Sei sulla pagina 1di 8

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

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 PtSn/Al2 O3 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 PtSn/Al2 O3 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 xed-bed quartz reactor. The kinetic expressions developed for the main and side reactions were veried by integral and a combination of integraldifferential analysis of
reactor data, respectively, and the kinetic parameters were obtained. The deactivation of the active sites for the three
reactions was found to follow a rst-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; PtSn catalysts; Cracking; Hydrogenolysis; Catalyst deactivation; Kinetics

1.

Introduction

Propylene is an important feedstock for producing a variety of petrochemicals such as polypropylene, acrolein and
acrylic acid. Due to ever-increasing demand and insufcient
supply by crackers, alternative production methods such
as propane dehydrogenation (PDH) has received attention
(Heinritz-Adrian et al., 2008):
C3 H8 C3 H6 + H2 ,

H298 = +129 kJ/mol

(1)

The reaction is highly endothermic and equilibriumlimited; therefore, higher temperatures and lower pressures
are necessary to achieve acceptable conversions. Unfortunately, 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

for parafn dehydrogenation. Platinum exhibits high catalytic


activity in dehydrogenation of parafns. To achieve high platinum 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; Sanlippo
and Miracca, 2006; Yu et al., 2006; Zhang et al., 2006a)
C3 H8 C2 H4 + 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)
C3 H8 + H2 C2 H6 + CH4 ,

H298 = 63.4 kJ/mol

(3)

In commercial catalysts, Sn is used as promoter to suppress hydrogenolysis reaction through reduction of surface
Pt ensemble size by dividing Pt surface to smaller ensembles

Corresponding author. Tel.: +98 21 44580100; fax: +98 21 44580505.


E-mail addresses: m.ravanchi@npc-rt.ir, maryamravanchi@gmail.com (M.T. Ravanchi).
Received 15 April 2011; Received in revised form 20 October 2011; Accepted 4 November 2011
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

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

Nomenclature
a
C
F
ki
kdi
Keq
ri
W
Xi
Xei
t

catalyst activity
concentration (mol/m3 )
feed rate (mol/h)
forward rate constant of reaction i
rate constant for catalyst deactivation in reaction i (h1 )
concentration equilibrium constant (mol/m3 )
rate of reaction i per mass of catalyst
(mol/(kg h))
weight of catalyst in the reactor (kg)
conversion of key reactant A in reaction i
equilibrium conversion of key reactant A in
reaction i
time-on-stream (h)

Greek symbols

parameter dened by Eq. (9)

parameter dened by Eq. (10)

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 parafn

a capacity factor (catalyst weight per volumetric feed ow rate) ((kg h)/m3 )
Subscripts
0
reactor inlet
A
key reactant, propane
B
key product, propylene
hydrogen
H
i
reaction number
reactor outlet
out

studied elsewhere (Moghimpour Bijani and Sahebdelfar, 2008).


Qing et al. (2011) studied the kinetics of propane dehydrogenation, cracking and coke formation over PtSn/Al2 O3 catalyst,
assuming same activity for all sites. The distinction of active
sites as different types has also been proposed in the literature
(Kumbilieva et al., 2006).
Compared to the studies on the mechanism of propane
dehydrogenation including coke formation and the efforts to
improve the activity and stability of the catalyst, relatively few
efforts have been made to establish a complete and reliable
kinetic model of engineering signicance, including kinetics 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 PtSn/Al2 O3 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 supplied by Roham Gas Co. with purity 99.5 wt.%, 99.99 wt.% and
99.99 wt.%, respectively. Commercial PtSn/-Al2 O3 catalyst
(Pt = 0.58 wt.%, 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 supplied by Procatalyse company. Quartz powder with grain size
0.1 mm was used as catalyst diluent.

2.2.
(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 reactions still occur to a rather appreciable extent on the catalysts, so that the selectivity of the UOP
Oleex 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 CC 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, reactions (2) and (3) can represent the side reactions.
It has been shown that side reactions result in signicant
inuence on the performance of commercial-sized dehydrogenation reactors (Sahebdelfar et al., 2011). Unfortunately, it
is difcult 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 difculty in commercial scale implementation of parafn dehydrogenation is rapid catalyst
deactivation due to coke formation (Moulijn et al., 2001). During the course 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

1091

Set-up

The experimental setup, a schematic diagram of which is


depicted in Fig. 1, was a fully automated system. All lines
and ttings of the setup were made of stainless steel 316 (SS316). A tubular xed-bed quartz reactor with inner diameter of
15 mm was used, the temperature of which was controlled by
a furnace. Space above and below the catalyst bed (1.5 g) was
packed with quartz powder (3 g) to ensure proper distribution
of uid ow. 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 diameter) 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 rst heated under hydrogen ow of


18.3 Nml/min at about 5 C/min to 530 C to desorb the
adsorbed moisture, if any, and then the catalyst was reduced
in hydrogen ow at 530 C for 1 h 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 h1 . The gas products were analyzed

1092

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

Fig. 1 Schematic diagram of the experimental setup.


using online gas chromatography Agilent 6890N RGA, which
was equipped with a capillary column, HP-Plot Al2 O3 /Na2 SO4 ,
50 m, 530 m, 15.0 m 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 i
was calculated as:
Conversion by reaction i
=

moles of propane in moles of propane converted by reaction i


mole of propane in

Larsson et al. (1996) found that among several kinetic models tested for dehydrogenation of propane on PtSn/Al2 O3
catalysts, simple power-law model of this type resulted
in the best t. In fact most mechanistic kinetic models,
e.g. LongmuirHinshelwoodHougenWatson (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-parafn dehydrogenation practice.
To account for catalyst deactivation, the reaction rate can
be written in terms of site activity as:

(4)


r 1 = k1 a1

The test run times were comparable to one catalyst cycle


(57 days) in the Oleex process when the catalyst being sent to
continuous catalyst regeneration (CCR) unit for regeneration.

3.

Kinetic and reactor model

3.1.

Kinetic expressions


r 1 = k1 CA k1 CB CH2 = k1

CB CH2
CA
Keq


(5)

where r1 is the rate of conversion of propane to propylene (reaction (1)) per catalyst weight, k1 and k1 are the
rate constants of forward and backward reactions, respectively, Keq is the equilibrium constant at reaction temperature,
C is the concentration, and, A and B, respectively, representing the key components, propane and propylene.

CB CH2
Keq


(6)

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


as (Scott Fogler, 1999):
a1 (t) =

Most previous studies have shown that dehydrogenation reaction is rst-order in parafn concentration and negative half
to zeroth-order in hydrogen concentration (Resasco, 2003).
Therefore to incorporate the rst-order dependence on parafn concentration, hydrogen-inhibition effect, and chemical
equilibrium limitation, the following expression is assumed
to represent kinetics of the reaction:

CA

r 1 (t)
r 1 (t = 0)

(7)

Considering volume change of reaction, following necessary algebraic manipulations, one arrives at the following
expression for reaction rate in terms of propane conversion
(Moghimpour Bijani and Sahebdelfar, 2008):
r 1 =

k1 a1 CA0 (Xe1 X1 )( + X1 )
(1 + A X1 )

(8)

in which
=

Xe1 + (H + B ) + H B (1 A + A Xe1 )


(H + Xe1 )(B + Xe1 )

(1 + A Xe1 Xe1 ) + A Xe1 (H + B ) + A H B


(H + Xe1 )(B + Xe1 )

(9)

(10)

where H , B and CA0 are hydrogen/parafn, olen/parafn


molar ratios and propane concentration in the feed, respectively, X1 is propane conversion to propylene, Xe1 is the

1093

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

equilibrium conversion under reaction conditions and A is


the volume expansion factor.
Similarly, assuming reaction (1) as the main propane
consuming reaction governing propane and hydrogen concentrations 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:
r 2 = k2 a2 CA0

r 3 = k3 a3 C2A0

(1 X1 )
(1 + A X1 )

(1 + A X1 )

where kdi is the respective rate constant of deactivation. Integrating this equation yields a correlation for ai as a function
of time, t:
ln ai = kdi t

(14)

Reactor model

The plug-ow reactor performance equation for reactions in


parallel is:
dW
dXi
=
FA0
r i

(15)


ln

ln


1

( A )
ln
( + Xe1 )
2

(1 + A Xe1 )

ln
( + Xe1 )

20

40

 + X

1,out

Xe1


2A X1,out

60

80

100

120

Conversion, %

Time-on-stream, h

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


propylene (), ethylene () and ethane () (T = 620 C,
H2 /HC = 0.6 mol/mol and WHSV = 2.2 h1 ).

The rate of reactions (2) and (3) depends on reaction (1) as


implied by Eqs. (11) and (12). Dividing Eq. (15) for reactions (2)
and (3) to that for reaction (1), and using Eq. (14) for deactivation terms, one obtains respectively;
(Xe1 X1 )( + X1 ) dX2
k2
=
exp((kd1 kd2 )t)
k1
(1 X1 )(1 + A X1 ) dX1

(17)

(Xe1 X1 )( + X1 ) dX3
CA0 k3
=
exp((kd1 kd2 )t)
k1
(1 X1 )(H + X1 ) dX1

(18)

Eqs. (17) and (18) hold for any point along the reactor.
Rearranging, integrating with respect to Xi and then differentiating with respect to t, the following correlations are
obtained for cracking and hydrogenolysis reactions, respectively:

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


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

X X

e1
1,out

= kd1 t + ln(k1   )

20

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


(1/X2,out )(dX1,out /dt)(((1 X1,out )(1 + A X1,out ))/((Xe1 X1,out )( + X1,out )))

As the propane and hydrogen concentration proles are


largely determined by the main reaction (Eq. (1)), its extent 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):

30

10

(12)

(13)

40

(1 X1 )(H + X1 )

ln

50

dai
= kdi ai
dt

3.2.

60

(11)

The deactivation rates could be considered as rst-order


and independent. Consequently, for reaction i:

70


= (kd1 kd2 )t + ln


= (kd1 kd3 )t + ln

k 
2

k1

k C 
3 A0
k1

(19)

(20)

Interested reader is referred to Appendix for detailed mathematical derivations.


The derivative terms could be obtained numerically from
time-on-stream conversion data for different reactions. Similarly, plots of Eqs. (19) and (20) should result in straight lines
with the slope giving the difference of deactivation rate constants 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.

(16)

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


of the feed to reactor and   , the ratio of catalyst weight
per volumetric feed ow rate, is a capacity factor known
as the weight-time. The subscript out refers to reactor outlet value of the parameter. Plots of Eq. (16) should result in
straight lines the slope of which giving kd1 and the intercept
giving k1 .

4.

Results and discussion

4.1.

Performance test results

Fig. 2 shows the overall conversion and conversion of individual reactions versus time-on-stream. The rst data points
showing the initial activity, characterized by large deviations, are never actually observed due to experimental

1094

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

Time-on-stream, h

y = 0.0439x + 2.4186
R = 0.9516

4
3.5

2
1.5

y = 0.063x - 0.2411
R = 0.9303

1
0.5

60

80

100

120

y = -0.0159x - 0.2855
R2 = 0.9543

-1
-1.5
-2

10

20

30

40

-2.5

Conversion to propylene, %

limitations. After a sharp initial decline in side reactions


accompanied 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 shows that 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, especially in early time-on-streams.
The fact that extrapolation of the line for hydrogenolysis in
Fig. 3 passes close the origin could be attributed to the fact that
both dehydrogenation and hydrogenolysis reactions occur on
platinum sites. This is not the case for cracking reaction which
occurs on different sites and also thermally. From Fig. 3 one
concludes that cracking reaction could proceed even when the
catalyst is fully deactivated and that more than half 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.

Modeling results

Fig. 4 shows a plot of Eq. (16) for a long-term run using experimental data with LHS showing the left-hand-side of that
equation. A favorable t is observed. The slope and intercept
give the deactivation and reaction rate constants for dehydrogenation reaction, respectively (Table 1).
In this way Eq. (16) provides a method to obtain timezero 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

ki
3

4.7 m /(kg h)
0.40 m3 /(kg h)
0.023 m6 /(mol kg h)

more accurate experimental data because of the appearance of time-derivative terms in these equations. To avoid
the uctuations encountered in numerical differentiation, it
has been proposed to t the data with an appropriate function and then differentiate the resulted interpolating function
(Levenspiel, 1999). The trends of time-data propose exponential functions as good candidates (Fig. 2) which in fact result
in fair ts. Figs. 5 and 6, respectively, show plots of Eqs. (19)
and (20) obtained by this approach after achieving convergence of deactivation rate constant difference terms by the
iterative procedure explained above, with LHS showing the
left-hand-side of these equations. Favorable ts 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 those
of 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 selectivity 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 deactivation 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
Time-on-stream, h
0

20

40

60

80

100

0
-0.5
-1
-1.5
-2
-2.5

y = 0.01318x - 2.45327
R = 0.97686

-3

Table 1 Calculated values of the rate constants


(T = 620 C).
Reaction no.

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


(T = 620 C, H2 /HC = 0.6 mol/mol, WHSV = 2.2 h1 ).

LHS of Eq. 19

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


() versus conversion to propylene at different
time-on-streams (T = 620 C, H2 /HC = 0.6 mol/mol and
WHSV = 2.2 h1 ).

1
2
3

40

-0.5

2.5

4.2.

20

LHS of Eq. 16

Conversion to by-products, %

4.5

-3.5

kdi
1

0.016 h
0.0027 h1
0.011 h1

-4

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


kd1 kd2 = 0.0132 (T = 620 C, H2 /HC = 0.6 mol/mol and
WHSV = 2.2 h1 ).

120

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

Time-on-stream, h
-2

20

40

60

80

100

120

-2.5
LHS of Eq. 20

-3
-3.5
-4

y = 0.0049x - 3.3575
R = 0.8367

-4.5

1095

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 reactions, 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 calculated
from these data and reported in Table 1 have a reasonable
accuracy.

-5
-5.5
-6

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


kd1 kd3 = 0.0049 (T = 620 C, H2 /HC = 0.6 mol/mol and
WHSV = 2.2 h1 ).

better ts for apparent deactivation of 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 partially deactivated catalyst will 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. 7bd) are
presented. These plots compare the calculated conversions
versus experimental conversions. As it can be seen, generally, the difference between experimental results and model
estimation is within 20% which conrms the accuracy of the
results.
It is noteworthy that in constructing these plots, the main
assumption of modeling (i.e. predominance of dehydrogenation reaction in propane consumption) is not applied to have

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 and differential analyses the accuracy of which is limited by the latter (that is, by time derivatives of conversions
data). As mentioned above, an approach is to t experimental
conversion data with an appropriate function and then differentiate the resulted tting 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 decrease of conversions within the specied step-size time
interval becomes smaller than that of experimental accuracy
and/or system disturbances. Therefore, using direct numerical differentiation formulas, these errors largely mask the true
value of the derivatives. The function should be checked by
the eye to give both close t to data and relevant slopes. The
exponential function is the simplest one to satisfy both these
requirements fairly. However, no simple function can t both
data points and their derivatives over a long range. Consequently, a small downward curvature is observed in plots of
Eqs. (19) and (20) with the exponential functions employed
for tting (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.

1096

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

not result in satisfactory values when applied to the whole of


time-on-stream domain, due to considerable uctuations 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 reactant (e.g. through pyrolysis reaction) and products (through
oligomerizationaromatization of the olenic products e.g. of
reactions (1) and (2)) can bring about coke formation and
catalyst deactivation (Qing et al., 2011). The independent deactivation 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 reactions due to rather low reaction temperature compared to the
melting point of Pt do not favor deactivation by sintering during 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 ethylene to ethane ratio in the product is much higher than
the equilibrium ratio implying that there is thermodynamic
driving force for hydrogenation of ethylene to ethane. However, 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 sufciently high to contribute an important role in selectivity to ethane and ethylene.
Consequently, ethane and ethylene production rates can be
adequately considered as measures for the rate of hydrogenolysis 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 selectivity approaches to unity. The propylene selectivities in the
data series employed in this work were mostly within the
range 8085%. This range is still sufciently large such that
the main reaction determines the concentrations of propylene
and hydrogen within the reactor. Therefore, the kinetic parameters obtained should be accurate to at least one signicant
gure.
Higher selectivities are not uncommon both on lab or commercial 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 PtSn/Al2 O3 catalyst were obtained when the
reactions occur simultaneously. Power law expressions and
rst order independent decay laws tted the kinetic data
of the reactions favorably. The rate constant of the main
reaction was found to be more than one order of magnitude larger than those of cracking and hydrogenolysis side
reactions. On the other hand, the rate constant of the deactivation 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 ndings 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 )( + X1 ) dX2
= k exp(At)
(1 X1 )(1 + A X1 ) dX1

(A1)

where
k=

k2
k1

(A2)

A = kd1 kd2

(A3)

Rearranging Eq. (A1), one obtains:


dX2 = k exp(At)

(1 X1 )(1 + A X1 )
dX1
(Xe1 X1 )( + X1 )

(A4)

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

X1,out

X2,out = k exp(At)
0

(1 X1 )(1 + A X1 )
dX1
(Xe1 X1 )( + X1 )

(A5)

Differentiating this equation with respect to t, gives:


dX2,out
= kA exp(At)
dt

+ k exp(At)
t

X1,out

X1,out

(1 X1 )(1 + A X1 )
dX1
(Xe1 X1 )( + X1 )

(1 X1 )(1 + A X1 )
dX1
(Xe1 X1 )( + X1 )

(A6)

in which, the rst term on the right hand side is AX2,out (according to Eq. (A5)). In the second term, as differentiation is with
respect to t and the integral limits are functions of time, the
Leibnizs rule (Bird et al., 2002) must be applied. According to
this rule, for function F(x, t), where

b(t)

F(x, t) =

f (x, t)dx

(A7)

a(t)

the time derivative is:


dF
=
dt

f
db
da
dx + f (b, t)
f (a, t)
t
dt
dt

(A8)

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


k exp(At)
0

dX1 +

X1,out

(1 X1 )(1 + A X1 )
(Xe1 X1 )( + X1 )

(1 X1,out )(1 + A X1,out ) dX1,out


(Xe1 X1,out )( + X1,out ) dt

(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 rst term of Eq. (A6)
is negligible. Consequently, Eq. (A6) is simplied to the below
equation:
(1 X1,out )(1 + A X1,out ) dX1,out
dX2,out
= AX2,out + k exp(At)
dt
(Xe1 X1,out )( + X1,out ) dt
(A10)

chemical engineering research and design 9 0 ( 2 0 1 2 ) 10901097

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 + A X1,out ))/((Xe1 X1,out )( + X1,out )))

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

References
Anderson, J.R., Pratt, K.C., 1985. Introduction to Characterization
and Testing of Catalyst, second ed. Academic Press, Australia.
Arora, V.K., 2004. Propylene via CATOFIN propane
dehydrogenation technology. In: Meyers, R.A. (Ed.), Handbook
of Petrochemicals Production Processes. McGraw-Hill, New
York, pp. 10.4310.49.
Bai, L., Zhou, Y., Zhang, Y., Liu, H., Sheng, X., 2009. Effect of
alumina binder on catalytic performance of PtSnNa/ZSM-5
catalyst for propane dehydrogenation. Ind. Eng. Chem. Res.
48, 98859891.
Barbier, J.B., Marecot, P., Martin, N., Elassal, L., Maurel, R., 1980.
Selective poisoning by coke formation on Pt/Al2 O3 . In:
Delmon, B., Froment, G.F. (Eds.), Catalyst Deactivation.
Elsevier, Amsterdam, pp. 5362.
Barias, O.A., Holmen, A., Blekkan, E.A., 1996. Propane
dehydrogenation over supported Pt and PtSn catalysts:
catalyst preparation, characterization, and activity
measurements. J. Catal. 158, 112.
Bhasin, M.M., McCain, J.H., Vora, B.V., Imai, T., Pujado, R.R., 2001.
Dehydrogenation and oxydehydrogenation of parafns to
olens to olens. Appl. Catal. A 221, 397419.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport
Phenomena, second ed. Wiley, USA.
Bricker, J.C., Jan, D.-Y., Foresman, J.M., 1990. Dehydrogenation
catalyst 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/Al2 O3 and PtSn/Al2 O3 catalysts. Ind. Eng.
Chem. Res. 40, 55575563.
Chin, S.Y., Radzi, S.N.R., Maharon, I.H., Shafawi, M.A., 2011.
Kinetic model and simulation analysis for propane
dehydrogenation in an industrial moving bed reactor. World
Acad. Sci. Eng. Technol. 76, 183189.
Heinritz-Adrin, M., Thiagarajan, N., Wenzel, S., Gehrke, H., 2003.
STAR-Uhdes dehydrogenation technology (an alternative
route to C3 and C4 olens). In: ERTC (European Rening
Technology Conference) Petrochemical, Paris.
Heinritz-Adrian, M., Wenzel, S., Youssef, F., 2008. Advanced
propane dehydrogenation. Pet. Technol. Q. 13, 8391.
Kogan, S.B., Herskowitz, M., 2001. Selective propane
dehydrogenation to propylene on novel bimetallic catalysts.
Catal. Commun. 2, 179185.
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, 2532.
Larsson, M., Hulten, M., Blekkan, E.A., Andersson, B., 1996. The
effect of reaction conditions and time on stream on the coke
formed during propane dehydrogenation. J. Catal. 164,
4453.
Larese, C., Campos-Martin, J.M., Fierro, J.L.G., 2000. Alumina- and
zirconiaalumina-loaded tinplatinum. Surface features and
performance for butane dehydrogenation. Langmuir 16,
1029410300.
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 over
a PtSnK/Al2 O3 catalyst. Appl. Catal. A 349, 156164.
Miracca, I., Piovesan, L., 1999. Light parafns dehydrogenation in
a uidized bed reactor. Catal. Today 52, 259269.


= At + ln(k)

(A11)

Moghimpour Bijani, P., Sahebdelfar, S., 2008. Modeling of a


radial-ow moving-bed reactor for dehydrogenation of
isobutene. Kinet. Catal. 49, 625632.
Moulijn, J.A., van Diepen, A.E., Kapteijn, F., 2001. Catalyst
deactivation: is it predictable? What to do. Appl. Catal. A 212,
316.
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-dodecane
dehydrogenation on promoted platinum catalyst. Chem. Eng.
Sci. 60, 41194129.
Pujado, P., Vora, B., 1990. Make C3C4 olens selectivity.
Hydrocarb. Process. Int. Ed. 69, 6570.
Qing, L., Zhijun, S., Xinggui, Z., Chen, D., 2011. Kinetics of
propane dehydrogenation over PtSn/Al2 O3 catalyst. Appl.
Catal. A 398, 1826.
Resasco, D.E., 2003. Dehydrogenation heterogeneous. In:
Verbeek, I.K. (Ed.), Encyclopedia of Catalysis. Wiley, New York,
pp. 4979.
Sahebdelfar, S., Tahriri Zangeneh, F., 2010. Dehydrogenation of
propane to propylene over PtSn/Al2 O3 catalysts: the
inuence of operating conditions on product selectivity. Iran.
J. Chem. Eng. 7, 5158.
Sahebdelfar, S., Moghimpour Bijani, P., Saeedizad, M., Tahriri
Zangeneh, F., Ganji, K., 2011. Modeling of adiabatic
moving-bed reactor for dehydrogenation of isobutane to
isobutene. Appl. Catal. A 395, 107113.
Sanlippo, D., Miracca, I., 2006. Dehydrogenation of parafns:
synergies between catalyst design and reactor engineering.
Catal. Today 111, 133139.
Scott Fogler, H., 1999. Elements of Chemical Reaction
Engineering, 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/Al2 O3 /FeCrAl
monolithic catalysts. J. Catal. 224, 404416.
Waku, T., Biscardi, J.A., Iglesia, E., 2003. Active, selective, and
stable Pt/Na[Fe]ZSM5 catalyst for the dehydrogenation of
light alkanes. Chem. Commun. (14), 17641765.
Waku, T., Biscardi, J.A., Iglesia, E., 2004. Catalytic
dehydrogenation of alkanes on Pt/NaFeZSM5 and staged O2
introduction for selective H2 removal. J. Catal. 222, 481492.
Weckhuysen, B.M., Schoonheydt, R.A., 1999. Alkane
dehydrogenation over supported chromium oxide catalysts.
Catal. Today 51, 223232.
Yu, C., Ge, Q., Xu, H., Li, W., 2006. Effects of Ce addition on the
PtSn/gamma-Al2 O3 catalyst for propane dehydrogenation to
propylene. Appl. Catal. A 315, 5867.
Yu, C., Xu, H., Ge, Q., Li, W., 2007. Kinetics of propane
dehydrogenation over PtSn/Al2 O3 catalyst. J. Mol. Catal. A:
Chem. 266, 8087.
Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006a. Effect
of Na addition on catalytic performance of PtSn/ZSM-5
catalyst for propane dehydrogenation. Acta Phys.-Chim. Sin.
22, 672678.
Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006b. Effect
of alumina binder on catalytic performance of
PtSnNa/ZSM-5 catalyst for propane dehydrogenation. Ind.
Eng. Chem. Res. 45, 22132219.
Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, Y., Wu, P., 2006c.
Propane dehydrogenation on PtSn/ZSM-5 catalyst: effect of
tin as a promoter. Catal. Commun. 7, 862868.
Zhang, Y., Zhou, Y., Liu, H., Wang, Y., Xu, Y., Wu, P., 2007. Effect of
La addition on catalytic performance of PtSnNa/ZSM-5
catalyst for propane dehydrogenation. Appl. Catal. A 333,
202210.

Potrebbero piacerti anche