far joo 2011

Upload: reclatis14

Post on 06-Jul-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/18/2019 Far Joo 2011

    1/7

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

    Sharif University of Technology

    Scientia IranicaTransactions C: Chemistry and Chemical Engineering 

    www.sciencedirect.com

    Kinetic modeling of side reactions in propane dehydrogenation overPt-Sn/γ -Al2O3 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, Iranb 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/γ -Al2O3wascarried out in a laboratory-scale plug-flowreactor at 580–620 °C under atmospheric pressure. Several

    kinetic models derived from different reaction mechanisms were testedusing 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 Langmuir–Hinshelwood mechanism, where the adsorption of propane

    was the ratecontrolling 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 someof the most important chemical products including polymers,

    synthetic rubbers and oxygenated compounds for reformulatedfuels [1,2]. Propylene is a major co-product from steam crackers

    and the economics of the process are highly dependent on thesupply/demand situation for both propylene and ethylene. In

    some geographic areas, propylene demand is increasing fasterthan ethylene demand. The dehydrogenation of propane with

    high selectivity to propylene gives the chance to de-couple theproductions 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: [email protected] (F. Khorasheh).

    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 [3–5], and, more recently, supported

    platinum catalysts   [6–9]; 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 platinum–tin catalyst has therefore received considerable

    attention [6–9].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 450–650 °C under atmospheric pressure. The following

    reactions occur during propane dehydrogenation [10–13]:

    C3H8 ↔ C3H6 + H2,   (R1)

    C3H8 → CH4 + C2H4,   (R2)

    C2H4 + H2 → C2H6,   (R3)

    C3H8 + H2 → CH4 + C2H6.   (R4)Reaction   (R1)   is the main reaction and reactions  (R2)–(R4)are the possible side reactions. Reaction   (R2)   can take place

    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

    http://dx.doi.org/10.1016/j.scient.2011.05.009http://www.sciencedirect.com/mailto:[email protected]://dx.doi.org/10.1016/j.scient.2011.05.009http://dx.doi.org/10.1016/j.scient.2011.05.009mailto:[email protected]://www.sciencedirect.com/http://dx.doi.org/10.1016/j.scient.2011.05.009

  • 8/18/2019 Far Joo 2011

    2/7

     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  g

    cm3

    Catalyst Density (ρc ) 1.12  g

    cm3

    Catalyst Surface Area (Sa) 200   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 catalyticdehydrogenation of propane over a commercial Pt-Sn/γ -Al2O3catalyst.

    The major side products were CH4, C2H4 and C2H6. The mainside product under the reaction conditions employed in this

    study was CH4, as it was produced via hydrogenolysis, thermal

    cracking and catalytic cracking reactions.

    2. Methods and materials

    The extent of side reactions in propane dehydrogenation

    were examined over a commercial Pt-Sn/γ -Al2O3  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 gasmixtures 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 

    thereactor, in betweentwo layersof quartsparticles. Theheight

    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 waszero. The temperature wasthen increased from 450 °Cto530 °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 temperaturewas then increased to the desired reaction temperatures of 

    580  °C, 600  °C or 620  °C at a rate of 5  °C per minute. Productgases 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,ENRO3015).

    Figure 1: The simplified sketch of propane dehydrogenation setup.

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

    Run Catalyst

    weight (g)

    Q C3(cc/min)

    Q H2(cc/min)

    WHSV

    (1/h)

    Conversion

    (%)

    1 1 39.96 31.968 4 20.45

    2 1.2 44.36 35.488 4 23.07

    3 1.5 55.45 44.36 4 23.32

    4 1.6 59.3 47.44 4 23.35

    3. Results and discussions

    To obtain proper kinetic data for determination of intrin-

    sic 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 differ-

    ent 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 exter-

    nal 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 h−1. If conversion wasunaffected by varying the propane flow

    rate at constant WHSV, it could be concluded that the reactionwas 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 resultsare presented in Table 3.  If the conversion did not change by

  • 8/18/2019 Far Joo 2011

    3/7

    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

    weight

    (g)

    Q C3(cc/min)

    Catalyst

    diameter

    (mm)

    Q H2(cc/min)

    WHSV

    (1/h)

    Conversion

    (%)

    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 24.15 0.75 55 0.35 44 9 24.3

    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.

    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 increasingreaction temperature (Figure2), 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 C3H6 =C3H6 Produced

    Input(C3H8)− Output(C3H8).   (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. Theamounts 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 3: Effect of temperature on propylene selectivity.

    Figure 4: Effect of temperature on the formation of side products (WHSV =8 (1/h)).

    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 dehy-

    drogenation of propane over the Pt-Sn/γ -Al2O3   catalyst canbe well represented in terms of Langmuir–Hinshelwood rate

  • 8/18/2019 Far Joo 2011

    4/7

     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 7: Effect of residence time on formation of side products (T  = 600 °C).

    Figure 8: Molar ratio of side products to all products at different WHSV and

    reaction temperatures.

    expressions   [13], where either an adsorption, surface reac-tion, or desorption step is considered as the rate controllingstep. Based on experimental data, Airaksinen et al. (2002) indi-cated that desorption of propylene from the surface sites wasnot likely to be the rate controlling step and suggested sev-eral mechanisms for propane dehydrogenation, where eitheradsorption or surface reaction was rate controlling  [14]. A re-cent study, where the same commercial catalyst as that used inthe current investigation was employed, suggested that mech-anisms involving propane adsorption either on single or dualsites as the Rate Controlling Step (RCS) were most consistentwith experimental data [15].

    The Langmuir–Hinshelwood mechanisms that were studiedin the current investigation to describe the kinetics of the

    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 thesurface reaction. It should be noted that competitive adsorption

    of side products, including CH4, C2H4   and C2H6, 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

    The estimation of the kineticparameters 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 forthe 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:

    dF i

    dW =−

     j

    −r i,

    i = C3H8, C3H6, C2H4, C2H6, CH4,   (2)where F i  is the molar flow rate of component i, W  is the massof catalyst,

     −r i   is the reaction rate for component   i, and the

    summation coversall reactions j leading to theformation and/ordisappearance 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−

    k

    =1

    5−i

    =1

    (F exp

    iE    − F pre

    iE    )2W i,   (3)

  • 8/18/2019 Far Joo 2011

    5/7

    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

    C3H8(g)+ 2S ←→ C3H7 · · · S+ H · · · S (RCS)   C3H8(g)+ S ←→ C3H7 · · · S · · ·H (RCS)C3H7 · · · S + S ←→ C3H6 · · · S+ H · · · S C3H7 · · · S · · ·H+ S ←→ C3H6 · · · S · · ·H+ H · · · SC3H6 · · · S ←→ C3H6(g)+ S C3H6 · · · S · · ·H ←→ C3H6(g)+ H · · · S2H · · · S ←→ H2(g)+ S 2H · · · S ←→ H2(g)+ 2SCH4 + S ←→CH4 · · · S CH4 + S ←→ CH4 · · · S

    C2H4 + S ←→ C2H4 · · · S C2H6 + S ←→ C2H6 · · · SC2H6 + S ←→ C2H6 · · · S C2H4 + S ←→ C2H4 · · · SS: Active sites on the catalyst S: Active site on the catalyst

    Mechanism III Mechanism IV

    C3H8(g)+ S ←→ C3H7 · · · S · · ·H (RCS)   C3H8(g)+ S ←→ C3H8 · · · SC3H7 · · · S · · ·H ←→ C3H6(g)+ H · · · S · · ·H C3H8 · · · S+ S ←→ C3H7 · · · S+ H · · · S (RCS)H · · · S · · ·H ←→ H2(g)+ S C3H7 · · · S+ S ←→ C3H6 · · · S+ H · · · SCH4 + S ←→ CH4 · · · S C3H6 · · · S ←→ C3H6(g)+ SC2H4 + S ←→ C2H4 · · · S 2H · · · S ←→ H2(g)+ 2SC2H6 + S ←→ C2H6 · · · S CH4 + S ←→ CH4 · · · SS: Active sites on the catalyst C2H4 + S ←→ C2H4 · · · S

    C2H6 + S ←→ C2H6 · · · SS: Active site on the catalyst

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

    Model Equation rate

    1 (I-a)   −r ′ A =k′

    P P −P Pr P H2

    K eq

    (1+K ′P PrP 1/2H2 +K PrP Pr+(K ′′′P H2 )

    1/2+K H2 P H2+K CH4 P CH4+K C2H4 P C2H4+K C2H6 P C2H6 )2

    2 (II-a)   −r ′ A =k′

    P  p−P Pr P H2

    K eq

    1+K PrH 2 P PrP H2+(K H2 P H2 )1/2+K ′P PrP 1/2H2

    +K CH4 P CH4+K C2H4 P C2 H4+K C2H6 P C2 H6

    3 (III-a)   −r ′ A =k′

    P  A−P B P C K eq

    1+K PrH 2 P C3 H6 P H2+K H2 P H2+K CH4 P CH4+K C2H4 P C2H4+K C2H6 P C2H6

    4 (IV-b)   −r ′ A =k′

    P  p−P Pr P H2

    K eq

    (1+K  pP  p+K PrP Pr+(K H2 P H2 )1/2+K ′P Pr√ 

    P H2+K CH4 P CH4+K C2 H4 P C2H4+K C2 H6 P C2H6 )2

    P: Propane (A)   k′: Rate constantPr: Propylene (B)   K eq: Equilibrium constant

    H2 = Hydrogen (C)   K : Adsorption constant

    where F exp

    iE    and F pre

    iE    are the experimental and predicted molarflow rates of component  i  at the reactor exit,  W i  are weight-ing factors, and summation k extends over experimental dataatdifferent WHSV. The optimized parameter values were highlyinfluenced by the choice of weighting factors, as propane andpropylene were major species, while methane, ethane andethylene were present in minor quantities. In the optimizationprocedure 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 revisedweighting factors were evaluated as follows:

    W i = Errori/Total error,   (4)

    Errori =4−

    k=1(F expiE    − F preiE    )2,   (5)

    Total error =5−

    i=1Errori.   (6)

    The normalized weighting factors were then used to optimizethe kinetic parameters of the main reaction for each reactiontemperature. The temperature dependence of the rate con-stants and equilibrium adsorption constants were determinedby the Arrhenius and van’t Hoff equations, respectively, asfollows:

    ln(ki) = ln(ki0)+ (−E i/RT ),   (7)ln(K i)

    =ln(K i0)

    +(

    −H i/RT ).   (8)

    Table 6: Proposed reaction rate equations for side reactions.

    Side reaction Kinetic equation

    Catalytic cracking (I)   r  = k1P C3 H8Ethylene hydrogenation (II)   r  = k2P C2 H4 P H2Hydrogenolysis (III)   r  = k2P C3 H8 P H2Thermal cracking (IV)   r  = k4P C3 H8

    Theconstantsare reported in Table 7 forthe main reaction using

    different reactions [10–12]. Proposed Langmuir–Hinshelwood

    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 thefollowing objective function:

    OF =4−

    k=1

    3−i=1

    (F exp

    iE    − F pre

    iE    )2,   (9)

    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 shouldbe 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.

  • 8/18/2019 Far Joo 2011

    6/7

     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), H  (kJ/mol).

    Model 1 Model 2 Model 3 Model 4

    k′ = 3.17× 103 exp −61.79

    RT 

      k′ = 1.21× 103 exp

    −53.71RT 

      k′ = 3.71× 103 exp

    −59.39RT 

      k′ = 4.25× 101 exp

    −70.66RT 

    K Pr = 7.44× 10−7 exp

    89.07

    RT 

      K H2 = 1.61× 10−3 exp

    36.03

    RT 

      K H2 = 9.71× 105 exp

    52.22

    RT 

      K  p = 6.70× 10−3 exp

    18.92

    RT 

    K H2 = 2.18× 10−1 exp

    9.36RT 

      K CH4 = 1.37× 10−5 exp

    52.63

    RT 

      K CH4 = 1.70× 103 exp

    21.28

    RT 

      K Pr = 1.42× 10−5 exp

    69.31

    RT 

    K CH4 = 9.07 × 10−3 exp

    16.06RT 

      K C2H6 = 1.34× 10−6 exp 64.39

    RT    K C2 H6 = 1.85× 104 exp

    38.23RT 

      K H2 = 2.01× 10−4 exp 51.24

    RT  K C2H6 = 3.60× 10−2 exp 5.75RT    K C2H4 = 5.15× 10−5 exp 42.84RT    K C2 H4 = 7.22× 105 exp 4.17RT    K CH4 = 1.98× 10−2 exp 6.83RT  K C2H4 = 6.97× 10−3 exp

    16.44

    RT 

      K Pr H2 = 3.70× 10−2 exp

    7.41RT 

      K Pr H2 = 1.67× 10−2 exp

    18.35

    RT 

      K C2H6 = 2.99×10−2 exp

    5.48

    RT 

    K ′ = 3.44× 10−2 exp

    12.63

    RT 

      K ′ = 3.47× 10−1 exp

    −9.16RT 

      K C2H4 = 3.39×10−2 exp

    5.61

    RT 

    K ′′′ = 1.02× 10+4 exp−49.97

    RT 

      K ′ = 7.73× 104 exp

    68.46

    RT 

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

    Model 1 Model 2 Model 3 Model 4

    k1 = 1.14 exp −6488

      K 1 = 7.53× 108 exp

    −273938T 

      k1 = 3.36× 1017 exp

    −40451T 

      k1 = 3.09× 104 exp

    −16661T 

    k2 = 1.34× 107 exp

    −27393T 

      k2 = 9.62× 107 exp

    −28187T 

      k2 = 3.94× 1011 exp

    −30754T 

      k2 = 1.25× 109 exp

    −25692T 

    k3 = 3.36× 104 exp

    −14723T 

      k3 = 5.57× 105 exp

    −17926T 

      k3 = 1.47× 101 exp

    −8498T 

      k3 = 2.85× 109 exp

    −5757.6T 

    k4

     =1.53E 

    +09 exp

    −1598.23T    k4 =

    1.53E 

    +09 exp

    −1598.23T    k4 =

    1.53E 

    +09 exp

    −1598.23T    k4 =

    1.53E 

    +09 exp

    −1598.23T 

    Table 9: Average % error between predictedand experimentalflow rates at

    reactor exit.

    Model Absolute error %

    Model 1  Main species 4.02

    Side products 14.68

    Model 2  Main species 6.22

    Side products 21.74

    Model 3  Main species 7.11

    Side products 17.74

    Model 4  Main species 12.18

    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

    the exit molar flow rates of various spices under different reac-

    tion conditions. Comparison between the experimental values

    andpredictedvalues 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-

    sonof 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 sideproducts, respectively.

    4. Conclusions

    The kinetics of propane dehydrogenation over a commercial

    Pt-Sn/γ -Al2O3  catalyst was best described in terms of a Lang-

    muir–Hinshelwood mechanism, where dual site adsorption of 

    propane was considered as the rate controlling step. The ki-

    netics of side reactions leading to the formation of methane,

    ethylene and ethane as minor products could reasonably be de-

    scribed by simplepower-law rate expressions. The proposed ki-

    netic model resulted in an average relative error of 4.0% and

    14.7% between predicted and experimental molar flow rates

    Figure 10: Comparison between experimental exit molar flow rates and

    predicted values from model 1 for major products at reaction temperature of 

    580 °C.

    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 therange of experimental conditionsemployedin this

    study.

  • 8/18/2019 Far Joo 2011

    7/7

    464   A. Farjoo et al. / Scientia Iranica, Transactions C: Chemistry and Chemical Engineering 18 (2011) 458–464

     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 fluidizedbed reactor’’, Catalysis Today, 52, pp. 259–269 (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 ISO-butane over platinum–potassium catalyst’’,  Kinetics and Catalysis, 36(4),pp. 504–510 (1995).

    [3] Pop, E.,Goidean, N., Giodean, D. andSerban, G. DE2401955 (17July 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. ‘‘Propanedehydrogenation on Pt–Sn/ZSM-5 catalyst: effect of tin as a promoter’’,Catalysis Communications, 7, pp. 860–866 (2006).

    [11] Bobrov, V.S., Digurov, N.G. and Skudin, V.V. ‘‘Propane dehydrogenationusing catalytic membrane’’, Journal of Membrane Science, 253, pp.233–242(2005).

    [12] Guo, J., Lou, H., Zhao, H., Zheng, L. and Zheng, X. ‘‘Dehydrogenationand aromatization of propane over rhenium-modified HZSM-5 catalyst’’, Journal of Molecular Catalysis A: Chemical, 239, pp. 222–227 (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. 5619–5627 (2002).

    [15] Mohagheghi, M. and Bakeri, G. ‘‘Kinetic studies of propane dehydrogena-tionover Pt-Sn/γ -Al2O3 catalyst’’,in: Presented at International Conferenceon Chemical Reactors, CHEMREACTOR-18, Malta(29 September–3 October2008).

     Afrooz Farjoo  was born in 1985. She obtained her B.S. degree from IsfahanUniversity of Technology in 2007 and her M.S. degree in 2009 from theDepartment 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 ChemicalEngineering in 1983 from Queen’s 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 inthe 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 fromthe Department of Petroleum and Chemical Engineering at Sharif University of Technology in 2007 and 2009, respectively, in the area of Thermo Kinetics andCatalysis.

    Mahnaz Soltani was born in 1974. She obtained her B.S. and M.S. degrees fromtheDepartmentof ChemicalEngineeringat Tehran Universityin 1996and 1998,respectively. Sheis nowworkingas a ChemicalEngineerin theNational IranianPetrochemical Company.