saheb del far 2012

Upload: reclatis14

Post on 06-Jul-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/18/2019 Saheb Del Far 2012

    1/8

    chemical engineering research and design 9 0 ( 2 0 1 2 ) 1090–1097

    Contents lists available at SciVerse ScienceDirect

    Chemical Engineering Research and Design

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

    Kinetic study of propane dehydrogenation and side

    reactions over Pt–Sn/Al2O3 catalyst 

    Saeed Sahebdelfar, Maryam Takht Ravanchi∗, Farnaz Tahriri Zangeneh,Shokoufeh Mehrazma, Soheila Rajabi

    Catalyst Research Group, Petrochemical Research and Technology Company, National Petrochemical Company, No. 27, Sarv Alley,

    Shirazi-south, Mollasadra, P.O. Box 14358-84711, Tehran, Iran

    a b s t r a c t

    The kinetics of reactions involved in dehydrogenation of propane to propylene over Pt–Sn/Al2O3 catalyst was studied.

    The simultaneous deactivation of individual dehydrogenation, hydrogenolysis and cracking sites was also studied.

    A model was developed to obtain the transient conversion of propane, product selectivity and catalytic site activity.

    The dehydrogenation reaction was considered as the main reaction governing propane and hydrogen concentrations

    along the reactor. Catalytic test runs were performed in a fixed-bed quartz reactor. The kinetic expressions devel-

    oped for the main and side reactions were verified by integral and a combination of integral–differential analysis of 

    reactor data, respectively, andthe kinetic parameters were obtained. Thedeactivation of the active sites for the three

    reactions was found to follow a first-order independent decay law. The rate constants of deactivation were found to

    decrease in the order of dehydrogenation, hydrogenolysis and cracking. Noncatalytic thermal cracking was found to

    be comparable to the catalytic route resulting in a very low apparent deactivation rate constant for cracking reaction.

    © 2011 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.Keywords: Propane dehydrogenation; Pt–Sn catalysts; Cracking; Hydrogenolysis; Catalyst deactivation; Kinetics

    1. Introduction

    Propylene is an important feedstock for producing a vari-

    ety of petrochemicals such as polypropylene, acrolein and

    acrylic acid. Due to ever-increasing demand and insufficient

    supply by crackers, alternative production methods such

    as propane dehydrogenation (PDH) has received attention

    (Heinritz-Adrian et al., 2008):

    C3H8  ⇔ C3H6 + H2, H298◦= +129 kJ/mol (1)

    The reaction is highly endothermic and equilibrium-

    limited; therefore, higher temperatures and lower pressures

    are necessary to achieve acceptable conversions. Unfortu-

    nately, these conditions favor side reactions and accelerate

    catalyst deactivation as well.

    In commercial practice both chromia (Arora, 2004; Miracca

    and Piovesan, 1999; Weckhuysen and Schoonheydt, 1999) and

    platinum (Bricker et al., 1990; Heinritz-Adrin et al., 2003;

    Pujado and Vora, 1990) based catalysts have been employed

    ∗ Corresponding author. Tel.: +98 21 44580100; fax: +98 21 44580505.E-mail addresses: [email protected], [email protected] (M.T. Ravanchi).Received15 April2011;Receivedin revisedform20 October 2011;Accepted4 November 2011

    for paraffin dehydrogenation. Platinum exhibits high catalytic

    activity in dehydrogenation of paraffins. To achieve high plat-

    inum dispersions, high-surface area supports are commonly

    used. Acidic sites on the support promote cracking (reaction

    (2)) and coke formation reactions. These sites are effectively

    neutralized by application of alkaline promoters (Bai et al.,

    2009; Bhasin et al., 2001; Padmavathi et al., 2005; Sanfilippo

    and Miracca, 2006; Yu et al., 2006; Zhang et al., 2006a)

    C3H8  ⇔ C2H4 + CH4, H298◦= +79.4 kJ/mol (2)

    Dehydrogenation virtually occurs on all platinum sites,

    while hydrogenolysis (reaction (3)) occurs on low coordination

    sites (steps and kinks) (Resasco, 2003)

    C3H8 +H2  ⇔ C2H6 + CH4, H298◦= −63.4 kJ/mol (3)

    In commercial catalysts, Sn is used as promoter to sup-

    press hydrogenolysis reaction through reduction of surface

    Pt ensemble size by dividing Pt surface to smaller ensembles

    0263-8762/$ – see front matter © 2011 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.doi:10.1016/j.cherd.2011.11.004

    http://www.sciencedirect.com/science/journal/02638762http://www.elsevier.com/locate/cherdmailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_7/dx.doi.org/10.1016/j.cherd.2011.11.004http://localhost/var/www/apps/conversion/tmp/scratch_7/dx.doi.org/10.1016/j.cherd.2011.11.004mailto:[email protected]:[email protected]://www.elsevier.com/locate/cherdhttp://www.sciencedirect.com/science/journal/02638762

  • 8/18/2019 Saheb Del Far 2012

    2/8

    chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097 1091

    Nomenclature

    a catalyst activity

    C concentration (mol/m3)

    F feed rate (mol/h)

    ki   forward rate constant of reaction i

    kdi   rate constant for catalyst deactivation in reac-tion i (h−1)

    Keq   concentration equilibrium constant (mol/m3)

    ri

      rate of reaction i per mass of catalyst

    (mol/(kg h))

    W    weight of catalyst in the reactor (kg)

    Xi   conversion of key reactant A in reaction i

    Xei   equilibrium conversion of key reactant A in

    reaction i

    t time-on-stream (h)

    Greek symbols

    ˛ parameter defined by Eq. (9)

    ˇ   parameter defined by Eq. (10)ε expansion factor, fraction change in volume

    resulting from change in total number of moles

    ratio of number of moles of species initially

    entering to that of paraffin

      a capacity factor (catalyst weight per volumet-

    ric feed flow rate) ((kgh)/m3)

    Subscripts

    0 reactor inlet

    A key reactant, propane

    B key product, propylene

    H hydrogen

    i   reaction numberout reactor outlet

    (Barbier et al., 1980; Carvalho et al., 2001; Larese et al., 2000;

    Takehira et al., 2004; Waku et al., 2003; Yu et al., 2007; Zhang 

    et al., 2006c, 2007).

    Nevertheless, side reactionsstill occur to a rather apprecia-

    ble extent on the catalysts, so that the selectivity of the UOP

    Oleflex process in dehydrogenation of propane and isobutane

    is 90% and 92%, respectively (Bhasin et al., 2001).

    Both cracking and hydrogenolysis reactions may occur by

    single or multiple C–C bond rupture according to the catalyst

    formulation and operating conditions. However,a recent study

    (Sahebdelfar and Tahriri Zangeneh, 2010) has shown that over

    the catalyst and under reaction conditions employed in this

    work, the former path is predominant. Consequently, reac-

    tions (2) and (3) can represent the side reactions.

    It has been shown that side reactions result in significant

    influence on the performance of commercial-sized dehydro-

    genation reactors (Sahebdelfar et al., 2011). Unfortunately, it

    is difficult to study individual reactions independent of other

    reactions. Hydrogenolysis occurs virtually on the same sites

    as dehydrogenation. Cracking can occur separately, but under

    conditions different from that of dehydrogenation reaction.

    Another important difficulty in commercial scale imple-

    mentation of paraffin dehydrogenation is rapid catalyst

    deactivation due to coke formation (Moulijn et al., 2001). Dur-

    ing thecourse of reaction, different sites deactivate at different

    rates resulting in a change in activity and selectivity with

    time-on-stream. The overall deactivation kinetics has been

    studied elsewhere (Moghimpour Bijani and Sahebdelfar, 2008).

    Qing et al.(2011) studied the kinetics of propane dehydrogena-

    tion, cracking and coke formation over Pt–Sn/Al2O3  catalyst,

    assuming same activity for all sites. The distinction of active

    sites as differenttypes hasalso been proposed in theliterature

    (Kumbilieva et al., 2006).

    Compared to the studies on the mechanism of propane

    dehydrogenation including coke formation and the efforts toimprove the activity and stability of the catalyst, relatively few

    efforts have been made to establish a complete and reliable

    kinetic model of engineering significance, including kinet-

    ics of dehydrogenation, hydrogenolysis and cracking as well

    as deactivation of the corresponding sites. These models are

    essential for optimization of the reactor performance through

    increasing propylene yield and catalyst lifetime.

    In the present work, reaction kinetics and deactivation of 

    different catalytic sites of a commercial Pt–Sn/Al2O3   catalyst

    in dehydrogenation of propane to propylene are studied. A

    model is developed to obtain the kinetics of the reactions

    involved and the deactivation of the corresponding sites and

    to obtain the related kinetic parameters.

    2. Experimental

    2.1. Materials

    The propane, hydrogen and nitrogen feed gases were sup-

    plied by Roham Gas Co. with purity 99.5wt.%, 99.99 wt.% and

    99.99wt.%, respectively. Commercial Pt–Sn/-Al2O3   catalyst

    (Pt= 0.58wt.%, Sn= 0.8 wt.% and surface area= 187 m2 /g) with

    trade name DP-803, the characteristics of which is reported

    elsewhere (Sahebdelfar and Tahriri Zangeneh, 2010), was sup-

    plied by Procatalyse company. Quartz powder with grain size0.1 mm was used as catalyst diluent.

    2.2. Set-up

    The experimental setup, a schematic diagram of which is

    depicted in Fig. 1, was a fully automated system. All lines

    and fittings of the setup were made of stainless steel 316 (SS-

    316). A tubular fixed-bedquartz reactor with inner diameter of 

    15mm was used, the temperature of which was controlled by

    a furnace. Space above and below the catalyst bed (1.5g) was

    packed with quartz powder (3g) to ensure proper distribution

    of fluid flow. A thermocouple was inserted into the center of 

    the catalyst bed to indicate bed temperature. The channeling and heat transfer effects in the reactor could be neglected as

    the radial aspect ratio (bed diameter to catalyst particle diam-

    eter) was >15. The axial aspect ratio i.e. the ratio of catalyst

    bed length to catalyst particle diameter was >30 and hence

    the dispersion effects can be also neglected (Anderson and

    Pratt, 1985). In all runs, except for the initial data points, the

    carbon and hydrogen balance was within±1.5%.

    2.3. Experimental procedure

    The reactor was first heated under hydrogen flow of  

    18.3 N ml/min at about 5 ◦C/min to 530 ◦C to desorb the

    adsorbed moisture, if any, and then the catalyst was reduced

    in hydrogen flow at 530◦C for 1h in the reactor. The kinetic

    experiments were carried out at 620 ◦C with molar ratio of  

    hydrogen to hydrocarbon equal to 0.6 and weight hourly

    space velocity (WHSV) 2.2 h−1. The gas productswere analyzed

  • 8/18/2019 Saheb Del Far 2012

    3/8

    1092 chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097

    Fig. 1 – Schematic diagram of the experimental setup.

    using online gas chromatography Agilent 6890N RGA, which

    was equipped with a capillary column, HP-Plot Al2O3 /Na2SO4,

    50m, 530m, 15.0m and FID detector. The concentration of 

    propane, propylene and lower hydrocarbons was measured

    in the product stream to calculate conversion, selectivity and

    yield of the reactions.The product selectivities were calculated based on moles

    of carbon converted. The conversion of propane by reaction iwas calculated as:

    Conversion by reactioni

    =moles of propane in− moles of propane converted by reaction i

    mole of propane in

    (4)

    The test run times were comparable to one catalyst cycle

    (5–7 days)in theOleflex process when thecatalyst being sent to

    continuous catalyst regeneration (CCR) unit for regeneration.

    3. Kinetic and reactor model

    3.1. Kinetic expressions

    Mostprevious studies have shown that dehydrogenation reac-

    tion is first-order in paraffin concentration and negative half 

    to zeroth-order in hydrogen concentration (Resasco, 2003).

    Therefore to incorporate the first-order dependence on paraf-

    fin concentration, hydrogen-inhibition effect, and chemical

    equilibrium limitation, the following expression is assumed

    to represent kinetics of the reaction:

    −r1  = k1CA − k−1CBCH2  = k1

    CA −

    CBCH2Keq

      (5)

    where r

    1

      is the rate of conversion of propane to propy-

    lene (reaction (1)) per catalyst weight, k1   and k−1   are the

    rate constants of forward and backward reactions, respec-

    tively, Keq is the equilibrium constant at reaction temperature,

    C is the concentration, and, A and B, respectively, rep-

    resenting the key components, propane and propylene.

    Larsson et al. (1996) f ound that among several kinetic mod-

    els tested for dehydrogenation of propane on Pt–Sn/Al2O3catalysts, simple power-law model of this type resulted

    in the best fit. In fact most mechanistic kinetic models,

    e.g. Longmuir–Hinshelwood–Hougen–Watson (LHHW) based

    model of Padmavathi et al. (2005) with surface reaction as

    rate-limiting step, approach to power law expressions at the

    high-temperature, low-partial pressure (low surface coverage

    (Scott Fogler, 1999)) prevalent in lower-paraffin dehydrogena-

    tion practice.To account for catalyst deactivation, the reaction rate can

    be written in terms of site activity as:

    −r1  = k1a1

    CA −

    CBCH2Keq

      (6)

    in which the activity of dehydrogenation sites, a1, is defined

    as (Scott Fogler, 1999):

    a1(t) =−r1(t)

    −r1(t = 0)  (7)

    Considering volume change of reaction, following neces-sary algebraic manipulations, one arrives at the following 

    expression for reaction rate in terms of propane conversion

    (Moghimpour Bijani and Sahebdelfar, 2008):

    −r1  =k1a1CA0(Xe1 − X1)(˛+ ˇX1)

    (1 + εAX1)2

      (8)

    in which

    ˛ =Xe1 + (H + B)+ HB(1 − εA + εAXe1)

    (H + Xe1)(B + Xe1)  (9)

    ˇ =(1 + ε

    AX

    e1−X

    e1)+ ε

    AX

    e1(

    H+

    B)+ ε

    A

    H

    B(H +Xe1)(B + Xe1)   (10)

    where H, B  and CA0  are hydrogen/paraffin, olefin/paraffin

    molar ratios and propane concentration in the feed, respec-

    tively, X1   is propane conversion to propylene, Xe1   is the

  • 8/18/2019 Saheb Del Far 2012

    4/8

    chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097 1093

    equilibrium conversion under reaction conditions and εA   is

    the volume expansion factor.

    Similarly, assuming reaction (1) as the main propane

    consuming reaction governing propane and hydrogen con-

    centrations along the reactor and noting that side reactions

    are far from equilibrium under dehydrogenation conditions

    (Waku et al., 2004), one may use the following power-law rate

    expressions for cracking (Lobera et al., 2008; Qing et al., 2011)and hydrogenolysis (Chin et al., 2011) reactions, respectively:

    −r2  = k2a2CA0(1 −X1)

    (1 + εAX1)  (11)

    −r3  = k3a3C2A0

    (1 −X1)(H + X1)

    (1 + εAX1)2

      (12)

    The deactivation rates could be considered as first-order

    and independent. Consequently, for reaction i:

    −dai

    dt  = kdiai   (13)

    where kdi is the respective rate constant of deactivation. Inte-

    grating this equation yields a correlation for ai  as a function

    of time, t:

    ln   ai  = −kdit (14)

    3.2. Reactor model

    The plug-flow reactor performance equation for reactions in

    parallel is:

    dW 

    FA0=

    dXi

    −ri (15)

    As the propane and hydrogen concentration profiles are

    largely determined by the main reaction (Eq. (1)), itsextent can

    be obtained independent of other reactions. Consequently, the

    integral analysis of the conversion data for reaction (1) along the reactor results in (Moghimpour Bijani and Sahebdelfar,

    2008):

    ln

    ˇ−1

      (ˇ − εA˛)

    2

    ˇ(˛ + ˇXe1) ln

    ˛ + ˇX1,out˛

    −ˇ(1 + εAXe1)

    2

    (˛ + ˇXe1)  ln

    Xe1 − X1,out

    Xe1

    − ε2AX1,out

    = −kd1t + ln(k1 ) (16)

    where W  is the catalyst weight, FA0   is the molar flow rate

    of the feed to reactor and  , the ratio of catalyst weight

    per volumetric feed flow rate, is a capacity factor known

    as the weight-time. The subscript out refers to reactor out-

    let value of the parameter. Plots of Eq. (16) should result in

    straight lines the slope of which giving kd1   and the intercept

    giving k1.

    0

    10

    20

    30

    40

    50

    60

    70

    0  20  40  60  80  100 120

    Time-on-stream, h

       C  o  n  v  e  r  s

       i  o  n ,

       %

    Fig. 2 – Overall conversion of propane (  ) and conversion to

    propylene (  ), ethylene (  ) and ethane (  ) ( T =620 ◦C,

    H2 /HC=0.6mol/mol and WHSV=2.2h−1 ).

    The rate of reactions (2) and (3) depends on reaction (1) asimplied by Eqs. (11) and (12). Dividing Eq. (15) f or reactions (2)

    and (3) to that for reaction (1), and using Eq. (14) f or deactiva-

    tion terms, one obtains respectively;

    (Xe1 −X1)(˛ + ˇX1)

    (1 − X1)(1 + εAX1)

    dX2dX1

    =k2k1

    exp((kd1 − kd2)t) (17)

    (Xe1 −X1)(˛ + ˇX1)

    (1 −X1)(H +X1)

    dX3dX1

    =CA0k3

    k1exp((kd1 − kd2)t) (18)

    Eqs. (17) and (18) hold for any point along the reactor.

    Rearranging, integrating with respect to Xi   and then dif-

    ferentiating with respect to t, the following correlations are

    obtained for cracking and hydrogenolysis reactions, respec-tively:

    ln

      ((1/X2,out)(dX2,out/dt)) − (kd1 − kd2)

    (1/X2,out)(dX1,out/dt)(((1− X1,out)(1 + εAX1,out))/((Xe1 − X1,out)(˛ + ˇX1,out)))

    = (kd1 − kd2)t + ln

    k2k1

      (19)

    ln

      ((1/X3,out)(dX3,out/dt))− (kd1 − kd3)

    (1/X3,out)(dX1,out/dt)(((1−X1,out)(H + X1,out))/((Xe1−X1,out)(˛ + ˇX1,out)))

    = (kd1 − kd3)t + ln

    k3CA0

    k1

      (20)

    Interested reader is referred to Appendix f or detailed mathe-

    matical derivations.

    The derivative terms could be obtained numerically from

    time-on-stream conversion data for different reactions. Sim-

    ilarly, plots of Eqs. (19) and (20) should result in straight lineswith the slope giving the difference of deactivation rate con-

    stants and the intercept giving the ratio of rate constants.

    Since the parameters kd1 −kd2  and kd1 −kd3  appear on both

    sides of Eqs. (19) and (20), respectively, an iterative procedure

    is necessary for their determination.

    Because of low deactivation rate compared to chemical

    conversion rates, the pseudo-steady condition was assumed

    to be valid in the above derivations.

    4. Results and discussion

    4.1. Performance test results

    Fig. 2 shows the overall conversion and conversion of indi-

    vidual reactions versus time-on-stream. The first data points

    showing the “initial activity”, characterized by large devi-

    ations, are never actually observed due to experimental

  • 8/18/2019 Saheb Del Far 2012

    5/8

    1094 chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097

    y = 0.063x - 0.2411

    R² = 0.9303

    y = 0.0439x + 2.4186

    R² = 0.9516

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    4.5

    0  10  20  30  40

    Conversion to propylene, %

       C  o  n  v  e  r  s

       i  o  n

       t  o   b  y  -  p  r  o

       d  u  c

       t  s ,

       %

    Fig. 3 – Conversion of propane to ethylene (  ) and ethane

    (  ) versus conversion to propylene at different 

    time-on-streams ( T =620 ◦C, H2 /HC=0.6mol/mol and

    WHSV=2.2h−1 ).

    limitations. After a sharp initial decline in side reactionsaccompanied by an increase in dehydrogenation selectivity,

    a gradual decline of the conversions is observed which is due

    to deactivation of the active sites involved. The initial period

    is due to the presence of too many active side-reaction sites

    which deactivate quickly leaving moderate sites for longer

    times-on-stream.

    Fig. 3 showsthat plot of conversions to ethylene and ethane

    versus that to propylene results in reasonably straight lines.

    This, along with Fig. 2, shows that the contribution of side

    reactions in overall consumption of propane is small, espe-

    cially in early time-on-streams.

    The fact that extrapolation of the line for hydrogenolysis in

    Fig. 3 passes close theorigin could be attributed to thefactthat

    both dehydrogenation and hydrogenolysis reactions occur on

    platinum sites. This is notthe case forcracking reactionwhich

    occurs on different sites and also thermally. From Fig. 3 one

    concludes thatcracking reaction could proceed even when the

    catalystis fully deactivatedand thatmore thanhalf of cracking 

    reaction originates from noncatalytic thermal cracking route.

    Cracking reaction occurs mainly on acidic sites of the carrier

    (Zhang et al., 2006b) and also proceeds thermally.

    4.2. Modeling results

    Fig. 4 showsa plot of Eq. (16) f or a long-term run using exper-

    imental data with LHS showing the left-hand-side of that

    equation. A favorable fit is observed. The slope and intercept

    give the deactivation and reaction rate constants for dehydro-

    genation reaction, respectively (Table 1).

    In this way Eq. (16) provides a method to obtain time-

    zero conversion to propylene i.e. an estimate of conversion in

    the absence of deactivation effects which is useful for kinetic

    study of the main reaction.

    Unlike the integral method of analysis used in Fig. 4,

    the plots of Eqs. (19) and (20) require a higher number and

    Table 1 – Calculated values of the rate constants( T =620 ◦C).

    Reaction no. ki   kdi

    1 4.7 m3 /(kg h) 0.016 h−1

    2 0.40 m3 /(kg h) 0.0027h−1

    3 0.023 m6 /(mol kg h) 0.011 h−1

    y = -0.0159x - 0.2855

    R2 = 0.9543

    -2.5

    -2

    -1.5

    -1

    -0.5

    0

    120100806040200

    Time-on-stream, h

       L   H   S  o

       f   E  q .

       1   6

    Fig. 4 – Typical plot of Eq. (16) using experimental data

    ( T =620 ◦C, H2 /HC=0.6mol/mol, WHSV=2.2h−1 ).

    more accurate experimental data because of the appear-

    ance of time-derivative terms in these equations. To avoid

    the fluctuations encountered in numerical differentiation, ithas been proposed to fit the data with an appropriate func-

    tion and then differentiate the resulted interpolating function

    (Levenspiel, 1999). The trends of time-data propose exponen-

    tial functions as good candidates (Fig. 2) which in fact result

    in fair fits. Figs. 5 and 6, respectively, show plots of Eqs. (19)

    and (20) obtained by this approach after achieving conver-

    gence of deactivation rate constant difference terms by the

    iterative procedure explained above, with LHS showing the

    left-hand-side of these equations. Favorable fits are observed.

    The resulted rate constants for side reactions are also given in

    Table 1.

    Table 1 reveals that the rate constants of side reactions

    are more than one order of magnitude smaller than thoseof the main reaction as required by a selective catalyst. Also,

    the dehydrogenation sites deactivate more rapidly than those

    of side reactions. This explains the observed drop of selectiv-

    ity to propylene with time-on-stream. The seemingly smaller

    deactivation rate constant in the case of cracking reaction can

    be attributed to the simultaneous occurrence of non-catalytic

    thermal cracking. The numerical values of rate and deactiva-

    tion constants are also consistent with the experimental data

    trends observed in Figs. 2 and 3.

    The existence of a noncatalytic component in cracking 

    activity implies that higher orders of deactivation could give

    y = 0.01318x - 2.45327

    R² = 0.97686

    -4

    -3.5

    -3

    -2.5

    -2

    -1.5

    -1

    -0.5

    0

    120100806040200

    Time-on-stream, h

       L   H   S  o

       f   E  q .

       1   9

    Fig. 5 – Plot of Eq. (19) using experimental data with

    kd1  − kd2 =0.0132 ( T =620◦C, H2 /HC=0.6mol/mol and

    WHSV=2.2h−1 ).

  • 8/18/2019 Saheb Del Far 2012

    6/8

    chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097 1095

    y = 0.0049x - 3.3575

    R² = 0.8367

    -6

    -5.5

    -5

    -4.5

    -4

    -3.5

    -3

    -2.5

    -2

    120100806040200

    Time-on-stream, h

       L   H   S  o

       f   E

      q .

       2   0

    Fig. 6 – Plot of Eq. (20) using experimental data with

    kd1  − kd3 =0.0049 ( T =620◦C, H2 /HC=0.6mol/mol and

    WHSV=2.2h−1 ).

    better fits for apparentdeactivationof cracking sites for longer

    times-on-stream. This could complicate the corresponding 

    formulations. However, too much long times-on-stream are

    not of practical interest as partiallydeactivated catalystwill be

    sent to regeneration unit before complete deactivation could

    occur.

    In Fig. 7 parity plots for total conversion of propane

    (Fig. 7a) and propane conversion to species (Figs. 7b–d) are

    presented. These plots compare the calculated conversions

    versus experimental conversions. As it can be seen, gener-

    ally, the difference between experimental results and model

    estimation is within 20% which confirms the accuracy of the

    results.

    It is noteworthy that in constructing these plots, the main

    assumption of modeling (i.e. predominance of dehydrogena-

    tion reaction in propane consumption) is not applied to have

    a better insight of the capability of the modeling. The best

    results are observed for the total conversion and conversion

    to propylene. In case of side reactions, however, the approach

    of experimental and calculated results occurs at shorter and

    longer time-on-streams for cracking and hydrogenolysis reac-

    tions, respectively. As in Fig. 7a, which is for total conversion

    of propane to products, there is a good correlation between

    experimental and model results, the rate constants calculatedfrom these data and reported in Table 1 have a reasonable

    accuracy.

    4.3. Issues on validity and accuracy

    While the method of data analysis of the main reaction is

    purely integral, that of side reactions is a combination of 

    integral anddifferential analysesthe accuracy of which is lim-

    ited by the latter (that is, by time derivatives of conversions

    data). As mentioned above, an approach is to fit experimental

    conversion data with an appropriate function and then differ-

    entiate the resulted fitting function. The use of this approach

    for evaluation of the derivatives is inevitable in analysis of 

    long-term deactivation data, as after certain time-on-stream

    the decreaseof conversions within the specifiedstep-size time

    interval becomes smaller than that of experimental accuracy

    and/or system disturbances. Therefore, using direct numeri-

    cal differentiation formulas, these errors largely maskthe true

    value of the derivatives. The function should be checked by

    the eye to give both close fit to data and relevant slopes. The

    exponential function is the simplest one to satisfy both these

    requirements fairly. However, no simple function can fit both

    data points and their derivatives over a long range. Conse-

    quently, a small downward curvature is observed in plots of 

    Eqs. (19) and (20) with the exponential functions employed

    for fitting (see Figs. 5 and 6). Alternatively, direct numerical

    differentiation of data using up to fourth-order formulas did

    Fig. 7 – Correlation between experimental data and model predictions for total propane conversion and propane

    conversions to species in operating conditions used.

  • 8/18/2019 Saheb Del Far 2012

    7/8

    1096 chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097

    not result in satisfactory values when applied to the whole of 

    time-on-stream domain, due to considerable fluctuations in

    calculated derivatives.

    The apparent concentration-independent deactivation in

    decay laws implies that deactivation might be caused both by

    reactant and products(Levenspiel, 1999). In fact, both the reac-

    tant (e.g. through pyrolysis reaction) and products (through

    oligomerization–aromatization of the olefinic products e.g. of reactions (1) and (2)) can bring about coke formation and

    catalystdeactivation(Qing et al., 2011). The independent deac-

    tivation is also a characteristic of catalyst decay by thermal

    sintering (Levenspiel, 1999). However, the observed low orders

    of deactivation and negligible Pt crystal growth during reac-

    tions due to rather low reaction temperature compared to the

    melting point of Pt do not favor deactivation by sintering dur-

    ing reaction.

    Finally, ethylene/ethane hydrogenation/dehydrogenation

    could occur as additional side reactions. The rate of these

    reactions should be very small due to the low concentration

    of C2   products within the reactor. Furthermore, the ethy-

    lene to ethane ratio in the product is much higher thanthe equilibrium ratio implying that there is thermodynamic

    driving force for hydrogenation of ethylene to ethane. How-

    ever, in an earlier work no appreciable decrease in ethylene

    to ethane ratio observed upon decreasing the space-velocity

    (Sahebdelfar and Tahriri Zangeneh, 2010). This illustrates that

    hydrogenation reaction rate is not sufficiently high to con-

    tribute an importantrole in selectivity to ethaneand ethylene.

    Consequently, ethane and ethylene production rates can be

    adequately considered as measuresfor the rateof hydrogenol-

    ysis and cracking reactions, respectively.

    The applicability of the simple yet practical approach

    employed in this work depends on the selectivity to the main

    product, propylene, becoming more accurate as the selectiv-ity approaches to unity. The propylene selectivities in the

    data series employed in this work were mostly within the

    range 80–85%. This range is still sufficiently large such that

    the main reaction determines the concentrations of propylene

    andhydrogen within thereactor. Therefore, the kinetic param-

    eters obtained should be accurate to at least one significant

    figure.

    Higher selectivities are not uncommon both on lab or com-

    mercial scale runs (Barias et al., 1996; Kogan and Herskowitz,

    2001). This indicates that the proposed model could be used

    in most of the cases of practical interest.

    5. Conclusions

    The activity and deactivation kinetics of dehydrogenation,

    hydrogenolysis and cracking sites in dehydrogenation of 

    propane over Pt–Sn/Al2O3   catalyst were obtained when the

    reactions occur simultaneously. Power law expressions and

    first order independent decay laws fitted the kinetic data

    of the reactions favorably. The rate constant of the main

    reaction was found to be more than one order of magni-

    tude larger than those of cracking and hydrogenolysis side

    reactions. On the other hand, the rate constant of the deac-

    tivation of dehydrogenation reaction was found to be larger

    than those side reactions which explain the loss of selectivity

    to propylene with time-on-stream. The findings of this work

    could be applicable in modeling of commercial size reactors

    where side reactions and catalyst deactivation play important

    roles.

    Appendix.

    Eq. (17) can be written as below:

    (Xe1 −X1)(˛ + ˇX1)

    (1 − X1)(1 + εAX1)

    dX2dX1

    = k exp( At) (A1)

    where

    k =k2k1

    (A2)

     A = kd1 − kd2   (A3)

    Rearranging Eq. (A1), one obtains:

    dX2  = k exp( At)  (1 − X1)(1 + εAX1)

    (Xe1 −X1)(˛ + ˇX1)dX1   (A4)

    Integrating this equation, gives X2,out as a function of X1,out as:

    X2,out  = k exp( At)

       X1,out0

    (1 − X1)(1 + εAX1)

    (Xe1 −X1)(˛ + ˇX1)dX1   (A5)

    Differentiating this equation with respect to t, gives:

    dX2,outdt

      = kA exp( At)

       X1,out0

    (1 −X1)(1 + εAX1)

    (Xe1 − X1)(˛ + ˇX1)dX1

    +k exp( At) ∂

    ∂t

       X1,out0

    (1 − X1)(1 + εAX1)

    (Xe1 −X1)(˛ + ˇX1)dX1

      (A6)

    inwhich,thefirsttermontherighthandsideis AX2,out

     (accord-

    ing to Eq. (A5)). In the second term, as differentiation is with

    respect to t and the integral limits are functions of time, the

    Leibniz’s rule (Bird et al., 2002) must be applied. According to

    this rule, for function F(x, t), where

    F(x, t) =

       b(t)a(t)

     f (x, t)dx (A7)

    the time derivative is:

    dF

    dt  =

       ba

    ∂f 

    ∂t dx + f (b, t)

    db

    dt  − f (a, t)

    da

    dt  (A8)

    Hence, the second term of the right hand side of Eq. (A6) is

    k exp( At)

       X1,out0

    ∂t

     ( 1−X1)(1 + εAX1)

    (Xe1 −X1)(˛ + ˇX1)

    ×dX1 +(1 − X1,out)(1 + εAX1,out)

    (Xe1 −X1,out)(˛+ ˇX1,out)

    dX1,outdt

      (A9)

    As the catalyst lifetime is much larger than the residence time

    within the reactor, a pseudo-steady state condition can be

    assumed within the reactor by which the first term of Eq. (A6)

    is negligible. Consequently, Eq. (A6) is simplified to the below

    equation:

    dX2,outdt

      = AX2,out + k exp( At) ( 1−X1,out)(1 + εAX1,out)

    (Xe1 − X1,out)(˛ + ˇX1,out)

    dX1,outdt

    (A10)

  • 8/18/2019 Saheb Del Far 2012

    8/8

    chemical engineering researchand design 9 0 ( 2 0 1 2 ) 1090–1097 1097

    By rearranging Eq. (A10), one obtains:

    ln

      ((1/X2,out)(dX2,out/dt))− A

    (1/X2,out)(dX1,out/dt)(((1− X1,out)(1 + εAX1,out))/((Xe1 −X1,out)(˛+ ˇX1,out)))

    = At + ln(k) (A11)

    Similar approach can be used to obtain Eq. (20).

    References

    Anderson, J.R., Pratt, K.C., 1985. Introduction to Characterizationand Testing of Catalyst, second ed. Academic Press, Australia.

    Arora, V.K., 2004. Propylene via CATOFIN propanedehydrogenation technology. In: Meyers, R.A. (Ed.), Handbookof Petrochemicals Production Processes. McGraw-Hill, NewYork, pp. 10.43–10.49.

    Bai, L., Zhou, Y., Zhang, Y., Liu, H., Sheng, X., 2009. Effect of alumina binder on catalytic performance of Pt–Sn–Na/ZSM-5catalyst for propane dehydrogenation. Ind. Eng. Chem. Res.48, 9885–9891.

    Barbier, J.B., Marecot, P., Martin, N., Elassal, L., Maurel, R., 1980.Selective poisoning by coke formation on Pt/Al2O3. In:

    Delmon, B., Froment, G.F. (Eds.), Catalyst Deactivation.Elsevier, Amsterdam, pp. 53–62.

    Barias, O.A., Holmen, A., Blekkan, E.A., 1996. Propanedehydrogenation over supported Pt and Pt–Sn catalysts:catalyst preparation, characterization, and activitymeasurements. J. Catal. 158, 1–12.

    Bhasin, M.M., McCain, J.H., Vora, B.V., Imai, T., Pujado, R.R., 2001.Dehydrogenation and oxydehydrogenation of paraffins toolefins to olefins. Appl. Catal. A 221, 397–419.

    Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. TransportPhenomena, second ed. Wiley, USA.

    Bricker, J.C., Jan, D.-Y., Foresman, J.M., 1990. Dehydrogenationcatalyst composition. US Patent 4914075, UOP.

    Carvalho, L.S., Reyes, P., Pecchi, G., Figoli, N., Pieck, C.L., Rangel,M.C., 2001. Effect of the solvent used during preparation on

    the properties of Pt/Al2O3  and Pt–Sn/Al2O3 catalysts. Ind. Eng.Chem. Res. 40, 5557–5563.

    Chin, S.Y., Radzi, S.N.R., Maharon, I.H., Shafawi, M.A., 2011.Kinetic model and simulation analysis for propanedehydrogenation in an industrial moving bed reactor. WorldAcad. Sci. Eng. Technol. 76, 183–189.

    Heinritz-Adrin, M., Thiagarajan, N., Wenzel, S., Gehrke, H., 2003.STAR-Uhde’s dehydrogenation technology (an alternativeroute to C3 and C4 olefins). In: ERTC (European Refining Technology Conference) Petrochemical, Paris.

    Heinritz-Adrian, M., Wenzel, S., Youssef, F., 2008. Advancedpropane dehydrogenation. Pet. Technol. Q. 13, 83–91.

    Kogan, S.B., Herskowitz, M., 2001. Selective propanedehydrogenation to propylene on novel bimetallic catalysts.Catal. Commun. 2, 179–185.

    Kumbilieva, K., Gaidai, N.A., Nekrasov, N.V., Petrov, L., Lapidus,A.L., 2006. Types of active sites and deactivation features of promoted Pt catalysts for isobutane dehydrogenation. Chem.Eng. J. 120, 25–32.

    Larsson, M., Hulten, M., Blekkan, E.A., Andersson, B., 1996. Theeffect of reaction conditions and time on stream on the cokeformed during propane dehydrogenation. J. Catal. 164,44–53.

    Larese, C., Campos-Martin, J.M., Fierro, J.L.G., 2000. Alumina- andzirconia–alumina-loaded tin–platinum. Surface features andperformance for butane dehydrogenation. Langmuir 16,10294–10300.

    Levenspiel, O., 1999. Chemical Reaction Engineering, third ed.Wiley, New York.

    Lobera, M.P., Tellez, C., Herguido, J., Menendez, M., 2008.

    Transient kinetic modeling of propane dehydrogenation overa Pt–Sn–K/Al2O3   catalyst. Appl. Catal. A 349, 156–164.

    Miracca, I., Piovesan, L., 1999. Light paraffins dehydrogenation ina fluidized bed reactor. Catal. Today 52, 259–269.

    Moghimpour Bijani, P., Sahebdelfar, S., 2008. Modeling of aradial-flow moving-bed reactor for dehydrogenation of 

    isobutene. Kinet. Catal. 49, 625–632.Moulijn, J.A., van Diepen, A.E., Kapteijn, F., 2001. Catalyst

    deactivation: is it predictable? What to do. Appl. Catal. A 212,3–16.

    Padmavathi, G., Chaudhuri, K.K., Rajeshwer, D., Sreenivasa Rao,G., Krishnamurthy, K.R., Trivedi, P.C., Hathi, K.K.,Subramanyam, N., 2005. Kinetics of n-dodecanedehydrogenation on promoted platinum catalyst. Chem. Eng.Sci. 60, 4119–4129.

    Pujado, P., Vora, B., 1990. Make C3–C4 olefins selectivity.Hydrocarb. Process. – Int. Ed. 69, 65–70.

    Qing, L., Zhijun, S., Xinggui, Z., Chen, D., 2011. Kinetics of propane dehydrogenation over Pt–Sn/Al2O3  catalyst. Appl.Catal. A 398, 18–26.

    Resasco, D.E., 2003. Dehydrogenation – heterogeneous. In:

    Verbeek, I.K. (Ed.), Encyclopedia of Catalysis. Wiley, New York,pp. 49–79.

    Sahebdelfar, S., Tahriri Zangeneh, F., 2010. Dehydrogenation of propane to propylene over Pt–Sn/Al2O3  catalysts: theinfluence of operating conditions on product selectivity. Iran.

     J. Chem. Eng. 7, 51–58.Sahebdelfar, S., Moghimpour Bijani, P., Saeedizad, M., Tahriri

    Zangeneh, F., Ganji, K., 2011. Modeling of adiabaticmoving-bed reactor for dehydrogenation of isobutane toisobutene. Appl. Catal. A 395, 107–113.

    Sanfilippo, D., Miracca, I., 2006. Dehydrogenation of paraffins:synergies between catalyst design and reactor engineering.Catal. Today 111, 133–139.

    Scott Fogler, H., 1999. Elements of Chemical ReactionEngineering, third ed. Prentice-Hall, New York.

    Takehira, K., Ohishi, Y., Shishido, T., Kawabata, T., Takaki, K.,Zhang, Q., Wang, Y., 2004. Oxidative dehydrogenation of ethane with CO2 over novel Cr/SBA-15/Al2O3 /Fe–Cr–Almonolithic catalysts. J. Catal. 224, 404–416.

    Waku, T., Biscardi, J.A., Iglesia, E., 2003. Active, selective, andstable Pt/Na–[Fe]ZSM5 catalyst for the dehydrogenation of light alkanes. Chem. Commun. (14), 1764–1765.

    Waku, T., Biscardi, J.A., Iglesia, E., 2004. Catalyticdehydrogenation of alkanes on Pt/Na–Fe–ZSM5 and staged O2introduction for selective H2 removal. J. Catal. 222, 481–492.

    Weckhuysen, B.M., Schoonheydt, R.A., 1999. Alkanedehydrogenation over supported chromium oxide catalysts.Catal. Today 51, 223–232.

    Yu, C., Ge, Q., Xu, H., Li, W., 2006. Effects of Ce addition on thePt–Sn/gamma-Al2O3 catalyst for propane dehydrogenation to

    propylene. Appl. Catal. A 315, 58–67.Yu, C., Xu, H., Ge, Q., Li, W., 2007. Kinetics of propane

    dehydrogenation over Pt–Sn/Al2O3 catalyst. J. Mol. Catal. A:Chem. 266, 80–87.

    Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006a. Effectof Na addition on catalytic performance of Pt–Sn/ZSM-5catalyst for propane dehydrogenation. Acta Phys.-Chim. Sin.22, 672–678.

    Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006b. Effectof alumina binder on catalytic performance of Pt–Sn–Na/ZSM-5 catalyst for propane dehydrogenation. Ind.Eng. Chem. Res. 45, 2213–2219.

    Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006c.Propane dehydrogenation on Pt–Sn/ZSM-5 catalyst: effect of tin as a promoter. Catal. Commun. 7, 862–868.

    Zhang, Y., Zhou, Y., Liu, H., Wang, Y., Xu, Y., Wu, P., 2007. Effect of La addition on catalytic performance of Pt–Sn–Na/ZSM-5catalyst for propane dehydrogenation. Appl. Catal. A 333,202–210.