You are on page 1of 12

Ind. Eng. Chem. Res.

2008, 47, 9183–9194 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 and desorption parameters) were determined by specific


Downloaded by CHULALONGKORN UNIV at 20:34:33:086 on May 23, 2019

experimentation.
Styrene (ST) is one of the most important monomers in the
To accurately predict the reactor performance, the develop-
chemical industry. More than 2.5 × 107 MT/year of styrene
ment of an intrinsic HougensWatson type kinetic model, i.e.,
monomer is produced worldwide.1 The dehydrogenation of
accounting for the adsorption and desorption of the reacting
ethylbenzene (EB) on iron oxide catalysts promoted by potas-
from https://pubs.acs.org/doi/10.1021/ie071098u.

species, is indeed required. This was attempted in the present


sium accounts for 85% of the commercial production.2 Benzene
work starting from an extensive set of kinetic experiments. The
(BZ), toluene (TO), methane, and ethylene are the main
intrinsic kinetic model is inserted into a heterogeneous reactor
byproducts.3 The dehydrogenation of EB is an endothermic and
model, accounting for diffusion limitation inside the catalyst
reversible reaction with an increase in the number of moles, so
particles and also for the thermal reactions, to simulate industrial
that high conversion requires high temperatures and a low EB
axial and radial flow multibed adiabatic reactors. The kinetic
partial pressure.
model for the main reactions is also combined with that for the
Potassium promotes the catalyst activity and enhances the
coke formation and gasification to calculate the dynamic
selectivity to ST. The role of potassium is attributed to the
equilibrium coke content of the catalyst for different steam-to-
existence of an active phase, potassium ferrite (KFeO2).4–10
hydrocarbon feed ratios and to investigate the influence of the
Potassium also promotes the gasification of the carbonaceous
coke formation on the reactor performance.
deposits on the catalyst and would maintain the catalyst activity,
even at relatively low steam-to-hydrocarbon ratios (i.e., below
2. Experimental Unit
a molar ratio of 11.8:1).5,11,12
High steam-to-hydrocarbon ratios favor the selectivity to ST Two liquid feeds, i.e., hydrocarbon mixture (EB and ST) and
but also the lifetime and stability of the catalyst by decreasing water, were separately pumped and controlled by two Harvard
the undesired carbon production. Devoldere and Froment13 precise syringe pumps. N2 was used as a diluent for the reaction
studied the influence of the formation of coke on the catalyst and as an internal standard for the gas chromatography (GC)
and of its gasification by steam on the operation of ST plants analysis. The flow rate of N2 was controlled by an Omega mass
and introduced the notion of a “dynamic equilibrium coke flow controller. Great care was taken to have the liquids and
content”. gases well-mixed through two preheaters in series before they
The kinetics of EB dehydrogenation have been widely were fed to the reactor. The reactor was fabricated from a
investigated14–18 but seldom in a fundamental way, generating stainless steel tube and had an internal diameter of 1 in. and a
empirical polynomial correlations for the optimization of the length of 18 in. The inner surface of the reactor was plated with
commercial unit.19–21 Furthermore, the reaction rates reported chromium, to suppress coke formation on the walls. The reactor
in most of the papers are not intrinsic, but rather are effective was heated by an electric furnace surrounding the reactor tube.
(i.e., including the effects of diffusional limitations).14,16,22 Three Omega type-K thermocouples were located on the surface
Recently, however, Schüle et al.23 derived a mechanistic model of the internal wall of the furnace. The temperature inside the
using a single-crystal unpromoted iron oxide film. It includes reactor was monitored by a Omega type-K thermocouple that
EB dehydrogenation into ST, but it does not consider BZ and slid inside a thermowell. No temperature gradient was observed
TO formation and thermal reactions. The redox reactions on over the catalyst bed.
the catalyst and coke formation were included in the model that The commercial iron catalyst was crushed to a particle size
contains 31 parameters, 13 of which were estimated from overall of 0.25-0.42 mm, to avoid internal diffusion limitations, and
kinetic measurements, whereas the others (such as adsorption 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
* To whom correspondence should be addressed. Tel.: +1-979-845-
3361. Fax: +1-979-845-6446. E-mail: g.froment@che.tamu.edu. tube diameter to catalyst pellet diameter is .10, the flow pattern

Present address: Corporate R&D, LG Chem, Ltd./Research Park, inside the reactor can be considered to be of the plug-flow type.
Moonji-Dong, Yuseong, Daejeon, 305-380, Korea. 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 Figure 1. Effect of temperature on the ethylbenzene (EB) conversion to
switching timing program were matched to accomplish the styrene (ST) for T ) 600, 620, and 640 °C; PT ) 1.04 bar; PN2 ) 0.432
desired separation of all compounds. Three capillary columns bar; H2O-to-EB molar ratio ) 11 mol/mol; ST-to-EB molar ratio ) 0; H2-
were used: HP PLOT Molecular Sieve 5A column (Agilent, to-EB molar ratio ) 0.
0.53 mm ID × 25 µm × 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 µm ×
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 selectivi-
ties to product j were calculated using the following definitions:

F0EB - FEB
EB conversion (%) ) 100 × Figure 2. ST selectivity as a function of EB conversion for T ) 600, 620,
F0EB 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.
Fj - F0j
Conversion of EB to product j (%) ) 100 × 3. Experimental Results
F0EB
Experimental data were collected by injecting the exit sample
Fj - F0j 6-10 times into the online GC setup under the same reaction
Selectivity to product j (%) ) 100 ×
F0EB - FEB conditions. The data plotted in the following figures are averages
of those values. The standard deviation of each point is ∼1%
Prior to conducting the experiments, the fresh iron catalysts were of the average value.
activated. The temperature was increased to 620 °C under a N2 For all the temperatures, the EB conversion did not increase
flow for 12 h. A typical partial pressure of N2 was 0.432 bar. appreciably when the space times exceeded 70 g-cat h/(mol EB),
Water was injected to the preheater 1 or 2 min before the EB because the reactions approach equilibrium at high space time.
was added. A typical H2O-to-EB feed ratio was 11 mol/mol. The solid lines in the following figures are drawn to fit the data.
Several days were required for the catalyst to be fully activated. Figure 1 shows the effect of temperature on the EB conversion
During the night, the feed of EB and water was always shut to ST for a molar steam-to-EB ratio of 11. The rate of formation
off, while the temperature was maintained at 620 °C under N2 of ST from EB decreased as the space time increased. The
flow. calculated equilibrium conversions of EB to ST are 80.4%,
Kinetic data were collected at various temperatures, space 85.0%, and 88.8% at T ) 600, 620, and 640 °C, respectively.
times, and feed molar ratios of H2O and EB, ST and EB, and The corresponding experimental values at W/FEB 0
) 62 g-cat
H2 and EB. Space time is defined as the weight of catalyst h/mol shown in Figure 1 were 60.0%, 71.6%, and 79.1%,
divided by the feed molar flow rate of EB. Space time was in respectively, i.e., well below the equilibrium values. Figure 2
the range of 6-70 g-cat h/(mol EB). Experiments were shows the ST selectivity as a function of the EB conversion for
performed at three different temperatures: 600, 620, and 640 the complete temperature range. The ST selectivity evolves in
°C. The kinetic experiments were always conducted at relatively an opposite way to the EB conversion. The ST selectivity has
low conversions, far away from equilibrium. A reference a tendency to decrease as the temperature increases, because
reaction condition was used to check whether the catalyst was the competitive reactions that produce byproducts become
not deactivated by coke accumulation, potassium loss, or pronounced with increasing temperature. Figures 3 and 4 show
reduction before conducting kinetic experiments. If any loss of the BZ and TO selectivity, as a function of EB conversion. The
activity was detected, the catalyst was replaced. The standard rate of BZ formation is only slightly affected by the EB
activity was easily reproduced. Interfacial gradients of temper- conversion (or space time), but the rate of TO formation is
ature and partial pressure were negligible in all experiments significantly enhanced as the EB conversion (or space time)
reported here. 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.

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
conversion or the ST selectivity for W/FEB 0
< 30 g-cat h/mol.
Even for W/FEB > 30 g-cat h/mol, the effect of increasing the
0

H2O-to-EB feed ratio on the EB conversion was insignificant.


The effect of the H2O-to-EB feed ratio on the styrene selectivity,
Figure 4. Toluene (TO) selectivity as a function of EB conversion for T )
however, becomes pronounced as the EB conversion increases.
600, 620, and 640 °C, PT ) 1.04 bar; PN2 ) 0.432 bar; H2O-to-EB molar Effect of the Styrene-to-Ethylbenzene Feed Ratio. Figure
ratio ) 11 mol/mol; ST-to-EB molar ratio ) 0; H2-to-EB molar ratio ) 0. 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


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
Figure 5. Effect of H2O-to-EB feed ratio on (a) the EB conversion and (b) free-radical mechanisms; however, given the low thermal
the ST selectivity for T ) 620 °C; PT ) 1.04 bar; ST-to-EB molar ratio ) conversions, they were approximated in the present work by a
0; H2-to-EB molar ratio ) 0. simple molecular scheme.24,25 The kinetic parameters for these
Effect of the Steam-to-Ethylbenzene Feed Ratio. Figure 5 reactions are given in Table 1. They were obtained by matching
shows the influence of the H2O-to-EB feed ratio on the EB simulations based on the detailed radical scheme shown by the
conversion and the ST selectivity at 620 °C. The experimental 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 nation. The reaction ST + H2 f BZ + C2H4 in Scheme 1 was
parameters given in Table 1 and accounting for the temperature dropped early in the kinetic modeling procedure of the present
profiles or levels in the preheater, catalytic, and bottom sections work. The selected rate equations are
of the reactor, the thermal conversion of EB to ST amounted k1KEB[PEB - (PSTPH2 ⁄ Keq)]
to 5.7%, whereas the measured EB conversion to ST shown in rc1 ) (1)
Figure 1 was 62.4%. (1 + KEBPEB + KH PH 2 2
+ KSTPST)2
4.2. Catalytic Reactions. Scheme 1 shows the catalytic
reaction scheme that is generally accepted for EB dehydroge- k2KEBPEB
rc2 ) (2)
(1 + KEBPEB + KH2PH2 + KSTPST)2
k3KEBPEBKH2PH2
rc3 ) (3)
(1 + KEBPEB + KH PH 2 2
+ KSTPST)2

k4KSTPSTKH2PH2
rc4 ) (4)
(1 + KEBPEB + KH PH 2 2
+ 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 conclu-
sions of Devoldere and Froment.26 Twenty-four rival models
based in Scheme 1 and comprised of models with rate-
determining steps that could be single-site or dual-site adsorp-
tion, 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, reflect-
ing 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 B
) η1rc1 + η2rc2 + η3rc3 + (rt1 + rt2 + rt3) (5a)
d(W ⁄ F0EB ) FB

dXST B
) η1rc1 - η4rc4 + rt1 (5b)
Figure 7. Effect of H2-to-EB feed ratio on (a) the EB conversion, (b) the
d( W ⁄ F0EB ) FB
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.
dXBZ B
) η2rc2 + rt2 (5c)
Table 1. Pre-exponential Factors and Activation Energies for the d(W ⁄ F0EB ) FB
Thermal Reactions
i Ati [kmol/(m3 h bar)] Eti [kJ/mol] dXH2 B
) η1rc1 - η3rc3 - 2η4rc4 + (rt1 - rt2) (5d)
1 2.2215 × 10 16
272.23 d( W ⁄ F0EB ) FB
2 2.4217 × 1020 352.79
3 3.8224 × 10 17
313.06 with initial conditions
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 bar-1 8.466 1.01 8.37 6.460 10.47
KST bar-1 34.00 1.51 22.6 31.02 36.99
KH2 bar-1 3.091 0.447 6.91 2.204 3.977
k1 kmol/(kg-cat h) 0.2725 0.0171 15.9 0.2385 0.3065
k2 kmol/(kg-cat h) 0.00544 0.000504 10.8 0.00444 0.00644
k3 kmol/(kg-cat h) 0.0184 0.00874 2.11 0.001095 0.03571
k4 kmol/(kg-cat h) 0.0302 0.00565 5.66 0.0190 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)]
A1 4.594 × 109
A2 1.060 × 1015
A3 1.246 × 1026
A4 8.024 × 1010
Pre-exponential Factor of Adsorption Species j [bar-1]
AEB 1.014 × 10-5
AST 2.678 × 10-5 Figure 8. Effect of temperature on (a) the rate coefficients ki and (b) the
AH2 4.519 × 10-7 adsorption equilibrium constants Kj. Symbols represent estimated values
per temperature; lines represent calculated values from estimates at all
Activation Energy [kJ/mol] temperatures.
E1 175.38
E2 296.29
E3 474.76
E4 213.78
Adsorption Enthalpy [kJ/mol]
∆Ha,EB -102.22
∆Ha,ST -104.56
∆Ha,H2 -117.95

W
Xj ) 0 at )0
F0EB

Small catalyst particles were used to eliminate internal


diffusion limitations and, thus, obtain intrinsic kinetics. In that
case, the effectiveness factors (ηi) are equal to 1. The thermal
reactions inside the pores of the catalyst were not taken into
account. Figure 9. Comparison of calculated and experimental conversions, as a
function of space time. Symbols represent experimental data; lines represent
5.2. Parameter Estimation. The parameter estimation was calculated values using the estimates of the kinetic parameters obtained
based on the integral method as described by Froment and from all temperatures simultaneously. T ) 620 °C; H2O-to-EB molar ratio
Bischoff27 and Froment.28 The set of stiff differential equations ) 11 mol/mol; PT ) 1.044 bar; PN2 ) 0.432 bar.
(eqs 5a-d) were integrated numerically using Gear’s method.29
The parameters were estimated through the minimization of the kinetic parameters lead to an excellent fit of the experimental
multiresponse objective function, which was performed by data. By way of example, Figure 9 compares the experimental
means of the Marquardt algorithm. The minimization of the sum and calculated conversions, as a function of space time at 620
of squares of residuals can be represented by °C.
5.3. Physicochemical Tests on the Model Parameters.
n Boudart and coauthors30–32 proposed several rules for validating
∑ [y - f(x , β)]
β
S(β) ) i i
2 98 Min (6) kinetic parameters. The following test procedure was guided
i)1 by the work of Mears and Boudart,33 Van Trimpont et al.,34
Xu and Froment,35 and Froment and Bischoff.27
The parameter values, the statistical tests, and the approximate
(1) For an endothermic reaction, the thermodynamics require
95% confidence intervals derived from the experimental data
the activation energy of reaction (Ei) to exceed the heat of
at 620 °C are given in Table 2. The parameter values obtained
reaction (∆Hr,i):
from the data at all temperatures simultaneously are shown in
Table 3. The agreement between the two sets of parameters can Ei > ∆Hr,i (7)
be observed in Figure 8, which shows the temperature depen-
dence of the adsorption equilibrium constants and rate coef- The activation energies for reactions of EB to ST and EB to
ficients. The symbols represent the values of the kinetic BZ, which are endothermic reactions, are given in Table 3. They
parameters estimated from various sets of data collected at a are indeed larger than the corresponding heats of reaction at
given temperature, whereas the lines correspond to the values 893.15 K, which amount to 124.8 and 101.5 kJ/mol, respec-
estimated from the complete set of data at all temperatures. The tively, 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)
-Sa,j
0
-Sg,j
0
51 - 0.0014∆Ha,j
f)
1 - B
B 3 [ a+
b(1 - B)
Re ] (14)

component [J/(mol K)] [J/(mol K)]a [J/mol] For cylindrical packings, the coefficients a and b are 1.28
ethylbenzene, EB 95.61 361.65 194.1 and 458, respectively.38 The pressure drop between the catalyst
styrene, ST 87.53 346.25 197.4 beds is neglected.
hydrogen, H2 121.5 186.1 216.1 6.2. Continuity Equations for the Components Inside
a
Values are obtained from Stull et al.36 the Porous Catalyst. The continuity equations for the compo-
nents inside a porous catalyst that account for the thermal
(2) The heat of adsorption (-∆Ha,j) must be positive, because reactions occurring in the void space inside the catalyst particle
the adsorption is exothermic. All the estimates of the heats of are given as follows:

( )
adsorption satisfy this constraint.
(3) The adsorption entropy must satisfy 1 d 2 dPs,EB RgT
r ) [F (r + rc2 + rc3) + s(rt1 + rt2 + rt3)]
r2 dr dr De,EB s c1
0 < -∆Soa,j < Sog (8) (15a)
The inequality comes from the relation

∆Soa ) Soa - Sog (9) r


(
1 d 2 dPs,ST
2 dr
r
dr
)-
RgT
De,ST
)[Fs(rc1 - rc4) + src1] (15b)

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, r2 dr
r (
1 d 2 dPs,BZ
dr
)-
RgT
De,BZ s c2 )
(F r + srt2) (15c)

( )
because of the translational contribution to Sgo.33 The gas-phase
standard entropies of EB, ST, and H2 in Sg,j o
can be obtained 1 d 2 dPs,H2 RgT
2 dr
r )- [Fs(rc1 - rc3 - 2rc4) + s(rt1 - rt3)]
36 o
from Stull et al. ∆Sa,j was calculated from the relation r dr De,H2

(15d)
∆Soa,j ) R ln Aj (10)
with boundary conditions
Table 4 shows that the rule is satisfied.
(4) The last criterion has been applied by Everett,37 Vannice Ps,j ) Pj at r ) R
et al.,31 and Boudart et al.30
dPs,j
41.8 < -∆Soa,j e 51 - 0.0014∆Ha,j (11) ) 0 at r ) 0
dr
Table 4 shows that this rule also is satisfied. where Ps,j is the partial pressure of component j inside the
catalyst.
6. Simulation of a Three-Bed Adiabatic Reactor with The effective diffusivities are calculated from the weighted
Axial Flow binary molecular diffusivities, the void fraction of the catalyst
6.1. Continuity, Energy, and Momentum Equations. A particle (0.4), and the tortuosity factor (3.0) along the lines
multibed industrial adiabatic reactor with axial flow was explained in Froment and Bischoff.27
simulated, based on a heterogeneous reactor model (i.e., The numerical integration of this set of equations yields the
accounting for internal diffusion limitations). The steady-state profiles of the reacting species inside the catalyst particle at a
continuity equations for the reacting species have already been given position in the reactor and provides insight into the
given in eqs 5. With the catalyst particle sizes used in industrial importance of diffusion limitations on the various reactions.
production units, the effectiveness factors ηi are different from These limitations can also be expressed in terms of a single
1. For the simplified pseudohomogeneous reactor model in number: the effectiveness factor (ηi). Accounting for the rates
which diffusion limitations are not taken into account, ηi ) 1. of both the catalytic reactions and the thermal reactions in the
With the set of rate equations used here, there is no analytical void space inside the porous catalyst, the effectiveness factors
expression for the calculation of the effectiveness factors, in ηi can be calculated from these profiles by means of
terms of a modulus, so that the solution can only be obtained

V
by explicitly considering the transport equations inside the 0
[rci(Ps,j)Fs + rti(Ps,j)s] dV
catalyst particle. ηi ) (16)
[rci(Pj)Fs + rti(Pj)s]V
The energy equation is written as
6.3. Numerical Procedures. The continuity, energy, and

[ ( )
6
B momentum equationsseqs 5, 12, 13, and 14swere solved
∑ ṁ c dT
j pj 0
d(W ⁄ FEB)
) F0EB -∆Hr1 η1rc1 + rt1
FB
- numerically, using Gear’s method. For each integration step
j)1 along the reactor length, the set of equations described by eqs

(
∆Hr2 η2rc2 + rt2
B
FB ) (
- ∆Hr3 η3rc3 + rt3
B
FB )
- ∆Hr4η4rc4 (12) ] 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.
and the momentum equation is given as Calculations using nine internal collocation points led to exactly
the same results.
dPt usGF0EB 6.4. Results and Discussion. The feed conditions and reactor
- ) fR (13)
d(W ⁄ F0EB) FBdpΩ dimension for the simulation of a three-bed adiabatic reactor
with axial flow are given in Table 5. The simulation results are
The friction factor (f) is calculated using the Ergun relation: 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] 72950 82020 78330
space time [kg-cat h/(kmol EB)] 103.18 219.19 329.98
length of each bed [m] 1.33 1.50 1.43
XEB [%] 36.89 65.78 83.76
SST [%] 98.49 95.10 90.43
SBZ [%] 1.000 1.423 1.754
STO [%] 0.507 3.480 7.809
Pin [bar] 1.25 1.06 0.783
Tin [K] 886 898.2 897.6
Tout [K] 811.36 845.71 873.6

parameter value
inner radius of reactor [m] 3.50
feed molar flow rate [kmol/h]
EB 707 Figure 11. Evolution of effectiveness factors in a three-bed adiabatic axial
ST 7.104 flow reactor for Tin ) 886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB
BZ 0.293 molar ratio ) 11 mol/mol; FEB0
) 707 kmol/h.
TO 4.968
H2O 7777
total feed molar flow rate [kmol/h] 8496.37

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 limita-
tions. 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-to-
EB molar ratio ) 11 mol/mol; FEB 0
) 707 kmol/h.

diffusion limitations, expressed in terms of the effectiveness


factors (η1, η2, and η3) 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
Figure 10. Comparison between profiles predicted by the heterogeneous effectiveness factor for reaction 4 (the formation of TO from
and the pseudohomogeneous model of (a) the EB conversion and (b) the ST) is very high at the entrance, because it is a consecutive
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/ reaction.
0
mol; FEB ) 707 kmol/h. Solid line represents the heterogeneous model, Figure 12 considers an isobaric reactor and shows how
dashed line represents the pseudohomogeneous model. 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] 72 950 82 020 78 330
space time [kg-cat h/(kmol EB)] 103.18 219.19 329.98
catalyst bed depth [m] 0.614 0.708 0.681
XEB [%] 36.59 64.18 81.19
SST [%] 98.43 93.92 83.24
SBZ [%] 1.01 1.53 2.12
STO [%] 0.56 4.54 14.60
Pin [bar] 1.25 1.22 1.21
Tin [K] 886 898.2 897.6
Tout [K] 812.04 850.26 890.37

parameter value
inner radius of catalyst bed [m] 1.5
length of each reactor [m] 7
Figure 13. Simplified radial flow reactor configuration. feed molar flow rate [kmol/h]
EB 707
ST 7.104
a constant steam-to-EB ratio, increases the ST selectivity at the BZ 0.293
TO 4.968
exit from 82.18% to 90.13%. This is a consequence of the shift H2O 7777
of the equilibrium conversion to a higher value. total feed molar flow rate [kmol/h] 8496.37

7. Simulation of a Reactor with Radial Flow and Three simulations of a radial flow reactor and an axial flow reactor
Adiabatic Beds with each of three adiabatic beds, using the heterogeneous
model. The operating conditions were identical. In the reactor
The pressure drop in a radial flow reactor is much smaller with radial flow, the EB conversion amounted to 81.19%,
than that in an axial flow reactor, because of the higher cross- compared to 83.76% in the axial flow reactor. The pressure drop
sectional area of the catalyst bed. It permits one to use smaller in the radial flow reactor, with its large cross-sectional area,
particles, which leads to higher effectiveness factors. The was 0.04 bar, whereas the pressure drop amounted to 0.95 bar
differences in the performance of these two types of reactor in the axial flow reactor, as shown in Figure 15. The lower EB
are discussed below.
7.1. Continuity, Energy, and Momentum Equations. Fig-
ure 13 schematically represents a radial flow reactor configu-
ration. 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
terms of space time, W/FEB 0
, with W ) πzFB(r2 - r02):
dXj
) ηiRj (17)
d(W ⁄ F0EB)
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 4

∑ ṁ cj pj
dT
0
d(W ⁄ FEB)

) F0EB (-∆Hri)ηiri (18)
j)1 i)1

and the momentum equation is given as

dPt F0EBFgus2
- ) fR (19)
d(W ⁄ F0EB) 2πzrFBdp
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 Figure 14. Comparison of simulated (a) EB conversion profiles and (b) ST
selectivity profiles using the heterogeneous model between a three-bed
dimensions are shown in Table 6. The length of each reactor adiabatic radial flow reactor and a three-bed adiabatic axial flow reactor
and the inner radius of the catalyst bed were assumed to be 7 for Tin ) 886, 898, and 897 K; Pin ) 1.25 bar; H2O-to-EB molar ratio )
and 1.5 m, respectively. Table 6 and Figures 14 and 15 compare 11 mol/mol; FEB0
) 707 kmol/h.
Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9191

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,
Figure 15. Comparison of simulated (a) temperature profiles and (b) pressure and 897 K; H2O-to-EB molar ratio ) 11 mol/mol; FEB 0
) 707 kmol/h.
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 Figure 17. Effect of H2O-to-EB molar feed ratio on dynamic equilibrium
permits operation at much lower feed pressures. At Pin ) 0.70 coke content profiles in a three-bed adiabatic radial flow reactor for Tin )
bar, which is a low value that cannot be used in the axial flow 886, 898, and 897 K; Pin ) 1.25 bar; FEB 0
) 707 kmol/h.
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 for the coke formation was based on a two-step mechanism:
axial flow reactor for the isobaric condition shown in coke precursor formation and coke growth.39 Coke gasification
Figure 12. was assumed to occur at the edges of the carbon, which were
oxidized by water.40 The effect of the operating conditions
8. Simulation of a Radial Flow Reactor with Three (particularly the steam-to-EB feed ratio) on the dynamic
Adiabatic Beds Accounting for Coke Formation and equilibrium coke content along the three-bed adiabatic reactor
Gasification and the effect of coke formation on the reactor performance
are discussed below.
Coke formed on the potassium-promoted iron oxide catalysts 8.1. Model Equations. 8.1.1. Rate Equation for the
is at least partially removed by gasification with steam.2 Previous Formation of Coke Precursor. The formation of an irreversibly
kinetic investigations have ignored coke formation and coke adsorbed coke precursor from EB and ST, both adsorbed up to
gasification; however, more recently, Devoldere and Froment13 equilibrium with the gas phase, is assumed to be the rate-
developed detailed kinetic models for these reactions. The model determining step.27
9192 Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008

dCCP The kinetic model for coke formation and gasification was
rs ) ) δrs0ΦCp ) coupled to the kinetic model for the main reactions in the
dt
kEB,pKEBPEB + kST,pKSTPST simulations of a three-bed adiabatic reactor with radial flow,
δ × (1 - RpCCP)ns(20) using the heterogeneous model. Equations 25a-d show the
(1 + KEBPEB + KSTPST)ns continuity equations for the components, accounting for the coke
The values of δ, ns, Rs, kEB,p, and kST,p were estimated by formation from both EB and ST.
Devoldere and Froment.13,41These values were obtained on a
dXEB B rc(EB)
catalyst that was not the same as that used in the work reported ) η1rc1 + η2rc2 + η3rc3 + (rt1 + rt2 + rt3) +
here. Nevertheless, their insertion in the present model leads to d( W ⁄ F0EB ) FB 8
useful insight and reliable trends. (25a)
8.1.2. Rate Equation for Coke Growth. Further dehydro-
genation of the coke precursor forms the sites on which the dXST B rc(ST)
) η1rc1 - η4rc4 + rt1 - (25b)
coke accumulates. The intrinsic rate of coke growth can be W ⁄ F0EB
d( ) FB 8
expressed as the product of three factors: the intrinsic rate of
coke growth per active site, the total number of active sites on dXBZ B
) η2rc2 + rt2 (25c)
the growing coke, and a deactivation function: d( W ⁄ F0EB) FB

rgr ) rgr
0
CtgrΦgr (21) dXH2 B
) η1rc1 - η3rc3 - 2η4rc4 + (rt1 - rt2) +
The model for the rate of coke growth is d( W ⁄ F0EB ) FB

rgr )
dCgr
dt
) (
21
rc(EB)
8 ) (
+ 20
rc(ST)
8 )
(25d)

kEB,grPEBnEB + kST,grPSTnST Ccn1 where rc(EB) represents the rate of coke formation from
1 - RgrCgr)ngr(22)
n2 (
ethylbenzene and rc(ST) represents that from styrene.
+ KH2√PH2
n3
(1 + KH OPH O ⁄ PH
2 2 2 ) PH2 The energy equation is written

{ [ ( )]
The values of nEB, nST, n1, n2, n3, ngr, Rgr, kEB,gr, and kST,gr were 6
B
estimated by Devoldere and Froment13,41 and are used in the ∑ ṁ c j pj
dT
0
d(W ⁄ FEB)
) F0EB -∆Hr1 η1rc1 + rt1
FB
-
present work. The intrinsic rate of coke formation, accounting j)1
for the coke precursor formation and coke growth, can be
expressed as the summation of eqs 20 and 22. [
∆Hr2 η2rc2 + rt2
B
FB ( )] [ ( )]
- ∆Hr3 η3rc3 + rt3
B
FB
- ∆Hr4η4rc4 -

( ) ( )}
8.1.3. Rate Equation for Coke Gasification. The rate
rc(EB) rc(ST)
equation for coke gasification was developed under the assump- ∆HC,EB - ∆HC,ST (26)
tion that the rate-determining step is the irreversible decomposi- 8 8
tion of an oxidized carbon complex to CO and free carbon.13 and the momentum equation is unchanged, with respect to
Using the pseudo-steady-state approximation for the surface eq 19.
intermediates, the rate of coke gasification is given by The set of continuity, energy, and momentum equationsseqs

{( }
k2PH2O 25, 26, and 19swas integrated simultaneously along the reactor.
rG ) CtG (23) 8.2. Results and Discussion. Figure 17 shows the effect of
1 + K3√PH2 [(PH2 ⁄ K1) + (k2 ⁄ k1)] + PH2O
) the H2O-to-EB molar ratio on the dynamic equilibrium coke
content in a three-bed adiabatic reactor with radial flow. The
The parameter values in eq 23 were also obtained from the work dynamic equilibrium coke content was low at high H2O-to-EB
of Devoldere and Froment.13,41 feed ratios. A high steam-to-EB ratio is not always preferred in
8.1.4. Coke Formation and Gasification: Dynamic industrial operation, because of the cost of steam generation.
Equilibrium Coke Content. The EB conversions in the main At this point, optimization is required to obtain the optimum
reactions decrease until the coke content of the catalyst reaches steam-to-EB feed ratio, also accounting for the lifetime of the
a steady state. The stabilization process is very fast, so that the catalyst.
deactivation of the catalyst is limited to a very early stage of Figure 18 shows the effect of coke formation on the simulated
the operation. After it is reached, the coke content, which is EB conversion and ST selectivity in a three-bed adiabatic reactor
called the dynamic equilibrium coke content and is dependent with radial flow. Accounting for coke formation from EB and
only on the temperature and the compositions, remains con- ST leads to a drastic decrease in the ST selectivity but, because
stant.13 No deactivation effect is observed from then onward. of the influence of the equilibrium, a slight increase in the EB
At dynamic equilibrium, the rate of formation of coke precursor conversion.
and the rate of coke growth are compensated by the gasification,
so that the dynamic equilibrium coke content can be obtained 9. Conclusion
from
The extensive set of experimental data obtained on a
dCCP commercial catalyst in the experimental part of the work
) δrs0ΦCP - rG ) 0 (24a)
dt reported here provides a comprehensive basis for a more
dCgr accurate evaluation of the effect of the various operating
) rgr
0
CtgrΦgr - rG ) 0 (24b) parameters on the selectivity of styrene production from
dt ethylbenzene.
These equations also express that coke formation, like any The detailed and rigorous kinetic model that has been derived
catalytic reaction, is subject to deactivation, but gasification is from the experimental database also accounted for the back-
not. ground thermal cracking, which strongly increases with tem-
Ind. Eng. Chem. Res., Vol. 47, No. 23, 2008 9193
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
ṁj ) 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/µ
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(β) ) objective function
-∆Sa,j 0
) standard entropy of adsorption of component j, J/(mol
K)
Sg0 ) standard entropy of the gas, J/(mol K)
Figure 18. Effect of coke formation on (a) EB conversion and (b) ST Sa0 ) standard entropy of the adsorbed molecule, J/(mol K)
selectivity in a three-bed adiabatic radial flow reactor for Tin ) 886, 898, T ) temperature, K
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 ethyl- Tr ) average temperature, K
benzene and styrene; dashed lines represent results neglecting this effect. us ) superficial velocity, mf3/(mr2 s)
V ) catalyst pellet volume, mp3
perature. The optimal operation of today’s large plants also must W ) weight of catalyst, kg-cat
consider the latter aspect: the kinetic study aimed at deriving XEB ) conversion of ethylbenzene
intrinsic rate equations. The diffusion limitations encountered Xj ) conversion of ethylbenzene into component j
with the catalyst particle sizes used in industrial reactors are Z ) length of radial flow reactor, mr
introduced through the modeling. The catalytic and thermal
kinetic models were applied in simulations of the operation of Greek Letters
multibed adiabatic commercial configurations with axial or radial R ) conversion factor in momentum equation
flow that also included diffusion limitations and coke deposition β ) model parameter
and gasification. Therefore, it becomes possible to investigate, δ ) conversion factor in the rate of coke site coverage, kmol/kg-
under realistic conditions, the complex influence of the various cat
operating variables, in particular, the operating pressure and the B ) void fraction of bed, mf3/mr3
steam-to-ethylbenzene ratio, which is an important factor in the s ) catalyst internal void fraction, mf3/mp3
economics of the process. ΦCP ) deactivation function for coke precursor
Φgr ) deactivation function for coke growth
Nomenclature η ) effectiveness factor
Ai ) pre-exponential factor of catalytic reaction i, kmol/(kg-cat h) FB ) bulk density of bed, kg-cat/mr3
Aj ) pre-exponential factor for adsorption of species j, bar-1 Fg ) gas density, kg/mf3
Ati ) pre-exponential factor of thermal reaction i, kmol/(mf3 h bar) Fs ) catalyst pellet density, kg-cat/mp3
CCP ) coke precursor content, kg coke/kg-cat Ω ) cross section of reactor, mr2
Cl ) molar concentration of vacant active sites l of catalyst, kmol/
kg-cat Acknowledgment
Cp ) specific heat of fluid, kJ/(kg K)
De,j ) effective diffusivity of component j, mf3/(mr s) The authors are grateful to Dr. R. G. Anthony, Artie McFerrin
dp ) catalyst equivalent pellet diameter, mp Department of Chemical Engineering, Texas A&M, for support
Ei ) activation energy of catalytic reaction i, kJ/mol and stimulating discussions.
Eti ) activation energy of thermal reaction i, kJ/mol
Fj ) molar flow rate of j, kmol/h
Literature Cited
Fjo ) feed molar flow rate of j, kmol/h
f ) friction factor in momentum equation (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 Ullmann’s Encyclopedia of Industrial (21) Yee, A. K. Y.; Ray, A. K.; Rangaiah, G. P. Multiobjective
Chemistry; Campbell, F. T., Pfefferkorn, R., Rounsaville, J. F., Eds.; Optimization of an Industrial Styrene Reactor. Comput. Chem. Eng. 2003,
WileysVCH: Weinheim, Germany, 1994; Vol. A25, p 329. 27, 111–130.
(3) Lee, E. H. Iron-Oxide Catalysts for Dehydrogenation of Ethylbenzene (22) Sheppard, C. M.; Maier, E. E. Ethylbenzene Dehydrogenation
in Presence of Steam. Catal. ReV.sSci. Eng. 1973, 8, 285–305. Reactor Model. Ind. Eng. Chem., Process Des. DeV. 1986, 25, 207–210.
(4) Coulter, K.; Goodman, D. W.; Moore, R. G. Kinetics of the (23) Schule, A.; Shekhah, O.; Ranke, W.; Schlogl, R.; Kolios, G.
Dehydrogenation of Ethylbenzene to Styrene over Unpromoted and K- Microkinetic Modelling of the Dehydrogenation of Ethylbenzene to Styrene
Promoted Model Iron-Oxide Catalysts. Catal. Lett. 1995, 31, 1–8. over Unpromoted Iron Oxides. J. Catal. 2005, 231, 172–180.
(5) Stobbe, D. E.; van Buren, F. R.; van Dillen, A. J.; Geus, J. W. (24) Sundaram, K. M.; Froment, G. F. Modeling of Thermal-Cracking
Potassium promotion of iron-oxide dehydrogenation catalysts supported on Kinetics. 1. Thermal-Cracking of Ethane, Propane and Their Mixtures.
magnesium-oxide. 1. Preparation and characterization. J. Catal. 1992, 135, Chem. Eng. Sci. 1977, 32, 601–608.
533. (25) Sundaram, K. M.; Sardina, H.; Fernandez-Baujin, J. M.; Hildreth,
(6) Muhler, M.; Schlogl, R.; Reller, A.; Ertl, G. The Nature of the Active J. M. Styrene Plant Simulation and Optimization. Hydrocarbon Process.
Phase of the Fe/K-Catalyst for Dehydrogenation of Ethylbenzene. Catal. 1991, (January), 93.
Lett. 1989, 2, 201–210. (26) Devoldere, K. R.; Froment, G. F. Unpublished results, 2000.
(7) Hirano, T. Active Phase in Potassium-Promoted Iron-Oxide Catalyst (27) Froment, G. F.; Bischoff, K. B. Chemical Reactor Analysis and
for Dehydrogenation of Ethylbenzene. Appl. Catal. 1986, 26, 81–90. Design, 2nd Edition; Wiley: New York, 1990.
(8) Hirano, T. Roles of Potassium in Potassium-Promoted Iron-Oxide (28) Froment, G. F. Model Discrimination and Parameter Estimation in
Catalyst for Dehydrogenation of Ethylbenzene. Appl. Catal. 1986, 26, 65– Heterogeneous Catalysis. AIChE J. 1975, 21, 1041–1057.
79. (29) Gear, C. W. Numerical Initial Value Problems in Ordinary
(9) Hirano, T. Dehydrogenation of Ethylbenzene over Potassium- Differential Equations; Prentice Hall: Englewood Cliff, NJ, 1971.
Promoted Iron-Oxide Containing Cerium and Molybdenum Oxides. Appl. (30) Boudart, M.; Mears, D. E.; Vannice, M. A. Ind. Chim. Belge. 1967,
Catal. 1986, 28, 119–132. 32, 281.
(10) Muhler, M.; Schutze, J.; Wesemann, M.; Rayment, T.; Dent, A.; (31) Vannice, M. A.; Hyun, S. H.; Kalpakci, B.; Liauh, W. C. Entropies
Schlogl, R.; Ertl, G. The Nature of the Iron Oxide-Based Catalyst for of Adsorption in Heterogeneous Catalytic Reactions. J. Catal. 1979, 56,
Dehydrogenation of Ethylbenzene to Styrene. 1. Solid-State Chemistry and 358.
Bulk Characterization. J. Catal. 1990, 126, 339–360. (32) Boudart, M. Two-Step Catalytic Reactions. AIChE J. 1972, 18, 465.
(11) Addiego, W. P.; Estrada, C. A.; Goodman, D. W.; Rosynek, M. P. (33) Mears, D. E.; Boudart, M. The Dehydrogenation of Isopropanol
An Infrared Study of the Dehydrogenation of Ethylbenzene to Styrene over on Catalysts Prepared by Sodium Borohydride Reduction. AIChE J. 1966,
Iron-Based Catalysts. J. Catal. 1994, 146, 407–414. 12, 313.
(12) Shekhah, O.; Ranke, W.; Schlogl, R. Styrene Synthesis: In situ (34) Van Trimpont, P. A.; Marin, G. B.; Froment, G. F. Kinetics of
Characterization and Reactivity Studies of Unpromoted and Potassium- Methylcyclohexane Dehydrogenation on Sulfided Commercial Platinum
Promoted Iron Oxide Model Catalysts. J. Catal. 2004, 225, 56. Alumina and Platinum-Rhenium Alumina Catalysts. Ind. Eng. Chem.
(13) Devoldere, K. R.; Froment, G. F. Coke Formation and Gasification Fundam. 1986, 25, 544–553.
in the Catalytic Dehydrogenation of Ethylbenzene. Ind. Eng. Chem. Res. (35) Xu, J. G.; Froment, G. F. Methane Steam Reforming, Methanation
1999, 38, 2626–2633. and Water-Gas Shift. 1. Intrinsic Kinetics. AIChE J. 1989, 35, 88–96.
(14) Wenner, R. R.; Dybdal, E. C. Catalytic Dehydrogenation of (36) Stull, D. R.; Westrum, E. F., Jr.; Sinke, G. C. The Chemical
Ethylbenzene. Chem. Eng. Prog. 1948, 44, 275. Thermodynamics of Organic Compounds; Wiley: New York, 1969.
(15) Carra, S.; Forni, L. Kinetics of Catalytic Dehydrogenation of (37) Everett, D. H. The Thermodynamics of Adsorptions. Part II.
Ethylbenzene to Styrene. Ind. Eng. Chem. Process Des. DeV. 1965, 4, 281. Analysis and Discussion of Experimental Data. Trans. Faraday Soc. 1950,
(16) Sheel, J. G. P.; Crowe, C. M. Simulation and Optimization of an 46, 957.
Existing Ethylbenzene Dehydrogenation Reactor. Can. J. Chem. Eng. 1969, (38) Handley, D.; Heggs, P. J. Momentum and Heat Transfer Mecha-
47, 183. nisms in Regular Shaped Packings. Trans. Inst. Chem. Eng. 1968, 46, T251.
(17) Elnashaie, S. S. E.H.; Abdalla, B. K.; Hughes, R. Simulation of (39) Beeckman, J. W.; Froment, G. F. Catalyst Deactivation by Active-
the Industrial Fixed-Bed Catalytic Reactor for the Dehydrogenation of Site Coverage and Pore Blockage. Ind. Eng. Chem. Fundam. 1979, 18, 245–
Ethylbenzene to StyrenesHeterogeneous Dusty Gas-Model. Ind. Eng. Chem. 256.
Res. 1993, 32, 2537–2541. (40) Mims, C. A.; Chludzinski, J. J.; Pabst, J. K.; Baker, R. T. K.
(18) Dittmeyer, R.; Hollein, V.; Quicker, P.; Emig, G.; Hausinger, G.; Potassium-Catalyzed Gasification of Graphite in Oxygen and Steam. J.
Schmidt, F. Factors Controlling the Performance of Catalytic Dehydroge- Catal. 1984, 88, 97–106.
nation of Ethylbenzene in Palladium Composite Membrane Reactors. Chem. (41) Froment, G. F. Unpublished work, March 2005.
Eng. Sci. 1999, 54, 1431–1439.
(19) Kolios, G.; Eigenberger, G. Styrene Synthesis in a Reverse-Flow ReceiVed for reView August 8, 2007
Reactor. Chem. Eng. Sci. 1999, 54, 2637. ReVised manuscript receiVed December 12, 2007
(20) Savoretti, A. A.; Borio, D. O.; Bucala, V.; Porras, J. A. Non- Accepted December 14, 2007
adiabatic Radial-flow Reactor for Styrene Production. Chem. Eng. Sci. 1999,
54, 205–213. IE071098U

You might also like