [7] R. Karinen, S. Airaksinen, P. Kiviranta, K. Keskinen,
u Superficial velocity of fluid phase (m s1) J. Linnekoski, P.U. Kyyny, A. Outi, I. Krause. Top. Catal.,
y Component mole fraction 2011; 54, 1206–1212.
[8] S. Sahebdelfar, P. Moghimpour Bijani, M. Saeedizad,
z Axial reactor coordinate (m) F. Tahriri Zangeneh, K. Ganji. Applied Catalysis A: Gen.,
2011; 395, 107–113.
Greek letters: [9] M. Farsi, A. Jahanmiri, M.R. Rahimpour. Journal
of Industrial and Engineering Chemistry, 2012; 18,
1676–1682.
r Density of fluid phase (kg m3) [10] W. Liang, R. Hughes. Catalysis Today, 2005; 104, 238–243.
a Catalyst activity constant (kgCat kgCoke1) [11] D. Sanfilippo, I. Miracca, F. Trifirò. Encyclopaedia of
Hydrocarbon-Dehydrogenation Processes, Eni Publisher,
 Effectiveness factor Rome, 2007.
e Bed void fraction [12] A. Abdi, V. Kiamanesh. Modeling and Simulation of the
ΔH Heat of reaction (j mole3) Catalytic Reactor for Propane Dehydrogenation in MTBE
Plant, Iranian petrochemical conference, Iran, 2008.
ΔP Pressure difference (Pa) [13] P. Moghimpour Bijani, S. Sahebdelfar, K. Catal. 2008; 49,
599–605.
Superscripts: [14] J.A. Tallmadge. AIChE J., 1970; 16, 1092–1103.
[15] K. Mohajeri, M. Khorashe, Development of a kinetic model
for catalyst deactivation in dehydrogenation of light paraffins,
B Reactor bed M.Sc Thesis in Sharif University, Iran, 2008.
g In the bulk gas phase [16] Operating Data Sheets of MTBE Plant, Bandar Imam
Khomeini Petrochemical Complex, Iran.
p Particle [17] A.L. Lindsay, L.A. Bromley. Ind. Eng. Chem., 1950; 42,
s At the catalyst surface 1508–1510.
[18] B.E. Poling, J.M. Prausnitz, J.P. O’Connell. The Properties
of Gases & Liquids, McGraw Hill, New York, 2001.
[19] E.L. Cussler. Diffusion Mass Transfer in Fluid Systems,
REFERENCES University Press, Cambridge, 1984.
[20] J.O. Hirschfelder, C.F. Curtis, R.B. Bird. Molecular Theory
[1] A.H. Weiss. The Manufacture of Propylene in: Refining of Gases and Liquids, John Wiley and Sons, New York,
Petroleum for Chemicals, American Chemical Society, 1952.
Washington, 1970. [21] C.R. Wilke. Chem. Eng. Progress, 1950; 46, 95–104.
[2] R.G. Craig, D.C. Spence. Handbook of Petroleum Refining [22] P.N. Dwivedi, S.N. Upadhyay. Ind. Eng. Chem. Proc. Des.
Processes, McGraw Hill, New York, 1986. Dev., 1977; 16, 157–165.

© 2013 Curtin University of Technology and John Wiley & Sons, Ltd. Asia-Pac. J. Chem. Eng. 2013; 8: 862–869
DOI: 10.1002/apj
免费论文查重:http://free.paperyy.com
3亿免费文献下载:http://www.ixueshu.com
超值论文自动降重:http://www.paperyy.com/reduce_repetition
PPT免费模版下载:http://ppt.ixueshu.com

-------------------------------------------------------------------------------

You might also like