Sunteți pe pagina 1din 12

Ind. Eng. Chem. Res.

2008, 47, 91839194

9183

Ethylbenzene Dehydrogenation into Styrene: Kinetic Modeling and Reactor


Simulation
Won Jae Lee and Gilbert F. Froment*
Artie McFerrin Department of Chemical Engineering, Texas A&M UniVersity, College Station, Texas 77843-3122

A set of intrinsic rate equations based on the HougensWatson formalism was derived for the dehydrogenation
of ethylbenzene into styrene on a commercial potassium-promoted iron catalyst. The model discrimination
and parameter estimation was based on an extensive set of experiments that were conducted in a tubular
reactor. Experimental data were obtained for a range of temperatures, space times, and feed molar ratios of
steam to ethylbenzene, styrene to ethylbenzene, and hydrogen to ethylbenzene. All the estimated parameters
satisfied the statistical tests and physicochemical criteria, and the kinetic model yielded an excellent fit of the
experimental data. The model was applied in the simulation of the dehydrogenation in industrial multibed
adiabatic reactors with either axial or radial flow and accounting also for thermal radical-type reactions, internal
diffusion limitations, coke formation, and gasification.
1. Introduction
Styrene (ST) is one of the most important monomers in the
chemical industry. More than 2.5 107 MT/year of styrene
monomer is produced worldwide.1 The dehydrogenation of
ethylbenzene (EB) on iron oxide catalysts promoted by potassium accounts for 85% of the commercial production.2 Benzene
(BZ), toluene (TO), methane, and ethylene are the main
byproducts.3 The dehydrogenation of EB is an endothermic and
reversible reaction with an increase in the number of moles, so
that high conversion requires high temperatures and a low EB
partial pressure.
Potassium promotes the catalyst activity and enhances the
selectivity to ST. The role of potassium is attributed to the
existence of an active phase, potassium ferrite (KFeO2).410
Potassium also promotes the gasification of the carbonaceous
deposits on the catalyst and would maintain the catalyst activity,
even at relatively low steam-to-hydrocarbon ratios (i.e., below
a molar ratio of 11.8:1).5,11,12
High steam-to-hydrocarbon ratios favor the selectivity to ST
but also the lifetime and stability of the catalyst by decreasing
the undesired carbon production. Devoldere and Froment13
studied the influence of the formation of coke on the catalyst
and of its gasification by steam on the operation of ST plants
and introduced the notion of a dynamic equilibrium coke
content.
The kinetics of EB dehydrogenation have been widely
investigated1418 but seldom in a fundamental way, generating
empirical polynomial correlations for the optimization of the
commercial unit.1921 Furthermore, the reaction rates reported
in most of the papers are not intrinsic, but rather are effective
(i.e., including the effects of diffusional limitations).14,16,22
Recently, however, Schule et al.23 derived a mechanistic model
using a single-crystal unpromoted iron oxide film. It includes
EB dehydrogenation into ST, but it does not consider BZ and
TO formation and thermal reactions. The redox reactions on
the catalyst and coke formation were included in the model that
contains 31 parameters, 13 of which were estimated from overall
kinetic measurements, whereas the others (such as adsorption
* To whom correspondence should be addressed. Tel.: +1-979-8453361. Fax: +1-979-845-6446. E-mail: g.froment@che.tamu.edu.

Present address: Corporate R&D, LG Chem, Ltd./Research Park,


Moonji-Dong, Yuseong, Daejeon, 305-380, Korea.

and desorption parameters) were determined by specific


experimentation.
To accurately predict the reactor performance, the development of an intrinsic HougensWatson type kinetic model, i.e.,
accounting for the adsorption and desorption of the reacting
species, is indeed required. This was attempted in the present
work starting from an extensive set of kinetic experiments. The
intrinsic kinetic model is inserted into a heterogeneous reactor
model, accounting for diffusion limitation inside the catalyst
particles and also for the thermal reactions, to simulate industrial
axial and radial flow multibed adiabatic reactors. The kinetic
model for the main reactions is also combined with that for the
coke formation and gasification to calculate the dynamic
equilibrium coke content of the catalyst for different steam-tohydrocarbon feed ratios and to investigate the influence of the
coke formation on the reactor performance.
2. Experimental Unit
Two liquid feeds, i.e., hydrocarbon mixture (EB and ST) and
water, were separately pumped and controlled by two Harvard
precise syringe pumps. N2 was used as a diluent for the reaction
and as an internal standard for the gas chromatography (GC)
analysis. The flow rate of N2 was controlled by an Omega mass
flow controller. Great care was taken to have the liquids and
gases well-mixed through two preheaters in series before they
were fed to the reactor. The reactor was fabricated from a
stainless steel tube and had an internal diameter of 1 in. and a
length of 18 in. The inner surface of the reactor was plated with
chromium, to suppress coke formation on the walls. The reactor
was heated by an electric furnace surrounding the reactor tube.
Three Omega type-K thermocouples were located on the surface
of the internal wall of the furnace. The temperature inside the
reactor was monitored by a Omega type-K thermocouple that
slid inside a thermowell. No temperature gradient was observed
over the catalyst bed.
The commercial iron catalyst was crushed to a particle size
of 0.25-0.42 mm, to avoid internal diffusion limitations, and
was mixed with the same particle size of R-Al2O3 (Saint-Gobain
NorPro, D-99) in a weight ratio of 1:6. Because the ratio of
tube diameter to catalyst pellet diameter is .10, the flow pattern
inside the reactor can be considered to be of the plug-flow type.
The catalyst-inerts mixture was placed in the middle section

10.1021/ie071098u CCC: $40.75 2008 American Chemical Society


Published on Web 02/06/2008

9184 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008

of the reactor, and its bed depth varied between 3 and 5 cm,
depending on the amount of catalyst. The upper and lower
sections of the reactor were filled with R-Al2O3 beads to serve
two functions: preheating and mixing of reactants and decreasing
the free volume in the reactor.
A small fraction of the exit gas was sent to a Shimadzu 17A
gas chromatograph with a thermal conductivity detector (TCD),
followed by a HewlettsPackard (HP) 5890 gas chromatograph
with a flame ionization detector (FID) for online analysis. Two
gas chromatographs were connected in series, and helium was
used as a carrier gas. The Shimadzu 17A gas chromatograph
was equipped with two valves to inject the gaseous products
and switched the valves by means of a timing program stored
in the gas chromatograph. The oven temperature programs of
the Shimadzu 17A and HP 5890 gas chromatographs and valve
switching timing program were matched to accomplish the
desired separation of all compounds. Three capillary columns
were used: HP PLOT Molecular Sieve 5A column (Agilent,
0.53 mm ID 25 m 15 m), for the separation of H2 and
N2; GS-Q capillary column (J&W, 0.53 mm ID 30 m), for
the separation of N2, CO, CO2, CH4, C2H4, and H2O; and a
HP-5 capillary column (Agilent, 0.53 mm ID 1.5 m
30 m), for the separation of aromatic compounds.
N2 was used as a primary internal standard for the TCD
analysis. EB was chosen as a secondary internal standard,
because it appeared on both TCD and FID as one of the major
compounds and could be used to tie the analyses on the two
detectors.
EB conversion, conversions of EB to product j, and selectivities to product j were calculated using the following definitions:
EB conversion (%) ) 100

Figure 1. Effect of temperature on the ethylbenzene (EB) conversion to


styrene (ST) for T ) 600, 620, and 640 C; PT ) 1.04 bar; PN2 ) 0.432
bar; H2O-to-EB molar ratio ) 11 mol/mol; ST-to-EB molar ratio ) 0; H2to-EB molar ratio ) 0.

F0EB - FEB
F0EB

Conversion of EB to product j (%) ) 100

Selectivity to product j (%) ) 100

Fj - F0j
F0EB

Fj - F0j
F0EB - FEB

Prior to conducting the experiments, the fresh iron catalysts were


activated. The temperature was increased to 620 C under a N2
flow for 12 h. A typical partial pressure of N2 was 0.432 bar.
Water was injected to the preheater 1 or 2 min before the EB
was added. A typical H2O-to-EB feed ratio was 11 mol/mol.
Several days were required for the catalyst to be fully activated.
During the night, the feed of EB and water was always shut
off, while the temperature was maintained at 620 C under N2
flow.
Kinetic data were collected at various temperatures, space
times, and feed molar ratios of H2O and EB, ST and EB, and
H2 and EB. Space time is defined as the weight of catalyst
divided by the feed molar flow rate of EB. Space time was in
the range of 6-70 g-cat h/(mol EB). Experiments were
performed at three different temperatures: 600, 620, and 640
C. The kinetic experiments were always conducted at relatively
low conversions, far away from equilibrium. A reference
reaction condition was used to check whether the catalyst was
not deactivated by coke accumulation, potassium loss, or
reduction before conducting kinetic experiments. If any loss of
activity was detected, the catalyst was replaced. The standard
activity was easily reproduced. Interfacial gradients of temperature and partial pressure were negligible in all experiments
reported here.

Figure 2. ST selectivity as a function of EB conversion for T ) 600, 620,


and 640 C, PT ) 1.04 bar; PN2 ) 0.432 bar; H2O-to-EB molar ratio ) 11
mol/mol; ST-to-EB molar ratio ) 0; H2-to-EB molar ratio ) 0.

3. Experimental Results
Experimental data were collected by injecting the exit sample
6-10 times into the online GC setup under the same reaction
conditions. The data plotted in the following figures are averages
of those values. The standard deviation of each point is 1%
of the average value.
For all the temperatures, the EB conversion did not increase
appreciably when the space times exceeded 70 g-cat h/(mol EB),
because the reactions approach equilibrium at high space time.
The solid lines in the following figures are drawn to fit the data.
Figure 1 shows the effect of temperature on the EB conversion
to ST for a molar steam-to-EB ratio of 11. The rate of formation
of ST from EB decreased as the space time increased. The
calculated equilibrium conversions of EB to ST are 80.4%,
85.0%, and 88.8% at T ) 600, 620, and 640 C, respectively.
0
The corresponding experimental values at W/FEB
) 62 g-cat
h/mol shown in Figure 1 were 60.0%, 71.6%, and 79.1%,
respectively, i.e., well below the equilibrium values. Figure 2
shows the ST selectivity as a function of the EB conversion for
the complete temperature range. The ST selectivity evolves in
an opposite way to the EB conversion. The ST selectivity has
a tendency to decrease as the temperature increases, because
the competitive reactions that produce byproducts become
pronounced with increasing temperature. Figures 3 and 4 show
the BZ and TO selectivity, as a function of EB conversion. The
rate of BZ formation is only slightly affected by the EB
conversion (or space time), but the rate of TO formation is
significantly enhanced as the EB conversion (or space time)
increases.

Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9185

Figure 3. Benzene (BZ) selectivity as a function of EB conversion for T )


600, 620, and 640 C, PT ) 1.04 bar; PN2 ) 0.432 bar; H2O-to-EB molar
ratio ) 11 mol/mol; ST-to-EB molar ratio ) 0; H2-to-EB molar ratio ) 0.

Figure 6. Effect of ST-to-EB feed ratio on (a) the EB conversion and (b)
the ST selectivity for T ) 620 C; PT ) 1.04 bar; H2O-to-EB molar ratio
) 11; H2-to-EB molar ratio ) 0.

Figure 4. Toluene (TO) selectivity as a function of EB conversion for T )


600, 620, and 640 C, PT ) 1.04 bar; PN2 ) 0.432 bar; H2O-to-EB molar
ratio ) 11 mol/mol; ST-to-EB molar ratio ) 0; H2-to-EB molar ratio ) 0.

results at 600 and 640 C are not included. The increase in the
H2O-to-EB feed ratio did not result in an increase of the EB
0
conversion or the ST selectivity for W/FEB
< 30 g-cat h/mol.
0
Even for W/FEB > 30 g-cat h/mol, the effect of increasing the
H2O-to-EB feed ratio on the EB conversion was insignificant.
The effect of the H2O-to-EB feed ratio on the styrene selectivity,
however, becomes pronounced as the EB conversion increases.
Effect of the Styrene-to-Ethylbenzene Feed Ratio. Figure
6 shows the effect of the ST-to-EB feed ratio on the EB
conversion and the ST selectivity at 620 C. As the ST-to-EB
ratio in the feed increases, the EB conversion decreases, because
of the competitive adsorption of ST and the approach to
equilibrium. Furthermore, the adsorbed ST on the surface
changes to a carbonaceous deposit, which causes catalyst
deactivation. As the ST-to-EB feed ratio increased, the ST
selectivity decreased.
Effect of Hydrogen-to-Ethylbenzene Feed Ratio. Figure 7
shows the effect of hydrogen addition on the EB conversion
and the ST and TO selectivity at 600 C. Hydrogen is involved
in the formation of TO from ST, so that when the H2-to-EB
feed ratio is increased, the TO selectivity is favored, while the
ST selectivity suffers from side reactions. The addition of
hydrogen further reduces the iron catalyst from hematite (Fe2O3)
to magnetite (Fe3O4), which has a lower activity.
4. Reaction Scheme and Rate Equations

Figure 5. Effect of H2O-to-EB feed ratio on (a) the EB conversion and (b)
the ST selectivity for T ) 620 C; PT ) 1.04 bar; ST-to-EB molar ratio )
0; H2-to-EB molar ratio ) 0.

Effect of the Steam-to-Ethylbenzene Feed Ratio. Figure 5


shows the influence of the H2O-to-EB feed ratio on the EB
conversion and the ST selectivity at 620 C. The experimental

4.1. Thermal Reactions. Thermal radical-type reactions


cannot be ignored in the derivation of the kinetic model for EB
dehydrogenation. They occur in the zones without catalyst or
in the voids of zones that contain only inert solids or in the
void fraction of the catalyst bed itself. These reactions involve
free-radical mechanisms; however, given the low thermal
conversions, they were approximated in the present work by a
simple molecular scheme.24,25 The kinetic parameters for these
reactions are given in Table 1. They were obtained by matching
simulations based on the detailed radical scheme shown by the
following molecular mechanism:

9186 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008

kt1

EB {\} ST + H2 rt1 ) kt1 PEB kt1 - 1

PSTPH2
Keq

Scheme 1. Catalytic Reaction Scheme of Ethylbenzene (EB)


Dehydrogenation

kt2

EB 98 BZ + C2H4 rt2 ) kt2PEB


kt3

EB + H2 98 TO + CH4 rt3 ) kt3PEB


By way of example, the conversion of EB to ST that results
from thermal reactions was simulated for a typical bench-scale
experiment at 620 C, as reported in Figure 1. Using the kinetic
parameters given in Table 1 and accounting for the temperature
profiles or levels in the preheater, catalytic, and bottom sections
of the reactor, the thermal conversion of EB to ST amounted
to 5.7%, whereas the measured EB conversion to ST shown in
Figure 1 was 62.4%.
4.2. Catalytic Reactions. Scheme 1 shows the catalytic
reaction scheme that is generally accepted for EB dehydroge-

nation. The reaction ST + H2 f BZ + C2H4 in Scheme 1 was


dropped early in the kinetic modeling procedure of the present
work. The selected rate equations are
k1KEB[PEB - (PSTPH2 Keq)]

rc1 )

(1 + KEBPEB + KH PH
2

(1)

+ KSTPST)2

k2KEBPEB
(1 + KEBPEB + KH2PH2 + KSTPST)2

rc2 )

(2)

k3KEBPEBKH2PH2

rc3 )

(1 + KEBPEB + KH PH
2

(3)

+ KSTPST)2

k4KSTPSTKH2PH2

rc4 )

(1 + KEBPEB + KH PH
2

(4)

+ KSTPST)2

These rate equations correspond to the rate-determining steps


on dual sites and involve an adsorbed hydrocarbon and a
molecularly adsorbed hydrogen, confirming the earlier conclusions of Devoldere and Froment.26 Twenty-four rival models
based in Scheme 1 and comprised of models with ratedetermining steps that could be single-site or dual-site adsorption, desorption, or reaction and with molecular or atomically
produced hydrogen in the reaction step proper were tested. The
discrimination was based on the F-test and the confidence
intervals of the parameters. A nonsignificant value of the
adsorption equilibrium constant for water was obtained, reflecting what was reported in the section on the influence of the
steam-to-EB feed ratio. Nonsignificant values were also obtained
for the parameters related to the conversion of ST to BZ, which
is the reason why this reaction does not appear in Scheme 1.
5. Parameter Values
5.1. Continuity Equations for the Reacting Species. Kinetic
analysis of the data obtained in the integral reactor with plug
flow previously described requires a set of continuity equations
for the reacting species, accounting for both catalytic and thermal
reactions in the catalyst bed and voids. For steady-state
operation, the following can be written for Scheme 1:
dXEB
W F0EB

d(

) 1rc1 + 2rc2 + 3rc3 + (rt1 + rt2 + rt3)


dXST

Figure 7. Effect of H2-to-EB feed ratio on (a) the EB conversion, (b) the
ST selectivity, and (c) the TO selectivity for T ) 600 C; PT ) 1.04 bar;
H2O-to-EB molar ratio ) 11; ST-to-EB molar ratio ) 0.

W F0EB

d(

dXBZ

Table 1. Pre-exponential Factors and Activation Energies for the


Thermal Reactions
i

Ati [kmol/(m3 h bar)]

1
2
3

2.2215 10
2.4217
1020
17
3.8224 10
16

) 1rc1 - 4rc4 + rt1

W F0EB

d(

Eti [kJ/mol]

dXH2

272.23
352.79
313.06

W F0EB

d(

) 2rc2 + rt2

B
FB

(5b)

B
FB

) 1rc1 - 3rc3 - 24rc4 + (rt1 - rt2)

with initial conditions

B
(5a)
FB

(5c)
B
FB

(5d)

Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9187
Table 2. Parameter Values and Statistical Tests Derived from the Data at 620 C
95% Confidence Interval
parameter

unit

estimate

standard deviation

t value

lower value

upper value

KEB
KST
KH2
k1
k2
k3
k4

bar-1
bar-1
bar-1
kmol/(kg-cat h)
kmol/(kg-cat h)
kmol/(kg-cat h)
kmol/(kg-cat h)

8.466
34.00
3.091
0.2725
0.00544
0.0184
0.0302

1.01
1.51
0.447
0.0171
0.000504
0.00874
0.00565

8.37
22.6
6.91
15.9
10.8
2.11
5.66

6.460
31.02
2.204
0.2385
0.00444
0.001095
0.0190

10.47
36.99
3.977
0.3065
0.00644
0.03571
0.0413

Table 3. Values of the Kinetic Parameters Derived from the


Estimation Based on the Data at All Temperatures
symbol

value

Pre-exponential Factor of Catalytic Reaction i [kmol/(kg-cat h)]


4.594 109
1.060 1015
1.246 1026
8.024 1010

A1
A2
A3
A4

Pre-exponential Factor of Adsorption Species j [bar-1]


1.014 10-5
2.678 10-5
4.519 10-7

AEB
AST
AH2

Figure 8. Effect of temperature on (a) the rate coefficients ki and (b) the
adsorption equilibrium constants Kj. Symbols represent estimated values
per temperature; lines represent calculated values from estimates at all
temperatures.

Activation Energy [kJ/mol]


E1
E2
E3
E4

175.38
296.29
474.76
213.78
Adsorption Enthalpy [kJ/mol]
-102.22
-104.56
-117.95

Ha,EB
Ha,ST
Ha,H2

Xj ) 0 at

W
)0
F0EB

Small catalyst particles were used to eliminate internal


diffusion limitations and, thus, obtain intrinsic kinetics. In that
case, the effectiveness factors (i) are equal to 1. The thermal
reactions inside the pores of the catalyst were not taken into
account.
5.2. Parameter Estimation. The parameter estimation was
based on the integral method as described by Froment and
Bischoff27 and Froment.28 The set of stiff differential equations
(eqs 5a-d) were integrated numerically using Gears method.29
The parameters were estimated through the minimization of the
multiresponse objective function, which was performed by
means of the Marquardt algorithm. The minimization of the sum
of squares of residuals can be represented by
n

S() )

[y - f(x , )]
i

98 Min

(6)

i)1

The parameter values, the statistical tests, and the approximate


95% confidence intervals derived from the experimental data
at 620 C are given in Table 2. The parameter values obtained
from the data at all temperatures simultaneously are shown in
Table 3. The agreement between the two sets of parameters can
be observed in Figure 8, which shows the temperature dependence of the adsorption equilibrium constants and rate coefficients. The symbols represent the values of the kinetic
parameters estimated from various sets of data collected at a
given temperature, whereas the lines correspond to the values
estimated from the complete set of data at all temperatures. The

Figure 9. Comparison of calculated and experimental conversions, as a


function of space time. Symbols represent experimental data; lines represent
calculated values using the estimates of the kinetic parameters obtained
from all temperatures simultaneously. T ) 620 C; H2O-to-EB molar ratio
) 11 mol/mol; PT ) 1.044 bar; PN2 ) 0.432 bar.

kinetic parameters lead to an excellent fit of the experimental


data. By way of example, Figure 9 compares the experimental
and calculated conversions, as a function of space time at 620
C.
5.3. Physicochemical Tests on the Model Parameters.
Boudart and coauthors3032 proposed several rules for validating
kinetic parameters. The following test procedure was guided
by the work of Mears and Boudart,33 Van Trimpont et al.,34
Xu and Froment,35 and Froment and Bischoff.27
(1) For an endothermic reaction, the thermodynamics require
the activation energy of reaction (Ei) to exceed the heat of
reaction (Hr,i):
Ei > Hr,i

(7)

The activation energies for reactions of EB to ST and EB to


BZ, which are endothermic reactions, are given in Table 3. They
are indeed larger than the corresponding heats of reaction at
893.15 K, which amount to 124.8 and 101.5 kJ/mol, respectively, as calculated from the thermodynamics.

9188 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008
Table 4. Adsorption Entropies and Standard Entropies for
Ethylbenzene (EB), Styrene (ST), and Hydrogen (H2)
component

0
-Sa,j
[J/(mol K)]

0
-Sg,j
[J/(mol K)]a

51 - 0.0014Ha,j
[J/mol]

ethylbenzene, EB
styrene, ST
hydrogen, H2

95.61
87.53
121.5

361.65
346.25
186.1

194.1
197.4
216.1

Values are obtained from Stull et al.36

(2) The heat of adsorption (-Ha,j) must be positive, because


the adsorption is exothermic. All the estimates of the heats of
adsorption satisfy this constraint.
(3) The adsorption entropy must satisfy
0 < -Soa,j < Sog

Soa ) Soa - Sog

For cylindrical packings, the coefficients a and b are 1.28


and 458, respectively.38 The pressure drop between the catalyst
beds is neglected.
6.2. Continuity Equations for the Components Inside
the Porous Catalyst. The continuity equations for the components inside a porous catalyst that account for the thermal
reactions occurring in the void space inside the catalyst particle
are given as follows:

RgT
1 d 2 dPs,BZ
r
)(F r + srt2)
2 dr
dr
De,BZ s c2
r

(15c)

B
dT
) F0EB -Hr1 1rc1 + rt1
0
F
d(W FEB)
B
j)1
B
B
- Hr3 3rc3 + rt3
- Hr44rc4 (12)
Hr2 2rc2 + rt2
FB
FB

and the momentum equation is given as


-

dPt
d(W F0EB)

) fR

usGF0EB
FBdp

RgT
1 d 2 dPs,H2
r
)[Fs(rc1 - rc3 - 2rc4) + s(rt1 - rt3)]
2 dr
dr
D
r
e,H2
(15d)
with boundary conditions
Ps,j ) Pj at r ) R
dPs,j
) 0 at r ) 0
dr

(11)

6.1. Continuity, Energy, and Momentum Equations. A


multibed industrial adiabatic reactor with axial flow was
simulated, based on a heterogeneous reactor model (i.e.,
accounting for internal diffusion limitations). The steady-state
continuity equations for the reacting species have already been
given in eqs 5. With the catalyst particle sizes used in industrial
production units, the effectiveness factors i are different from
1. For the simplified pseudohomogeneous reactor model in
which diffusion limitations are not taken into account, i ) 1.
With the set of rate equations used here, there is no analytical
expression for the calculation of the effectiveness factors, in
terms of a modulus, so that the solution can only be obtained
by explicitly considering the transport equations inside the
catalyst particle.
The energy equation is written as

(14)

(15b)

6. Simulation of a Three-Bed Adiabatic Reactor with


Axial Flow

[ (
) (

(9)

Table 4 shows that this rule also is satisfied.

j pj

b(1 - B)
Re

RgT
1 d 2 dPs,ST
r
)[Fs(rc1 - rc4) + src1]
2 dr
dr
D
r
e,ST

Table 4 shows that the rule is satisfied.


(4) The last criterion has been applied by Everett,37 Vannice
et al.,31 and Boudart et al.30

m c

a+

(8)

(10)

41.8 < -Soa,j e 51 - 0.0014Ha,j

RgT
1 d 2 dPs,EB
r
)
[F (r + rc2 + rc3) + s(rt1 + rt2 + rt3)]
2 dr
dr
De,EB s c1
r
(15a)

where Sog is the standard entropy of the gas and Soa is the entropy
of the adsorbed molecule. For adsorption, Soa is smaller than Sgo,
because of the translational contribution to Sgo.33 The gas-phase
o
standard entropies of EB, ST, and H2 in Sg,j
can be obtained
36
o
from Stull et al. Sa,j was calculated from the relation
Soa,j ) R ln Aj

B

The inequality comes from the relation

1 - B

f)

(13)

The friction factor (f) is calculated using the Ergun relation:

where Ps,j is the partial pressure of component j inside the


catalyst.
The effective diffusivities are calculated from the weighted
binary molecular diffusivities, the void fraction of the catalyst
particle (0.4), and the tortuosity factor (3.0) along the lines
explained in Froment and Bischoff.27
The numerical integration of this set of equations yields the
profiles of the reacting species inside the catalyst particle at a
given position in the reactor and provides insight into the
importance of diffusion limitations on the various reactions.
These limitations can also be expressed in terms of a single
number: the effectiveness factor (i). Accounting for the rates
of both the catalytic reactions and the thermal reactions in the
void space inside the porous catalyst, the effectiveness factors
i can be calculated from these profiles by means of

i )

[rci(Ps,j)Fs + rti(Ps,j)s] dV
[rci(Pj)Fs + rti(Pj)s]V

(16)

6.3. Numerical Procedures. The continuity, energy, and


momentum equationsseqs 5, 12, 13, and 14swere solved
numerically, using Gears method. For each integration step
along the reactor length, the set of equations described by eqs
15 was solved by means of the orthogonal collocation method
with six internal collocation points, whose coefficients were
obtained numerically from the Jacobian orthogonal polynomials.
Calculations using nine internal collocation points led to exactly
the same results.
6.4. Results and Discussion. The feed conditions and reactor
dimension for the simulation of a three-bed adiabatic reactor
with axial flow are given in Table 5. The simulation results are
given in Table 5 and shown in Figures 10 and 11.

Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9189
Table 5. Feed Conditions, Reactor Dimensions, and Simulation
Results of a Three-Bed Adiabatic Axial Flow Reactor Using the
Heterogeneous Model
value
parameter

bed 1

bed 2

bed 3

weight of catalyst [kg]


space time [kg-cat h/(kmol EB)]
length of each bed [m]
XEB [%]
SST [%]
SBZ [%]
STO [%]
Pin [bar]
Tin [K]
Tout [K]

72950
103.18
1.33
36.89
98.49
1.000
0.507
1.25
886
811.36

82020
219.19
1.50
65.78
95.10
1.423
3.480
1.06
898.2
845.71

78330
329.98
1.43
83.76
90.43
1.754
7.809
0.783
897.6
873.6

parameter
inner radius of reactor [m]
feed molar flow rate [kmol/h]
EB
ST
BZ
TO
H2O
total feed molar flow rate [kmol/h]

value
3.50
707
7.104
0.293
4.968
7777
8496.37

Figure 11. Evolution of effectiveness factors in a three-bed adiabatic axial


flow reactor for Tin ) 886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB
0
molar ratio ) 11 mol/mol; FEB
) 707 kmol/h.

The comparison of simulated profiles of EB conversion and


ST selectivity between the pseudohomogeneous model and the
heterogeneous model is plotted against the space time in Figure
10. The EB conversion at the exit of the reactor simulated by
means of the heterogeneous model was 83.76%, versus 86.82%
for the pseudohomogeneous model, ignoring diffusion limitations. The ST selectivity for the heterogeneous model was
90.43%, versus 91.43% for the pseudohomogeneous model. The

Figure 12. Effect of total pressure on (a) the EB conversion and (b) the ST
selectivity in a three-bed adiabatic axial flow reactor using the heterogeneous
model for isobaric conditions and for Tin ) 886, 898, and 897 K; H2O-to0
EB molar ratio ) 11 mol/mol; FEB
) 707 kmol/h.

Figure 10. Comparison between profiles predicted by the heterogeneous


and the pseudohomogeneous model of (a) the EB conversion and (b) the
ST selectivity profiles in a three-bed adiabatic axial flow reactor for Tin )
886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB molar ratio ) 11 mol/
0
mol; FEB
) 707 kmol/h. Solid line represents the heterogeneous model,
dashed line represents the pseudohomogeneous model.

diffusion limitations, expressed in terms of the effectiveness


factors (1, 2, and 3) are shown in Figure 11. At the entrance
of the beds, the temperature is high and the intrinsic reaction
rate is fast. Accordingly, the effectiveness factors for reactions
1 and 2 (the formation of ST from EB and formation of BZ
from EB, respectively) are low, meaning that the process is
diffusion-controlled. These effectiveness factors increase along
the bed as the intrinsic reaction rates decrease. In contrast, the
effectiveness factor for reaction 4 (the formation of TO from
ST) is very high at the entrance, because it is a consecutive
reaction.
Figure 12 considers an isobaric reactor and shows how
reducing the pressure from 1.25 to 0.70 bar, while maintaining

9190 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008
Table 6. Feed Conditions, Reactor Dimensions, and Simulation
Result of a Three-Bed Adiabatic Radial Flow Reactor Using the
Heterogeneous Model
value
parameter

bed 1

bed 2

bed 3

weight of catalyst [kg]


space time [kg-cat h/(kmol EB)]
catalyst bed depth [m]
XEB [%]
SST [%]
SBZ [%]
STO [%]
Pin [bar]
Tin [K]
Tout [K]

72 950
103.18
0.614
36.59
98.43
1.01
0.56
1.25
886
812.04

82 020
219.19
0.708
64.18
93.92
1.53
4.54
1.22
898.2
850.26

78 330
329.98
0.681
81.19
83.24
2.12
14.60
1.21
897.6
890.37

parameter

Figure 13. Simplified radial flow reactor configuration.

a constant steam-to-EB ratio, increases the ST selectivity at the


exit from 82.18% to 90.13%. This is a consequence of the shift
of the equilibrium conversion to a higher value.
7. Simulation of a Reactor with Radial Flow and Three
Adiabatic Beds
The pressure drop in a radial flow reactor is much smaller
than that in an axial flow reactor, because of the higher crosssectional area of the catalyst bed. It permits one to use smaller
particles, which leads to higher effectiveness factors. The
differences in the performance of these two types of reactor
are discussed below.
7.1. Continuity, Energy, and Momentum Equations. Figure 13 schematically represents a radial flow reactor configuration. Gas flows in a centrifugal direction across the catalyst
bed contained in a cylindrical basket. The cross-sectional area
of the catalyst bed varies with the radial coordinate r. The
continuity equation for the components can be expressed in
0
terms of space time, W/FEB
, with W ) zFB(r2 - r02):
dXj
d(W F0EB)

) iRj

inner radius of catalyst bed [m]


length of each reactor [m]
feed molar flow rate [kmol/h]
EB
ST
BZ
TO
H2O
total feed molar flow rate [kmol/h]

value
1.5
7
707
7.104
0.293
4.968
7777
8496.37

simulations of a radial flow reactor and an axial flow reactor


with each of three adiabatic beds, using the heterogeneous
model. The operating conditions were identical. In the reactor
with radial flow, the EB conversion amounted to 81.19%,
compared to 83.76% in the axial flow reactor. The pressure drop
in the radial flow reactor, with its large cross-sectional area,
was 0.04 bar, whereas the pressure drop amounted to 0.95 bar
in the axial flow reactor, as shown in Figure 15. The lower EB

(17)

where Rj is the total rate of reaction of component j.


The steady-state energy equation can be written in terms of
0
W/FEB
,
6

m c

j pj

j)1

dT
) F0EB (-Hri)iri
0
d(W FEB)
i)1

(18)

and the momentum equation is given as


F0EBFgus2
) fR
2zrFBdp
d(W F0EB)
dPt

(19)

The continuity, energy, and momentum equations (eqs 17,


18, and 19, respectively) must be integrated simultaneously. For
the radial flow reactor, the cross section of the catalyst bed is
dependent on the space time, i.e., radial position, so that the
superficial velocity (us) must be adapted in each integration step
through the reactor.
7.2. Results and Discussion. The feed conditions and reactor
dimensions are shown in Table 6. The length of each reactor
and the inner radius of the catalyst bed were assumed to be 7
and 1.5 m, respectively. Table 6 and Figures 14 and 15 compare

Figure 14. Comparison of simulated (a) EB conversion profiles and (b) ST


selectivity profiles using the heterogeneous model between a three-bed
adiabatic radial flow reactor and a three-bed adiabatic axial flow reactor
for Tin ) 886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB molar ratio )
0
11 mol/mol; FEB
) 707 kmol/h.

Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9191

Figure 15. Comparison of simulated (a) temperature profiles and (b) pressure
drop profiles using the heterogeneous model between a three-bed adiabatic
radial flow reactor and a three-bed adiabatic axial flow reactor for Tin )
886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB molar ratio ) 11 mol/
0
mol; FEB
) 707 kmol/h.

conversion in the radial flow reactor is caused by the lower


pressure drop, meaning that the conversion essentially occurs
close to the feed pressure. In the axial flow reactor, a substantial
fraction of the conversion occurs at lower pressures. The ST
selectivity, which is strongly dependent on the pressure, as
previously evidenced by Figure 12, was 83.24% for radial flow,
versus 90.43% for axial flow.
The difference in the TO selectivity (14.60% vs 7.89%) was
substantial, but the difference in the BZ selectivity (2.12% vs
1.75%) was insignificant.
These results might lead to the conclusion that the radial flow
reactor is less favorable than the axial flow version; however,
Figure 16 reveals the advantage of the radial flow reactor, which
permits operation at much lower feed pressures. At Pin ) 0.70
bar, which is a low value that cannot be used in the axial flow
reactor, the ST selectivity amounted to 91.32%, compared to
83.24% at 1.25 bar for essentially the same EB conversion. This
result is quite similar to that derived from the simulation of the
axial flow reactor for the isobaric condition shown in
Figure 12.
8. Simulation of a Radial Flow Reactor with Three
Adiabatic Beds Accounting for Coke Formation and
Gasification
Coke formed on the potassium-promoted iron oxide catalysts
is at least partially removed by gasification with steam.2 Previous
kinetic investigations have ignored coke formation and coke
gasification; however, more recently, Devoldere and Froment13
developed detailed kinetic models for these reactions. The model

Figure 16. Effect of feed pressure on simulated (a) EB conversion and (b)
ST selectivity in a three-bed adiabatic radial flow reactor Tin ) 886, 898,
0
and 897 K; H2O-to-EB molar ratio ) 11 mol/mol; FEB
) 707 kmol/h.

Figure 17. Effect of H2O-to-EB molar feed ratio on dynamic equilibrium


coke content profiles in a three-bed adiabatic radial flow reactor for Tin )
0
886, 898, and 897 K; Pin ) 1.25 bar; FEB
) 707 kmol/h.

for the coke formation was based on a two-step mechanism:


coke precursor formation and coke growth.39 Coke gasification
was assumed to occur at the edges of the carbon, which were
oxidized by water.40 The effect of the operating conditions
(particularly the steam-to-EB feed ratio) on the dynamic
equilibrium coke content along the three-bed adiabatic reactor
and the effect of coke formation on the reactor performance
are discussed below.
8.1. Model Equations. 8.1.1. Rate Equation for the
Formation of Coke Precursor. The formation of an irreversibly
adsorbed coke precursor from EB and ST, both adsorbed up to
equilibrium with the gas phase, is assumed to be the ratedetermining step.27

9192 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008

rs )

dCCP
) rs0Cp )
dt
kEB,pKEBPEB + kST,pKSTPST
(1 - RpCCP)ns(20)

(1 + KEBPEB + KSTPST)ns

The values of , ns, Rs, kEB,p, and kST,p were estimated by


Devoldere and Froment.13,41These values were obtained on a
catalyst that was not the same as that used in the work reported
here. Nevertheless, their insertion in the present model leads to
useful insight and reliable trends.
8.1.2. Rate Equation for Coke Growth. Further dehydrogenation of the coke precursor forms the sites on which the
coke accumulates. The intrinsic rate of coke growth can be
expressed as the product of three factors: the intrinsic rate of
coke growth per active site, the total number of active sites on
the growing coke, and a deactivation function:
0
Ctgrgr
rgr ) rgr

dCgr
)
dt
kEB,grPEBnEB + kST,grPSTnST

(1 + KH OPH O PH
2

+ KH2PH2

{(

W F0EB

d(

) 1rc1 + 2rc2 + 3rc3 + (rt1 + rt2 + rt3)

dXST

) 1rc1 - 4rc4 + rt1

W F0EB

d(

dXBZ
W F0EB

d(
dXH2
d(

n3

PH2

1 - RgrCgr)ngr(22)

n2 (

k2PH2O

1 + K3PH2 [(PH2 K1) + (k2 k1)] + PH2O

CtG (23)

The parameter values in eq 23 were also obtained from the work


of Devoldere and Froment.13,41
8.1.4. Coke Formation and Gasification: Dynamic
Equilibrium Coke Content. The EB conversions in the main
reactions decrease until the coke content of the catalyst reaches
a steady state. The stabilization process is very fast, so that the
deactivation of the catalyst is limited to a very early stage of
the operation. After it is reached, the coke content, which is
called the dynamic equilibrium coke content and is dependent
only on the temperature and the compositions, remains constant.13 No deactivation effect is observed from then onward.
At dynamic equilibrium, the rate of formation of coke precursor
and the rate of coke growth are compensated by the gasification,
so that the dynamic equilibrium coke content can be obtained
from
dCCP
) rs0CP - rG ) 0
dt

(24a)

dCgr
0
) rgr
Ctgrgr - rG ) 0
dt

(24b)

These equations also express that coke formation, like any


catalytic reaction, is subject to deactivation, but gasification is
not.

B rc(ST)
FB
8

) 2rc2 + rt2

B
FB

) 1rc1 - 3rc3 - 24rc4 + (rt1 - rt2)

21
Ccn1

B rc(EB)
+
FB
8
(25a)

W F0EB

The values of nEB, nST, n1, n2, n3, ngr, Rgr, kEB,gr, and kST,gr were
estimated by Devoldere and Froment13,41 and are used in the
present work. The intrinsic rate of coke formation, accounting
for the coke precursor formation and coke growth, can be
expressed as the summation of eqs 20 and 22.
8.1.3. Rate Equation for Coke Gasification. The rate
equation for coke gasification was developed under the assumption that the rate-determining step is the irreversible decomposition of an oxidized carbon complex to CO and free carbon.13
Using the pseudo-steady-state approximation for the surface
intermediates, the rate of coke gasification is given by
rG )

dXEB

(21)

The model for the rate of coke growth is


rgr )

The kinetic model for coke formation and gasification was


coupled to the kinetic model for the main reactions in the
simulations of a three-bed adiabatic reactor with radial flow,
using the heterogeneous model. Equations 25a-d show the
continuity equations for the components, accounting for the coke
formation from both EB and ST.

(25b)

(25c)
B
+
FB

) (

rc(EB)
rc(ST)
+ 20
(25d)
8
8

where rc(EB) represents the rate of coke formation from


ethylbenzene and rc(ST) represents that from styrene.
The energy equation is written
6

{ [ ( )]
( )] [ ( )]

B
dT
) F0EB -Hr1 1rc1 + rt1
0
F
d(W FEB)
B
j)1
B
B
- Hr3 3rc3 + rt3
- Hr44rc4 Hr2 2rc2 + rt2
FB
FB
rc(ST)
rc(EB)
- HC,ST
(26)
HC,EB
8
8

mc

j pj

)}

and the momentum equation is unchanged, with respect to


eq 19.
The set of continuity, energy, and momentum equationsseqs
25, 26, and 19swas integrated simultaneously along the reactor.
8.2. Results and Discussion. Figure 17 shows the effect of
the H2O-to-EB molar ratio on the dynamic equilibrium coke
content in a three-bed adiabatic reactor with radial flow. The
dynamic equilibrium coke content was low at high H2O-to-EB
feed ratios. A high steam-to-EB ratio is not always preferred in
industrial operation, because of the cost of steam generation.
At this point, optimization is required to obtain the optimum
steam-to-EB feed ratio, also accounting for the lifetime of the
catalyst.
Figure 18 shows the effect of coke formation on the simulated
EB conversion and ST selectivity in a three-bed adiabatic reactor
with radial flow. Accounting for coke formation from EB and
ST leads to a drastic decrease in the ST selectivity but, because
of the influence of the equilibrium, a slight increase in the EB
conversion.
9. Conclusion
The extensive set of experimental data obtained on a
commercial catalyst in the experimental part of the work
reported here provides a comprehensive basis for a more
accurate evaluation of the effect of the various operating
parameters on the selectivity of styrene production from
ethylbenzene.
The detailed and rigorous kinetic model that has been derived
from the experimental database also accounted for the background thermal cracking, which strongly increases with tem-

Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9193

Figure 18. Effect of coke formation on (a) EB conversion and (b) ST


selectivity in a three-bed adiabatic radial flow reactor for Tin ) 886, 898,
and 897 K; H2O-to-EB molar ratio ) 11 mol/mol; F0EB ) 707 kmol/h. Solid
lines represent the results accounting for the coke formation from ethylbenzene and styrene; dashed lines represent results neglecting this effect.

perature. The optimal operation of todays large plants also must


consider the latter aspect: the kinetic study aimed at deriving
intrinsic rate equations. The diffusion limitations encountered
with the catalyst particle sizes used in industrial reactors are
introduced through the modeling. The catalytic and thermal
kinetic models were applied in simulations of the operation of
multibed adiabatic commercial configurations with axial or radial
flow that also included diffusion limitations and coke deposition
and gasification. Therefore, it becomes possible to investigate,
under realistic conditions, the complex influence of the various
operating variables, in particular, the operating pressure and the
steam-to-ethylbenzene ratio, which is an important factor in the
economics of the process.
Nomenclature
Ai ) pre-exponential factor of catalytic reaction i, kmol/(kg-cat h)
Aj ) pre-exponential factor for adsorption of species j, bar-1
Ati ) pre-exponential factor of thermal reaction i, kmol/(mf3 h bar)
CCP ) coke precursor content, kg coke/kg-cat
Cl ) molar concentration of vacant active sites l of catalyst, kmol/
kg-cat
Cp ) specific heat of fluid, kJ/(kg K)
De,j ) effective diffusivity of component j, mf3/(mr s)
dp ) catalyst equivalent pellet diameter, mp
Ei ) activation energy of catalytic reaction i, kJ/mol
Eti ) activation energy of thermal reaction i, kJ/mol
Fj ) molar flow rate of j, kmol/h
Fjo ) feed molar flow rate of j, kmol/h
f ) friction factor in momentum equation

G ) superficial mass flow velocity, kg/(mr2 h)


-Ha,j ) adsorption enthalpy of adsorbed component j, kJ/mol
-Hr ) heat of reaction, kJ/mol
Kj ) adsorption equilibrium constant of component j, bar-1
Keq ) equilibrium constant, bar
ki ) rate coefficient of catalytic reaction i, kmol/(kg-cat h)
kti ) rate coefficient of thermal reaction i, kmol/(mf3 h bar)
l ) vacant active site on the catalyst
m
j ) mass flow rate of component j, kg/h
Pj ) partial pressures of component j in bulk fluid, bar
Ps,j ) partial pressure of component j inside the catalyst, bar
Pt ) total pressure, bar
R ) radius of catalyst particle, mp
Rj ) total rate of reaction of the component j, kmol/(kg-cat h)
Re ) Reynolds number based on particle diameter; Re ) dpusFg/
r ) radial coodinate of reactor, mr
ro ) inner radius of catalyst bed in a radial reactor, mr
rc ) rate of coke formation, kg coke/(kg-cat h)
rci ) rate of catalytic reaction i, kmol/(kg-cat h)
rG ) rate of coke gasification, kg coke/(kg-cat h)
rgr ) rate of coke growth, kg coke/(kg-cat h)
0
rgr
) initial rate of coke growth per active center, kg coke/(kg mol
h)
rs ) rate of site coverage, kg coke/(kg-cat h)
rs0 ) initial rate of site coverage, kg coke/(kmol h)
rti ) rate of thermal reaction i, kmol/(mf3 h)
S() ) objective function
0
-Sa,j
) standard entropy of adsorption of component j, J/(mol
K)
Sg0 ) standard entropy of the gas, J/(mol K)
Sa0 ) standard entropy of the adsorbed molecule, J/(mol K)
T ) temperature, K
Tr ) average temperature, K
us ) superficial velocity, mf3/(mr2 s)
V ) catalyst pellet volume, mp3
W ) weight of catalyst, kg-cat
XEB ) conversion of ethylbenzene
Xj ) conversion of ethylbenzene into component j
Z ) length of radial flow reactor, mr
Greek Letters
R ) conversion factor in momentum equation
) model parameter
) conversion factor in the rate of coke site coverage, kmol/kgcat
B ) void fraction of bed, mf3/mr3
s ) catalyst internal void fraction, mf3/mp3
CP ) deactivation function for coke precursor
gr ) deactivation function for coke growth
) effectiveness factor
FB ) bulk density of bed, kg-cat/mr3
Fg ) gas density, kg/mf3
Fs ) catalyst pellet density, kg-cat/mp3
) cross section of reactor, mr2

Acknowledgment
The authors are grateful to Dr. R. G. Anthony, Artie McFerrin
Department of Chemical Engineering, Texas A&M, for support
and stimulating discussions.
Literature Cited
(1) Product Focus: Styrene. Chem. Week 2002, (May 15), 36.

9194 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008
(2) James, D. H.; Castor, W. M. In Ullmanns Encyclopedia of Industrial
Chemistry; Campbell, F. T., Pfefferkorn, R., Rounsaville, J. F., Eds.;
WileysVCH: Weinheim, Germany, 1994; Vol. A25, p 329.
(3) Lee, E. H. Iron-Oxide Catalysts for Dehydrogenation of Ethylbenzene
in Presence of Steam. Catal. ReV.sSci. Eng. 1973, 8, 285305.
(4) Coulter, K.; Goodman, D. W.; Moore, R. G. Kinetics of the
Dehydrogenation of Ethylbenzene to Styrene over Unpromoted and KPromoted Model Iron-Oxide Catalysts. Catal. Lett. 1995, 31, 18.
(5) Stobbe, D. E.; van Buren, F. R.; van Dillen, A. J.; Geus, J. W.
Potassium promotion of iron-oxide dehydrogenation catalysts supported on
magnesium-oxide. 1. Preparation and characterization. J. Catal. 1992, 135,
533.
(6) Muhler, M.; Schlogl, R.; Reller, A.; Ertl, G. The Nature of the Active
Phase of the Fe/K-Catalyst for Dehydrogenation of Ethylbenzene. Catal.
Lett. 1989, 2, 201210.
(7) Hirano, T. Active Phase in Potassium-Promoted Iron-Oxide Catalyst
for Dehydrogenation of Ethylbenzene. Appl. Catal. 1986, 26, 8190.
(8) Hirano, T. Roles of Potassium in Potassium-Promoted Iron-Oxide
Catalyst for Dehydrogenation of Ethylbenzene. Appl. Catal. 1986, 26, 65
79.
(9) Hirano, T. Dehydrogenation of Ethylbenzene over PotassiumPromoted Iron-Oxide Containing Cerium and Molybdenum Oxides. Appl.
Catal. 1986, 28, 119132.
(10) Muhler, M.; Schutze, J.; Wesemann, M.; Rayment, T.; Dent, A.;
Schlogl, R.; Ertl, G. The Nature of the Iron Oxide-Based Catalyst for
Dehydrogenation of Ethylbenzene to Styrene. 1. Solid-State Chemistry and
Bulk Characterization. J. Catal. 1990, 126, 339360.
(11) Addiego, W. P.; Estrada, C. A.; Goodman, D. W.; Rosynek, M. P.
An Infrared Study of the Dehydrogenation of Ethylbenzene to Styrene over
Iron-Based Catalysts. J. Catal. 1994, 146, 407414.
(12) Shekhah, O.; Ranke, W.; Schlogl, R. Styrene Synthesis: In situ
Characterization and Reactivity Studies of Unpromoted and PotassiumPromoted Iron Oxide Model Catalysts. J. Catal. 2004, 225, 56.
(13) Devoldere, K. R.; Froment, G. F. Coke Formation and Gasification
in the Catalytic Dehydrogenation of Ethylbenzene. Ind. Eng. Chem. Res.
1999, 38, 26262633.
(14) Wenner, R. R.; Dybdal, E. C. Catalytic Dehydrogenation of
Ethylbenzene. Chem. Eng. Prog. 1948, 44, 275.
(15) Carra, S.; Forni, L. Kinetics of Catalytic Dehydrogenation of
Ethylbenzene to Styrene. Ind. Eng. Chem. Process Des. DeV. 1965, 4, 281.
(16) Sheel, J. G. P.; Crowe, C. M. Simulation and Optimization of an
Existing Ethylbenzene Dehydrogenation Reactor. Can. J. Chem. Eng. 1969,
47, 183.
(17) Elnashaie, S. S. E.H.; Abdalla, B. K.; Hughes, R. Simulation of
the Industrial Fixed-Bed Catalytic Reactor for the Dehydrogenation of
Ethylbenzene to StyrenesHeterogeneous Dusty Gas-Model. Ind. Eng. Chem.
Res. 1993, 32, 25372541.
(18) Dittmeyer, R.; Hollein, V.; Quicker, P.; Emig, G.; Hausinger, G.;
Schmidt, F. Factors Controlling the Performance of Catalytic Dehydrogenation of Ethylbenzene in Palladium Composite Membrane Reactors. Chem.
Eng. Sci. 1999, 54, 14311439.
(19) Kolios, G.; Eigenberger, G. Styrene Synthesis in a Reverse-Flow
Reactor. Chem. Eng. Sci. 1999, 54, 2637.
(20) Savoretti, A. A.; Borio, D. O.; Bucala, V.; Porras, J. A. Nonadiabatic Radial-flow Reactor for Styrene Production. Chem. Eng. Sci. 1999,
54, 205213.

(21) Yee, A. K. Y.; Ray, A. K.; Rangaiah, G. P. Multiobjective


Optimization of an Industrial Styrene Reactor. Comput. Chem. Eng. 2003,
27, 111130.
(22) Sheppard, C. M.; Maier, E. E. Ethylbenzene Dehydrogenation
Reactor Model. Ind. Eng. Chem., Process Des. DeV. 1986, 25, 207210.
(23) Schule, A.; Shekhah, O.; Ranke, W.; Schlogl, R.; Kolios, G.
Microkinetic Modelling of the Dehydrogenation of Ethylbenzene to Styrene
over Unpromoted Iron Oxides. J. Catal. 2005, 231, 172180.
(24) Sundaram, K. M.; Froment, G. F. Modeling of Thermal-Cracking
Kinetics. 1. Thermal-Cracking of Ethane, Propane and Their Mixtures.
Chem. Eng. Sci. 1977, 32, 601608.
(25) Sundaram, K. M.; Sardina, H.; Fernandez-Baujin, J. M.; Hildreth,
J. M. Styrene Plant Simulation and Optimization. Hydrocarbon Process.
1991, (January), 93.
(26) Devoldere, K. R.; Froment, G. F. Unpublished results, 2000.
(27) Froment, G. F.; Bischoff, K. B. Chemical Reactor Analysis and
Design, 2nd Edition; Wiley: New York, 1990.
(28) Froment, G. F. Model Discrimination and Parameter Estimation in
Heterogeneous Catalysis. AIChE J. 1975, 21, 10411057.
(29) Gear, C. W. Numerical Initial Value Problems in Ordinary
Differential Equations; Prentice Hall: Englewood Cliff, NJ, 1971.
(30) Boudart, M.; Mears, D. E.; Vannice, M. A. Ind. Chim. Belge. 1967,
32, 281.
(31) Vannice, M. A.; Hyun, S. H.; Kalpakci, B.; Liauh, W. C. Entropies
of Adsorption in Heterogeneous Catalytic Reactions. J. Catal. 1979, 56,
358.
(32) Boudart, M. Two-Step Catalytic Reactions. AIChE J. 1972, 18, 465.
(33) Mears, D. E.; Boudart, M. The Dehydrogenation of Isopropanol
on Catalysts Prepared by Sodium Borohydride Reduction. AIChE J. 1966,
12, 313.
(34) Van Trimpont, P. A.; Marin, G. B.; Froment, G. F. Kinetics of
Methylcyclohexane Dehydrogenation on Sulfided Commercial Platinum
Alumina and Platinum-Rhenium Alumina Catalysts. Ind. Eng. Chem.
Fundam. 1986, 25, 544553.
(35) Xu, J. G.; Froment, G. F. Methane Steam Reforming, Methanation
and Water-Gas Shift. 1. Intrinsic Kinetics. AIChE J. 1989, 35, 8896.
(36) Stull, D. R.; Westrum, E. F., Jr.; Sinke, G. C. The Chemical
Thermodynamics of Organic Compounds; Wiley: New York, 1969.
(37) Everett, D. H. The Thermodynamics of Adsorptions. Part II.
Analysis and Discussion of Experimental Data. Trans. Faraday Soc. 1950,
46, 957.
(38) Handley, D.; Heggs, P. J. Momentum and Heat Transfer Mechanisms in Regular Shaped Packings. Trans. Inst. Chem. Eng. 1968, 46, T251.
(39) Beeckman, J. W.; Froment, G. F. Catalyst Deactivation by ActiveSite Coverage and Pore Blockage. Ind. Eng. Chem. Fundam. 1979, 18, 245
256.
(40) Mims, C. A.; Chludzinski, J. J.; Pabst, J. K.; Baker, R. T. K.
Potassium-Catalyzed Gasification of Graphite in Oxygen and Steam. J.
Catal. 1984, 88, 97106.
(41) Froment, G. F. Unpublished work, March 2005.

ReceiVed for reView August 8, 2007


ReVised manuscript receiVed December 12, 2007
Accepted December 14, 2007
IE071098U

S-ar putea să vă placă și