You are on page 1of 7

Scientia Iranica C (2011) 18 (3), 458–464

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/γ -Al2 O3 catalyst
A. Farjoo a , F. Khorasheh a,∗ , S. Niknaddaf a , M. Soltani b
a
Department of Chemical and Petroleum Engineering, Sharif University of Technology, Tehran, Iran
b
National Iranian Petrochemical Company, Research and Technology Division, Tehran, Iran

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

KEYWORDS Abstract The kinetics of side reactions in the dehydrogenation of propane over a supported platinum
Propane dehydrogenation; catalyst modified by tin, were investigated. Catalytic dehydrogenation over a commercial Pt-Sn/γ -Al2 O3
Side reactions; was carried out in a laboratory-scale plug-flow reactor at 580–620 °C under atmospheric pressure. Several
Platinum catalyst; kinetic models derived from different reaction mechanisms were tested using experimental data obtained
Kinetics. under a range of reaction conditions. It was found that the kinetics of the main dehydrogenation reaction
was best described in terms of a Langmuir–Hinshelwood 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.
Open access under CC BY license.

1. Introduction The dehydrogenation of light alkanes is an important


reaction from an industrial point of view, since it is a selective
The current chemical scenario shows an increased interest process to produce the corresponding short-chain alkenes
in the production of light olefins as starting materials for some via direct catalytic dehydrogenation. Two types of catalyst
of the most important chemical products including polymers, have been developed and patented for this reaction, including
synthetic rubbers and oxygenated compounds for reformulated chromia-based catalysts [3–5], and, more recently, supported
fuels [1,2]. Propylene is a major co-product from steam crackers platinum catalysts [6–9]; both suffer from deactivation due
and the economics of the process are highly dependent on the to coke deposition, and often a feed of hydrogen is used to
supply/demand situation for both propylene and ethylene. In improve the catalyst stability. The addition of tin to platinum
some geographic areas, propylene demand is increasing faster catalysts has proved to be an effective way to reduce undesired
than ethylene demand. The dehydrogenation of propane with reactions and prevent rapid deactivation due to coke formation.
high selectivity to propylene gives the chance to de-couple the The platinum–tin catalyst has therefore received considerable
productions of ethylene and propylene, and will be of great attention [6–9].
interest in the coming years if the trend in the market for Propane dehydrogenation is an endothermic process that
polymers persists [1]. requires relatively high temperatures to obtain a high yield of
propylene, and is commercially carried out in a temperature
∗ range of 450–650 °C under atmospheric pressure. The following
Corresponding author.
E-mail address: khorashe@sharif.edu (F. Khorasheh).
reactions occur during propane dehydrogenation [10–13]:
C3 H8 ↔ C3 H6 + H2 , (R1)
1026-3098 © 2011 Sharif University of Technology. Production and hosting by
Elsevier B.V. Open access under CC BY license.
Peer review under responsibility of Sharif University of Technology. C3 H8 → CH4 + C2 H4 , (R2)
doi:10.1016/j.scient.2011.05.009
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) 458–464 459

Table 1: Specifications of the commercial catalyst for propane dehydro-


genation.

Diameter 1.8 mm
Bulk Density (ρb ) 0.65 cm3
g

Catalyst Density (ρc ) 1.12


g
cm3
m2
Catalyst Surface Area (Sa) 200 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
Figure 1: The simplified sketch of propane dehydrogenation setup.
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/γ -Al2 O3 Table 2: Propane conversion for catalyst at different feed flow rates.
catalyst. Run Catalyst QC3 QH2 WHSV Conversion
The major side products were CH4 , C2 H4 and C2 H6 . The main weight (g) (cc/min) (cc/min) (1/h) (%)
side product under the reaction conditions employed in this 1 1 39.96 31.968 4 20.45
study was CH4 , as it was produced via hydrogenolysis, thermal 2 1.2 44.36 35.488 4 23.07
cracking and catalytic cracking reactions. 3 1.5 55.45 44.36 4 23.32
4 1.6 59.3 47.44 4 23.35

2. Methods and materials

The extent of side reactions in propane dehydrogenation 3. Results and discussions


were examined over a commercial Pt-Sn/γ -Al2 O3 catalyst. A
simplified schematic of the laboratory scale setup used for the To obtain proper kinetic data for determination of intrin-
propane dehydrogenation experiments is shown in Figure 1. sic reaction rates, it was necessary to perform experiments in
Using a three-lined setup, propane, hydrogen and nitrogen gas the absence of external and internal mass transfer limitations.
mixtures could be introduced, whose composition and total Experiments were conducted using catalyst particles of differ-
flow was set using Brooks mass flow controllers. The inside ent sizes to investigate internal mass transfer limitations. Feed
diameter of the reactor was 12 mm, the length of the reactor flow rates were also varied over a wide range, keeping WHSV
was 90 cm, and the catalyst was loaded in the middle section of (Weight Hourly Space Velocity) constant to investigate exter-
the reactor, in between two layers of quarts particles. The height nal mass transfer limitations under reaction conditions.
of the catalyst zone was approximately 2 cm. The specifications
of the catalyst are given in Table 1. 3.1. External mass transfer limitations
In each experiment, the catalyst sample was heated under
nitrogen flow from ambient temperature to 150 °C in 25 min. In order to eliminate the external mass transfer limitations,
The temperature was then held constant for 3 h. The catalyst several experiments were conducted in which the size of the
was then heated from 150 °C to the desired reduction catalyst pellet was kept constant and the feed flow rate was
temperature of 450 °C in 3 h under hydrogen flow. During this varied over a wide range, while keeping WHSV constant at
period, nitrogen flow was decreased gradually, such that when 4 h−1 . If conversion was unaffected by varying the propane flow
the catalyst temperature reached 450 °C, the nitrogen flow was rate at constant WHSV, it could be concluded that the reaction
zero. The temperature was then increased from 450 °C to 530 °C was independent of the external mass transfer of the fluid
in 16 min under hydrogen flow only. The temperature was then outside the catalyst pellet. The results of these experiments are
held constant for 1 h at 530 °C, and then reduced to 350 °C in shown in Table 2, indicating that when propane flow rate, Qc3 ,
3 h under hydrogen flow at which point the flow of propane was greater than 44.36 cc/min, there were no external mass
was started. The feed to the reactor consisted of propane and transfer limitations.
hydrogen with H2 /propane molar ratio of 0.8. The temperature
was then increased to the desired reaction temperatures of 3.2. Internal mass transfer limitations
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 In order to eliminate internal mass transfer limitations,
outlet stream from the reactor was analyzed by an online several experiments were performed using different sizes of
Gas Chromatograph (model PERICHROM 2100 Packed column, catalyst pellet at a feed flow rate where external mass transfer
SS316, 6 m, 1/8 in, 28% DC200 on Chromsorb PAW 60/80, ENRO limitations were absent. (Qc3 > 55.45 cc/min) The results
3015). 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) 458–464

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

Run Catalyst QC3 Catalyst QH2 WHSV Conversion


weight (cc/min) diameter (cc/min) (1/h) (%)
(g) (mm)

1 0.75 55 1.8 44 9 13.72


2 0.75 55 1 44 9 23.1
3 0.75 55 0.71 44 9 23.9
4 0.75 55 0.50 44 9 24.1
5 0.75 55 0.35 44 9 24.3

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 =


3.3. Effect of temperature on conversion and selectivities
8 (1/h)).

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):

C3 H6 Produced
Selectivity to C3 H6 = . (1)
Input(C3 H8 ) − Output(C3 H8 )
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 Figure 5: Comparison of formation of thermal catalytic side products.

compared with those in the presence of the catalyst. The


amounts of side product formed via thermal side reactions and ratio of side products to all products at the reactor exit
catalytic side reactions are shown in Figure 5, indicating that was found to increase with decreasing WHSV (Figure 8),
the catalytic reactions are more dominant and, with increasing resulting in a decrease in propylene selectivity with decreasing
reaction temperature, the extent of both catalytic and thermal WHSV (Figure 9). Propylene selectivities presented in Figures 3
reactions would increase. and 9 indicated that the reaction temperature had a more
pronounced effect on propylene selectivity in comparison with
3.4. Effect of residence time on conversion and selectivities WHSV.

Temperature and WHSV are two parameters that affect 3.5. Kinetic modeling
propane conversion. WHSV is the inverse of residence time
and as indicated by Figures 6 and 7, respectively, both propane Intrinsic kinetics of the main reaction in catalytic dehy-
conversion and the formation of side products increase with drogenation of propane over the Pt-Sn/γ -Al2 O3 catalyst can
decreasing WHSV. For a given reaction temperature, the molar be well represented in terms of Langmuir–Hinshelwood rate
A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458–464 461

Figure 6: Effect of residence time on propane conversion. Figure 9: Effect of residence time on propylene selectivity.

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 IV-
b’’, 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 −
= −ri ,
dW j

i = C3 H8 , C3 H6 , C2 H4 , C2 H6 , CH4 , (2)
Figure 8: Molar ratio of side products to all products at different WHSV and where Fi is the molar flow rate of component i, W is the mass
reaction temperatures.
of catalyst, −ri is the reaction rate for component i, and the
summation covers all reactions j leading to the formation and/or
expressions [13], where either an adsorption, surface reac-
disappearance of component i. Experimental data for each run
tion, or desorption step is considered as the rate controlling
included the inlet molar flow rate of propane and the exit
step. Based on experimental data, Airaksinen et al. (2002) indi-
molar flow rates of propane, propylene, methane, ethylene
cated that desorption of propylene from the surface sites was
and ethane. Hydrogen flow rates were obtained by difference
not likely to be the rate controlling step and suggested sev-
from an overall material balance. Numerical integration of
eral mechanisms for propane dehydrogenation, where either
the system of Eqs. (2) was carried out by the ode15s, and
adsorption or surface reaction was rate controlling [14]. A re-
the optimization procedure was performed with the aid of
cent study, where the same commercial catalyst as that used in
the current investigation was employed, suggested that mech- the Genetic Algorithm function of MATLAB. For each reaction
anisms involving propane adsorption either on single or dual temperature, the optimized kinetic parameters were obtained
sites as the Rate Controlling Step (RCS) were most consistent by minimizing the following objective function:
with experimental data [15]. 4 −
5
The Langmuir–Hinshelwood mechanisms that were studied

OF = (FiEexp − FiEpre )2 Wi , (3)
in the current investigation to describe the kinetics of the k=1 i=1
462 A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458–464

Table 4: Mechanisms for propane dehydrogenation.

Mechanism I Mechanism II

C3 H8 (g) + 2S ←→ C3 H7 · · · S + H · · · S (RCS) C3 H8 (g) + S ←→ C3 H7 · · · S · · · H (RCS)


C3 H7 · · · S + S ←→ C3 H6 · · · S + H · · · S C3 H7 · · · S · · · H + S ←→ C3 H6 · · · S · · · H + H · · · S
C3 H6 · · · S ←→ C3 H6 (g) + S C3 H6 · · · S · · · H ←→ C3 H6 (g) + H · · · S
2H · · · S ←→ H2 (g) + S 2H · · · S ←→ H2 (g) + 2S
CH4 + S ←→CH4 · · · S CH4 + S ←→ CH4 · · · S
C2 H4 + S ←→ C2 H4 · · · S C2 H6 + S ←→ C2 H6 · · · S
C2 H6 + S ←→ C2 H6 · · · S C2 H4 + S ←→ C2 H4 · · · S
S: Active sites on the catalyst S: Active site on the catalyst
Mechanism III Mechanism IV
C3 H8 (g) + S ←→ C3 H7 · · · S · · · H (RCS) C3 H8 (g) + S ←→ C3 H8 · · · S
C3 H7 · · · S · · · H ←→ C3 H6 (g) + H · · · S · · · H C3 H8 · · · S + S ←→ C3 H7 · · · S + H · · · S (RCS)
H · · · S · · · H ←→ H2 (g) + S C3 H7 · · · S + S ←→ C3 H6 · · · S + H · · · S
CH4 + S ←→ CH4 · · · S C3 H6 · · · S ←→ C3 H6 (g) + S
C2 H4 + S ←→ C2 H4 · · · S 2H · · · S ←→ H2 (g) + 2S
C2 H6 + S ←→ C2 H6 · · · S CH4 + S ←→ CH4 · · · S
S: Active sites on the catalyst C2 H4 + S ←→ C2 H4 · · · S
C2 H6 + S ←→ C2 H6 · · · S
S: Active site on the catalyst

Table 5: Langmuir–Hinshelwood rate expressions for the main propane dehydrogenation reaction.

Model Equation rate


 
PPr PH
k′ PP − K 2
′ eq
1 (I-a) −rA = 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
2 (II-a) −rA′ = 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 2 2 2  2 4 4 2 4 2 4 2 6 2 6
P P
k′ P − B C
A Keq
3 (III-a) −rA′ = 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
PPr PH
k′ Pp − Keq
2

4 (IV-b) −rA′ = (1+Kp Pp +KPr PPr +(KH2 PH2 )1/2 +K ′ PPr



PH +KCH PCH +KC H PC H +KC H PC H )2
2 4 4 2 4 2 4 2 6 2 6
P: Propane (A) k′ : Rate constant
Pr: Propylene (B) Keq : Equilibrium constant
H2 = Hydrogen (C) K : Adsorption constant

exp pre
where FiE and FiE are the experimental and predicted molar Table 6: Proposed reaction rate equations for side reactions.
flow rates of component i at the reactor exit, Wi are weight-
Side reaction Kinetic equation
ing factors, and summation k extends over experimental data at
different WHSV. The optimized parameter values were highly Catalytic cracking (I) r = k1 PC3 H8
influenced by the choice of weighting factors, as propane and Ethylene hydrogenation (II) r = k2 PC2 H4 PH2
Hydrogenolysis (III) r = k2 PC3 H8 PH2
propylene were major species, while methane, ethane and Thermal cracking (IV) r = k4 PC3 H8
ethylene were present in minor quantities. In the optimization
procedure for parameter estimation, first a preliminary esti-
mate for the kinetic parameter was obtained using a weigh-
ing factor of one for each component. Subsequently, the revised The constants are reported in Table 7 for the main reaction using
weighting factors were evaluated as follows: different reactions [10–12]. Proposed Langmuir–Hinshelwood
rate expressions. Having evaluated the parameters for the main
Wi = Errori /Total error, (4) reaction, the second step involved the estimation of the rate
4
− exp expressions of the optimized kinetic parameters for the main
Errori = (FiE − FiEpre )2 , (5) reaction, and the rate constants for the side reactions using the
k=1 following objective function:
5
− 4 −
3
Total error = Errori . (6)

OF = (FiEexp − FiEpre )2 , (9)
i=1
k=1 i=1
The normalized weighting factors were then used to optimize
where summation i extends for side products only. The
the kinetic parameters of the main reaction for each reaction
temperature. The temperature dependence of the rate con- optimized rate constants for each of the side reactions were
stants and equilibrium adsorption constants were determined obtained for different temperatures and the corresponding
by the Arrhenius and van’t Hoff equations, respectively, as Arrhenius parameters are reported in Table 8. It should be noted
follows: that the Arrhenius parameters for the rate constants of the
thermal cracking side reaction were obtained from separate
ln(ki ) = ln(ki0 ) + (−Ei /RT ), (7) experimental data at different temperatures, under similar
ln(Ki ) = ln(Ki0 ) + (−1Hi /RT ). (8) conditions, but in the absence of the catalyst.
A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458–464 463

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

Model 1 Model 2 Model 3 Model 4


 .79   .71   .39   .66 
k = 3.17 × 10 exp −61
′ 3
k = 1.21 × 10 exp −53
′ 3
k = 3.71 × 10 exp −59
′ 3
k = 4.25 × 101 exp −70

RT  RT
 RT  RT
KPr = 7.44 × 10−7 exp 89RT.07 KH2 = 1.61 × 10−3 exp 36RT.03 KH2 = 9.71 × 105 exp 52RT.22
 18.92 
Kp = 6.70 × 10−3 exp
 
 .36   RT 
KH2 = 2.18 × 10−1 exp 9RT KCH4 = 1.37 × 10−5 exp 52RT.63 KCH4 = 1.70 × 103 exp 21RT.28 KPr = 1.42 × 10−5 exp 69RT.31
   

KCH4 = 9.07 × 10−3 exp 16RT.06 KC2 H6 = 1.34 × 10−6 exp 64RT.39 KC2 H6 = 1.85 × 104 exp 38RT.23 KH2 = 2.01 × 10−4 exp 51RT.24
       
. .
 4.17  .
KC2 H6 = 3.60 × 10−2 exp RT 5 75
KC2 H4 = 5.15 × 10−5 exp RT 42 84
KC2 H4 = 7.22 × 10 exp RT
5
KCH4 = 1.98 × 10−2 exp RT 6 83
     
 16.44   7.41 
KPr H2 = 1.67 × 10−2 exp 18RT.35
 5.48 
KC2 H4 = 6.97 × 10 exp RT KPr H2 = 3.70 × 10 exp RT KC2 H6 = 2.99 × 10 exp RT
−3 −2
  −2

K ′ = 3.44 × 10−2 exp 12RT.63


 9.16   .61 
K ′ = 3.47 × 10−1 exp −RT KC2 H4 = 3.39 × 10−2 exp 5RT
 

 .97 
K ′′′ = 1.02 × 10+4 exp −49 K ′ = 7.73 × 104 exp 68RT.46
 
RT

Table 8: Rate constants for proposed rate expressions for side reactions.

Model 1 Model 2 Model 3 Model 4

k1 = 1.14 exp K1 = 7.53 × 10 exp k1 = 3.36 × 10 k1 = 3.09 × 104 exp −16661


 −6488  8
 −273938  17
 −40451   
T T
exp T T

k2 = 1.34 × 107 exp k2 = 9.62 × 107 exp k2 = 3.94 × 1011 exp k2 = 1.25 × 109 exp −25692
 −27393   −28187   −30754   
T T T T
.6
k3 = 3.36 × 104 exp k3 = 5.57 × 105 exp k3 = 1.47 × 101 exp k3 = 2.85 × 109 exp −5757
 −14723   −17926   −8498   
T T T T
−1598.23 −1598.23 −1598.23 .23
k4 = 1.53E + 09 exp k4 = 1.53E + 09 exp k4 = 1.53E + 09 exp k4 = 1.53E + 09 exp −1598
       
T T T T

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


reactor exit.
Model Absolute error %

Main species 4.02


Model 1
Side products 14.68
Main species 6.22
Model 2
Side products 21.74
Main species 7.11
Model 3
Side products 17.74
Main species 12.18
Model 4
Side products 25.98

The kinetic constants reported in Table 7 for the main reac-


tion and in Table 8 for the side reactions were used to predict
Figure 10: Comparison between experimental exit molar flow rates and
the exit molar flow rates of various spices under different reac- predicted values from model 1 for major products at reaction temperature of
tion conditions. Comparison between the experimental values 580 °C.
and predicted values from each model would serve as a measure
to differentiate between the predictive ability of different mod-
els. 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 compari-
son of experimental exit molar flow rates, with predicted values
from model 1 as a function of residence time at 580 °C, is pre-
sented in Figures 10 and 11 for the major species and the side
products, respectively.

4. Conclusions

The kinetics of propane dehydrogenation over a commercial


Pt-Sn/γ -Al2 O3 catalyst was best described in terms of a Lang-
muir–Hinshelwood mechanism, where dual site adsorption of Figure 11: Comparison between experimental exit molar flow rates and
predicted values from model 1 for side products at reaction temperature of
propane was considered as the rate controlling step. The ki-
580 °C.
netics of side reactions leading to the formation of methane,
ethylene and ethane as minor products could reasonably be de-
scribed by simple power-law rate expressions. The proposed ki- of major species and side products, respectively, at the reactor
netic model resulted in an average relative error of 4.0% and exit, over the range of experimental conditions employed in this
14.7% between predicted and experimental molar flow rates study.
464 A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458–464

Acknowledgments [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
The authors acknowledge financial support from the Iranian dehydrogenation of isobutane on chromia/alumina catalyst’’, Industrial &
National Petrochemical Research and Technology Company for Engineering Chemistry Research, 41, pp. 5619–5627 (2002).
the experimental work involved in this study. [15] Mohagheghi, M. and Bakeri, G. ‘‘Kinetic studies of propane dehydrogena-
tion over Pt-Sn/γ -Al2 O3 catalyst’’, in: Presented at International Conference
on Chemical Reactors, CHEMREACTOR-18, Malta (29 September–3 October
References 2008).

[1] Miracca, I. and Piovesan, L. ‘‘Light paraffin dehydrogenation in a fluidized Afrooz Farjoo was born in 1985. She obtained her B.S. degree from Isfahan
bed reactor’’, Catalysis Today, 52, pp. 259–269 (1999). University of Technology in 2007 and her M.S. degree in 2009 from the
[2] Loc, L.C., Gaidai, N.A., Kiperman, S.L., Thoang, H.S. and Novikov, P.B. Department of Petroleum and Chemical Engineering at Sharif University of
‘‘Kinetics of side reactions in the dehydrogenation of n-butane and ISO- Technology in the area of Thermo Kinetics and Catalysis.
butane over platinum–potassium catalyst’’, Kinetics and Catalysis, 36(4),
pp. 504–510 (1995).
[3] Pop, E., Goidean, N., Giodean, D. and Serban, G. DE 2401955 (17 July 1975). Farhad Khorasheh was born in 1961. He took his B.S. degree in Chemical
[4] Schramm, B., Kern, J., Schwahn, H., Preuss, A.W., Gottlieb, K. and Engineering in 1983 from Queen’s University, Ontario, Canada and his M.S.
Bruderreck, H. DE 3739002 A1 (24 May 1989). and Ph.D. degrees in Chemical Engineering, in 1986 and 1992, respectively,
[5] Kirner, J.F. GB 2162082 A1 (29 January 1986). from the University of Alberta, Edmonton, Canada. He is now a Professor in
[6] Box, E. US 3692701 (19 September 1972). the Department of Petroleum and Chemical Engineering at Sharif University of
[7] Wilhelm, F.C. US 3998900 (21 December 1976). Technology. His research interests include Modeling and Reactor Design.
[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). Saeed Niknaddaf was born in 1985. He obtained his B.S. and M.S. degrees from
[10] Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y. and Wu, P. ‘‘Propane the Department of Petroleum and Chemical Engineering at Sharif University of
dehydrogenation on Pt–Sn/ZSM-5 catalyst: effect of tin as a promoter’’, Technology in 2007 and 2009, respectively, in the area of Thermo Kinetics and
Catalysis Communications, 7, pp. 860–866 (2006). Catalysis.
[11] Bobrov, V.S., Digurov, N.G. and Skudin, V.V. ‘‘Propane dehydrogenation
using catalytic membrane’’, Journal of Membrane Science, 253, pp. 233–242
(2005). Mahnaz Soltani was born in 1974. She obtained her B.S. and M.S. degrees from
[12] Guo, J., Lou, H., Zhao, H., Zheng, L. and Zheng, X. ‘‘Dehydrogenation the Department of Chemical Engineering at Tehran University in 1996 and 1998,
and aromatization of propane over rhenium-modified HZSM-5 catalyst’’, respectively. She is now working as a Chemical Engineer in the National Iranian
Journal of Molecular Catalysis A: Chemical, 239, pp. 222–227 (2005). Petrochemical Company.

You might also like