Vous êtes sur la page 1sur 4

Ind. Eng. Chem. Process Des. Dev.

the effective light path length instead of the film thickness,


S, agree well with the experimental results as shown in
Figures 7 and 8.
The relative volumetric rate of light absorption can also
be determined from the correlation for a correction factor
based on the diffuse light model shown in Figure 10.

We are grateful to the students at the University of


Tokushima, Mitsuyoshi Hara, Makoto Etoh, Masahiro
Yokota, Kazuo Shiramizu, and Makoto Takahashi for their
assistance in the experimental work.
Nomenclature
C = concentration, M or mol L-l
C0 = initial concentration, M
DT = diameter of reactor, l

=
=

=
=

207

liquid phase
single phase
two-phase
at reactor wall

Greek Letters

molar absorptivity, L-l M-1


parameter defined by eq 9
angle of rotation with coordinate center at radial position
r, radian
= wavelength, l
= attenuation coefficient, {-1
= quantum yield of ferrous production, mol einstein-1
correction factor defined by eq 10
Registry No. Potassium ferrioxalate, 14883-34-2.

a =

/3

Acknowledgment

IJ

J.t

t/;

Literature Cited
Akehata, T.; ltoh K.; Inokawa,

= relative radiant energy


light intensity, einstein t-2 t-1
average light intensity over reactor cross-sectional area,
einstein 1-2 t-1
(I) = volumetric light intensity, einstein 1-2 r1
Ua>
volumetric rate of light intensity, einstein t-3 t-1
L. = effective light path length, l
r = radial distance, l
R = radius of reactor, l
RL = liquid holdup
S
liquid film thickness, l
T(X) = transmittance
U
velocity, l r1
z = axial distance, l
.:lz
axial length of optical window, l

E(X)

I
l

1986, 25, 207- 2 10

Subscripts

dif = diffuse light model


Fe3+ = ferric ion
Fe2+ = ferrous ion

A. Kagaku Kogaku Ronbunshu 1976, 2, 583.


Cassano, A. E .; Silveston,P. L.; Smith, J. M. Ind. Eng. Chem. 1967, 59, 19.
Governale, L. J.; Clarke, J. T. Chem. Eng. Prog. 1956, 52, 281.
Hill, F. B.; Reiss, N.; Shendalman, L. H. AIChE J. 1968, 14, 798.
Itoh, M.; Hara, Y. JpnPatent, 3 925 041, 1964.
Matsuura, T.; Smith, J. M. AIChE J. 1970, 16, 321.
Otake, T.; Tone, S.; Kono, K.; Nakao, K. J. Chem. E ng. Jpn. 1979, 12, 289.
Otake, T.; Tone, S.; Higuchi, K.; Nakao, K. Kagaku Kogaku Ronbunshu 1981,
7, 57.
Otake, T.; Tone, S.; Higuchi, K.; Nakao, K. Int. Chem. Eng. 1983, 23, 288.
Parker, C. A. Proc. R. Soc. London, Ser. A 1953, A 220 , 104.
Ramage, M.P.; E ckert, R. E . Ind. Eng. Chem. Fundam. 1975, 14, 214.
Shlrotsuka, T.; Sutoh, M. Kagaku Kogaku Ronbunshu 1978, 4, 502.
Spadoni, G.; Bandini, E .; Santarelli, F. Chem. Eng. Sci. 1978, 33, 517.
Tomlda, T.; Yoshida, M.; Okazaki, T. J. Chem. Eng. Jpn. 1976, 9, 464.
Tomida, T.; Yusa, F.; Okazaki, T. Chem. Eng. J. 1978, 16, 81.
Tomida, T.; Tabuchi, F.; Okazaki, T. J. Chem. Eng. Jpn. 1982, 15 , 434.
Troniewski, L.; Ulbrich, R. Chem. Eng. Sci. 1984, 39, 751.
Yokota, T.; Iwano, T.; Deguchi, H.; Tadaki, T. Kagaku Kogaku Ronbunshu
1981a, 7, 157.
Yokota, T.; Iwano, T.; Saito, A.; Tadaki, T. Kagaku Kogaku Ronbunshu
1981b, 7, 164.
Williams, J. A. AIChE J. 1978, 24, 335.
Zolner, W. J.; Williams, J. A. AIChE J. 1971, 17, 502.

Received for review October 1, 1984


Accepted June 24, 1985

Ethylbenzene Dehydrogenation Reactor Model


Charles M. Sheppard* and Edward E. Maler
USS Chemicals, Monroeville, Pennsylvania

15146

Hugo S. Caram
Department of Chemical Engineering, Lehigh University, Bethlehem, Pennsylvania

18015

A model of an industrial catalytic dehydrogenation reactor was developed. Several kinetic models were calibrated
by using catalyst manufacturers data. The calibrated Langmuir-Hinshelwood models did not represent the data
well, so an empirical model was selected and a comparison was made between simulated and actual industrial
operations for two different plants. The optimum operating conditions were explored for one- and two-bed reactor
configurations by using two industrial catalyst systems.

The purpose of this work was the development of a


model to simulate an industrial ethylbenzene dehydroge
nation reactor and the use of this model to perform eco
nomic analyses of different operating conditions, reactor
configurations, and catalysts systems.
0 196-4305/86/ 1 125-0207$0 1.50/0

Styrene is commercially produced by the dehydroge


nation of ethylbenzene over an iron oxide catalyst in the
presence of steam. Steam is used to supply heat for this
endothermic reaction, to inhibit coke formation, and to
reduce the partial pressure of the products and thus in 1985 American Chemical Society

208

Ind. Eng. Chem. Process Des. Dev., Vol. 25, No. 1, 1986

Table I. Summary of the Kinetic Models Investigated

driving force for the


reaction generating
model

author

3
3b
4
5'

"z

styrene

toluene

styrene

benzene

toluene

E.

11.5 24.1
PEB
PEB - (PsTYPH2/ PEBPH2
K.q)
Carra
18.3 38.2
PEB +
PEB
PEB - (PsTvPHz! PEB
zPsTY
K.q)
(1 +
17.5 30.6
Lebedev
PEB
PEB - (PsTYPH2/ PEBPH2
zPsTY)2
K.q)
(1 +
17.2 30.2
PEB
PEB - (PSTYPHz/ PEB
zPsTY)2
K.q)
Sheppard PEB +
38.1 65.9
PEB
PEB - (PsTYPH2/ PEB
(zPsTv)2
K.q)
42.3 74.5
Sheppard PEB +
PEB
PEB - (PSTYPH2/ PEB
(zPsTY)2
K.q)
(0.0218 atm) exp(T/6995 K). 6Toluene reaction driving force modified.
Wenner

inhibitor
none

crease the equilibrium conversion.


The three key reactions are

E.

benzene

E.

16.6

31.3 13.2

33.2

20.6

46.8

45.2

14.4 16.02

16.9

styrene toluene

benzene

9.6

2.4

0.6

8.0

4.0

0.8

0.6

34.1

3.1

2.0

0.5

51.4 18.3

37.8

2.7

1.3

0.6

47.9

87.4 39.9

74.9 100

2.3

0.9

0.6

61.4

115.5 61.6

114.4 100

2.0

0.6

0.7

9.1

18.9

'Model4 fit for Shell 015 catalyst.

10 -------r--
9
8

EB =STY+ H2

(1)

EB+ H2 =TOL+ CH4

(2)

EB =BEN+ C2 H4

(3)

In addition, other aromatics present may be dehydroge


nated or dealkylated, and coke on the catalyst surface
reacts with the steam. Oxidation and the water gas shift
reaction also occur; these reactions involve species which
are vapor at room temperatures and will be referred to as
the gas reactions. The following three reactions adequately
represent these reactions and complete the set of reactions
for the model:
CH4+ H20 =CO+ 3H2

(4)

1/ 2C2H4+ H20 =CO+ 2H2

(5)

CO+ H20 =C02+ H2

(6)

A model was written to simulate a plug flow reactor


using a published integration package (EPISODE, 1975)
to integrate the differential component, momentum, and
energy balance down the reactor (Sheppard, 1982). No
dispersion effects were included in the component or en
ergy balances. The model can be used to simulate either
an isothermal or an adiabatic reactor.
Selection of a Kinetic Model

av absolute deviation
between the predicted
and
manufacturer's data

kinetic const for the


calibrated models

In seeking to simulate an ethylbenzene dehydrogenation


reactor, the first order of business is to select a kinetic
model. The kinetic data in the literature include a simple
Arrhenius relationship proposed by Wenner and Dybdal
(1948) and Langmuir-Hinshelwood mechanisms proposed
by Carra and Forni (1965) and by Lebedev et al. (1978).
An empirical model was proposed in order to reconcile the
observation of Carra and Forni (1965) that the initial re
action rate was independent of the partial pressure of
ethylbenzene and the observation of Lebedev et al. (1978)
that the inhibition term was approximately proportional
to the partial pressure of styrene squared. These rate
expressions (see Table I) were evaluated by using iso
thermal integral data provided by the catalyst manufac
tures to see which best describes the kinetic behavior of
the catalyst. Data for Shell 105 were used because this
catalyst was being employed. Also, both Shell and United
Catalyst Inc. supplied information for this catalyst.
The computer model was used to simulate the isother
mal conditions of the manufactures conversion data.
Kinetics taken from the literature for the gas reactions (eq

-- LEAST SQUARES FIT

0 SHELL- SET 1
c SHELL- SET 2
e lJIIITEOCATALYST -SET 3
lJIIITEOCATALYST- SET4
lolJIIITEOCATALYST-SET5

Y lJIIITEOCATALYST-SET6
SHELL105 C A T ALYST

kg-cat hr

0.9
0.8
0.7

n.s
0.5
0.4

0.3

0.2

.__
....____
..
....____
._
.L-__--I____....L.__
.
........
.llt,
.._
__,
__
0.1 ....__
1.10

1.12

1.14

1.16

1.18

1.20

1.22

103/T IK'1J

Figure 1. Arrhenius plot of the apparent reaction rate constants for


the main reaction of model 4.

4-6) were used in the model (Akers and Camp, 1955; Moe,
1962). The model was used in conjunction with a modified
Leverberg-Marquart search algorithm which located the
apparent reaction rate constants for the three reactions
involving ethylbenzene. The analysis includes the effect
of the gas reactions and the inhibition term. These ap
parent rate constants were used in conjunction with a
linear least-squares regression to calculate the Arrhenius
constants (see Figure 1). After the four kinetic models
were calibrated, they were used to simulate these same
manufactures conversion data, and the average deviation
between the models and data are summarized in Table I.
From this table, it is apparent that model 4 is the best
model of those considered. Table I also contains the pa
rameters for the Arrhenius equation and the inhibition
term. The activation energies calculated for models 4 and
5 are higher than can be easily justified on physical

Ind. Eng. Chern. Process Des. Dev., Vol. 25, No. 1, 1986

209

Table II. Comparison of Model 4 and Plant Data


pressure
in, atm

pressure
out, atm

% conversion

run
no.

temp
in, C

temp
out, C

model
data
model
data
model
data
model
data

1.
1
2
2
3
3
4
4

649.44
649.44
648.85
648.85
647.18
647.18
649.44
649.44

570.43
593.33
570.10
597.83
569.83
592.62
568.0

Polymer Corporation Data


40.63
1.953
2.37
39.80
2.29
2.37
40.71
2.37
1.953
40.72
2.73
40.30
1.955
2.31
38.35
2.37
40.60
2.71
1.795
43.22
2.71

model
data
model
data
model
data
model
data
model
data

5
5
6
6
7
7
8
8
9
9

617.78
617.78
632.22
632.22
623.33
623.33
628.89
628.89
628.89
628.89

555.44
558.33
568.33
571.11
555.37
560.00
559.38
565.56
561.20
562.22

1.952
1.952
2.429
2.429
1.884
1.884
1.680
1.680
1.816
1.816

styrene

benzene

toluene

styrene
selectivity, %

1.91
2.99
1.90
2.99
1.87
2.53
1.89
2.67

4.65
2.29
4.62
3.17
4.50
2.40
4.58
2.79

86.11
87.09
86.18
86.86
86.35
88.60
86.25
88.78

1.77
1.52
1.81
1.43
1.78
1.23
2.16
1.34
2.16
1.87

3.96
3.62
4.48
3.17
4.03
2.13
5.38
2.52
5.70
3.95

88.39
89.23
86.72
89.15
87.99
92.11
86.68
91.67
86.94
89.16

USS Chemicals Data

grounds. The two-site styrene adsorption term dominates


the inhibitor. So part of the high activation energies may
be the contribution of the styrene adsorption/desorption
temperature dependence.
It was found that the reactions were relatively insensitive
to the kinetics of the gas reactions. However, a possible
refinement to the model would be the empirical correction
of the kinetics for reactions 4--fi, since the predicted gaseous
product rate, when simulating plant operating conditions,
is about 2.5 times less than the observed rate. Lee (1973)
suggested that intraparticle diffusion may be important,
but effectiveness factor calculations did not show a sig
nificant diffusion effect even when a tortuosity factor of
20 (instead of 3 which is typical) was used (Sheppard,
1982).
Testing the Kinetic Model against Plant Data

The reactor model was then run by using these kinetic


constants to simulate industrial reactors. Plant conditions
for the USS Chemicals plant of Houston, TX (Sheppard,
1982), and for the Polymers Corp. of Ontario, Canada
(Sheel, 1968), were used in this comparison. As expected,
the range of plant operating conditions is narrow. The
inlet temperature, pressure, flow rate, and composition
were specified for the model as well as adiabatic reactor
operation. The pressure drop was calculated by using the
Ergun equation and a void fraction of 0.325. Table II gives
a comparison of the ethylbenzene conversion and selec
tivity for several representative points. With the exception
of the Polymer Corp. conversion to benzene, the predicted
ethylbenzene conversions were higher than the actual
conversions for all three reactions. The styrene selectivity,
however, agrees well. The higher predicted conversions
may be due to catalyst deactivation with age. One cause
for this deactivation may be the potassium migration as
described by Lee (1973) and by Herzog and Rase (1984).
For the almost 100 cases simulated, the average difference
between the predicted and measured conversions to sty
rene was less than 10%; this also confirms the model
ability to predict plant behavior. Also, 85% of these cases
showed an actual conversion lower than the predicted
conversion. This again suggests catalyst deactivation.
Economic Analysis of Catalysts and Reactor
Configurations

The styrene industry trend in the U.S.A. has been to


move to multibed reactor systems (tO approach isothermal

1.766
1.755
2.231
2.170
1.734
1.721
2.476
1.572
1.621
1.707

43.66
42.59
41.08
41.16
42.61
39.18
49.07
43.17
48.99
47.87

operating conditions) and to higher selectivity catalysts.


These high selectivity catalysts typically can be operated
with lower steam-to-oil ratios and not experience deacti
vation problems caused by coke buildup. The lower steam
to oil ratio results in an energy savings. The USS Chem
icals ethylbenzene dehydrogenation unit employs a single
reactor bed and the high activity, lower selectivity, Shell
105 catalyst.
An economic analysis to quantify the benefits of using
a high selectivity catalyst was performed. This involved
using manufacturer's data to fit the kinetic parameters for
kinetic expressions of the same form as those used for Shell
105. Shell 015 was chosen since most of the data available
for high selectivity catalysts were for this catalyst. The
kinetic parameters calculated which best represent the
catalyst behavior are given in Table I under model 5. The
economic analysis also involved choosing an objective
function. The objective function (also referred to as
"profit") was (the styrene production rate) X (the styrene
selling price)- (the styrene production cost). A styrene
selling price of 55.1/kg (25 /lb) was used. The styrene
production cost was calculated based on a fixed cost of
$190/h, a material cost of ethylbenzene of 34.6/kg
(15.7/lb), a vent gas credit of 4.63/std m3 (0.131 /std
ft3), and an energy cost of $3.94/billion J ($4.16/million
BTU). The vent gas credit was multiplied by a factor of
2.5 in the analysis to account for the fact that the model
consistently underpredicted the vent gas production. The
plant energy requirements are dominated by the energy
lost in condensing the reactor effluent and the steam used
in the recycle column reboilers. The design condenser duty
was 1000 kJ/kg (945 Btujlb) of the reactor effluent, and
the design recycle columns reboiler duties were a total of
1740 kJ/kg (1650 Btujlb) of feed. These costs are calcu
lated by scaling the heat duties by the flow rates and using
the design values as the base point. A fixed cost of 1.0/kg
(0.458/lb) is used for the B/T column reboiler and the
styrene finishing energy costs. The other energy user
considered is the steam boiler duty. Mo.e details of this
analysis are given by Sheppard (1982).
The economic analysis of styrene production in a sin
gle-bed reactor (having of the same configuration as the
USS Chemical reactor) included investigating three dif
ferent steam-to-oil ratios for the two different catalysts
over a range of reactor inlet temperatures. The results are

210

Ind. Eng. Chern. Process Des. Dev., Vol. 25, No. 1, 1986
-------T---r----
CATAL VST
015

MOLAR STEAM
TO Oil RATIO

14.1

10.0

"

60

200

100
"'..

"'..

..:

t::'

ii:
0
0::

"'
0
0::
0..

0..

MOLAR STEAM TO
OIL RATIO
STAGE 1

-100

STAGE 2

INTERSTAGE
HEATING VIA

8.0

8.0

8.0

11.8

STEAM INJECTION

"'

11.8

11.8

HEAT EXCHANGER

HEAT EXCHANGER

650
REACTOR INLET TEMPERATURE, oC

Figure 2.
reactor.

Profit vs. reactor inlet temperature for a single-stage

presented in Figure 2. From this figure, it is seen that


some gains in profitability for the high activity catalyst
can be made by optimizing the steam-to-oil ratio but the
gains in profitability by shifting to a higher selectivity
catalyst are much greater.
For the analysis of a two-bed reactor system, there is also
the question of how to perform the interstage heating.
Two options are an interstage heat exchanger or interstage
steam injection. If steam is to be injected between the
stages, then it has been suggested to use a lower steam
to-oil ratio in the first bed. These suggestions were in
vestigated over a range of inlet temperature for the Shell
015 catalyst (since the economics favor high selectivity
catalyst); see Figure 3. From the graph, it can be seen that
the high steam-to-oil ratio case using an interstage heat
exchanger is the most profitable. However, this analysis
does not consider the capital cost difference between the
two options or the possibility of polymer buildup in the
heat exchanger.
Conclusions

A model was developed which simulates an industrial


ethylbenzene dehydrogenation reactor. Kinetic expressions
were calibrated for both a high activity catalyst and a high
selectivity catalyst. An empirical model provided the best
fit on manufactures and plant data, but because of the high
calculated activation energies, doubt remains about the
theoretical foundation of the model. This model was then
used to locate the optimum inlet temperature and steam
to-oil ratio for a specified styrene selling price and a set
of material and operating costs. The model was also used
to investigate the economics of installing a two-bed reactor
system. The economics of using a high selectivity catalyst

REACTOR INLET TEMPERATURE, C

Figure 3. Profit vs. reactor inlet temperature for a two-stage reactor


packed with Shell 015.

were superior to the high activity catalyst.


Nomenclature

frequence factor for reaction j


activation energy for reaction j, kcal/ (kg mol)
kj = kinetic constant for reaction j, kg mol/ (kg of catalyst h
atmn) = exp(Aj - EaJfRaT)
rj = reaction rate constant, kg mol/ (kg of catalyst h) = kj
(driving force)/ (inhibitor)
Ra ideal gas constant
styrene conversion or conversion to styrene = (styrene pro
duced) j (ethylbenzene initially present)
styrene selectivity = (styrene produced) j(ethylbenzene consumed)
T = absolute temperature, K
z = absorption coefficient
f = bed void fraction used in the Ergun equation = 0.325
A

F!aJ

Registry No.

EB, 100-41-4; STY, 100-42-5.

Literature Cited
Akers, W. W.; Camp, D.P. AIChE J. 1955, 1, 4 71.
Carra, Forni Ind. Eng. Chem. Process Des. Dev. 1965, 4, 28.
EPISODE (ExperimentalPackage for Integration of Systems of Ordinary Dif
ferential Equations), June 24, 1975 Version; an Argonne National Labora
tory (Chicago, IL) modification of an earlier version by G. D. Bryne and A.
C. Hlndmarsh.
Herzog, B. D.; Rase H. F. Ind. Eng. Chem. Prod. Res. Dev. 1984, 23, 187.
Lebedev, N. N.; Odabashyan, G. V.; Lebedev, V. V.; Makorov, M. G. Kinet.
Kalal. 1978, 18, 1177-1182.
Lee, H. Catal . Rev. 1973, 8, 285.
Moe, J. M. Cham. Eng. Prog. 1962, 58 (3), 33.
Sheel, J. G. P. Eng. Thesis, McMaster University, Hamilton, Ontario, Canada,
1968.
Sheppard C. M. M.S. Thesis, Lehigh University, Bethleham, PA, 1982.
Wenner R. R.; Dybdal, E. C. Cham. Eng. Prog. 1948, 44 (4), 275.

Received for review Ja nuary 10, 1985


Accepted June 21, 1985

Vous aimerez peut-être aussi