You are on page 1of 9

Chemical Engineering Science 54 (1999) 205—213

Non-adiabatic radial-flow reactor for styrene production


A.A. Savoretti, D.O. Borio*, V. Bucalá, J.A. Porras
Planta Piloto de Ingenierı& a Quı& mica, UNS-CONICET, 12 de Octubre 1842, (8000) Bahı& a Blanca, Argentina
Received 9 December 1997; accepted 23 June 1998

Abstract

A non-adiabatic radial reactor is proposed to carry out the dehydrogenation of ethylbenzene to styrene. Radial flow and
continuous heating (using superheated steam) are the main features of the new design. Steam used as heating medium flows through
tubes, which are radially installed in the catalyst bed. By means of steady-state simulations, this new design has been compared with
two adiabatic beds with radial flow and reheating between stages (similar to those used in industry). For equal steam consumption, the
proposed design leads to higher selectivity to styrene than the industrial adiabatic design. This enhancement in selectivity (which is
observed for different conversion levels) would significantly improve the economics of the styrene production process.  1998
Elsevier Science Ltd. All rights reserved.

Keywords: Styrene production; Non-adiabatic reactors; Radial fixed-bed reactors; Dehydrogenation of ethylbenzene

1. Introduction existing ethylbenzene dehydrogenation reactors. On the


other hand, since radial reactors allow lower pressure
Dehydrogenation reactions require high-temperature drops and good flow distribution, they are specially suit-
levels to favor the chemical equilibrium and the reaction able for processes requiring low pressures to increase the
rate. However, the temperature cannot be excessively equilibrium conversion. They present as additional ad-
increased because hydrocarbon cracking, selectivity vantage the possibility of using smaller catalyst pellets
drops or catalyst deactivation can occur. The rate of with the consequent increase of plant capacity. For all
dehydrogenation reactions diminishes as conversion in- these reasons, radial reactors are being preferred to per-
creases, not only because equilibrium is approached, but form the dehydrogenation of ethylbenzene (Sundaram
also in many cases reaction products act as inhibitors et al., 1991).
(Voge, 1982). Pure dehydrogenations are endothermic Due to the endothermicity of the main reaction, all the
(15,000—35,000 kcal/kmol), and hence have large heat adiabatic reactors used for styrene production lead to
requirements. decreasing temperature profiles along the catalyst bed.
Most of the industrial styrene plants carry out the However, Voge (1982) and Eigenberger (1992) suggested
dehydrogenation reaction adiabatically in multiple reac- that the ideal temperature profile (to maximize the
tors operating in series. The heat of reaction is supplied at styrene yield) would probably show an increase along the
the inlet of each stage by means of superheated steam or reactor length. Therefore, a continuous heating along the
by heat exchangers (James and Castor, 1994). reaction path can be thought as an adequate strategy to
From the standpoint of the reactants flow configura- approach the desirable shape of the temperature profile.
tion, two reactor designs are commonly used in styrene Non-adiabatic reactor designs have already been pro-
plants: axial and radial flow reactors. Axial flow posed for styrene production. The called ‘isothermal de-
adiabatic reactors have been analyzed by Sheel and hydrogenation’ was pioneered by BASF. The reactor is
Crowe (1969) and Sheppard et al. (1986). These authors a shell and tube heat exchanger, the reactants flow axially
carried out studies of simulation and optimization of through the packed tubes, while a heating medium (hot
flue gas) circulates through the shell side (James and
* Corresponding author. Tel.: 00 54 91 861 700; fax: 00 54 91 861600; Castor, 1994). On the other hand, Lurgi GmbH operated
e-mail: reborio@criba.edu.ar. an ‘isothermal reactor system’ (a multitubular fixed-bed

0009-2509/99/$ — see front matter  1998 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 9 8 ) 0 0 1 9 5 - X
206 A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213

reactor) that uses a molten salt mixture (molten carbon- Table 1


ates) as heating medium (Short and Bolton, 1985; James Kinetic information
and Castor, 1994).
k (p !(p p /K ))
A novel non-adiabatic fixed-bed reactor is proposed in C H CH CH   C H CHCH #H r"     CO
    \      p #(80p )
this paper. Radial flow and continuous heating (using  
superheated steam) are the main characteristics of the k p
C H CH CH P C H CH CH r "  
new design. A continuous heating along the radial coor-          p #(80p )
 
dinate is proposed to increase the reactor selectivity. k p
 
C H CH CH #H P  C H CH #CH r "
Radial flow is selected to minimize the pressure drop and           p #(80p )
 
to improve the reactor capacity. The performance of the 0.5CH CH #H O P  CO#2H r "k p p
       
CH #H O P  CO#3H r "k p p
proposed design is compared with that of an industrial       
unit consisting of two adiabatic radial fixed beds with 
CO#H O  CO #H r "(k p p
 \      
interstage heating. !k p p )
\  
kmol
k "3.3489;10 exp (!275.9/R¹),
 kgcat hr
2. Kinetic scheme for styrene production kmol
k "2.2168;10 exp (!313.6/R¹),
 kgcat hr
A system of six catalytic reactions is used to represent
kmol
the industrial process of dehydrogenation of ethylben- k "6.54236;10 exp (!365.86/R¹),
 kgcat hr
zene. Thermal cracking reactions are neglected. The kin-
etic expressions used in this paper were obtained by kmol
k "0.124 exp (!66.14/R¹),
 kgcat hr kPa 
Sheppard (1982) for a commercial catalyst composed by
Fe O , Cr O and K O. Information about this kinetic kmol
     k "0.0123 exp (!66.14/R¹),
model is given in Table 1.  kgcat hr kPa
kmol
k "864.9 exp (!132.03/R¹),
 kgcat hr kPa
3. Optimal temperature profiles kmol
k "65731.50 exp (!255.39/R¹),
\ kgcat hr kPa
In this work we seek a new reactor design capable to
  
!29814.8 1 1
maximize the styrene production, which strongly de- K "38.09 exp ! , kPa
 R ¹ 900
pends on the temperature profile along the reactor.
Therefore, it is important to know the shape of the
optimal temperature profile (which allows one to achieve
the optimal outlet product distribution) before designing
The subscript i denotes all the chemical species present
the reactor.
in the kinetic scheme given in Table 1 [i.e., (1) ethylben-
A general method for process optimization, developed
zene, (2) styrene, (3) hydrogen, (4) benzene, (5) ethylene,
by Pontryagin et al. (1962), was selected to find the
(6) toluene, (7) methane, (8) water, (9) carbon monoxide,
optimal temperature trajectories. Using just the reactor
(10) carbon dioxide]. Chemical reactions are noted by the
mass balances, this method allows to obtain a continuous
subscript j. The initial conditions are F (0)"F .
thermal profile which maximizes the objective function G G
The optimization problem consists in the calculation
without exceeding the maximum allowable temperature
of the temperature profile ¹(q) which maximizes the
value (¹ ). This upper constraint was specified to limit
outlet reactor yield without surpassing ¹ . The objective
the influence of the thermal cracking reactions.
function is defined as
To determine the optimal temperature profiles, only
the main reactions (1)—(3) of Table 1 were considered.
F (q )!F
The reliability of this simplification was verified compar- g" 127 D 127 . (2)
F
ing the results (for some particular conditions) with those # 
obtained using the whole set of six chemical reactions
given in Table 1. The optimal temperature profiles were obtained ap-
The reactor was represented by a pseudohomo- plying the maximum principle to Eqs. (1) and (2) follow-
geneous plug-flow model and a constant pressure was ing the guidelines given by Kirk (1970) and Mufti (1970).
assumed. The mass balances for an ideal fixed bed are The optimal trajectory must verify that the Hamiltonian
reaches its maximum value at any 0(q(q . The
given by D
Hamiltonian is defined as the product of the adjoint
dF  l functions and the right-hand sides of Eq. (1) (Westerterp
G"!¼ GH r , i"1, 10 (1)
dq E l H et al., 1982).
H H
A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213 207

Fig. 2. Typical temperature trajectories in industrial adiabatic reactors


Fig. 1. Influence of ¹ on optimal temperature profiles. Inlet with intermediate heating.

molar flows: F "0.039 kmol/s, F "1.1;10\ kmol/s,
#  127
F "4;10\ kmol/s, F "3.2;10\ kmol/s, F "0.47 kmol/s;
 2 &-
catalyst mass"54,892 kg.
the hydrocarbons stream (mainly, steam and ethylben-
zene). The resulting reactant mixture enters the reactor
Optimal profiles obtained for different maximum from the bottom and flows in the centripetal and centri-
allowable temperatures are shown in Fig. 1. The optimal fugal direction in the first and second bed, respectively.
trajectories are rising curves ending by ¹(q)"¹ . Since The global flow configuration is countercurrent. Para-
meters of this reactor are shown in Table 2.
the activation energies verify E (E (E (see Table 1),
  
temperatures must be low at the reactor entrance to limit
the extent of the side reactions (i.e. toluene and benzene
formation). As residence time increases, the temperature 5. Mathematical model
is raised to shift the equilibrium and therefore to increase
the outlet conversion. The values of ethylbenzene conver- A pseudohomogeneous plug-flow model is used to
sion, selectivity and yield are given for each ¹ level in simulate the steady-state operation of the reactor. The

Fig. 1. pressure drop in the catalyst bed is estimated from the
Ergun’s equation (Sundaram et al., 1991). The pressure in
the heating tubes is assumed to be constant. The reactant
4. The non-adiabatic radial-flow reactor mixture and steam are assumed as ideal gases. The gas
density is assumed to be dependent on temperature,
In industrial practice, the adiabatic reactors for styrene pressure and composition. All the other physical proper-
production show decreasing temperature profiles (see ties are considered to be a function of temperature. The
Fig. 2). Despite of the usual heating between stages the simulations are carried out for a constant value of the
shape of the industrial temperature profiles is very differ- reactor outlet pressure, which is in practice determined
ent to that of the optimal temperatures profiles (see by the suction pressure at the off-gas compressor (Sun-
Fig. 1). For this reason, a non-adiabatic reactor (based daram et al., 1991).
on a continuous heating) is proposed to achieve increas- The differential equations expressing the steady-state
ing temperature profiles. mass balances, heat balances for the catalyst bed and the
A schematic drawing showing the main features of the heating tubes, and the momentum balance are:

 
proposed design is presented in Fig. 3. The catalyst is dF  l  l
disposed in the shell side and the heating medium (super- G"$o n(2rH!r N ) GH r # GH r ,
dr @ R R l H l H
heated steam) flows through tubes which are located H H H H
radially in the catalyst bed [see Fig. 3(b) and (c)]. The i"1, 10 (3)
reactor is composed of two radial fixed beds in series. d¹
"
Superheated steam entering from the top [see Fig. 3(a)], dr
flows in the centripetal direction through the heating

tubes; and is collected in an internal tube. Steam leaving $ (*H r )o n(2rH!r N )#2nr N º(¹ !¹)
H H @ R R R R Q
the first bed enters the second one and flows in the H (4)
centrifugal direction to be finally collected in the external 
FC
annular channel. The outlet water vapor is used to dilute G NG
G
208 A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213

culated as described in Table 2. Although the calculated


internal heat transfer coefficients (tube side) are low,
superheated steam has been selected as heating medium
because it is available nowadays in styrene plants to
dilute the hydrocarbon stream and reheat the reactants
between stages (Sundaram et al., 1991).
The system of Eqs. (3)—(6) were solved with the follow-
ing set of boundary conditions:
at r"R (top bed), ¹ "¹
 Q Q
at r"R (top bed), P"P (7)
 
at r"R (bottom bed), F "F for i"1, 10,
 G G
¹"¹ .

The boundary-value problem defined by Eqs. (3)—(7) is
solved using a shooting method. The integration along
the radial coordinate is performed by means of a Gear
algorithm. Parameters used in the calculations are shown
in Tables 1 and 2.

6. Comparison of non-adiabatic and adiabatic reactors

The non-adiabatic radial-flow reactor is compared


with two adiabatic radial reactors in series with reheating
between stages (similar to those found in industry). Geo-
metrical parameters of the adiabatic reactors are in-
cluded in Table 2. To compare the new design with the
adiabatic one, a common basis was chosen. The set of
parameters kept constant for both reactor designs and
the common physical constraints are given in Table 3.
Figs. 4 and 5 show schematic representations of the
styrene plant including the proposed non-adiabatic reac-
tor (plant I) and the current adiabatic catalyst beds (plant
II), respectively. For both plants the total amount of
steam entering the reactor is a sum of the steam added to
the recycle stream and an extra amount of steam. For the
non-adiabatic reactor, this additional steam flows
through the reactor heating tubes before to be mixed
with the reactants stream. In plant II, this steam is used in
the intermediate heat exchanger to rise the temperature
Fig. 3(a) Sketch of the non-adiabatic radial flow reactor. (b) Cross sec- of the reactant mixture, and is reheated before the react-
tion of fixed bed in the direction A—B of Fig. 3(a). (c) Enlargement of ants dilution.
a cut in the direction C—D fo Fig. 3(a) (through the bundle of tubes The performance of the two reactor designs is analyzed
surrounded by catalyst pellets). assuming that the fresh feed of ethylbenzene to the
styrene plant F is specified. This case study is selected
# D
to compare the performance of plants I and II when the
d¹ 2nr N º
Q"$ R R (¹ !¹) (5) flowrate of fresh ethylbenzene is determined by the
dr ¼C Q process located upstream the styrene plant (e.g. the ethyl-
Q NQ
benzene production rate cannot be increased). The
 
dP f ¼ 
"$  . (6) comparison is performed for different conversion levels
dr d o 1000 n(2rH!N r )
N E R R (i.e. different reactor loads), considering the conditions
The symbols $ in Eqs. (3)—(6) are used to adapt the given in Table 3 and the following additional assumptions:
model to centripetal flow (#) or centrifugal flow (!). If
N "0 the Eqs. (3)—(6) represent the model of an E Equal total steam consumption for both designs (at
R
adiabatic unit. The heat transfer coefficients are cal- each conversion level).
A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213 209

Table 2
Geometrical parameters for the non-adiabatic and adiabatic reactors

Non-adiabatic bed Adiabatic bed

» (m) 19.02 19.02


A
r (m) 0.609 0.305

r (m) 1.227 1.067

H (m) 6.838 5.790
N 24181 —
R
d (m), nominal diameter: 0.5 0.0213 —
R
d (m) 0.0158 —
R 
¸ (m) 0.618 —
R

"» /» 0.2185 —
R R 0
A (m) 999.98 —
R
Tubes array Triangular —
Tube pitch (m) '0.02538 —
Clearance between heating tubes (m) '4d "0.01268 —
N
º\"h\#h\
U 
h calculated using correlations reported by De Wasch and Froment (1972)
U
h estimated from the Colburn’s equation (Perry and Chilton, 1982)


Table 3
Basis for comparison between the adiabatic and the proposed design

Parameters kept constant for both designs


Catalyst mass per bed (M "27,446 kg)
A
Recycle composition based on dry basis (y "0.8996; y "0.0972; y "3.51e!4; y "2.81e!3)
0 0 0 0
Outlet pressure (P "101.3 kPa)


 
F
Dilution ratio n"  "11.7

F
G
G
Constraints for the process variables
¹ )1150 K (Sneyder and Subramaniam, 1994)
Q
¹ )800 K (Clough and Ramı́rez, 1976)
D
¹(r))¹ "900 K (Sheppard, 1982; Voge, 1982)

 
*P
5.6( (23 kPa/m (Savoretti, 1995)
*r

E Constant value for the mass flowrate through the Fig. 6 shows the reactor styrene selectivity (p) as
heating tubes (¼ "30657.6 kg/h). a function of ethylbenzene conversion for F
1 # D
E Constant inlet temperature (1150 K) of the steam used "68.75 kmol/h. For both reactor designs, the selectivity
as heating medium in the non-adiabatic reactor (see drops as the conversion increases. This behavior is
Fig. 4). caused by the progressive temperature increase, which is
E Constant dilution ratio at the reactor inlet (n). The necessary to rise the conversion. However, the non-
value n"11.7 is obtained by adjusting, for the adiabatic reactor exhibits higher selectivity values than
F selected and for each conversion level, the steam the adiabatic unit for the analyzed range of x . The
# D #
flowrate added to the recycle stream (see Figs. 4 and 5). selectivity improvements range between 2.87 and 4.14%.
At conditions of total recycle of non-converted reactant,
For the non-adiabatic reactor, the desired conversion reactor selectivity is equivalent to plant yield (p "g ).
is achieved by changing the inlet reactants temperature Since F was kept constant, any increase in selectivity
# D
(¹ ). In the case of the adiabatic reactor, each level of leads to a proportional increase in the production rate of

conversion is obtained by varying the inlet temperatures styrene (F "g F "pF ).
127 # D # D
(at the first and second bed) so that the highest selectivity Fig. 7 shows the temperature profiles inside the two
was reached. For both cases the maximum allowable reactors for the same outlet conversion, i.e. x "0.736
#
temperature in the catalyst bed is not surpassed (points A and B in Fig. 6). The corresponding selectivity
(¹ "900 K). profiles are shown in Fig. 8. Conversely to the adiabatic

210 A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213

Fig. 4. Scheme of a plant for styrene production with a non-adiabatic


reactor (Plant I).

Fig. 7. Temperature profiles for points A (adiabatic) and B


(non-adiabatic) of Fig. 6. F "68.75 kmol/h, x "0.736.
# D #

Fig. 5. Scheme of an industrial plant for styrene production with two


adiabatic beds (Plant II).

Fig. 8. Styrene selectivity profiles for points A and B of Fig. 6


F "68.75 kmol/h, x "0.736.
# D #

bed (Fig. 7). These lower temperatures lead to higher


selectivities at conditions of high concentration of ethyl-
benzene. In spite of the selectivity decrease in the last part
of the reactor, the non-adiabatic design shows a higher
outlet selectivity than the adiabatic beds (Fig. 8).
The selectivity for a value of F 30% higher than that
# D
of Fig. 6 is given in Fig. 9. For this new value of fresh
feed, the non-adiabatic reactor again exhibits higher sel-
ectivities than the adiabatic unit. As it can be seen in Figs.
6 and 9, both designs show different ranges of feasible
conversions. Due to the increase in F , the feasible
Fig. 6. Styrene selectivity (p) as a function of ethylbenzene conversion. # D
x ranges in Fig. 9 are narrower than those of Fig. 6. At
F "68.75 kmol/h.
# D
#
the minimum attainable conversion the maximum pres-
sure drop per unit length is violated at the internal
reactor, the non-adiabatic reactor shows a small temper- diameter of the reactor because of the increase in the
ature drop near the reactor entrance followed by a mon- reactor load. At the maximum allowable conversion, the
otonous temperature increase up to the reactor outlet. temperature reaches its maximum allowable value
The continuous heating along the reaction path allows to (900 K) at a particular radial position (at the reactor inlet
reach the desired conversion level with lower temper- for the adiabatic beds, and at the reactor outlet for the
atures in the first half of the reactor than the adiabatic non-adiabatic design).
A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213 211

Fig. 11. Increased profit by using the non-adiabatic reactor instead of


the adiabatic units.
Fig. 9. Styrene selectivity (p) as a function of ethylbenzene conversion
F "89.36 kmol/h.
# D

because of the catalyst deactivation. In practice, the inlet


temperature is progressively raised to avoid excessive
conversion losses, up to temperature levels for which the
operation must be shut down to replace the catalyst. Due
to the lower inlet temperatures, the new design would
extend the catalyst life.
For analogous operating conditions, the non-adiabatic
reactor leads to higher selectivities than the adiabatic
unit. To show how the increase in selectivity influences
the economy of the styrene plant, the following simplified
function is defined:

u"F C !F C . (8)
127 127 # D #
Eq. (8) represents the difference between the income for
selling styrene and the cost of the main raw material
(ethylbenzene). The function defined by Eq. (8) will be
Fig. 10. Inlet reactants temperature F "68.75 kmol/h.
# D used to compare the performance of the industrial unit
and the non-adiabatic reactor. Assuming that, for a given
x value and the same steam consumption, the recycle
Fig. 10 shows the inlet reactants temperature as a func- #
and separation costs are approximately the same, the
tion of conversion. Since F
# D
and the steam consump- comparison of u for both designs allows one to know the
tion are the same for both designs, the inlet reactants economical benefits introduced by the operation of the
temperature becomes the manipulated variable to new reactor. For a constant value of F and from Eq.
achieve the desired conversion level. For x '0.58, the # D
# (8), the profit for each x can be calculated as:
non-adiabatic reactor allows to achieve a given x with #
#
¹ values up to 77 K lower than those of the adiabatic P "u !u "(F C !F C )
 P ,  127 127 # D # ,
beds. This capability of the non-adiabatic reactor would
allow to minimize the thermal cracking reactions which !(F C !F C )
are favored at high temperatures (Sundaram et al., 1991). 127 127 # D #  (9)
Therefore, if the thermal cracking reactions were in- "C [(F ) !(F ) ]
127 127 , 127 
cluded in the mathematical model, the non-adiabatic
design would predict higher improvements in selectivity "C F [(p) !(p) ].
127 # D , 
than the enhancements showed in Figs. 6 and 9. In
industry, the homogeneous reactions become important Fig. 11 shows the profit given by Eq. (9) as a function of
at the inlet zone of the first adiabatic bed where high x for a C "79.15 $/kmol and F "68.75 kmol/h.
# 127 # D
temperatures and the highest ethylbenzene concentra- For the conditions of Fig. 11 the styrene production
tions occur. This operating problem is aggravated given by the adiabatic unit and the non-adiabatic
212 A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213

reactor ranges between 43,300—51660 and 49,900— H height of fixed bed, m


53,000 ton/year, respectively. For a plant of this capacity, h heat transfer coefficient (tube side),

the non-adiabatic reactor would introduce an increased kW/m K
revenue between 800,000 and 1,550,000 $/year. h wall heat transfer coefficient (bed side),
U
kW/m K
k rate coefficient for reaction j, see Table 2
H
K equilibrium constant, kPa

7. Summary and conclusions ¸ tube length, m
R
N number of heating tubes
R
A non-adiabatic radial-flow reactor design is proposed p partial pressure, kPa
to carry out the dehydrogenation of ethylbenzene to P total pressure, kPa
styrene. Radial flow and continuous heating (using P profit, $/year
P
superheated steam) are the main features of this design. r radial coordinate, m
The radial flow configuration leads to good flow distri- r rate of reaction j, kmol/kg s
H 
bution and low pressure drops. Low pressures favor the r radius of heating tubes, m
R
styrene production, therefore low pressure drops across R gas constant, kJ/kmol K
the reactor are particularly important for this process. ¹ temperature, K
Moreover, the proposed design does not require the heat º overall heat transfer coefficient, kW/m K
exchanger between stages which is needed in the indus- » catalyst volume, m
A
trial adiabatic radial reactors, and consequently avoids » tube volume, m
R
the additional pressure drop in the heat exchanger (see » total reactor volume, m
0
Fig. 5). ¼ mass flow, kg/s
The selected heating medium of the proposed design is x ethylbenzene conversion
#
superheated steam, which is available in industrial y molar fraction
styrene plants, in fact, it is used in large quantities to
dilute the reactants and to reheat them between stages.
The steam consumption for the non-adiabatic reactor is Greek letters
not higher than that used in an adiabatic process. There-
fore, the steam usage would enable an easier revamping *H heat of reaction j, kJ/kmol
H
of existing styrene plants than the use of alternative
tube fraction
R
heating mediums. u economical function
For a given fresh feed, the non-adiabatic reactor leads g ("(F !F )/F ) styrene yield at the
127 127 # 
to operations with higher styrene selectivities and lower reactor outlet
inlet temperatures than the adiabatic design. The in- l stoichiometric coefficient of component i in
GH
crease in selectivity, and consequently, in styrene produc- reaction j
tion levels would improve significantly the economics of o density, kg/m
the process. This increased revenue given by the opera- p ("(F !F )/(F !F )), selectivity
127 127 #  #
tion of the new design justifies to perform an economical q ("V o /¼ ) space time, s kg /kg
A @ E 
analysis including the investment costs. q final space time, s kg /kg
D 
The design of the non-adiabatic reactor has not been
optimized, therefore further studies could be performed
to improve the internal heat transfer coefficient (tube Subscripts
side), and consequently the reactor performance.
A adiabatic
b bed
B benzene
cat catalyst
Notation EB ethylbenzene
ext external
A heat transfer area, m f fresh feed
R
C specific heat, kJ/kmol K g gas
N
d tube diameter, m int internal
R
d particle diameter, m i component i
N
E activation energy of reaction j, kJ/kmol NA non-adiabatic
H
f friction factor ma maximum allowable
F molar flow, kmol/s out at the reactor outlet
A.A. Savoretti et al. /Chemical Engineering Science 54 (1999) 205—213 213

R recycle stream Pontryagin, L.S., Boltyanski, V.G., Gamkrelidze, R.V., & Nishenko,
s steam E.F. (1962). ¹he Mathematical ¹heory of Optimal Processes. New
York: Interscience.
STY styrene Savoretti, A. (1995). Reactores Radiales de ¸echo Fijo: Ana& lisis Com-
T toluene parativo y Propuestas de DisenJ o. Ph.D. Thesis, Departamento de
0 at the reactor inlet Quı́mica e Ingenierı́a Quı́mica, Universidad Nacional del Sur,
Bahı́a Blanca, Argentina.
Sheel, J.G.P., & Crowe, C.M. (1969). Simulation and optimization of an
existing ethylbenzene dehydrogenation reactor. Can. J. Chem.
References Engng, 47, 183—187.
Sheppard, C.M. (1982). Kinetic and Reactor Model for the Ethylbenzene
Clough, D.E., & Ramı́rez, W.F. (1976). Mathematical Modeling and Dehydrogenation Reaction. M.S. thesis, Lehigh University, Be-
Optimization of the Dehydrogenation of Ethylbenzene to Form thleham, PA.
Styrene. A.I.Ch.E. J., 22, 1097—1105. Sheppard, C.M., Maler, E.E., & Caram, H.S. (1986). Ethylbenzene
De Wasch, A.P., & Froment, G.F. (1972). Heat Transfer in Packed dehydrogenation reactor model. Ind. Engng Chem. Process Des.
Beds. Chem. Engng Sci., 27, 567—576. Dev., 25, 207—210.
Eigenberger, G. (1992). Fixed bed reactors. In Elvers, Hawkins, Short, H.C., & Bolton, L. (1985). New styrene process pares production
& Schulz (Eds.), ºllmann’s encyclopedia of industrial chemistry (vol. costs. Chem. Engng, 92(17), 30—31.
B4, pp. 199—238). Weinheim, Germany: VCH. Snyder, J.D., & Subramaniam, B. (1994). A novel reverse flow strategy
James, C.H., & Castor, W.M. (1994). Styrene. In Elvers, Hawkins, for ethylbenzene dehydrogenation in a packed-bed reactor. Chem.
& Schulz (Eds.), ºllmann’s encyclopedia of industrial chemistry (vol. Engng Sci. 49, 5585—5601.
A25, pp. 329—344). Weinheim, Germany: VCH. Sundaram, K.M., Sardina, H., Fernandez-Baujin, J.M., & Hildreth,
Kirk, D. (1970). Optimal Control ¹heory: An Introduction. Englewood J.M.S. (1991). Plant simulation and optimization. Hydrocarbon Pro-
Cliffs, NJ: Prentice-Hall. cessing, 70, 93—97.
Mufti, I. (1970). Computational methods in optimal control problems. Voge, H.H. (1982). Dehydrogenation, In: J.J. McKetta (Ed.), McKetta
¸ecture notes in operations research and mathematical systems (vol. Encyclopedia of Chem. Proc. Des., (vol. 14, pp. 276—290).
27). Berlin: Springer Westerterp, K.R., van Swaaij, W.P.M., & Beenackers, A.A.C.N.
Perry, R., & Chilton, C. (1982). Manual del Ingeniero Quı́mico (5th (1982). Chemical Reactor Design and Operation. New York:
Spanish ed.). New York. Wiley.

You might also like