Vous êtes sur la page 1sur 13

This article was downloaded by: [Istanbul Universitesi Kutuphane ve Dok]

On: 20 December 2014, At: 09:27


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Petroleum Science and Technology


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lpet20

Modeling of Coke Formation and


Catalyst Deactivation in Propane
Dehydrogenation Over a Commercial PtSn/-Al2O3 Catalyst
a

S. Niknaddaf , M. Soltani , A. Farjoo & F. Khorasheh

Department of Chemical and Petroleum Engineering , Sharif


University of Technology , Tehran , Iran
b

National Iranian Petrochemical Company, Research and Technology


Division , Tehran , Iran
Published online: 30 Oct 2013.

To cite this article: S. Niknaddaf , M. Soltani , A. Farjoo & F. Khorasheh (2013) Modeling of Coke
Formation and Catalyst Deactivation in Propane Dehydrogenation Over a Commercial Pt-Sn/-Al2O3
Catalyst, Petroleum Science and Technology, 31:23, 2451-2462, DOI: 10.1080/10916466.2011.565296
To link to this article: http://dx.doi.org/10.1080/10916466.2011.565296

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the
Content) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/termsand-conditions

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

Petroleum Science and Technology, 31:24512462, 2013


Copyright Taylor & Francis Group, LLC
ISSN: 1091-6466 print/1532-2459 online
DOI: 10.1080/10916466.2011.565296

Modeling of Coke Formation and Catalyst Deactivation


in Propane Dehydrogenation Over a Commercial
Pt-Sn/ -Al2 O3 Catalyst
S. Niknaddaf,1 M. Soltani,2 A. Farjoo,1 and F. Khorasheh1
1

Department of Chemical and Petroleum Engineering, Sharif University of Technology,


Tehran, Iran
2
National Iranian Petrochemical Company, Research and Technology Division,
Tehran, Iran

Propane dehydrogenation was carried over a commercial Pt-Sn/ -Al2 O3 catalyst at atmospheric
pressure and reaction temperatures of 580, 600, and 620C and WHSV of 11 h 1 in an experimental
tubular quartz reactor. Propane conversions were measured for catalyst time on stream of up to nine
days. The amounts of coke deposited on the catalyst were measured after one, three, six, and nine
days on stream using a thermogravimetric differential thermal analyzer (TG-DTA) for each reaction
temperature. The coke formation kinetics was successfully described by a coke formation model
based on a monolayer-multilayer mechanism. In addition, catalyst deactivation was presented by a
time-dependant deactivation function. The kinetic order for monolayer coke formation was found to
be two, which would support a coke formation step involving two active sites. The kinetic order for
multilayer coke formation was found to be zero. The activation energy for monolayer coke formation
was found to be 29.1 kJ/mol, which was lower than the activation energy of about 265.1 kJ/mol for
multilayer coke formation indicating that the presence of metals can promote coke formation on the
catalyst surface.
Keywords: catalyst deactivation, coke formation, kinetics, mathematical modeling, propane dehydrogenation

1. INTRODUCTION
The predicted annual extensive demand of propylene will increase by 5% (Corma et al., 2005) and
is anticipated to reach 75 million tons per year in 2010 (Bhasin et al., 2001). The dehydrogenation
of paraffins has a significant role in the production of monomers for synthetic rubber and synthesis
of other organic products. The increasing availability of lower alkanes resulting from more strict
environmental legislation has made dehydrogenation processes economically attractive (Resasco
and Haller, 1994) and intensive work has been done to develop catalytic dehydrogenation processes
for light alkanes, especially for propane and isobutene, on platinum-based catalysts (Assabumrungrat et al., 2000; Herguido et al., 2005). Dehydrogenation reactions of lower alkanes are
Address correspondence to F. Khorasheh, Department of Chemical and Petroleum Engineering, Sharif University of
Technology, Azadi Avenue, 113659465, Tehran, Iran. E-mail: khorashe@sharif.ir

2451

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

2452

S. NIKNADDAF ET AL.

highly endothermic and because of thermodynamic constraints are carried out at high temperatures
generally in the range of 530630C using chromium (Gascn et al., 2003), and more recently
due to environmental considerations by platinum-based catalysts (Lobera et al., 2008). At these
high temperatures carbonaceous deposits, collectively termed coke, are rapidly formed and as a
consequence catalyst deactivation and regeneration are important considerations for commercial
processes. Catalyst deactivation in propane dehydrogenation over alumina supported platinum
catalysts has received considerable attention in the past few years (Baris et al., 1996; Larsson
et al., 1996). Coke deposition may cause a decrease in catalytic activity by either site coverage
or pore blockage, or both.
The aim of this work was to develop a kinetic rate expression to quantitatively describe the
formation of coke deposits on a commercial Pt-Sn/ -Al2O3 catalyst in propane dehydrogenation as
a function of time-on-stream and to develop a time-dependent relationship for catalyst deactivation.

2. METHODS AND MATERIALS


A laboratory scale reactor set up was used for the propane dehydrogenation (PDH) experiments.
A simplified schematic of the setup is shown in Figure 1. Using a three-lines setup, propane,
hydrogen, and nitrogen gas mixtures could be introduced whose composition and total flow was
set using Brooks mass flow controllers. The inside diameter of the reactor was 12 mm, the length
of the reactor was 90 cm, and the catalyst was loaded in the middle section of the reactor in
between two layers of quarts particles. The height of the catalyst zone was approximately 2 cm.
The specifications of the catalyst are given in Table 1.

FIGURE 1

The simplified sketch of propane dehydrogenation setup.

COKE FORMATION AND CATALYST DEACTIVATION

2453

TABLE 1
Specifications of the Commercial Catalyst for
Propane Dehydrogenation

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

1.8 mm
0.65 (gr/cm3 )
1.12 (gr/cm3 )
200 (m2 /gr)

Diameter
Bulk Density, b
Catalyst Density, c
Catalyst Surface Area, Sa

In each experiment the catalyst sample was heated under nitrogen flow from ambient temperature to 150C in 25 min. The temperature was then held constant for 3 hours. The catalyst was
then heated from 150C to the desired reduction temperature of 450 C in 3 h under hydrogen flow.
During this period nitrogen flow was decreased gradually such that when the catalyst temperature
reached 450 C, the nitrogen flow was zero. The temperature was then increased from 450 C
to 530 C in 16 min under hydrogen flow only. The temperature was then held constant for 1
h at 530C. The temperature was then reduced to 350C in 3 h under hydrogen flow at which
point the flow of propane was started. The feed to the reactor consisted of propane and hydrogen
with H2 /propane molar ratio of 0.8. The temperature was then increased to the desired reaction
temperatures of 580, 600, or 620C at a rate of 5 C per minute. The reactor temperature was
measured by three thermocouples externally attached to the top, middle, and bottom of the reactor,
and one thermocouple located inside the reactor at the bottom of the catalyst bed. Product gases
were analyzed after a 90-min stabilization period. The outlet stream from the reactor was analyzed
by an online gas chromatograph (model PERICHROM 2100 Packed column, SS316, 6m, 1/8 in,
28% DC200 on Chromsorb PAW 60/80, ENRO 3015) equipped with thermal conductivity and
flame ionization detectors to calculate the propane conversions. Dehydrogenation of propane was
carried out at a weight hourly spaced velocity (WHSV) of 11 h 1 . To determine the amount of
coke deposited on the catalyst, propane dehydrogenation experiments at each reaction temperature
were performed four times terminating the reaction after one, three, six, and nine days. For
each catalyst sample, the amount of coke deposited on the catalyst was measured by thermal
gravimetric/differential thermal analysis (TG/DTA) using a Dimond-Perkin Elmer apparatus with
Chan Win DRB software.
3. RESULTS AND DISCUSSIONS
To obtain proper kinetic data for determination of intrinsic coke formation kinetics, it was
necessary to perform experiments in the absence of external and internal mass transfer limitations.
Experiments were conducted using catalyst particles of different sizes to investigate internal mass
transfer limitations. Feed flow rates were also varied over a wide range keeping WHSV constant
to investigate external mass transfer limitations under reaction conditions.
3.1. External Mass Transfer Limitations
In order to eliminate the external mass transfer limitations several experiments were conducted in
which the size of the catalyst pellet was kept constant and the feed flow rate was varied over a
wide range while keeping WHSV constant at 4 h 1 . If conversion was unaffected by varying the
propane flow rate at constant WHSV, it could be concluded that the reaction was independent of
the external mass transfer of the fluid outside the catalyst pellet. The results of these experiments
are shown in Table 2 indicating that when propane flow rate, Qc3 , was greater than 44.36 cc/min,
there was no external mass transfer limitations.

2454

S. NIKNADDAF ET AL.

TABLE 2
Propane Conversion at Different Feed Flow Rates

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

Run
1
2
3
4

Catalyst
Weight, g

QC3 ,
cc/min

QH2 ,
cc/min

WHSV,
1/h

Conversion
%

1
1.2
1.5
1.6

39.96
44.36
55.45
59.3

31.968
35.488
44.36
47.44

4
4
4
4

20.45
23.07
23.32
23.35

3.2. Internal Mass Transfer Limitations


In order to eliminate internal mass transfer limitations, several experiments were performed using
different sizes of catalyst pellets at a feed flow rate where external mass transfer limitations were
absent (QC3 > 55:45 cc/min). The results are presented in Table 3. If the conversion did not
change by further reduction in the size of the catalyst pellet, there would be no internal mass
transfer limitations. Subsequent experiments were performed using catalysts pellets of 0.35 mm
in diameter.
3.3. Coke Formation
Figure 2 illustrates the amounts of coke (as % weight of catalyst) deposited on the catalyst
surface versus time-on-stream for different reaction temperatures. The results clearly indicate an
increase in coke formation with increasing reaction temperature in accordance with coking as an
endothermic reaction and with positive activation energy. The coke content increased sharply with
time-on-stream during an initial stage followed by a more moderate linear increase during the
later stage of the reaction. The coke content for both the initial and final stages of the reaction
increased with increasing temperature. These results suggest that in the initial stage the coke
deposition itself inhibits further coke formation and after this stage, a nearly constant residual
coking activity is attained in the later stage. Similar trends have been observed by Gascn et al.
(2003), Pea et al. (1993a, 1993b), and van Sint et al. (2001).
3.4. Propane Conversion
Figure 3 shows the effect of reaction temperature on propane conversion. For each reaction
temperature, a decrease in conversion was observed with time-on-stream. The increase in catalyst
coke content and the decrease in propane conversion with time-on-stream suggest that coke
TABLE 3
Propane Conversion for Catalyst Pellets of Different Sizes

Run
1
2
3
4
5

Catalyst
Weight, g

QC3 ,
cc/min

Catalyst
Diameter, mm

QH2 ,
cc/min

WHSV,
1/h

Conversion
%

0.75
0.75
0.75
0.75
0.75

55
55
55
55
55

1.8
1
0.71
0.50
0.35

44
44
44
44
44

9
9
9
9
9

13.72
23.1
23.9
24.1
24.3

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

COKE FORMATION AND CATALYST DEACTIVATION

2455

FIGURE 2 Amount of coke (% weight of catalyst) versus time-on-stream at different reaction temperatures.
(color figure available online)

accumulation on the catalyst surface and the resulting active site blockage is responsible for
the decrease in propane conversion. Figure 3 also demonstrates that the initial propane conversion
increases with increasing reaction temperature. The decrease in propane conversion with timeon-stream, however, is also more pronounced at higher reaction temperature at higher reaction
temperatures as coke formation leading to a loss in catalyst activity is enhanced at higher
temperatures. In other words, higher reaction temperatures result in both higher initial catalyst
activity as well as faster loss of activity. These results are in good accordance with those reported
by Lobera et al. (2008).
3.5. Effect of Time on Stream on Side Reaction Side Products
Figure 4 illustrates the molar flow rates of reaction side products, namely, methane, ethylene, and
ethane, at the reactor exit with time-on-stream. While the flow rates for methane and ethylene
increased with time-on-stream, those for ethane decreased. This observation could be explained

FIGURE 3
online)

Propane conversion versus time-on-stream at different reaction temperatures. (color figure available

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

2456

S. NIKNADDAF ET AL.

(a)

(b)

(c)
FIGURE 4 Molar flow rates of (a) ethane, (b) methane, and (c) ethylene at reactor exit versus time-on-stream
at different reaction temperatures. (color figure available online)

COKE FORMATION AND CATALYST DEACTIVATION

2457

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

by the fact that hydrogenolysis of propane leading to the formation of ethane and methane, and
hydrogenation of ethylene to ethane are side reactions that could take place on the platinum active
sites (Zhang et al., 2006). The catalyst deactivation by active site blockage due to coke formation
would lead to a decrease in the formation of ethane as a side product. The main cracking products
(ethylene and methane), on the other hand, are mainly formed by cracking on the catalyst support
and to some extent, by thermal cracking, that are unaffected by the deactivation of the catalyst.
3.6. Kinetic Modeling
3.6.1. Coke Formation Model
The kinetic model used to describe the coke formation on the catalyst surface was based on
the monolayermultilayer coke growth model (MMCGM) that is a simple mechanistic model
first proposed by Nam and Kittrell (1984), and generalized and used successfully by various
investigators including Monzn et al. (1999), Pea et al. (1993a, 1993b), Rodrguez et al. (1997),
and Romero et al. (1996). In this model, the rate of coke formation on the catalyst surface with
time-on-stream is given by the sum of coke formation rates for monolayer coke which is formed
on the surface of the catalyst and the simultaneous multilayer coke deposition rate:
dCm
dCM
dCC
D
C
dt
dt
dt

(1)

For the monolayer coke, the rate of coke deposition on the surface of catalyst is related to the
fraction of sites on the first layer which are not covered by carbonaceous species:
rC m D

dCm
D k1 .Cmax
dt

Cm /m

(2)

where Cm is the monolayer coke concentration, Cmax is the maximum coke concentration in
monolayer, k1 is a kinetic coefficient, and m is the kinetic order pertinent to monolayer coke
formation. The multilayer coke formation initiate as soon as the monolayer coke is available and
is related to sites that are covered by monolayer coke:
rCM D

dCM
D k2 Cmn
dt

(3)

where CM is the multilayer coke concentration, k2 is a kinetic coefficient, and n is the kinetic
order pertinent to coke formation in multilayer. It is assumed that kinetic constants k1 and k2 are
functions of temperature only and an Arrhenius-type relation is used for both monolayer .i D 1/
and multilayer coke .i D 2/ formation:



EaiC 1
1
ki D koi exp
(4)
R
T
T0
where Eai c and T0 D 873 K are the activation energy for coke formation and the average
temperature at which experimental data were obtained, respectively.
The proposed coke formation rate equations can be evaluated for different sets of parameters
m and n related to the kinetic order of coke formation for monolayer and multilayer coke,
respectively. The integrated form of the model equations are presented in Table 4 for three different
sets of values for m and n. Experimental data for amounts of coke deposited on catalyst surface for
different reaction temperatures and time-on-streams were used to estimate the model parameters by

2458

S. NIKNADDAF ET AL.

TABLE 4
Kinetic Models for Coke Formation and Catalyst Deactivation

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

Model
Coke Formation

Model Equation

A1 .n D 1; m D 1/

dCC
D k1 .Cmax
dt

Cm / C k2 Cm

A2 .n D 0; m D 2/

dCC
D k1 .Cmax
dt

Cm /2 C k2

A3 .n D 1; m D 2/

dCC
D k1 .Cmax
dt

Cm /2 C k2 Cm

Integrated Equation

k1 k2
CC D Cmax
1 e. k1 t/ C k2 Cmax t
k1


k1 t
2
CC D Cmax
C k2 t
1 C Cmax k1 t


2
Cc D Cmax

k1 t
1 C Cmax k1 t

k2
ln1 C k1 Cmax t
k1

C k2 Cmax t
ki D k0i exp
Deactivation

EaiC
R

1
T

1
T0



Model Equation
a D exp. t p /
rA D a.time  Temperature/  rA0



1
1
 D 0 exp . Ea =R/
T
T0

minimizing the following objective function using the fmin search function in Matlab optimization
toolbox:
O:F D

N
X

.CW model

CW exp/

(5)

j D1

where CW model and CW exp are the predicted and experimental wt% of coke deposited on the
catalyst surface, respectively, and the summation index j extends for all experimental data points
.N D 12/. The optimal values of the objective function for each case are reported in Table 5
indicating that among the three models, model A2 with corresponding values of m D 2 and n D 0
would best describe the experimental amounts of coke deposited on the catalyst surface. For this
model, the optimized parameter values, namely the Arrhenius parameters for kinetic constants, k1
and k2 , and Cmax are presented in Table 6 and as indicated by Figure 5, there is a good agreement
between predicted and experimental amounts of coke deposited on the catalyst surface versus
time-on-stream for different reaction temperatures.

TABLE 5
Values of Objective Function for Different Coke Formation Models
Model
A1 .n D 1; m D 1/
A2 .n D 0; m D 2/
A3 .n D 1; m D 2/

Optimal Objective Function Value


2.58
2.43
2.83

COKE FORMATION AND CATALYST DEACTIVATION

2459

TABLE 6
Optimized Parameter Values for Coke Formation and
Deactivation Models
Parameter

Value
Coke Formation Model

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

Cmax
k01
Ea1
k02
Ea2

mg coke
mg cat.

0.001


mg cat:
mg coke hr


kJ
mol


mg coke
mg cat: hr


kJ
mol

0.001
29.12
0.0027
265.10
Deactivation Model

p
Ea
0

0.784


kJ
mol

220.012
0.083

The kinetic order of m D 2 indicates that the monolayer coke formation step involves two active
sites. For multilayer coke, however, the kinetic order is zero. The activation energy for monolayer
coking is 29.1 kJ/mol, which is significantly lower than activation energy of about 265.1 kJ/mol
for multilayer coking indicating that Pt can promote coke formation on the catalyst surface. Lobera

FIGURE 5 Predicted and experimental amount of coke (% weight of catalyst) versus time-on-stream at different
reaction temperatures. (color figure available online)

2460

S. NIKNADDAF ET AL.

et al. (2008) and van Sint et al. (2001) reported similar results for propane dehydrogenation over
Pt-Sn-K/Al2O3 and Pt/Al2O3 , respectively.

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

3.6.2. Time-dependent Deactivation Model


A reduction in catalyst activity with time-on-stream is expected as coke formation leads to
the coverage of active sites on the surface of catalyst by carbonaceous species. Experimental
data indicate a gradual decrease in catalyst activity up to and beyond the initial stage of coke
deposition. A time-dependent deactivation model (Romero et al., 1996) was used to describe
the loss in catalyst activity with time-on-stream. The model contains an exponential relationship
between catalyst activity, a, and catalyst coke content, CC , as well as a time-dependent relationship
for catalyst coke content as follows:
a D exp. BC C /

(6)

CC D At p

(7)

where A, B, and p are adjustable model parameters. Hence, the activitytime relationship can be
expressed by:
a D exp. t p /

(8)

where  is given by:


 D 0 exp



Ea
R



1
T

1
T0



(9)

The model parameters, namely, 0 , Ea, and p, were obtained by minimizing the following
objective function using the fmin search function in Matlab optimization toolbox:

N
X
Xexp Xmodel

O:F D

X
j D1

(10)

exp

where Xexp and Xmodel are the experimental and predicted conversion of propane, respectively.
Propane conversions at the reactor exit were obtained by solving the integrated form of the plug
flow reactor model (Eq. [11]) using the reaction rate expression proposed by Lobera et al. (2008)
as the reaction rate for the fresh catalyst and a time- and temperature-dependent activity defined
by Equation (12):
Z

dXA
W
D
rA
FA

(11)

rA D a.time  Temperature/  rA0

(12)

The optimized parameters for the time-dependent deactivation model are represented in Table 6.
Figure 6 illustrates the experimental and predicted propane conversions from the deactivation
model for different reaction temperatures as a function of time-on-stream. Figure 7 shows a very
good agreement between predicted versus experimental propane conversions for the entire range
of reaction conditions.

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

COKE FORMATION AND CATALYST DEACTIVATION

2461

FIGURE 6 Predicted and experimental fractional conversion of propane versus time-on-stream at different
reaction temperatures. (color figure available online)

FIGURE 7 Predicted versus experimental fractional conversion of propane with deactivation model. (color figure
available online)

4. CONCLUSIONS
Extensive experiments on propane dehydrogenation reaction were performed to investigate the
catalyst deactivation by coking. A monolayer-multilayer coke deposition model was proposed to
describe the kinetics of coke formation on a commercial Pt-Sn/ -Al2O3 catalyst. The activation
energies for monolayer and multilayer coke formation were found to be 29.1 and 265.1 kJ/mol,
respectively, indicating that the presence of metals could promote coke formation on the catalyst
surface during propane dehydrogenation. A time-dependent deactivation model was also developed
to describe the loss of catalyst activity as function of time-on-stream and reaction temperatures.
Predicted values of coke deposition and catalyst activity from the two models were in good
agreement with experimental data.

2462

S. NIKNADDAF ET AL.

ACKNOWLEDGMENTS
The authors acknowledge financial support from the Iranian National Petrochemical Research and
Technology Company for the experimental work involved in this study.

Downloaded by [Istanbul Universitesi Kutuphane ve Dok] at 09:27 20 December 2014

REFERENCES
Assabumrungrat, S., Jhoraleecharnchai, W., Praserthdam, P., and Goto, S. (2000). Kinetics for dehydrogenation of propane
on Pt-Sn-K/gamma-Al2 O3 catalyst. J. Chem. Eng. 33:529.
Baris, O. A., Holmen, A., and 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., and Pujado, P. R. (2001). Butane dehydrogenation over Pt/alumina:
Activation, deactivation and the generation of selectivity. Appl. Catal. A: Gen. 221:397.
Corma, A., Melo, F. V., Sauvanaud, L., and Ortega, F. (2005). Light cracked naphtha processing: Controlling chemistry
for maximum propylene production. Catal. Today 699:107108.
Gascn, J., Tllez, C., Herguido, J., and Menndez, M. (2003). Propane dehydrogenation over a Cr2 O3 /Al2 O3 catalyst:
Transient kinetic modeling of propene and coke formation. Appl. Catal. A: Gen. 248:105.
Herguido, J., Menndez, M., and Santamara, J. (2005). On the use of fluidized bed catalytic reactors where reduction and
oxidation zones are present simultaneously. Catal. Today 100:181.
Larsson, M., Hultn, M., Blekkan, E. A., and Andersson, B. (1996). The effect of reaction conditions and time on stream
on the coke formed during propane dehydrogenation. J. Catal. 164:4453.
Lobera, M. P., Tllez, C., Herguido, J., and Menndez, M. (2008). Transient kinetic modelling of propane dehydrogenation
over a PtSnK/Al2 O3 catalyst. Appl. Catal. A: Gen. 349:156164.
Monzn, A., Romeo, E., Royo, C., Trujillano, R., Labajos, F. M., and Rives, V. (1999). Use of hydrotalcites as catalytic
precursors of multimetallic mixed oxides. Application in the hydrogenation of acetylene. Appl. Catal. A: Gen. 185:53.
Nam, I. S., and Kittrell, J. R. (1984). Use of catalyst coke content in deactivation modeling. Ind. Eng. Chem. Process.
Des. Dev. 23:237.
Pea, J. A., Monzn, A., and Santamara, J. (1993a). Coking kinetics of fresh and thermally aged commercial Cr2 O3 /Al2 O3
catalyst. Appl. Catal. A: Gen. 101:185.
Pea, J. A., Monzn, A., and Santamara, J. (1993b). Deactivation by coke of a Cr2 O3 /Al2 O3 catalyst during butene
dehydrogenation. J. Catal. 142:59.
Resasco, D. E., and Haller, G. L. (1994). Catalytic dehydrogenation of lower alkanes. J. Catal. 11:379411.
Rodrguez, J. C., Guimon, C., Borona, A., and Monzn, A. (1997). Effect of Zn content on catalytic activity and
physicochemical properties of Ni-based catalysts for selective hydrogenation of acetylene. J. Catal. 171:268.
Romero, E., Rodrguez, J. C., Pea, J. A., and Monzn, A. (1996). Coking kinetics of a Cr2 O3 /Al2 O3 catalyst during
1-butene dehydrogenation: Effect of H2 partial pressure. J. Chem. Eng. Data 74:1034.
van Sint, M., Kuipers, J. A. M., and van Swaij, W. P. M. (2001). A kinetic rate expression for the time-dependent coke
formation rate during propane dehydrogenation over a platinum alumina monolithic catalyst. Catal. Today 66:427.
Zhang, Y., Zhou, Y., Qiu, A., Wang, Y., Xu, B. Y., and Wu, P. (2006). Propane dehydrogenation on Pt-Sn/ZSM-5 catalyst:
Effect of tin as a promoter. Catal. Commun. 7:860866.

Vous aimerez peut-être aussi