Anda di halaman 1dari 7

Scientia Iranica C (2011) 18 (3), 458464

Sharif University of Technology


Scientia Iranica
Transactions C: Chemistry and Chemical Engineering
www.sciencedirect.com

Kinetic modeling of side reactions in propane dehydrogenation over


Pt-Sn/ -Al2 O3 catalyst
A. Farjoo a , F. Khorasheh a, , S. Niknaddaf a , M. Soltani b
a
b

Department of Chemical and Petroleum Engineering, Sharif University of Technology, Tehran, Iran
National Iranian Petrochemical Company, Research and Technology Division, Tehran, Iran

Received 2 January 2010; revised 17 June 2010; accepted 18 September 2010

KEYWORDS
Propane dehydrogenation;
Side reactions;
Platinum catalyst;
Kinetics.

Abstract The kinetics of side reactions in the dehydrogenation of propane over a supported platinum
catalyst modified by tin, were investigated. Catalytic dehydrogenation over a commercial Pt-Sn/ -Al2 O3
was carried out in a laboratory-scale plug-flow reactor at 580620 C under atmospheric pressure. Several
kinetic models derived from different reaction mechanisms were tested using experimental data obtained
under a range of reaction conditions. It was found that the kinetics of the main dehydrogenation reaction
was best described in terms of a LangmuirHinshelwood mechanism, where the adsorption of propane
was the rate controlling step. Simple power low rate expressions were used to express the kinetics of side
reactions.
2011 Sharif University of Technology. Production and hosting by Elsevier B.V. All rights reserved.

1. Introduction
The current chemical scenario shows an increased interest
in the production of light olefins as starting materials for some
of the most important chemical products including polymers,
synthetic rubbers and oxygenated compounds for reformulated
fuels [1,2]. Propylene is a major co-product from steam crackers
and the economics of the process are highly dependent on the
supply/demand situation for both propylene and ethylene. In
some geographic areas, propylene demand is increasing faster
than ethylene demand. The dehydrogenation of propane with
high selectivity to propylene gives the chance to de-couple the
productions of ethylene and propylene, and will be of great
interest in the coming years if the trend in the market for
polymers persists [1].

Corresponding author.
E-mail address: khorashe@sharif.edu (F. Khorasheh).

1026-3098 2011 Sharif University of Technology. Production and hosting by


Elsevier B.V. All rights reserved. Peer review under responsibility of Sharif
University of Technology.
doi:10.1016/j.scient.2011.05.009

The dehydrogenation of light alkanes is an important


reaction from an industrial point of view, since it is a selective
process to produce the corresponding short-chain alkenes
via direct catalytic dehydrogenation. Two types of catalyst
have been developed and patented for this reaction, including
chromia-based catalysts [35], and, more recently, supported
platinum catalysts [69]; both suffer from deactivation due
to coke deposition, and often a feed of hydrogen is used to
improve the catalyst stability. The addition of tin to platinum
catalysts has proved to be an effective way to reduce undesired
reactions and prevent rapid deactivation due to coke formation.
The platinumtin catalyst has therefore received considerable
attention [69].
Propane dehydrogenation is an endothermic process that
requires relatively high temperatures to obtain a high yield of
propylene, and is commercially carried out in a temperature
range of 450650 C under atmospheric pressure. The following
reactions occur during propane dehydrogenation [1013]:
C3 H8 C3 H6 + H2 ,

(R1)

C3 H8 CH4 + C2 H4 ,

(R2)

C2 H4 + H2 C2 H6 ,

(R3)

C3 H8 + H2 CH4 + C2 H6 .

(R4)

Reaction (R1) is the main reaction and reactions (R2)(R4)


are the possible side reactions. Reaction (R2) can take place

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

459

Table 1: Specifications of the commercial catalyst for propane dehydrogenation.


Diameter
Bulk Density (b )

1.8 mm
g
0.65 cm3

Catalyst Density (c )

1.12

Catalyst Surface Area (Sa)

200

g
cm3
m2
g

both in the gas phase (thermal conversion) and on the


surface of the catalyst (catalytic conversion). Thermal cracking
and catalytic cracking are both side reactions in propane
dehydrogenation, resulting in formation of methane and
ethylene as major side products. These reactions, however,
occur via different mechanisms. Thermal cracking involves a
free radical chain mechanism, while catalytic cracking proceeds
via surface species. Reaction (R4) is the hydrogenolysis
reaction which occurs on the surface of the catalyst. The side
reactions can be divided into two distinct categories; thermal
side reactions including thermal cracking, and catalytic side
reactions including propane hydrogenolysis, catalytic cracking
and ethylene hydrogenation. The objective of this study is to
investigate the kinetics of side reactions during the catalytic
dehydrogenation of propane over a commercial Pt-Sn/ -Al2 O3
catalyst.
The major side products were CH4 , C2 H4 and C2 H6 . The main
side product under the reaction conditions employed in this
study was CH4 , as it was produced via hydrogenolysis, thermal
cracking and catalytic cracking reactions.

Figure 1: The simplified sketch of propane dehydrogenation setup.

Table 2: Propane conversion for catalyst at different feed flow rates.


Run
1
2
3
4

Catalyst
weight (g)
1
1.2
1.5
1.6

QC3
(cc/min)
39.96
44.36
55.45
59.3

QH2
(cc/min)

WHSV
(1/h)

Conversion
(%)

31.968
35.488
44.36
47.44

4
4
4
4

20.45
23.07
23.32
23.35

2. Methods and materials


The extent of side reactions in propane dehydrogenation
were examined over a commercial Pt-Sn/ -Al2 O3 catalyst. A
simplified schematic of the laboratory scale setup used for the
propane dehydrogenation experiments is shown in Figure 1.
Using a three-lined setup, propane, hydrogen and nitrogen gas
mixtures could be introduced, whose composition and total
flow was set using Brooks mass flow controllers. The inside
diameter of the reactor was 12 mm, the length of the reactor
was 90 cm, and the catalyst was loaded in the middle section of
the reactor, in between two layers of quarts particles. The height
of the catalyst zone was approximately 2 cm. The specifications
of the catalyst are given in Table 1.
In each experiment, the catalyst sample was heated under
nitrogen flow from ambient temperature to 150 C in 25 min.
The temperature was then held constant for 3 h. The catalyst
was then heated from 150 C to the desired reduction
temperature of 450 C in 3 h under hydrogen flow. During this
period, nitrogen flow was decreased gradually, such that when
the catalyst temperature reached 450 C, the nitrogen flow was
zero. The temperature was then increased from 450 C to 530 C
in 16 min under hydrogen flow only. The temperature was then
held constant for 1 h at 530 C, and then reduced to 350 C in
3 h under hydrogen flow at which point the flow of propane
was started. The feed to the reactor consisted of propane and
hydrogen with H2 /propane molar ratio of 0.8. The temperature
was then increased to the desired reaction temperatures of
580 C, 600 C or 620 C at a rate of 5 C per minute. Product
gases were analyzed after a 90 min stabilization period. The
outlet stream from the reactor was analyzed by an online
Gas Chromatograph (model PERICHROM 2100 Packed column,
SS316, 6 m, 1/8 in, 28% DC200 on Chromsorb PAW 60/80, ENRO
3015).

3. Results and discussions


To obtain proper kinetic data for determination of intrinsic reaction rates, it was necessary to perform experiments in
the absence of external and internal mass transfer limitations.
Experiments were conducted using catalyst particles of different sizes to investigate internal mass transfer limitations. Feed
flow rates were also varied over a wide range, keeping WHSV
(Weight Hourly Space Velocity) constant to investigate external mass transfer limitations under reaction conditions.
3.1. External mass transfer limitations
In order to eliminate the external mass transfer limitations,
several experiments were conducted in which the size of the
catalyst pellet was kept constant and the feed flow rate was
varied over a wide range, while keeping WHSV constant at
4 h1 . If conversion was unaffected by varying the propane flow
rate at constant WHSV, it could be concluded that the reaction
was independent of the external mass transfer of the fluid
outside the catalyst pellet. The results of these experiments are
shown in Table 2, indicating that when propane flow rate, Qc3 ,
was greater than 44.36 cc/min, there were no external mass
transfer limitations.
3.2. Internal mass transfer limitations
In order to eliminate internal mass transfer limitations,
several experiments were performed using different sizes of
catalyst pellet at a feed flow rate where external mass transfer
limitations were absent. (Qc3 > 55.45 cc/min) The results
are presented in Table 3. If the conversion did not change by

460

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

Table 3: Propane conversion for catalyst pellets of different sizes.


Run

1
2
3
4
5

Catalyst
weight
(g)
0.75
0.75
0.75
0.75
0.75

QC3
Catalyst
(cc/min) diameter
(mm)

QH2
WHSV
(cc/min) (1/h)

Conversion
(%)

55
55
55
55
55

44
44
44
44
44

13.72
23.1
23.9
24.1
24.3

1.8
1
0.71
0.50
0.35

9
9
9
9
9

Figure 3: Effect of temperature on propylene selectivity.

Figure 2: Effect of temperature on propane conversion.

further reduction in the size of the catalyst pellet, there would


be no internal mass transfer limitations. Kinetic studies were
performed using catalyst pellets of 0.35 mm in diameter.
Figure 4: Effect of temperature on the formation of side products (WHSV =
8 (1/h)).

3.3. Effect of temperature on conversion and selectivities


Propane dehydrogenation is an endothermic reaction,
and increasing the temperature increases the equilibrium
conversion. Propane conversion increased with increasing
reaction temperature (Figure 2), but increasing the temperature
also caused the side reactions to become more significant,
thus decreasing propylene selectivity (Figure 3). Propylene
selectivity is defined by Eq. (1):
Selectivity to C3 H6 =

C3 H6 Produced
Input(C3 H8 ) Output(C3 H8 )

(1)

Figure 4 shows the effect of temperature on the formation


of side products. Temperature increase beyond 650 C is not
recommended, as it would result in a significant decrease in
propylene yield. To investigate the extent of thermal cracking
reactions, dehydrogenation was carried out in the absence of
the catalyst at 580 C, 600 C and 620 C, and results were
compared with those in the presence of the catalyst. The
amounts of side product formed via thermal side reactions and
catalytic side reactions are shown in Figure 5, indicating that
the catalytic reactions are more dominant and, with increasing
reaction temperature, the extent of both catalytic and thermal
reactions would increase.
3.4. Effect of residence time on conversion and selectivities
Temperature and WHSV are two parameters that affect
propane conversion. WHSV is the inverse of residence time
and as indicated by Figures 6 and 7, respectively, both propane
conversion and the formation of side products increase with
decreasing WHSV. For a given reaction temperature, the molar

Figure 5: Comparison of formation of thermal catalytic side products.

ratio of side products to all products at the reactor exit


was found to increase with decreasing WHSV (Figure 8),
resulting in a decrease in propylene selectivity with decreasing
WHSV (Figure 9). Propylene selectivities presented in Figures 3
and 9 indicated that the reaction temperature had a more
pronounced effect on propylene selectivity in comparison with
WHSV.
3.5. Kinetic modeling
Intrinsic kinetics of the main reaction in catalytic dehydrogenation of propane over the Pt-Sn/ -Al2 O3 catalyst can
be well represented in terms of LangmuirHinshelwood rate

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

461

Figure 9: Effect of residence time on propylene selectivity.

Figure 6: Effect of residence time on propane conversion.

main reaction in catalytic dehydrogenation of propane are


given in Table 4. The corresponding rate expressions given in
Table 5 are referred to in the form model I-a or model IVb, where the Roman numeral indicates the mechanism and
the letter (a or b) indicates the RCS, with a referring to the
adsorption of propane on active sites and b referring to the
surface reaction. It should be noted that competitive adsorption
of side products, including CH4 , C2 H4 and C2 H6 , were also
considered in the above mechanisms. Simple power law rate
expressions, Table 6, were used to describe the kinetics of the
side.
3.6. Estimation of kinetic parameters
Figure 7: Effect of residence time on formation of side products (T = 600 C).

The estimation of the kinetic parameters involved a two step


optimization procedure. The first step involved the estimation
of parameters appearing in the rate expression for the main
propane dehydrogenation reaction (as specified in Table 5).
Having evaluated the parameters for the main reaction, the
second step involved the estimation of kinetic constants for
the side reactions (as specified in Table 6). This two-step
procedure was necessary, since the side products were present
in very small amounts compared with propane and propylene
as major species. The governing differential equations for the
isothermal plug flow reactor employed in the experiments
are:
dFi
dW

ri ,

i = C3 H8 , C3 H6 , C2 H4 , C2 H6 , CH4 ,
Figure 8: Molar ratio of side products to all products at different WHSV and
reaction temperatures.

expressions [13], where either an adsorption, surface reaction, or desorption step is considered as the rate controlling
step. Based on experimental data, Airaksinen et al. (2002) indicated that desorption of propylene from the surface sites was
not likely to be the rate controlling step and suggested several mechanisms for propane dehydrogenation, where either
adsorption or surface reaction was rate controlling [14]. A recent study, where the same commercial catalyst as that used in
the current investigation was employed, suggested that mechanisms involving propane adsorption either on single or dual
sites as the Rate Controlling Step (RCS) were most consistent
with experimental data [15].
The LangmuirHinshelwood mechanisms that were studied
in the current investigation to describe the kinetics of the

(2)

where Fi is the molar flow rate of component i, W is the mass


of catalyst, ri is the reaction rate for component i, and the
summation covers all reactions j leading to the formation and/or
disappearance of component i. Experimental data for each run
included the inlet molar flow rate of propane and the exit
molar flow rates of propane, propylene, methane, ethylene
and ethane. Hydrogen flow rates were obtained by difference
from an overall material balance. Numerical integration of
the system of Eqs. (2) was carried out by the ode15s, and
the optimization procedure was performed with the aid of
the Genetic Algorithm function of MATLAB. For each reaction
temperature, the optimized kinetic parameters were obtained
by minimizing the following objective function:
OF =

4
5

k=1 i=1

(FiEexp FiEpre )2 Wi ,

(3)

462

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

Table 4: Mechanisms for propane dehydrogenation.


Mechanism I

Mechanism II

C3 H8 (g) + 2S C3 H7 S + H S (RCS)
C3 H7 S + S C3 H6 S + H S
C3 H6 S C3 H6 (g) + S
2H S H2 (g) + S
CH4 + S CH4 S
C2 H4 + S C2 H4 S
C2 H6 + S C2 H6 S
S: Active sites on the catalyst

C3 H8 (g) + S C3 H7 S H (RCS)
C3 H7 S H + S C3 H6 S H + H S
C3 H6 S H C3 H6 (g) + H S
2H S H2 (g) + 2S
CH4 + S CH4 S
C2 H6 + S C2 H6 S
C2 H4 + S C2 H4 S
S: Active site on the catalyst

Mechanism III

Mechanism IV

C3 H8 (g) + S C3 H7 S H (RCS)
C3 H7 S H C3 H6 (g) + H S H
H S H H2 (g) + S
CH4 + S CH4 S
C2 H4 + S C2 H4 S
C2 H6 + S C2 H6 S
S: Active sites on the catalyst

C3 H8 (g) + S C3 H8 S
C3 H8 S + S C3 H7 S + H S (RCS)
C3 H7 S + S C3 H6 S + H S
C3 H6 S C3 H6 (g) + S
2H S H2 (g) + 2S
CH4 + S CH4 S
C2 H4 + S C2 H4 S
C2 H6 + S C2 H6 S
S: Active site on the catalyst

Table 5: LangmuirHinshelwood rate expressions for the main propane dehydrogenation reaction.
Model

Equation rate

PPr PH
k PP K 2

1 (I-a)

rA =

2 (II-a)

rA =

3 (III-a)

rA =

4 (IV-b)

rA =

P: Propane (A)
Pr: Propylene (B)
H2 = Hydrogen (C)

k : Rate constant
Keq : Equilibrium constant
K : Adsorption constant

exp

1/2

(1+K PPr PH +KPr PPr +(K PH2 )1/2 +KH2 PH2 +KCH4 PCH4 +KC2 H4 PC2 H4 +KC2 H6 PC2 H6 )2
2

PPr PH
k Pp K 2
eq

1/2
1+KPr H PPr PH +(KH PH )1/2 +K PPr PH +KCH PCH +KC H PC H +KC H PC H
2 6 2 6
2 4 2 4
4
4
2
2
2
2
2

P P
k P B C
A

k Pp

(1+Kp Pp +KPr PPr +(KH2 PH2 )1/2 +K PPr

PPr PH
2
Keq

PH +KCH PCH +KC H PC H +KC H PC H )2


2 6 2 6
2 4 2 4
4
4
2

pre

Wi = Errori /Total error,

(4)

exp
(FiE FiEpre )2 ,

(5)

k=1

Total error =

Keq

1+KPr H PC H PH +KH PH +KCH PCH +KC H PC H +KC H PC H


2 3 6
2
2
2
4 4
2 4
2 4
2 6 2 6

where FiE and FiE are the experimental and predicted molar
flow rates of component i at the reactor exit, Wi are weighting factors, and summation k extends over experimental data at
different WHSV. The optimized parameter values were highly
influenced by the choice of weighting factors, as propane and
propylene were major species, while methane, ethane and
ethylene were present in minor quantities. In the optimization
procedure for parameter estimation, first a preliminary estimate for the kinetic parameter was obtained using a weighing factor of one for each component. Subsequently, the revised
weighting factors were evaluated as follows:

Errori =

eq

Errori .

(6)

i=1

Table 6: Proposed reaction rate equations for side reactions.


Side reaction
Catalytic cracking (I)
Ethylene hydrogenation (II)
Hydrogenolysis (III)
Thermal cracking (IV)

Kinetic equation
r
r
r
r

= k1 PC3 H8
= k2 PC2 H4 PH2
= k2 PC3 H8 PH2
= k4 PC3 H8

The constants are reported in Table 7 for the main reaction using
different reactions [1012]. Proposed LangmuirHinshelwood
rate expressions. Having evaluated the parameters for the main
reaction, the second step involved the estimation of the rate
expressions of the optimized kinetic parameters for the main
reaction, and the rate constants for the side reactions using the
following objective function:
OF =

4
3

(FiEexp FiEpre )2 ,

(9)

k=1 i=1

The normalized weighting factors were then used to optimize


the kinetic parameters of the main reaction for each reaction
temperature. The temperature dependence of the rate constants and equilibrium adsorption constants were determined
by the Arrhenius and vant Hoff equations, respectively, as
follows:
ln(ki ) = ln(ki0 ) + (Ei /RT ),

(7)

ln(Ki ) = ln(Ki0 ) + (1Hi /RT ).

(8)

where summation i extends for side products only. The


optimized rate constants for each of the side reactions were
obtained for different temperatures and the corresponding
Arrhenius parameters are reported in Table 8. It should be noted
that the Arrhenius parameters for the rate constants of the
thermal cracking side reaction were obtained from separate
experimental data at different temperatures, under similar
conditions, but in the absence of the catalyst.

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

463

Table 7: Kinetic constants for the proposed LangmuirHinshelwood rate expressions, E (kJ/mol), 1H (kJ/mol).
Model 1

Model 2

.79
k = 3.17 10 exp 61
RT
KPr = 7.44 107 exp 89RT.07
.36
KH2 = 2.18 101 exp 9RT

KCH4 = 9.07 103 exp 16RT.06

5
.
75
KC2 H6 = 3.60 102 exp RT
16.44
3
KC2 H4 = 6.97 10 exp RT

K = 3.44 102 exp 12RT.63


.97
K = 1.02 10+4 exp 49
RT
3

Model 3

.71
k = 1.21 10 exp 53
RT

KH2 = 1.61 103 exp 36RT.03

KCH4 = 1.37 105 exp 52RT.63

KC2 H6 = 1.34 106 exp 64RT.39

42
.
84
KC2 H4 = 5.15 105 exp RT
7.41
2
KPr H2 = 3.70 10 exp RT
9.16
K = 3.47 101 exp RT
3

Model 4

.39
k = 3.71 10 exp 59
RT
KH2 = 9.71 105 exp 52RT.22

KCH4 = 1.70 103 exp 21RT.28

KC2 H6 = 1.85 104 exp 38RT.23


4.17
5
KC2 H4 = 7.22 10 exp RT

KPr H2 = 1.67 102 exp 18RT.35

.66
k = 4.25 101 exp 70
RT

18.92
RT
KPr = 1.42 105 exp 69RT.31

KH2 = 2.01 104 exp 51RT.24

6
.
83
KCH4 = 1.98 102 exp RT
5.48
2
KC2 H6 = 2.99 10 exp RT
.61
KC2 H4 = 3.39 102 exp 5RT

K = 7.73 104 exp 68RT.46


Kp = 6.70 103 exp

Table 8: Rate constants for proposed rate expressions for side reactions.
Model 1
k1 = 1.14 exp

Model 2
K1 = 7.53 10 exp

6488

k2 = 1.34 107 exp


k3 = 3.36 104 exp
k4 = 1.53E + 09 exp

k2 = 9.62 107 exp

27393
T
14723

T
1598.23
T

k3 = 5.57 105 exp

k4 = 1.53E + 09 exp

Model 3

273938
T
28187

exp

k2 = 3.94 1011 exp

T
17926

T
1598.23
T

k1 = 3.36 10

17

k3 = 1.47 101 exp

k4 = 1.53E + 09 exp

Model 4

40451
T
30754

T
8498

T
1598.23
T

k1 = 3.09 104 exp 16661


T

k2 = 1.25 109 exp 25692


T

.6
k3 = 2.85 109 exp 5757
T

.23
k4 = 1.53E + 09 exp 1598
T

Table 9: Average % error between predicted and experimental flow rates at


reactor exit.
Model

Absolute error %

Model 1

Main species
Side products

4.02
14.68

Model 2

Main species
Side products

6.22
21.74

Model 3

Main species
Side products

7.11
17.74

Model 4

Main species
Side products

12.18
25.98

The kinetic constants reported in Table 7 for the main reaction and in Table 8 for the side reactions were used to predict
the exit molar flow rates of various spices under different reaction conditions. Comparison between the experimental values
and predicted values from each model would serve as a measure
to differentiate between the predictive ability of different models. Table 9 presents the average percent absolute error for each
model for both major species (propane and propylene) and side
products, indicating that model 1, where dual site adsorption
of propane is considered as the rate controlling step, provides
the best agreement with experimental data. A typical comparison of experimental exit molar flow rates, with predicted values
from model 1 as a function of residence time at 580 C, is presented in Figures 10 and 11 for the major species and the side
products, respectively.

Figure 10: Comparison between experimental exit molar flow rates and
predicted values from model 1 for major products at reaction temperature of
580 C.

4. Conclusions
The kinetics of propane dehydrogenation over a commercial
Pt-Sn/ -Al2 O3 catalyst was best described in terms of a LangmuirHinshelwood mechanism, where dual site adsorption of
propane was considered as the rate controlling step. The kinetics of side reactions leading to the formation of methane,
ethylene and ethane as minor products could reasonably be described by simple power-law rate expressions. The proposed kinetic model resulted in an average relative error of 4.0% and
14.7% between predicted and experimental molar flow rates

Figure 11: Comparison between experimental exit molar flow rates and
predicted values from model 1 for side products at reaction temperature of
580 C.

of major species and side products, respectively, at the reactor


exit, over the range of experimental conditions employed in this
study.

464

A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458464

Acknowledgments
The authors acknowledge financial support from the Iranian
National Petrochemical Research and Technology Company for
the experimental work involved in this study.
References
[1] Miracca, I. and Piovesan, L. Light paraffin dehydrogenation in a fluidized
bed reactor, Catalysis Today, 52, pp. 259269 (1999).
[2] Loc, L.C., Gaidai, N.A., Kiperman, S.L., Thoang, H.S. and Novikov, P.B.
Kinetics of side reactions in the dehydrogenation of n-butane and ISObutane over platinumpotassium catalyst, Kinetics and Catalysis, 36(4),
pp. 504510 (1995).
[3] Pop, E., Goidean, N., Giodean, D. and Serban, G. DE 2401955 (17 July 1975).
[4] Schramm, B., Kern, J., Schwahn, H., Preuss, A.W., Gottlieb, K. and
Bruderreck, H. DE 3739002 A1 (24 May 1989).
[5] Kirner, J.F. GB 2162082 A1 (29 January 1986).
[6] Box, E. US 3692701 (19 September 1972).
[7] Wilhelm, F.C. US 3998900 (21 December 1976).
[8] Barri, S.A.I. and Tahir, R. EP 351066 A1 (17 January 1990).
[9] Cottrell, P.R. and Fettis, M.E. US 5087792 (11 February 1992).
[10] Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y. and Wu, P. Propane
dehydrogenation on PtSn/ZSM-5 catalyst: effect of tin as a promoter,
Catalysis Communications, 7, pp. 860866 (2006).
[11] Bobrov, V.S., Digurov, N.G. and Skudin, V.V. Propane dehydrogenation
using catalytic membrane, Journal of Membrane Science, 253, pp. 233242
(2005).
[12] Guo, J., Lou, H., Zhao, H., Zheng, L. and Zheng, X. Dehydrogenation
and aromatization of propane over rhenium-modified HZSM-5 catalyst,
Journal of Molecular Catalysis A: Chemical, 239, pp. 222227 (2005).

[13] Fogler, H.S., Elements of Chemical Reaction Engineering, 2nd ed., p. 581.
Prentice Hall International, New York, US (1999).
[14] Airaksinen, S.M.K., Harlin, M.E. and Krause, A.O.I. Kinetic modeling of
dehydrogenation of isobutane on chromia/alumina catalyst, Industrial &
Engineering Chemistry Research, 41, pp. 56195627 (2002).
[15] Mohagheghi, M. and Bakeri, G. Kinetic studies of propane dehydrogenation over Pt-Sn/ -Al2 O3 catalyst, in: Presented at International Conference
on Chemical Reactors, CHEMREACTOR-18, Malta (29 September3 October
2008).
Afrooz Farjoo was born in 1985. She obtained her B.S. degree from Isfahan
University of Technology in 2007 and her M.S. degree in 2009 from the
Department of Petroleum and Chemical Engineering at Sharif University of
Technology in the area of Thermo Kinetics and Catalysis.
Farhad Khorasheh was born in 1961. He took his B.S. degree in Chemical
Engineering in 1983 from Queens University, Ontario, Canada and his M.S.
and Ph.D. degrees in Chemical Engineering, in 1986 and 1992, respectively,
from the University of Alberta, Edmonton, Canada. He is now a Professor in
the Department of Petroleum and Chemical Engineering at Sharif University of
Technology. His research interests include Modeling and Reactor Design.
Saeed Niknaddaf was born in 1985. He obtained his B.S. and M.S. degrees from
the Department of Petroleum and Chemical Engineering at Sharif University of
Technology in 2007 and 2009, respectively, in the area of Thermo Kinetics and
Catalysis.
Mahnaz Soltani was born in 1974. She obtained her B.S. and M.S. degrees from
the Department of Chemical Engineering at Tehran University in 1996 and 1998,
respectively. She is now working as a Chemical Engineer in the National Iranian
Petrochemical Company.

Anda mungkin juga menyukai