Scale-up Criteria for Stirred Tank Reactors
z
J. J. EVANGELISTA, STANLEY KATZ, and REUEL SHINNAR
T h e City College, CUNY, N e w York, N e w Yark
zyxwvutsrqpon
zyxwvutsrqpon
A method is derived for the design of stirred tank reactors for homogeneous reactions. A simple
mixing model proposed previously by Curl (4) is used to compute the effects of finite mixing
time on complex chemical reactions. It is also shown how the parameters of the model can be
obtained by tracer experiments, or estimated theoretically by the assumption of isotropic
turbulence. It i s shown that in many practical cases the assumption of ideal mixing i s a good
approximation. However, the effects of imperfect mixing are more likely to be felt in a large
reactor than in a pilot plant. Some quantitative examples are discussed. Methods are derived
to compute the average outlet concentration for complex systems such as autothermic reactions, polymerization, crystallization, etc.
In a number of recent publications the effect of mixing
on homogenous chemical reaccions is discussed. In this
paper an attempt is made to summarize this work into a
design and scale up procedure for stirred tank reactors.
The method as described is limited to reactants which
completely mix though it can be extended to certain
heterogeneous reaction systems.
Now in any such scale up problem the first question
that one has to answer is what are the dangers involved.
If there is only a simple one path reaction, then the only
question is how much is t,he conversion affected, and this
can be relatively easily handled either by safety factors or
by some of the estimation procedures described in the
following. The more complex and challenging case is the
one in which the product quality itself might be affected
by the scale up, and in this case there is no way of compensating for our lack of precise knowledge by a larger reactor volume. This is true especially in complex reactions,
where side reactions might be favored by local overconcentrations. Other systems very sensitive to local overconcentration are systems involving nucleation, such as
crystallization and certain polymerization processes.
The problem of scaling up a homogeneous reaction is
therefore basically to evaluate the effect of mixing on the
reaction. The first question that we have to ask ourselves
is what changes when we scale up. One property we want
to maintain during scale up is a similarity in the basic
flow regime. It was shown in previous work (8, 9) that
this criterion can be fulfilled by choosing the size of the
pilot plant or bench reactor such that the Reynolds number is large (> 104 or preferably 105). This ensures that
the average velocity distribution is a function of space
coordinates only and fairly independent of Reynolds number, resulting in a similarity in the overall velocity distribution between geometrically similar reactors. The second
criterion commonly used and explained elsewhere (8, 9)
is that the energy input per unit volume should be constant, as this gives a similarity of the turbulent flow regime
in the high wave number range of the turbulent velocity
spectrum. This is important for heterogeneous systems, as
the velocity field that a single particle sees around its
periphery remains constant during scale up. It was pointed
out that the time scale of mixing changes during scale up
at constant energy input per unit volume. To keep this
time constant during scale up of geometrically similar
vessels (with large Reynolds numbers) one would have to
keep the agitator speed constant. For homogeneous reactions the overall mixing time is really the important criterion, and as long as we can keep both the Reynolds
number high and the revolutions per minute constant,
then we could scale up with considerable confidence. This
is however in most cases imwactical. At constant revolutions per minute the Reynolds number increases linearly
with the characteristic length L and the energy input per
unit volume increases proportionally to Lz, If we want to
keep the Reynolds number above 104 in the small vessel
we either have a relatively small scale up ratio or we end
up with an unrealistic energy consumption in the large
vessel. We therefore often have to live with the fact that
during scale up the mixing time increases and try to estimate this effect in a quantitative way. These considerations apply not only for the design of continuous reactors
but also to the design of batch reactors, if one of the reactants is added continuously.
Zweitering ( 1 7 ) has shown that if the reaction is of
nth order and the feed is premixed one can estimate
bounds on the conversion. One bound ( a maximum for
n > 1 and a minimum for n < 1) is the case of maximum
segregation, where all particles are assumed to mix only
with particles having the same age and waiting time, the
other bound is the case of maximum mixedness, which in
the case of a stirred reactor with a Poisson residence time
distribution is the same as t,he ideally mixed reactor. Obviously for a first-order reaction the conversion is independent of mixing and depends only on residence time
distribution. For more complex reactions these limiting
cases do not result in any bounds on conversion, and other
methods to obtain these bounds for more complex systems
have been proposed ( 1 5 ) . While these methods are not
rigorous they still give a fairly good estimate of the
bounds.
Such bounding is especially useful if the bounds are
close together. This indicates that the system is insensitive
to mixing and can be scaled up quite safely. However, in
the cases where it matters most, namely, in complex reactions, nucleation, etc., the bounds are very far apart,
and thus bounding methods based on residence time distribution alone are not very useful. Luckily most agitated
reactors are, in their behavior, very close to ideally stirred
tanks. Mixing times in most reactors are measured in seconds whereas residence times are normally measured in
minutes or hours. Unless we deal with a system sensitive
to mixing or with very large reactors, it is very hard to
measure any deviations from complete mixing [see for
example (IS)].We therefore normally deal with the case
of a reactor which at least in the pilot plant is very close
to ideal mixing and our main problem is to estimate
1. How large is the deviation from ideal mixing likely
to become in the large scale reactor?
2. How sensitive is the process to small deviations from
complete mixing?
A quantitative approach to these two problems is outlined in the following sections.
zyxw
zyxwvu
zyxwv
zyxwvutsrq
Vol. 15, No. 6
THE MODEL
The mixing model for a stirred tank reactor used here
AlChE Journal
Page 843
zyxwvuts
zyxwvu
is the one used by Curl ( 4 ) , Spielman and Levenspiel
(11),Kattan and Adler ( 1 8 ) , and others in this and related contexts. Briefly, it regards the reacting mass as
made up of a large number of equally-sized parcels of
material (particles), which from time to time undergo
independent pair collisions, equalize concentrations, and
then separate. Between collisions, each parcel of fluid
behaves like a little batch reactor. Fresh material is fed
at a constant rate, and the withdrawal takes a representative cut of the contents of the vessel. One speaks of these
particles as though they had a definite material identity.
This is reasonable for dispersed phase systems but here
where we are dealing with homogenous phase systems,
they might perhaps be better regarded for the present
purpose as primitive representations of turbulent eddies.
If we consider, for concreteness, a single reaction where
the concentration c of reagent behaves in batch according
to
x’
+ x” .) (-y’-
y”
2
y
) dx‘dy”d’dy’’
zyxwvutsrqp
zyxwvutsr
zyxwvuts
zyxwvutsr
zyxwvutsr
zyxwvu
zyxwvutsr
then we may describe the contents of the reactor at any
time t by the concentration distribution p ( c , t ) of particles at that time. Technically speaking, p is a probability
density in c, with
p(c, t ) d c giving the proportion of
S,
particles having concentration between a and b at time t.
The distribution p, according to what has been said,
satisfies the integro-differential equation ( 4 )
Calculations of reactor performance according to Equations (2) and (4) for some common one and two reaction
schemes will be found below.
The basic equation in p may be applied as well in the
study of certain polymerization reactions, where we are
concerned, to describe the molecular weight distribution.
In a homogeneous radical polymerization, for example, we
are concerned to describe the concentration of initiator
(catalyst) i ( t ) , of monomer m ( t ) , of growing radicals
dr, + ( r , t ) d r ,
having molecular weight between r and r
and of terminated polymer having molecular weight between T and T + dr, + ( r , t ) d r . Assuming termination by
combination, the batch performance of such a polymerization system may be described following reference 19 by
+
a$(r, t )
ar
+Gm-
at
= d i(t)d(r)
- D +(r, t ) ,f+dr
=-Bi(t)
where po(c, t ) is the concentration distribution for the
feed, l/a the nominal residence time of material in the
tank, and ,f3 a measure of the agglomerative mixing intensity. The mean concentration of reagent in the vessel
and outlet is simply the first moment of p , and this is the
working measure of overall reactor performance for a
given kinetic system. Information about the random fluctuations in a turbulent mixing system is given by the
higher moments.
It should be noted that the concentration variable c in
Equation ( 2 ) may be scaled in any one of a number of
convenient ways, provided only that the basic property
of averaging on agglomeration is preserved. Thus, we may
write (2) with yield, conversion, extent of reaction, and
so on, in place of c. Further, when there is more than one
reaction going on, so that the kinetics are described by
several simultaneous equations in place of Equation (I),
we find natural generalizations of ( 2 ) . Thus, with two independent reactions whose progress variables (say) follow the kinetic scheme
=
- G m ( t ) S#Jdr
Here, the molecular weight is reckoned in monomer
units, v is the number of radicals formed by each molecule
of decomposing initiator, and B, G, D are rate constants
for initiation, propagation, and termination, respectively.
Again, introducing the moments of the size distributions
we find, (19) a set of batch kinetic equations in these
moments and the concentrations d, m:
dxo
= uBi - DxO‘
dt
dxi
= Gmxodt
DX&~
= 2Gmxl - D x g 2
dt
d3c2
-=
dyo
dt
D
2
-=
dy2
D (xoxz
xo2
(3)
we have
dt
Page 844
AlChE Journal
+ x12)
November, 1969
zyx
zyz
zyxwvutsrqp
di
-=-Bi
dt
drn
dt
-=-
INTERPRETATION OF TRACER EXPERIMENTS
zyx
zyxwvuts
Gntx,
These serve as an eight dimensional version of ( 3 ) , and
we have corresponding to them an eight dimensional
version of (4) in the distribution p ( ~~,~~,~~,y~,y~,y2,$rn,t,i,m,t).
From the distribution p , we may recover, say, the weight
average molecular weight of radical together with polymer as
Entirely similar considerations may be applied to the
study of such particulate systems as crystallization reaction.
Now the methods of solution of Equations ( 2 ) and ( 4 )
that are developed below apply regardless of the number
of independent reaction variables, provided only that the
kinetic expressions may be taken as polynomials in these
reaction variables, They may accordingly furnish a useful
guide to how the mixing interacts with strong nonlinearities in the kinetics, as in the crystallization nucleation
rate, or with an independently varied critical feed stream,
as in the initiator feed to polymerization reactors. We do
not however present any results along these lines here,
but reserve these studies for later report.
Solutions to E uation (2) for several simple reactions
have been publis ed ( 4 , 7, 11 ) In the literature (11) a
Monte Carlo method was used and ( 7 ) a direct numerical
method was employed. A method will be described in
this paper which allows one to obtain the moments of
p ( c ) in a simpler way.
The authors want in no way to imply that the above
model represents the real mixing processes in a turbulent
reactor. These are far more complex and defy as yet an
analytical description. The above model has, however,
one important similarity to turbulent mixing. As will be
shown below a concentration disturbance (in the absence
of reaction) as computed from this model shows the same
sort of decay as in isotropic turbulence, the variance (or
the second moment) of the concentration in fluctuations
in both cases decreasing exponentially with time. We
furthermore do not claim that it is possible to predict a
conversion accurately if B/a is not very large. Our claim
that this simplified model is useful in design is based o n
the fact that the behavior of the conversion with respect
to B/a reaches an asymptotic value for high values of
p/a. As long as B/a during scale up stays lar e enough
so that we remain in a region where our resu ts are insensitive to p/a, we have good justification to assume
that a real mixing process will also be insensitive to mixing
in the same range of mixing times. In the design of real
stirred tank reactors we mostly deal with small deviations
from complete mixing. To say that the results are insensitive to the value of B/a provided B/a is large enough
is the same as assuming the vessel is ideally mixed. The
method described above allows one to estimate the mixing
intensity required to approach ideal mixing. Now even if
the vessel is not ideally mixed, and the results depend on
p/a, as long as the deviations from ideal mixing are small,
the above method should estimate these effects fairly accurately. The question that we still have to answer is how
can we relate p/a to an experimentally measurable quantity, or how can we estimate it in the absence of suitable
measurements.
x
We are concerned here to attach am empirical interpretation to the mixing intensity parameter B of Equation
(2) and its higher-dimensional analogues. This interpretation is to be found in the analysis of tracer mixing
experiments, and we accordingly begin by writing ( 2 ) ,
without the kinetic term, in the form
We may note at once that the information we want is
contained in the statistical fluctuations of concentration
about its mean. As far as mean values go, the mixing
system (without reaction) behaves like a completely
mixed vessel with input-output time constant (Y.
Indeed, if we denote the mean concentration by
we find from ( 6 ) that
equation
p
satisfies the ordinary differential
(7)
.
independent of 8, where
is the mean of the feed distribution po. And if we put a
step of tracer into the feed of an initially tracer-free vessel, so that
po(c, t ) = 6 ( c - c o ) ; t > 0
(8)
and
do) = 0
(9)
z
zyxwvutsrq
zyxwvut
zyxwvuts
Y
Vol. 15, No. 6
we find from ( 7 ) that the mean tracer concentration in
the vessel and its outlet is
p(t) =~
~ ( 1 - e ~ “ ~ )
(10)
the familiar stirred tank exponential response.
T.he first measure of the concentration fluctuation is
given by its variance
V2(t) = J { c - P ( t H 2 p ( c , t ) d c
and we find from Equations (6) and ( 7 ) that it satisfies
the differential equation
is the variance of the feed distribution. Again, with a step
of tracer (8) in the feed of a vessel that is initially tracerfree, so that we have, besides, ( 9 ) , the initial condition
AlChE Journal
$(o) = 0
(12)
Page 845
zyxwvutsrqp
zyxwvutsrqpo
zyx
zyxwvut
zyxwvu
zyxwvut
we find by solving (11) and consulting (10) that the
variance of tracer concentration in the vessel and its outlet
is given by
g-at
u"t)
=a
zyx
zyxwvut
- e-Pt
co2 e - a t
I3-a
(13)
With careful monitoring of an output line, the expression
for the variance in (13) may serve as a guide to the
estimation of /3. This variance rises from 0 to a maximum
at
1
8+a
t=In
8-.
2a
-
before decaying again to 0, and, with CY known, the bare
location of this maximum will give some estimate of 8,
even without detailed knowledge of the variance history.
The step input experiments described above furnish
perhaps the most convenient means for an empirical determination of p, although the underlying Equation (6)
can readily be made to yield results appropriate to quite
different experimental situations. Indeed, the most direct
way of visualizing the effect of /3 is via a batch experiment
with no input-output at all. If, in ( 6 ) , we drop the inputoutput terms, we are left with ( 4 )
S ( T - c )
and it is hoped that in the near future more measurements
will be available, enabling one to correlate /3 with agitator
design.
I t may be noted finally that each value of the mixing
intensity /3 corresponds to a definite value of Zweiterings
degree of segregation (17) in the reactor. Indeed, if we
denote the degree of segregation by s, we have simply
dc'dc"-p(c,t)
}
a+l3
with vanishing @ giving complete segregation, and infinite
8, complete mixing. Thus, an estimate- of the degree of
segregation, obtained perhaps by studying a particular
reaction system (not first order), can be translated directly
into an estimate of 8.
(14)
APPLICATION OF TURBULENCE THEORY
The work of Batchelor (1) and Corrsin (3) and others
on the application of the theory of turbulence to the scale
up of stirred tank reactors leads directly to an estimate of
the mixing intensity B in terms of the large-scale properties of the turbulence. Accordingly, we recapitulate the
basic results here, and discuss their bearing on the practical estimation of 8. The analysis is all for a homogenous
isotropic turbulence.
The development begins with the standard partial differential equation for Fick's law diffusion superimposed
on turbulent convection in an isolated system, from
which one argues in the usual way that
dc2
-=-
dt
dt
and that the variance decays according to
Thus, in a pure batch experiment, where a quantity of
tracer is injected initially into some portion of the vessel,
the resulting inhomogeneity in the distribution of tracer
in the vessel decays, as far as variance goes, according to
that is, simply with the time constant Js. According to this
direct interpretation, /3 can be related to the characteristics of the turbulent flow in the vessel, and we shall review below tshese relations and their implication for the
translation of j3 from one scale of operation to another.
In order to get an experimental measurement of /3 we
can therefore perform three types of experiments. One is
to introduce an amount of tracer and measure the decay
of the concentration fluctuations. The second is to introduce a step in tracer concentration and perform the same
measurement. The third is in the case of multiple feeds to
introduce a tracer in one of the feed streams only and
measure the variance of the concentration fluctuations
over the vessel at steady state. (The latter can also readily
be computed.) In each case we need a method which is
able to measure concentration fluctuations on a very small
scale.
Two methods have been recently developed which
should make an important contribution to our understanding of the mixing process in stirred tanks. One developed by Kim and Wilhelm ( 5 ) is based on the scattering of light due to gradients in the refractive index. The
second by Brodkey (2) uses fiber optics and colored
tracers. As yet there have been very few actual studies
2 0 (grad c ) 2
Here D is the molecular diffusivity, c is a concentration
fluctuation, and the overbar denotes an average so that
c2 is simply the variance uz of concentration in a batch
experiment. The mean square gradient in concentration
fluctuation is related back to 02 by introducing a microscale 1, for the turbulent mixing, according to which the
mean square components of the concentration gradient all
have the common value
d ( t ) = a(0)e-Pt
Page 846
a
zyx
and we may see from this equation that the mean tracer
concentration is constant
-dP
=o
s=-
2
so that
and
dc2
-=--
dt
120 -
L2
C2
This represents a simple exponential decay in variance,
and consulting (15), we see that
Information about the turbulent velocity field itself is
however more naturally expressed in terms of the Taylor
(dissipation) microscale l d , and of a corresponding Reynolds number R. The characteristic velocity appearing in
this Reynolds number is the root mean square velocity
fluctuation u. For an isotropic turbulence, this root mean
square fluctuation is the same in each coordinate direction.
AlChE Journal
u5 = uy = uz =
ib
fl
November, 1969
zyxwvutsrqpo
zyxwvutsr
zyxw
zyxwvut
zyxwvut
zyxw
and the appropriate Reynolds number is accordingly defined as
R=-- 1
* *
and, consulting (16)
kdu
(17)
where v is the (molecular) kinematic viscosity of the
working fluid, The two microscales are related (approximately) by the fact that
this relation being valid for R large, but D / v not too large.
We may note, as a guide, that difEusivities in water are
of the order of
to
sq. cm./sec., while the kinematic viscosity v is 10-2 sq. cm./sec. In most practical
liquid reaction problems D / v is less than or equal to
unity. [Should D / v be large the literature (1 ) shows how
the following analysis might be modified.]
Now the Taylor microscale $I may be related to a characteristic linear dimension L of the mixing system as a
whole (the paddle diameter, say, or the vessel diameter
itself) by taking
-l d= - A
L
R
or, applying the definition (17) of the Reynolds number
The first inference commonly drawn from ( 2 2 ) is that
in order to keep the same reactor mixing properties on
scale up, one should maintain similarity, and also keep
~
In geometrically similar vessels,
the ratio E / L constant.
as e is proportional to N3L2, this means keeping the agitator speed N constant during scale up. Also, since c is
the power input per unit mass, keeping e/L2 constant
amounts to making the actual power input proportional
to L5, a demand which often results in uncommonly high
power requirements. More realistically, we may use ( 2 2 )
to give a numerical estimate of B for a specified reactor
size L and power input per unit mass t, and then use the
methods developed below to estimate the effect of this B
on the reaction system in hand. To do this requires of
course a knowledge of the constant A f , but since this is
scale-free, it can be estimated economically in small scale
by means of the tracer experiments discussed above.
The dependence of on L and L shown in ( 2 2 ) might
of course also have been obtained by a straightforward
dimensional argument. But the analysis based on turbulence theory leads also to some idea of the actual numerical magnitudes involved that may serve as a useful guide
to practice in the absence of the tracer experiments suggested above. From work in wind tunnels and in pipe
flows, Corrsin cites the (approximate) values
A
FJ
2 0. ., ~ 1/2
FJ
so that ( 2 2 ) gives
Here A is an empirical parameter, independent of the
scale of the mixing system, but varying from one mixing
configuration to another. One would expect A to depend,
for example, on the physical properties of the working
fluid, on the shape of the stirrer in a mixed vessel, and
on the ratio of stirrer diameter to vessel diameter, but to
have the same value for geometrically similar mixing systems working on the same fluid.
The Taylor microscale l d may also be related to the
power input to the turbulent fluid. Using a conservative
numerical factor, one takes the energy dissipation per unit
mass of fluid to be 10/3 v u2/I?d2, and, with an empirical
efficiency factor 7, one may identify this as the actual
power input t to the mixing system per unit mass of working fluid
10 7pu2
E =
(20)
3 262
1
In a typical commercial application, with L = 100 cm.,
E = 2 x 104 sq.~m./sec.~
( 1 h.p./100 gal. of water),
Equation ( 2 3 ) gives /3
.63 Vsec. It may be noted that
with these same numerical values, ( 2 1 ) gives for the root
mean square velocity fluctuation, u
340 cm./sec., a
plausible figure for the circulation velocity in the vessel
(which in this context we identify with the turbulent
velocity fluctuation).
In agitated vessels both A and 7 should strongly depend
on agitator design and one would need more accurate
studies before one could really predict A and 7 on the
basis of the geometrical design of the vessel.
Tshere are several studies available on the affect of
agitator design on overall mixing time. Van De Vusse ( 1 3 )
measured mixing time by observing the disappearance of
large scale refractive index gradients via a Schlieren
method. While this is not the same as the tracer experiment described in this section, we can still get an estimate
of B, by writing
-
FJ
zyxwvutsrq
zyx
zyxwvutsrq
--
The efficiency 7 measures that fraction of the power input to the system which goes directly to the turbulent
velocity field (and is only later dissipated as heat), as
distinct from that fraction which goes directly to heat.
One would expect 7 also to be independent of the scale
of the mixing system, but varying from one mixing configuration to another.
The basic physical relations in question are now contained in (17), (19), ( 2 0 ) , and these may be solved directly for u, &, 1,.
One finds
u=
0.27A2
(->
1,= (80.0 A2q)
Vol. 15, No. 6
B=- m
7m
where the constant of proportionally m depends on the
sensitivity of the method used to determine the time of
complete mixing ( m should vary from 1 to 5 ) . Choosing
m = 1 gives a real lower bound on /3, and this is exactly
what is needed in our case.
For a baffled completely stirred turbo or paddle mixer,
Van De Vusse recommends a scale up equation derived
by dimensional arguments which is exactly equivalent to
Equation ( 2 2 ) . If one compares values of #? estimated
from Van De Vusse's work to Equation ( 2 3 ) one finds
that the values from ( 2 3 ) are too high. However, one
should remember that Van De Vusse measured disappearance of a dye, and that by choosing B = 1/Tm7 we
made a very conservative estimate of 8. In view of this
zyxw
(&)'I3
1
D3L2
(7
J
AlChE Journal
Page 847
zyxw
zyxwvutsr
zyxwvutsrqpo
zyxwvut
zyxwvutsrqpon
zyxwvuts
zyxwvutsrq
zyxwvuts
zyxwvuts
the experimental results are really in surprisingly good
agreement with our a prwri estimate.
Van De Vusse’s data correlate well with
(3)
1/3
B-0.1
As said before the constant in this correlation strongly
depends on agitator design and should preferably be
either measured or in the absence of more accurate data
estimated from the literature ( 1 3 ) . In case a high value
of 6 is desirable strong emphasis should be given to correct agitator design and correct placement of the feed
pipes. Multiple agitators and multiple feed points can
increase the value of 6 considerably.
CALCULATION OF REACTOR PERFORMANCE
We are concerned here to establish a systematic way of
solving Equation (2) and its higher-dimensional analogues such as Equation (4) under conditions of steady
state reactor operation. These equations seem in general
to be very difhcult to solve. and since our interest commonly centers on high mixing rates, we proceed by expansion in powers of a//?.What results is a succession of
linear integral equations in the successive contributions
to the overall p , and while these don’t seem to be easily
solvable either, they do lead to sets of linear algebraic
equations in the moments of the successive contributions
to p . The procedure closes, that is, it gives at each level
of approximation a self-contained set of equations in a
finite number of leading moments, provided the rate expressions ( 1), (3) and so on are polynomials. The resulting linear equations in the moments can be solved by
standard means. The overall average reactor performance
is found finally by adding up the contributions to the first
moments of p .
We begin with the one dimensional Equation ( 2 ) . We
drop the time dependence, introduce a neutral variable x
to stand for concentration, yield or whatever, and divide
through by the input-output time constant
to find for
the distribution p
If f is a polynomial of degree N , then (26) furnishes
an overall relation among the first N moments of p . In
particular, for a first order reaction with rate constant k,
where x is the concentration of reagent so that f ( x ) =
-kx/a, we find for the mean reagent concentration in
vessel and outlet the standard result
1+-
k
a!
(independent of the mixing intensity f i ) , where p1;o is the
mean reagent concentration in the feed.
In general, however, p (but not po) and its moments
depend on the mixing intensity, and we proceed by expanding (24), (26), (27) in powers of l/X. We remind
ourselves that in almost all practical applications A is huge
as compared to unity, and we therefore expect such an
expansion to converge fairly rapidly.
We write formally
and
m
,
. -
so that
=
pn(j)
Jxn p(*)(x)dx
(30)
Bringing (28) to (24) and equating coefficients in
gives for p ( 0 )
p‘0’
(x)
- JJp‘O’
( x ’ ) p‘”’ (x”)
(Y
for dl)
x‘
+ xn
x
2
+ A S J p ( x ’ ) p ( x ” ) a(?-
x) dx‘dx”
(24)
Here po ( x ) represents the feed distribution, commonly
6 ( x - xo) for a suitable xo, f ( x ) is the rate expression
normalized on a, and
and for the higher p ( j )
zyxwvutsr
is a dimensionless mixing intensity. The first result we
want from (24) is an overall average material balance. If
we denote the moments of p by
fi =
and those of po by
CL~;O=
s
s
- pu-1) (x)
x n p ( x ) dx
,
xn po(x) dx
~=Pl;o-cL
(34)
and in the higher p ( j )
(26)
We note also here that p ( x ) is to be a properly normalized probability density, so that
Page 848
d
-{ f ( x ) p ( + I ) ( x ) } ; i = 2 , 3 . . .. .. (33)
dx
Similarly, bringing (28) and (29) to (26) and equating
coefficients in A gives in p(O)
we find this material balance by multiplying (24) through
by x and integrating
-Jf(x)p(x)
) dx‘dx”
pl(j)-
Jf(x)p(j)(x)dx
= 0; i = 1 , 2 , .
...
(35)
Finally from (27), we have that
AlChE Journal
November, 1969
while
zyxwvutsrqpon
zyxwvu
zyxwvutsrqp
zyxwvuts
zyxwvu
zyxwvutsrq
zy
zyxwvu
po'o' = 1
po(j) = 0;
j
=
(36)
1,2, . . . .
(37)
Now the equation (31) in p(O) simply represents the
behavior of the perfectly mixed reactor (infinite #), and
its solution is
p'O)(x) = 8 [ x - p 1 ' 0 ' ]
(38)
which satisfies the normalization (36). The mean value
p1(O) is determined by the overall material balance (34)
- f [ p l ' O ' l = Pl;O
pl'o'
(39)
With p ( 0 ) in hand, we may attack the equation (32) in
the first correction p") by taking partial moments (30),
that is, by multiplying through by successive powers of x
and integrating. We find, taking due account of (37),
linear algebraic equations in the p n C 1 )of the form
moments can be found at will, this causes no special difficulty. Finally, having evaluated the partial moments for
as many i as desired, we may see explicitly the effect of
the mixing intensity on the overall performance of the
reactor from (29).
p1 = p1'0'
+ 1x
-p1(1)
+ -1
A2
p1'2'
+ . ..
(43)
This completes our solution of the one-dimensional
equation (24). Entirely similar considerations apply to
the higher dimensional versions of ( 2 ) such as ( 4 ) . For
the working equation, we have in place of (24)
.
-
-pn;o
+ n[pl(o)]n-l
-
f [ . ~ 1 ' ~ )n
] ;=
2,3, . . . .
(40)
there being no information in the moment equations for
n = 1. This loss of information is however made up by
the overall material balance (35) for j = 1.
-
s
pl(l)
S
.
zyxw
(41)
and taking it together with the Equations (40) for n =
2 , 3 , . . . ,N, we find a set of N linear algebraic equations
in ~ l ( ~. ).,, p N ( l ) which can be solved by standard numerical means. With these first N moments in hand, as
many further moments as desired can be generated by taking (40) for n = N
1, N+ 2, etc., in turn.
The same procedure applies to the higher p ( j ) of (33).
Taking moments, we find a set of algebraic equations very
much like (40), indeed having the same left hand sides
p
~ 1 ~ 1 x +2 ~ X2 R ~ R ( X)
dx
..
nR,o=
Jx1nlxn2.
. .XnR p o ( x ) c l j ,
There are now R overall material balances in place of
(26)
- Jfl(x)p(x)dx
.
+
. . nR =
pn1nZ
with
pn1nz
f ( x ) p ( l ) ( x ) d x= 0
If f is a polynomial of degree N, Equation (41) is another
independent relation among the first N moments of pcl),
.
where x stands for (XI,
xz, . . , x R ) , dx for the volume
element dxl, dxz, . . . , d X R , and the 6-symbol for the appropriate product of &functions. The moments of p are
now multidimensional, and we denote them
-
= p1o
Jf I t ( x ) p ( x ) d x
= Po0
. . O;O - P I O . . . o
. . .1;o-poo..
1
(45)
and these determine entirely the average behavior of the
system for first order kinetics. The single normalization
condition (27) stands.
Tqheexpansion (28) applied to (44) leads to (31) for
,
A
for p ( l ) , and to
...
i = 2,3,. . . .
n = 2,3,.
(42)
the right hand sides for each i involving moments of the
earlier p ( j ) . Again, there is no information in the moment
equations for n = 1, but i f f is a polynomial of degree N ,
we find for each i a closed set of equations in p l ( j ) , pz(j),
. . , p ~ ( 5by
) taking Equation (42) for n = 2, 3 , . . . , N
together with (35). As before, once the first N moments
of p'i) have been found, as many higher moments as desired can be generated by taking (42) for n = N
I,
N
2 etc. It should be noted that the integral term on
the right hand side of (42) leads to a certain cascading
in the number of earlier moments required as we go to
successive values of i. Thus, with f of degree N, we need:
2N - 1 moments of p ( l ) to find N moments of p ( 2 ) ;
3N - 2 moments of p ( 1 ) to find (2N - 1) moments of
p ( 2 ) to find N moments of pC3); etc. But since higher
.
+
Vol. 15, No. 6
+
for the higher p ( j ) . The moment expansions
m
-
with
(j)
PninZ
. . . nR =
S
...
dx
X ~ ~ ~ X Z ~ ~R Zn ~ p (( X)
j )
(49)
applied to the material balance (45) give R equations in
P'O'
AlChE Journal
p1;o:.
0
- Jfl(x)p'"(x)dx
= p 1 0 . . .o;o
Page 849
zyxwvutsrqponmlk
zyxwvutsrqpo
zyxwvutsrqpo
zyxwvutsrqponm
zyxwv
zyxwvuts
zyxwvuts
zyxwv
zyxw
S f ~ ( ~ ) p ( ~ ) ( x ) d x = ~ ~ ~ . (50)
..i;o
~o:P:.i-
and R equations in each of the higher p(j)
-J
Pld!'.
.o
/LOO(!),
.1 -
=0
f1(x)p"'(x)dx
fR(X)p'')(X)dx = 0
j = 1,2,. . .
(51)
Finally, the moilleiit expansions applied to the normalization ( 2 7 ) give
(0)
... o =
PO0
and
(j)
POO. . . 0 = 0 ;
(52)
1
i= I,%...
(53)
The Equation (31) in p(O) is solved just as for one dimension. We have
(0)
...0).
= S(xl--/LlO
p'O'(X)
*
(0)
*8(xR-P00...1)
(S4)
satisfying the normalization ( 5 2 ) , with the R first moments determined by the R material balances (50).
(0)
PlO
(0)
. . . 0 - f l C P l 0 . . . 0,
(0)
(0)
POO.. . 1 - f R [ b ' + l O . .
. 0,
.
*
.
(0)
*
,poo..
(0)
1
?
POO.
Fig. 1. Second order reactions; 2A
values.
3
- - - - - - - Asymptotic
28.
(j)
Pnin2..
nR-
2nl+
. . . + n ~ - l ml=O
.ll = P l O . . .o;o
. .11 = P O O . .
. l;O
(55)
For the first correction p"), we take moments. We find
from (46)
(1)
Pnln2.
'
n
-~2 n l + . . + n R - l
ml=0
r=l
111
+ n2 + . . . +
j = 2,3,. . .
nR
>1
.
(57)
and for polynomial f r these equations can again be closed
off by applying the overall balance (51) and the normalization condition (53). Tthere is again, as we take successive values of j , a cascading requirement for the number of
moments of the earlier pj, but these can as before be generated in turn once each basic set of equations for the
leading moments has been solved. Finally, having evaIu-
Again, the leading term in the sum on the left hand side
disappears by virtue of the normalization (53), and we
find no information in the equations for n1 n2 . . .
nR = 1. Assuming these f7 to be polynomials, each of
degree L N in each argument, this deficiency is made up
as in the 1 dimensional case by taking the Equations (56)
with nl, n2, , . . , n R each running from 0 to N (but exn2
. . . n~ = 0, l ) , and adjoining to
cluding nl
them the R balance equations (51) for j = 1. What results
is a self-contained set of N R
R - 1 linear algebraic
equations in the leading moments of p"). Even in complex
cases, this is by no means a very large number for standard
numerical methods. Thus, for the polymerization kinetics
( 5 ) , with N = 2 and R = 8, we would have only 23
equations.
The procedure may as before be extended to the higher
p'j) by taking moments in (47).We find
+ +
+
+
+
+
+
Page 850
+ B + 2C. Separate feeds for A
Fig. 2. Bimolecular reactions; A
and B, pO(x, y) = 1/2[6 (x-2)
= 1/2x0 = 1/2y0 = 1. - -
AlChE Journal
6 (y)
+6
( 4 6 (y-2)1,
- - - - - - Asymptotic values.
CO
November, 1969
zy
zyxwv
zy
zyx
not quoted here. In practical design problems X is almost
invariably large as compared to unity, and the results for
small values of A are therefore of little interest in our context.
DISCUSSION AND EXAMPLES
Fig. 3. Adiabatic reactions; x = concentration of product, yield =
x averaged over contents of vessel. - - - - - - - - Asymptotic values.
ated the partial moments for as many j as desired, we may
see explicitly the effect of the mixing intensity on the overall reactor performance by assembling the R first moments
from (48).
1
(0)
P10.1 0 = PlO
PO0
.
l=POo
1 . .
0
+--lo..
h
(1)
.o
+1
(2)
0
= p O .
+-A1 ~ O(1). . . l + - 1P l o
(0)
..l
A2
(2)
+ , ,.
o+
...
(58)
This concludes our development of the solution to the
reactor performance equations for high mixing intensity,
and the series expansions and moment methods developed
here are applied below to the calculation of reactor behavior for some common reaction systems. It might be
noted that similar results can be developed for low mixing
intensity by expanding the basic equations in powers of
f3/a, that is, in powers of the A of ( 2 5 ) rather than 1/k.
What results is a hierarchy of differential rather than integral equations in the successive contributions to p , the
leading term being that corresponding to a completely
segregated reactor. Moment methods are no longer appropriate, and indeed the successive (linear) differential
equations can readily be solved: directly, for the onedimensional case; via characteristics, for higher dimension.
The methods are quite straightforward, and the results are
One can now summarize the preceding discussions into
a design procedure for stirred tank reactors.
The first and most important point is to determine
whether deviations from complete mixing will have a
detrimental effect. Quite often the opposite is true. Thus
for example in a homogeneous premixed second-order reaction a degree of segregation different from zero would
improve conversion. Several authors have therefore pointed
out that assuming complete mixing will lead to an overestimate of the required volume. The authors do not agree
to this as in our experience most agitated vessels are very
close to complete mixing. It is not ractical to control the
mixing so that it should be incomp ete, while maintaining
good agitation heat transfer. If no agitation is necessary
then we should a priori use a plug flow reactor in such a
case. If mixing is necessary and it is further desirable to
minimize the variance of the residence time distribution
then this can be achieved by using a cascade of stirred
tanks or similar devices. Therefore the main problem in
designing or scaling up a stirred tank reactor is to decide
whether small deviations from complete mixing have a
detrimental effect or not.
In Figure 1 the effect of A = 2f3/0r on conversion is
plotted for some typical cases of a second-order reaction.
We note that for values of A larger than 100 the conversion
is already insensitive to variation of A. This is not the case
lor A close to unity, but in stirred tank reactors such low
values are quite uncommon. Consider again a practical example. For a 10 liter vessel with a standard turbine agitator and power input the low estimate of /3 according to
the literature ( 1 3 ) would be 0.05 to 0.1 sec.-l. A value of
X of 5 would mean a residence time of 50 sec. For a 10,000
liter vessel the lowest estimate of f3 under the same condition would be 0.01 to 0.02 and a value of A = 5 would
correspond to 250 sec. residence time.
As we stated already our main problem in the design
and scale up of stirred tank reactors concerns reactions on
which a too low value of X is detrimental. The simplest
such case would be a reaction whose total order is less
than unity. Such reactions are very scarce in liquid
homogenous systems and again the effect of A is small.'
Another fairly simple case is where two reactants enter
the vessel in two separate feed pipes. In Figure 2 the conversion for a simple irreversible second-order reaction with
separate feeds is plotted as a function of X. Again the
effect is small for values of x over 100.
Here however, in very large vessels the effect would
lead to a significant reduction of A and should be evaluated
quantitatively.
A second case with similar features is an autothermic
reaction, in which a cold feed is introduced into a hot reaction mixture. Again in this case incomplete mixing always reduces the conversion. In Figure 3 a numerical
example of the effect of A on one typical autothermic reaction is given. In computing the effect of a on the reaction
we made use of a method introduced by Spalding (lo),
namely the fact that for most exothermic adiabatic reac-
fl
zy
zyxwv
zy
Fig. 4. Parallel reactions; A + B; rate constant kl, 2A + 2C; rate
constant kz. 5 = CB/CO-CA, selectivity.
- - - - - - - Asymptotic
values.
-
Vol. 15,
No.
6
0 Several authors have discussed the case of a zero order reaction. The
assumption of a zero order reaction is however a simplified description
of the limiting reaction velocity of catalytic reaction at high pressures.
These reactions are not zero order at high conversions, where the effect
of A would be noticeable. Furthermore, our method (as well as Zwietering's) does not apply to heterogeneous catalytic reactions.
AlChE Journal
Page 851
zyxwvuts
zyxwvutsrqpo
zyxwvuts
zyxwvutsrqpo
zyxwvutsr
tions the effect of conversion on reaction rate can with
sufficient accuracy be described by
-dx
_ -A
dt
+
(l-x)"(x
E
)
~
(59)
zyxw
zyxwvutsrqp
zyxwvutsrqponmlkj
The constants n and m are determined by a best fit to the
actual reaction rate dependence which normally contains
an exponential term, such as
dx
-=
dt
A( 1- x) n c E I R T
- x)ne-E/RTo(1 + ax)
=A( 1
(60)
Spalding has shown that for most exothermic reactions
the mistake made in writing (59) instead of an Arrhenius
relation is small. Equation (59) has the advantage that it
can be treated by the moment method of the preceding
section.
A low value of leads always to a lower conversion and
the effect is quite pronounced. However for values of a
larger than 100 the effect of /3 again becomes small.
These last cases are examples of a broad class of cases
in which the only effect of a low value of a is a reduced
conversion. We can always compensate for this by a larger
residence time, and in fact we might want to compare the
cost of intense mixing to the cost of an increased residence
time. A quantitative evaluation, similar to the one in the
examples given, will lead to a conservative design procedure as long as we use a conservative estimation procedure for A.
A simple example of the second class of problems in
which a low value of A might change the quality of the
product is given in Figure 4, where the reaction is assumed to consist of two parallel reactions of different order
[for an example see ( 1 4 )1. As the undesired reaction is of
a higher order, the concentration of the reactant should
be as low as possible everywhere. A low value of A will
therefore favor the undesirable side reaction.
If the side product is not separated, and the conversion
is high the reduction of A due to scale up might have a
serious effect. One notes here, that just increasing the
residence time without changing the intensity of agitation
will not be sufficient, as the side reaction will occur in the
concentrated region near the inlet. There is a whole class
of reactions of this type, and what complicates this case
further is the fact that in many cases the exact mechanism
and kinetics of all the possible side reactions is unknown.
As another example of this type one might mention reactions in which the pH is controlled by acid addition and
where high pH in the nonmixed region might cause side
reactions.
If the kinetics of all possible undesired side reactions is
at least approximately known then the above methods can
be used to estimate the minimum value of A necessary to
guarantee the specifications. Sometimes it might turn out
to be impractical to obtain a sufficiently high value of A in
a very large tank and several parallel smaller reactors
might provide simpler and cheaper solution.
One might also want to investigate ways of increasing A
by improving both the design of the agitator and the design of the feed distribution. It is unfortunate that we do
not possess sufficiently accurate data for the effect of the
feed distribution on agitator design. But it is apparent that
using multiple agitators (single or multiple shaft) and
multiple feed injection for each agitator, we can considerably increase A without increasing the energy consumption.
Similar considerations apply to the case of systems with
nucleation, in which the effect of A might be the most pronounced, as even with relatively very low concentration
Page 852
fluctuations all the nucleation might occur in the small
unmixed region. This will be discussed in detail in a future
paper.
One last and somewhat discouraging conclusion that we
can draw from the above discussion is that direct empirical
evaluation of the effect of agitator design on complex reactions is impossible in the pilot plant. If a reaction is sensitive to mixing time then this sensitivity is a strong function of /3 and therefore a reaction might be insensitive to
agitator design in the pilot plant while it is very sensitive
to correct agitator design in the large scale plant. This by
the way is true of other systems sensitive to agitator design (9).
ACKNOWLEDGMENT
The work reported here was supported under AFOSR Grant
No. 921-67. Some of the work reported here is a part of the
research carried out by J. J. Evangelista in partial fulfillment of
the requirements for the Ph.D. at The City University of
New York.
N OTATl ON
A
A
= empirical parameter appearing in Equation (19)
= rate constant in autothermic reaction expression
a, b
B
= bounds on concentration
= rate constant for initiation step in polymerization
CA
= outlet concentration of
CB
= outlet concentration of B
= outlet concentration of C
scheme
cc
A
= concentration
cf, Cf' = dummy variables
co = concentration of A in feed
= molecular diffusivity
D
= rate constant for termination step in polymerizaD
C
tion scheme
f
f,
G
= rate expression
= rate expression, T = 1, 2 , . ,R
= rate constant for propagation step in polymeriza-
..
tion scheme
i( t )
L
= initiator concentration
= reaction rate constant
= characteristic length of the mixing system
m
m
= microscale for turbulent mixing
= constant of proportionality
= exponent in approximate rate expression for auto-
k
&
L,
= Taylor (dissipation) microscale
thermic reaction
m ( t ) = monomer concentration
N
N
n
= agitator speed, rev./min.
= degree of polynomial f
= exponent in approximate rate expression for auto-
n,
= integer in moment definition;
thermic reaction
p
po
pj
r
=
1, 2,
= concentration distribution
...R
s
= concentration distribution for the feed
= jth term in expansion of p
= Reynolds number
= number of independent concentration variables
= batch rate expression
= linear dimension of crystal
= molecular weight
= degree of segregation
s
= batch rate expression
t
=time
= root mean square of velocity fluctuation
= x-component of u
= y-component of u
= z-component of u
R
R
r
r
r
u
u,
uy
uZ
AlChE Journal
November, 1969
x
x
xj
zyxwvutsrqpon
zyxwvut
zyxwvuts
zyxwvutsrq
zyxwvutsr
zyxwvutsrqpo
= Cartesian coordinate
= progress variable
= jth moment of size distribution of growing poly-
mer chains
x’,
X”
y
y
yj
LITERATURE CITED
= dummy variables
=
Cartesian coordinate
= progress variable
= jth moment of size distribution of terminated polymer
y’, y” = dummy variables
z
= Cartesian coordinate
Greek Letters
a
= inverse of nominal residence time
/I
=concentration distribution of growing polymer
chains
= concentration distribution of terminated polymer
1. Batchelor, G . K., J. Fluid Mech., 5, 113 (1959).
2. Brodkey, R. S., and J. 0. Nye, Reo. Sci. lnstr., 34, 1086
(1963).
Lee, AIChE I., 10,187 (1964).
3. Corrsin: g:%d., 3, 329 (1957); 10,870 (1964).
4. Curl, R. L., AlChE I., 9, 175 ( 1963).
5. Kim, Y. G., and R. H. Wilhelm, Preprint 26B, Fifty-sixth
Annual Meeting of the AIChE, Houston, Texas.
6. Rietema, K., Aduun. Chem. Eng., 5, (1964).
7. Shain, S., AlChE J., 12,806 (1966).
8. Shinnar, R., J. Fluid Mech., 10, Pt. 2, 259 (1961).
, and J. M. Church, Ind. Eng. Chem., 52, 253
9.
( 1960); 53,479 (1961).
10. Spalding, D. B., Combustion Flume, 1,287,296 (1957).
11. Spielman, L. A,, and 0. Levenspiel, Chem. Eng. Sci., 20,
247 1965).
12. Tadmor, Z:, and A. Biesenberger, lnd. Eng. Chem. Fundamentals, 5, 336 (1966).
13. Van de Vusse, J. G., Chem. Eng. Sci., 4, 178,209 (1955).
14. lbid., 17,507,1962.
15. Weinstein, H., and R. J. Adler, Chem. Eng. Sci., 22,
65 (1967).
16. Worrell, G . R., and L. C. Eagleton, Can. I. Chem. Eng.,
43,254 (Dec. 1964).
17. Zwietering, Th. N., Chem. Eng. Sci., 11, No. 1, 1 (1959).
18. Kattan, A., and R. J. Adler, AlChE J., 13,580 ( 1967).
19. Saidel, G. M., and S . Katz, ibid., 319 (1967).
zyxwvu
zyxwv
= measure of agglomerative mixing intensity
= actual power input to mixing system per unit mass
of fluid
7
= empirical efficiency parameter
A
= mixing intensity parameter, = 2/3/a
p
= mean concentration
= mean feed concentration
~ c n = nth moment of p
f i ; o = nth moment of PO
r-nf = nth moment of pj
v
= kinematic viscosity of workin fluid
= number of radicals formed y each molecule of
v
decomposing initiator
= variance of concentration distribution
$
u.02 = variance of feed distribution
T
= mean residence time
T~
= mixingtime
c
P;
zyx
Manuscript receitied October 11, 1967; retiision receitied June 26,
1968; paper accepted July 18, 1968.
The Dynamic Modeling, Stability, and Control
of a Continuous Stirred Tank Chemical Reactor
W. FRED RAMIREZ and BRIAN A. TURNER
University of Colorado, Boulder, Colorado
A realistic CSTR model was developed and verified experimentally. The reaction studied was
the exothermic, base sodium
model was used to evaluate
analysis, local linearizotion
The effect of control valve
hydroxide catalyzed decomposition of hydrogen peroxide. The
the usefulness of the following stability analyses: steady state
and Liapunov’s direct method through Krasovskii’s theorem.
hysteresis on the system was olso investigated.
The purpose of this investigation was to study both experhentally and theoretically the dynamics and control
of a cantinuous stirred tank reactor, a C.S.T.R. There exist
several analytical methods for determining stability in a
C.S.T.R. (1 to 4, 7). These include steady analysis (71,
hearization techniques (31, and Krasovskii’s Theorem
with the Liapunov function it suggests ( 2 ) .
Brian A. Turner is now with the Shell Oil Company, Martinez, Califomis.
Vol. 15, No. 6
D Y N A M I C EQUATloNs
The system considered was a continuous stirred
reactor. The reaction was first order and homogeneously
catalyzed. The reactant, hydrogen peroxide, and the cats$% sodium hydroxide, were fed into the reactor, while
the product, water, and unreacted peroxide were continuously removed. The solution density and heat capacity
varied with concentration, but their product is nearly constant and was considered constant in formulating the dynamic model. The exothermic heat of reaction was continuously removed through a cooling water belt which
AlChE Journal
Page 853