Vernikovskaya2011 Banyak

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Chemical Engineering Journal 176–177 (2011) 158–164

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Dehydrogenation of propane–isobutane mixture in a fluidized bed reactor over


Cr2 O3 /Al2 O3 catalyst: Experimental studies and mathematical modelling
N.V. Vernikovskaya a,b,∗ , I.G. Savin c , V.N. Kashkin a , N.A. Pakhomov a,b , A. Ermakova a , V.V. Molchanov a ,
E.I. Nemykina a , O.A. Parahin d
a
Boreskov Institute of Catalysis SB RAS, Novosibirsk 630090, Russia
b
Novosibirsk State University, Novosibirsk 630090, Russia
c
JSC “Tobolsk-Neftehim”, Tobolsk, Russia
d
JSC “NPK Sintez”, Barnaul, Russia

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

Article history: An influence of propane addition to the inlet feed on the performance of industrial fluid bed reactor for
Received 15 December 2010 isobutane dehydrogenation was experimentally and numerically studied. Experiments on dehydrogena-
Received in revised form 24 May 2011 tion of propane–isobutane mixture in a pilot fluidized and in a lab fixed bed reactors were performed
Accepted 25 May 2011
over Cr2 O3 /Al2 O3 industrial catalyst. Adding C3 H8 to the reactor inlet was found to increase experimen-
tal conversion of C3 –C4 mixture and the total process selectivity to olefins. Results of the mathematical
Keywords:
modelling of the industrial-scale fluidized bed reactor show some benefits of C3 H8 addition. Selectivity
Fluidized bed reactor
to i-C4 H8 was found to be high enough and grows slightly from 86 to 89% on increasing inlet C3 H8 frac-
Dehydrogenation
Propane–isobutane mixture
tion from 0 to 60 wt%. Inlet concentrations of C3 H8 up to 20 wt% lead to the apparent selectivity to C3 H6
Mathematical modelling exceeding 100%. Coke yield rises slowly allowing safe industrial fluid bed reactor operation.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction (1) thermodynamics restrictions for propane dehydrogenation are


not too hard;
The processes of hydrocarbons dehydrogenation play an impor- (2) propene is widely used for polypropylene production.
tant role in the world industry of olefins production. Using fluid bed
technology for hydrocarbons dehydrogenation at subatmospheric
pressure has definite advantages, such as low capital and operation Main goal of the present study was to predict the effect of addi-
cost, relatively simple reactor design, continuous operation and tion of propane to the inlet reactor feed on the industrial fluid bed
easy process control. Fluid bed technology is successfully applied at reactor performance.
industrial scale primarily for isobutene and isopentene production
in Russian Federation and Saudi Arabia. In case of isobutene pro-
duction fluid bed reactor productivity varies greatly and depends on 2. Experimental
the MTBE (Methyl Tertiary Butyl Ether) market demands, because
isobutene produced is used further in MTBE synthesis. If it is nec- 2.1. Catalyst preparation
essary to reduce the fluid bed reactor productivity one have to
decrease the feed of hydrocarbons into the reactor. This results in A modified sample of industrial fluid bed Cr2 O3 /Al2 O3 catalyst
reduction of reactor stability and olefin yield because of decreas- “KDM” (which is commercially produced by JSC “NPK Sintez”, Rus-
ing the quality of the gas distribution and appearing the stagnant sia) was prepared by incipient wetness impregnation of gibbsite
regions with intensive coke formation. So, to raise the reactor sta- thermal activation product [1] with an aqueous solution of CrO3
bility and to maximize olefins yield in this case addition of another with KOH and ZrO2 additives. A product of centrifugal thermal
paraffin at the fluid bed reactor inlet looks attractive. Propane activation of gibbsite (the CTA-product), obtained by applying the
seems to be the most suitable co-feed because: centrifugal flash reactor technology, was used as a support for this
catalyst [1,2]. The calculated loadings of Cr2 O3 , K2 O and ZrO2 in
calcinated catalyst were 15.0, 1.5 and 1.0 wt%, respectively. The
impregnation has been carried out for 1 h at continuous mixing in
∗ Corresponding author at: Boreskov Institute of Catalysis SB RAS Novosibirsk Z-shaped commingler at room temperature. The obtained samples
630090, Russia. Tel.: +7 383 3269438; fax: +7 383 3306878. were dried overnight at room temperature followed by drying at
E-mail address: vernik@catalysis.ru (N.V. Vernikovskaya). 120 ◦ C for 4 h and calcination at 750 ◦ C for 1 h.

1385-8947/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.05.115
N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164 159

Nomenclature
Subscripts
Ar Archimed number e dense phase
cp,g , cp,s apparent gas phase (catalyst particles) heat capacity b bubble phase
(J/(kg K))
0 , c0
cp,g p,s gas phase (catalyst particles) heat capacity (J/(kg K))
Dt reactor diameter (m) 2.2. Catalyst properties
Dz effective gas dispersion coefficient (m2 /s)
Dp catalyst particle diameter (m) The surface area of obtained catalyst was 100 m2 /g. There are
Ei activation energy (J/mol) ␥-Al2 O3 phase and traces (less then 5 wt%) of ␹-Al2 O3 phase in the
H height of extended bed (m) sample according to the XRD data. Four different chromium states
−r Hj jth reaction heat effect (kJ/mol) are specified in sample calcinated on the air [3]. They are two types
ke effective thermal conductivity (J/(m2 K)) of six-valent chromium (water-insoluble Cr6+ species (about 1 wt%)
ki the rate constant of ith reaction (mol/(m3 s atm)) grafted on the support surface, water-soluble Cr6+ species as chro-
ki0 the pre-exponential factor of the ith reaction con- mates (about 1 wt%)) and two types of trivalent chromium: Cr3+
stant (mol/(m3 s atm)) species in the lattice of alumina support (solid solution of Cr3+ in
Keq1(2) equilibrium constant of propane (isobutane) dehy- ␥-Al2 O3 ) and fine-dispersed Cr2 O3 particles on the support surface.
drogenation reactions (atm) Fine-dispersed Cr2 O3 particles are active sites in dehydrogenation.
Ki adsorption parameter (atm−1 or atm−0.5 ) Reducing of Cr6+ also leads to formation of active sites. However,
N number of experiments the particles obtained by reduction of soluble Cr6+ species are more
Nc number of components active and selective. Majority of Cr3+ species, which forms solid
NR number of reactions solution with support, evidently does not take part in dehydro-
P, T pressure (atm) and temperature (K) under work genation process, but it plays the important role in stabilization of
conditions support structure at calcination.
P0 , T0 pressure (atm) and temperature (K) under normal
conditions 2.3. Fixed bed reactor
Pi ith component partial pressure (atm)
Rg gas constant (J/(mol K)) The kinetics reactions study was carried out using a fixed bed
Re Reynolds number plug-flow reactor at atmospheric pressure. The reactor was a 12-
Sc Schmidt number mm-i.d., 130-mm-long quartz tube placed inside a sand bath with
ub gas velocity in bubble phase under normal condi- constant temperature. The reactor was equipped by thermocou-
tions (m/s) ples inside a thermowell of 6-mm o.d. installed at the reactor
ue gas velocity in dense phase under normal conditions axis, flow mass controllers and gas chromatography. The mixture
(m/s) of 0.08–0.12 mm-i.d. catalyst diluted by 0.25–0.5 mm-i.d quartz
Ug = εb ub + (1 − εb )ue superficial overall gas velocity under sand at a ratio 1:1 ÷ 1:7 for the purpose of eliminating axial
normal conditions (m/s) temperature gradients was loaded into the reactor. The fresh cat-
u0g inlet superficial overall gas velocity under normal alyst was heated to the required temperature under the flow of
conditions (m/s) N2 . The i-butane–propane mixture was fed into the reactor dur-
Us velocity of downward catalyst moving (m/s) ing 10 min. Then the reaction products were analysed with gas
Wj the jth reaction rate (mol/(m3 s)) chromatography. The catalyst was regenerated with air at 650 ◦ C
x reactor axial coordinate (m) during 10–30 min. The temperature reduction after regeneration
Yi total (on two phases) value of ith compound weight up to required values was done under the flow of N2 . The opera-
fraction tion conditions were varied within the next values: temperature,
ybi , yei mole fraction of ith compound in bubble and dense 565–590 ◦ C; residence time, 0.2–20 s; propane: isobutane volume
phases ratio, 0–100 mol%; hydrogen concentration, 0–80 mol%.
yije experimental concentration of ith component in jth
experiment, mole fraction 2.4. Pilot fluidized bed reactor
yijc calculated concentration of ith component in jth
experiment, mole fraction Experiments were performed to show the effect of addition of
propane to the inlet reactor feed on the pilot fluidized bed reactor
Greek letters performance. The fluidized bed reactor with Cr2 O3 /Al2 O3 catalyst
aˇ undimensional coefficient was a 35-mm-i.d., 600-mm-long quartz tube inside an electrical
ˇ mass transfer coefficient (s−1 ) furnace. The reactor was equipped with a distributor plate. The
 PT0 /P0 T temperature was measured by thermocouples inside a thermowell
εb fraction of bubble phase cross section at the centre of the catalyst bed. The catalyst volume was 100 cm3 .
εbc void fraction of bubble phase The height of expanded catalyst bed was 15–20 cm. The reactor
εd void fraction of dense phase feed contained either isobutene or propane–isobutane mixture.
ε0d void fraction of dense phase at the beginning of bub- The range of operating conditions was as follows: temperature,
ble formation 555–585 ◦ C; feed velocity, 28–42 l/h.
ε0 void fraction of fixed bed
ij stoichiometric coefficients 3. Mathematical model of industrial fluidized bed reactor
s apparent catalyst bed density (kg/m3 ) for i-C4 H8 dehydrogenation
S0 catalyst bed density (kg/m3 )
A microspheric catalyst used in the fluid bed dehydrogena-
ton process circulates between reactor and regeneration unit. The
160 N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164

Fig. 1. The scheme of mass and heat transfer processes in the industrial fluid bed reactor for i-C4 dehydrogenation.

regeneration unit provides coke combustion. The regenerated hot The experimental results of dehydrogenation of
catalysts fed into the top of the reactor unit thus increasing reactor propane–isobutane mixture in pilot fluidized bed reactor are
top temperature. Inlet reactor gas feed flows upwards in the reac- shown in Table 1. One can see the increase of isobutene yield
tor unit and contacts with catalyst, the schemes of mass and heat because propane is diluent for isobutane. This leads to decreasing
transfer in the reactor are shown in Fig. 1. The reactor is supplied of isobutene partial pressure and shifting the dehydrogenation
with internals for bed staging. So, the following issues of fluid bed equilibrium of the mixture components to the desired isobutene.
reactor for alkane dehydrogenation should be taken into account: So at 560 ◦ C one can see the isobutene yield increase from 44.4 wt%
in the case of dehydrogenation of pure isobutene to 51 wt% in
the case of dehydrogenation of the propane–isobutane mixture
(1) Gas velocity range in the reactor is 0.15–0.45 m/s at the opera-
(compare rows 2 and 5 in Table 1). At that the value of isobutene
tion conditions. This velocity corresponds to bubbling fluidized
selectivity remains almost the same in these two experiments and
bed conditions.
makes up 93.3 and 93.9%, respectively. At 581 ◦ C (see row 4 in
(2) Linear gas velocity is varied throughout the height due to reac-
Table 1) apparent propylene selectivity exceeds 100% because of
tion volume change through the dehydrogenation reactions.
occurring the main reaction of propane dehydrogenation and side
(3) The reducing the catalyst back mixing due to the presence of
reaction of C4 cracking to propylene and methane.
internals leads to axial temperature gradients existence in the
reactor.
4.2. Mathematical modelling of industrial fluid bed reactor

Therefore a two-phase non-isothermal reactor model taking The comparison of simulation results with experimental data
into account axial gas and heat dispersion and reactions occur- from industrial fluid bed pure isobutane dehydrogenation reac-
ring in these two phases was used for the industrial-scale reactor tor at JSC “Tobolsk-Neftechim” is shown in Fig. 2. The simulated
modelling (Appendix B). The model was combined with proposed
kinetic model (Appendix A).

4. Results and discussion

4.1. Pilot fluid bed experiments

Thermodynamic calculations for dehydrogenation of a model


C3 –i-C4 paraffins mixture indicate that there is a complex mutual
effect of the components on the equilibrium yield of olefins [2].
Each component can act simultaneously: (i) as a diluent shifting
the dehydrogenation equilibrium of the mixture components to the
desired olefin and (ii) as an additional source of hydrogen, which
shifts the dehydrogenation equilibrium to the left.
Experiments on dehydrogenation of model C1 –C4 paraffin mix-
tures on Cr2 O3 /Al2 O3 catalysts in the pilot fluidized and fixed beds Fig. 2. An influence of the average bed temperature on isobutylene selectivity and
show an increase of the isobutane conversion, an increase of the yield: lines – modelling results, symbols – experimental data from industrial fluid
process selectivity and the yield of isobutene. bed reactor. u0g = 0.15 m/s, Ptop = 1.2 atm, i-C4 H10 in = 95.5 wt%.
N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164 161

Conversion (%)i-C4 H10 (C3 H8 )

58.3 (33.6)
65.1 (29.4)
54.3 (34.8)
57.5 (40.6)
47.8
Selectivity (mass%)i-C4 H8 (C3 H6 )

Fig. 3. Modelling results: the temperature and concentrations in bubble phase


throughout the height of fluid bed reactor. C3 H8 in = 45 wt%, i-C4 H10 in = 51 wt%. u0g =
0.15 m/s, Ptop = 1.2 atm.
79.0 (104.8)
91.8 (81.3)

89.0 (92.3)

93.9 (89)
93.3
Product yield (mass%)i-C4 H8 (C3 H6 )
Reactor performance

51.5 (30.8)
52.8 (33)

51.9 (31)

51 (31)

Fig. 4. An influence of average bed temperature and inlet propane mass fraction on
44.6

C3 H6 fraction in PPF at the reactor outlet.

isobutylene yield and selectivity are in a good agreement with the


Temperature in the middle of fluidized bed (◦ C)

experimental data.
The results of mathematical modelling of dehydrogenation of
propane–isobutane mixture in industrial fluid bed reactor over
Cr2 O3 /Al2 O3 catalyst are shown in Figs. 3–9. The temperature
and concentration profiles one can see in Fig. 3. The temperature
increases with height due to hot regenerated catalyst is fed into the
top of the reactor. The paraffin concentration decreases and olefins
one increases with reactor height.
The separation of propylene from the propane–propylene frac-
tion (PPF) is economically efficient if propylene weight fraction
is more than 35% in PPF. The influence of average temperature
of fluidized bed reactor and propane mass fraction at the reac-
582
560
571
581
560
The results of the pilot fluidized bed reactor experiments.

velocity (l/h)
Feed space

42
28
28
28
28
C3 H8

44.1

43.6
42.2
43.6

concentrations of

propane (wt%)
isobutane and

i-C4 H10
Table 1

Inlet

55.9
99.9
56.6
57.8
56.4

Fig. 5. An influence of average bed temperature and inlet propane mass fraction on
the isobutylene yield.
162 N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164

Fig. 6. An influence of average bed temperature and inlet propane mass fraction on Fig. 9. An influence of average bed temperature and inlet propane mass fraction on
the isobutylene selectivity. the coke yield.

in the IIF, that is satisfied under the whole range of the process
temperatures and inlet propane concentration (Fig. 5).
The isobutylene yield goes up along with increasing average
layer temperature and propane mass fraction (Fig. 5). The rise of the
inlet propane concentration leads to reduction of hydrogen concen-
tration in the reaction mixture shifting isobutane dehydrogenation
equilibrium to the right. The isobutylene selectivity goes down on
increasing temperature due to increasing the rate of cracking reac-
tion (Fig. 6). And the isobutylene selectivity goes up on increasing
inlet propane concentration because of the partial pressure of the
isobutylene (the main cracking products source) is decreased.
The dependences of propylene yield (Fig. 7) and selectivity
(Fig. 8) have more complicated behavior, that is caused by the
propylene is the target product of propane dehydrogenation reac-
Fig. 7. An influence of average bed temperature and inlet propane mass fraction on tion and it is the product of the side isobutylene cracking reaction.
the apparent propylene yield. This is the reason for the apparent propylene selectivity is raised
on temperature increasing and exceeds 100% at low inlet propane
concentration (Fig. 8). The large amount of propylene formed by
tor inlet mixture on C3 Н6 fraction in PPF at the reactor outlet are
isobutylene cracking reaction is summarized with the propylene
shown in Fig. 4. To provide weight fraction of C3 Н6 > 35% in PPF
generated by the propane dehydrogenation reaction resulting to
the average bed temperature must be 575–585 ◦ C, that is typi-
enormous high C3 H6 selectivity. Increasing inlet propane concen-
cal temperature for pure isobutane fluid bed dehydrogenation. To
tration (other conditions being kept constant) leads to decrease the
keep C3 H6 fraction constant (35 wt%) on increasing i-C4 at the inlet
apparent propylene selectivity, but the selectivity values are still
the bed temperature have to be increased. This effect results from
sufficiently high up to the inlet propane concentration of 45 wt%
increasing of hydrogen concentration in the reaction mixture and
(Fig. 8).
increasing i-C4 at the inlet thus shifting the equilibrium of dehy-
The influence of average bed temperature and inlet propane
drogenation reaction of propane to the left.
mass fraction on the coke yield is presented in Fig. 9. The coke yield
In case of separation of isobutylene from isobutane–isobutylene
increases on increasing bed temperature and inlet propane weight
fraction (IIF) the isobutylene mass fraction must be higher 45%
fraction due to kinetics reason. Coke formation rate was experimen-
tally found to decrease in our kinetics experiments on increasing
hydrogen concentration. Increasing inlet propane concentration
results to decreasing hydrogen concentration in the reaction mix-
ture, so coke formation rate is increased. Nevertheless, calculated
coke yield not exceeds 1.5 wt% under the whole of temperature and
propane inlet concentration ranges, thus allowing safe industrial
fluid bed reactor operation.

5. Conclusions

Experiments on dehydrogenation of model C1 –C4 paraffin mix-


tures on Cr2 O3 /Al2 O3 catalysts in pilot fluidized and fixed beds
show an increase of the total conversion of C3 –C4 paraffin to olefins
and an increase of the total process selectivity to olefins.
The mathematical modelling and optimization of the
Fig. 8. An influence of average bed temperature and inlet propane mass fraction on propane–isobutane mixture in the industrial fluid bed reactor
the apparent propylene selectivity. over Cr2 O3 /Al2 O3 catalyst was carried out. The influence of such
N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164 163

Table A1
The list of operation conditions used for kinetics investigation. k20 exp(−E2 /RT )PiC4 H8
W2 =
Parameter Value Dimension 1 + K4 PH2 + K5 PiC4 H8

Temperature 565–590 C
Pressure 101 kPa k30 exp(−E4 /RT )(PC3 H8 − (PC3 H6 PH2 /Keq2 ))
Catalyst load 1.4–5 ml W3 =
Residence time 0.2–20 s
0.5
1 + K1 PiC H
+ (K2 PH2 )0.5 + K3 PC0.5H
4 8 3 6
Reactor type Plug flow
k40 exp(−E4 /RT )PC3 H6
W4 =
parameters as average reactor temperature and inlet mixture 1 + K4 PH2 + K5 PiC4 H8
composition on the process performance was studied.
As a result of the present modelling the using k60 exp(−E5 /RT )PC3 H6
W5 =
propane–isobutane mixture for the industrial-scale fluidized 1 + K4 PH2 + K5 PiC4 H8
bed olefins production was shown to have the following benefits:
k60 exp(−E6 /RT )PiC4 H8
• Selectivity to i-C4 H8 is high enough and grows slightly from 86 W6 =
1 + K4 PH2 + K5 PiC4 H8
to 89% on increasing inlet C3 H8 fraction from 0 to 60 wt%.
• Inlet concentrations of C3 H8 up to 20 wt% lead to the apparent
k70 exp(−E7 /RT )PC2 H6
selectivity to C3 H6 exceeding 100%. This effect arises due to i-C4 W7 =
cracking leading to additional propene formation. 1 + K4 PH2 + K5 PiC4 H8
• Reasonably high propene concentration of 35 wt% in C3 H8 –C3 H6
mixture could be obtained at relatively low bed temperature k80 exp(−E8 /RT )PiC4 H10
W8 =
575–585 ◦ C, which is close to the typical temperature for pure 1 + K4 PH2 + K5 PiC4 H8
i-C4 fluid bed dehydrogenation.
• Coke yield rises slowly (from 0.7–1.0 wt% at 100% i-C4 at the inlet k90 exp(−E9 /RT )PC3 H8
to 1.2–1.4 wt% at 45% C3 H8 at the inlet) allowing safe industrial W9 =
1 + K4 PH2 + K5 PiC4 H8
fluid bed reactor operation.
The equilibrium constants of the reactions of propane and isobu-
Appendix A. Kinetics tane dehydrogenation were calculated using “HYSYS” software
package.
The kinetics and mechanism of either isobutane dehydrogena- −3 2
Keq1 = e−(118 505−130.321T −6.6369×10 T )/8.31/T
tion or propane one over Cr2 O3 /Al2 O3 catalyst have been explored
in various publications [4–9]. The Langmuir–Hinshelwood model
2 )/8.31/T
[5] or the Langmuir–Hinshelwood–Hougen–Watson one [6] gave Keq2 = e(124 912−127.9T −6.751E−3T
the best fit to experimental isobutane dehydrogenation data. A
The kinetic parameters were estimated using gradient method
Langmuir–Hinshelwood model with strong adsorption of propene
of the minimization of the functional:
was found to provide the best fit to describe the reaction kinetics
1  e
N Nc −1
of propane dehydrogenation to produce propene over Cr2 O3 /Al2 O3
2
catalyst in fixed bed [9]. F= (yij − yijc )
Nc · N
In the present work the data obtained in our experiments in the j=1 i=1
fixed bed plug-flow reactor was used for kinetics investigation of
The kinetic parameters obtained cannot be listed in the present
dehydrogenation of propane–isobutane mixture over Cr2 O3 /Al2 O3
work due to secrecy agreements with JSC “NPK Sintez”.
catalyst. Several sets of experiments have been performed, varying
the temperature, gas flow rate and feed composition (see Table A1).
Appendix B. Mathematical formulation
The following reaction set was chosen to describe
isobutane–propane dehydrogenation:
The material balances for the bubble and dense phases:
(1) iC4 H10 ↔ iC4 H8 + H2 ∂2 ybi ∂(ub ybi ) ˇ
Dzb − − · (ybi − yei )
(2) iC4 H8 + H2 → C3 H6 + CH4 ∂x2 ∂x εb
(3) C3 H8 ↔ C3 H6 + H2 ⎛
(4) C3 H6 + 2H2 → C2 H6 + CH4 
NR

(5) C3 H6 → 3Ccoke + 3H2 +


Rg T0
⎝ 1 − εbc · ij Wjb
P0 1 − ε0
(6) iC4 H8 → 4Ccoke + 4H2 j=1
(7) C2 H6 + H2 → 2CH4 ⎞
(8) iC4 H10 → C3 H6 + CH4
1 − εd 1 − εb  
Nc NR

(9) C3 H8 + H2 → C2 H6 + CH4 + yei · ij Wjb ⎠ = 0 (1)


1 − ε0 εb
i=1 j=1
The dehydrogenation equilibrium, the concentrations of paraf-
fin, olefins and hydrogen was shown [5–9] to have effect on the ∂2 yei ∂y ˇ
propane–isobutane dehydrogenation over chromia–alumina cata- Dze − ue ei − · (yei − ybi )
∂x2 ∂x (1 − εb )
lyst. So the next kinetic equations were used for dehydrogenation, ⎛ ⎞
cracking and coke formation reaction rates: 
NR
 
Nc NR
Rg T0 1 − εd
+ ⎝ ij Wje − yei ij Wje ⎠ = 0
k10 exp(−E1 /RT )(PiC4 H10 − (PiC4 H8 PH2 /Keq1 )) P0 1 − ε0
W1 = j=1 i=1 j=1
0.5
1 + K1 PiC + (K2 PH2 )0.5 + K3 PC0.5H
H
4 8 3 6 (2)
164 N.V. Vernikovskaya et al. / Chemical Engineering Journal 176–177 (2011) 158–164

Table B1
The parameters used in the simulation of industrial fluid bed reactor.
The heat balance:
Parameter Value Dimension
∂2 T ∂T
ke · − (cp,g · g · Ug − cp,s · s · Us ) 0
cp,s 0.808 kJ/K/kg
∂x2 ∂x
Dt 4.5 m

NR 1 − ε 1 − εbc
 H 4.9 m
d s 0 kg/m3
− r Hj · Wje + W =0 (3) 2400
1 − ε0 1 − ε0 jb ue 0.01 m/s
j=1 u0g 0.15 m/s
us 0.01–0.1 m/s
The equation for gas velocity dependence on mole flow changes: Dp 8E−5 m
Ptop 120 kPa

P0 ∂ub  Nc
1 − εb 1 − εd 
NR
1 − εbc
Nc NR εb 0.35 Dimensionless
= ij Wje + ij Wjb εbc 0.95 Dimensionless
Rg T0 ∂x εb 1 − ε0 1 − ε0 ε0d 0.48 Dimensionless
i=1 j=1 i=1 j=1 ε0 0.4 Dimensionless
(4) ˛ˇ 0.87 Dimensionless

To 630 C

The corresponding boundary conditions are:


The reactor performance is evaluated through the following
dyei 0 ), dy 0 ),
x=0: Dze dx
= ue (yei − yei Dzb dxbi = ub (ybi − ybi quantities:
(u0 − (1 − ε )u )
g b e

u0b = , dT dx
= 0, YiC4 H8
εb (5) Yield of isobutene = × 100%
dyei dybi 0
YiC
x=H: = 0; = 0; − ke dT dx 4 H10
dx dx
= (cp,g · g · Ug − cp,s · s · Us ) · (T − T0 )

The catalyst balance in bubble and dense phases: YiC4 H8


Isobutene selectivity = 0
× 100%
YiC − YiC4 H10
4 H10
(1 − εb ) · (1 − ε0d ) − εb (1 − εbc )
εd = 1 −
1 − εb

YC3 H6
The model parameters Yield of propene = × 100%
Averaged over reactor cross section heat capacities of gas and YC0
3 H8
solid phases are:

0
cp,g = cp,g · (εb εbc + εd (1 − εb )) YC3 H6
Propene selectivity = × 100%
YC0 − YC3 H8
3 H8
0
cp,s = cp,s · ((1 − εb )(1 − εd ) + εb (1 − εbc ))
Coke yield = Ycoke × 100%
The density of solid material is:

s = s0 · ((1 − εb )(1 − εd ) + εb (1 − εbc )) References


Coefficients estimation [10]: [1] L.A. Isupova, Y.Y. Tanashev, I.V. Kharina, E.M. Moroz, et al., Physico-chemical
Axial gas dispersion coefficient nearby the stages: properties of TseflarTM -treated gibbsite and its reactivity in the rehydration
process under mild conditions, Chem. Eng. J. 107 (2005) 163–169.
Ug · H 0.715 · Dt0.285 [2] V.V. Molchanov, N.A. Pakhomov, L.A. Isupova, V.A. Balashov, et al., RF Patent
Dz,g = No. 2,322,290, 2008 (Priority of 18.12.2006).
3.47Ar 0.149 Re0.0234 Sc −0.231 [3] E.I. Nemykina, N.A. Pakhomov, V.V. Danilevich, V.A. Rogov, V.I. Zajkovskij,
T.V. Larina, V.V. Molchanov, Influence of chromium content on the properties
Axial gas dispersion coefficient between the stages:
of microspherical chromia–alumina catalyst for isobutene dehydrogenation
obtained by applying the product of centrifugal thermal activation of gibbsite,
Ug · H 0.715 · Dt0.285 Kinet. Catal. 51 (6) (2010) 929–937.
Dz,g = 0.05
3.47Ar 0.149 Re0.0234 Sc −0.231 [4] I. Miracca, L. Piovesan, Light paraffins dehydrogenation in a fluidized bed reac-
tor, Catal. Today 52 (1999) 259.
Interphase mass transfer coefficient: [5] J. Happel, K. Kamholz, D. Walsh, V. Strangio, Kinetics of the
isobutane–isobutene–hydrogen system using tracers, Ind. Eng. Chem.
ˇ = aˇ 1.631 · Sc 0.37 · Ug Fundam. 12 (3) (1973) 263.
[6] A.G. Zwahlen, J.B. Agnew, Isobutane dehydrogenation kinetics determina-
The values of parameters are listed in Table B1. tion in a modified Berty gradientless reactor, Ind. Eng. Chem. Res. 31 (1992)
2088–2093.
Solution of the model equations: [7] S.M.K. Airksinen, M.E. Harlin, A.O.I. Krause, Kinetic modeling of dehydrogena-
To solve the differential equations of the mathematical model tion of isobutane on chromia/alumina catalysts, Ind. Eng. Chem. Res. 41 (2002)
the iteration technique to steady-state problem using the time 5619.
[8] S.M.K. Airksinen, A.O.I. Krause, Effect of catalyst prereduction on the dehydro-
step as a parameter was applied. This procedure is closely related genation of isobutane over chromia/alumina, Ind. Eng. Chem. Res. 44 (2005)
to method for solving the unsteady-state problem. Since the sec- 3862.
ond variable (t) was added the partial differential equations were [9] J. Gascón, C. Téllez, J. Herguido, M. Menéndez, Propane dehydrogenation over
a Cr2 O3 /Al2 O3 catalyst: transient kinetic modeling of propene and coke forma-
obtained. In order to solve the system of partial differential equa- tion, Appl. Catal. A: Gen. 248 (2003) 105–116.
tions the method of lines was used. As a result of applying the [10] H.T. Bi, N. Ellis, I.A. Abba, J.R. Grace, A state-of-the-art review of gas–solid
method of lines the large system of ODE was obtained. A special case turbulent fluidization, Chem. Eng. Sci. 55 (2000) 4789–4825.
[11] E.A. Novikov, Numerical methods for solution of differential equations in chem-
of a second-order Rosenbrock method and an automatic adjust-
ical kinetics, in: Mathematical Methods in Chemical Kinetics, Novosibirsk,
ment of the integration step algorithm were employed for solving Nauka, 1990, pp. 53–68 (in Russian).
this ODE system [11].

You might also like