Вы находитесь на странице: 1из 63

Catalysis Reviews

ISSN: 0161-4940 (Print) 1520-5703 (Online) Journal homepage: http://www.tandfonline.com/loi/lctr20

Reactor Optimization in the Presence of Catalyst


Decay

Frank S. Kovarik & John B. Butt

To cite this article: Frank S. Kovarik & John B. Butt (1982) Reactor Optimization in the Presence
of Catalyst Decay, Catalysis Reviews, 24:4, 441-502, DOI: 10.1080/03602458208079661

To link to this article: http://dx.doi.org/10.1080/03602458208079661

Published online: 24 Feb 2007.

Submit your article to this journal

Article views: 15

View related articles

Citing articles: 17 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lctr20

Download by: [Nanyang Technological University] Date: 10 June 2016, At: 06:59
CATAL. REV.-SCI. E N G . , 24(4), 441-502 (1982)

Reactor Optimization in the Presence


of Catalyst Decay
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

FRANK S. KOVARIK
Ipatief f Laboratory

JOHN B . B U T T
D e p a r t m e n t of Chemical E n g i n e e r i n g

Northwestern University
E v a n s t o n , Illinois 60201

I . INTRODUCTION ... ............................... 442


11. THE EFFECTIVE RATE CONSTANT AND T H E
CONSTANT CONVERSION POLICY ................. 443
111. ON OPTIMAL CONTROL AND THE MAXIMUM
PRINCIPLE METHOD OF ANALYSIS ................ 449
I V . ON THE PERTURBATION THEORY METHOD
OF ANALYSIS ..................................... 463
V. THE ECONOMIC O B J E C T I V E : MAXIMIZING
PROFIT ........................................... 468
V1. REVERSE OPERATION ............................. 478
VII. SOME ADDITIONAL INDUSTRIALLY RELATED
OPTIMIZATION AND SIMULATION STUDIES ........ 48 3
V I I I . CONCLUSION ..................................... 498
REFERENCES ..................................... 500

441

Copyright 0 1983 by Marcel Dekker, Inc.


44 2 KOVARIK AND BUTT

I. INTRODUCTION

Interest over the past decade or so in the more quantitative


aspects of catalyst deactivation has led to a body of work in the
literature concerned with various reactor design loperation opti-
mization problems that include the effects of catalyst mortality.
From a chemical reaction engineering point of view these efforts
provide a rather unique combination of reactor design, chemical
kinetics, and optimal control theory. As was the case for much
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

of the early work on optimization, this literature gives a num-


ber of cases of examples and counterexamples, assumptions that
are sometimes stronger than the strong maximum principle, and
plenty of "academic chemical reaction I" (A -+ B ) . For the hard-
hats among u s , some questions thus may be raised as to the
practical import of much of this in terms of engineering appli-
cation. Nonetheless, the sum of work is now approaching suf-
ficient proportion to make at least a partial accounting of it worth-
while.
A s in all reviews of this sort, we make no claims at universal-
i t y , but hope the reader'? interest will be retained by what is in-
cluded and hislher curiosity piqued by what is left out or only
partially described. Much of the work is, of necessity, mathe-
matical, and the actual chemical/physical processes of deactiva-
tion sometimes assume a secondary role. For those who have not
already done so, a brief tour of a prior review [ 11 , dealing with
more general problems of deactivation, would be worthwhile. A
particular problem in the present effort has been the classifica-
tion of the various categories of topics, since many of the sources
treat somewhat overlapping subjects. A s a result, we have cate-
gorized the contents here primarily on the basis of what we, or
the original authors, consider to be the major thrust of the work.
The origins of research in this area, and in catalyst deactiva-
tion in general as an identifiable topic of interest in reaction en-
gineering for that matter, seem only to go back to the work of
Szdpe and Levenspiel 1 2 , 3 1 . They pointed out that many of the
problems of chemical reaction engineering, considered to be steady-
state problems, a r e , in the face of catalyst deactivation, in fact
unsteady-state problems. In the field of chemical reactor design
and analysis, this leads to a host of well-defined optimization
problems balancing factors such as yield or selectivity against
catalyst activity. This early work considered the temperature
policies required for optimization of the final conversion for a
specified reaction time and final catalyst activity. Both batch
and fixed-bed reactors [ 41 were investigated, in which the de-
activation rate was taken to be independent of the concentra-
tions of any component of the reaction mixture. The results
REACTOR OPTIMIZATION AND CATALYST DECAY 443

indicated that a constant conversion policy was optimal. A con-


siderable amount of work followed concerning various aspects of
the constant conversion policy, and we can begin the review with
a summary of this in the next section.

11. THE EFFECTIVE RATE CONSTANT AND THE


CONSTANT CONVERSION POLICY
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Lee and Crowe [ 51 demonstrated that the temperature policy


of S d p e and Levenspiel [ 2 ] was not valid when the deactivation
rate is considered to be concentration-dependent . ( W e note a t
the out set here that concentration or conversion-independent
deactivation kinetics can be considered a variant of time-on-
stream formulations, popular for correlation of coke formation.
The utility of the results of such an assumption often depends
on how the problem is formulated.) That work, on batch reac-
t o r s , proposed that Jlk(T) = constant, where k ( T ) is the rate
constant and JI is the catalyst activity. From the relationship
Jlk[hn(x)]P = constant ( 1)

with x concentration and the deactivation rate given by


d$/dt = -kd(T)h,(@)h,(x) (2)

and
kakd P

it is obvious that only when h,(x) is constant (p is constant) can


Jlk(T) be constant.
According to S d p e [ 3 ] , t h e optimal T ( t ) policy implied a con-
stant exit conversion for the CSTR and PFR. The assumptions
incorporated in that analysis were a single, irreversible reaction
with separable deactivation kinetics [ 11, the rates given as y =
r ( t , $, x ) and (dJlldt) = f ( T , $) , and constant inlet concentra-
tion.
Subsequently, Crowe [ 61 established a necessary and suffi-
cient condition for the CSTR constant conversion optimal policy
and a proof of the PFR constant conversion policy when tempera-
t u r e control is distributed. The deactivation rate was assumed
to be conversion-dependent in both cases. For the CSTR (single
u n i t ) , constant conversion is optimal i f :

1. A single irreversible reaction is considered


2. Deactivation kinetics are separable
44 4 KOVARIK AND BUTT

3. The exponent, p , on the temperature function must be con-


stant, where :

in which x and xo are fractional outlet and inlet conversions and


u(T) is a temperature function. This contrasts with the batch
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

reactor and the constant effective rate coefficient policy.


For the PFR with catalyst decay conversion-dependent , the
activity will be a function of both position and time. In this
case we have the following set of equations:

ax/az = +q(u)PF(x) ( 6)

where u is a strictly monotonic function of temperature, F(x) and


f(x) express the conversion dependency of the two reaction rates,
and g(+) is the activity dependence of the decay rate. The initial
conditions require that @ ( z , 0) = + i ( Z ) and x(0, t ) = xo. The ob-
jective function to be maximized is given by

with a bounded optimal control u ( z, t ) :


*
U*(U U'

The first integral is a profit term for conversion up to the final


time t f , and the second a profit term for residual activity. For
the optimal distributed control of a PFR described by these equa-
tions, an unconstrained optimal temperature in R ( Za, l) [where
R ( Q , a)defines the rectangle (Za, Q ) ' ( t a , t b ) , Za < 1, Za < Z b ,
ta < tb] implies constant exit conversion for p = 1 and for p # 1
and Za = 0, constant conversion over the entire bed.
Earlier work on optimum temperature policies, including some
of those already cited here, generally stated that for an isother-
mal PFR (n-th order reaction kinetics and m-th order decay, in-
dependent of conversion) the operational optimum was to maintain
the effective rate constant at its initial value, a s stated in Eq. ( 1).
This in turn leads to temperature-time trajectories of the type
REACTOR OPTIMIZATION A N D CATALYST DECAY 445

FALLING X
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

t t t
F

I
t t
F
FIG. 1. Temperature-time policies for deactivation in fixed-bed
reactors : (a) constant conversion, (b) constant conversion-falling
conversion, (c) temperature constraint at end-of -run.

shown in Figs. l(a) and l ( b ) . Levenspiel and Sadana [41 have


pointed out some intuitive difficulties with this result, since the
optimal trajectory becomes a function of the starting temperature
and it is unlikely that a policy starting at the upper constraint,
for example, would give the same global optimum as those illus-
trated in the figure. In fact, for ED > ER, they demonstrate
that the path shown in Fig. l ( c ) , in which the rising tempera-
ture reaches the limiting value T * only at the end of the r u n ,
446 KOVARIK AND BUTT

is optimal. However, Crowe [ 7 ] has pointed out that the analysis


requires a fixed final activity; when final activity and conversion
are not specified, trajectories a s Fig. l ( b ) can be optimal.
The fixed effective rate constant policy has also come under
fire from Sadana [8], who claims it to be optimum only for liquid
systems (concentrations independent of T ) o r zero-order ideal
gas reactions. For an n-th order irreversible reaction in an iso-
thermal reactor with fixed feed FA and catalyst loading W , con-
version i s given by
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

where J, is the activity and k is the r a t e constant of t h e main r e -


action (this result also assumes quasi-steady state conditions).
The objective function is maximum conversion of reactant :

J x ( t ) dt
0

subject to T, 5 T 5 T* (12)

The Hamiltonian along the unconstrained temperature arc will be


constant, and can be derived a s :

n # 1:

n = 1:

H = x + (1 - [
x) In ( 1 - (31
x). - = constant

where RT < < ED in both cases. This should be reasonable in


many systems, although we note that ED can be small in some
instances of rapid irreversible poisoning. In that event, how-
ever, the requirement that ED > ER is likely not to be m e t .
Equations (13) and (14) in t u r n require that x be constant,
which further requires that J,kCOnbe constant. Since for an
ideal gas C o a ( l / T ) , then:

k@/Tn= constant (15)


forms the optimal temperature policy for t h e n - t h order reaction.
REACTOR OPTIMIZATION AND CATALYST DECAY 44 7

The batch reactor problem of Sze'pe and Levenspiel [ 2 ] has


been again revisited, this time by Pommersheim and Chandra
[ 91, with concentration-dependent catalyst decay. For the ki-
netic models :

m n
-r = dC Idt = - k f $ C
A A
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

it is shown that the same calculus of variations approach used in


R e f . 2 can be applied, although the analysis i s rendered more
complex by the appearance of Ck in Eq. ( 1 7 ) . The resulting op-
timal temperature policy is defined by t h e exceedingly awkward
form :
-1

ERk/E rn
D

['
kER/ED + 1 - n

+1 - n)]

The corresponding effective rate constant k t + is defined b y :

where y = 1/T in both equations. The effective rate in this case


does not remain fixed in value when 6 = f ( C A ) . In the event that
decay is concentration-independent, then k = 0 and E q . (18) b e -
comes
448 KOVARIK AND BUTT
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

TIME

TIME

FIG. 2. Effect of the concentration dependence of decay on


conversion, activity, and optimal temperature policy.

which constitutes an extension of Ref. 2 to arbitrary rn. Figure


2 demonstrates the effect of k from C A on ~ activity, conversion,
and temperature policy, where other reaction orders are unity.
It is seen that a s k increases, the temperature policy required
REACTOR OPTIMIZATION AND CATALYST DECAY 449

becomes steeper, eventually reaching a slope unrealistic for prac-


tical reactor operation. Increased k is also coincident with higher
final activities and conversions, Activity is independent of k until
t h e final temperature constraint is attained. An additional finding
of interest for this problem is that optimal conversion levels do not
depend on II (from Ji%) in the deactivation rate equation. The r e -
striction ED > ER applies to all these results.
In this work Pommersheim and Chandra claimed that the four
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

sets of criteria listed in Table 1 all amounted to the same optimiza-


tion problem. Indeed, solutions to the optimization problem are
identical for all four criteria [ 101. However, details of the opti-
mization may differ since for the optimal policy equations to be
valid the appropriate Legendre-Clebesh conditions must be sat -
isfied. These conditions can differ for the four criteria, and
place some restrictions on the general validity of the analytical
solution given in Ref. 9. Also, the policy of maintaining tem-
perature constant at the maximum value, once attained, was de-
termined not to be optimal [ 101, and for cases in which the op-
timal temperature policy exceeds feasible reactor conditions the
overall solutions given in R e f s . 2 and 9 will be invalid. There
i s , however, some disagreement on the details of this argument
[ 11, 121. For unconstrained temperature limits, though, those
solutions are valid so long as all other qualifying conditions are
satisfied. And all this for a batch reactor.

111. ON OPTIMAL CONTROL AND THE MAXIMUM PRINCIPLE


METHOD OF ANALYSIS

Crowe and co-workers published a number of papers on reac-


tor optimization with catalyst decay in the 1970s using a maximum
principle analysis, two of which we have already cited 15, 61. The
topics considered include optimization of both single [ 131 and multi-
bed [ 141 tubular reactors with spatially uniform temperature, dis-
tributed control optimization of a tubular fixed-bed reactor con-
sidering first -order [ 151 and nonlinear [ 161 catalyst deactivation,
optimization of distributed parameter systems with boundary con-
trol [ 171 , further investigation of constant conversion policies
161 , and the effect of micromixing an optimal conversion [ 181.
Some highlights from these investigations are worthwhile pur-
suing.
In Ref. 13 a search was made for the optimal T ( t ) policy that
would maximize the reaction production in a fixed time within a
single uniform temperature tubular reactor. A single irreversible
4 50 KOVARIK AND BUTT

TABLE 1
Mat hematical Formulations ; Catalyst Decay-Single
Reaction, Concentration-Dependent
Deactivation

Functions: CA(t), $ ( t > , y(t)

Constrsints: = CA + r ( C A , J1, y) = 0
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

$2 = 4 +- 6 ( c A , $, Y ) = 0

End conditions and objective:


a
I. to = 0 , q0, CA given
Q

t f , JI, given

6 = [CAI If =objective
0

11. t o = 0 , J1,,, CA given


0

t f ’ cAf given
6 = [-$I [ = objective

111. to = 0, $0, CA given


0

JIf, C given
Af
6 = [ t l [f = objective
0

b
IV.

a
Used in Ref. 2 for concentration independent decay.
bPresented in Ref. 10.
REACTOR OPTIMIZATION AND CATALYST DECAY 451

reaction with separable deactivation kinetics was considered. The


integrated form of the material balance can be written as

with space time z = 0 at the end of the reactor, x = x,(t) at z = 0.


Quasi-steady state is assumed in this balance, and F ( x ) represents
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

the conversion-dependent factor in the rate equation of the main


reaction. The decay rate is separable and independent of conver-
sion :

with + = q0 at t = 0 for all t and 0 5 g ( + ) I 1. It is also assumed


that the proportionality KakP exists between the two rate con-
stants, which are of Arrhenius form ( i . e . , p = E R / E D ) . In the
event that p = 0 , the reaction rate is not a function of tempera-
ture which means the lowest feasible temperature should be
chosen to maximize + ; when p = O D , ED = 0 , k ( t ) is constant
and temperature has no effect.
In general the object is to maximize the total amount of reac-
tion over a fixed reaction time, T , via choice of temperature policy.
Thus

subject to 0 5 k, 5 k 2 k*. Now in the maximum principle format,


one writes:

where +(O) = Jl0 and + ( T I 2 0, and the adjoint variable, p , is

p ( ~ )= 0 if +(T) > 0
452 KOVARIK AND BUTT

The Hamiltonian is:

and the maximization of P is equivalent to


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

with k+(T) the optimal policy and $+, )1+ obtained from Eqs. ( 2 4 )
.
and (25) using k+(T) In result, one of the following three con-
ditions is necessary at t < T:

aH +
a) - (k ) = 0 and - (k') 5 0 if k, < k+ < k*
3k
Ok2

b)
aH + +
- ( k ) 2 0 if k = k *
ak

C) (k+) 5 0 if k + = k,
ak

The sub and super * in the above refer to minimum and maximum
bounding values, respectively. Crowe refers to these in order
as curves S, C*, and C,.
For H along curve S to be locally maximum at any time t :

-
a2H a2x 1
(-1 + p ( 1 + JIKF'))
ax
- < 0
ak2 = 3= ak -

where F' is the derivative of the conversion dependent factor in


Eq. (211, F ( x ) , and is < 0. From Eq. (27) it is seen that for sub-
policy S to be maximal, either

JIKF'L- 1

or

Subpolicy S is part of an optimal policy if 0 5 p 5 Pm.


Any optimal policy is composed of one or more subpolicies
among the group S , C * . or C,. Figure 3A summarizes some op-
timal policies for different initial temperatures and p greater and
less than unity and constant inlet conversion. The parameter
REACTOR OPTIMIZATION AND CATALYST DECAY 453

TABLE 2
Parameter Values for Example Calculation of Fig. 3

g ( $ ) = 1; k* = 0 . 1 s space time/d real time

$,, = 1, xo = 0 . 0 , K * = 2.303, x* = 0.90, t * = 900 K


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

(XS refers to subpolicy S in Fig. 3A)

values employed in this calculation are given in Table 2 . It is ob-


served that lower initial temperature means less time spent on con-
straint and, when initial 'temperature is fixed, total t i m e increases
with increasing p. If total r u n time is fixed, optimal temperature
policies have the initial temperature increasing with p , a s shown
in Fig. 3B. Zero activity halts the operation for p = 1 . 5 before
t h e time constraint is attained.
A sensitivity analysis performed on the objective function dem-
onstrated that P+ (max P) was somewhat sensitive to initial tem-
perature. Thus, perturbating To by 21%resulted in a decrease
in Pf from 5 to 9%. This would suggest that searching for an op-
timal initial temperature in an industrial reactor could indeed lead
to meaningful performance improvements.
A final comparison, given in Fig. 3C, is that of constant con-
version operation to the best isothermal policy. The objective in
this example appears to be more sensitive to the 1%T operturba-
tion mentioned above than to a change from the optimal policy to
t h e best isothermal policy.
In summary, the essential characteristics of these optimal tem-
perature policies appear to be a s follows :

a. The policy usually ends on the upper temperature constraint, T*


b. N o unconstrained policy can be optimal below a limiting conver-
sion which is a function of activity and x,,
C . In an unconstrained subpolicy, t h e conversion depends upon
the initial conversion and the starting point
d . If inlet conversion is time-dependent , discontinuous optimal
policies a r e possible (not shown here)
e . A simple one-dimensional search on inlet temperature can be
used to find the optimal policy; however, an optimal isother-
mal policy may be preferable to most suboptimal temperature
policies with erroneous values in To
4 54 KOVARIK A N D B U T T

p = 1.5, $,= 1.0

-
1
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

'800 xs
1 - .90 .77

7 00
0 5 10 15
Time, daysleec space time

p = 0.5, 9 o = 1.0

7 00
0 10 20
Time, d a y s f s e c space time

F I G . 3A. Optimal temperature policies for various initial tem-


perature for p > and < 1. Conversions for the bottom figure a r e
i n t h e o r d e r 1, 2, 3, 4, 5: x s = .90, .84, .76, .56, .556; Z =
.84, .80, .74, .65, .552 for T o = 900, 875, 850, 825, 800 K . T o
for the top figure a r e 900, 875, and 862 K , the lowest conver-
sion for which a stationary policy exists.
REACTOR OPTIMIZATION AND CATALYST DECAY 4 55

I P = 1.5

p = 1.0
8
b
- 800
p = 0.5
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

7001 I , ,
0 20 40
Time, days/sec

F I G . 3B. Optimal temperature policies for a fixed total time-on-


stream of 4 1 . 7 d as a function of the parameter p .

F I G . 3C. Comparison of isothermal and constant conversion


policies as a function of initial temperature.
4 56 KOVARIK AND BUTT

An extension of the analysis has also been given [ 141 for tubu-
lar reactor sequences. Each reactor is considered to be at uni-
form temperature and activity, with deactivation rate independent
of conversion, and the problem posed is to determine the T(t)
policy for all beds that maximizes the total amount of reaction
( A + B ) over a fixed time period. The reactor contains M beds
at uniform temperature, with ( a$ / at) negligible during one resi-
dence time. Conversion in and out of bed i is
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

xi dx
QiKiei = 1 F O
X
i-1

where

m
c e 1. = i
i=1

The objective is again

Results of the optimization are somewhat similar to those for a


single bed. If inlet conversion is constant and all beds are in
the unconstrained stationary policy, Si, the exit conversion from
each bed is constant. Also, the final temperature of each bed must
be at the upper constraint unless $i(T) = 0 or X M ( T ) = 1. Figure 4
gives an example of a typical three bed policy. Note the increased
slope (dki/dt) of the as yet unconstrained beds when an upstream
bed reaches the constraint.
A numerical example was solved for a two-bed reactor in which
the influence of p was again examined. The calculations were per-
formed by assuming the final activity and then marching backwards
in time. For p < 1 (0.5 in the calculation) the solution was the
same as for a single bed; the two are held at the same tempera-
ture on the same program. For p = 1 it was found that there are
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

0 -r
T1 72. 73'

FIG. 4. Illustrative three-bed policy for p < 1.

many possible policies giving the same final conversion, and for
p 2 1 (1.5 in the calculation) the single bed policy is not optimal.
It was calculated that when p 2 1, it may be advantageous to hold
some beds inactive early in reactor operation, thereby conserving
catalyst activity; also, if p > 1, only one bed should be uncon-
strained at any specific time.
Thus far w e have discussed optimization based on PFR models,
which leads to the logical question of what effect does a deviation
in residence time distribution (RTD) from this ideal limit have on
optimal policy. This is treated b y Lee and Crowe [ l a ] for con-
tinuous reactors with arbitrary RTD (macromixing) and segre-
.
gated flow (minimum micromixing) A basis is well established,
since for a single irreversible reaction maximum conversion for
a fixed T ( t ) policy is attained in a PFR. A general formulation
in terms of the exit age distribution, E(6) do, with 6 = number
of residence times, follows directly from simple reactor theory :
m

x($K(k), xo) = 1 x (6, $ K ( k ) , xo)E(6) d6 (29)


o p
4 58 KOVARIK AND BUTT

where x is now average conversion and xp is conversion for ele-


ments of reaction mixture with exit ages between 8 and 0 + do.
The optimization problem is as in Eq. ( 2 3 ) , where x ( t ) is now ob-
tained from Eq. (29).
The CSTR sequence provides a convenient model for exploring
deviations from limiting PRF RTD's. For an N tank sequence, the
exit age distribution is
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

E ( 8 ) dO= - NN ON-' exp (-NO) dB


N ( N - l)!

where N is the effective mixing parameter. The authors give a


numerical example for the case in which g ( + ) = . I The initial
)
temperature and conversion both increase with N since micro-
mixing is reduced as N increases, hence profit also increases
with N . While it may be dangerous to develop overly general
conclusions from numerical examples, it does appear that mix-
ing effects are not large. For the illustration given, if one
goes from the limit of the CSTR to the PFR, the profit change
was on the order of 20%. Using the plug flow optimal policy
( N = m ) for reactor R N in the sequence ( N < m ) results in
a loss of profit of less than 1%over a range of p both >1 and
<1. One would conclude from these results that a reactor could
be operated nearly optimally without foreknowledge of its RTD
unless the deviations from the PFR limit are large indeed. This,
of course, may be the case if severe flow nonidealities such as
channeling are involved.
Finally, in Ref. 18 a comparison was made between optimal
isothermal and temperature-programmed operation. For a single
CSTR , as for a single PFR, there turns out to be very little dif-
ference between these modes of operation, at least in terms of
optimal average conversion. Accordingly, is it worth manipu-
lating reactor temperature in such complicated fashion in order
to obtain what appear to be rather disappointing increments in
conversion?
Further application of the strong maximum principle has been
given for boundary control optimization of a distributed parameter
system by Gruyaert and Crowe [ l ? ] . While the Pontryagin maxi-
m u m principle is useful in determining optimal control of lumped
parameter systems, a corresponding strong maximum principle
attributed to Degtyarev and Sirazetdinov has been developed for
analysis of distributed parameter systems. For the case where
the control enters through the boundary conditions ("boundary
control"), a weaker form of the necessary conditions showed
that the Hamiltonian must only be stationary with respect to the
REACTOR OPTIMIZATION AND CATALYST DECAY 459

optimal control, unconstrained, rather than having to reach a


maximum with respect to control at the optimum ("distributed
controll'). I t has been stated in the literature, however, that
the strong maximum principle is a necessary condition for opti-
mization with boundary control.
Gruyaert and Crowe examine the validity of this maximum
principle for slow deactivation in a PFR. Optimization is via
inlet temperature control ; for this system the mathematics re-
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

duces to a set of first -order hyperbolic partial differential


equations. Now for the parameters of this system the maximum
principle for boundary control requires that t h e optimal control
be entirely constrained; however, it is shown that a totally con-
strained policy under certain circumstances may not be optimal.
A numerical comparison of the best constrained policy with an
unconstrained policy obtained via Pontryagin in the limiting case
when the characteristics are orthogonal showed the latter to be
superior. Accordingly, the strong maximum principle is incor-
rect for boundary control of this class of systems.
The optimal temperature policy T ( z, t ) distributed in both
space and t i m e was investigated by Therien and Crowe [16] for
a single tubular reactor using assumptions and problem objec-
tives similar to those previously stated [ 131 . Both analytical
expressions for the optimal policies and the character of the
optimal control were obtained. In this study the deactivation
rate equation was

where g($) is a continuous and twice differentiable, nonlinear,


monotonically increasing function 0 5 g ( $ ) 5 1. The adjoint
variables X and LI are defined from

where

X(z, t > = t It at z = 1 for all t E [ O , 11


s f
and

t = tbltf

t =time on stream; tb €10. t f l


b
460 KOVARIK AND BUTT

t = final reaction time


f
t = space time
S

a p / a t = x - F ( ~ ) . K- p - k . g ' ( J I ) (31)

with

~ ( z t, ) = 0 at. t = 1 for all z E [ O , 11


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

F'(x) = d F ( x ) / d x

The Hamiltonian-like function is:

The control policies, similar to R e f . 10, a r e again referred to a s


stationary, S , and constrained, C * and C,, where we have

The extremal control policy, k+(z, t ) , can once again be composed


of one or more of these subpolicies. A lengthy analysis is given
for various combinations of subpolicies and how they a r e affected
by reactor operating parameters, values of the kinetic parameters
m and n , and in particular the value of p . A s is stated, this pa-
rameter is critical to the admissibility of a control subpolicy to a n
extremal control policy.
Numerical calculations were performed for the reaction system
parameters given in Table 3. This set of parameters is such that
from the control analysis various combinations can lead either to
extremal control switching lines o r a stationary policy with uni-
form reactor temperature varying with time. For example, in
Fig. 5(a) where p > 1, extremal control switching lines a r e de-
fined in the z-t plane, separating it into regions that define the
controls k * and k, a s a function of n . As n decreases t h e s u r -
face area upon which k , is defined decreases, and in t h e limit
REACTOR OPTIMIZATION AND CATALYST DECAY 461

TABLE 3
Parameters for Optimal Temperature Policy T( z, t )

F(x) = (1 - XI"; n = 0.5, 1 . 0 , 2.0

g($) m = 0.75, 1.0, 2 . 0

k ( T ) , K(T) = Arrhenius forms; ED/R = 15,000 K


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

% ( t ) = 0 for all t E [ O , 11

t = 1.0 s , tf = 2 5 d

T* = 900 K , T, = 700 K

k, = 20.245 s-l

for n -+ 0 , disappears altogether. However, Fig. 5(b) shows


that for p < 1 a stationary policy is called for and the sensitiv-
ity to n is rather small. In both cases it was found that high
deactivation orders ( m ) are less susceptible to decay than lower
ones under identical conditions of operation, thus higher order
deactivation processes call for higher average optimal operating
temperatures.
The sensitive effect of the parameter p on the stationary policy
is illustrated in Fig. 5(c) for 0 < p < 1, the optimal temperature
required being very strongly increased with increasing p. Cor-
responding calculations for p > 1 indicated only a very small ef-
fect on the extremal control switching lines. Similarly, exit con-
version is very sensitive to changes in p for 0 < p < 1 (a reflec-
tion of temperature sensitivity) , but is essentially constant for
p > 1. Comparison of the extremal policies with best isothermal
policies was also carried out for a range of values of p , n , and
m. Improvements on the order of 5%were noted for p > 1, low
order decay, and high order reaction.
Extension of the analysis to the case where the initial catalyst
activity profile is piecewise continuous (but decay kinetics are
first order) is given by Crowe and Thierien in Ref. 15.
462 KOVARIK AND BUTT

m = 1.0
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

880

860

840

820

-800 10 1.0
t

760 -
1 4
-C
I
7 00
0 1.0
t

F I G . 5. ( a ) Effect of the reaction order n on the extremal con-


trol switching line in the plane of T. (b) Effect of t h e reaction or-
der n on the stationary temperature policy in time, uniform tempera-
t u r e . For 1, 2 , 3 , n = 0.5, 1.0, 2 . 0 in both (a) and ( b ) . ( c ) E f -
fect of the kinetic parameter p on the stationary temperature policy
in time, uniform temperature. For 1, 2 , 3, p = 0 . 1 , 0 . 5 , 0.9.
REACTOR OPTIMIZATION AND CATALYST DECAY 463

IV. ON THE PERTURBATION THEORY METHOD


OF ANALYSIS

Bankoff and co-workers have employed a rather different


mathematical approach, via perturbation and asymptotic meth-
ods, to the optimization of decaying catalyst systems. A very
brief discussion of perturbation analysis may be in order at this
point. When we have a system described by the ODE:
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

dxldt = f ( x , t , a)

where a is a parameter along with other necessary conditions,


and the solution to this system is known, x ( t , k), then the so-
lution to the same ODE with a small perturbation, &, from a, can
be obtained from a regular perturbation solution of

dxldt = f ( x , t , a, + E) =f(x, t, E); E << 1

However, when the following set of ODE'S are derived :

dxldt = f ( x , y , t )

as E -f 0, the equations are reduced to a lower dimensional space


and the solution is obtained via singular perturbation analysis.
I t is clear that the two equations above have t h e general form
that w e have encountered in formulating reactor mass balances
with associated catalyst decay rates.
Dalcorso and BRnkoff [ 19, 201 used a regular perturbation an-
alysis for study of both the fixed-time catalyst decay problem,
subject to control and state inequality constraints, and the more
common problem where the total operation time is determined by
minimum conversion requirements and not preset. It is generally
assumed in the literature that t h e kinetics of t h e main and deac-
tivation reactions are well known, which may be the case for the
former but is almost always not t r u e for the latter. It is s u g -
gested that a more realistic way of determining the optimal con-
trol policy for these systems would be the combination of an op-
timal feedback control with an estimation scheme where important
variables ( e . g . , concentration, temperature) ' I . . .are continuously
measured, the unknown parameters sequentially re-estimated by
a nonlinear filter, and the optimum control determined from the
filter output.. . .I1 For nonlinear systems the control-estimation
problem can become messy indeed; however, some relief is often
464 KOVARIK AND BUTT

available in that for slow decay the estimation procedure need only
be considered occasionally.
Neighborinp optimal paths for deactivating systems, which were
run for a fixed time to maximize total yield in a PFR under quasi-
stationary state conditions, are derived in Ref. 19. General n-th
order deactivation kinetics are included. The algorithm developed
is for optimal feedback control that is based on an update of the
currently estimated system parameters, and the method can be ap-
plied either to deactivation or to yield loss by parasitic side reac-
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

tions in undeactivated systems. In the latter case the parameter


E is taken to be (k,/k,), where k, and k2 are rate constants for
the main and side reactions, respectively.
A s an illustration of the application of singular perturbation
theory, we review Kao and Bankoff 1211, who applied t h e theory
to optimal control problems with fixed time and free endpoint. In
thiscase the reaction system did not involve deactivation kinetics ;
however, the method can be used to decompose any large-scale
system with substantially different time constants. In essence,
the singular perturbation method provides an efficient correction
procedure to the zeroth order solution corresponding to the pseudo-
steady state hypothesis (PSSH) for reactions with highly reactive
intermediates. Consider the scheme

kl k2 k3 k4
A -A*+ B- B*-c

with k 2 , k, >> k,, k,. We seek the temperature profile in a PFR


of fixed length which will maximize the production of B , given
feed composition and flow rate. Although the reactive interme-
diates A* and B* may not be kinetically significant in terms of
the PSSH, it is shown in Ref. 2 1 that the optimal temperature
profile is affected by the fast reactions, even with rates more
than an order of magnitude greater than the slower ones. Fig-
ure 6 shows the difference in the perturbation solution (first o r -
der) and a solution based on PSSH (A +. B * C) for the concen-
tration profile of B under optimal control. In the example shown,
E = kl(To)/k2(To) = 0.01, B 2 = E , / E l , E 3 = E,/E, and K , = k,(To)/
k,(To). The increase shown for the first-order solution is about
0,5%, which the authors deem "small but significant. Another
example presented for E = 0 . 1 also demonstrated about a 0.58 dif-
ference, hence the absolute levels of improvement attained are
rather insensitive to the details of the optimal temperature pro-
file.
An interesting extension of the singular perturbation analysis
[22] employs a scaling parameter to transform the free time prob-
lem characteristic of reactor deactivation into an equivalent fixed
REACTOR OPTIMIZATION AND CATALYST DECAY 465

dO.4

0.5

0.4
O
- -1.0
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

0.3

- -2.0
0.2
0 0.2 0.4 0.6 0.8 1.0

Dimensionless Residence T irne

FIG. 6. Zeroth- and first-order concentration functions for


the case B2 = 0.5, / j 3 = 2.0, and K, = 0 . 5 . CB = C g 0 + C g l t .
Residence time = k,(To)z/V with z bed length.

time problem. This was then applied to control of a CSTR with


rapidly decaying catalyst where the characteristic time for reac -
tion may be only an order of magnitude smaIler than that for de-
activation [ 231 , conditions under which the often overworked QSS
assumption is not valid. In that analysis, reactor operation was
terminated when a lower limit on activity, JI *, was reached, and
the optimization problem was maximization of profit via choice of
the correct T ( t ) . First-order corrections to the zeroth-order
solutions ( Q S S ) can be determined ; these corrections correspond
to zeroth-order inner solutions at both ends of the operating in-
terval plus first order outer solutions. The equations for the
main and deactivation reactions a r e

' 1 (35)
dC /dt = - ( C - CA) - kRJICA
A A0

and

d + / d t ' = -k
DJI
4 66 KOVARIK AND BUTT

with t' real time and 8 CSTR residence time. In nondimensional


form :

with +(O) = 1 , 5 ( 0 ) = 0 , and q ( 1 ) = J1,. In t h e perturbation an-


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

alysis, E = kD(T-.)/kR(To), while E = 0 pertains to the normal


Qss analysis. Other terms are defined as p = ekR(T0), t =
kD(To)t', CA = C A o ( 1 - 5) , where 5 is extent of reaction. The
dimensionless temperature control, u , is

u = exp [1- X
1 + blX

with

b, = RTo/ER

The objective function is

where C = cost per time of operation relative to the product sales


price. According to the theoretical development of R e f . 2 2 , the
Hamiltonian is

with 1-1 the fast adjoint variable and A t h e slow adjoint variable.
The final forms of the state and adjoint equations a r e
REACTOR OPTIMIZATION AND CATALYST DECAY 467

The scaling parameter B is defined from

t=to+Be
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

and the optimization is given by the optimal value of B . The nec-


essary condition for optirnality [ 221 is

where v and r- are Lagrange multipliers associated with the upper


and lower control constraints, respectively. The optimal value of
B is given by either of the forms

1
1 H da = 0 or H(1) = 0 (43)
0

From Eq. (42) one can extract the following optimal control pol-
icies :
if v = O , Q > O

u* if v > O , Q = O

Setting E = 0 in Eqs. ( 4 0 ) - ( 4 3 ) gives variational equations for the


zeroth-order coefficients of the asymptotic expansion of the vari-
ables. Examples given in Ref. 23 indicate that perturbation cor-
rections of the zero-order, quasi-steady state (QSS) solution in-
crease the profit ; conversely, the QSS control when applied to
the reduced equations always yields an overestimated profit.
Once again, however, the gains appear to be modest, on the
order of several percent at most.
Kao [ 241 has recently extended the singular perturbation
method to the problem of CSTR start-up with rapid deactiva-
tion. In general the policies here will be rather different than
those we have typically encountered (cf. Figs. 1 and 2 ) . For
an irreversible reaction suffering from independent or product -
468 KOVARIK AND BUTT

dependent deactivation, Kao proposes. that the reactor should be


started up at maximum allowable reaction temperature and then
switched to the QSS optimal constant conversion policy subse-
quently. The conversion level at which this occurs is estab-
lished by operating cost, product sales price, and reactor resi-
dence time. When operating costs approach sales price the in-
fluence of the modified policy becomes larger; some example
cases demonstrated profit increases of about 10%over the Q S S
policy.
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

V. THE ECONOMIC OBJECTIVE : MAXIMIZING PROFIT

The economics detailed by Kao as mentioned above lead to the


concerns of the present section, since thus far we have been pri-
marily considering formulations that base the optimization on
product yield or conversion with other economic factors gen-
erally not included. One might suspect that if the cost of the
stationary optimal temperature control policy were included in
the foregoing, one would always (or almost always, anyway) opt
for the isothermal optimum. An interesting study including
some of these additional factors is the optimization of a vinyl
chloride reactor reported by Ogunye and Ray [ 251. Here the
feed rate, molar ratio of reactants, cooling jacket temperature,
and inlet temperature and pressure were determined that would
maximize an objective containing product values and catalyst
costs. The algorithm was based on a distributed maximum prin-
ciple that was described separately [ 26, 271.
The work is interesting not only because of the nature of the
objective function, but also because the reaction kinetics are
more complex and the PFR reactor model accounts for volume
contraction, pressure drop, and velocity variation. The vinyl
chloride reaction is a sequence, catalyzed by mercuric chloride
on carbon, with the rate-determining step being the reaction of
the complex HgC12-C2H2.HC1 with HCl:

The resultant rate of reaction is given by

where k is the overall velocity constant, P,total pressure, x con-


version, 6 molar ratio H2/C2H2in feed, and k’ the reciprocal of
REACTOR OPTIMIZATION AND CATALYST DECAY 469

the adsorption constant for HC1 on the carbon support. Deactiva-


tion of the catalyst occurs via sublimation of the HgC1, due to re-
actor hot spots. The kinetics of deactivation are similar to what
we have seen earlier here, and propose

aJ,/at = -k J, = p exp ( - A H E / R T ) +
d
with $(z, 0) = w(z), where z = dimensionless bed length, AHE is
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

the heat of evaporation of HgCl,, and p is a normalized preexpo-


nential factor. The reactor mass balance is

< z 5 1 and 0 5 t 5. 1, and the energy balance is


for 0 -

with T(0, t ) = V,(t), the inlet temperature. Additional notation


in the above is B = AoL with L reactor length and A. the preex-
ponential part of k d , E l activation energy for the main reaction,
Vo initial molar velocity of C 2 H 2 , (-AH) heat of reaction, over-
all heat transfer coefficient, A' heat transfer area per unit vol-
ume of catalyst, Uc coolant temperature, and V, the velocity of
the reacting mixture. Equations (45) and (46) were solved si-
multaneously with an expression for pressure drop based on the
Ergun equation. In the optimization it was assumed that the op-
timal operating time could be determined separately, hence the
objective function is based on maximization of profit :

1
I = J x{x(l, t ) - C , [ 1 - x ( 1 , t ) ] - C,[S - x(1, t ) l -
0
1
C,x(l, t ) - C,x(l, t ) } dt - a J w ( z ) dz (47)
0

The first term gives the value of vinyl chloride product (x is


mass velocity of feed), the second and third separations costs,
the fourth and fifth the cost of reactants, and the second in-
tegral the catalyst cost. Optimization can be conducted with
4 70 KOVARIK AND BUTT

1.0 200

X
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

0 0.2 0.6 1.0


t

2.0 u

N
m 1.6 >
3 u
3
x 1.2

0.8
I I I 1
0 0.2 0.4 0.6 0.8 1.0
L

FIG. 7. ( a ) Optimal policy under coolant temperature, Uc,


control; V 2 = 423 K, V 3 (inlet p r e s s u r e ) = 1.5 atm, x = 1 . 5 k g l
m 2 - s . ( b ) Optimal policy u n d e r Uc, V 2 control; V, and x as in
( a ) . ( c ) Optimal policy under UC, Vz, and V 3 control with V,
upper bound of 2 atm.
REACTOR OPTIMIZATION AND CATALYST DECAY 471

TABLE 4
Values of Constants used for Computations [ 251

a 0.4 c, 0.4 (-AH) 26,000 cal/mol

A’ 640.0 C, 0 . 0 5 AHE 14,000.0 cal/mol

AR 1.05 X lo5 dp 0.00625 m h, 0.0016 kcal/(m2)


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

(S)(OC)
c1 0.2 ER 6,000.0 cal/mol
L 3
c2 0.2 ED 7,700.0 cal/mol
E~ 0.60

a = catalyst cost V 0.002 kg/(m2)(s)


AR = preexponential factor,
main reaction P 5 x lo5

respect to any of the operation parameters such a s coolant tempera-


t u r e , inlet conditions, feed composition and rate, and catalyst dis-
tribution.
Figure 7 demonstrates some different forms of the optimization
problem for parameters a s given in Table 4 , in increasing degree
of control. Constant or near-constant conversion appears to be
a result of any of these optimal policies. An important related
problem has to do with catalyst distribution. Since the mechan-
i s m of catalyst deactivation in this case is by volatilization, the
reduction or elimination of hot spots could increase operating
cycles significantly. A s might be expected, whether nonuni-
form distribution is desired or not is a sensitive function of cat-
alyst cost. For a catalyst cost over the cycle of operation less
than 0.076 (units not given), the optimal policy calls for uniform
loading. If, however, cost is increased to 0 . 4 , the optimum load-
ing profile assumes a roughly sinusoidal shape, with high loading
at the immediate entrance to the bed, thence to a minimum in ac-
tivity (loading) of about 0 . 4 of uniform (nondiluted) activity in
the vicinity of z = 0.1, finally increasing to the uniform loading
value at z = 0.8. It is shown that this procedure effectively
damps the hot spot encountered with pure catalyst, and al-
though the total production is 9% lower, t h e savings in catalyst
cost results in an increase in objective of 35%.
412 KOVARIK AND BUTT

The optimization routine of Refs. 25 and 27 was subsequently


extended to include the cycle time associated with semiregenera-
tive reactor operation [28]. In the normal case, one would have
a bank of parallel reactors, N + M , of which N are operational and
M are being regenerated, as shown in Fig. 8 A , and it is desired
to maintain continuous product flow. The total length of the mul-
tibed process, L , contains a number of discontinuities, a k , sym-
bolizing points of separation between beds. The generalized
forms of the reaction and deactivation equations are
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

for 0 5 z 5 1 and 0 5 t 5 1. Also


+
x.(cxk
1 , t ) = hi[qk, x ( a k , t ) l (49)

for k = 1, 2, . .. , N + 1. The catalyst decay function is

..
with m = 1, 2 , . , p. In the above x is conversion, J, is activity,
uk are controls in the k-th bed ( i . e . , temperature), wk is the ex-
tent of catalyst regeneration, and &c is a feed rate to the k-th bed.
Equation (51) is a "blending equation" that relates activity at the
new time, 6,+, to activity at the end of the last time period,
+(z, Am-), and the freshly regenerated Catalyst, wk(z, 6,').
The objective is to maximize the profit over a cycle of operation,
8 , assuming in this case that the cycle time and regeneration se-
quence are fixed. Hence

1 1
+ J J G I x , y. uk, g a l dz dt (52)
0 0

where G , includes product value, heat credits for exit temperature,


fixed costs, etc. ; G 2 is catalyst andfor regeneration cost, old cat-
alyst salvage value, etc. ; and G is heat exchange costs, etc.
REACTOR OPTIMIZATION AND CATALYST DECAY 473

A computational example was developed for the specific deac-


tivation rate expression :

in which k' is dimensionless temperature in the k-th bed, RTk/


ER , p = ER /ED, and p1 = Ao8, where A, is the deactivation fre-
quency factor. Details of the optimization and computational pro-
cedure are given in Refs. 2 7 and 28. The calculations pertain to
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

an adiabatic, endothermic, first-order irreversible reaction. Equa-


tions (48) and (50) can then be written as

+
aJl/at = - p l exp (-{p[uk+ J l [ x ( z , t ) - x ( a k , t ) l l (54)

in which J1 = (-AHR/C~)(R/ER) and B1 = Ao'L with Aol the pre-


exponential factor of the main reaction. These equations make
use of the relationship between extent of reaction and tempera-
ture in adiabatic reactors :

The reactor sequence illustrated consisted of a train of three re-


actors with an additional "swing" reactor, and operation was con-
ducted such that one of the reactors was being regenerated at
any given time. How does one choose the regeneration sequence?
Ogunye and Ray point out the determination of an optimal re-
generation sequence is a separate and formidable optimization
problem in its own right. Hence they detail a number of simple
rules-of-thumb, one of which says the most inactive reactor is
regenerated at any time and a freshly regenerated reactor is r e -
placed in service and the next in line is regenerated, filling in
with the swing or spare reactor. This type of sequencing is il-
lustrated in Fig. 8B. (Alternative schemes are also described,
but not computed.) The primary results of the computation,
using optimal inlet temperature control, are shown in Figs. 8C
and 8D. From Fig. 8C it is seen that the optimal control main-
tains, as far as possible, constant conversion in the downstream
beds and only when temperature is at the upper constraint (Uk* =
0.085 in the example) in all three beds simultaneously does con-
version decrease. In Fig. 8D the catalyst activity distribution
after each switch in the cycle is demonstrated. Note that each
bed is fully regenerated. This is because a low (or no) penalty
was included in the objective function for the cost of regeneration.
474 KOVARIK AND BUTT

Operating Reactors

‘111 912 ‘113


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Re gene rating Reactors

FIG. 8A. A cyclic reactor operation-regeneration scheme.

0.25

1.00

0.75
t
0.50

0.25

FIG. 8B. An example regeneration sequence.

Increasing the penalty for regeneration does not qualitatively af-


fect the policy shown in Fig. 8 ; however, the beds are then not
completely regenerated in each cycle and the constant conversion
falls to a lower level. Further extension of this general approach
to semiregenerative reactors with recycle has also been reported
[ 291. This turns out to be a rather more difficult problem be-
cause of the complex boundary conditions on the system.
In these studies of Ogunye and Ray on semiregenerative op-
eration, the cycle time was externally specified, as mentioned
REACTOR OPTIMIZATION AND CATALYST DECAY 475
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

0.073

0.9
u2
X
0.8 0.077

0.7

I I I I I
0 .25 .50 .75 1
t
z

FIG. 8C (left). Optimal inlet temperature control for sequence


(0.
FIG. ED (right). Catalyst activities resulting from optimal in-
let temperature control for sequence (i).

above. A one variable search method for determining optimum op-


erating cycles has been reported by Park and Levenspiel 130, 311,
and applied to reactors with three different types of gas-solid con-
tacting, i) mixed f l o w gas-mixed batch solids, ii) plug f l o w gas-
mixed batch solids, and iii) plug f l o w gas-packed batch solids
.
(PFR) In general the relationship between operational and re-
generation cycles can be treated in terms of some characteristic
"relay variable" (or variables) over a production cycle. For ex-
ample, Sz6pe [31 suggested three relay variables, qi, +f and op-
eration time top, while Miertschin and Jackson [ 321 suggest two,
Jli and top. The formulation of a single relay variable adequate
to describe the cyclic operation has large advantages, since one
476 KOVARIK AND BUTT

variable search possesses considerable computational time advan-


tages over multivariable systems.
The economic objective employed [ 301 is profit maximization,
assuming no intercyclic dependency :
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

in which the first term in the numerator is net income during the
operational phase and the second is regeneration cost. These are
expressed on a total time basis per cycle via division by operating
and regeneration times, top and tR , respectively. In optimization
we seek

%
where T(+) is the optimal temperature progression. Details of the
functions Gop, CR , and tR are given in Ref. 30. For the overall
optimization it is required to find both the best interval of cat -
alyst activity (regeneration problem) and the best temperature
programming within that interval (operational problem) .
Consider an isothermal reactor with catalyst of uniform activ-
ity throughout the bed, Q S S , and constant feed during the op-
erational phasR. Let u s assume that there is a temperature policy,
suboptimal to T ( $ ) , over a small interval (Jl0 + 6 / 2 ; Jl0 - 6 / 2 ) for
which we can define an operational index:

where
REACTOR OPTIMIZATION AND CATALYST DECAY 477

The optimal temperature policy is that which maximices the opera-


tional index as the catalyst deactivates. Obtaining P fixes the
optimal temperature policy for a specific interval ef catalyst ac-
tivity. The regeneration problem is to maximize P and determine
the corresponding range of catalyst activity.
A regeneration index can be defined a s
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

where C R ~ ( $0)
, is the cost for regenerating from zero activity
to $, and tRv is the variable regeneration time. The following
condition must be satisfied :

'i ,opt 5
I
.I. J($* Pmax
)d$=C
Rf
+ 8maxt Rf
vf .opt

where

and

The last terms in these expressions reflect the fact that certain
fixed times and costs are associated with the regeneration, in-
dependent of the levels of catalyst acti3ty encompassed.
The overall calculation to determine P max with corresponding
('i,opt and $f,opt, satisfying Eq. (60), is given by the following
procedure :
%
1. Guess 8max = Pguess
% %
2. Determine I ($, P) from Eq. (58b)
5 %
3. Evaluate I ( $ , Pguess) from Eq. (58a)
4. Calculate J($, 8guess) from Eq. (59)
5. Repeat steps (1)-( 4 ) until the optimal regeneration criterion,
Eq. ( 6 0 ) , is met
4 78 KOVARIK AND BUTT

2nd try too low


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

$fopt
0

activity.$
'L
FIG. 9. Graphical search for the maximum profit, Pmax.

An illustration of this procedure is given in Fig. 9. The analysis


can be adjusted to include other control variables such a s feed
rate, composition, o r reactor pressure. Details of application of
the method to the three types of mixing behavior mentioned earlier
are given by Park and Levenspiel in R e f . 31. Both parallel and
series deactivation models are included in the analysis, a s well a s
zero-order and first -order regeneration kinetics. The operating
policies for the mixed flow gas-mixed batch solids model and the
plug flow gas-mixed batch solids model are similar, due to the
similarity of mixing behavior in the solids phase. However, in
the plug flow gas-fixed batch solids case the catalyst activity is
not uniform in the reactor and the method does not directly apply.
It can, however, be modified by considering the mean catalyst ac-
tivity in place'of uniform activity a s an approximation.

VI. REVERSE OPERATION

Various alternative modes of reactor operation have been con-


sidered from time to time a s a means of coping with deactivation.
One of the more intriguing of these is the reverse operation sug-
gested by Earp and Kershenbaum [ 331. In this case a high tem-
perature reaction wne moves countercurrent to the feed mixture
i n a PFR , traveling by conduction or radiation (or both) through
REACTOR OPTIMIZATION AND CATALYST DECAY 479

the solid phase. A naturally occurring phenomenon, it can be in-


duced, in an exothermic reaction, by externally heating the down-
stream end of the reactor. The velocity of propagation of the
zone can be controlled by changes in feed rate and concentra-
tion of reactants. In this transient period of operation the re-
action zone is established near the reactor exit with the remainder
of the bed inactive, containing fresh o r nearly fresh catalyst.
In order to determine under what conditions reverse operation
would be preferable to conventional modes of operation, a tem-
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

perature-time-space optimal policy, considering both concentra-


tion-dependent and independent deactivation kinetics, was in-
vestigated. The goal was to maximize the time-averaged yield:

m
'
S =c I Y ( 1 , t ) do
0

, ) a proportionality constant, Y the extent of reac-


with c ( C A ~ 8,
tion, and ,9 a maximum time. The reactor model employed was of
standard PFR formulation except that expansion effects with tem-
perature were included. Otherwise the state, adjoint, and gra-
dient equations were derived from standard first variational tech -
niques. Eight example cases were investigated, a s shown in
Table 5.
Figure 10 shows the convergence behavior from initial estimate
to final solution; in Fig. 10(a) an isothermal policy in space and
time and in Fig. 10(b) a profile distributed uniformly along the
reactor with time. In some cases there was some dependence of
the final solution with initial estimates; however, overall per-
formance is the same and discrepancies are probably due to lo-
cal maxima. The decay was concentration-independent in both of
these cases; however, in Fig. 10(a), p = ER /ED < 1 and in Fig.
10(b), p > 1. When p < 1, the reaction wne settles across the
entire bed and the general temperature level is lower; however,
for p > 1 the reaction zone only occupies a small segment of the
reactor, with much higher peak temperatures, and propagates
along the bed either cocurrent or countercurrent to the feed di-
rection. This characteristic is shown clearly in Fig. 10(b) in the
lower panel, where propagation is in the direction of feed flow.
Figure 1O(c) gives an example of reverse propagation. In these
figures, t is a dimensionless t i m e of operation parameter, and the
calculations are conducted for the example of CO oxidation on CuO
catalyst.
When the problem formulation consisted of a decay law indepen-
dent of conversion, policies were generated that could have the
480 KOVARIK AND BUTT

TABLE 5
Operating Parameters for Examples of Reverse
Operation [ 331

Initial
Decay equation Kl K2 p temperature Run

3000 32.4 <1 Isothermal ... .. F 1


. .. . . R 1
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Reverse
d+ = -+K, exp
dt
(-') 0.003925 8.0 >1 Isothermal
Reverse
..... F2
. .. .. R 2
12,000 32.4 <1 Isothermal ..... F 3
. . . .. R 3
d~ = -$K1Y2 exp
dt
(-7) 0.0157 8.0 >1
Reverse

Isothermal ..... F4
Reverse ..... R 4
Other parameters :

Frequency factor, main reaction = To = 298 K


2.2 X lo6 m3/kmol-s

Length = 1.0 m Total run time =


5 ~ 1 0 5 s

Inlet concentration = 0.00275 kmol/m3

Velocity = 0 . 1 mls

ER = 3.9 X lo4 kJ/kmol

Main reaction = first order, irreversible

reaction zone propagating either cocurrently o r countercurrently ;


however, introducing conversion-dependence into the deactivation
kinetics resulted in only a countercurrent policy, the t r u e optimal
policy. It was postulated that the main objective in choosing the
optimal temperature policy is to maintain a high value of the ratio
of the rate of the main reaction to the deactivation reaction eval-
uated over the reaction zone. For p < 1, low reaction tempera-
tures will accomplish this and the reaction zone wants to be as
REACTOR OPTIMIZATION AND CATALYST DECAY 481

long as possible in order to get to as low a temperature a s possible


and still maintain conversion. This is reflected in Fig. 10(a) ; it
was further shown that the policy in this example resulted in es-
sentially constant conversion.
It was also observed that for p > 1 and reverse propagation, the
maximum in the temperature wave and the maximum temperature at
the final time are approximately constant, and this does not depend
upon the nature of the deactivation kinetics. This hints that these
problem formulations are naturally bounded. i f we examine this in
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

terms of temperature maximization of F , the ratio of the rate of re-


action to the rate of deactivation, then

where z, and Z r are the boundaries of the reaction zone and only a
small time increment is considered. Evaluation of this expression
from the kinetic models yields the following requirement for a
stationary point (i.e., optimal isothermal policy) :

(ER/RTo) - (ED/RTo)
o= 2 (63)

where (I = (T/To) and To is the feed temperature. From the pa-


ramaters of the examples for p > l, the value of @ is 3.8, and it
was shown for F2 and F, that indeed the temperature maximum in
the reverse propagating wave approached this asymptotic maxi-
mum. Since the main and deactivation reactions occur almost en-
tirely within a small mne near the maximum of the temperature
wave, the assumption of an isothermal reaction zone may be jus-
tifiable.
The second limit mentioned above deals with the end of run
condition. Here the catalyst has no future to contemplate so we
might as well maximize the rate of reaction everywhere:

aR/a+ = 0 (64)

where R = rate of reaction per unit mass of catalyst. For a sta-


tionary point in this case, the rate equation yields

or Q = 7.8. Again for F2 and F1,it is shown that this is indeed the
tendency of the end-of-operation profiles.
482 KOVARIK A N D BUTT

Case F - 1 Case F - 2

Iteration = 0 Iteration = 0
3.2

-
b

w
2.6
1
w
v
2.0 all t v
all
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

1.4

0.8

= 5 = 95
3.2
8a
2.6

2.0
I t = 1.0
0.8
I
6-o
4.0
0.8 I
1.4
t 0.0 I 2.0
0.0 \ 0.4
0.8

= 65 ( F i n a l ) = 390 ( F i n a l )

8.01
~. t = 1.0

2.0

1.4 0.0

L I I

0 .2 .4 . 6 .8
Distance
1 0 1
0 .2
I
.4
1
.6
Distance
I
.8
I
?

The near-isothermal policies ( p < 1) presented should be eco-


nomically practical in an industrial environment. Temperature
control on the cocurrent reaction zone profiles seems a bit r e -
mote of practical accomplishment ; however, for systems that
generate a policy that is amenable to reverse operation, where
very little control effort should be necessary, further study is
warranted.
REACTOR OPTIMIZATION AND CATALYST DECAY 483

Case R - 2

Iteration = 0

61 -C
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

- 0

= 20

-
H
%
0

! A 4
6
c
L t = 1.0 I
2

= 905
8
t = 1.0

0
0 .2 .4 .6 .8 1
Distance
FIG. 10. Evolution of temperature profiles in some cases of re-
verse operation.

VII. SOME ADDITIONAL INDUSTRIALLY RELATED


OPTIMIZATION AND SIMULATION STUDIES

Direct application of some of the papers we have been discuss-


ing to specific industrial problems has been somewhat slim with
4 84 KOVARIK AND BUTT

the exception of the vinyl chloride reactor of Ogunye and Ray


[ 251 and the possible use of some of the methods developed to
the operational-regeneration cycle problem [ 2 9 , 311. A detailed
consideration of the latter for a specific reaction, dehydrogena-
tion of 1-butene to butadiene, has been reported by Dumez and
Froment [ 341. While we are primarily concerned with their r e -
actor simulation and cycle optimization in the context of this re-
view, it should be noted that the paper contains a large amount
of experimental work on kinetics of the dehydrogenation reac-
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

tion, kinetics of coke deposition, and the dynamics of the re-


actor undergoing deactivation. No academic reactions with
power-law kinetics here; in fact kinetics of both main and de-
activation reactions were fit with rather complex kinetics. For
the dehydrogenation rate

__ -
r O = k ~ ° K ~ \ P ~KJ
H ( 1 + K P + K P +KDPD)’
B B H H

where sub H , B , and D refer to hydrogen, butene, and butadiene,


K is the equilibrium constant, and K i are the normal Langmuir-
Hinshelwood adsorption constants. The rate constant K H O obeys
an Arrhenius law with an activation energy of about 29 kcal/mol.
The rate of coke formation was best f i t by

n m
dCc kcBoPB +kcDoPD
r =-- - exp (-aCc)
c dt
(1+ KcH J P H )

in which the exponential function reflects current activity level $.


In accord with the normal separable formulation, multiplication of
Eq. (66) by $ is also used to reflect the influence of deactivation
on the main reaction.
The reactor model is quite complex, a s intraparticle diffusion
was present in the experimentation and hence a two-phase bal-
ance is necessary. For the fluid phase mass balance:

Butene :
REACTOR OPTIMIZATION AND CATALYST DECAY 485

Hydrogen :
\

a (uScH) = - a k D H ( G )
az
RP

Butadiene :
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

where z is reactor length and r is particle radius. Note the quasi-


steady-state assumption here. Heat transfer-solid to fluid-is
given by

= -[%hk(T - Tk) + a.h.(T - Till (69)


az 1 1

where CT = CB + CH + C D
US = linear velocity

ak = external catalyst surface per volume

ai = external surface of inert packing per volume

Tk, T i = temperatures corresponding to k and i

= fluid molar heat capacity


cpf
h k , hi = heat transfer coefficient
Solid phase heat balances are given for catalyst and inert packing
as :

Catalyst :

aTk
'B.kCk at = akhk (T - Tk) + aikhik (Ti - Tk)
486 KOVARIK AND BUTT

Inert :

aT.
p
Bi
C -1 = a.h.(T
i at 1 1
- Ti> - a. h. (Ti - Tk)
ik ik

where aik = inert -catalyst contact area p e r volume

c k , Ci = specific heat of catalyst and inert per weight


Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

= bulk density of catalyst and inert


'Bk' 'Bi
For the intraphase (catalyst) mass balances we have, for the e x -
ample of butene (others are similar),

a2CB,k
ar2
+ 2 aCB,k
r ar
P acB ,k = -pk (r
D~ at D~
+*)
r
J ~ ~ B ~ B
(72)

where I-cB is the rate of coke deposition from butene, E~ is par-


ticle porosity, and r H is the rate of dehydrogenation overall (in-
cluding the activity function $). The values of $ c and ~ rcB a r e
broken out individually from t h e overall coke formation rate of
Eq. (67).
Parameters for the industrial reactor example and t h e optimiza-
tion are given in Table 6. Typical reactor behavior is shown in
Figs. 11A and 11B. The large initial decrease in reactor tempera-
t u r e is the result of active catalyst and the substantial endotherm
associated with the dehydrogenation. This travels down t h e reac-
t o r for some time until the apparent minimum passes out the e n d ;
subsequently the temperature profile assumes a Q S S behavior.
The decreasing coke profile i s due primarily to the higher inlet
temperature of the reactor, and inhibition of coke formation by
hydrogen near the reactor exit.
The system of equations was simplified for the process optimi-
zation study by eliminating t h e consideration of intraparticle dif-
fusion and substituting an effective reaction rate correlation. The
objective was to maximize the yield I through selection of optimal
on-stream time, t f . where the yield is defined a s

J FD d t
0
T = (73)
t + t
f R
REACTOR OPTIMIZATION AND CATALYST DECAY 487

TABLE 6
Characteristics of an Industrial Reactor for
Butene Dehydrogenation [ 341
~ ~~

Length 0.8 m
C r o s s section 1 m2
Catalyst and inert diameter 0.0046 m
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Catalyst bulk density 400 k g of cat / m 3 diluted bed


Inert bulk density 900 k g of inert/m3 diluted bed
Catalyst external surface area 274 m 2 / m 3 diluted bed
Inert surface area 411 m 2 / m 3 diluted bed
Inlet total pressure 0.25 atm
Inlet butene pressure 0 . 2 5 atm
Molar flow rate 15 kmol / m2 h -
Feed temperature 6OOOC
Initial bed temperature 6OOOC

In Eq. (73) t R is regeneration t i m e and FD is outlet butadiene


flow rate. The regeneration time is taken to consist of fixed
purge and evacuation time contributions, t p a_nd t,, plus a con-
tribution proportional to average coke level, C, :

t = t + t +AC
R P V C

The average coke content and yield are presented in Fig. 12 as a


function of the total cycle time, t,, for a value of the proportion-
ality constant A = 0.05 h / %coke. Reactor operating conditions
a r e inlet temperature 6OO0C, inlet pressure 0.030 atm, and inlet
flow rate 2 0 . 4 kmol/m*-h. The maximum shown corresponds to
t, = 1 . 1 h with an average of 3 . 7 %coke content. Often in prac-
tice the dehydrogenation cycle is halted when t h e coke concen-
tration is sufficient to compensate, in regeneration, for the heat
loss during the operational phase. In the present instance, this
t u r n s out to be 1 . 5 to 2 . 0 % , which is clearly suboptimal. See also
Noda et al. [35] for an earlier example of optimization based on
experimental data (dehydrogenation of isopentane) using a dis-
tributed maximum principle.
4 88 KOVARIK AND BUTT

870

*
860
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

!Yi

850

840
1::;
0.075

I I 1 1
830 0 0.2 0.4 0.6 0.8
Distance. m

870

860

M
850

840
9
0.10
0.25
0.20 h

0.15
830
0 0.2 0.4 0.6 0.8
Distance, m
FIG. 11A. Development of temperature profiles in the dehydro-
genation reactor.
REACTOR OPTIMIZATION AND CATALYST DECAY 489

0.03

u
m
h
d
a
4J
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

a
0 0.02

M
Y 0.01
0 0.05
U

0 a 1 I
0 0.2 0.4 0.6 0.8
Distance, m
FIG. 11B. Development of coke profiles in the dehydrogena-
tion reactor.

3.0

Ec,%
2.5

2.0
Cycle T i m e , tC, h

FIG. 12. Optimal cycle time, industrial butene dehydrogena-


tion reactor.
490 KOVARIK AND BUTT

A recent study of series reaction networks ( A -+ B -+ C) has


been reported by Reiff and Kittrell [ 361. In the case that the
activation energy for the two steps are the same they were able
to show that a constant conversion policy was rigorously optimal
for establishing a temperature-time policy maximizing the yield
of B over the total operating time. A maximum principle calcula-
tion was used for a CSTR with concentration-independent ki-
netics and feed consisting only of reactant A . While restric-
tion of the exact calculation to equal activation energies may
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

seem a rather severe l i m i t , the authors claim that for commer-


cial hydrocracking this is essentially the case. The activity-
time relationship obtained is
0

kD
*
and the temperature-time policy. is

where 6 = E l l E ~and k D a i s the initial value of the deactivation


rate constant. The corresponding constant conversion policy
is

with T the space time. Since yield in this system is directly a


function of conversion, the corresponding optimal constant yield
is

with K the ratio of frequency factors, A2 /A,.


An expression was also derived for optimal start-of-run tem-
perature :
REACTOR OPTIMIZATION AND CATALYST DECAY 491

with a = A , / ( A D ) 8 and y = A 2 / ( A ~B) . As can be seen, this is in-


dependent of the total operating time. An example calculation of
these policies is given for a set of parameters representative of
hydrocracking conditions and kinetics.
Perhaps the most extensive study of a reactor optimization
problem associated with catalyst decay is that of Weekman and
co-workers [ 37-42] concerned with catalytic cracking. This work,
appearing in the literature over a number of years, encompassed
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

a wide range of problems including kinetic lumping, deactivation


rates, and reactor modeling. We shall confine our attention to a
specific problem similar to that of Dumez and Froment : the opti-
mization of operation-regeneration cycles in fixed bed operation.
For the kinetic model of the cracking process, it was proposed
that the gas-oil charge is cracked to a gasoline fraction together
with low molecular weight products and coke directly, while the
gasoline fraction itself can also be cracked to low molecular weight
products and coke. Schematically:

C, -
kO
alC2 + a,C,

with C, = gas-oil charge, C 2 = C 5 - 410°F gasoline, and C 3 = C+'s,


d r y gas and coke. Kinetics of the cracking of C, are second o r -
der due to the wide range of molecular weights contained within
that lump, while the cracking of C, is first order. A time-on-
stream correlation was used for coke deposition:
b
C =at (80)
c c

where tc is time-on-stream and a and b are empirically determined


constants. This can further be reduced to a catalyst activity, q ,
as
-at
C
$ = e (81)

where a is some characteristic decay constant dependent upon cat -


alyst and feedstock. For separable deactivation kinetics and reac-
tion orders as specified above, then

(82)
492 KOVARIK AND BUTT

where y1 and y2 are fractions of gas-oil and gasoline, respectively.


Incorporating these into a plug flow reactor model and time-aver-
aging for conversion yields the following expression :
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

In Eq. ( 8 4 1 , A and X are lumped parameters that are conveniently


referred to as extent of reaction and extent of decay, where

x = at
0

In E q s . (85) and (86), e is bed porosity, p is vapor density in


masslvolume, is liquid density, and to is a characteristic time
of catalyst decay (operating cycle length) used to define a di-
mensionless time variable. The importance of E q . (84) is that it
allows characterization of the kinetic lumping, decay rate, and
reactor model in terms of only the two parameters A and A that
can be determined for specific conditions of operation, catalyst,
and feedstock by fit to experimental data.
In processes in which the time scale of deactivation is long,
such as in hydrotreating, the regeneration cycle occupies only
a small portion of total operating time and close control of the
cycle length in regeneration may not be critically important to
the overall operating strategy. However, the rapid decay en-
countered in catalytic cracking implies that operating and r e -
generation cycles may be of the same order of magnitude in
length and a well-specified optimization problem appears, as was
the case discussed earlier for butene dehydrogenation.
A s usual, the first requirement is to define a performance cri-
terion by which the overall operation may be evaluated. A logi-
cal one i s the ratio of total product obtained in actual operation

*Note that here, similar to Dumez and Froment , an optimal op-


erational policy is not sought. The operation here is isothermal
with declining conversion with time-on-stream, and the optimiza -
tion deals only with operation-regeneration cycle time with this
fixed operating policy.
REACTOR OPTIMIZATION AND CATALYST DECAY 493

to that obtained at 100%conversion with no catalyst deactivation.


The corresponding reactor efficiency is
-
E =--NFotox - total actual product
total ideal product
Fott

where N is the number of operation cycles of duration to each, Fo


is mass fed/time, and t t is the total operating cycle length. Note
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

also that X is the time averaged conversion corresponding to the


end of cycle time, to. Total cycle length is

tt = N(t + t )
O R
where t R is regeneration. Combining Eqs. (871, (88) , and the ex
pression for X, Eq. (84) gives
-
E
r
=
a(to + t R )
In [ + l+A
(1 A exp (-ato) 3
where

x = at0
For the case in which regeneration time is independent of the
operating time per cycle (for example, catalyst discharge for ex-
ternal regeneration) , it is apparent from inspection of Eq. (89)
that an optimum value of to will exist. Figure 13(a) presents the
results of parametric calculations* with Eq. (89) for representa-
tive values of KO and a , and a space velocity of 8 V/V/hr, while
Fig. 13(b) presents similar results for fixed regeneration time
and variable space velocity. While the particular values of the
optima shown are dependent upon the parameters used in the cal-
culation, the trends depicted are not. For fixed space velocity,
a s one increases the regeneration time, optimum cycle length in-
creases. For constant regeneration time, as space velocity is
decreased, the value of E r increases and occurs at larger to.

*Analytical differentiation with respect to to to determine the


extremum results in an awkward transcendental form, so it is
easiest in this case to proceed via direct numerical evaluation.
494 KOVARIK AND BUTT

q = 18.8 h - l

tR= 0 h
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

0 .1 .2 .3 .4 .5 .6
Operate Cycle Time , h

FIG. 13. (a) Reactor efficiencies for regeneration time inde-


pendent of operation time. (b) Efficiencies at various space ve-
locities for fixed regeneration time.
REACTOR OPTIMIZATION AND CATALYST D E C A Y 495

However, it should be remembered that for a fixed feed rate a s


space velocity is decreased, the size of the reactor increases;
hence the trend shown in Fig. 13(b) may be a bit deceptive since
it does not reflect the increased capital cost of equipment required
to attain the improvement in E r . It is interesting to note that in
both figures the locus of the maximum in E r does not vary very
much with respect to the abcissa even for relatively large ranges
of the parameter (save the trivial case of tR = 0 in Fig. 13(a).
Variation of the optimum cycle length with respect to the cat-
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

alyst activity and decay parameters is shown in Figs. 14(a) and


14(b). For a fixed decay velocity (Fig. 14a), the operation
time decreases as the catalyst is less active in initial (intrinsic)
operation, a n d , of course, for a catalyst of initial infinite activ-
ity no regeneration i s required. For a fixed intrinsic activity
(Fig. 14b), cycle length decreases a s the decay constant in-
creases. Unlike the other examples h e r e , in this case there is
a pronounced change in cycle length with the variation of c1 a n d ,
a s one might expect, this appears to be the single most sensitive
parameter indicating relative operative and regeneration times.
When regeneration time and operating cycle time are not in-
dependent, a somewhat different analysis may be used. Regen-
eration and operating cycles become interdependent in a process
such as catalytic cracking when the amount of coke deposited on
the catalyst is dependent upon the length of the operating cycle
and the time required for regeneration is in t u r n dependent upon
the amount of coke. We can start with the Voorhies relationship
for coke content:

b
C = at
c o

which is obtained from integration of


-n
dC Idt = k C
C c c
with initial conditions Cc = 0 when to = 0. In the integrated form
of Eq. ( 8 0 ) the constants a r e

and

b = 1
--
l + n (92)

Further, for the kinetics of coke burning we can use the model
m
dCc/dtR = kbCO,Cc ( 93)
496 KOVARIK AND BUTT

0~-
tR = 0.1 h
15 h"
s = 8 V/V/h
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

50

I I I r I
.1 .2 -3 -4 -5 5

El-

*l>OO Operate Cycle T i m e , h .5

F I G . 14. (a) Effect of intrinsic activity on reactor efficiency.


(b) Effect of decay velocity on reactor efficiency.
REACTOR OPTIMIZATION AND CATALYST DECAY 497

where normally m = 1, but can be less than this for high concen-
trations of coke (multilayer deposits). Now for constant oxygen
concentration we pose the question of what values of m and n will
define an optimum operation-regeneration cycle. Solving Eq. (93)
for CO, = constant and C, = atob at tR = o gives

1 - m
n + l
bl-m
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

(to)
- mS1
* (94)
t R - k C (1 - m ) ’
b 0 2

From the reactor efficiency Eq. (89) it can be shown that if E r +. 0


a s to -+ 0, an optimum value for E r will exist. Hence, substituting
tR from Eq. (94) into (89) and evaluating the limit as to -+ 0, we
obtain

For the efficiency to appzdach zero as to -+ 0 , the exponent on to


must be negative, therefore if n > 0 and m 2 0 , an optimum in E r
with to will exist. Thus, for zero- or positive-order coke-burn-
ing kinetics, an optimum operation-regeneration cycle combination
will be found provided the Voorhies parameter b is between zero
and unity. In Fig. 15 is shown just such a relationship, illus-
trated for m = 1, b = 0.5, and a final coke content at the end of
regeneration of 0 . 0 3 wt 8 coke on catalyst. The locus of the maxi-
mum in E r is space velocity-dependent and is moderately sensitive
to that variable. Corresponding regeneration times are not shown
in the figure but can be calculated directly from Eq. (95) using
the optimum value for to.
Finally, Douglas et al. [ 4 3 ] also have pointed out that the in-
terrelationship between the chemical reactor, the time scale of
deactivation, and other pIant components can have important ef-
fects on optimal conversion policies. For example, several com-
mercial catalyst deactivation strategies are summarized in Table 7.
It is clear that each of these stems from a different design basis,
and that the time scale of deactivation is important in determining
strategy. We have perhaps been remiss in not emphasizing more
strongly the extreme variatiolls encountered in practice for the
time scale of deactivation; for an excellent discussion the reader
is referred to the review of Denny and Twigg [ 4 4 ] .
4 98 KOVARIK AND BUTT

.7
kb = 20 h"
.6

.-c

.4
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Er
.3

.2

.I

C
0 .1 .2 .3 .4 -5 ~
i

Operation Cycle Time, h

FIG. 15. Reactor efficiency with regeneration time dependent


upon operation time, first -order burning kinetics. Voorhies co-
efficient = 0 . 5 , final coke = 0.03%.

VIII. CONCLUSION

We have briefly toured some of the interesting work accomplished


over the last decade in this area, rather specialized, of catalyst de-
cay-reactor optimization. If nothing else, one can see that a myriad
of factors are involved in what various workers consider to be im-
portant or not. Hence we have questions of activation energy ratio,
reaction and decay order, on-stream correlations, residence time
distribution, concentration independent decay, the effective rate
coefficient and the constant conversion policy, isothermal vs op-
timal control policy, regeneration cycles, quasi-steady state as-
sumptions, and on, and on, and on. Many apparent disagreements
over optimal policies really are not, but arise from variations in
the nature of the objective function or constraints that are not
always precisely stated.
Almost all of this work has been theoretical and to some extent
runs the risk of drifting away from practical application. One
REACTOR OPTIMIZATION AND CATALYST DECAY 499

TABLE 7
Summary of Operational Strategies for a Reactor Subject
to Catalyst Deactivation

1. Vary reactor temperature with time to maintain a constant con-


version with a constant reactor feed flow rate. A typicaI pol-
icy for large throughput (400 MM pounds per year) and slow
deactivation rates (months to years of catalyst life).
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

2. Vary throughput of the reactor feed while holding the reactor


temperature and conversion constant. A possible policy for
medium deactivation rates (weeks to months catalyst life) and
small to medium throughput (about 25 MM pounds per year)
systems.
3. Allow the conversion to fall while holding the reactor feed flow
rate and reactor temperature constant. Similar applications as
in Item 2 .
4. Maintain the fresh feed rate and reactor temperature constant
and let the recycle flow increase. Similar application a s in
Item 2.
5. Use a combination of reactors in parallel and the policies of
Items 1 or 3 . Usually, with two reactors in parallel, one will
be off-line for catalyst regeneration while the other is opera-
ting. A typical policy for large throughput and medium to
fast deactivation rates (days to months of catalyst life).
6. Continuous catalyst regeneration while maintaining constant
conversion, throughput, and reactor temperature. A typical
policy for large throughput, rapid deactivation systems (hours
to days of catalyst life).

worries that the costs of optimal control are not often included in
the optimization, and that the problems introduced through inter-
actions in a large-scale plant, a s addressed in Table 7 , are not
considered. Whether or not a constant conversion policy is op-
timal, in whatever sense, is secondary to the fact that the por-
tion of a refinery downstream of a hydrocracker, for example,
likes to be at steady state. While justification is not necessary
for using strictly analytical tools to improve the process model
conception, a more penetrating insight into optimization would
seem assisted by way of more experimentation. Certainly the
link between mathematical analysis and empiricism has not been
500 KOVARIK AND BUTT

closed here; the analytical foundation still nee !s a bit of experi-


mental massage. Perhaps the 1980s will prove to be the time when
this experimental link will be strengthened. Very recently Ooshima
and Harano [45] have used the approach of S d p e and Levenspiel
to investigate optimal temperature operation for the hydrolysis of
sucrose in a batch reactor. Romero et al. [ 461 have investigated
fixed-bed optimization for the dehydration of benzyl alcohol on a
CulSiO, catalyst in light of the model developed by Krishnaswamy
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

and Kittrell [47], and Brisk and Barton studied fixed-bed hydro-
genation of ethylene on Ni/Al,O, [48]. In all three cases relatively
good agreement with theory was obtained and these studies appear
to provide a strong and encouraging foundation for the future.

Acknowledgment

Preparation of this review was assisted in part by the support


of the Exxon Education Foundation.

REFERENCES

r11 J . B . Butt, Adv. Chem. Ser., 109, 259 (1972).


r21 S. Sz6pe and 0. Levenspiel, Chem. Eng. Sci., 23, 881
(1968).
31 S. S d p e , PhD Dissertation, Illinois Institute of Technol-
ogy, Chicago, 1966; A . Chou, W. H . Ray and R . A r i s .
Trans. Inst. Chem. Eng., 45, 153 (1967), predates the
publication but not the thesis of Sz6pe.
[ 41 0. Levenspiel and A . Sadana, Chem. Eng. Sci., - 33, 1393
(1978).
[ 51 S. I. Lee and C . M. Crowe, K d . , g ,743 (1970).
[ 61 C . M. Crowe, - Ibid.,-31, 959 (19176).
71 C . M. Crowe, K d . , 34. 1176 (1979).
[81 A. Sadana, Chem. Eng. Commun., 4, 51 (1980).
r91 J. M. Pommersheim and K . Chandra, AIChE J . , 2-1, 1029
(1975).
101 J. M. Pommersheim, L. L . Tavlarides, and S. Mukhavilli,
Ibid., 26, 327 (1980).
r 111 F M . Frowe, E d . , 27, 877 (1981).
r 121 J. M . Pommersheim, L . L . Tavlarides, and S . Mukhavilli,
Ibid., 27,' 877 (1981).
131 T M . Fmwe, Can. J. Chem. Eng., 48, 576 (1970).
141 C . M . Crowe and S. I . Lee, I&., 49, 385 (1971).
[ 151 C . M. Crowe and N . Thierien, Ibid., 52, 822 (1974).
[ 161 - -E, 810 (1974).
N . Thierien and C . M . Crowe, Ibid.,
REACTOR OPTIMIZATION AND CATALYST DECAY 50 1

I171 F. Gruyaert and C . M . Crowe, AIChE J . , 20, 1124 (1974).


[ 181 S. I. Lee and C . M. Crowe, Can. J . ChemTEng., - 48, 192
(1970).
1191 J. P . Dalcorso and S . G . Bankoff, Chem. Eng. J . , 3, 62 (1972).
1201 J . P. Dalcorso and S . G . Bankoff, - Ibid.,
- 3, 74 (1972).
[21] Y . K . Kao and S . G . Bankoff, Chem. Eng. Commun., - 1, 141
(1973).
r 221 Y . K.Kao and S . G . Bankoff, Int. J . Syst. Sci., 1,335 (1974).
Y . K . Kao and S . G . Bankoff, Chem. Eng. Sci., - 30, 1315
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

[ 231
(1975).
I241 Y . K. Kao, Chem. Eng. J . , In Press.
[ 251 A . F. Ogunye and W . H . Ray, Ind. Eng. Chem., Process
D e s . Dev., 9, 619 (1970).
A . F. Ogunye and W . H . Ray, AIChE J . , 17, 43 (1971).
A . F. Ogunye and W . H . Ray, I&., 17, 365 (1971).
A . F. Ogunye and W . H . Ray, Ind. Eng. Chem., Process
Des. Dev., lo, 40 (1971).
A. F. Ogunye and W. H. Ray, K d . , lo, 416 (1971).
J . Y . Park and 0. Levenspiel, - Ibid., 15, 534 (1976).
J. Y. Park and 0. Levenspiel, Ibid., E, 538 (1976).
G. N . Miertschin and R . J a c k s o r c a n . J . Chem. Eng.,
-
48, 702 (1970).
331 R . G . Earp and L . S . Kershenbaum, Chem. Eng. Sci.,
30, 35 (1975).
-
341 F. J . Dumez and G . F. Froment, Ind. Eng. Chem.,
Process Des. Dev., 15, 291 (1976).
351 H. Noda, S. Kanehara, S . Tone, and T . Otake, Chem.
Eng. Sci., 30, 887 (1975).
[ 361 E . K. R e i f f a n d J . R . Kittrell, Chem. Eng. J . , In Press.
371 V. W . Weekman, J r . , Ind. Eng. Chem., Process Des.
Dev., 1, 90 (1968).
[ 381 V . W. Weekman, J r . , Ibid., 8 , 388 (1969).
[ 391 V . W. Weekman, J r . , 3 D . M . Nace, AIChE J., Is, 397
(1970).
[ 401 D. M . Nace, S . E . Voltz, and V. W . Weekman, J r . , K.
Eng. Chem., Process Des. Dev., lo, 530, 538 (1971).
[ 411 S . E . Voltz, D . M . Nace, S . M . Jacob, and V . W . Weekman,
J r . , g., 11, 261 (1972).
[ 421 B . Gross, D . M . Nace, and S . E . Voltz., I s . , 13, 199 (1974).
[ 431 J. M . Douglas, E . K . Reiff, and J. R . KittrellTChem.
Eng. Sai., 35, 322 (1980).
441 P. J. D e n n y a n d M. V . Twigg, in Studies on Surface
Science and Catalysis, Vol. 6, Catalyst Deactivation,
( B . Delmon and G . F. Froment, e d s . ) , Elsevier, Arnster-
dam, 1980, p . 577.
502 KOVARIK AND BUTT

[45] H . Ooshima and Y. Harano, J . Chem. Eng. J p n . ,13, 484


( 1980).
[461 A . Romero, J . Bilbao, and J . R . Gonzalez-Velasco, Chem.
Eng. Sci., 36, 797 (1981).
[47] S. Krishnaswamy and J . R . Kittrell, Ind. Eng. Chem.,
Process Des. Dev., 1 8 , 399 (1979).
[48] M. L. Brisk and G . w. Barton, Trans. Inst. Chem. Eng.,
-
56, 113 (1978).
Downloaded by [Nanyang Technological University] at 06:59 10 June 2016

Вам также может понравиться