Chemical Reactor Design Mathematical Modeling and Applications
Chemical Reactor Design Mathematical Modeling and Applications
Juan A. Conesa
Author All books published by Wiley-VCH
are carefully produced. Nevertheless,
Prof. Juan A. Conesa authors, editors, and publisher do not
Universidad de Alicante warrant the information contained in
Ingenieria Quimica these books, including this book, to
Carr. San Vicente del Raspeig be free of errors. Readers are advised
s/n to keep in mind that statements, data,
03690 Alicante illustrations, procedural details or other
Spain items may inadvertently be inaccurate.
10 9 8 7 6 5 4 3 2 1
In loving memory of my parents, Paco y Rosita.
vii
Contents
Preface xiii
Nomenclature xv
1 Nonideal Flow 3
1.1 Introduction 3
1.2 Residence Time Distribution (RTD) Function 3
1.2.1 Measurement of the RTD 4
1.2.1.1 Pulse Input 4
1.2.1.2 Step Input 6
1.2.2 RTD Concept in Heterogeneous Systems 7
1.2.3 Characteristics of RTD 8
1.2.3.1 Mean Residence Time 9
1.2.3.2 Second and Third Moments of the RTD 10
1.3 RTD in Ideal Reactors 11
1.3.1 RTD of the Batch and PFR Reactors 11
1.3.2 RTD of an ideal CSTR 12
1.3.3 RTD of PFR/CSTR in Series 13
1.4 Modeling the Reactor with the RTD 15
1.4.1 Models with One Parameter: Tanks-in-series and Dispersion
Models 15
1.4.1.1 Tanks-in-series Model 15
1.4.1.2 The Dispersion Model 18
1.4.2 Models with Two Parameters 22
1.4.2.1 Two CSTR with Exchange of Matter 22
1.4.2.2 CSTR with Dead Volume and Short Circuit 23
1.5 Other Models of Real Reactors Using CSTR and PFR 25
Bibliography 32
2.2 Convolution 35
2.2.1 Convolution Properties 37
2.2.2 Application to a Reactor RTD 38
2.2.3 Calculating Convolution Functions 38
2.3 Deconvolution 41
2.4
2.5
®
Computer Program Using Matlab (Convolution) 44
Computer Program Using MATLAB (Deconvolution) 47
2.6 Convolution of Signals in Reactors Connected in Series 50
Bibliography 55
Index 327
xiii
Preface
The main objective of this book is to introduce the reader to the analysis of com-
plex chemical reactors. In them, reactions with complex kinetics will be carried
out and/or they will work in unusual situations, such as unsteady state operation.
The mathematical tools necessary for the characterization of flow and kinetic
models that can solve complex problems in the design of reactors are given in
the book. Reactors operating in a transitory regime are also described, and their
design equations are analyzed. Special attention is paid to the description and
design of catalytic reaction systems, in which the presence of two or more phases
makes their analysis and design complex.
Throughout the book there are many examples of application of the concepts
and equations studied, to strengthen the content treated. In the examples, com-
plex reactors are solved in terms of the characterization of their flow, and other
situations that are not usually dealt with in textbooks.
®
For the resolution of many examples, programming in Matlab (or its equiva-
lent in freeware, GNU Octave) is used. The necessary level of knowledge of this
program is very low. The use of big programs that link two or more Matlab scripts
is avoided, and simple programs are offered as alternatives, which, although more
rudimentary, are capable of performing complex calculations.
The book covers some aspects that are not treated in any textbook dedicated
to the reactor design. There are some chapters with material similar to other
texts but in many facets the work presents aspects that are not treated at all
in any other book, as is the use of numerical methods for solving engineering
problems (unsteady state regime included), the use of transfer functions to study
residence time distributions, the convolution and deconvolution curves for reac-
tor characterization, forced-unsteady-state-operation, scale-up of chemical reac-
tors, design of multiphasic reactors, and biochemical reactors design (not only
Michaelis–Menten nor Monod kinetics). Other aspects considered, mainly in
part two, are the design of multiphase gas–liquid–solid reactors, including bub-
ble reactors, agitated and trickle flow reactors.
A special emphasis is done to the numerical solution of differential equations
using the finite differences approximation. I know there are available more com-
plex tools for solving such situations, but in my opinion, it is important that
students have in mind this simple system for solving PDE. This would give a back-
ground for understanding other more complex methods.
xiv Preface
At the University of Alicante, the text is the basis for a course of 4.5 credit points
in the last year of ChemEng MS. The course is complemented with the study of
electrochemical, photochemical, and sonochemical reactors.
Nomenclature
Subscripts:
i Actual position
i+1 Position of the following interval
i−1 Position of the preceding interval
S Surface
0 Inlet conditions
mm Maximum of the maxima curve (critical point)
Superscripts
t Actual time
t+1 Time of the following interval
t−1 Time of the previous interval
1
Part I
Nonideal Flow
1.1 Introduction
Basic chemical reactors (plug flow reactor or PFR, and continuously stirred tank
reactor or CSTR) are studied considering their behavior is that of an ideal reactor.
Unfortunately, in practice, we often find behaviors that are far from that consid-
ered ideal. Consequently, working with them, the chemical engineer must be able
to handle and diagnose the behavior of these reactors. At the time of describing
the nonideal behavior of a reactor, three concepts are introduced: the residence
time distributions (RTDs), the quality of the mixture (not discussed in this book),
and the models that can be used to describe the reactor. These three concepts are
used to describe the deviations of the mixing assumed in the ideal models and
are considered as attributes of the mixture in nonideal reactors.
One way of approaching the study of nonideal reactors is to consider them, in
a first approximation, as if the flow model were the one corresponding to a CSTR
or a PFR. However, in real reactors, the nonideal flow model implies a minor con-
version, so a method that allows for this conversion loss to be considered must
be available. Therefore, a higher level of approximation implies the use of infor-
mation about the RTD.
For example, consider an ideal CSTR; the input flow that is introduced to the
reactor at a given moment mixes instantaneously and completely with the rest of
the material that already exists inside the reactor. In this way, some of the particles
that enter the reactor abandon this one almost immediately with the exit current,
whereas other atoms remain of almost indefinite form, since all the material is
never dragged. Of course, many of the particles leave the reactor after a period
close to the average residence time.
The RTD of a reactor is a feature of the mixture that is taking place inside the
reactor. Thus, in an ideal PFR there is no axial mixing, and this absence is reflected
in the RTD that this type of reactors exhibit. In contrast, in an ideal CSTR there
is a great degree of mixing, so the RTD that these reactors exhibit is very differ-
ent from that of the plug flow. However, not all RTDs are unique to one type of
reactor; reactors with marked differences can give identical RTDs. Despite this,
the RTD of a certain reactor presents distinctive keys with respect to the type of
mixture that is taking place inside it and is one of the ways to characterize the
reactor that provides more information.
Tracer
input
Feed Output
Reactor
Tracer
Tracer detection
concentration Tracer
concentration
Pulse input
Pulse
response
0 Time
0 Time
Tracer Tracer
concentration concentration
0 Time 0 Time
That is, ΔM is the amount of tracer (in moles or grams, for example) that has
remained in the reactor for a time interval comprising between “t” and “t + Δt.”
Dividing by the total amount of tracer injected into the reactor (M0 ), we obtain
the tracer fraction whose residence time in the reactor is between “t” and “t + Δt”:
ΔM C(t) ⋅ Q
= ⋅ Δt (1.2)
M0 M0
For a pulse injection, we define the RTD function, E(t), as
Q ⋅ C(t)
E(t) = (1.3)
M0
This expression describes in a quantitative way how long the different elements
of fluid have passed inside the reactor. In consequence:
ΔM
= E(t) ⋅ Δt (1.4)
M0
If M0 is not directly known, it may be obtained from the output concentrations,
adding the different amounts of the tracer that have exited the reactor between 0
and infinity. Expressing in differential form:
dM = Q ⋅ C(t) ⋅ dt
and integrating, we obtain
∞
M0 = Q ⋅ C(t) ⋅ dt (1.5)
∫0
Since the volumetric flow (Q) is generally constant, we will define E(t) as
C(t)
E(t) = ∞ (1.6)
∫0 C(t) ⋅ dt
6 1 Nonideal Flow
In that expression, the integral in the denominator is the area under the curve,
C(t). In this way, from the tracer concentration, C(t), it is possible to find the curve
E(t), as long as that curve is obtained from a perfect pulse of the input tracer.
Another way to interpret the function of residence times is in its integral form:
[ ] t2
Amount of tracer exiting the reactor after
= E(t) ⋅ dt (1.7)
passing inside it a time between t1 and t2 ∫t1
Now, the fraction of tracer that passed inside the reactor a time t, between 0
and infinity is equal to 1; therefore:
∞
E(t) ⋅ dt = 1 (1.8)
∫0
The main drawback of the use of the pulse technique lies in the difficulty in
achieving a tracer input to the reactor that is reasonably pulsed (as we explain
in Chapter 2, deconvolution of curves). The injection should take place in a very
short period compared to the residence times and the tracer dispersion between
the injection point and the reactor inlet should be negligible. If these conditions
are satisfied, the technique is a simple and direct way to obtain the RTD.
The tracer concentration in the feed should be maintained at this value until
the concentration of tracer in the effluent is practically C 0 , i.e. equal to that of the
feed, at which time the test can be interrupted. Figure 1.1 shows a typical output
concentration curve for this type of input.
Since the input concentration (C 0 ) remains constant over time, we can extract
it from the integral using Eq. (1.9):
t
′ ′
Cout (t) = C0 E(t )dt (1.10)
∫0
Obviating that t′ is mathematically equal to t, we see that the signal at the output
of this experiment is a cumulative function of E(t), as it evaluates the integral of
all E(t) from t = 0 to the instant “t.” This cumulative distribution is called the “F
curve” and can be directly determined from a step input.
[ ] t
C(t)
= E(t)dt = F(t) (1.11)
C0 step input ∫0
Differentiating this expression, we obtain the function of RTD, E(t):
[ ]
d C(t)
E(t) =
dt C0 step input
The determination of the E(t) using a step input is, in general, easier to carry
out experimentally than the pulse input. It has some other advantages, as in that
it is not necessary to know the total amount of tracer introduced during the
test period. In return, it has some disadvantages: (i) sometimes, it is not easy
to maintain a constant concentration of tracer in the feed; (ii) for calculating the
E(t) curve, this procedure implies the differentiation of the data, which can lead,
sometimes, to errors; (iii) as the feed should be maintained for a long period of
time, the amount of tracer required is usually very large, so if the tracer is expen-
sive, a pulse input is usually used.
Other techniques to introduce the tracer are possible: negative step (dilution of
the food), a periodic signal, or a random signal. However, they tend to be more
difficult to carry out and are discussed in the following chapters.
Gas in Solid
Solid
input Time
Figure 1.2 Two E(t) functions are needed to characterize a reactor with two moving phases.
extremely fast. Nevertheless, the gas will flow through this bed, and will present
a specific and complicated flow pattern, as we know. This pattern will depend on
different factors and usually is between the plug flow and the CSTR behavior. The
measurement of the precise RTD of both phases will give important information
for the design and scale-up of the reactors.
0.8
0.4
0.2
0
40
Time
1.2 Residence Time Distribution (RTD) Function 9
it can be affirmed that 80% of the input tracer spends less than 40 units of time
in the reactor.
u ⋅ dv = u ⋅ v − v ⋅ du
∫ ∫
Applying to Eq. (1.15):
1
V
= t[1 − F(t)]∞
0 +∫ t ⋅ dF (1.16)
Q 0
10 1 Nonideal Flow
tm t
1.3 RTD in Ideal Reactors 11
E(t) E(t)
Negative Positive
skew skew
t t
Figure 1.5 Form of the RTD curve for positive and negative skewness.
The third moment is also calculated around the average and is known as skew-
ness. It is defined as
∞
1
s3 = (t − tm )3 ⋅ E(t)dt (1.22)
𝜎 3∕2 ∫0
The greatness of this moment measures the extent to which the distribution
with respect to the mean is displaced in one direction or another. A positive value
of the skewness indicates that a tail to the right is expected in the distribution,
and to the left if the skewness is negative. Figure 1.5 shows a scheme.
0 x0 x
In this way, the variance of this distribution is zero, by the fact that all values of
(t − t m ) are zero (all the signal is obtained at t = t m ). The function can be seen in
Figure 1.6.
t t
Feed
CSTR PFR
CA, T
Product
tp t
This output concentration will exit the PFR in series delayed by a time tp . There-
fore, the RTD of the series reactor system will be given by (Figure 1.9)
E(t) = 0 if t < t p
( )
1 (t − t p )
E(t) = ⋅ exp − if t ≥ t p (1.34)
tt tt
Note that Eq. (1.34) is the same as Eq. (1.31), but the time scaling has been
moved tp units.
Let us see now the case of a CSTR preceded by a PFR. If a signal tracer is intro-
duced into the input pulse PFR, the same signal appears at the input of the CSTR,
but with a delay of tp seconds, so the system’s RTD will be the same as when the
CSTR is the first reactor and is followed by a PFR. That is, the order in which both
reactors are placed is not important, and the resulting RTD is the same provided
that the sum of residence times in the two sections is the same.
However, this is not the only thing that should be considered; in case the reac-
tion that takes place in the system is of the second order, the conversion that
would be obtained with both dispositions would be different, as the conversion
depends on the concentration. Contrarily, a first-order reaction will produce the
same conversion both for PFR + CSTR and CSTR + PFR. This means that the
RTD is not a complete description of what happens in a reactor or reactor sys-
tem. The RTD is unique for a particular reactor; however, the reactor or reactor
system is not unique to a particular RTD. In this way, in nonideal reactors, the
RTD gives information that is not enough to characterize their behavior, and we
need more information.
1.4 Modeling the Reactor with the RTD 15
Feed
CA0
…..
CA1, V1
CAn, Vn
CA2, V2
If an impulse tracer signal is injected into the first tank, the tracer fraction that
leaves the third reactor system after remaining in the system for a time between
t and t + dt is given by E(t) ⋅ dt, which can be estimated from the concentration
obtained in a pulse tracer experiment:
CA3 (t)
E(t) = ∞ (1.35)
∫0 CA3 (t) ⋅ dt
In this expression, C A3 (t) is the concentration of the tracer at the outlet of the
third reactor. Now, we must obtain how this concentration varies with time. For
a unique CSTR, the mass balance will be
V1 ⋅ dCA1
= −Q ⋅ CA1 (1.36)
dt
Integrating, we obtain the expression of the tracer concentration at the exit of
that reactor:
( ) ( )
t t
CA1 = CA0 exp −Q ⋅ = CA0 exp − (1.37)
V1 t1
Since the volumetric flow is constant Q = Q0 and that the volume of all the
reactors is the same (V 1 = V 2 = V 3 ), the average times will be identical (t1 = t2 =
t3 = t𝜄 ), t𝜄 being the residence time in each one of the reactors, not in the whole
system. Posing a balance of the tracer in the second reactor:
V2 dCA2
= Q ⋅ CA1 − Q ⋅ CA2 (1.38)
dt
1.4 Modeling the Reactor with the RTD 17
Taking into account the expression that we have previously obtained for C A1 ,
we arrive at the differential equation:
( )
dCA2 CA2 t
+ = CA0 exp − (1.39)
dt t2 t1
which can be solved using an integration factor together with the initial condition
C A2 = 0 for t = 0:
( )
CA0 t t
CA2 = ⋅ exp − (1.40)
ti ti
Using the same procedure for the third reactor, we obtain the expression for the
tracer concentration at the exit of the third tank (and, therefore, of the system):
( )
CA0 t 2 t
CA3 = 2
⋅ exp − (1.41)
2t i ti
Substituting in the equation for the curve E(t):
( )
C3 t2 t
E(t) = = 2 ⋅ exp − (1.42)
C0 2t i ti
If we generalize for nt equal tanks in series:
( )
t nt −1 t
E(t) = nt ⋅ exp − (1.43)
(nt − 1)!t i ti
since t i = t/nt , where t is the quotient of the total volume of the system by the
volumetric flow rate Q.
Figure 1.11 shows the RTD for different CSTR numbers in series. As nt
increases, the behavior is closer to piston flow.
The number of reactors in series can be calculated from the dimensionless
variance 𝜎 2 :
∞ 2
t
𝜎2 = (t − t)2 ⋅ E(t)dt = … = (1.44)
∫0 nt
t t
18 1 Nonideal Flow
That is an expression that gives us the number of tanks needed to model the
nonideal reactor as a series of nt CSTR connected in series.
Let us now calculate the conversion of a reaction in the tanks in series. If the
reaction is of the first order:
CA0
1st reactor ∶ CA1 = (1.45)
1 + t1 k
CA1 CA0
2nd reactor ∶ CA2 = = (1.46)
1 + t2 k (1 + t 1 k)(1 + t 2 k)
As all residence times are equal, t i , and the temperature is constant (k 1 = k 2 = k),
we can generalize the expression as
CA0
CAn = (1.47)
(1 + t i k)nt
Therefore, we can express the conversion as
1
XA = 1 − (1.48)
(1 + t i ⋅ k)nt
In general, the value of nt obtained from the variance is considered as a non-
integer number when calculating the conversion. In this sense, equations of the
model can be applied to fractional number of tanks. Nevertheless, if the reaction
is not of the first order, sequential molar balances must be made in each reactor
(see Example 1.3 in the following sections).
Note that the term corresponding to the dispersion of the component “A” is based
on Fick’s law for diffusion. If we do an inert tracer balance in a differential volume:
In − Out = Accumulation
dCT dCT
− dnT = dV ⋅ = S ⋅ dz ⋅ (1.50)
dt dt
and, using partial derivatives:
𝜕nT 𝜕CT
− =S⋅ (1.51)
𝜕z 𝜕t
Substituting for nT (Eq. (1.49)) and dividing by the cross-section S:
𝜕 2 CT 𝜕(u ⋅ CT) 𝜕CT
De − = (1.52)
𝜕z2 𝜕z 𝜕t
If we divide by (u⋅L):
De 𝜕 2 CT 𝜕CT 𝜕CT
− = (1.53)
u ⋅ L 𝜕z 2 L ⋅ 𝜕z u ⋅ L ⋅ 𝜕t
The parameter De /uL is the so-called recipient dispersion module or the
Peclet–Bodenstein module (Bo = De /uL) that measures the degree of axial
dispersion. When this module tends to zero, the system is close to piston flow,
and when it tends to infinity (large dispersion), we have complete mixing. In the
case of a packed bed, the module would be (𝜀⋅De /u⋅dp ), where dp is the particle
diameter and 𝜀 the porosity of the bed.
Equation (1.52) is only solvable for small values of Bo number, usually Bo < 0.01.
In that case, the dispersion modifies the input signal in the reactor, but the tracer
widening does not vary in the measuring point with time, in such a way that the
boundary conditions are well known. In this situation, the expression for the RTD
curve of the reactor is
( )2
⎡ t ⎤
⎢ 1 − ⎥
1 t
E= √ ⋅ exp ⎢− ⎥ (1.54)
t ⋅ 4 ⋅ 𝜋 ⋅ Bo ⎢ 4 ⋅ Bo ⎥
⎣ ⎦
V
tm = t = (1.55)
Q
𝜎 2 = 2 ⋅ Bo (1.56)
Nevertheless, if the value of the Bodenstein module Bo is higher than 0.01, the
response of the tracer to the impulse is wide and passes through the measuring
point so slowly that it can change its form during the time of measuring. This
produces an asymmetrical E curve that, on some occasions, does not have an
analytical expression for the E curve.
In this case, the E curve also depends on what happens in the input and output
sections of the reactor vessels. We consider two cases: closed boundary condi-
tions (where there is plug flow behavior outside of the system), and open bound-
ary conditions (where the flow is not affected when passing through the system).
20 1 Nonideal Flow
Plug flow
For simplifying all possibilities, let us consider only two cases: the closed–closed
containers in which there is neither dispersion nor radial variation of the con-
centration, both upstream and downstream of the reaction zone; and open–open
containers in which there is dispersion both before and after the reaction zone.
Both cases are shown in Figure 1.12, where it is observed that the fluctuations
of the concentration due to dispersion overlap the piston flow velocity profile. A
closed–open container would have no dispersion at the entrance but only at the
exit of the reaction zone.
Boundary Conditions for a Closed–Closed Vessel (Bo > 0.01) In this case, immediately
before the reactor entrance zone (z = 0− ) and immediately after the exit zone
(z = L+ ), we have piston flow (without dispersion). However, between z = 0+ and
z = L− , there is dispersion and convection by the flow movement. The corre-
sponding boundary condition at the input is
nT (0− , t) = nT (0+ , t) (1.57)
Substituting the value nT of at each side, we obtain
[ ]
𝜕CT
uSCT (0 , t) = −SDe
−
+ uSCT (0+ , t) (1.58)
𝜕z z=0+
Taking into account that at the input C T (0− , t) = C T0 (known concentration):
[ ]
De 𝜕CT
CT0 = − + CT (0+ , t) (1.59)
u 𝜕z z=0+
Also, at the exit of the reactor considered, we can write C T (L− , t) = C T (L+ , t),
this being later the measured exit concentration. In this way, when z = L:
CT (L− , t) = CT (L+ , t)
( )
𝜕CT C (L− , t) − CT (L+ , t)
= T =0 (1.60)
𝜕z z=L 𝜕z
The combination of Eqs. (1.59) and (1.60) are known as the Danckwerts’ bound-
ary conditions.
On the other hand, the initial condition of the reactor at t = 0 is
t = 0, z > 0 CT (0+ , 0) = 0 (1.61)
1.4 Modeling the Reactor with the RTD 21
Boundary Conditions for an Open–Open Container (Bo > 0.01) These conditions
would be applicable in the case of a packed bed in which the tracer was injected
at a point downstream of the inlet, a distance around two to three times the
diameter, and whose concentration was measured at a certain distance before
the exit. A solution of the differential equation (Eq. (1.52)) could be obtained in
the case of a pulse injection.
For an open–open system, the boundary condition at the input is
nT (0− , t) = nT (0+ , t) (1.64)
Note that the expression is the same as that obtained in the previous case. If
the dispersion coefficient is the same at the entrance as in the reaction zone, we
will have
[ ] [ ]
𝜕CT 𝜕CT
−De + u ⋅ CT (0− , t) = −De + u ⋅ CT (0+ , t) (1.65)
𝜕z z=0− 𝜕z z=0+
As we can imagine, the derivatives at z = 0+ and z = 0− are the same, as no
discontinuity is included in the model, so:
CT (0− , t) = CT (0+ , t) (1.66)
while on the exit:
[ ] [ ]
𝜕CT 𝜕CT
−De + u ⋅ CT (L , t) = −De
−
+ u ⋅ CT (L+ , t) (1.67)
𝜕z z=L− 𝜕z z=L+
CT (L− , t) = CT (L+ , t) (1.68)
In addition to these boundary conditions, many other modifications may occur.
For example, the dispersion coefficient can have different values in each of the
three regions (before the entrance, in the reaction zone, and after the output)
and/or the tracer can be injected at a point other than z = 0. However, we consider
only the case that the dispersion coefficient is the same for any value of z and that
the pulse tracer is injected at the point z = 0 at time t = 0.
22 1 Nonideal Flow
Q1
CSTR CSTR
CA2, V2
CA1, V1 Q1
Product CA1, Q0
dCT2
V2 = Q1 CT1 − Q1 CT2 (in the second tank) (1.73)
dt
where C T1 and C T2 are, respectively, tracer concentrations in both reactors. These
two differential equations are coupled and should be solved simultaneously.
In this model, the two adjustable parameters are the flow rate exchanged (Q1 )
and the volume of the most agitated region (V 1 ). Remember that the measured
volume (V ) is the sum of V 1 and V 2 . We will call 𝛽 the fraction of the total flow
that is transferred between both reactors:
Q1 = 𝛽 ⋅ Q0 (1.74)
and 𝛼 the fraction of the total volume that corresponds to the most agitated
area:
V1 = 𝛼 ⋅ V → V2 = (1 − 𝛼) ⋅ V (1.75)
On the other hand, the average time (t) is given by the quotient V /Q0 .
The initial conditions (t = 0) for this model are (i) C T1 = (C T1 )0 , and (ii)
(C T2 )0 = 0
An analytical solution is possible in this case, and is as follows:
( ) ( )
[ ] (𝛼m1 + 𝛽 + 1) exp
m2 t
− (𝛼m2 + 𝛽 + 1) exp
m1 t
CT1 t t
=
(CT1 )0 pulse 𝛼(m1 − m2 )
(1.76)
being:
[]⎡ √
⎤
1−𝛼+𝛽 ⎢ 4𝛼𝛽(1 − 𝛼) ⎥
m1 , m2 = −1 ± 1− (1.77)
2𝛼(1 − 𝛼) ⎢ (1 − 𝛼 + 𝛽)2 ⎥
⎣ ⎦
However, for more complicated models, an approximate solution would be nec-
essary.
Equation (1.76) shows that, if tank 1 is small compared to 2 (𝛼 small) and the
transfer speed between both reactors is small (𝛽 small), the second exponential
term tends to 1 during the first part of the response to an impulse injection. Dur-
ing the second part, the first exponential term tends to 0. If we represent the
logarithm of the tracer concentration vs. time, the response curve will tend to a
straight line at both ends of the curve and the parameters will be obtained from
the slopes (m1 for t → ∞ and m2 for t → 0) and the cut points of both lines (for
t → ∞, the cut point is – {𝛼m2 + 𝛽 + 1}/𝛼{m1 − m2 }).
Feed
CT0, Q0 CT0, Qb
CT0, Qt
Bypass CSTR Vd
Dead volume CTs, Vt
Q0 = Qt + Qb
Product CTs, Qt
CT
Figure 1.14 Real reactor and modeling using a single CSTR with dead volume and short
circuit.
In this case, the derivation of the equations is simpler if we consider the injec-
tion of a tracer in positive step. Let us use the scheme in Figure 1.14; the balance
in nonstationary regime (at t > 0, some tracer is still entering the system) of non-
reactive tracer in the volume of reactor V t , is
dMTs dC
Qt ⋅ CT0 − Qt ⋅ CTs = = Vt Ts (1.78)
dt dt
Remembering that for a positive step entry it is fulfilled that:
t < 0 → CT = 0
t ≥ 0 → CT = CT0
The tracer balance at the point of union of both currents will be
Qb ⋅ CT0 + Qt ⋅ CTs
CT = (1.79)
Q0
If we define:
Vt = 𝛼V and Qb = 𝛽Q0 , with t = V ∕Q0
Integrating and replacing in the expression of tracer balance in the reactor,
we get
[ ( )]
CTs 1−𝛽 t
= 1 − exp − (1.80)
CT0 𝛼 t
So, the expression of the tracer concentration that leaves the system will be
[ ( )]
CT 1−𝛽 t
= 1 − (1 − 𝛽) ⋅ exp − (1.81)
CT0 𝛼 t
If we reorder the equation, we can obtain the parameters of the model
(Qt and V t or similarly 𝛼 and 𝛽) from a plot of the tracer concentration at the
output as a function of time. Representing ln[C T0 /(C T0 − C T )] vs. time, if the
model is correct, a straight line of slope (1 − 𝛽)/t should be obtained 𝛼 and an
ordinate in the origin of value ln[1/(1 − 𝛽)].
In order to have in mind the possible RTD curves obtained with this model,
Figure 1.15 shows three curves corresponding to three different cases: the ideal
1.5 Other Models of Real Reactors Using CSTR and PFR 25
0.3
0.2
0.1
0
0 5 10 15 20
Figure 1.15 E(t) predicted by this model with different values of the parameters, and the one
for ideal CSTR for comparison.
CSTR, a reactor where 𝛼 = 0.5 and 𝛽 = 0.3 and the other reactor with 𝛼 = 0.5 and
𝛽 = 0.1; all of them with the same average residence time of five minutes. As we
can see, the differences are quite small and, in practice, it is difficult to affirm with-
out a little more information if the system has or not a dead volume and/or bypass.
Q Vp Q E(t) F(t)
Vd
Vp t Vs t
t= t=
Q Q
Q Vs Q E(t) F(t)
Vd
Vs t Vs t
t= t=
Q Q
Q1
Q Vp Q E(t) F(t)
Q2
t expected
Vp t Vp t
t= t=
Q1 Q1
Q1
Q Vs Q E(t) F(t)
Q2
t expected
Vs t Vs t
t= t=
Q1 Q1
Q1 Vp1
E(t) F(t)
Q Q
t expected
Q2 Vp2
(1 + r)Q
E(t) F(t)
Q Vp Q
rQ
Solution
During the design of the reactor, a plug flow must be assumed. In this case, the
molar balance of the reacting species “A,” for a first-order reaction, gives
XA = 1 − exp(−kt)
In the present case:
V 0.1 ⋅ 3
t= = = 10 s
Q 0.03
Coinciding with the value of t m . From the previous equations: k = 0.39 s−1
Assuming a dispersion model for the real reactor, Eq. (1.63) is fulfilled, so:
( )2 [ ( )]
𝜎 1
= 2 ⋅ Bo − 2 ⋅ Bo2 ⋅ 1 − exp −
tm Bo
Iterating, we can find the value of the dispersion module: Bo = (De /u⋅L) = 0.667
We have not seen in the previous sections the equations for the balances in
reacting systems. Let us see this now. If we consider a tubular reactor in which
we simultaneously have dispersion and reaction and we can do a molar balance
of component A, in a range Δz of the reactor:
Input − Output + Generation = Accumulation
nA − (nA + dnA ) + rA dV = 0
1 dnA
− + rA = 0 (1.82)
S dz
S ⋅ dz = dV
Combining this expression with the molar flow of substance A:
𝛿C
nA = −De S A + uSCA (1.83)
𝛿z
we obtain a differential equation of the second order:
De d2 CA dCA rA
− + =0 (1.84)
u dz2 dz u
which is only linear when the reaction rate is of order 0 or 1.
When the kinetics is of the first order (rA = −kC A ), the following expression is
obtained:
De d2 CA dCA kCA
− − =0 (1.85)
u dz2 dz u
which considers the flow, dispersion, and reaction. The solution of this second-
order differential equation can be done analytically. If we consider a closed–
closed system, we will apply the Danckwerts’ boundary conditions at the input
and at the exit of the reactor:
• Taking into account that at the input C A (0− ,t) = C A0 (known concentration).
D 𝜕CA ||
CA0 = − e ⋅ + CA (0+ , t)
u 𝜕z ||z=0+
28 1 Nonideal Flow
• At the exit of the reactor considered, we can write C A (L− , t) = C A (L+ , t), this
being later the measured exit concentration. In this way, when z = L:
( )
𝜕CA
=0
𝜕z z=L
Finally, the solution for the conversion in the reactor can be expressed by
( )
4a ⋅ exp 21uL
De
1 − XA = ( ) ( ) (1.86)
(1 + a)2 ⋅ exp 2 D − (1 − a)2 ⋅ exp − 2auL
auL
D
e e
where a = [1 + 4(t ⋅ k) ⋅ (De ∕uL)] . This expression allows knowing the conver-
1∕2
sion that would be obtained for a first-order reaction to be carried out in a tubular
reactor or in a bed reactor packed with dispersion.
For the system of the example, the value of “a” results to be 3.376 and the con-
version in the real reactor is 0.88
Time (min) 0 2 4 6 7 8 10 12 14 16 18 20
C T (mol/l) 3.0 5.3 7.2 8.0 8.3 8.6 9.2 9.7 9.7 9.8 9.9 10
The measured volume of the reactor is 1 m3 and the flow rate to the reactor is
0.1 m3 /min. The reaction rate constant is 150 l/(kmol min). The feeding contains
a concentration of A at the input of 2 kmol/m3 . Calculate the conversion that can
be expected in this reactor.
Solution
Let us first have a look of the experimental results. The graph C T -time corre-
sponding o the data is shown in Figure 1.17a.
As we have seen, Eq. (1.81) shows that the tracer concentration for a step input
in this model can be expressed by
[ ( )]
CT 1−𝛽 t
= 1 − (1 − 𝛽) ⋅ exp − (1.87)
CT0 𝛼 t
From the data in the table, it is easy to calculate the slope and intercept
( of the)
representation ln(C T /(C T0 − C T )) vs. t. The slope would correspond to − 1−𝛽𝛼t
and the intercept to ln(1/(1 − 𝛽)). In Figure 1.17b, we can see the corresponding
straight line. Taking into account an intercept of −0.4549 and a slope of 0.2875,
as solution of the model we find 𝛼 = 0.3 and 𝛽 = 0.3 bearing in mind that the
1
experimental value of average residence time t = VQ = 0.1 = 10 minutes should
be used as it was defined for the whole system.
1.5 Other Models of Real Reactors Using CSTR and PFR 29
12.0
10.0
Concentration (mg/l)
8.0
6.0
Correlation Experimental
4.0
2.0
0.0
0 5 10 15 20 25 30
(a) Time (min)
3
y = 0.2875x – 0.4549
2 R2 = 0.9804
–1
–2
0 2 4 6 8 10 12 14 16 18 20
(b) Time (min)
Figure 1.17 (a) Concentration of tracer obtained in a step input experiment and (b)
calculation of the parameters.
For the system with reaction, let us see how to obtain the conversion with this
model in the case of the first-order reaction (not in this case, but will be useful).
For a first-order reaction, the molar balance of A in the reactor where the reaction
takes place (V t ) gives
Qt CA0 − Qt CAs − kCAs Vt = 0
CA0 (1 − 𝛽)Q0
CAs =
(1 − 𝛽)Q0 + 𝛼Vk
If we do a reagent balance A in the point where the short circuit current and
the output current of the reactor are mixed, we will obtain
CA0 Qb + CAs Qt = CA (Qb + Qt )
30 1 Nonideal Flow
Using the data in the example, the concentration at the exit of the CSTR is
0.927 kmol/m3 . The value of C A at the exit of the system can be calculated from
Eq. (1.80). Finally, in the example, C A = 1.249 kmol/m3
Solution
First of all, we can plot the data obtained with the tracer.
9,00E–04
8,00E–04
7,00E–04
6,00E–04
5,00E–04
E(t) s–1
4,00E–04
3,00E–04
2,00E–04
1,00E–04
0,00E+00
0 5000 10 000 15 000 20 000 25 000
Time (s)
As we can see, the curve is similar to a CSTR, but small differences can be
accounted with the use of the tanks-in-series model. From the data, it is easy to
calculate:
∞
tm = t ⋅ E(t) ⋅ dt = 1129 s
∫0
∞
𝜎2 = (t − tm )2 ⋅ E(t) ⋅ dt = 5.420 ⋅ 105 s2
∫0
We can calculate the number of tanks in the model, that is, nt = t m 2 /𝜎 2 = 2.221
tanks.
We have seen in the previous sections the equations for the balances in react-
ing systems or first order. In that case, Eq. (1.48) with nt = 2.221 would give us
the expected conversion. Let us see what occurs if the system is nth order. If
we consider a single tank reactor (of volume V 1 ) we can do a molar balance of
component A:
V1 C − CA C − CA1
= A0 = A0 (1.93)
nA0 CA0 (rA ) CA0 (k ⋅ CA1
n
)
If the system has a constant density:
V C −C
t1 = = A0 n A1 (1.94)
Q (k ⋅ CA1 )
CA0
CA1 = (1.95)
n−1
(1 + kt 1 CA1 )
As we can check, the concentration at the exit can be calculated from this non-
linear equation, and an approximation can be used.
32 1 Nonideal Flow
Solution
(a) If the three reactors are connected in series, the equations just discussed in
previous example are valid, so the concentration at the exit of each reactor
will be given by Eqs. (1.95) and (1.96). The one corresponding to the third
reactor is equivalent.
(b) On the contrary, if the reactors are connected in parallel, each of them will
have a flow rate equal to 1/3 of the total. The concentration at the exit of the
reactors will be the same in all three. And the mix of three currents of the
same concentration in the same proportion, will give a current with exactly
the same concentration, calculated by Eq. (1.95) bearing in mind the volume
and the flow rate at each reactor.
Bibliography
Aris, R. (1969). Elementary Chemical Reactor Analysis. Englewood Cliffs: Prentice
Hall.
Bourne, J. (1999). Turbulent mixing and Chemical Reactions. New York: Wiley.
Bibliography 33
Carberry, J.J. and Varma, J. (eds.) (1987). Chemical Reaction and Reactor
Engineering. New York: Marcel Dekker.
Coulson, J.M. and Richardson, J.F. (1979). Chemical Engineering III. Chemical
Reactor Design, 2a ed. Oxford: Pergamon Press.
Fogler, H.S. (1998). Elements of Chemical Reaction Engineering, 3a ed. New Jersey:
Prentice Hall.
Holland, C.D. and Anthony, R.G. (1992). Fundamentals of Chemical Reaction
Engineering, 2a ed. Englewood Cliffs: Prentice-Hall.
Kayode Coker, A. (2001). Modelling of Chemical Kinetics and Reactor Design.
Elsevier.
Levenspiel, O. (1999). Chemical Reaction Engineering, 3rde. New York: Wiley.
Levenspiel, O. (1962). Chemical Reaction Engineering: An Introduction to the Design
of Chemical Reactors. New York: Wiley.
Levenspiel, O. (1979). The Chemical Reactor Omnibook. OSU. Bookstores, Corvalis.
Westerterp, K.R., Van Swaaij, W.P.M., and Beenackers, A. (1984). Chemical Reaction
Design and Operation, 2nde. Chichester: Wiley.
35
2.1 Introduction
As discussed in Chapter 1, both pulse and step tracer input can be used to char-
acterize the residence time distribution (RTD) in chemical reactors. One of the
problems of these techniques is that the pulse or the step is usually not perfect, in
the sense that deviations from the ideal pulse or step are usually present. In this
chapter, the evolution of the output signal in a reactor is studied for an arbitrary
input of tracer.
Also important are the systems where more than one reactor is combined in
series. In this case, the signal entering the first reactor is modified according to
its RTD, and the modified signal enters the second reactor, where it is modified,
and so on.
In this sense, the convolution of the signals enables us to calculate the response
to an arbitrary stimulus function, as is explained later. It is also possible to calcu-
late the impulse response E(t) by deconvolution from the measured stimulus and
response.
2.2 Convolution
First of all, let us focus on the idea of the convolution operation. This is very
important (and maybe difficult) to understand the behavior of signals (tracer
curve response) throughout a series of reactors.
During the convolution of two functions “f” and “g,” we will calculate a third
function that expresses how the sum of one of the functions is modified by the
other. Mathematically, a convolution is defined as the integral over all space of
one function “f (x)” at “x” times displacing another function g(x) given at “u − x.”
The integration is taken over the variable x, typically from minus infinity to infin-
ity over all the dimensions (in case the variable is more than one dimension). In
this way, the convolution is a new function of a new variable u, as shown in the
following equations:
g(x) f(x)
2 2
1 1
–2 0 2 4 6 –2 0 2 4 6
g(u – 1) f(2)g(u – 2)
2 2
1 1
–2 0 2 4 6 –2 0 2 4 6
g(u – 4) f(4)g(u – 4)
2 2
1 1
–2 0 2 4 6 –2 0 2 4 6
f(x)g(u – x) f(x)⊗g(u – x)
2 2
1 1
–2 0 2 4 6 –2 0 2 4 6
The cross in the circle is used to indicate the convolution operation. Another
way of figuring out the concept of convolution is to see it as an integral stating
the amount of overlap of one function “g” as it is moved over another function
“f .” In some manner, the convolution “blends” one function with another.
Figure 2.1 shows a graphical example of the convolution of two given functions,
f (x) and g(x). At the top of the figure, we have the original functions. Basically, the
figure shows how we can consider the convolution, as a weighted sum of copies
of one function displaced in the ordinate axis. In the sum, the fractions are given
by the function value of the second function at the shift vector. The next three
pairs of graphs show (on the left) the function “g” shifted by various values of x
and, on the right, that shifted function “g” multiplied by “f ” at the value of x.
The last two graphs show the superposition of all weighted and displaced copies
of “g”, and, on the right, the integral (i.e. the sum of all the weighted, shifted copies
of “g”). We can see that the biggest contribution comes from the copy shifted by
4, the position of the peak of “f .”
If one of the functions is a Dirac delta centered at x1 , the other function will be
shifted by a vector equivalent to the position of the peak, i.e. it will be centered at
2.2 Convolution 37
(x1 + x2 ), x2 being the original position of the second function. This is discussed
later when treating the convolution of RTD signals in chemical reactors.
Other interesting properties of the convolution of curves is that the moments
of the convolution are the sum of the moments of the individual curves, that is,
the mean residence time of the convoluted curve is the sum of the mean residence
times, and this is also true for the variances.
As pointed out later, it is not possible to calculate the convolution of two given
functions if both are given at points with different Δx.
Assume that f (x) is a vector of “n” values, i.e. f (x) = [f 1 , f 2 , f 3 , …, f n ] and g(x)
has “m” values so that g(x) = [g 1 , g 2 , g 3 , …, g m ].
As an example, if n = 5 and m = 4, for the convolution operation the desired
result is
C1 = f 1 ⋅ g1
C2 = f 2 ⋅ g1 + f 1 ⋅ g2
C3 = f 3 ⋅ g1 + f 2 ⋅ g2 + f 1 ⋅ g3
C4 = f 4 ⋅ g1 + f 3 ⋅ g2 + f 2 ⋅ g3 + f 1 ⋅ g4
C5 = f 5 ⋅ g1 + f 4 ⋅ g2 + f 3 ⋅ g3 + f 2 ⋅ g4
C6 = f 5 ⋅ g2 + f 4 ⋅ g3 + f 3 ⋅ g4
C7 = f 5 ⋅ g3 + f 4 ⋅ g4
C8 = f 5 ⋅ g4
2.2 Convolution 39
(as mentioned before), it is necessary that the time increments in both vectors
(Δt) are equal, i.e.:
tm,Cin − t0,Cin tn,E − t0,E
Δt = = (2.18)
m−1 n−1
where “n” and “m” are, respectively, the number of points of vectors E and C in .
Let us note that the convoluted signals C out would contain (v = n + m − 1)
values in a vector, and the corresponding time vector could be calculated by
t0,convolution = t0,Cin + t0,E (initial time is the sum of the initial times)
tv,convolution = tm,Cin + tn,E (final time is the sum of the final times)
tv,convolution − t0,convolution
Δtconvolution = (2.19)
m+n−2
Applying Eq. (2.15), we can also see that the following is true:
(m − 1) ⋅ Δt + (n − 1) ⋅ Δt ( m + n − 2 )
Δtconvolution = = ⋅ Δt = Δt
m+n−2 m+n−2
(2.20)
That is, the time increment of the convoluted curve is the same as those
of the curves (remember that this time increment should be the same for
both curves).
t (s) 0 1 2 3 4
C (mol/l) 0 8 4 6 0
t (s) 2 3 4 5 6
E 0 0.05 0.50 0.35 0
Solution
We must calculate both the concentration and the time vectors. The time vector
is easy to calculate. We have m = 5 points and n = 5 points, so the resulting vector
will have 9 points (m + n − 1). The first value of time is (0 + 2) = 2 and the last one
is (6 + 4) = 10. The corresponding Δt convolution is (10 − 2)/(m + n − 2) = 8/8 = 1;
therefore:
t convolution (s) 2 3 4 5 6 7 8 9 10
2.3 Deconvolution 41
2.3 Deconvolution
Once we have outlined the convolution operation, let us propose the inverse oper-
ation, i.e. the deconvolution of curves. This is a more complicated issue and is
rarely treated in specialized books and papers. Let us assume that we want to
know the E curve of a reactor, where we feed a given amount of tracer in the form
C in (associated to a time vector) and where we are measuring the exit concentra-
tion C out . This is a very common problem because the calculation of E from an
impulse input signal (C in = Dirac delta function) or a step function is not always
possible or easy. Assuming the dimension of vectors is “m” for C in and “v” for C out
(as used earlier), we have that C out = E ⊗ C in and
Cout(v×1) = [A(v×m) ⋅ Cin(m×1) ] (2.21)
Note that the dimensions of the convolution matrix A are (v × m) and from
that we will calculate (n = v − m + 1) values of the vector E curve (according to
Eq. (2.16)) (check Example 2.2 for details).
42 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors
tv,E − t0,E
Δtdeconvolution = = Δt (2.23)
v−1
Since (v − 1) = (m + n − 2), then Eq. (2.23) is the same as Eq. (2.20).
t out (min) 4 5 6 7 8 9 10 11 12 13
C out (mg/l) 0.9 1.8 2.1 5.2 3.6 4.5 1.7 0.8 0.7 0.5
Indicate how to calculate the E(t) of the reactor, together with the time vector
associated to the E(t).
Solution
For doing this deconvolution, the easiest method is to calculate discrete values
of C in and then deconvolute. We can choose the number of points we use for
the calculation (with limits), and also at the time values, but the time increment
should be the same so that the C out signal (i.e. one minutes) and the C in cannot be
zero in all points. Another restriction is that we will need enough data for C in . We
have v = 10 points of the convoluted curve (C out ). For example, if we use m = 10
points to calculate C in , as we have v = 10 values of C out , we can only calculate
n = v − m + 1 = 1 point of the RTD. If we use m = 9 points in C in , we will be able
2.3 Deconvolution 43
t in (min) t 0 1 2 3 4
C in (mg/l) t*exp(−t/2) 0 0.61 0.74 0.67 0.54
Following the reasoning presented before, the deconvoluted signal will have
(10 − 5 + 1) = 6 points. The time vector of these points is easy to calculate, as
t E,0 = t out,0 − t in,0 and the time increment is one minutes; therefore:
t E (min) 4 5 6 7 8 9
For calculating the E(t), we should solve the following equations system:
⎛0.9⎞ ⎡E1 0 0 0 0 ⎤
⎜1.8⎟ ⎢⎢E2 E1 0 0 0 ⎥
⎥
⎜ ⎟ ⎢
⎜2.1⎟ ⎢E3 E2 E1 0 0 ⎥⎥ ⎛ 0 ⎞
⎜5.2⎟ ⎢E4 E3 E2 E1 0 ⎥ ⎜
⎜ ⎟ ⎢E 0.61 ⎟
⎜3.6⎟=⎢ E E E E ⎥ ⎜ ⎟
5 4 3 2 1
⋅ 0.74 ⎟
⎜4.5⎟ ⎢E6 E5 E4 E3 E2 ⎥⎥ ⎜⎜
⎜1.7⎟ ⎢0 0.67 ⎟
⎜ ⎟ ⎢ E6 E5 E4 E3 ⎥ ⎜ ⎟
⎥ ⎝0.54 ⎠
⎜0.8⎟ ⎢0 0 E6 E5 E4 ⎥
⎜0.7⎟ ⎢0 0 0 E6 E5 ⎥
⎜ ⎟ ⎢
⎝0.5⎠ ⎣0 0 0 0 E6 ⎥⎦
As we can see, we have 6 unknowns and 10 equations, so the system is overde-
termined. An optimization in needed, giving the desired values of E(t):
t E (min) 4 5 6 7 8 9
E(t) 1.79 2.33 2.58 0.00 1.24 0.00
5
Cout
4 E(t)
Cin
Cin, Cout, E
0
0 2 4 6 8 10 12 14
Time (min)
44 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors
t (s) 0 1 2 3 4 5 6 7 8 9 10 11
C in 0 0 8 6 4 5 6 3 1 0 0 0
t (s) 0 1 2 3 4 5
E 0 0.05 0.5 0.35 0.1 0
Calculate the convolution vector and the corresponding time vector to the con-
voluted signal.
Solution
The GNU code for the resolution of this example would be the following:
2.4 Computer Program Using Matlab
® (Convolution) 45
for i=1:n
A(i:(m-1)+i,i)=Er; % Construction of the matrix
convolution
end
plot(tin,Cin,tr,Er,tconvolution,Cout)
legend('Cin','RTD','Cout')
xlabel('time(s)')
In the Matlab program, first of all, there are the inputs of the experimental
curves. The time increment is then calculated and following that the matrix A is
constructed. Finally, the convoluted curve is calculated, and the time vector cor-
responding to the convoluted curve and the graphs are constructed. The result is
shown in Figure 2.3.
t (s) 0 1 2 3 4 5 6 7 8 9 10 11
C in 0 0 8 6 4 5 6 3 1 0 0 0
In the CSTR, the residence time is known to be t = 5 s. What is the signal at the
exit of the reactor?
46 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors
0
0 5 10 15
Time (s)
Solution
In this case, the E (RTD) curve should be calculated for each value of the time
vector associated to C in , according to the equation:
( )
1 t
E = exp −
t t
In the program, the only necessary change is to introduce the following instead
of the experimental RTD curve:
tr=tin; % Time vector associated to E
Er=(1/5)*exp(-tr/5); % RTD distribution in the reactor
The result of this simple calculation is given in Figure 2.4. This is very illustrative
and we can “play” with the given parameters to see the effect in the convoluted
curve.
Example 2.5 Convolution of a Given Vector with an Ideal Reactor. Plug Flow
Reactor
Let us have the following data for the input signal to a plug flow reactor (PFR):
t (s) 0 1 2 3 4 5 6 7 8 9 10 11
C in 0 0 8 6 4 5 6 3 1 0 0 0
In the PFR, the residence time is known to be t= 5 s. What is the signal at the
exit of the reactor?
Solution
In this case, the E (RTD) curve is a Dirac delta centered at five seconds. This is
®
very easy to simulate in GNU Octave (and Matlab ) by just using a vector with
zeros at any position, but 1 at the corresponding position to five seconds (= t).
We will use logical variables for that purpose.
2.5 Computer Program Using MATLAB (Deconvolution) 47
0
0 2 4 6 8 10 12 14 16 18 20
Time (s)
0
0 2 4 6 8 10 12 14 16 18 20
Time (s)
As in the former example, in the GNU Octave code, the only necessary change
is to change the RTD distribution:
tr=tin; % Time vector associated to E
Er=(tr==5); % RTD distribution in the reactor
In this case, the double equal (==) is a logical operator answering 1 if the con-
dition is met, and zero in the other case. The result of this calculation is given in
Figure 2.5.
to give a good result. For the calculation, we need to solve the system of linear
equations given by Eq. (2.13) in such a way that the first step is to obtain the con-
volution matrix A. As mentioned before, the system is overdetermined, and the
minimization of a function is necessary. The O.F. to minimize can be programmed
in the following way:
In that function, the first step is to share the variables C in and C out with other
scripts, in order to be able to get the corresponding values of the vectors. Then,
the function calculates the corresponding convolution matrix to give the calcu-
lated differences between C in and the product (A*C out ). Finally, O.F. is obtained
for a given vector f 0 that will be optimized.
On the other hand, the main script for calculation has the following structure:
Cout = [0 0 0 0.4 4.3 6 5.1 4.8 5.3 4.1 2.1 0.6 0.1 0 0 0];
tout= [0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15];
Erd=fminsearch(@minimal,f0);
t0deconvolution=(tout(1)-tin(1));
tenddeconvolution=(tout(end)-tin(end));
inctdeconvolution=(tenddeconvolution-t0deconvolution)/(n-1);
tdeconvolution=t0deconvolution:inctdeconvolution:tenddeconvolution;
figure
plot(tin,Cin,tout,Cout,tdeconvolution,Erd)
legend('Cin','Cout','Er')
xlabel('time (s)')
2.5 Computer Program Using MATLAB (Deconvolution) 49
In the script, after giving values to C in and C out , the calculation of the time
increment and dimension of the vectors is done. The initial value for f 0 is taken
to be a vector with all values equal to 0.5. Finally, the optimization is done by
minimizing the O.F. using “fminsearch” and the graphs are constructed.
t (s) 0 1 2 3 4 5 6 7 8 9 10 11
C in 0 0 8 6 4 5 6 3 1 0 0 0
And at the exit of the reactor, the concentration follows the following data:
t (s) 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
C out 0 0 0 0.4 4.3 6 5.1 4.8 5.63 4.1 2.1 0.6 0.1 0 0 0
Calculate the RTD at the reactor by deconvolution of the vectors and the cor-
responding time vectors.
Solution
This is an unusual deconvolution case. The GNU code in this example would be
the following:
Cout = [0 0 0 0.4 4.3 6 5.1 4.8 5.3 4.1 2.1 0.6 0.1 0 0 0];
tout= [0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15];
t0deconvolution=(tout(1)-tin(1));
tenddeconvolution=(tout(end)-tin(end));
inctdeconvolution=(tenddeconvolution-t0deconvolution)/(v-1);
tdeconvolution=t0deconvolution:inctdeconvolution:tenddeconvolution;
figure
plot(tin,Cin,tout,Cout,tdeconvolution,Erd)
legend('Cin','Cout','Er')
xlabel('time (s)')
50 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors
0
0 5 10 15
Time (s)
As in the previous cases, first of all, there is the input of the experimental curves.
Then the time increment is calculated from C in or C out . The dimensions of the
vectors are read and an arbitrary first value of E(t) is given. The optimization is
done and the time vector corresponding to the deconvoluted curve, according to
the equations shown before, is calculated. The result is shown in Figure 2.6.
The practice with these techniques could be completed by calculating the con-
volution of the calculated (deconvoluted) E curve with the input signal C in , just
to check everything is correct and that we obtain again the C out vector.
Pulse Cout3
input
Reactor 1 Reactor 2 Reactor 3
Cout1 Cout2
Solution
We can calculate the area over the C f (t) curve:
∞ 5 8 11
Cf (t)dt = (t − 2)dt + 3dt + (11 − t)dt = 18
∫0 ∫2 ∫5 ∫8
In this way:
E = (t − 2)∕18 when 2 ≤ t ≤ 5
E = 3∕18 when 5 ≤ t ≤ 8
E = (11 − t)∕18 when 8 ≤ t ≤ 11
E=0 elsewhere
Also, we can calculate the following integral:
∞ 5 8 11
t ⋅ C(t)dt = t(t − 2)dt + 3t dt + (11 − t)t dt = 117
∫0 ∫2 ∫5 ∫8
And then:
117
tm = = 6.5 (units of time)
18
If this signal is fed to a CSTR, the balance in the reactor would be
⎧ (t − 2) t = [2, 5] ⎫
dC ⎪ ⎪
t + C = Cf = ⎨ 3 t = (5, 8) ⎬
dt ⎪(11 − t) t = [8, 11]⎪
⎩ ⎭
We can do the numerical calculation of the output signal replacing the incre-
ments by differences, as we study later. For example, in the first interval (super-
scripts “t” and “t + 1” refer to the actual position in time and the next one):
C t+1 − C t
t + C t = (t − 2), with C = 0 for t = 2
Δt
i.e.
(t − 2 − C t ) ⋅ Δt
C t+1 = C t +
t
Choosing a correct value of Δt, we can calculate the evolution of the concen-
tration. Doing a similar reasoning for the other two time intervals, the solution
is given in the following plot:
C input C output
3.5
3
2.5
Concentration
2
1.5
1
0.5
0
0 2 4 6 8 10 12 14
Time
2.6 Convolution of Signals in Reactors Connected in Series 53
Solution
(a) The first reactor of this system was designed to be a completely mixed flow
reactor. Let us see in a graph the form of the C(t) function obtained:
C(t)
10·a
t (min)
0 5 10
Note that the function for t > 10 is zero. As we can see, the response is some-
what similar to a CSTR, so the reactor was quite well scaled-up and con-
structed, at least in the aspect of the flow of solids. To calculate the E(t)
function, we need to calculate the area under the C(t) curve. This can be
done in several ways (for example, considering the area is that of a triangle),
but the general way is
∞ 10 ( )]10
t2
C(t)dt = a ⋅ (10 − t) ⋅ dt = a ⋅ 10t −
∫0 ∫0 2
( 2
) 0
10
= a ⋅ 10 ⋅ 10 − = 50a
2
So, for the RTD we have
C(t) a ⋅ (10 − t)
E1 (t) = ∞ = = 0.02 ⋅ (10 − t)
∫0 C(t)dt 50 ⋅ a
The mean residence time for this reactor is
∞ 10
tm = t ⋅ E1 (t) ⋅ dt = t ⋅ 0.02 ⋅ (10 − t) ⋅ dt
∫0 ∫0
( )]10
0.2t 2 t 3
1
= − 0.02 = = 0.̂
3
2 3 0 3
(b) The E(t) function of the solid being fed to a fluidized bed is that of a CSTR
(actually, the reactor closest to the perfectly mixed flow is the fluidized bed),
so we can write:
( )
1 t
E2 = exp −
t2 t2
54 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors
t (min) E1 E2
0 0.200 0.333
1 0.180 0.239
2 0.160 0.171
3 0.140 0.123
4 0.120 0.088
5 0.100 0.063
6 0.080 0.045
7 0.060 0.032
8 0.040 0.023
9 0.020 0.017
10 0.000 0.012
11 0.000 0.009
12 0.000 0.006
Let us use the Matlab program shown before, and now the code is given.
for i=1:n
A(i:(m-1)+i,i)=E2; % Construction of the matrix convolution
end
plot(t,E1,t,E2,tconvolution,Cout)
legend('E1','E2','Cout')
xlabel('time(s)')
0.25
0.2
0.15
0.1
0.05
0
0 5 10 15
Time (s)
Bibliography
Anderssen, A.S. and White, E.T. (1969). The analysis of residence time distribution
measurements using Laguerre functions. Canadian Journal of Chemical
Engineering 47: 288–295.
Levenspiel, O. (1999). Chemical Reaction Engineering, 3e. New York, NY: Wiley.
Mann, U. (2009). Principles of Chemical Reactor Analysis and Design. New York,
NY: Wiley.
Mecklenburg, J.C. and Hartland, S. (1975). The Theory of Backmixing. London:
Wiley.
57
3.1 Introduction
Transfer function is a very useful concept for studying nonideal chemical reac-
tors. This function indicates how the changes in the input to a system affect the
exit stream. It is valid for several systems (computers, PIPs, control systems, …)
and is based on the following definition:
Response to stimulus Y (s)
H(s) = =
Forcing function X(s)
This definition is given using the variable “s,” which is related to the Laplace
transform of the time-dependent functions in the reactor system.
Using a block diagram, we can say that:
Entrance to the process Process output
(forcing or stimulus) (response to stimulus)
System
H(s)
X(s) Y(s)
The transfer function also can be considered as the response of a system initially
inert to a pulse input signal (as we show later, the Laplace transform of a pulse is
the unity):
∞
H(s) = L{h(t)} = e−st ⋅ h(t)dt (3.2)
∫0
The exit is then calculated as
Y (s) = H(s) ⋅ X(s) (3.3)
and the response as a function of time can be calculated through the inverse of
the Laplace transform:
y(t) = L−1 {Y (s)} (3.4)
Linearity:
L{af (t) + bg(t)} = aL{f (t)} + bL{g(t)}
Product by scalar:
L{Kf (t)} = KL{f (t)} = Kf (s)
Derivatives:
′
L{f (t)} = sL{f (t)} − f (0)
′′ ′
L{f (t)} = s2 L{f (t)} − sf (0) − f (0)
L{f n (t)} = sn L{f (t)} − sn−1 f (0) − · · · − f n−1 (0)
The values at t = 0 are usually taken as zero; therefore:
L{f (t)} = sL{f (t)} = s ⋅ f (s)
′
Therefore:
f (∞) = lim[sf (s)]
s→0
Ramp function
Slope = K
Time
60 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems
Sinu function
Time
Suppose f (t) = sin(wt), where “w” is the frequency (rad/time) of the sinusoidal
wave. Applying the Laplace transform:
∞
L{sin(wt)} = e−st sin(wt)dt
∫0
and integrating:
w
L{sin(wt)} =
s2 + w2
3.3.1.3 Pulse Function
Pulse function
Time
showing that the Laplace transform of a pulse is the unity, as stated before.
Tracer
concentration Feed
Step input
C A, T
Product
0 Time
The result is, as we expected, the RTD of the CSTR, as a step input was used
as input signal.
(4) Pulse input, using Laplace transform:
This later case can also be treated using integral transform. Following
Eq. (3.23):
Input + Generation = Output + Accumulation
( )
dCout
M ⋅ 𝛿(t) + 0 = V + Q ⋅ Cout = (Cin (0) ⋅ V ) ⋅ 𝛿(t) (3.26)
dt
Taking Laplace transform:
V ⋅ s ⋅ Cout (s) + Q ⋅ Cout (s) = (Cin (0) ⋅ V ) ⋅ 1 (3.27)
3.4 Use of Laplace Transform in Chemical Reactor Characterization 65
Rearranging:
Cin (0) ⋅ V 1 1
Cout (s) = ⋅ = Cin (0) ⋅ t ⋅ (3.28)
Q (1 + t ⋅ s) (1 + t ⋅ s)
Doing the inverse:
[ ] ( )
1 1 t
L−1 = exp − (3.29)
(1 + t ⋅ s) t( ) t
Cin (0) t
Cout (t) = exp − (3.30)
t t
and, finally:
Cout ( )
Cin (0) 1 t
E(t) = ∞ Cout
= exp − (3.31)
∫0 C (0)
dt t t
in
Eq. (3.31) obviously represents the RTD in the mixed flow reactor. At this
point, we have to ask for the transfer function in the CSTR itself, i.e. what is
the function H(s) transforming the input signals X(s) into Y (s)? If we take a
look at the previous equation, the transfer function in a CSTR can be derived
from the E(t):
[ ( )]
1 t 1
E(s) = L exp − = (3.32)
t t 1 + ts
In this equation, u(t) represents the step input function. That, taking Laplace
transform, is
}
C(0, s) = Δu
s (3.36)
C(V , 0) = 0
Eq. (3.34) should also be transformed with respect to “t,” and we get
( )
𝛿C
Q⋅ + s ⋅ C(s) = 0
𝛿V
or, more commonly:
Q ⋅ dC = −s ⋅ C(s) ⋅ dV
The solution of this linear differential equation is
( )
Δu V
C (s) = exp −s (3.37)
s Q
and doing the inverse transform:
[ ( )] ( )
C 1 V V
F(t) = =L exp −s =u t− = u(t − t) (3.38)
Δu s Q Q
This function implies that the step input delays a time equivalent to t = V ∕Q
in the reactor, i.e. the residence time in the plug flow.
As we have done for the back mixed flow reactor, we will need a function H(s)
characteristic of the reactor. From the previous discussion, we see that, in the
PFR, the transfer function is
E(s) = L(𝛿(t − t)) = exp(−ts) (3.39)
Solution
From the definition of the transfer function, we can write
Y (s)
H1 (s) =
X(s)
Z(s)
H2 (s) =
Y (s)
where H 1 (s) and H 2 (s) are the transfer functions of the reactors. Easily we can get
( )
t 1
Z(s) = H1 (s) ⋅ H2 (s) ⋅ X(s) = X(s) ⋅ exp − ⋅
t1 1 + t2 ⋅ s
Note that we have substituted the transfer functions by the ones corresponding
to PFR (reactor 1) and CSTR (reactor 2). In this equation, the residence times t1
and t2 refer to the flow and volume of each reactor, i.e.:
𝛼V
t1 = = 𝛼t
Q
(1 − 𝛼)V
t2 = = (1 − 𝛼)t
Q
In this simple way, we see that a convolution of signals is transformed into a
product in the Laplace space. Let us continue with more complex systems.
System System
X1(s) H1(s) Y1(s) H2(s) Y2(s)
where Ej (s) is the L.T. of a pulse introduced in the individual system having a
volume V .
Feed
.....
CA1, T1
CAn, Tn
Product
CA2, T2
The overall impulse response then is obtained by inversion of the Laplace trans-
form:
⎧ ⎫ ( )N ⎧ ⎫
⎪ 1 ⎪ N ⎪ 1 ⎪
E(t) = L−1 ⎨ ( )N ⎬ = L−1 ⎨ ( )N ⎬
⎪ 1+ t s ⎪ t ⎪ s+ N ⎪
⎩ N ⎭ ⎩ t ⎭
( )N N−1
N t
= e−Nt∕t (3.48)
t (N − 1)!
This equation is equivalent to that found in Chapter 1 for the tanks-in-series
model, as it could not be otherwise.
βQ αV
S M
Q Q
(1–β)Q (1–α)V
Figure 3.4 Branches in parallel with the split “S“ and mix “M“ points.
70 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems
In this system, the upper branch contains a fraction 𝛼 of the total volume, and
the lower one (1 − 𝛼). Through the upper branch passes a flow 𝛽Q and the lower
one (1 − 𝛽)Q. We will assume that the volume of the split point (S) and the mix
point (M) are nil. Let E1 (s) and E2 (s) be the transfer functions (Laplace transform
of an impulse and therefore Laplace transform of the reactor RTD) of the upper
and lower branches, in the case that this branch had all the volume and all the
flow would pass through it. Since only volume 𝛼V is in the upper branch and
passes 𝛽Q through it, the Laplace transform of the response is
( )
𝛼
E1 s (3.49)
𝛽
Similarly, for the other branch:
( )
1−𝛼
E2 s (3.50)
1−𝛽
The overall response is obtained by the weighted average of the two, since at
the point M there is a mixture proportional to the flow:
( ) ( )
𝛼 1−𝛼
E(s) = 𝛽 E s + (1 − 𝛽)E2 s (3.51)
𝛽 1−𝛽
This can be generalized to M-branches in parallel, so that:
( )
∑M
𝛼j
E(s) = 𝛽 Ej s (3.52)
j=1
𝛽j
where 𝛽 j is the fraction of the total flow passing through branch “j” and 𝛼 j is the
volume fraction that this branch has.
This implies that if the dimensionless response of a system with volume V and
flow Q produces a certain flow pattern E(t), then the response to an impulse when
the same system when it is a subsystem (part of a network) containing a fraction
𝛼 of the volume and a passing fraction 𝛽 of the flow through it, is
( )
𝛼
Esub t (3.53)
𝛽
Using the L.T.s:
∞ ∞
1
L{E(t)} = e−st E(t)dt = e−st E𝜃 (𝜃)dt = E(s) (3.54)
∫0 t ∫0
∞ ( ) ( )
𝛼1 −st 𝛼 𝛼
L{Esub (t)} = e E t dt = Esub s (3.55)
𝛽 t ∫0 𝛽 𝛽
βQ
S αV M
Q Q
(1– β)Q
(1–α)V
where 𝛽 and (1 − 𝛽) are the fractions of the flow and 𝛼 and (1 − 𝛼) are the fractions
of the volume of each branch. In the same way:
E(s) = 𝛽E1 (s) + (1 − 𝛽)E2 (s) (3.57)
If the reactors of the system shown in the last figure were CSTRs:
1
E1 (s) = E2 (s) = j = 1, 2 (3.58)
1 + tj s
and then:
1 1
E(s) = 𝛽 𝛼
+ (1 − 𝛽) (3.59)
1 + 𝛽 ts 1 + 1−𝛼
1−𝛽
ts
This means that:
𝛽2 𝛽t (1 − 𝛽)2 (1−𝛽)t
− (1−𝛼)t
E(t) = L−1 {E(s)} = e− 𝛼t + e (3.60)
𝛼t (1 − 𝛼)t
RQ
F G2
RQ
(1–α)V
72 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems
is joined by recycle flow rate, RQ, so that the flow rate of (R + 1)Q flows through
the forward branch of the system that contains volume 𝛼V . At splitting point S,
RQ is recycled through the recycle branch of volume (1 − 𝛼)V when flow rate Q
leaves the system.
Let F, U, Y , and E be the signals of the system in the branches shown in
Figure 3.6. By definition, we can write
E
G1 =
Y
F
G2 =
E
G1 and G2 being the transfer functions of the systems 1 and 2, respectively. By
addition of the signals, we have the following at the point M:
Q ⋅ U + RQ ⋅ F = (1 + R)Q ⋅ Y
Eliminating the value of Q:
U + R ⋅ F = (1 + R) ⋅ Y
If we consider U = 1 (pulse input), and rearranging:
1+R⋅F 1 + R ⋅ G2 ⋅ E
E = G1 ⋅ Y = G1 = G1
1+R 1+R
G1
E=
1 + R − R ⋅ G1 ⋅ G2
which can be also written as
1 G1
E=
1 + R 1 − R G1 G2
R+1
To get the impulse response in the time domain, inversion of the equation from
the Laplace space is necessary. Often, it is interesting to expand the last equation
by binomial theorem, to get
∞ ( )n
1 ∑ R
E= G1 G1 n G2 n (3.61)
1 + R n=0 R + 1
and obtain the impulse response in the time domain by inverting each term in
the expression.
C(t)
A1 = 3A2 = 9A3 = 27A4
σ1 = 2/3
A1 A2 A3 A4
1 4 7 10
t (min)
Solution
(a) The mean residence time can be calculated, according to Chapter 1, by
∞
A1 A A
t ⋅ C(t)dt 1 ⋅ A1 + 4 ⋅ + 7 ⋅ 1 + 10 ⋅ 1
∫0 3 9 27
tm = ∞ =
A1 A1 A1
C(t)dt A1 + + +
∫0 3 9 27
94
= = 2.357 min
40
(b) The RTD shown is typical of a system in which recirculation is occurring.
Thus, we propose a model that is schematized in the figure. V p represents a
PFR. Let us call C 0 , C 1 , and C 2 , respectively, to the signals of the tracer fed,
entering the reactor model and in the recirculation current. Let us call R to
the ratio of flow rate in the recirculation.
Q C1 C2 Q
Vp
C0 (1 + R)Q C2
RQ C2
“G1 ” being the transfer function of the reactor vessel. We can clear, looking for
signal exit:
(C0 + RC2 ) ⋅ G1 = (1 + R) ⋅ C2
C0 G1 + RC2 G1 = C2 + RC2
C0 G1
C2 =
(1 + R) − RG1
Note that the transfer function of the whole system is given by
C2 G1
Gglobal = =
C0 (1 + R) − RG1
In this case, G1 = exp(−tp s). The residence time in the vessel can be related to the
average residence time in the system:
Vp t
tp = =
(1 + R) ⋅ Q 1 + R
From this equation, we can calculate the recirculation flow:
5
= 2.357 → R = 1.10
1+R
Taking into account these equations, finally:
( )
t
exp − 1+R s
Gglobal = ( )
t
(1 + R) − R exp − 1+R s
As one can expect, it is very difficult to do the anti-transform
( of this func-
)
t
tion. For a transfer function such as G1 = exp(−tp s) = exp − 1+R s we can do
the anti-transform, and it is
( )
t
E1 (t) = 𝛿 t −
1+R
The anti-transform of Gglobal gives a E(t) having a lot of peaks, as expected:
( ) ( ) ( )
t t t
E(t) = 𝛿 t − +𝛿 t−2 𝛿+ t−3 +…
1+R 1+R 1+R
Note that this reactor can also be modeled using a combination of the four PFRs
of different volumes (V 1 , V 2 , V 3 , and 1–V 1 –V 2 –V 3 ) connected in parallel, each
one with a flow rate Q1 , Q2 , Q3 , and (Q–Q1 –Q2 –Q3 ). The volume and flow rate
in each PFR can be calculated bearing in mind that the reactor residence time is
t i = V i /Qi and that the area of each peak is Ai = Qi /Qtotal . The result will give the
same E(t).
t (min) 6 8 10 12 14 16 18 20
c (mg/l) 4.3 3.2 2.7 2.1 1.7 1.4 1.0 0.8
Besides this, initially we observe rapid fluctuations for five seconds, measuring
a very high tracer concentration in this period of time.
(a) Propose a model for the reactor.
(b) If the reactor behaves as a perfect CSTR, what volume do we need for
XA = 0.75?
(c) What volume of a perfect CSTR do we need to get conversion that currently
is produced by the well-mixed region?
Solution
(a) When a sharp increase of tracer is observed, the first possibility to test is the
presence of a bypass. We cannot a priori discard the possibility of stagnancy
either. We can propose the following model based on the available data:
Vd
S M
Q Q
Vm
βQ
(1–β)Q
In the figure, V m represents the volume of the perfectly mixed region, and V d
the “dead” volume, not accessible to flow. Let V m = 𝛼⋅V and V = V m + V d .
The fraction of the flow in the bypass is (1 − 𝛽). In this way, the transfer func-
tion for the whole system will be given by the sum of the bypass and the stirred
tank:
𝛽
E(s) = (1 − 𝛽) +
1 + 𝛼𝛽 ts
αQo
CSTR
PFR PFR
βVr γVr
3.5 Complex Network of Ideal Reactors 77
A very deep stirred reactor is agitated with two agitators on a single axis. A plau-
sible model of the reactor is two PFRs partially in parallel and partially in series,
followed by a CSTR, as indicated in the scheme. Let 𝛼 be the fraction of the total
flow that is fed into the first PFR and let 𝛽 and 𝛾 be the fractions occupied by each
PFR. Determine the values of the average residence times in each of the three
reactors of the model, based on the total average time of residence. Calculate the
transfer function of the system to a pulse input.
Solution
The volume of the CSTR is (1 − 𝛽 − 𝛾)⋅V r and the relation of the residence times
with the total t would be
𝛽
t1 = t (flow 𝛼Q0 is flowing through this vessel)
𝛼
t2 = 𝛾 t (the total flow Q0 is flowing through this vessel)
G2 = exp(−t2 s)
1
G3 =
1 + t3 s
By definition:
C1 = G1 ⋅ C0
Doing a balance, we can write
Q0 C2 = G2 C1 𝛼Q0 + C0 (1 − 𝛼)Q0
and then:
[𝛼G1 + (1 − 𝛼)]C0 G2 G3 = C3
From that:
C3
= [𝛼G1 G2 G3 + (1 − 𝛼)G2 G3 ]
C0
And, finally:
[ ]
C3 𝛼 exp(−t1 s) exp(−t2 s) exp(−t2 s)
Gtotal = = + (1 − 𝛼)
C0 1 + t3 s 1 + t3 s
Solution
The input signal would have the following expression:
Cin
Input =
[u(t) − u(t − a)]
a
where u(t) is the step input and u(t − a) is the same input delayed “a” units. This
equation gives the squared input signal. We can calculate the Laplace transform:
Cin C C 1 C exp(−s ⋅ a)
L(u(t)) − in L(u(t − a)) = in − in
Input(s) =
a a a s a s
Cin
= (1 − exp(−s ⋅ a))
a⋅s
Bearing in mind that the CSTR has the transfer function shown in Eq. (3.32),
at the output we will have
Cin 1
Output(s) = Input(s) ⋅ E(s) = (1 − exp(−s ⋅ a)) ⋅
a⋅s 1 + ts
Rearranging:
Cin Cin ⋅ exp(−s ⋅ a)
Output(s) = −
a ⋅ s ⋅ (1 + ts) a ⋅ s ⋅ (1 + ts)
And doing the inverse of Laplace transform:
[ ( )] [ ( )]
C t C t−a
Output(t) = in 1 − exp − − in 1 − exp −
a t a t
With this equation, we can simulate the curve for the values of a = 0, 1, and 2:
a=0 a=1
2 2
1.8 1.8
1.6 1.6
1.4 Input/Cf 1.4 Input/Cf
1.2 1.2
C/Cf
C/Cf
1 Output/Cf 1 Output/Cf
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 1 2 3 4 0 1 2 3 4
t/tm t/tm
a=2
2
1.8
1.6
1.4 Input/Cf
1.2
C/Cf
1 Output/Cf
0.8
0.6
0.4
0.2
0
0 1 2 3 4
t/tm
3.5 Complex Network of Ideal Reactors 79
Solution
In the CSTR we have, for a non-reacting species:
dCout
t + Cout = Cin with Cout = 0 at t = 0
dt
In the interval 0 ≤ t ≤ 5:
dCout
t + Cout = 0.06t
dt
Doing the Laplace transform:
0.06
t ⋅ s ⋅ C(s) + C(s) =
s2
Solving for C(s):
0.06
s2
C(s) =
1+t⋅s
And doing the inverse of the Laplace transform:
[ ( )]
t t
Cout (t) = 0.06t − 1 + exp −
t t
This equation gives at t = 5 → C out = 0.11
In the interval when 5 < t ≤ 10:
dCout
t + Cout = 0.06 ⋅ (10 − t) with Cout = 0.11 at t = 5
dt
Let us call C out (t = 5) = C 5 . Transforming:
[ ]
10 1
t ⋅ s ⋅ (C(s) − C5 ) + C(s) = 0.06 + 2
s s
Solving for C(s) and doing the inverse of Laplace transform:
( ) [ ( )]
t−5 t−5
Cout (t) = 0.11 ⋅ exp − + 0.3 ⋅ 1 − exp −
[ t ( )] t
t−5 t−5
− 0.06 ⋅ t − 1 + exp −
t t
This gives C out (t = 10) = C 10 = 0.122.
80 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems
Reactor 3: C2
PFR
p·Q0 V3 = (1 – a – b)·V p·γ·Q0
(a) Determine the values of the average residence times in each of the three reac-
tors of the model, based on the average time of total residence.
(b) Taking into account the transfer function in the CSTR and PFR, find the
transfer function of the proposed reactor network.
Solution
(a) Let us calculate the average residence times of each reactor as a function of
the total residence time:
V1 aV a
t1 = = = t
(1 − p)Q0 (1 − p)Q0 (1 − p)
V2 bV b
t2 = = = t
(1 − p)Q0 + p(1 − 𝛾)Q0 (1 − p𝛾)Q0 (1 − p𝛾)
V (1 − a − b)V 1−a−b
t3 = 3 = = t
pQ0 pQ0 p
(b) For calculating the global transfer function, let us use the definitions of the
transfer functions in the three volumes:
C1 = G1 ⋅ C0
C2 = G3 ⋅ C0
C4 = G2 ⋅ C3
Let us write a balance at the mix point between reactors 1 and 2:
(1 − p)C1 + p(1 − 𝛾)C2 = (1 − p + p − p𝛾)C3
So:
(1 − p)C1 + p(1 − 𝛾)C2
= C3
(1 − p𝛾)
3.6 Transfer Function for the Dispersion Model 81
and then the values of r1 and r2 are distinct real numbers, with the following
solution:
√( )
2
u
De
± u
De
+ 4⋅s
De
r1,2 = (3.68)
2
The solution in the Laplace domain is then:
√( )
⎡⎛ u u
2
4⋅s
⎞ ⎤
⎢⎜ D + De
+ De
⎟ ⎥
C(s) = M1 ⋅ exp ⎢⎜ ⎟ t⎥
e
⎢⎜ 2 ⎟ ⎥
⎢⎜ ⎟ ⎥
⎣⎝ ⎠ ⎦
√( )
⎡⎛ u 2 ⎞ ⎤
⎢⎜ D − D
u
+ 4⋅s
D
⎟ ⎥
+ M2 ⋅ exp ⎢⎜ ⎟ t⎥
e e e
⎢⎜ 2 ⎟ ⎥
⎢⎜ ⎟ ⎥
⎣⎝ ⎠ ⎦
[( √ ) ]
u u 4 ⋅ De ⋅ s
= M1 ⋅ exp + 1+ t
2De 2De u2
[( √ ) ]
u u 4 ⋅ De ⋅ s
+ M2 ⋅ exp + 1− t (3.69)
2De 2De u2
The constants “M1 ” and “M2 ” should be calculated from the boundary conditions.
0.2 0.02
= −
(s + t2 ⋅ s ) (s + t2 ⋅ s3 )
2 2
3.6 Transfer Function for the Dispersion Model 83
0.300
E1
0.250 E2
Cout (Laplace)
E1, E2, Cout
0.200
0.150
0.100
0.050
0.000
0 2 4 6 8 10 12 14
Time (min)
(a) Calculate the mean residence time of the system, considering the same frac-
tion of the flow to each reactor.
(b) Deduce the form of the E(t) function, in the general case.
Reactor 1
αQ
Q Q
Reactor 2
βQ
γQ
Reactor 3
Solution
(a) Considering the system as a whole, we will write
Cout = 𝛼C1 + 𝛽C2 + 𝛾C3 = 𝛼 ⋅ exp(−5t) + 𝛽 ⋅ t exp(−t) + 𝛾 ⋅ t2 exp(−3t)
84 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems
We can then calculate the integrals to get the mean residence time:
∞ ∞ ∞
Cout dt = 𝛼 ⋅ exp(−5t)dt + 𝛽 ⋅ t ⋅ exp(−t)dt
∫0 ∫0 ∫0
∞
+𝛾 t2 ⋅ exp(−3t)dt = 0.2 ⋅ 𝛼 + 1 ⋅ 𝛽 + 0.075 ⋅ 𝛾
∫0
In the case proposed, where 𝛼=𝛽=𝛾=0.33, the area is 0.4243.
On the other hand:
∞ ∞ ∞
t ⋅ Cout dt = 𝛼 ⋅ t ⋅ exp(−5t)dt + 𝛽 ⋅ t 2 ⋅ exp(−t)dt
∫0 ∫0 ∫0
∞
+𝛾 t3 ⋅ exp(−3t)dt = 0.04 ⋅ 𝛼 + 2 ⋅ 𝛽 + 0.075 ⋅ 𝛾
∫0
Using the proposed values of 𝛼, 𝛽 and 𝛾, we get 0.7 as the value of this integral.
Then:
∞
∫0 t ⋅ Cout dt 0, 7
tm = ∞ = = 1.64 min
∫0 Cout dt 0.4243
(b) For deducing the E(t) function, we can use the definition of this curve when
a pulse input is introduced:
C(t) 𝛼 ⋅ exp(−5t) + 𝛽 ⋅ t exp(−t) + 𝛾 ⋅ t2 exp(−3t)
E(t) = =
∞
∫0 Cout dt 0.2 ⋅ 𝛼 + 1 ⋅ 𝛽 + 0.075 ⋅ 𝛾
Also, the system can be treated using transfer functions. In this case, the
transfer functions of each reactor would be
C1 (s) 1
G1 = = C1 (s) =
L{𝛿(t)} s+5
C2 (s) 1
G2 = = C2 (s) =
L{𝛿(t)} (s + 1)2
C3 (s) 2
G3 = = C3 (s) =
L{𝛿(t)} (s + 3)3
and then:
𝛼 𝛽 2𝛾
G = 𝛼G1 + 𝛽G2 + 𝛾G3 = + 2
+
(s + 5) (s + 1) (s + 3)3
That obviously will give the same E(t) combining the three signals. Never-
theless, if the system were to be connected in series, the convolution of the
three signals would be difficult without the aid of the transfer function, In
that case, we would have
E = E1 ⊗ E2 ⊗ E3
1 1 2
G = G1 ⋅ G 2 ⋅ G 3 = ⋅ ⋅
s + 5 (s + 1) (s + 3)3
2
Bibliography
Bourne, J. (1999). Turbulent Mixing and Chemical Reactions. New York, NY: Wiley.
Doraiswamy, L.K. (1984). Recents Advances in the Engineering Analysis of
Chemically Reacting Systems. New York, NY: Wiley.
Kayode Coker, A. (2001). Modelling of Chemical Kinetics and Reactor Design.
Elsevier.
Spiegel, M.R. (1965). Theory and Problems of Laplace Transforms, Schaum’s Outline
Series. New York, NY: McGraw-Hill.
Wallas, S.M. (1995). Chemical Reaction Engineering Handbook. Amsterdam:
Gordon and Breach.
87
4.1 Introduction
In mathematics an equation in partial derivatives (sometimes abbreviated as
PDE) is a relation between a function F of several independent variables x, y, z,
t, … and the partial derivatives of F with respect to those variables. The partial
differential equations are used in the mathematical formulation of processes of
physics and other sciences that are usually distributed in space and time. Typical
problems are the propagation of sound or heat, electrostatics, electrodynamics,
fluid dynamics, elasticity, quantum mechanics, and many others.
In chemical reaction engineering, this is especially important in the transient
regime during the start-up and shut down of chemical reactors (Chapter 5). The
method will permit to calculate, among others, the time to get a stationary regime
in each case.
Also important is the application of the methods to the differential equations
obtained in heterogeneous systems, where the reactant must be transferred from
one phase to another, where it finally reacts.
We work also with models for nonideal flow that are based in PDE, where the
initial values and boundary conditions must be proposed.
For this chapter, we are going to focus on the parabolic equations, although the
results obtained can be extended without problems to equations of elliptical or
hyperbolic type.
Examples of PDE related to chemical engineering are as follows:
Hyperbolic:
dC dC
−u = 0 (Advection equation)
dt dx
d2 C d2 C
2
− a 2 = 0 (wave equation)
dt dx
Elliptical:
𝜕2T 𝜕2T
+ 2 =0 (Laplace equation)
𝜕x2 𝜕y
Parabolic:
𝜕C 𝜕2C
− De 2 = 0 (diffusion equation)
𝜕t 𝜕x
(
) ( 2 )
𝛿f (xi ) 1 𝛿 f (xi )
f (xi+3 ) = f (xi ) + (3Δx) + (3Δx)2
𝛿x 2! 𝛿x2
( 3 )
1 𝛿 f (xi )
+ (3Δx)3 + · · · (4.9)
3! 𝛿x3
Multiplying by a, b, c and adding:
a ⋅ f (xi+1 ) + b ⋅ f (xi+2 ) + c ⋅ f (xi+3 ) = (a + b + c)f (xi )
+ f ′ (xi ) Δx (a + 2b + 3c)
1 ′′
+ f (xi )Δx2 (a + 4b + 9c)
2!
1 ′′′
+ f (xi )Δx3 (a + 8b + 27c) (4.10)
3!
Taking:
a + 2b + 3c = 0
a + 4b + 9c = 1
a + 8b + 27c = 0
We get a = − 52 ; b = 2; c = − 12
With what you get:
2f (xi ) − 5f (xi+1 ) + 4f (xi+2 ) − f (xi+3 )
′′
f (xi ) = (4.11)
Δx2
The first-order approximation for the first and second derivative of a function
is shown in Table 4.1. Table 4.2 shows the second-order approximation for the
first and second derivatives of a function.
Front differences:
f (xi+1 ) − f (xi )
f ′ (xi ) =
Δx
′′ f (xi ) − 2f (xi+1 ) + f (xi+2 )
f (xi ) =
Δx2
Rear differences:
f (xi ) − f (xi−1 )
f ′ (xi ) =
Δx
Front differences:
4f (xi+1 ) − 3f (xi ) − f (xi+2 )
f ′ (xi ) =
2Δx
′′ 2f (xi ) − 5f (xi+1 ) + 4f (xi+2 ) − f (xi+3 )
f (xi ) =
Δx2
Rear differences:
3f (xi ) − 4f (xi−1 ) + f (xi−2 )
f ′ (xi ) =
2Δx
′′ 2f (xi ) − 5f (xi−1 ) + 4f (xi−2 ) − f (xi−3 )
f (xi ) =
Δx2
Central differences:
f (xi+1 ) − f (xi−1 )
f ′ (xi ) =
2Δx
′′ 2f (xi+1 ) − 2f (xi ) + f (xi−1 )
f (xi ) =
Δx2
dCA
[nA ]x ⋅ S + rA ⋅ S ⋅ dx = [nA ]x+dx ⋅ S + S ⋅ dx ⋅ (4.12)
dt
nA is the molar flow density (kmol/(s m2 )) and rA the reaction rate, which will be
considered to be of first order:
dCA
([nA ]x − [nA ]x+dx ) ⋅ S − k CA ⋅ S ⋅ dx = S ⋅ dx ⋅ (4.13)
dt
92 4 Partial Differential Equations in Reactor Design
interface interface
Gas Solid catalyst Gas
CAs CAs
A A
Using dnA = [nA ]x + dx − [nA ]x for the small increment of this variable and
eliminating “S” we get
dCA
(−dnA ) − k CA ⋅ dx = dx ⋅ (4.14)
dt
This is
dnA dCA
− − kC A = (4.15)
dx dt
And, using Fick’s law for the molar flow density:
dCA
nA = −DAB ⋅
dx
So,
dnA d2 CA
= −DAB ⋅
dx dx2
We can rearrange (Eq. (4.14)) using partial derivatives:
𝜕 2 CA 𝜕CA
DAB ⋅ − kCA = (4.16)
𝜕x2 dt
The numerical methods of solving this type of problem calculate the value of the
dependent variable (in our example, the concentration) only in discrete points,
called nodes. In this way, the region existing between the limit points is divided
into nodal points. On the other hand, the time variable is discretized, resulting in
a two-dimensional network, in which an axis represents the spatial variable and
the other the temporary variable.
4.4 Approaching the Problem Using Finite Differences 93
Temporal
location ‘t’ Known initial condition
Known boundary condition
ti + 1
Δt
ti
ti – 1
xi – 1 xi xi + 1
Δx
Spatial location ‘i”
Rearranging:
Δt ⋅ DAB t Δt ⋅ DAB t Δt ⋅ DAB t
t+1
CAi = CAi+1 − 2 ⋅ CAi − CAi−1
Δx2 Δx2 Δx2
+ k ⋅ Δt ⋅ CAi
t t
+ CAi (4.21)
We finally get
( )
Δt ⋅ DAB t Δt ⋅ DAB Δt ⋅ DAB t
t+1
CAi = CAi+1 + 1 − k ⋅ Δt − 2 ⋅ t
CAi + CAi−1
Δx2 Δx2 Δx2
(4.22)
With Eq. (4.22), we can obtain the values of the concentration in the nodes in
time (t + Δt) depending on the value at time “t,” which, together with the initial
and boundary conditions, allows us to solve the problem.
t = 0 → CA = 0, for all points but the one contacting the other phase,
where CA = CAs
t = 0; CA = CAS ; for ∀x = 0
And the limit conditions will be given by (constant limit value):
CA = CAs ; for x = 0
[ t t
]
CAi − CAi−1 t t
= 0 i.e. CAN = CAN−1 (4.23)
Δx
x=L
Note that the derivative in the last point must be done in the form of a rear deriva-
tive, i.e. using the points “i − 1” and “i.” If central or front differences are used,
the information in the point “N + 1” is needed, and the system is not solvable.
We can also use a second-order approximation of the first derivative, and in
this case for the last node (rear differences) we will have
[ t+1 t t
]
3CAi − 4CAi−1 + CAi−2 t+1 1 t t
= 0 i.e. CAN = (4CAN−1 + CAN−2 )
2 ⋅ Δx 3
i=N
(4.24)
The equation represents the relationship between the concentrations at the
final node and the two nodal points closest to it. The extension of this result to
other types of insulating surfaces is simple.
Another interesting example may be the transport of a quantity (heat, mass, or
momentum) from the sine of a fluid to a solid surface, which can be described by
a transfer coefficient. For example, the heat flow from a surface to the fluid can
be represented by
qS = h(TS − T0 ) (4.25)
where T 0 is the temperature of the wall, T S the temperature of the sine of the
fluid, and “h” is the heat transfer coefficient. Applying the energy balance, the
equation is given by
]
dT
qs = −k ⋅ = h ⋅ (T − T0 ) (4.26)
dx S
Using a first-order approximation of the derivative:
]
Ti+1 − Ti
−k ⋅ = h ⋅ (TS − T0 ) (4.27)
Δx i=0
we can get
h ⋅ Δx
T1 = ⋅ (TS − T0 ) + T0
k
96 4 Partial Differential Equations in Reactor Design
h ⋅ Δx
T1 − ⋅ TS
k
T0 = ( )
h ⋅ Δx
1−
k
If the derivative were approximated to the second order:
]
4Ti+1 − 3Ti − Ti+2
−k ⋅ = h ⋅ (TS − T0 )
2Δx i=0
we would obtain
2 ⋅ h ⋅ Δx
−4T1 + T2 − ⋅ TS
k
T0 = ( )
2h ⋅ Δx
−3 −
k
4.4.3 Stability
In order to correctly solve a system of differential equations by the finite differ-
ence method, three conditions must be met: that the difference scheme be con-
sistent, that it be stable, and that it converge to the correct solution. If a scheme
is used in differences of those commented on earlier, the system will be consis-
tent. However, guaranteeing stability and convergence (that is, converging to the
solution of the differential equation and not to another solution) is not trivial.
There are different alternatives to establish a scheme in finite differences. How-
ever, as noted, the increments of the independent variables (position and time in
our example) cannot be chosen independently. Even taking very small values of
Δx – forcing a great precision – it can happen that the system is not stable.
One way to ensure stability is through the following theorem. Given a scheme
like the following:
Cit+1 = A ⋅ Ci+1
t
+ B ⋅ Cit + D ⋅ Ci−1
t
(4.28)
If A, B, D are all positive and also (A + B + D) ≤ 1, the scheme is stable; the errors
tend to decrease as we proceed with the iterations and converge to the solution
of the differential equation that we are trying to solve.
So, for our example on the gas diffusing in a solid catalyst:
( )
Δt ⋅ DAB Δt ⋅ DAB Δt ⋅ DAB
A= 2
B = 1 − k ⋅ Δt − 2 ⋅ 2
D=
Δx Δx Δx2
Studying the first condition, the coefficients A and D are always positive. The
coefficient B must, therefore, meet:
( )
Δt ⋅ DAB
1 − k ⋅ Δt − 2 ⋅ ≥0
Δx2
which leads us to
1
Δt ≤ (4.29)
2 ⋅ DAB
k+
Δx2
4.4 Approaching the Problem Using Finite Differences 97
In other words, you cannot choose the time and position increments
completely independently of each other. Usually, the increment of position is
fixed, and an increment of time that allows the system to be stable is chosen.
To check the second condition:
[ ( ) ]
Δt ⋅ DAB Δt ⋅ DAB Δt ⋅ DAB
+ 1 − k ⋅ Δt − 2 ⋅ + ≤1 (4.30)
Δx2 Δx2 Δx2
In this example, it is always fulfilled.
while t <50
t = t + inct;
Cn (1) = CAs; % BOUNDARY CONDITION SURFACE
Cn (N) = 1/3 * (4 * C (N-1) -C (N-2)); % BOUNDARY CONDITION CENTER
for i = 2: N-1
Cn (i) = A * C (i-1) + (1-k * inct-2 * A) * C (i) + A * C (i + 1);
end
C = Cn;
plot (x, C)
drawnow
end
In the first lines of the program, the variables of the system are defined, and
the number of intervals “N” is fixed. The corresponding x-increment (length) is
calculated by L/(N − 1), and the time increment is selected by multiplying a num-
ber between 0–1 (in this case 0.9, although this can be changed) by the condition
given in Eq. (4.29). If the time increment does not fulfill this condition, the system
will be unstable. This can be checked using 1.1 (or higher) instead of 0.9.
The program now defines the initial value of the vector “C” (concentration)
and the values to be used of the vector “x,” at time equal to zero. A loop is defined
98 4 Partial Differential Equations in Reactor Design
1 1
0.9 0.9
0.8 0.8
0.7 0.7
CA (mol/L)
CA (mol/L)
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
Distance (cm) Distance (cm)
Final time = 0,5 s Final time = 1 s
1 1
0.9 0.9
0.8 0.8
0.7 0.7
CA (mol/L)
CA (mol/L)
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
Distance (cm) Distance (cm)
Final time = 3 s Final time = 5 s
Figure 4.3 Graphical representation of the concentration of gas in the solid particle at
different times.
until time reaches a certain value (that can be, obviously, changed). In the loop, a
new vector of concentration “Cn” is calculated, corresponding to the value of the
concentration in a time incremented by a value of “inct.” Conditions for points at
the input (Cn(1)) and the end (Cn(N)) are defined, and the points going from 2 to
N − 1 are calculated by an equation corresponding to Eq. (4.22). Finally, the new
value of the vector concentration to be used as the last one calculated is changed,
by doing C = Cn, and the graph is plotted.
This program permits to visualize the evolution of the concentration profile
with time, until time reaches the fixed value.
Figure 4.3 shows some of the results obtained with this program at different
values of the final time.
Dp Exchange
between zones.
Rate constant ‘γ’
βdyn·S βstat·S
Flow rate = Q
metals. The system shows an unusual behavior, as the residence time distribution
(RTD) would depend on the width of the carbonaceous felt used.
To develop a model for the observed behavior, different nonideal flow mod-
els have been tested. Different authors consider the hydrodynamic mixing in a
porous medium with the help of a model containing a series of perfect mixers
with stagnant areas. The porous reactor showed a distribution with a single peak
in the RTD, but it can be represented as the sum of two distributions: a normal
distribution and an exponentially decaying one. The proposed model for the flow
inside this electrochemical reactor is represented in Figure 4.4.
In our case, the model will consider only one pathway, for simplicity. In that
pathway we will assume that the electrolyte flows following a dispersed plug-flow
model, and that a stagnant area exists (such as shown in Figure 4.4). In this way,
the electrolyte in the stagnant areas will be refreshed by the electrolyte flowing
in the region considered in an axially dispersed plug flow.
The electrolyte holdup in this zone, 𝛽 tot = m3 liquid/m3 column, is divided into
a dynamic holdup, 𝛽 dyn , and a static or stagnant holdup, 𝛽 stat (see Figure 4.4). The
rate of exchange between the static (s) and dynamic (d) zones is assumed to be
proportional to the concentration difference. An exchange coefficient, 𝛾 (s−1 ), is
defined for the transport:
moles exchanged
= 𝛾(cd − cs )
m3 liquid s
where cd and cs are the concentrations in the dynamic and static phase, respec-
tively. “𝛾” can be considered as the product of a mass transfer coefficient and the
specific interfacial area between the flowing and the stagnant zones, k L a (in units
of time−1 ).
The model equations will be deduced applying the mass balance equation to
the dynamic and the static zones over a thin slice perpendicular to the direction
of the flow. We will assume the density and other variables to be constant over
100 4 Partial Differential Equations in Reactor Design
( )
Δt ⋅ Dd Δt ⋅ ud
t+1
Cd,i = t
Cd,i+1 ⋅ −
Δz2 Δz
( )
2 ⋅ Δt ⋅ Dd Δt ⋅ ud 𝛽tot ⋅ Δt ⋅ 𝛾
+ Cd,i ⋅ 1 −
t
+ −
Δz2 Δz 𝛽dyn
( )
Δt ⋅ Dd 𝛽 ⋅ Δt ⋅ 𝛾
t
+ Cd,i−1 + tot ⋅ Cs,i
t
(4.36)
Δz 2 𝛽dyn
and renaming:
( )
′
Δt ⋅ Dd Δt ⋅ ud
A = −
Δz2 Δz
( )
′ 2 ⋅ Δt ⋅ Dd Δt ⋅ ud 𝛽tot ⋅ Δt ⋅ 𝛾
B = 1− + −
Δz2 Δz 𝛽dyn
( )
′ Δt ⋅ Dd
D =
Δz2
we get
𝛽dyn
= A ⋅ Cd,i+1
′
+ B ⋅ Cd,i
′
+ D ⋅ Cd,i−1
′
t+1
Cd,i t t t
+ B ⋅ Cs,i
t
(4.37)
𝛽stat
To study the stability of the equations, we must check for the stability condi-
tions, which, as mentioned before, are
A > 0 (from Eq. (4.34))
A′ > 0; B′ > 0; D′ > 0; (A′ + B′ + D′ ) ≤ 1 (from Eq. (4.36))
If the theorem is applied to Eq. (4.35), bearing in mind the terms that accom-
pany each value of C d and C s , we find that, if A > 0, then:
𝛽stat
Δt < ≡ Condition 1
𝛽tot ⋅ 𝛾
For the analysis of Eq. (4.36), we have that (A′ + B′ + D′ ) is always minor or
equal to one, and also always D′ > 0. Considering only A′ > 0 and B′ > 0, then:
Dd
Δz < ≡ Condition 2
ud
and also:
1 Δz
Δt < ( )=( )
2 ⋅ Dd ud 𝛽tot ⋅ 𝛾 2 ⋅ Dd 𝛽tot ⋅ 𝛾 ⋅ Δz
− + − 1 +
Δz2 Δz 𝛽dyn Δz 𝛽dyn
≡ Condition 3
So, for the system of differential equations to be stable, we will need to compare
the time increments given by conditions 1 and 3, and choose the lower one.
the exit provide what are usually termed the “Danckwerts’ boundary conditions”
(see Chapter 1):
( ) 𝜕C (0+ )
D A
CA0 = CA (0+ ) − (4.38)
u 𝜕z
where C A0 is the input concentration of species A and the point 0+ represents
the first differential point inside the reactor. If we apply Eq. (4.38) to our case, we
have
( ) Ct − Ct
t t
Dd d,1 A0
CA0 = Cd,1 − (4.39)
ud Δz
that rearranging results in:
t t
Cd,1 = CA0 (4.40)
That will constitute the boundary condition at the input for the problem. On the
other side, at the exit, we can write (following Eq. (1.68))
t t
Cd,N = Cd,N−1 (4.41)
For the calculation of the E curve (or the RTD curve), we will simulate an input
signal as a step experiment:
Concentration = 0 if t < 0
Concentration = 1 if t ≥ 0
These relations constitute the other boundary condition of the problem, at the
input of the reactor. This would give what is called the F curve, i.e. the integral
residence time curve. For calculating the E curve corresponding to the reactor,
the derivative of F will be taken.
In the following lines, a Matlab program is given to simulate a particular case
of the system shown in Figure 4.4, with a group of constants with a given value
(listed in the program).
%DEFINITIONS OF CONSTANTS:
Bdyn=0.38; % m ̂ 3/m ̂ 3
Btot=0.98; % m ̂ 3/m ̂ 3
Bstat=Btot-Bdyn; % m ̂ 3/m ̂ 3
gamma=0.1; % 1/s
Dd=0.01;% m ̂ 2/s
ud=1; % m/s
L=1; % m Total length
Az=L/(N-1); % Length-increment
% Stability conditions:
At1 = 0.5 * Bstat/Btot/gamma; % Following Condition 1
At2 = 0.5 * Az / (2*Dd/Az - 1 + Btot*gamma*Az/Bdyn);
% Following condition 3
4.5 Other Applications of PDE – Numerical Methods 103
A = 1+ Btot/Bstat*At*gamma;
B = Btot/Bstat*At*gamma;
Aprime = At*Dd/Az ̂ 2-At*ud/Az;
Bprime = 1-2*At*Dd/Az ̂ 2+At*ud/Az-Btot*At*gamma/Bdyn;
Dprime = At*Dd/Az ̂ 2;
t=0;
tend=2; % Ending time to be used, secs
% Initial values:
Cdo= 1; %mol/m3 (step experiment simulation)
Cd(1:N+1)=zeros(size(1:+1));
Cd(1) =Cdo;
Cs(1:N+1)=zeros(size(1:N+1)) ;
tv=[]; % All time values to be used
Cout=[]; % All concentration values
Z=0:Az:L; % Values of z-variable
while t<tend
t = t+At
Cdn(1)=Cdo; %Boundary condition at the input
Csn(N)=Cs(N-1); %Boundary condition at the exit
Cdn(N)=Cd(N-1); %Boundary condition at the exit
for i=2:N-1
Cdn(i) =Aprime*Cd(i+1) + Bprime*Cd(i) + Dprime*Cd(i-1)
+ B*Bdyn/Bstat*Cs(i);
Csn(i) = A*Cs(i) + B*Cd(i);
end
Cd=Cdn;
Cs=Csn;
tv=[tv t]; % Initially is void but it is filled with the
corresponding value
Cout=[Cout Cd(N)];% Initially is void but it is filled with the
corresponding value
end
plot(Z,Cd(1:N),Z,Cs(1:N))
legend('Cd','Cs')
xlabel('Z (m)')
ylabel('Cd (mol/m ̂ 3); Cs (mol/m ̂ 3)')
drawnow
% Calculation of the RTD
E=diff(Cout)/At ; %E(t) of the system (derivative exit concentra-
tion of the step-input run)
figure
plot(tv, [0 E])
xlabel('t(s)')
ylabel('E(t)')
E=[0 E];
% Calculation of the average time and variance:
tE= tv.*E ;
t_m=trapz(tv,tE)
var = trapz(tv,(((tv - tm). ̂ 2 ) .* E ))
The structure of the program is similar to those explained before. First, the
different variables of the system are fixed, the value of space intervals is fixed,
104 4 Partial Differential Equations in Reactor Design
and the corresponding space increment is calculated. Then, the minimum time
increment is calculated, by using conditions 2 and 3. The initial values of the con-
centrations and times are then defined. Note that a step tracer input is considered,
as a value of C = 1 is defined. Void vectors “tv” and “C out ” are defined, which will
be used to store the values of the time and concentration at the exit of the reac-
tor that will be used in the calculation. At this point, the time loop begins, where
we use two vectors Cdn and Csn to represent the concentration in dynamic and
stagnant zones at an incremented time. Finally, the new value of the vector con-
centration to be used as the last ones calculated are changed, by doing Cd = Cdn
and Cs = Csn. The concentration at the exit of the reactor is given by the last
numbers of these vectors and are stored in the C out variable.
In order to calculate the RTD vector of the reactor, at the final part of the pro-
gram a numerical derivation of the concentration found is done. This is because
the simulation is done using a step input for the tracer, so the E(t) would corre-
spond to the derivative of the obtained signal.
At the last part of the program, the mean residence time t m and the variance
of the RTD obtained is calculated. The graphs shown in Figure 4.5 are obtained
with different values of 𝛽 dyn , i.e. with different fractions of the reactor vessel with
a static behavior. As we can see, as the value of 𝛽 stat increases, the RTD presents
a long tail, due to the slow mass transfer to the dynamic phase.
3 3.5
2.5 3
2.5
2
2
E(t)
E(t)
1.5
1.5
1
1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
(a) t(s) (b) t(s)
Figure 4.5 RTD calculated in the reactor, for (a) 𝛽 dyn = 0.38 and (b) 𝛽 dyn = 0.68.
4.5 Other Applications of PDE – Numerical Methods 105
expression
𝜕C 𝜕 2 C 𝜕C
=M 2 −
𝜕t 𝜕x 𝜕x
where C is the concentration, t is dimensionless time, M is the diffusion module,
and x is a dimensionless length.
With contour condition at the point of departure, 𝜕C𝜕x
= 0 at x = 1
Using the explicit finite difference method, we can calculate the profile of con-
centrations that the tube would have after a dimensionless time unit for three
different values of M, for example, 0.01, 0.001, and 0.0001. We would like, in this
example, to represent the profiles obtained in a single figure with different colors.
Using the first-order approximation of the derivatives, we can write
Cit+1 − Cit t
Ci+1 − 2 ⋅ Cit + Ci−1
t t
Ci+1 − Cit
=M −
Δt Δx2 Δx
Rearranging to solve for the concentration at the time “t + 1,” we get
( ) ( )
MΔt Δt 2MΔt Δt MΔt t
Cit+1 = − C t
i+1 + 1 − − Cit + C
Δx2 Δx Δx2 Δx Δx2 i+1
From this equation we can calculate the stability conditions:
( )
MΔt Δt
− >0
Δx2 Δx
( )
2MΔt Δt
1− − >0
Δx2 Δx
where the first one is equivalent to:
Δx < M
and the second one:
⎛ ⎞
⎜ 1 ⎟
Δt < ⎜ ⎟
⎜ 2M + 1 ⎟
⎝ Δx 2 Δx ⎠
In the example, the initial condition of the system is clear, as all the tube is empty
of the substance (all C(x,0) = 0 except in the first increment that will be C 0 ).
The contour conditions are mentioned in the statement, and are
Ci=1 = C0 In the first interval
t
CN−1 − CNt t
= 0 In the last interval, i.e.CN−1 = CNt
Δx
The corresponding Matlab program is the following one, which is based on the
same structure as that of the previous examples.
incx = L / (N-1);
inct = 0.9 / (2 * M / incx ̂ 2-1 / incx); %PONER AQUI LO QUE SEA
EXPLICANDO
C (1) = exp (0);
C (2: N) = 0;
x = 0: incx: L;
r = M * inct / incx ̂ 2;
t = 0;
while t <1
t = t + inct;
Cnnew (1) = exp (- (t-0.4) ̂ 2 / 0.005);
Cnnew (N) = C (N-1); % Implies that dC / dx = 0 in i = N
for i = 2: N-1
Cnnew (i) = r * C (i + 1) + (1 + inct / incx-2 * r) * C (i)
+ (- inct / incx + r) * C (i-1);
end
C = Cnew;
plot (x, C)
axis ([0 0.1 0 1])
drawnow
end
Cb = C0 in this surface
y
z
dy
General flow
direction, u
y=w
1
y = 0 (center)
2w
dz
y=w
z=0 z = zfinal
𝜕C
= u ⋅ Cb ⋅ dy − D b dz
𝜕y
[ ]
𝜕Cb
Output = Input + d −D dz + u ⋅ Cb ⋅ dy
𝜕y
Generation = rb dV = kb Cb∝ dy dz
𝜕C
Accumulation = − b dy dz
𝜕t
Putting them together and dividing by “dy dz,” we get
𝜕 2 Cb 𝜕C 𝜕Cb
−D + u b + kb Cb∝ + =0
𝜕y2 𝜕z 𝜕t
Taking the steady state:
𝜕 2 Cb 𝜕C
−D + u b + kb Cb∝ = 0
𝜕y2 𝜕z
the boundary conditions are as follows:
• At the inlet: C(y,0) = C A0 (all the section
[ ] is at the same concentration)
𝜕C
• At the center, by symmetry: y = 0, 𝜕yb =0
y=0
• At (
the )wall, y = 0 and y = w, the rate of diffusion equals rate of reaction:
−D 𝜕C𝜕y
= kw Cw
y=0
n n n
𝜕 2 Cb Cb,m+1 − 2Cb,m + Cb,m−1
=
𝜕y2 Δy2
where subscript “m” refers to the position on the y-axis, and superscript “n” to
the position on the z-axis.
Considering the mole balance:
𝜕 2 Cb 𝜕C
−D + u b + kb Cb∝ = 0
𝜕y2 𝜕z
it is easy to get
n n n n+1 n
Cb,m+1 − 2Cb,m + Cb,m−1 Cb,m − Cb,m
n
−D +u + kb (Cb,m )∝ = 0
Δy2 Δz
Rearranging:
[ ]
n+1 Δz D n n n n ∝ n
Cb,m = (C − 2Cb,m + Cb,m−1 ) − kb (Cb,m ) + Cb,m
u Δy2 b,m+1
And finally, for first-order (𝛼 = 1) reaction:
( )
n+1 Δz ⋅ D n Δz ⋅ D Δz n Δz ⋅ D n
Cb,m = C + −2 − k + 1 Cb,m + C
u ⋅ Δy2 b,m+1 u ⋅ Δy2 Δy2 b u ⋅ Δy2 b,m−1
Renaming:
Δz ⋅ D
𝛾=
u ⋅ Δy2
we get
( )
Δz
n+1
Cb,m = 𝛾Cb,m+1
n
+ −2𝛾 − k + 1 n
Cb,m + 𝛾Cb,m−1
n
Δy2 b
( )
For the system to be stable, 𝛾 ≥ 0 (always fulfilled), and −2𝛾 − Δz
k
Δy2 b
+1
≥ 0, so:
√ ( )
2D
Δy ≥ Δz + kb
u
At the center (y = 0), we will have
n n
𝜕Cb Cb,center − Cb,center−1
= =0
𝜕z Δz
where subscripts refer to the center and the position just before the
center(“center-1”). And then:
n n
Cb,center = Cb,center−1
This is true for any position on the z-axis.
On the other hand, at the wall (y = w):
n n n
𝜕Cw Cb,wall − Cb,wall−1 kw Cb,wall
= =
𝜕y Δy D
4.5 Other Applications of PDE – Numerical Methods 109
We can use the later equations to program in Matlab the case required. The
following program can be used.
kb=0.0001; % mol/l⋅s
CA0=1; % mol/l
kw=0.000001; % l/molS
w=0.01; % m half of the total lenght
u=0.1; % m/s
D=1e-3; %m2/s
while z<zfinal
z=z+incz;
CAn(N)=CA(N)/(1+kw*incy/D); % Boundary condition at the walls
CAn(1)=CA(2); % Boundary condition in the center
for i=2:N-1
CAn(i)=A*CA(i+1)+B*CA(i)+A*CA(i-1);
end
CA=CAn;
CAv=[CAv CA']; % To save the CA vector in a matrix
end
z=0:incz:zfinal;
mesh(z,y,CAv)
xlabel('z')
ylabel('y')
zlabel('C_A')
110 4 Partial Differential Equations in Reactor Design
1
1
0.8
0.8
0.6
0.6
CA
0.4
CA
0.4
0.2 0.2
0 0
0.01 0.01
0.008 0.01
0.006 0.005 0.008
0.004 0.01 0.006
0.002 0.006 0.008 0.004
0.004 0.002
y 0 0 0.002 y 0 0
z z
(a) (b)
Figure 4.7 Concentration profile obtained with different values of the constants. (a)
kw = 10−6 l/(mol s), (b) kw = 10−4 l/(mol s).
The program plots a 3D figure that can be rotated representing half of the
reactor, where we can check the evolution of the concentration with positions.
Figure 4.7a shows the results.
In this figure, we can see that the input concentration of reactant “A” (in the
position z = 0) is quickly consumed in the center of the reactor (y = 0). Never-
theless, in the walls the reaction is being produced, but at a lower rate giving a
characteristic profile. If the kinetic constant of the catalytic reaction is increased,
for example, at k w = 1⋅10−3 l/(mol s), the reaction will also take place at the wall,
giving a concentration profile similar to that in Figure 4.7b.
Bibliography
Carrillo Ledesma, A., González Rosas, K.I., and Mendoza Bernal, O. (2019).
Introducción al Método de Diferencias Finitas & su Implementación
Computacional. http://132.248.182.159/acl/Textos/ (accessed 2 May 2019).
LeVeque, R.J. (2007). Finite Difference Methods for Ordinary and Partial Differential
Equations: Steady-State and Time-Dependent Problems. Society for Industrial
and Applied Mathematics.
Hjortso, M.A. and Wolenski, P. (1954). Linear Mathematical Models in Chemical
Engineering, 2e. World Scientific.
Shankar, P.M. (2016). Pedagogy of second order homogeneous differential
equations: a holistic approach using a Matlab workbook. Computer Applications
in Engineering Education 24 (1): 114–121.
Wallas, S.M. (1995). Chemical Reaction Engineering Handbook. Amsterdam:
Gordon and Breach.
111
5.1 Introduction
In this chapter, we discuss the behavior of ideal reactors during the start-up and
shutdown of the flow, i.e. the unsteady state behavior of such systems.
We will increase the complexity of the differential equations to be solved, and
in this way we will begin with a continuous stirred tank reactor (CSTR) working
in a dynamic regime. Later we discuss the plug flow reactor (PFR) without and
with dispersion, and finish the chapter discussing the existence of multiple states
in CSTR in which an exothermal reaction is produced.
CA, T
Product
®
For simulating this behavior, a simple Matlab program can be used. For
example, in case we want to simulate a CSTR reactor with t = 10 s and
C A0 = 1 mol/m3 , where k = 1 s−1 , the corresponding Matlab program can be that
shown in Program Listing 5.1.
The main result is given in Figure 5.2. As you can see, the concentration at
the exit varies with time, while the steady state is established. In the CSTRs, by
definition, the concentration is the same in all points of the reactor and equals
the exit concentration.
5.3 PFR Working in Dynamic Regime (No Dispersion) 113
1
0.9
0.8
0.7
CA (mol/m3)
0.6
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8
Time (s)
Figure 5.2 CSTR working in unsteady state regime. The concentration at the exit of the CSTR is
constant in steady-state regime, but varies when considering the start-up or shutdown.
input = nA
output = nA + dnA
In a differential volume
accumulation = dV · dCA/dt
generation = rA · dV
114 5 Unsteady State Regime Simulation in Reactor Design
incV=Vtotal/(N-1); % Increment V
inct=0.95*incV/(Qv0+k*incV); % Must be lesser than defined in Eq 5.11.
V=0:incV:Vtotal; % Values of V that Will be used in the calculation
5.4 PFR Working in Dynamic Regime (with Dispersion) 115
% Initial values:
tend=[];
CAend=[];
t=0;
CA=zeros(1,N);
CA(1)=CA0;
tfinal=1; % secs, Final time
A=1-k*inct-Qv0*inct/incV;
while t<tfinal
t=t+inct
CAn(1)=CA0; % Boundary condition for the first increment
CAn(N)=CA(N-1); % Boundary condition for the last increment.
Eq. 5.13
for i=2:N-1
CAn(i)=A*CA(i)+(Qv0*inct/incV)*CA(i-1); % PDE given in Eq. 5.10
end
CA=CAn;
CAend=[CAend CA(end)]; % Saving the concentration at the exit
tend=[tend t]; % Saving the values of time used
plot(V,CA)
xlabel('Volume (m ̂ 3)')
ylabel('C_A (mol/m ̂ 3)')
drawnow
end
figure
plot(tend,CAend)
xlabel('Time (s)')
ylabel('C_A at the exit (mol/m ̂ 3)')
In the program, the last value of the concentration at each time, i.e. the con-
centration at the exit, is saved in a vector (C end ) in order to see the evolution of
this concentration with time.
Some results are presented in Figure 5.4, using different values of final time in
order to see the evolution of the concentration profile.
1 1
0.9 0.8
0.6 0.2
0.5 0
0.4 –0.2
0.3 –0.4
0.2 –0.6
0.1 –0.8
0 –1
0 2 4 6 8 10 12 14 16 18 20 0 0.02 0.04 0.06 0.08 0.1 0.12
Volume (m3) Time (s)
Final time = 0.1 s
1 0.04
0.9 0.035
0.6 0.025
0.5 0.02
0.4 0.015
0.3
0.01
0.2
0.1 0.005
0 0
0 2 4 6 8 10 12 14 16 18 20 0 0.05 0.1 0.15 0.2 0.25
Volume (m3) Time (s)
Final time = 0.2 s
1 0.04
0.9 0.035
0.8
0.03
CA at the exit (mol/m3)
0.7
CA (mol/m3)
0.6 0.025
0.5 0.02
0.4 0.015
0.3
0.01
0.2
0.1 0.005
0 0
0 2 4 6 8 10 12 14 16 18 20 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Volume (m3) Time (s)
Final time = 0.3 s
Figure 5.4 PFR working in unsteady state regime (without dispersion) at different times.
In the previous equation, two terms are present, the first one corresponding to
the convection (general movement of the flow) and the second one corresponding
to the diffusion by the Fick’s law. Deriving and doing Q = (u⋅S), then:
𝜕nA 𝜕CA 𝜕 2 CA
=u⋅S⋅ − De ⋅ S ⋅ (5.17)
𝜕z 𝜕z 𝜕z2
and introducing in the mole balance:
𝜕 2 CA 𝜕CA 𝜕CA
De ⋅ −u⋅ + rA = (5.18)
𝜕z 2 𝜕z 𝜕t
5.4 PFR Working in Dynamic Regime (with Dispersion) 117
Obviously, these two conditions must fulfill, so the lower value of Δt should be
selected.
The procedure can be applied to different designs. We will apply the method
for the simulation of a 1 m length reactor where a feed of 0.01 mol/m3 of reactive
is fed, with a gas velocity of 0.01 m/s and values of k = 1.7⋅10−5 mol/(m3 s) and
De = 10−4 m2 /s. Program Listing 5.3 shows a possible program for Matlab.
incZ
De/u
disp('Something is wrong')
return % This condition must be fulfilled (Eq. 5.21)
end
inct=0.95/(-u/incZ+2*De/incZ ̂ 2+k); % Must be lower according
to Eq. 5.22.
alfa=De*inct/incZ ̂ 2
beta=1-2*De*inct/incZ ̂ 2+u*inct/incZ-k*inct
gamma=De*inct/incZ ̂ 2-u*inct/incZ
L=0:incZ:Ltotal; % Values of V to be used in the calculation
% Initial condition:
t=0;
CA=zeros(1,N);
CA(1)=CA0;
tfinal=100; % Secs, final time
alfa=De*inct/incZ ̂ 2;
beta=1-2*De*inct/incZ ̂ 2+u*inct/incZ-k*inct;
gamma=De*inct/incZ ̂ 2-u*inct/incZ;
while t<tfinal
t=t+inct
CAn(1)=CA0; % Boundary condition for the first increment
CAn(N)=CA(N-1); % Boundary condition for the last increment
for i=2:N-1
CAn(i)=alfa*CA(i-1)+beta*CA(i)+gamma*CA(i+1);
end
CA=CAn;
plot(L,CA)
drawnow
end
The program is similar to those used in Chapter 2 and in the previous section.
Figure 5.5 shows some interesting results.
0.01 0.01
0.009 0.009
Concentration (mol/m3)
Concentration (mol/m3)
0.008 0.008
0.007 0.007
0.006 0.006
0.005 0.005
0.004 0.004
0.003 0.003
0.002 0.002
0.001 0.001
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Distance from the input (m) Distance from the input (m)
Final time = 10 s Final time = 40 s
0.01 0.01
0.009 0.009
Concentration (mol/m3)
Concentration (mol/m3)
0.008 0.008
0.007 0.007
0.006 0.006
0.005 0.005
0.004 0.004
0.003 0.003
0.002 0.002
0.001 0.001
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Distance from the input (m) Distance from the input (m)
Final time = 80 s Final time = 100 s
Figure 5.5 PFR working in unsteady state regime (with dispersion) at different times.
Qc Tc, Qc
CA, T
Product
as steady conditions are usually applied. V represents the volume of the reactor,
which is constant.
In the same way, an energy balance at the reactor can be done:
dT
𝜌cp Q(T0 − T) − (ΔHr )(kCA ) ⋅ V = UA(T − Tc ) + 𝜌cp V
(5.25)
dt
In this case, the heat of reaction is ΔH r and the heat capacity of the reacting
fluid is C p . Energy balance at the cooling jacket can be written as
dTc
𝜌c cpC Qc (Tc0 − TC ) + UA(T − Tc ) = 𝜌c cpC Vc (5.26)
dt
120 5 Unsteady State Regime Simulation in Reactor Design
where V c is the volume of the fluid in the jacket that is also constant.
In many textbooks and manuals, the variation of C A , T, and T c with time are
not considered and only the steady state is studied. In that case, the balance
equations “loss” the term depending with time, in such a way that:
QCA0 − (kC A ) ⋅ V = QCA (5.27)
𝜌cp Q(T0 − T) − (ΔHr )(kCA ) ⋅ V = U ⋅ A ⋅ (T − Tc ) (5.28)
𝜌c cpc Qc (Tc0 − Tc ) + U ⋅ A ⋅ (T − Tc ) = 0 (5.29)
In this way, Eqs. (5.27)–(5.29) represent the usual CSTR at steady state. We can
solve these equations for C A and T c . This gives:
Q0 CA0
CA = (5.30)
Q0 + V ⋅ k
UA
Tc0 + ⋅T
𝜌c Cpc Qc
Tc = (5.31)
UA
1+
𝜌c Cpc Qc
We can define the heat generated (qG ) and the negative of the heat removed
(−qR ) as
qG = −ΔHr ⋅ (kC A ) ⋅ V (5.32)
−qR = −[𝜌cp Q0 (T − T0 ) − UA(T − Tc )] (5.33)
We can easily simulate the behavior of the reactor with a simple Matlab pro-
gram. For illustrating this situation, we will use the following constants:
Qv0=0.010; % m3/s
Qr=0.01;% m3/s
CA0=1; %mol/m3
V=1; % m3
VR=0.01; % m3
k0=7e11; % seg-1
E_R=10500; % K-1
R=8.314; %J/mol⋅K
U=40; % W/m2⋅K
A=2; % m2
TC0=300; % K
T0=300; % K
5.5 Multiple Steady States in CSTR with Exothermal Reaction 121
DHr=-10000; % J/mol
rCp=720; % J/K
rCpR=1000; % J/K
QR=[];
QG=[];
Tv=280:430; % Temperature range, in K
plot (Tv,QG,'x',Tv,QR)
xlabel ('Temperature (K)')
ylabel ('J/s')
legend ('Q_g_e_n_e_r_a_t_e_d','Q_r_e_m_o_v_e_d')
In the program, the variables are fixed to the mentioned values, and the tem-
perature of the reactor is varied from 280 to 430 K. The corresponding C A and T c
are calculated using the expressions (5.30) and (5.31). As a result of the program,
the following figure is obtained.
As we can see, three different solutions to Eqs. (5.27)–(5.29) are found, i.e. three
different steady states are possible. The exact values of the temperatures and com-
position are presented in Table 5.1.
In principle, the one giving a lower concentration of C A is the best, as it implies
a higher conversion.
But it is important to be aware of the stability of these three solutions, because
the solutions may not be stable, and then the reactor will work in an unintended
state. If we look only at the results in Figure 5.7, and the reactor is started at,
for example, 330 K (between intermediate and upper steady states), the heat
generated by the reaction is greater than the heat removed. This will cause the
temperature in the reactor to rise, and the rise will continue until the upper
140
Qgenerated
120 Qremoved
100
80
60
J/s
40
20
–20
–40
280 290 300 310 320 330 340 350 360 370 380
Temperature (K)
Figure 5.7 Three possible steady states in the exothermal cooled CSTR.
steady state is reached. It is easy to show, using similar arguments, that the upper
and lower states are stable, but it is not the intermediate. Let us show that this is
not completely true.
In order to simulate the dynamical behavior of the reactor, we will solve numer-
ically Eqs. (5.24)–(5.26). To do that, let us use the first-order approximations of
the derivatives:
Ct+1
A
− CtA
QCA0 − (kCtA ) ⋅ V = QCtA + V
Δt
(Eq.(5.34), from Eq.(5.24))
T t+1 − T t
𝜌cp Q(T0 − T t ) − (ΔHr )(kCtA ) ⋅ V = UA(T t − Tct ) + 𝜌cp V
Δt
(Eq.(5.35), from Eq.(5.25))
Tct+1 − Tct
𝜌c cpC Qc (Tc0 − Tct ) + UA(T t − Tct ) = 𝜌c cpC Vc
dt
(Eq.(5.36), from Eq.(5.26))
We must solve to find the calculated Ct+1
A
, T t + 1 , and Tct+1 :
Q(CA0 − CAt ) − (kCAt ) ⋅ V
CAt+1 = Δt + CAt (5.34)
V
𝜌cp Q(T0 − T t ) − (ΔHr )(kCtA ) ⋅ V − UA(T t − Tct )
T t+1 = + Tt (5.35)
𝜌Cp V
𝜌c cpC Qc (Tc0 − Tct ) + UA(T t − Tct )
Tct+1 = + Tct (5.36)
𝜌c cpC Vc
5.5 Multiple Steady States in CSTR with Exothermal Reaction 123
We can now simulate the behavior starting at the different steady states. In the
simulation, we will start at a steady state. Note that the values of concentration at
the input and temperatures are the same, as the system has not changed. The ini-
tial state of the system is the steady state, and the evolution with time is followed.
The Matlab program can be the following one.
while t<=tfinal
i=i+1;
t=t+inct;
k=k0*exp(-E_R/T(i-1));
CA(i)=inct*((Qv0*(CA0-CA(i-1))-k*CA(i-1)*V)/V)+CA(i-1);
T(i)=inct*(rCp*Qv0*(T0-T(i-1))-DHr*V*k*CA(i-1)-U*A*(T(i-1)
-TC(i-1)))/(rCp*V)+T(i-1);
TC(i)=inct*(rCpR*Qr*(TC0-TC(i-1))+U*A*(T(i-1)
-TC(i-1)))/(rCpR*VR)+TC(i-1);
end
plot(tv,T(1:i-1),'b-')
xlabel('Time (s)')
ylabel('Temperature in the reactor (K)')
The calculation gives the evolution of temperature seen in Figure 5.8a–c. “a”
corresponds to the upper steady state, “b” to the intermediate, and “c” to the lower
one. As can be seen, actually the upper and intermediate states are not stable at
all, and the temperature tends to reach the only stable state, which is the lower
one. Figure 5.8c represents this stable state, and, as can be checked, a very small
temperature change is observed. In this sense, the study of the stationary states is
useless, as it can give unreal steady states. Actually, instead of talking about three
124 5 Unsteady State Regime Simulation in Reactor Design
360 330
340 320
330 315
320 310
310 305
300 300
0 20 40 60 80 100 120 140 160 180 200 0 20 40 60 80 100 120 140 160 180 200
(a) Time (s) (b) Time (s)
304
304
303.5 303
303
302
302.5
301
302
301.5 300
301
299
300.5
298
300 0 20 40 60 80 100 120 140 160 180 200
0 20 40 60 80 100 120 140 160 180 200
Time (s)
(c) Time (s) (d)
Figure 5.8 Temperature evolution with time in the system starting at different conditions.
(a) Upper steady state. (b) Intermediate steady state. (c) Lower steady state. (d) Introducing a
perturbation in the lower steady state.
“steady states,” we would better off talking about three “different solutions of the
balances in stationary conditions.”
We can check for the robustness of the stationary state with another simulation.
Starting from the stationary state, we can introduce a small perturbation in one
of the input conditions, for example, T input = (T 0 − 5) in such a way that the feed
is 5 ∘ C cooler. In this case, the (−Qr ) curve will move to higher values (Figure 5.9)
and the new steady states will be somewhat different (states 1′ , 2′ , and 3′ in the
figure). The usual behavior is then that, if we start from steady states 1, 2, and 3,
the system will evolve to states 1′ , 2′ , and 3′ , respectively. Nevertheless, with the
data proposed in this example, as we have seen, states 2 and 3 are unstable and
all cases finish in the state 1′ .
Matlab programming of this case is quick and simple. The same program as that
in the previous case can be used, but the perturbation changes the temperature
of the feed:
1′ 2
Temperature
Bibliography
LeVeque, R.J. (2007). Finite Difference Methods for Ordinary and Partial Differential
Equations: Steady-State and Time-Dependent Problems. Society for Industrial
and Applied Mathematics.
Hjortso, M.A. and Wolenski, P. (1954). Linear Mathematical Models In Chemical
Engineering, 2e. World Scientific.
127
6.1 Introduction
A key problem in the development of a new process is scaling. A definition of scal-
ing could be “correct operation and start up a commercial size unit whose design
and operating procedures are partly based on experimentation and demonstra-
tion in the operation of a smaller system.”
The reactor is the nucleus of many chemical plants. It is also the part of the plant
where there are most scaling problems, particularly when the reaction involves
several phases, because all phenomena are not equally affected by the dimensions
of the plant.
Until around World War II, scaling was an empirical process. The common
way of doing this was to enlarge the process in small steps, as shown schemati-
cally in Figure 6.1 for a chemical reactor. Figure 6.1 also shows the approximate
production rate and the time intervals associated with the design/construction
and operation. Depending on the nature of the process, these time intervals may
differ greatly from those shown.
Initially, large safety margins were incorporated, often producing plants with a
much higher capacity than designed. Clearly, scaling in small steps is very expen-
sive and insecure, as long as a predictive model is not used. Today, however, the
scaling process uses the data of the laboratory and/or pilot plant, supplemented
by studies in models and mathematical studies, to determine the size and dimen-
sions of the industrial units.
Due to the different predictive power of different mathematical models within
the production plant, there are different equipment for which the maximum scale
factor, from which the operation is unreliable, is different. Table 6.1 shows this
maximum factor for several equipment.
Let us assume that we have a reactor in which 30% conversion is obtained.
We want to design a reactor with 90% conversion. Someone could think about
making the reactor three times larger, but we know that many reactions show a
rate that decreases with increasing conversion. Also, thermodynamic limitations
could exist, preventing higher conversion of, for example, 30%. In fact, a conver-
sion of 30% may be achievable in a three times smaller reactor. In many cases, the
scaling is complicated by the existence of lateral reactions, thermal effects, phase
separation, etc.
Relative production
Design
>1000
Operation
1000
100
1
0.1
0 2 4 6 8 10 12
Time (years)
Figure 6.1 Conventional scaling process. Source: Moulijn et al. (2013). Adapted with
permission from John Wiley & Sons.
Table 6.1 Maximum typical values of the scale factor of several equipment.
Reactors
Multitubular >10 000
Homogeneous plug flow and stirred tank >10 000
Columns of bubbles <1000
Fluidized bed 50–100
Separation processes
Distillation and rectification 1000–50 000
Absorption 1000–50 000
Extraction 500–1000
Drying 20–50
Crystallization 20–30
A scaling factor with a low level of risk can only be achieved in situations in
which there are known correlations or when an engineering approach is possi-
ble. Thus, the unit operations of distillation, absorption, and rectification can be
scaled without an intermediate step. The behavior of the gases and the liquid–gas
equilibrium can be explained well in physical terms, for which the calculations are
direct. For solids, the values of the maximum scale factor are much lower, due to
complications such as deposits and abrasion.
One of the common objectives during scaling is to convert a batch process
(tested in the laboratory) to continuous operation. In this sense, a familiar con-
cept in terms of scaling is to assume the residence time in the continuous reactor
identical to the reaction time of the batch reactor.
Moreover, if we have a continuous process with a flow rate Q1 that we want
to increase to a Q2 ≫ Q1 producing a material with the same characteristics, the
scale factor is S = Q2 /Q1 and volume of the system will be given by S = V 2 /V 1 so
6.2 Scaling the Batch Tank Stirred Reactor 129
Case A: Reaction
under control
Time
Figure 6.2 Temperature control in batch reactors manipulates the temperature of the
external fluid.
reaction temperature (T). When the exothermic reaction begins, the temperature
of the medium that transfers heat is decreased, often to the lowest possible level.
After the reaction is under control (case A), the temperature can be increased to
maintain the reaction mixture at the optimum temperature to achieve a reaction
rate sufficiently high. If a reaction is faster or very exothermic, it may be impossi-
ble to keep the temperature under control even with maximum cooling (case B).
Obviously, such an event is not desirable, and a thermal runaway is produced.
Thermal control is so important in many reactors that, usually, they have the
appearance of heat exchangers. The reaction rate and the equilibrium are greatly
affected by temperature, and therefore also are affected by side reactions, forma-
tion of other products, yields, selectivities, etc.
Many of the agitated reactors found in the chemical industry are of the
jacketed type. As a consequence, the area of heat transfer per unit volume
decreases with increasing the reactor volume. Generally, the heat transfer could
be improved with a higher agitation; however, usually, this does not produce a
dramatic increase in the heat transference.
The heat transfer can also be improved using a coolant with lower temperature,
but the use of a medium at very low temperature is expensive, and the compo-
nents of the reaction mixture could precipitate on the walls.
The solution to this problem is to separate the reaction volume and the heat
exchange area, i.e. using external heat exchangers or by introducing heating/
cooling coils in the reactor. Thus, the batch reactor may be scaled from laboratory
to industrial, simply based on the reaction time. In addition, if necessary, the rate
of heat production is lowered by reducing the concentrations of the reactants
or catalysts (the operation in semi-batch could be attractive, as discussed in
Section 6.2.2).
®
A valid Matlab program is presented here. Figure 6.3 shows the results of the
calculations for a small and large reactor.
132 6 Scaling and Stability of Chemical Reactors
CP (mol/m3)
CP (mol/m3)
350 1000 500
T (K)
T (K)
330 800 450 600
310 T 600 400
Tc 400
290 CP 400 350
200
270 200 300
250 0 250 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Time (s) Time (s)
Figure 6.3 Concentration and temperature profiles in the small and the large reactor.
% Initial conditions:
CA(1)=CA0;
CP(1)=CP0;
T(1)=T0;
t=0;
N=1000; % Number of time increments
tfinal=600; % sec, final time to be used
inct=tfinal/(N-1);
% Integration by finite increments:
for i=2:N
t(i)=t(i-1)+inct;
if t(i)<360
Tc=345;
6.2 Scaling the Batch Tank Stirred Reactor 133
else
Tc=295;
end
k1=k10*exp(-E1/(0.008314*T(i-1)));
k2=k20*exp(-E2/(0.008314*T(i-1)));
R1=-k1*CA(i-1);
R2=-k2*CP(i-1);
CA(i)=CA(i-1)+R1*inct;
CP(i)=CP(i-1)+(R2-R1)*inct;
T(i)=T(i-1)+((R1*DH1)+(R2*DH2)+(U*A_V*(Tc-T(i-
1))))*inct/(d*cp);
end
plot(t,CA,t,CP,t,T)
In the previous figure, we can check that the selectivity of the process (and
the resultant quantity of the desired compound for a total time of 10 min
[600 s]) in a large-scale reactor is significantly lower than that of a pilot plant
reactor. The reason is that the relationship between the area of exchange and the
reactor volume, in the large-scale reactor is much smaller than that in the
small reactor, which produces incorrect control of the temperature. Thus,
the value of the reaction rate at the second reaction increases, thus lowering the
selectivity.
It is possible to opt for a semi-batch operation when the temperature is diffi-
cult to control. Initially, only a small part of the reaction is charged, and when
the desired temperature is reached, more reactant is added over time. The heat
production rate can be controlled effectively by careful dosing so that a loss of
control temperature is prevented.
To simulate the semi-batch mode, we will assume the same conditions as in
the batch reactor. However, the initial concentration of A will be minor, and
the remaining amount A is added at a rate of Q0 (m3 /s) starting at time t d . It is
assumed for simplicity that the physical properties of the mixture are unchanged
during dosing. Equations (6.1)–(6.3) describe the course of the process before
the dosing of the additional amount of A begins. The equation of the molar
balance during the period t > t d was modified as
dCA Q
= r1 + o ⋅ (CA0 − CA ) (6.4)
dt V
where Q0 is the addition flow rate of A and V is the volume at each instant. We
will assume the following values:
Q0 = 1 m3 ∕h td = 100 s CA0 = 150 mol∕m3
Figure 6.4 shows the results of the calculations for a large semi-batch reac-
tor with A/V = 1.00 m−1 with an initial volume of 500 l with a concentration
of 1000 (mol A)/m3 . Table 6.3 compares the results obtained in the batch and
semi-batch reactors. A valid program for computing the conditions is given.
134 6 Scaling and Stability of Chemical Reactors
1000 400
Concentration (mol/m3) 900 CA
800 CP 380
Temperature (K)
700 360
600
500 340
400
300 320
200 300
100
0 280
0 100 200 300 400 500 600 700 0 100 200 300 400 500 600 700
(a) Time (s) (b) Time (s)
Figure 6.4 Concentration of reactive A (a), product P, and reactor temperature (b).
Table 6.3 Results in the batch and semibatch reactor after 600 seconds.
% Initial conditions:
CA(1)=1000; % Assuming initial condition in the reactor
CP(1)=CP0;
6.2 Scaling the Batch Tank Stirred Reactor 135
T(1)=T0;
V(1)=0.5; % Initial volume, m3
t=0;
N=10000; % Number of time increments
tfinal=600; % sec, final time to be used
tchange=360; % sec, time where Tc is changed
td=100; % sec, time when adding the input current
inct=tfinal/(N-1);
% Integration by finite increments:
for i=2:N
t(i)=t(i-1)+inct;
if t(i)<tchange
Tc=345;
else
Tc=295;
end
k1=k10*exp(-E1/(0.008314*T(i-1)));
k2=k20*exp(-E2/(0.008314*T(i-1)));
R1=-k1*CA(i-1);
R2=-k2*CP(i-1);
if t(i)<td
CA(i)=CA(i-1)+R1*inct;
CP(i)=CP(i-1)+(R2-R1)*inct;
T(i)=T(i-1)+((R1*DH1)+(R2*DH2)+(U*A_V*(Tc-T(i-
1))))*inct/(d*cp);
V(i)=V(1);
end
if t(i)>=td
CA(i)=CA(i-1)+(R1+Q0*(CA0-CA(i-1)))*inct;
CP(i)=CP(i-1)+(R2-R1-Q0*CP(i-1))*inct;
T(i)=T(i-1)+((R1*DH1)+(R2*DH2)+(U*A_V*(Tc-T(i-
1))))*inct/(d*cp);
V(i)=V(i-1)+Q0*inct; % Volume changing with time
end
end
figure(1)
plot(t,CA,t,CP,'-.')
legend('C_A', 'C_P')
ylabel('Concentration, mol/m∧ 3')
xlabel('Time, s')
figure(2)
plot(t,T)
ylabel('Temperature, K')
xlabel('Time, s')
136 6 Scaling and Stability of Chemical Reactors
The data presented show that even in the case of an incorrect cooling pro-
cess, the yield and selectivity in a large semi-batch reactor are better than those
obtained in a smaller batch reactor. With this example it is also clear that the use
of a model helps in the design and operation process of a reactor of this type.
To ensure that a piston flow is obtained in the tubular reactor, two criteria have
to be met:
𝜌uD
Re = > 10 000 and L∕D > 100 (6.8)
𝜇
The typical liquid velocities in the pipes oscillate between 2 and 5 m/s. In
this example, we choose a velocity (u) of 2.5 m/s, so the first criterion is met if
D > 0.004 m. By choosing a maximum residence time of 100 s, a reactor length
of 250 m results. Thus, the second criterion is satisfied for D < 2.5 m.
A Matlab program valid for this simulation is presented.
k2=k20*exp(-E2/(0.008314*T(i-1)));
R1=-k1*CA(i-1);
R2=-k2*CP(i-1);
CA(i)=CA(i-1)+R1*incL;
CP(i)=CP(i-1)+(R2-R1)*incL;
T(i)=T(i-1)+((R1*DH1)+(R2*DH2)+(U*A_V*(Tc-T(i-
1))))*incL/(u*d*cp);
end
plot(L,CA,L,CP,L,T)
Firstly, the PFR can operate in an isothermal or adiabatic mode, or with a tem-
perature profile along the reactor. Figure 6.5 shows the concentration and tem-
perature profiles in reactors of various diameters. When there is no heat transfer,
i.e. in an adiabatic reactor, the diameter is not important. The optimum reactor
length (for a higher selectivity) decreases as the diameter increases. The yield of
the product P also decreases when the diameter of the reactor increased, because
the cooling is less efficient. Table 6.4 resumes the results.
Figure 6.5 and Table 6.4 show that the larger the diameter of the reactor, the
more it resembles the adiabatic operation. Therefore, instead of increasing the
diameter of the reagent or in order to increase the production capacity, it is better
to use a large number of parallel tubes of smaller diameter, so that the temperature
control is improved.
As mentioned before, the isothermal operation is also possible, in principle, in
a tubular reactor. Figure 6.6 shows the concentration of P and the temperature of
the hot medium necessary to maintain the reactor at the desired temperature for
large and small diameter reactors.
Figure 6.6 shows that, in practice, the cooling temperature to achieve
the isothermal operation of the reactor would have to be unacceptably low
CP (mol/m3)
Temperature (K)
1000 450
(a) (b)
0.05 m Adiabatic
800
400 0.1 m
0.067 m
600
350 0.067 m
0.1 m
400
0.05 m
300
200
Adiabatic Tc
0 250
0 50 100 150 200 250 0 50 100 150 200 250
Distance from the inlet (m)
Figure 6.5 Concentration of P (a) and reactor temperature (b) in adiabatic and jacketed piston
flow reactors of different diameters.
6.3 Rapid Exothermic Reaction in a Tubular Reactor 139
Table 6.4 Optimum length, conversion and performance for tubular reactors.
Diameter (m) A/V (m−1 ) Optimal L (m) Conversion (% mol) Yield of P (% mol)
CP (mol/m3) Tc(K)
1000 400
(a) T = 330 K (b)
800 T = 345 K
T = 295 K D = 0.05 m
300
T = 295 K
600
T = 345 K
200 T = 295 K
400
100
200 T = 360 K D = 0.2 m T = 345 K
0 0
0 50 100 150 200 250 0 50 100 150 200 250
Distance from the inlet (m)
Figure 6.6 (a) Concentration of P as a function of the temperature of the reactor (T) in
isothermal tubular reactors (diameter not relevant). (b) Required temperature of the
refrigerant to maintain isothermal reactor at temperature for a large reactor and a small one.
(even impossible in large diameter reactors). The reason for needing this
extreme cooling in the first part of the reactor is that the reaction rate is very
high initially, with a large production of heat in the exothermic reaction.
This section tries to calculate the maximum temperature (T max ) that will be
achieved in a tubular reactor system where an exothermic reaction occurs. Con-
sider that a single reaction is occurring in a constant density system. The energy
balance in this case would be, according to Eq. (6.7):
dT A
u ⋅ 𝜌 ⋅ CP ⋅ = (−rA ) ⋅ ΔHr + U ⋅ (Tc − T) ⋅ (6.9)
dL V
The mole balance in the reactor tells us that:
dCA dXA
(−rA ) = −u ⋅ = u ⋅ CA0 ⋅
dL dL
and then:
dXA
dL = u ⋅ CA0 ⋅ (6.10)
(−rA )
where X A represents the molar conversion. Clearing dT from the first equation
and making use of the second, one can easily be reached:
A
CA0 ⋅ (−ΔHr ) U ⋅ (Tc − T) ⋅ V
dT = ⋅ dXA + ⋅ dL (6.11)
𝜌 ⋅ CP u ⋅ 𝜌 ⋅ CP
The term that accompanies dX A would be the temperature increase that would
occur if the reactor were adiabatic, so it is usually called ΔT ad :
CA0 ⋅ (−ΔHr )
ΔTad = (6.12)
𝜌 ⋅ CP
Therefore, we can write
A
dT U ⋅ CA0 ⋅ (Tc − T) ⋅ V
= ΔTad + (6.13)
dXA (−rA ) ⋅ 𝜌 ⋅ CP
From this equation, if T = T max , we would have dT = 0; therefore:
ΔTad ⋅ (−rA ) ⋅ 𝜌 ⋅ CP
Tmax = Th − A
(6.14)
U ⋅ CA0 ⋅ V
In this equation, (−rA ) should be calculated at the temperature and the molar
conversion obtained at the T max . These quantities are determined by the whole
history of the reaction mixture, so that Eq. (6.14) does not allow the direct calcu-
lation of T max . However, it is possible to give an approximate solution to Eq. (6.13)
using some approximations.
To do this, let us express Eq. (6.13) using dimensionless numbers:
A
U⋅ V
Nc =
kc ⋅ 𝜌 ⋅ CP
E ⋅ ΔTad
Nad = (6.15)
R ⋅ Tc2
and defining a dimensionless temperature difference:
E
Δ𝜐 = (T − Tc ) (6.16)
RTc2
6.3 Rapid Exothermic Reaction in a Tubular Reactor 141
At the reactor inlet, the conversion is zero and the difference in dimensionless
temperature will be
E
Δ𝜐0 = (T0 − Tc )
RTc2
Only when the inlet temperature (T 0 ) equals the temperature of the cooling
medium (T c ) will we have that Δ𝜐0 = 0.
The dimensionless number N ad takes into account the adiabatic temperature
increase, while N c is a measure of the cooling capacity of the medium. In these
numbers, k c represents the kinetic constant at the temperature of the cooling
medium. The number N c is very sensitive to changes in the cooling temperature,
as it contains a term exponentially dependent on T c , while N ad is not much tem-
perature dependent. The relationship between k c and the kinetic constant at any
temperature is
( ) ( ) ( )
E E E
k = k0 ⋅ exp − = k0 ⋅ exp − ⋅ exp (T − Tc )
RT RTc RTTc
( )
E
≈ kc ⋅ exp (T − Tc ) = kc ⋅ exp(Δ𝜐) (6.17)
RTc2
We can use this approach with first-order kinetics:
(−rA ) = kc ⋅ exp(Δ𝜐) ⋅ CA0 ⋅ (1 − XA ) (6.18)
Making use of the definitions (Eq. (6.15)) and of Eqs. (6.13) and (6.18):
( )
E ⋅ (Tc − T) ⋅ exp − RTE 2 ⋅ (T − Tc )
E dT
⋅ = Nad + Nc c
(6.19)
2
RTc dXA (1 − XA ) ⋅ RTc2
Finally, Eq. (6.19) is expressed as follows:
d(Δ𝜐) Δ𝜐 ⋅ exp(−Δ𝜐)
= Nad − Nc (6.20)
dXA (1 − XA )
Qualitative information about reactor stability can be obtained by examining
the Eq. (6.20), which we recall is a simplified form of the energy balance in a cooled
tubular reactor. For positive values of Δ𝜐, the slope d(Δ𝜐) dX
can never be greater
A
5
Nc/Nad = 2
4
Nad = 28
3
Δυ
2 Nad = 16
Nad = 26
1 Nad = 24
Nad = 8
0
0 0.2 0.4 0.6 0.8 1
XA, molar conversion
Figure 6.7 Temperature increase for exothermic reactions in a tubular reactor. Nc /Nad = 2 < e.
We see in the graph that above a certain value of N ad , (in Figure 6.7, between
26 and 28) the curve Δ𝜐 − X A completely loses the lower part observed at high
conversion values and the temperature increase is close to that obtained in the
adiabatic regime (in the graph it is not observed, but all the curves return to
Δ𝜐 = 0 for X A = 1).
Let us calculate the locus of the maxima of the paths given by Eq. (6.20). To do
this, we equal the expression (Eq. (6.20)) to zero and clear:
Nc
XA = 1 − ⋅ Δ𝜐max ⋅ exp(−Δ𝜐max ) (6.21)
Nad
For each value of conversion, Eq. (6.21) has either two solutions (one with
Δ𝜐max > 1 and another with Δ𝜐max < 1) or none (in the interval X A = 0–1).
Figure 6.7 shows (in red) the locus of the points represented by (6.21) in the case
N c /N ad = 2.
Figure 6.8 shows (in red) examples of loci given by Eq. (6.21), for N c /N ad = 5,
and 2, with arrows (blue) representing the value of the slope d(Δ𝜐)/dX A given by
Eq. (6.20). The value of this slope is always positive to the right of the locus given
by Eq. (6.21), and negative to the left. In Figure 6.8, the position of the “mm”
points is also indicated.
When N c /N ad > e, the place of the maxima intersects the axis X A = 0 (no
conversion) in two points, for example, the values 2.52 and 0.26 for N c /N ad = 5,
as shown in Figure 6.8. In this case, if a reaction begins with Δ𝜐0 < 0.26, it
will follow a path similar to that shown in Figure 6.9 (blue line): the value
of Δ𝜐 will not increase very much (in fact, Δ𝜐 will always be less than 0.26)
because it must decrease as soon as the solid line is passed, which marks the
maximums in the derivative, and the reaction can be easily controlled. The same
happens if 0.26 < Δ𝜐0 < 2.52 (black line), since the temperature must decrease
monotonically. However, if Δ𝜐0 > 2.52 (see green line), the temperature initially
6.3 Rapid Exothermic Reaction in a Tubular Reactor 143
Δυ
2.52
2 Nc/Nad = 5
1 0.26
0
0 0.2 0.4 0.6 0.8 1
XA, molar conversion
5 5
4 4
3 3
Δυ
Δυ
2 2
Maxima of the curve of Nc/Nad = 2
1 maximums “mm” 1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
XA, molar conversion XA, molar conversion
Figure 6.8 Geometric place of the maximum temperature differences for several values
of Nc /Nad .
increases and will continue to grow until the maximum area is crossed. Clearly,
the point Δ𝜐0 = 2.52 is a stability limit (for the case N c /N ad = 5).
As shown in Figure 6.8, by decreasing the value of N c /N ad the curve present-
ing Δ𝜐max is closer to the right, so that the possibility that the reactor temper-
ature runaway increases. If N c /N ad = e, the place of maxima is tangent to the
vertical axis.
Let us analyze carefully the case N c /N ad = 2 (Figure 6.7). We see that the tem-
perature shoots up when the path crosses the line after its maximum (after the
point we call “mm”). Thus, the path that passes through “mm” can be considered
as critical, in the sense that at the moment that some point of the reactor is located
above this path, there are many possibilities for the temperature in the reactor to
be triggered. For example, if in the case of Figure 6.7 the value of the tempera-
ture at the inlet makes Δ𝜐0 = 2, probably the reactor to be thermally unstable and
temperature increase indefinitely.
Let us see what equation defines the loci of the maximums of the curve of max-
imum values (points “mm”). For this, we derive Eq. (6.21) and equate to zero:
dXA N
= − c [exp(−Δ𝜐mm ) − Δ𝜐mm ⋅ exp(−Δ𝜐mm )] (6.22)
dΔ𝜐max Nad
The solution to this equation is Δ𝜐mm = 1 (it can be perfectly seen in
Figures 6.7–6.9), no matter the value of N c /N ad , or what is the same:
E RTc2
Δ𝜐mm = 1 = (Tmm − Tc ) ⇒ Tmm = Tc + (6.23)
RTc2 E
144 6 Scaling and Stability of Chemical Reactors
2 Nc/Nad = 5
Δυ0 = 2
1
0.26
0
Δυ0 = 0 0 0.2 0.4 0.6 0.8 1
XA, molar conversion
5
Nc/Nad = 2
4
Nad = 28
3
Δυ
Δυ0 = 2 2 Nad = 16
Nad = 26
1
Nad = 24
Nad = 8
0
0 0.2 0.4 0.6 0.8 1
Δυ0 = 0 XA, molar conversion
what conditions at the reactor inlet will produce those situations. Van Welsenare
and Froment empirically obtained a second criterion for thermal stability, care-
fully studying the trajectory of the curves T–X A , valid only in the case T c = T0
(i.e. Δ𝜐0 = 0):
√( )
Nad ΔTad Nc N
= ≤1+ + c (6.27)
Δ𝜐mm Tmm − Tc e e
The condition (Eq. (6.27)) is helpful when trying to determine the maximum
allowable concentrations of reagents as C A0 = 𝜌C P ΔT ad /(−ΔH r ). The applica-
bility of the method of van Welsenare and Froment is limited, since the second
stability criterion is derived with the condition T 0 = T c . In practice, for very
exothermic reaction at elevated temperature, the feed is preheated by the refrig-
erant, so T 0 ≅ T c .
Equation (6.27) is also useful to estimate the value of minimum N c necessary
(or what is the same, the cooling conditions needed by the system) to keep the
system stable given the conditions of entry.
Solution
For the batch reactor, the material and energy balances have the following form:
dCA
= r1 = −k1 CA
dt
dCP
= r2 − r1 = −k2 CP − (−k1 CA )
dt
dT A
𝜌CP = (−r1 )(ΔH1 ) + (−r2 )(ΔH2 ) + U(Tc − T)
dt V
Solving previous balances knowing that dT/dt = 0, T c is cleared:
A
0 = k1 CA0 (1 − XA )(ΔH1 ) − k2 CA0 XA (ΔH2 ) + U(Tc − T)
V
Knowing that the temperature must be T=300 K, we can calculate k 1 and k 2 :
( )
20
k1 = 0.5 ⋅ exp − = 1.56 ⋅ 10−4 s−1
0.008 314 ⋅ 300
146 6 Scaling and Stability of Chemical Reactors
( )
60
k2 = 25 000 ⋅ exp − = 8.9255 ⋅ 10−7 s−1
0.008 314 ⋅ 300
and, finally:
−1.56 ⋅ 10−4 ⋅ s−1 1000 mol∕m3 (1 − XA )(−300 000 J∕mol)
+8.9255 ⋅ 10−7 s−1 1000 mol∕m3 ⋅ XA (−250 000 J∕mol)
Tc = A
+ 300 K
500 J∕(s m2 K) ⋅ V
m−1
and the following profiles are obtained, using a spreadsheet:
Variation of cooling temperature needed
330
310
290
Temperature Tc (K)
In the previous figure, it can be seen that in the large-scale reactor the refrig-
eration to be supplied is greater than that in the smaller reactor and that in the
large reactor the A/V ratio is lower.
Solution
(a) Applying the second Welsenare criterion, the T mm that makes the system is
thermally unstable can be obtained by
√( )
RTmm ⎡ 4RTc ⎤⎥
≤ 0.5 ⎢1 − 1−
E ⎢ E ⎥
⎣ ⎦
Using the known data:
[ √( ]
)
Tmm 4 ⋅ 300
≤ 0.5 1 − 1−
15 300 15 300
We obtain T mm = 596.2 K. The maximum concentration can be obtained
from
𝜌CP ⋅ ΔTad
CA0 =
(−ΔHr )
where ΔT ad can be obtained from the relationship:
√( )
ΔTad Nc N
≤1+ + c
Tmm − Tc e e
And the value of N c is
A
U⋅ V
Nc =
kc ⋅ 𝜌 ⋅ CP
In a tubular reactor, A/V = 2/R, which equals 16 m−1 in the present example.
The value of N c is then 728.9, and a value of ΔT ad = 6633 K is obtained and
therefore a C A0 = 4.8 mol/m3 .
(b) First, the new T c is calculated from the following equation:
E
(Tmm − Tc ) ≤ 1 i.e.Tc = 758.2 K
RTc2
The ΔT ad is obtained from a C A0 = 4.8 mol/m3
𝜌CP ⋅ ΔTad
CA0 = i.e.ΔTad = 6633.3 K
(−ΔHr )
Next, N c is calculated from
√( )
ΔTad Nc N
≤1+ + c
Tmm − Tc e e
Finally, U can be obtained from
A
U⋅ V
Nc =
kc ⋅ 𝜌 ⋅ CP
Therefore, for T mm = 800 K, a value of N c = 4.36 is obtained, what implies
a U = 79.4 W/(m2 K) and for T mm = 900 K we obtain N c = 4.17 and
U = 99.5 W/(m2 K).
148 6 Scaling and Stability of Chemical Reactors
Solution
(a)–(d) In this system, Eq. (6.20) can be applied:
d(Δ𝜐) Δ𝜐 ⋅ exp(−Δ𝜐)
= Nad − Nc (6.20)
dXA (1 − XA )
Using the finite differences method, we can write
( )
Δ𝜐 ⋅ exp(−Δ𝜐 t
)
Δ𝜐t+1 = Δ𝜐t + Nad − Nc ⋅ ΔXA
(1 − XAt )
For calculating the requested curves, we can use the following Matlab program.
while XA<1
i=i+1;
Nad=11400*dTad/Tc∧ 2;
kc=(3.94*10∧ 12)*exp(-11400/Tc);
Nc=0.2/kc;
dv(i)=dv(i-1)+(Nad-Nc*(dv(i-1)*exp(-dv(i-1))/(1-
XA)))*incXA;
T(i)=((dv(i)*Tc∧ 2)/11400)+Tc;
XA=XA+incXA;
end
XAv=0:incXA:1;
hold on
figure(1)
plot(XAv,T). % Temperature-conversion
xlabel('Conversion')
ylabel('Temperature (K)')
figure (2)
plot(XAv,dv). % dv-conversion
xlabel ('Conversion')
ylabel('dv')
DVmax=[DVmax max(dv)]; % Vector saving the maximum
value of dv for each value of Tc
NcdivNad=[NcdivNad Nc/Nad]; % Values of Nc/Nad
XAmaxima=[XAmaxima 1-Nc/Nad*max(dv)*exp(-max(dv))];
% Values of XA corresponding to max(dv)
end
figure (3)
plot(XAmaxima,DVmax,'+')
xlabel('Conversion')
ylabel('DVmax')
figure (4)
plot(Thv, NcdivNad)
xlabel('Cooling temperature (K)')
ylabel('Nc/Nad')
figure (5)
plot(XAmaxima, DVmax, 'or')
xlabel ('conversion')
ylabel('DVmax')
150 6 Scaling and Stability of Chemical Reactors
480 14
460 12
440 10
Temperature (K)
420 8
400 6
Δυ
380 4
360 2
340 0
320 –2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) Conversion (b) Conversion
14 2.2
12 2
1.8
10
1.6
8
Nc/Nad
1.4
Δυmax
6 1.2
1
4
0.8
2 0.6
0 0.4
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 335 340 345 350
(c) Conversion (d) Cooling temperature (K)
14 15
12
10 10
8
Δυmax
Δυ
6
5
4
0 0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(e) Conversion (f) Conversion
(e) If T c = 335 K and diameter is four times higher, then we will have that rela-
tion of cooling area to volume “A/V ” will decrease four times. In this situation,
N c = 0.2/(4⋅k c ).
The Matlab program is very similar, using only T c = 335 K. The corresponding
plot can be observed in Figure 6.10f.
(f ) The corresponding Matlab is presented, and the results are shown in the
figure.
In the figure, the arrows represent the value of the slope dΔ𝜐 dXA
. As can be
observed, if the reaction begins with Δ𝜐0 < 1, the value of Δ𝜐 will not increase
much, never exceeding the value of 1 and will decrease following the red line,
that is, the reaction is easily controllable. The point Δ𝜐 = 1 is the stability
limit, and if the reaction starts with higher values of Δ𝜐, it will be out of
control. This point is known as point “mm” and is the maximum of the curve of
maximums.
If we are inside the curve (to the right), it is observed that the arrows indicate
a decrease in the value of Δ𝜐, that is, if the reaction is between the red lines, the
value of Δ𝜐 will decrease, following the trajectory of the lower red line.
dta=ones(length(ta)).*0.1;
dv=(Nad-Nc.*V.*exp(-V)./(1-Ta)).*dta;
quiver(Ta,V,dta,dv);
axis([0,1,0,5])
xlabel('Conversion A')
ylabel('Dv')
text(0.02,1.2, 'Maximum of the curve')
text(0.02,1,'of maxima')
title('Nc/Nad=e')
zoom
hold on
% Curve of maxima
v=linspace(0,5,100);
ta=linspace(0,0.8,100);
tta=-Nc/Nad.*v.*exp(-v)+1;
plot(tta,v,'r')
hold off
152 6 Scaling and Stability of Chemical Reactors
Nc/Nad = e
5
4.5
4
3.5
3
2.5
Δυ
2
1.5
Maximum of the curve
1 of maxima
0.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Conversion A
Solution
(a) Let us focus on the heat balance in a batch reactor:
dT A
𝜌 ⋅ CP ⋅ = (−r1 ) ⋅ ΔH1 + U ⋅ (Tc − T) ⋅
dt V
That is equivalent to Eq. (6.9) representing the balance in a tubular reactor,
substituting dt by dz/u.
6.3 Rapid Exothermic Reaction in a Tubular Reactor 153
end
axis([0 1 0 12])
legend('Situation 1','Situation 2')
xlabel('X_A')
ylabel('Dimensionless temperature difference')
10 Situation 1
Situation 2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
XA
Nc
In the limit Nad
= e = 188.3∕(CA0 ⋅ A) we can deduct that the following must
be fulfilled:
CA0 = 69.27∕A
For example, if the total area of heat transfer is A = 1 m2 , the maximum ini-
tial concentration will be 69.27 mol/l. If the initial concentration is higher,
the adiabatic temperature increment will be higher than the limit with the
given heat transfer characteristics. The result is somewhat surprising because
a higher area will permit a lower concentration, this is because the total heat
capacity is maintained constant, which is not a usual situation.
Bibliography
7.1 Introduction
This chapter tries to give an overview of the operation of chemical reactors
in an unsteady regime, which is usually known by its acronym: FUSO (forced
unsteady state operation). The four most used forms in the practice of carrying
out an operation in an unsteady regime are analyzed: periodic variation of the
input conditions, periodic flow inversion, operation with variable volume, and
use of oscillating pressure. We review how to carry out this operation, giving
their advantages and disadvantages.
It will be concluded that FUSO is not less efficient than the equivalent operation
in steady regime and, on many occasions, its practical interest can be consid-
erable. It is true that some of the possibilities discussed in this work are rarely
used, some have only been tested on a laboratory scale, and others are at the
pilot plant level, while only a few have been applied on an industrial scale. How-
ever, it is expected to see greater use in the future. An important exception is the
system used for automotive emission control of NOx, called diesel oxidation cat-
alyst (DOC, see Chapter 8), where a periodical switching between fuel-rich and
fuel-lean combustion generates pulses that are more efficient.
On the other hand, the consideration of operations in an unsteady state is an
interesting exercise to better understand the operation of chemical reactors.
When developing a chemical process, the production capacity marks whether
to carry it out continuously or in batch systems. Once made, by the high produc-
tion, the choice of a continuous process tends to consider the steady state as the
best option and even as the only one possible. This is usually due to the fact that,
in a continuous industrial process, a constant product output is sought, which
maintains its characteristics over time, so that an elementary analysis leads to
the conclusion that it is best to use a continuous process in steady state. And this
choice is supported, in addition, by certain additional practical considerations,
such as that the control of the process is much easier in steady state and it is eas-
ier for the operators to understand, and that a good prediction of the behavior
can be achieved (more stable and safe), since the equations that govern the pro-
cess are simpler, the control of the temperature is simpler, the assemblies can be
considerably cheaper, etc.
Storage tank
a particular case of the periodic operation, when the period tends to zero (or to
infinity); therefore, it does not seem logical to think that the permanent regime
will be the best option in all cases.
In fact, the studies carried out suggest that the optimal state of the catalyst sur-
face to achieve a high conversion or selectivity occurs in an unsteady operation, in
which the inertia of the catalyst is used. Increasing catalyst activity, expressed as
an increased conversion or reaction rate, is often the reason for studying the peri-
odic operation. However, as already indicated, FUSO is more complicated and
more expensive than the operation in steady state, so this increase in conversion
could be achieved in a more economical way by increasing the size of the reactor
or the amount of catalyst used. Consequently, working with a periodic operation
is only reasonable when the increase in catalyst activity is considerable.
Still, there are situations in which this conversion increase is a reasonable goal,
for example, in reactions with extreme temperatures and/or pressures or with
an enormously expensive catalyst. It is also interesting when it is necessary to
increase the capacity of an existing reactor, and it could also be considered in pro-
cesses whose conversion is limited, for example, by the equilibrium. The objective
can be even more ambitious, focusing on improving selectivity and/or produc-
tion, where operation in an unsteady regime proves to be perhaps the best tech-
nique to achieve it, except for the development of new catalysts. There is also
evidence that the periodic operation can decrease the rate of deactivation of the
catalyst, and this can also be an objective to be considered when considering the
application of a FUSO.
Operation in an unsteady regime has been studied for more than 70 years, and
during this time numerous studies have been published. They deal with ways to
get a periodic operation, describing the improvements that have been achieved.
According to these works, it can be seen that the alternatives to achieve a periodic
operation can be classified into four types, which are described in the following
sections:
1. Periodic variation of the input (concentration, flow or temperature),
2. Periodic flow inversion (periodic flow reversal),
3. Operation with variable volume (variable volume operation, “VVO”),
4. Operation with oscillating pressure (pressure swing).
PB (PB) average
Time
PA (PA) average
Reactant A
Time
Valves
Inert
Reactant B
Reactor
Figure 7.2 Flow variation of reagents A and B periodically between two values.
A series of terms are associated with this type of systems that operate in peri-
odic conditions when attempting to describe their periodicity. These terms can
be better understood by referring to Figure 7.3, in which the input signals and
process rate are schematically shown.
The variables that appear are as follows:
(a) Period (𝜏): Refers to the time passing between the repetitions of the variation
in the input condition.
(b) Division (s) (split): The duration of one of the parts of the cycle related to the
period. A value of s = 0.5 indicates that each of the two parts of the cycle has
the same duration, and it will be a symmetrical cycle. The split measures the
duration of a portion of the cycle in which one of the reactants is at its highest
concentration; for this reason, when we speak of a split, we must refer to one
of the reactants.
(c) Amplitude (A): The change in the value of the entry condition over the average
value. The amplitude will have a single value when there is symmetry, but if
s = 0.5, two amplitudes should be given, one for each portion of the cycle.
(d) Phase lag: The variations of the composition in the figure are 180∘ or 𝜋 radians
out of phase, but another delay could have been used. A special case would
be that of a pulse operation, in which the phase shift would be zero.
Reactor rate
with permission of Elsevier.
Instantaneous cycle-
invariant rate
A1
Mole fraction of reactants in feed
A2
sτ
Cycle
period
τ
Mean mole
fraction
Cycling operation
achieved, directing the separation between the reagent and the product, and
causing improvements in production and/or selectivity when limitations are
imposed for the balance of the reaction.
of the cycle. Therefore, the mode and the establishment of certain variables for
the system are practically the same.
Among all the different signals that can be developed when establishing the
periodic cycles of the inputs, three types stand out: rectangular signal (when the
range of the input values oscillates instantaneously from the maximum to the
minimum and vice versa) such as that shown in Figure 7.2; triangular signal (when
the value of the input decreases with time from the maximum to the minimum
and goes back up); and sinusoidal signal (when the function of the input variable
oscillates between the maximum and minimum values describing a sinusoid).
Of all of them, it has been shown that the most significant effects in the
unsteady state are produced by the application of a rectangular signal, being
even greater at a greater amplitude of the cycle (greater difference between the
maximum and minimum values of the entrance).
CO
O
CO + O
(a) (b)
Figure 7.4 Island model. (a) Steady- state operation, (b) FUSO. Source: Zhou et al. (1986).
Adapted with permission of Elsevier.
164 7 Forced Unsteady State Operation of Chemical Reactors
Steady operation is generally not economical, because the acid obtained has a
low concentration. In the dynamic operation, the SO2 and oxygen are introduced
to the reactor continuously, and the oscillation is applied to the water. In this way,
during the operation there is a period when the water flows, and then it does not
flow in the next period. This causes an increase in the rate of mass transfer, on
the order of 40%, and the production of concentrated sulfuric acid. Sulfuric acid
accumulates in cavities during the period when there is no water flow. Because
sulfuric acid is easily absorbed by water, it takes a long time for the catalyst to
be deactivated by the accumulation of SO3 , and high concentrations of acid are
obtained.
Temperature XA
400 1
1
300 2 3
3
0.5 2
200
100 0
Axial position Axial position
Figure 7.5 Operation diagram of an operation in adiabatic stages and one of “reverse flow.”
t=0 t=τ
Temperature
Flow direction
Output Feed
Fixed-bed
reactor
Three-way valve
168 7 Forced Unsteady State Operation of Chemical Reactors
greatly lower the final concentration of the product. Vanden Bussche and Fro-
ment (1996) devised an association of three reactors in which each bed operates
cyclically in three phases (inlet, recirculation, and exit), and in each cycle, each
reactor goes through the three phases with the same time in each of them. The
inversion of flow in the reactor occurs when passing from the recirculation phase
to the exit phase.
Time
7.6 Oscillating Pressure 169
The reagent is not completely converted into products, as part of it is used for
the growth and maintenance of microorganisms. In processes in steady state, it
is often not possible to work with this type of catalysts, since the microorganisms
mutate spontaneously, which results in a reactor not working as expected.
In some conventional batch reactors, a high initial concentration of the reac-
tants inhibits the conversion carried out by the microorganisms. For this reason,
in the VVO, the reagent is added to the reactor during the feeding period, and
subsequently it is discharged in the “batch” period. The volume of the reactor
that remains in the “rest” period contains enough microorganisms to keep the
four cycles as short as possible: if almost all the microorganisms were eliminated
in the discharge period, the feeding period would be too long. Usually, after a few
cycles, the reactor is completely drained to get rid of all the microorganisms that
no longer work satisfactorily.
So far, the VVO has also been applied to the preparation of fine chemical prod-
ucts, in order to prevent runaway reactions. In the fed-batch mode, the reagent
is slowly added to the reactor so that the heat released by the exothermic reac-
tion can be effectively dissipated, for example. Another application that the VVO
operation could have is in plants where multiple products of fine chemistry are
manufactured, in which different types of chemical reactions must be carried out
in the same reactor.
Scape
easier to implement is the square variation, and since it has not been shown that
another type is more effective is the most studied.
When changing the input pressure of a system with catalyst and adsorbent, the
system passes through three different stages:
(i) Feed is introduced under pressure.
(ii) The catalytic reaction takes place. The adsorption of the product in the
adsorbent material also occurs, which will be mixed in the bed with the
catalyst.
(iii) The bed is depressurized and the reactants and the products are separated
at low pressure. The desorption of the products occurs.
We can consider the whole process as a combination of two processes: the
periodic operation of the reactor and the cyclic separation process of adsorp-
tion/desorption by oscillating pressure. Large pressure variations occur near the
feed, but due to the low permeability of the bed, a current (which we will call
output current) can be continuously drawn from the other end of the reactor.
In addition, in the feeding phase of each cycle, the concentration can be varied
(as we have seen in a previous section), there being two limits: (i) introduce one
reagent and then the other in stages (case of reaction with two reagents) and (ii)
introduce a constant composition mixture. Although this second form is chosen,
we do not avoid the cyclic variation of the concentration due to the separation of
reactants because of the adsorbent.
The use of this method, combining chemical reaction and adsorption by oscil-
lating pressure, is relatively new. Although a new type of periodic operation, it
is more complicated to carry out than the periodic inversion of flow, and more
expensive, so that the application on an industrial scale seems far away. However,
it is theoretically interesting because the conversion can exceed what is predicted
by equilibrium.
Bibliography
Creaser, D., Andersson, B., Hudgins, R.R., and Silveston, P.L. (1999). Cyclic
operation of the oxidative dehydrogenation of propane. Chemical Engineering
Science 54: 4437–4448.
Haure, P.M., Hudgins, R.R., and Silveston, P.L. (1989). Periodic operation of a
trickle-bed reactor. AIChE Journal 35: 1437–1444.
Lange, R. and Hanika, J. (1994). Investigations of periodically operated trickle-bed
reactors. Chemical Engineering Science 49: 5615–5621.
Matros, Y.S. (1989). Catalytic Processes Under Unsteady State Conditions.
Amsterdam: Elsevier Science B.V.
Reshetnikov, S.I., Ivanov, E.A., Kiwi-Minsker, L., and Renken, A. (2003).
Performance enhancement by unsteady-state reactor operation: theoretical
analysis for two-sites kinetic model. Chemical Engineering and Technology 26:
751–758. https://doi.org/10.1002/ceat.200301640.
Silveston, P.L., Hudgins, R.R., and Renken, A. (1995). Periodic operation of catalytic
reactors-introduction and overview. Catalysis Today 25: 91–112.
Bibliography 171
Part II
Industrial Catalysis
It is important for the chemical engineer who must design a catalytic reactor to
know not only the design equations but also the chemical fundaments of the cat-
alysts that will be used in the industry. Thus, the chemical bases of catalysis are
discussed in this chapter, with special reference to the most important industrial
catalytic processes. The structure of the catalysts is detailed (with special inter-
est those that are supported on ceramic materials). With this information, we
proceed to the design of the reactors, dealt with specifically in Chapter 9.
8.1 Introduction
8.1.1 Reactors for Solid-Catalyzed Reactions
Reactions catalyzed by solids usually involve two or three phases. Solid catalyst
and gaseous and liquid reactants are brought in contact to achieve the desired
conversion. Some of the reactor types used are briefly presented here for
background information, with generalized remarks on their advantages and
disadvantages.
Gas–solid reactors are the most well-known two-phase catalytic reactor types
operated continuously. The major reactors for solid-catalyzed gas-phase reac-
tions are the fixed-bed, the fluidized-bed, and the entrained-flow reactors.
The fixed-bed reactor is either operated adiabatically or isothermally. In the
latter case, a multitubular reactor system is often chosen where the parallel reac-
tor tubes are surrounded by a heat-exchanging fluid. The fixed character of the
catalyst bed necessitates a long catalyst life. If deactivation occurs, it should be
possible to regenerate the catalyst inside the reactor.
Advantages of the adiabatic fixed-bed reactor are its simplicity, the high catalyst
load per unit volume, little catalyst attrition, and little back mixing. The disad-
vantages are the high pressure drop, difficult temperature control, and the long
diffusion distances.
A reactor system similar to the fixed-bed reactor is the moving-bed reactor,
where the deactivation rate is relatively low, but too high for pure fixed-bed oper-
ation, e.g. in some catalytic reforming processes. Catalyst particles should be
spherical and of uniform size for smooth flow through the bed.
The circulation of gases through fixed beds approximates the plug flow and is
completely different from the circulation in fluidized beds, where the gas flow is
not easily defined. The gas flow in a fluidized bed is very different from the plug
flow in that it can present a large bypass, because part of the gas passes in the
form of bubbles. This makes the contact not effective, needing a high amount of
catalyst if high conversions are desired. If, in a particular case, the effectiveness
of the contact is very important, it is better to use a fixed-bed reactor.
The fluidized-bed reactor is mainly chosen because of its good heat exchange
properties and the uniform temperature distribution. Furthermore, due to the
fluid-like behavior of the bed, the catalyst can be added and withdrawn during
operation for regeneration or replacement in case of fast deactivation. In addi-
tion, due to the relatively small particle size, the diffusion distances are short. On
the other hand, the disadvantages are the occurrence of catalyst attrition and back
mixing, and the fact that fluid-bed reactors are difficult to scale-up.
It is difficult to control the temperature in the fixed beds, especially if they are
large since their heat conductivity is low; therefore, in very exothermic reactions,
hot zones or mobile hot fronts will form, which can damage the catalyst and even
deactivate it. This does not happen in fluidized beds where the temperature can be
easily homogenized when there is a rapid mixture, and the bed can be considered
practically isothermal.
If the catalyst used in the process is easily deactivated and it has to be regener-
ated very often, it is better to use a fluidized bed since the behavior of the solids
as a fluid permits to handle them much more easily.
The entrained flow is used when very short contact times are required, for
example, in case of highly active catalysts that deactivate fast. In fluid catalytic
cracking (FCC), the recirculating catalyst also supplies part of the heat for the
endothermal reaction. Depending on the catalyst loading, one distinguishes
dilute and dense phase “risers.” Advantages of entrained-flow reactors are the
high mass transfer rates, short contact times, and the possibility of continuous
catalyst replacement. The disadvantages are the occurrence of catalyst attrition,
reactor erosion, and the requirement to separate the catalyst from the product.
Intermediate reactor types between fluidized-bed and entrained-flow reactors
also exist, like internally circulating bed reactors with an internal riser section
and an external annular section of a fluid/moving-bed type.
Relatively new are the monolithic reactor types, mainly used in environmental
applications. Examples include the three-way catalyst for the exhaust gas clean-
ing of gasoline engines, the selective catalytic reduction (SCR) for NOx reduc-
tion with ammonia, the oxidation of volatile organic compounds, etc. All have
extremely low pressure drops due to the structured character of the catalyst and
can be used at high flow rates.
Usually, no reactors of the fluid bed or slurry type are used for liquid–solid
reactions. In most cases, a gas phase is also involved, resulting in a three-phase
(“multiphase”) operation. Many applications of gas/liquid/solid reactors are
found in industry. Selective oxidations and hydrogenations, hydrotreating,
hydrations, and aminations of a liquid reactant (mixtures) catalyzed by a
solid are important industrial examples. Two extremes exist in three-phase
operations: the trickle-bed and the slurry reactor (Figure 8.1).
8.1 Introduction 177
Liquid
distributor
Liquid in
Gas out
Liquid out
In the trickle-bed reactor, the gaseous and liquid reactants flow cocurrently
downward over a packed bed of catalyst. The liquid phase must be distributed
such that a good wetting of the catalyst is obtained. Uneven wetting may lead to
local hot spots and runaways. The gas phase is the continuous phase. High gas
flow rates are used to remove heat. Since the liquid covers the catalyst particles,
reactant diffusion is much slower than that in the gas-phase operation. The nature
of the packed bed requires catalyst bodies of a few millimeters, so the catalyst
effectiveness is restricted.
Advantages of trickle-bed reactors are high catalyst load, no need for cata-
lyst separation, no attrition, and very limited back mixing. The disadvantages are
reduced catalyst effectiveness and selectivity, cocurrent operation, flow maldis-
tributions, and large pressure drop.
In slurry reactors, small catalyst particles (10–100 nm) and gas bubbles are
present suspended in a liquid phase by mechanical mixing. These reactors can
be operated in batch or semi-batch mode, with the gas phase usually supplied
continuously. Also, the continuous mode is possible.
Upflow reactors are sometimes used for small-scale testing. Here, the liquid
phase is the continuous phase through which the gas bubbles rise, and wetting
problems are absent.
Countercurrent operation may have certain advantages like removal of inter-
mediate products to avoid side reactions, to overcome equilibrium limitations
(hydrogenation of aromatics), or to avoid product inhibition. Also, temperature
profiles along the reactor length differ, which may be used advantageously. The
major problem is flooding, i.e. entrainment of the liquid by the gas phase. Larger
catalyst particles must be used to allow sufficiently high flow rates, thereby fur-
ther lowering the catalyst efficiency.
178 8 Industrial Catalysis
Resistance to
poisoning Intrinsic activity
(specially to and selectivity
coke)
Appropriate
distribution of
Stable structure pores (and pore
sizes)
Appropriate
Low cost mechanical
properties
8.1 Introduction 179
• They can improve the stability by separating the small crystals of the catalyst,
avoiding sintering.
• If a chemical that reacts with the catalyst is used as a support, a chemical with
higher catalytic activity can be formed.
• Increasing the surface area of the support reduces the sensitivity of poisons.
• If the support has a high thermal capacity and thermal conductivity, it can dis-
sipate the heat, avoiding local overheating that can cause sintering.
The most commonly used supports in the chemical industry are silica, alumina,
zeolites, and active carbon.
Preparation of these supports is very diverse, and depending on the necessary
properties, the preparation method can vary. Taking as support alumina, the tex-
ture will depend on many factors. The major phase in the material will have an
important effect on the surface characteristics and the porosity; the area can vary
from 500 m2 /g for γ-alumina type to 2 m2 /g for α-alumina.
The chemical structure of alumina is complicated due to the existence of differ-
ent phases. When alumina is generated by precipitating an aluminum solution,
the nature of the precipitate can be either a hydrogel of ammonium or a crys-
talline structure; and in the latter case, the particular structure is dependent on
the temperature, pH, and agitation of the initial solution.
Rarely, an industrial catalyst consists only in the active phase and a support
material, as mentioned before. The final catalyst will have different additives,
mainly promoters, substances with no catalytic properties by themselves, but
increasing the rate or the selectivity of the reactions, or both.
The catalytic stabilizers may help usually to prevent the sintering of the active
phase. For example, in the metallic catalysts, these compounds act as a barrier
among the different metallic points, in order to avoid sintering. Other additives
are used in the preparation of the catalysts, as is the case of those used to ease the
extrusion of the pellets or the cementation of the support granules.
During the preparation of the catalysts, the incorporation of the active phase
on the support can be done in different ways: precipitation, adsorption, impreg-
nation, ionic interchange, etc. But the important point is the arrangement of the
active phase on the support. The objective is to obtain an optimum profile of
the active phase over the support to be applied to a particular reaction system.
Figure 8.3. shows typical distribution profiles of the active phase for a spherical
pellet along its diameter.
In the uniform distribution, the material is dispersed along the diameter. This
is the most favorable case when the chemical step is controlling the reaction
rate, and no diffusion control exists. Nevertheless, if the reaction is limited by
mass transfer phenomena, it is better to use a catalyst where the active phase is
on the exterior surface of the sphere (second drawing in Figure 8.3 “Outer egg
shell”). On the contrary, in systems where the catalyst is exposed to different poi-
sons or extreme conditions, it may be better to use a geometry where the active
phase is on the center of the pellets (“inner egg yolk”), or a situation among them
(“middle egg white”). In order to get these active profiles in the support, usually
a competitive adsorption with other ions is used, which can later be eliminated
or minimized in any way.
180 8 Industrial Catalysis
Cylindrical or
spherical
geometry
Impregnation
profile
Catalysts with the same kind of global distribution of active phase can present
differences in the distribution at a smaller scale. In this way, the activity and selec-
tivity of the catalysts will be a function of the size of the metallic particle and the
morphology of the support.
The inorganic oxides are dissolved in acidic or basic media, so activated car-
bon is used in these cases. In addition, the acidity of supports such as silica and
alumina is not always desired, and more inert materials like anatase or activated
carbon are used. Furthermore, silica cannot be used under conditions of high
temperature and steam atmosphere, because evaporation takes place.
8.1.2.4 Zeolites
The majority of silica-aluminates are amorphous materials, that is, in which there
is no ordering of the atoms in a large distance forming a crystallographic phase.
However, zeolites are silica-alumina materials presenting a very interesting crys-
tallinity. Zeolites are crystalline and have a well-defined pore structure with pores
of molecular dimensions.
Zeolites are formed by assembling tetrahedral units of SiO4 or (AlO4 )− . These
units are first grouped in a building block composed of 24 tetrahedra, called
sodalite, and then in units of sodalite, giving rise to the superstructure of the
182 8 Industrial Catalysis
Al
Si
Si
Al
Al Si
0.74 nm
Figure 8.4 Structure of sodalite unit and superstructure of the zeolites. Source: Adapted from
https://commons.wikimedia.org/wiki/File:Faujasite_structure.svg.
zeolites. Figure 8.4 shows the structure called faujasite, in which the Si or Al atoms
are in the intersecting lines and the O atoms are between them.
In the inner part of the structure of the zeolite, we find cavities with large holes,
whose size depends on the way in which the sodalite units are assembled. The size
of these cavities determines the size of the molecules that can be adsorbed inside
and that can react within the “box” of zeolite. Other ways to assemble the sodalite
units include the following:
• Sodalite, formed by units of sodalite directly above each other, with a diameter
between the cavities of only 0.26 nm, where only small molecules such as water
or methanol can be adsorbed;
• Zeolite A, in which the units are separated by an oxygen layer forming cavities
of 0.42 nm in diameter that can adsorb somewhat larger molecules;
• Faujasite, with cavities of 0.74 nm (Figure 8.4);
• Mordenite, with cavity size of 0.80 nm;
• ZSM-5, with cavity of 0.60 nm.
More than a hundred zeolite structures are well defined on an atomic scale.
Thus, bond length and angles of the atoms comprising the catalyst material are
known in detail. Depending on the type of zeolite, the microporosity shows a
sharp, almost unique pore size distribution in the range of the molecular dimen-
sions, enabling molecular shape selectivity.
Based on the chemical and physical compatibility of alumina and silica on an
atomic scale, relatively strong Brønsted acidity is achieved by substitution of alu-
minum that is counterbalanced by protons. The catalytically active sites induced
by a post-synthesis procedure of the zeolites are accessible for reactant molecules,
since all the atoms in the zeolite framework are exposed to the pore volume.
8.2 Industrial Preparation of Catalysts 183
Many reactions in the oil refining and petrochemical industry are accompanied
by the formation of carbonaceous by-products (as coke) that usually deactivate
the catalysts. Zeolites are thermally stable and can, therefore, be regenerated at
high temperature by oxidation of these carbonaceous deposits.
The contribution of zeolites in the industry has been mainly developed in the
past 40 years, and it can be considered essential for the development of many
industrial processes. Table 8.2 shows the main processes where zeolites are
present in refineries.
is large and fragmented in numerous segments, each with its own characteristics.
These characteristics are generally determined by the chemical nature of the
catalyst and/or the market in which, normally, the segments are not competitive
with each other.
must be treated with ammonium salts, so that the ammonium salt of the zeolite
is produced, ending with a 350 ∘ C heating that releases the ammonia and leaves
the proton form of the zeolite (Brønsted acid). Subsequent heating to 450 ∘ C will
liberate the water from the structure forming the Lewis acid sites.
(1) Wet mixer: In this unit, the mix of the alumina powder with water is done,
and other ligands, with the aim of getting a mixture with some particular
rheological properties. It is obvious that one of the main steps to obtain a
good catalytic support is to get a good mixture in this step.
(2) Extruder: The mix, when reaching the desired rheological properties, is dosed
to an extruder. This equipment can be assembled with a double or a simple
spindle, and dispose of a system for evacuating the air that can be occluded
in the mix. Also, the extruder will have a system for heating or cooling the
mix, in order to regulate the working temperature. Also, the extruder head
should be modified in accordance with the desired physical form of the final
product. The geometry of the support is of special importance, modulating
the surface-to-volume ratio.
(3) Pre-drying oven: Usually, the extruded product exits with a high humidity,
and it is necessary to eliminate it before the calcination step.
(4) Calcination oven: Composed of a conveyer belt with programmable veloc-
ity, in order to vary the time the support is in the range of temperature
300–800 ∘ C. Heat is obtained by infrared resistances or with burners that
use natural gas as fuel.
(5) Sifting machine: The product leaves the calcination oven with different parti-
cle sizes, and in the sifting machine they are separated into three parts (fines,
thick material, and the desired product).
Nickel The industrial manufacture of catalysts that have nickel as active element
generally uses nickel nitrate (Ni(NO3 )2 ) as mineral salt, because it presents an
excellent solubility in water and has a very low cost. Contrarily, if this salt is used,
a system to eliminate NOx will be needed, and also a loss of activity as this is not
the salt giving the best activity.
Another possibility is to use nickel formate (C2 H2 NiO4 ), which presents some
advantages such as higher activity and that a system for eliminating contamina-
tion will not be necessary. Nevertheless, the salt has low solubility and implies
use of ammonia solutions.
186 8 Industrial Catalysis
Cobalt It is similar to nickel, that is, the main salt used is cobalt nitrate
(Co(NiO3 )2 ), but cobalt acetate (Co(CH3 CO2 )2 ) can also be used. Currently, the
drawback of these salts is the high cost.
Table 8.3 shows typical spatial velocity values for the three most important
types of catalytic applications, which are the production of chemical products,
operation in refinery, and exhaust gas conversion.
For instance, a typical catalytic reactor in a chemical plant may contain
500 000 kg of catalyst, which has a capacity of 500 000 kg per day. This represents
a weight/weight-h velocity of
WWH = 500 000 kg product∕(500 000 kg cat − day) = 1 day−1 = 0.04 h−1
This mass of catalyst may be also necessary to convert 100 000 barrels
(1.4⋅108 kg) of oil into gasoline and diesel oil in a refinery, which will be
equivalent to a space velocity of 1.2 h−1 .
The catalysts used in the exhaust gas cleaning typically fill a volume of 5 l and
need to convert flow rates of approximately 250 000 l/h. These amounts results in
a required gas hourly space velocity of GHSV = 50 000 h−1
Another usual way to measure the ability of a catalyst is the “turnover rate.”
The abovementioned parameters do not give an idea of how a particular catalyst
is functioning on a molecular scale. This type of information could be needed to
understand if a catalyst is working correctly from the chemical point of view. It is
also helpful to compare the performance of different catalysts of diverse physical
properties with different structures.
This type of measurement is carried out with the so-called turnover rate or the
turnover number, which is the number of catalytic cycles per active site of the
catalyst and per unit of time:
moles of product
Turnover rate (TR) =
[(moles of active sites) ⋅ (time)]
grams of reactive
MW reactive
=[ ( ) ]
(g cat) ⋅ gsites
cat
⋅h
grams of reactive
(g cat)⋅h WWH
=[ ( )] = (8.4)
sites
MW reactive ⋅ g cat MW ⋅ SD
Table 8.4 Important industrial catalyzed reactions with the corresponding catalyst.
Main Type of
Reactants products/coproducts Type of catalyst reactiona)
Main Type of
Reactants products/coproducts Type of catalyst reactiona)
Main Type of
Reactants products/coproducts Type of catalyst reactiona)
Alkynes cis-Alkenes Pd H
Alkynes Alkanes Pd H
Anilines Cyclohexylamines Rh, Pd H
Anilines Cyclohexanones Pd H
Epoxides Alcohols Pd H
Hydrazones Hydrazines Pt H
Imines Amines Pt H
Aliphatic ketones Alcohols Ru H
Aromatic ketones Aromatic alcohols Pd H
Benzyl compounds Aromatic Pd H
hydrocarbons
Benzyl compounds Cyclohexyl Rh, Ru H
derivatives
Aliphatic nitriles Primary amines Pd, Pt, Rh H
Aliphatic nitriles Secondary amines Rh (solvent) H
Aliphatic nitriles Tertiary amines Pd, Pt (solvent) H
Aromatic nitriles Benzylamines Pd H
Aromatic nitriles Dibenzylamines Pt H
Aromatic nitriles Aldehydes Pd H
Nitriles Secondary amines Rh H
Nitriles Tertiary amines Pd, Rh H
Phenols Cyclohexanones Pd H
Phenols Cyclohexanols Rh, Ru H
Phenols Cyclohexanes Pt, Ir H
but can accept electrons from the reactants. Commonly, the Brønsted acids can
be converted into the corresponding Lewis acid by loss of a molecule of water,
and the hydration reaction will produce the Lewis acid, in reactions of the type
(usually reversibly):
−−−−→
+H O
Lewis acid ←−−2−− Brønsted acid
−H2 O
Some of the most important industrial processes related to the acid catalyst are
those related to the refinery plants. In these, the main processes in which the acid
catalysts intervene are the following:
• Catalytic cracking (FCC);
• Disproportionation of aromatics (reaction in which multi-alkylated aromatics
are generated without substitutions, which is used to increase or decrease the
octane level of gasoline);
• Alkylation of isobutane/isobutylene in which a C8 molecule is produced from
two C4 , and super acids are required as catalysts;
• Synthesis of methyl tert-butyl ether (MTBE);
• Methanol-to-gasoline (MTG) process, in which a zeolite-type ZSM5 catalyzes
the formation of (CH2 )n and H2 O starting from methanol.
Acid catalysis is also present in the chemical industry, in general. Some of the
production processes of products in the chemical industry, which require acid
catalysts, are the following:
• Benzene ethylation, producing ethylbenzene;
• Production of cumene by reaction of benzene and propylene;
• Isomerization Beckman producing caprolactam by reacting a cyclohexanone
oxime with a strong acid;
• Esterification, forming ethyl acetate mainly.
Combustion gases
Stripper
Steam
Regenerator
Riser
Air
Feed
active in the combustion of the coke. Zeolite is the component determining the
selectivity of the reaction to the different products.
The present FCC catalysts incorporate zeolite of the type Y as the main active
component. In this, not all the active centers have the same acid strength, and
there exists a distribution depending on the amount of Al in the net, i.e. on the
ratio Si/Al.
Due to the high temperatures to which the catalyst is subjected, especially in the
regenerator and in the presence of steam, the zeolite in the fresh catalyst suffers a
de-aluminization process, whereby Al is extracted from the net and is deposited
on the surface of the zeolite.
Another characteristic of zeolite Y affecting its catalytic behavior is the size of
the particles. In fact, the smaller the particle size, the higher the amount of acid
sites in the surface and they are accessible to gasoil molecules. Also, the diffusion
of the cracking products is favored with small particle sizes, and the possibility
of producing secondary reactions is reduced.
Zeolite ZSM-5 has also been used in the FCC process. This is a zeolite with
smaller channels than type Y, and can be useful to obtain products of low molec-
ular weight (gas fraction). Also, ZSM-5 is more resistant to the deactivation by
coke. ZSM-5 is usually mixed in a low proportion (2–5%) with Y-zeolite, thus
optimizing the process.
8.3 Main Catalytic Processes in Industry 193
V-4
R-1
10
3
1
V-1 E-1
B-1
6
5
4
E-2
2 7
V-2
C-1
Figure 8.8 Flowchart of the ethylbenzene production process. V-1, V-2, and V-4 – valves;
C-1 – compressor; B-1 – pump; E-1 and E-2 – heat exchangers; R-1 – reactor.
194 8 Industrial Catalysis
The most important factor for modeling the reactor is the kinetics of the reac-
tions that occur in it. Two reactions occur, the main reaction, catalyzed by zeolite,
and the secondary reaction, where an undesired product is produced:
Zeolite Y
C6 H6 + C2 H4 −−−−−−−→ C6 H5 C2 H5
C6 H5 C2 H5 + C2 H4 → C6 H4 (C2 H5 )2
This reaction is carried out at 190 ∘ C in the presence of a zeolite catalyst.
from 1975 onwards, all the production of ethylene oxide in the United States was
exclusively by direct oxidation of ethylene.
The main causes that motivated the boom of the direct oxidation process were
the large number of chemical products required by the traditional method of
obtaining it, as well as the high cost of chlorine, which had a marked influence
on the total cost. Also, the considerable contamination of the water could be a
problem.
The direct oxidation process is carried out using oxygen or air on a silver cat-
alyst, as already mentioned. This reaction takes place in the gas phase, unlike
the traditional chlorohydrin process that was carried out in the liquid phase. The
main reactions that take place in obtaining ethylene oxide by direct oxidation are
shown here:
O2 + 2C2 H4 → 2C2 H4 O
O2 + 1∕3 C2 H4 → 2∕3 CO2 + 2∕3 H2 O
The process is usually carried out in fluidized-bed reactors. The catalyst used
in the reactor contains a wt. 5% of silver, using a support composed of alumina
and silica (8% by mass of Al2 O3 and 92% of SiO2 ). Spherical particles of very small
diameter are used, of a density close to 1800 kg/m3 . The efficiency factor of such
a catalyst under the conditions of the reaction is practically unity. The process is
done at 550 K and 10 atm. of pressure, and the input mix to the reactor must have
less than 6 vol% of oxygen, to avoid the total combustion of the ethylene. Under
these conditions, the total conversion of the ethylene is close to 0.40, so it is con-
venient to improve the global conversion by recirculating the unreacted ethylene.
For this reason, a unit for the separation is needed, as shown in Figure 8.9. On the
other hand, as the gas velocity is usually high, the catalyst particles are dragged
and a cyclone at the exit separates the solid from the gas, returning the valuable
catalyst to the bed.
Ethylene recirculation
Separation
Ethylene Air
Cyclone
Catalyst
recirculation
Compressor
Water
3CO2 + H2O
r3 r4
+11/2 O2 +9/2 O2
r1 r2 OH
H + H 2O
+ O2
+1/2 O2
Propylene O
O
Acrylic acid
Acrolein
Noble metals Basic metals Alkaline earth metals Rare earth metals
The composition of the exhaust gases varies widely depending on the condi-
tions of the engine and its type. Table 8.6 shows a set of typical values for gasoline
and diesel engines.
The main reason for these differences is that typical temperatures are much
higher in gasoline engines. This produces a high emission of NOx (that is a
product of excessive combustion) and CO, but less soot (carbonaceous material,
the main component of particulate matter), consistent with the higher level
of combustion. The much greater presence of O2 in the exhaust of a diesel
engine makes the simultaneous elimination of coal, CO, HC, and NOx very
complicated. In addition, the high carbon content causes the catalysts to become
poisoned, which will require treatment.
The gases could be divided into oxidants and reducers. In the first category
would be O2 , NOx, and SOx, while the reducers would be HC, carbon, CO, and
H2 (that is presented usually in very small proportion). In addition, ammonia
(NH3 ) produced by decomposition of urea (CO(NH2 )2 ) can be added to the
exhaust gases as a reducing agent. This will permit elimination of NOx in a
process called “selective catalytic reduction,” or SCR.
8.3.4.1 Catalytic Reactions for the Removal of Pollutants in the Exhaust Gases
NOx reduction can occur through various chemical processes:
2NO + 2CO → N2 + 2CO2 (3 − way catalyst)
Cm Hn + (2m + n∕2) NO
→ (m + n∕4) N2 + mCO2 + n∕H2 O (lean NOx catalyst)
Cm Hn + NOx + O2
→ N2 + CO2 + H2 O (lean NOx catalyst − oxygen rich)
NO + NH3 → N2 + H2 O (selective catalytic reduction)
In the three-way catalysts used in gasoline engines, the first reaction is involved,
NO being removed by the presence CO at low concentrations of O2 , producing
N2 and CO2 . Hydrocarbons (HCs) can also react with NOx to form N2 , in the lean
NOx catalyst that is described later. Also, in the presence of O2 , some HCs will
be oxidized to form extra CO2 and water. In this case, all HCs may be eliminated
and do not react with NO. Finally, urea can be added to produce NH3 , which
reduces NOx by SCR. Since most NOx (>90% in diesel) are present as NO (with
10% NO2 ), the main reaction will be that shown earlier.
On the other hand, the main process for the elimination of CO will be the oxi-
dation to CO2 , which occurs in the three-way catalysts in gasoline engines, and
also in a “diesel oxidation catalyst” (DOC):
2CO + O2 → 2CO2 (3 − way catalyst, DOC)
CO + H2 O → H2 + CO2 (3 − way catalyst, DOC)
A minority reaction that also follows in these catalysts is the WGSR forming
N2 and CO2 from CO and water, discussed in the previous topics. Hydrogen is a
very potent reductant that can be used to reduce NOx to N2 and water. The HCs
are also removed by oxidation to CO2 with the three-way catalyst.
202 8 Industrial Catalysis
NOx
3-WAY
SCR
CATALYST
(NH3)2CO +
High temp. Al2O3/LaO2 H2O → 2NH3 NH3 + NO +
Low O2 N2 + H2O
Pt/Rh/Pd + CO2 NO2 + O2 >>
N2 + H2O
Gasoline Urea/H2O
Exhaust
engine injection
Urea/H2O
DOC injection
Exhaust
NH3 + NO +
Low temp. BaO, CeO, (NH3)2CO + NO2 + O2 >>
Pt/Rh/MgO N2 + H2O
High O2 Rh (Al2O3 H2O → 2NH3 N2 + H2O
support) + CO2
Diesel engine DPF SCR
Figure 8.11 Some catalytic processes present in the gas exhaust of diesel and gasoline
engines.
75 NOx
Emission (% of maxima)
HC
50
25
S∗ CO
12 14 16 18
Air/fuel ratio
In the case of a low amount or air (small air/fuel ratio), there is an excess of fuel
and the conversion of CO and HC does not occur because there is not enough
oxygen. Under these conditions, the elimination of NOx is favored, and the power
of the engine is high, but the fuel consumption is very high. Under these condi-
tions, there is much more CO than NO, so the NO is not enough to eliminate all
the CO present.
On the other hand, when the ratio air/fuel is very high (very oxidant condi-
tions), there is a drastic increase of the emission of HC and a loss of engine power.
At values of ratio higher than 24, the ignition cannot be produced. In usual driv-
ing conditions, engines usually run at a ratio somewhat poor, minimizing fuel
consumption.
In gasoline engines, this is achieved by a careful control of the oxygen content
by means of sensors (“lambda probe”), which keeps the oxygen level very close to
the stoichiometric. Since there is no excess oxygen, it does not compete with NO
for the oxidation of CO or HC and therefore it cannot be effectively reduced to
N2 in commercial three-way catalysts.
The last two systems provide NOx reduction at a lower and higher temperature
than V2 O5 . Pt-based systems are better for heavy diesel engines that have a lower
exhaust temperature.
The SCR process is also present in many industries, in processes where NH3 or
urea is injected in the gas before emission, in order to decrease the NOx emission.
An example of the design is discussed in Chapter 9.
Figure 8.13 Flow dynamics in a DPF and general aspect. Source: Adapted from https://
commons.wikimedia.org/wiki/File:Dpf_filter.png and https://commons.wikimedia.org/wiki/
File:DPF_Cordierite_Alumunium-Titanat.jpg.
act as a filter, eliminating the carbon formed (and other PM), which is deposited
on the walls. These walls are coated with a compound that disperses the noble
metal (Pt/Rh/MgO) or the basic metal (V2 O5 ). Figure 8.13 shows a scheme of the
flow dynamics.
The appearance of these filters is the one seen on the right side. The filter acts in
two periods. The first period is passive, and the particles are accumulating while
they are slowly oxidized by the air. Later, when the pressure loss is too important,
the temperature of the catalyst increases due to the heat produced by the com-
bustion of the particles. The second period occurs when the temperature exceeds
a certain threshold at which the rate of combustion of the soot is high and, in
contact with the catalyst, CO2 is produced. Thus, begins a new cycle. A typical
temperature of the first period is 350 ∘ C, while at 500 ∘ C combustion will start
(“soot light off temperature”).
Systems based on noble metals (Pt and Pd) are the most used in the DPF. This
is related to the high activity presenting at the low temperatures of the diesel
systems. However, Pt and Pd are very sensitive to the presence of sulfur in the
exhaust gases. One of the unwanted processes in DPF and DOC is the formation
of sulfate salts from SO2 and SO3 . In this way, the efficiency of a DPF dramatically
decreases in the presence of small amounts of S in the fuel, from about 50 ppm.
Bibliography
Cifuentes, B., Valero, M., Conesa, J., and Cobo, M. (2015). Hydrogen production by
steam reforming of ethanol on Rh-Pt catalysts: influence of CeO2 , ZrO2 , and
La2 O3 as supports. Catalysts 5: 1872–1896. https://doi.org/10.3390/catal5041872.
Clark, J.H. (2001). Catalysts for Green Chemistry. Pure Appl. Chem. 73: 103–111.
Cobo, M., Becerra, J., Castelblanco, M. et al. (2015). Catalytic hydrodechlorination
of trichloroethylene in a novel NaOH/2-propanol/methanol/water system on
ceria-supported Pd and Rh catalysts. J. Environ. Manage. 158: 1–10. https://doi
.org/10.1016/j.jenvman.2015.04.035.
Gates, B.C. (1992). Catalytic Chemistry, 1992. Wiley.
González Velasco, J.R. and Gutiérrez Ortiz, M.A. (1997). Catálisis: una ciencia
multidisciplinar con presente & futuro. Universidad del País Vasco.
Greenwood, N.N. and Earnshaw, A. (1997). Chemistry of the Elements.
Butterworth-Heinemann.
Kent, J.A. (2012). Handbook of Industrial Chemistry and Biotechnology. London:
Springer https://doi.org/10.1007/978-3-319-52287-6.
Reshetnikov, S.I., Ivanov, E.A., Kiwi-Minsker, L., and Renken, A. (2003).
Performance enhancement by unsteady-state reactor operation: Theoretical
analysis for two-sites kinetic model. Chemical Engineering and Technology 26 (7):
751–758. https://doi.org/10.1002/ceat.200301640.
Satterfield, C.N. (1991). Heterogeneous Catalysis in Industrial Practice, 2e.
McGraw-Hill.
White, M.G. (1990). Heterogeneous Catalysis. Upper Saddle River, NJ: Prentice Hall.
207
9.1 Introduction
As mentioned in Chapter 8, the rate of a reaction can be modified by the presence
of certain substances, which are usually neither initial reactants nor products.
These substances are called catalysts, and they produce an increase in the reac-
tion rate.
A catalyst can vary the reaction rate of a process in thousands or millions of
times, so its use in industrial processes is usually very interesting. The character-
istics of a catalyst can be summarized as
1. Selectivity of the catalysts. It refers to the ability of these substances to act in
certain reactions and not to do so in different ones.
2. A catalyst increases the rate of reaction or favors a reaction vs. another, but in
no case determines the equilibrium or the end point thereof , which will always
be conditioned by thermodynamic issues.
3. The activity of a catalyst depends not only on its chemical constitution but also
on its physical or crystalline structure, since the catalyst can lose its activity
above certain temperatures. Therefore, an exhaustive study of a catalyst must
include an investigation of its surface. In this way, it is necessary to ensure that
the catalysts have a large active solid surface per unit volume.
4. In the presence of a catalyst, the reactant molecules undergo a weakening of
their bonds or form intermediates in the vicinity of the surface of the solid.
Several models that try to explain this behavior have been proposed:
(a) The intermediate product is an association of the reactive molecule with
a region of the surface of the solid.
(b) Reagent molecules are under the influence of surface forces, moving in an
environment close to it.
(c) Free radicals form on the surface of the catalyst, which move toward the
gas causing a series of reactions, until finally they are destroyed; in this
last theory, the reaction will take place within the gas, the surface of the
catalyst being a generator of free radicals.
5. According to the transition state theory, the catalyst reduces the potential
energy barrier that must be exceeded in order for the reactants to pass into
products. This will, in turn, cause a decrease in the activation energy, and
therefore, an increase in the reaction rate, as can be seen in Figure 9.1.
Initial state
To be able to carry out a design of reactors for this type of reactions, it will be
necessary to study in detail the rate equation in these processes from the study
of the kinetics of the reaction.
Table 9.1 Different properties used for defining reaction rate and their corresponding valid units.
Kinetic constant
Rate equation (valid units) assuming Relationship with
Extensive property symbol (valid units) first-order kinetics generation term
Note that the units of the kinetic constants will depend on the expression used
to define the reaction rate, even for first-order kinetics. The table shows valid
units for first-order kinetics in the different rate expressions; Eq. (9.6) could serve
for changing units of kinetic constants or the reaction rate, if necessary.
Solution
First of all, let us calculate the pre-exponential factor and activation energy from
the given data. We can do it using the Arrhenius equation:
( )
E
k ′ = k0′ ⋅ exp −
RT
210 9 Catalytic and Multiphase Reactor Design
Knowing k ′ (at 250 ∘ C) = 0.050 and k ′ (at 400 ∘ C) = 2 m6 /(mol kgcat s), we calculate:
m6 E
k0′ = 7.71 ⋅ 105 = 8656.1 K
mol kg s R
(a) The mass balance in isothermal conditions in a plug flow reactor with cata-
lyst is
dnA = dgA = (−rA′ ) ⋅ dW
where “dW ” is the weight of the catalyst in a differential volume. From the
data we know that (−rA′ ) = k ′ ⋅ CA2
In this case, the concentration depends on the molar conversion of the reac-
tant (X A ) and also on the possible change in volume (or flow rate) due to the
change in moles of the reaction. In this way:
nA n ⋅ (1 − XA )
CA = = A0
Q QO ⋅ (1 + 𝜀 ⋅ XA )
where 𝜀 is the expansion factor of the reaction, i.e. the number of moles pro-
duced by mol of reactants, in this case 𝜀 = 2; therefore:
(1 − XA )
CA = CA0
(1 + 2 ⋅ XA )
( )
( ) 500
atm ( )
P0 1013 mol
CA0 = = atm l
= 0.115
RT0 0.082 mol ⋅ (250 + 273) K l
K
®
With initial condition X A = 0, we will use Matlab to solve the equation, as
shown in Program listing 9.1:
u0=3;% m/s
porosity=0.4;
dcat=2000;% kg/m3 catalyst
k0=7.707e5;% m6/mol ⋅ kg ⋅ s
E_R=8656.1; % K
P=500000; % Pa
Ca0=P/(8.314*523); % mol/m3
T=250+273; % K
k=k0*exp(-E_R/T); % m6/mol ⋅ kg ⋅ s
for i=2:n
L(i)=L(i-1)+incx;
XA(i)=XA(i-1)+(dcat*(1-porosity)*k*Ca0*(1-XA(i-
1))∧ 2/(u0*(1+2*XA(i-1))∧ 2))*incx;
end
figure(1)
plot(L,XA)
xlabel(’Length (m)’)
ylabel(’Molar conversion’)
In the program, the different values of the variables are given, and the length
increment is defined. The number of points is then calculated, and the differential
equation is solved numerically using the finite differences method. The values of
212 9 Catalytic and Multiphase Reactor Design
length and conversion are saved in two vectors that are plotted in the last part of
the program. The following figure shows the result.
1
0.9
0.8
0.7
Molar conversion
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Length (m)
(b) The mass balance in adiabatic conditions is similar to that of the isothermal
case, but the temperature change must be borne in mind. The energy balance
in a PFR, with only one reaction and without transfer abroad (adiabatic), is
∑
s
dXA
dT ⋅ nj0 Cpj + ΔHR ⋅ =0
j=1
nA0
Bearing in mind that only reactant A is fed, and rearranging:
dT ΔHR
=−
dXA CpA
In the isothermal case, we checked that Q = QO ⋅ (1 + 2 ⋅ X A ) but now a cor-
rection of the flowrate with temperature is needed, and the following results:
T
Q = QO ⋅ (1 + 2 ⋅ XA )
T0
and then:
dT dT dXA
= ⋅
dx dXA dx
ΔHR 𝜌cat ⋅ (1 − Porosity of the bed) ⋅ k ⋅ CA0 (1 − XA )2
=− ⋅ ( )2
CpA
u0 ⋅ TT ⋅ (1 + 2XA )2
0
Some small changes are needed in the program listing used in the previous
isothermal case. A valid program is given, as shown in Program listing 9.2.
u0=3;% m/s
porosity=0.4;
dcat=2000;% kg/m3 catalyst
k0=7.707e5;% m6/mol ⋅ kg ⋅ s
CpA=15; %J /mol ⋅ K
DHR=-800;% J/mol
E_R=8656.1; % K
P=500000; % Pa
Ca0=P/(8.314*523); % mol/m3
for i=2:n
L(i)=L(i-1)+incx;
dXAdt=(dcat*(1-porosity)*k*Ca0*(1-XA(i-
1))∧ 2/(u0*ftemp*(1+2*XA(i-1))∧ 2));
XA(i)=XA(i-1)+dXAdt*incx;
T(i)=T(i-1)+(-DHR/CpA)*dXAdt*incx;
k=k0*exp(-E_R/T(i)); % m6/mol ⋅ kg ⋅ s
ftemp=(T(i)/523)∧ 2;
end
figure(1)
plot(L,XA)
xlabel(’Length (m)’)
ylabel(’Molar conversion’)
figure(2)
plot(L,T)
xlabel(’Length (m)’)
ylabel(’Temperature’)
The results are shown in the following two figures (conversion and temperature
evolution vs. length of the reactor).
214 9 Catalytic and Multiphase Reactor Design
1 580
0.9
570
0.8
0.7
Molar conversion
560
Temperature
0.6
0.5 550
0.4
540
0.3
0.2
530
0.1
0 520
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Length (m) Length (m)
Reactant “A”
concentration
CAg
Void volume
CAs
External Internal
diffusion diffusion
Figure 9.2 Porosity and void volumes in catalyst. Reaction steps in catalytic systems.
9.3 Rate Equation When Chemical Step Is Limiting Reaction Rate 215
1. External diffusion (or film resistance): Diffusion of the reactants toward the
outer surface of the catalyst;
2. Internal diffusion (or resistance to diffusion in the pores): Diffusion of the reac-
tants to the interior of the catalyst, where the reaction will mainly occur;
3. Chemical adsorption (or physical in certain occasions) on the surface: During
the union of the reagent molecules somewhere on the surface of the catalyst;
4. Surface chemical reaction: Reaction on the surface of the catalyst particle;
5. Chemical desorption on the surface: During the separation of the product
molecules from the surface of the catalyst;
6. Internal diffusion (or resistance to diffusion in the pores): Diffusion of products
toward the exterior of the catalyst;
7. External diffusion (or film resistance): During the exit of the products toward
the gas phase.
Of all these reaction steps, that of internal diffusion has some particular char-
acteristics. This step is considered as a controlling step, in the sense that if it is
very slow the rate will be affected by the internal diffusion, since the reaction
could also take place on the surface of the solid. Therefore, it is said that the resis-
tance to internal diffusion is “in parallel” with the other resistances, all of them
“in series.”
Catalyst
Hydrocarbon Carbonium ion + 1e–
Catalyst
Catalyst
At the end, the carbonium ion can react with another initial hydrocarbon
molecule to give a carbonium ion with a greater number of carbon atoms and
start the breaking steps again. Among the currently used cracking catalysts
are zeolites, as shown in Chapter 8. These compounds are crystalline alumi-
nosilicates containing “boxes,” often of molecular dimensions, which can block
branched chains of the molecules and thereby favor breakage by a given bond,
increasing the selectivity in a given component.
Metal catalysts are frequently used in dehydrogenations and hydrogenations.
An example can be the synthesis of NH3 , which would follow the following
scheme:
Fe
N2(g) 2N(ad)
Fe
N(ad) + H(ad) NH(ad)
Fe NH3(g)
H2(g) 2H(ad)
+H(ad)
NH2(ad) NH3(ad)
+H(ad)
For the oxidation reactions of hydrocarbons, the active components of the cata-
lyst are metal oxides (molybdenum, vanadium, chromium, etc.), capable of having
several valences. In this case, it seems that the following reactions take place:
Hydrocarbon + Oxidized Catalyst
→ Reaction products + Partially reduced catalyst
Partially reduced catalyst + O2 → Oxidized catalyst
It has been possible to verify these phases, marking the oxygen atoms (isotopes)
of the catalyst, and subsequently checking the presence of these atoms among the
products of the reaction.
The catalytic activity has been related to the electrical conductive character of
the catalyst. It has been found that when the catalyst was a little conductive, no
reaction took place; if it was very conductive, the most developed reaction was to
total the carbon oxides and water, and for an intermediate situation the desired
oxidized products were frequently formed (benzoic acid and benzaldehyde from
toluene).
For many reactions, the reaction mechanism, which includes the adsorption
and desorption of the active centers, can be represented by schemes such as the
following (where L is the active center):
Slope ∼
−1
0
CA
and therefore
CO2 ⋅L = Kad CO2 CL (9.22)
since the adsorption step is the slowest one, the reactants and the products will
be in equilibrium. It follows that:
Ct = CL + Kad CO2 CL = CL (1 + Kad CO2 ) (9.23)
So:
k Kad Ct CRH CO2 k ′ CRH CO2
r = k CO2 ⋅L CRH = k (Kad CO2 CL )CRH = =
1 + Kad CO2 1 + Kad CO2
(9.24)
The two deducted expressions have been checked experimentally. For any
other mechanism, the corresponding equations can be obtained. The rate
expressions will coincide with those obtained experimentally, provided that
the accepted hypotheses are correct (Langmuir mechanism, controlling step,
etc.), in addition to the fact that the remaining steps of the reaction are faster.
The value of rate of the given equations can be matched with any of the rate
expressions, based on the catalyst mass, reactor volume, etc.). The units of the
constant or constants will depend on it.
From the equations obtained, it can be predicted, if the model is correct, how
the reaction rate will vary with the total pressure of the system, or with the con-
centrations of the reactants.
The problem, when external diffusion is the controlling step, is usually the cal-
culation of the mass transfer coefficient, although it can be estimated by semiem-
pirical equations where dimensionless modules are used. We can relate k C to the
Sherwood module, as follows:
(k d )
Sh = C P (9.26)
DA
where DA = coefficient of molecular diffusion, which must be known or esti-
mated by semiempirical equations, dp = effective diameter of the particle, that is,
the diameter of a sphere that has the same outer area.
The Sherwood module is related to the Reynolds and Schmidt numbers, by
means of the parameter jD :
Sh
jD = 1 (9.27)
Sc 3 ⋅ Re
where
dp ⋅ 𝜌 ⋅ u
Re = (9.28)
𝜇
and
𝜇
Sc = (9.29)
𝜌 ⋅ DA
where “𝜇” is the superficial rate, calculated as the flow/total section of the reac-
tor.
For fixed-bed and for type A → B reactions in the absence of inert, the value
of jD depends mainly on the Reynolds modulus. There are numerical expressions
that relate jD to the Reynolds module, such as the Thodos equation:
0.725
jD = 0.41 (9.30)
Re − 0.15
With all these relations, we can estimate the value of the coefficient k C .
Solution
The basic equation of this system for calculating the reaction rate is given in
Eq. (9.25).
Let us estimate the mass transfer coefficient k C . For that, let us use the corre-
lations in Eqs. (9.27) and (9.30).
We know the value of the Reynolds number, so we can estimate jD = 4.913 using
Eq. (9.30). The value of Sc can also be calculated using Eq. (9.29):
0.094
Sc = = 1.483
0.66 ⋅ 0.096
Using Eq. (9.27), we can calculate Sh = 0.2913.
From the value of the Reynolds number, we can estimate the particle diameter:
kg
dp ⋅ 𝜌 ⋅ u dp ⋅ 56.47 m2 h
Re = → 0.052 =
𝜇 kg
0.094 m h
Therefore, dp = 8.6⋅10−5 m. The definition of Sh (Eq. (9.26)) can be used
to calculate the value of the mass transfer coefficient, obtaining a value of
k C = 323.0 m/h.
The value of the difference in concentrations (C Ag − C AS ) can be then calcu-
lated:
g
(−rA ) = − A = kC ⋅ (CAg − CAs )
Sex
kmol 1 kgcat m kmol
0.153 ⋅ = 323.0 ⋅ (CAg − CAs ) 3
kgcat h 45 m2cat h m
So, (C Ag − C As ) = 1.05⋅10−5 kmol/m3 = 1.05⋅10−2 mol/m3
The value of C Ag can be also calculated, using the ideal gas equation:
C Ag = PA /RT = 1.322 mol/m3 , so the relative difference is (C Ag − C As )⋅100/
C Ag = 0.79%.
counterdiffusion and this only takes place in the X direction, it is fulfilled that
nA = − nB and the above equation becomes:
dxA
nA = −c ⋅ DAB (9.32)
dX
and if the concentration is constant:
dCA
nA = −DAB (9.33)
dX
In multicomponent systems, the treatment is complicated because the term
of global displacement is not negligible, and the diffusivities may depend on the
composition. In gaseous systems, in the presence of a large amount of inert at rest,
and in liquid systems, Eqs. (9.32) and (9.33) are applicable. Note that the units
of the diffusion coefficient, by definition, are m3 fluid /(mcat s), which are usually
indicated as m2 /s.
The value of DAB varies with temperature and pressure in different ways
depending on the type of diffusion. In porous catalysts, and for a process A ↔ B,
or in systems with a large proportion of inert, the diffusion law is applied as
dCA
nA = −De (9.34)
dX
nA being the density of flow, defined as the kmol of A/s that are diffused per unit
of total normal area, including that occupied by the solids. De is the effective dif-
fusion coefficient that takes into account the free passage section of the holes.
Usually, De is related to the coefficient DAB by relations of the type:
DAB ⋅ 𝜀
De =
𝜑
where 𝜀 is the fraction of holes of the catalyst (porosity, m3 holes /m3 cat ) and 𝜑 is the
tortuosity factor (frequent value 3 or 4 or 1/𝜀).
x=0
2L
A A
0.9
0.8
0.7
0.6 mL = 0.5 mL = 1
CA/CAS
mL = 2 mL = 5
0.5
mL = 10 mL = 20
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/(l)
Exterior surface Center of the catalyst
V
∫0 (−rA ′′′ )dV L
(−rA ′′′ ) V
∫0 (−rA ′′′ )dL
𝜂= = =
(−rA ′′′ )no diffusion problems k ′′′ CAs k ′′′ CAs L
( )
L dC
∫0 −De dxA dL [ ]
De dCA
= = ′′′ (9.46)
k ′′′ CAs L k CAs L dx x=0
where (−rA ′′′ )and (−rA ′′′ ) are expressed in (kmoles of A/(m3 cat s)), V is the volume
of the catalytic particle, and Stotal represents the total surface through which there
is input of reagent A (in the case considered of a laminar flat catalyst with entry
of reagent by both sides, Stotal = 2S). Note the difference between the average real
rate (−rA ′′′ ) , which can be determined experimentally, and the rate (−rA ′′′ ) in
each layer of the porous particle.
Then:
(−rA ′′′ ) = k ′′′ CAs 𝜂 (9.47)
which is a very comfortable expression to work with, since the efficiency factor
modifies the reaction rate according to the effect of diffusion.
Taking the equation of the concentration profile deduced before (Eq. (9.45)) we
obtain
dCA CAs
= (−m) sinh(m(L − x)) (9.48)
dx cosh(mL)
which for x = 0 provides
]
dCA −m CAs
= sinh(mL) = −m CAs tanh(mL) (9.49)
dx x=0 cosh(mL)
Introducing this equation in (9.46), we obtain
tanh(mL)
𝜂= (9.50)
mL
This function can be approximated in two different regimes:
– When there is no resistance to diffusion in the pellet, (Φ = mL < 0.4), 𝜂 ≅ 1,
and
(−rA ′′′ ) = k ′′′ CAs (9.51)
– When resistance to diffusion is important, (Φ = mL > 4), 𝜂 ≅ 1/mL, and
1
k ′′′ CAs C (k ′′′ ⋅ De ) 2
′′′
(−rA ) = 𝜂 = As (9.52)
mL L
At first sight, it could seem that the diffusion in the pellet can be considered as
a corrective multiplicative factor of any of the previous cases, using 𝜂. But this is
not the case, since this factor considers both the reaction term on the surface and
the diffusion term.
products on one side, cylindrical pellets, and spherical particles. In the case of
the permeable single porous pellet on one side, the solution is exactly the same
as the problem solved previously but considering “L” at the total thickness.
In the case of cylindrical particles, represented in Figure 9.6, a mass balance of
the reactive A would be
Input − Output + Accumulation = Generation
The function J 0 is called the modified Bessel function of the first kind of order
zero. “R” represents the total radius of the cylindrical particle. Let us call L = R/2,
and then:
CA (r) J0 (2 ⋅ m ⋅ L ⋅ r)
= (9.60)
CAs J0 (2 ⋅ m ⋅ L)
As we have defined the effectiveness as the quotient between the actual reaction
rate and those we would have if no diffusional limitations are present:
(−rA ′′′ )
𝜂= (9.46)
(−rA ′′′ )no diffusion problems
We can then write
V V
∫0 (−rA ′′′ ) ⋅ dV ∫0 k ′′′ ⋅ CA (r) ⋅ dV
(−rA ′′′ ) = = (9.61)
V V
Then, for calculating the effectiveness factor we must do
V
∫0 k ′′′ ⋅CA (r)⋅dV V J0 (2⋅mL⋅r)
∫0 k ′′′ ⋅ CAs ⋅ ⋅ dV J (2 ⋅ mL)
V J0 (2⋅mL) 1
𝜂= = = ⋅ 1
k ′′′ ⋅ CAs V⋅ k ′′′ ⋅ CAs m ⋅ L J0 (2 ⋅ mL)
(9.62)
The solution of the integral of the Bessel function can be consulted in differ-
ent books treating mathematics. If we do the balance in a spherical geometry,
following the same procedures, the expression for the effectiveness factor is
V
(−rA ) ∫0 k ⋅ CA (r) ⋅ dV
𝜂= = V
(9.63)
(−rA )no diffusion problems ∫0 k ⋅ CAs ⋅ dV
Evaluation
( of the)integral in the denominator is straightforward, and obviously
4
equals 3 kCAs 𝜋R3 . The integral in the numerator is more complicated to eval-
uate and it is easier to calculate the observed rate from the flow into the pellet. A
balance of the reactant over the whole spherical pellet can be written as
V ( )
dC
k ⋅ CA (r) ⋅ dV = −(−De 4𝜋R )
′′′ 2
(9.64)
∫0 dr r=R
That is, the observed reaction rate equals the rate of diffusion from the pellet
surface into the pellet. The minus sign outside the parenthesis is needed because
228 9 Catalytic and Multiphase Reactor Design
–2
–4
% difference
–6
Cylinder
–8 Sphere
–10
–12
–14
0 5 10 15 20
Thiele modulus (m l)
Figure 9.7 Error implied in the use of the approximation Eq. (9.50) instead of Eq. (9.62)
(cylinder) and Eq. (9.65) (sphere).
the flow into the pellet has a negative radial direction. Differentiation is generally
easier to do than integration, so from Eq. (9.64) we can get
( )
V
∫0 k ′′′ ⋅CA (r)⋅dV
V 1 1 1
𝜂= = ⋅ − (9.65)
k ′′′ ⋅ CAs m⋅L tanh(mL) 3 ⋅ mL
On this occasion, for the spherical particles we must use L = R/3.
As we see, the use of these characteristic lengths greatly simplifies the calcula-
tion of the efficiency factors. One more approximation is usually done frequently:
since Eqs. (9.50), (9.62), and (9.65) give approximately the same values of 𝜂 for a
given value of (m⋅L), the simplest Eq. (9.50) is taken as valid, making for a rea-
sonably low error. In Figure 9.7, we can check the relative error implied in the use
of Eq. (9.50) instead of the corresponding equation for cylindrical and spherical
geometries. As we can see, the maximum error is approximately −14% at inter-
mediate values of (m⋅L). Errors are negative, i.e. the corresponding effectiveness
factor calculated using (9.50) is higher than that of Eqs. (9.62) and (9.65).
In this way, we see that the effectiveness factor for different geometries can be
calculated in a totally equivalent form to the case of plane geometry, only taking
care to choose the correct value of L. Actually, for a given catalyst with arbitrary
geometry, or for a catalyst of various mixed forms, the use of Eq. (9.50) is extended
by particularizing the characteristic length using the relation:
Volume of the particles
L=
External surface available for entry and difussion of the reagent
(9.66)
As mentioned before, this is valid only for first-order kinetics, which was
assumed for deducting the equations. If we pretend to generalize the Thiele
9.5 Reaction Rate When Internal Diffusion Is Slow 229
modulus in such a way that (Eq. (9.50)) continues being correct, we must use the
following definition:
(−r′′′ As ) ⋅ L
Φ=m⋅L= [ ]1∕2 (9.67)
C
2De ∫C As (−r′′′ A ) ⋅ dCA
Ae
Solution
(a) In the data, it is said that a feed rate of 60 h−1 produces a conversion of
X A = 0.5. The average reaction rate would be given by
As the Weisz modulus is higher than 4, the influence of the internal diffusion
is quite important.
(c) In order to calculate the efficiency factor, as the internal diffusion is impor-
tant, we can do
tanh(mL) 1
𝜂= ≅
mL mL
k ′′′ ⋅C
(−r′′′ A ) ⋅ L2
As
⋅ L2
We = = mL = 𝜂 ⋅ (mL)2 ≅ mL
De ⋅ CA De ⋅ CA
9.5 Reaction Rate When Internal Diffusion Is Slow 231
where m2 = k ′′′ /De . And then, with the value of the Thiele’s modulus:
1
𝜂 ≅ 6.2 = 0.161
(d) We would like to estimate the kinetic constant, so:
(−rA′′′ )
(−rA′′′ ) = k ′′′ ⋅ CA ⋅ 𝜂 i.e. k ′′′ =
CA ⋅ 𝜂
mol
0.122 cm3 reactor h
= mol
= 1.05 ⋅ 105 h−1
0.6 ⋅ 10−5 cm 3
⋅ 0.161
Solution
From the data at the experimental tank reactor, we can calculate the value of the
kinetic constant, as in the CSTR we have
n − nA
nA0 − nA = (k ′ ⋅ CA ) ⋅ W i.e. k ′ = A0
CA ⋅ W
The inlet molar feed rate can be calculated from the data:
P⋅Q 1 atm ⋅ 5 l∕s mol
nA0 = = = 0.1
R⋅T 0.082 atm l
⋅ 606 K s
mol K
mol
nA = 0.1 ⋅ (1 − 0.7) = 0.03
s
nA 0.03 mol
CA = = = 0.006
Q 5 l
nA0 − nA (0.1 − 0.03) l m3
′
k = = = 11.66 = 1.166 ⋅ 10−3
CA ⋅ M 0.006 ⋅ 10 s gcat s gcat
For the bed of big particles, considering a PFR:
dn
rA = A , i.e. − (k ′ ⋅ CA ⋅ 𝜂) ⋅ dW = −nA0 ⋅ dX A
dM
+ (k ′ ⋅ CA0 (1 − XA ) ⋅ 𝜂) ⋅ dW = +Q ⋅ CA0 ⋅ dX A
XA final
Q
W= ⋅ dX A
∫0 (k ′ ⋅ (1 − XA ) ⋅ 𝜂)
tanh(mL)
The value of the efficiency factor can be estimated: 𝜂 = mL
, with values:
√
k ′′′
m= and L = R∕3
De
232 9 Catalytic and Multiphase Reactor Design
Solution
First, we can estimate the rate of external diffusion of the reactive:
mol
rA = kc (CAg − CAs ) ≅ kc CAg = 300 ⋅ 20 = 6000
h m2cat
This rate represents the maximum transfer through the external film in the
catalytic system. This is quite high and, at first sight, probably there will not be
external diffusion problems, but this must be compared to the other steps.
The Weisz modulus can be calculated to estimate the internal diffusion prob-
lems. For that, the reaction rate should be expressed in the correct units:
( ( )2 ) ( ( )2 )
dp 2.4 ⋅ 10−3
gA = rA ⋅ Sp = rA ⋅ 4𝜋 = 0.042 ⋅ 4𝜋
2 2
= 7.6 ⋅ 10−7 mol A∕h
g g mol A
r′′′ A = A = ( (A ) ) = … = 105
Vp 4 dp 3 h m3c
3
𝜋 2
The concentration in the catalyst surface is the same as that in the bulk fluid if
no problems of external diffusion are present. Then:
Solution
(a) The design equation for a PFR is given by
Solution
In this case, the reaction rate is modified by the internal diffusion, so:
(−r′ A ) = k ′ ⋅ 𝜂 ⋅ CA2
9.5 Reaction Rate When Internal Diffusion Is Slow 235
( ) ( )1
Rp 3 2
m⋅L= ⋅ k ′′′ ⋅ CA ⋅
3 2 ⋅ De
tanh(mL)
𝜂=
mL
The problem is very similar to that presented in Example 9.1, and just the pre-
vious equations must be included. A valid program for that calculation is given.
For changing the units of the constants, the following relation will be used:
k ′′′ (m3L ∕(m3cat s)) = k ′ (m3L ∕(kgcat s)) ⋅ 𝜌cat (kgcat ∕m3cat )
u0=3;% m/s
porosity=0.4;
dcat=2000;% kg/m3 catalyst
k0=7.707e5;% m6/mol ⋅ kg ⋅ s
E_R=8656.1; % K
P=500000; % Pa
Ca0=P/(8.314*523); % mol/m3
T=250+273; % K
k=k0*exp(-E_R/T); % m6/mol ⋅ kg ⋅ s
for i=2:n
L(i)=L(i-1)+incx;
XA(i)=XA(i-1)+(dcat*(1-porosity)*k*Ca0*(1-XA(i-
1))∧ 2/(u0*(1+2*XA(i-1))∧ 2))*incx;
CA=Ca0*(1-XA(i))/(1+2*XA(i));
mL=(dp/6)*(k*dcat.*CA*3/(2*De)).∧ 0.5;
effectivity=tanh(mL)./mL
end
236 9 Catalytic and Multiphase Reactor Design
figure(1)
plot(L,XA)
xlabel(’Length (m)’)
ylabel(’Molar conversion’)
The variation of the conversion with length is given in the following figure. It
is quite similar to that presented in Example 9.1, but the conversion increases
slowly. This is due to the correction of the constant done by the effectivity, which
is approximately 0.5 in this design (it can be checked along the program; Program
listing 9.3).
0.9
0.8
0.7
Molar conversion
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Length (m)
For the last equality, we have added the numerators and the denominators of
the previous quotients. Note that the final equation has two summands in the
denominator,
( )corresponding to the resistance to the diffusion through the exter- ( )
nal layer k1 and that of the combined step of pore diffusion and reaction k𝜂1 .
C ( )
In the case that k C is infinity (no resistance to external diffusion), k1 = 0 and
C
the reaction rate is given by (9.74). Contrarily, if k𝜂 is infinity, Eq. (9.75) equals
Eq. (9.25) with C As = 0 (logically, as the diffusion and reaction is infinitely fast).
Solution
For calculating the effectiveness factor, both the diffusion and the chemical reac-
tion rate are needed at the system temperature; then:
{ ( )}
E 1 1
k ′ 623K = k ′ 500K exp − = 105 cm3 ∕(gcat s)
R 623 500
238 9 Catalytic and Multiphase Reactor Design
2·L
Alumina washcoat
Monolith substrate
Solution
First of all, we will assume there is no influence of the external diffusion on the cat-
alyst. For estimating the effect of internal (pore) diffusion, we will need the value
of the kinetic constant at the temperature of the reactor: k ′′′ (623 K) = 19.63 s−1
Bearing in mind that the characteristic length of this catalytic system is
−4
L = 5.5⋅10
2
m, we can write
√
19.63 s−1 5.5 ⋅ 10−4
∅=m⋅L= = 0.803
2.3 ⋅ 10−6 m2 ∕s 2
This value of the Thiele modulus can be used to calculate the efficiency factor:
tanh(mL)
𝜂= = 0.83
mL
In these conditions, the reaction rate would be calculated with:
(−r′′′ NO ) = k ′′′ ⋅ CNO ⋅ 𝜂 = 15.773 ⋅ CNO
In the system, C NOin is the input concentration of NO to the reactor, i.e.
C NOin = 1000 mg/Nm3 .
We will assume the reactor is plug flow, so in this reactor we can do a mass
balance in a differential volume:
Q·CNOin·(1 – XNO)
dV
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
Volume (m3)
11 cells
11 cells
1.35 m
9.7 Monolithic Catalytic Reactors 241
Calculate the amount of catalyst needed and the molar flow rate profiles of all
species through the reactor. Data: Molecular weights: CO2 = 44, C2 H4 O = 44,
H2 O = 18, O2 = 32.
Solution
Initially, we will assume the plug flow reactor for the gas and that the catalyst is
fixed in a bed. The entering currents will depend on the conversion of the reactor,
which must produce
kg 1 kmol 1 h mol
nBfinal = 100 ⋅ = 0.6313 ethylene oxide
h 44 kg 3600 s s
The high value of velocity of the gases through the bed would indicate that there
are no problems of external diffusion, we will have
−r1 = k1 CA 𝜂 = k1 CA0 (1 − XA ) 𝜂
where “A” refers to oxygen. For the effectiveness:
√
tanh(mL) ( ) k ′′′ i
R
𝜂= m⋅L= ⋅
mL 3 De
with i = 1, 2. The value of L is related to the spherical particles and is calculated
using their radius. For the first reaction:
( ) √
75 ⋅ 10−6 0.255
m⋅L= ⋅ = 0.01 85
3 4.64 ⋅ 10−7
and then the effectiveness is approximately unity. The same is true for the sec-
ond reaction, so internal diffusion does not interfere with the reaction. This is an
expected result, as the catalytic particles are so small that it is predictable not to
have internal diffusion problems.
Using the plug flow approximation, in a differential volume of catalyst
dnA = (−r1 − r2 ) ⋅ dVcat
dnA = (−k ′′′ 1 CA0 (1 − XA ) − k ′′′ 2 CA0 (1 − XA )) ⋅ dVcat
dWcat
nA0 ⋅ (−dX A ) = (−k ′′′ 1 CA0 (1 − XA ) − k ′′′ 2 CA0 (1 − XA )) ⋅
𝜌cat
(−dX A ) C ⋅ (−k ′′′ 1 − k ′′′ 2 ) dWcat
= A0 ⋅
(1 − XA ) nA0 𝜌cat
nA0 𝜌cat
dX = dWcat
CA0 (1 − XA )(k ′′′ 1 + k ′′′ 2 ) A
This last equation is the one we must integrate through the reactor (increasing
the mass of catalyst) and it will give the change in the molar conversion through
the bed.
The molar flow rate of oxygen (“A” in the former equations) can be calculated
from the needed production of ethylene oxide. In this way, we can relate the molar
flow rates of all species:
Oxygen → nA = nA0 (1 − X A )
Ethylene oxide → nB = 2 ⋅ nA0 ⋅ X A ⋅ S1
Carbon dioxide (and water) → nC = 23 nA0 (1 − XA ) ⋅ (1 − S1 )
9.7 Monolithic Catalytic Reactors 243
%Calculation
L=R/3;
m1=sqrt(k1/De);
m2=sqrt(k2/De);
mL1=m1*L;
mL2=m2*L;
Fef1=(tanh(mL1)/mL1);
Fef2=(tanh(mL2)/mL2);
CA0=(P*Cox)/(Rg*T) %mol/m3
nBFinal=Q*(1000/3600)/MW ; %mol/s
S1=k1/k2;
nA0= nBFinal/(2*S1*0.4)
Mcat=0;
n=1000; % Number of points to be used in the integra-
tion
incXA=0.4/(n-1);
% Initial values
XA(1)=0;
244 9 Catalytic and Multiphase Reactor Design
for i=2:n
XA(i)=XA(i-1)+incXA;
dWcat=-(((1-XA(i))*nA0*denscat)/(CA0*(-k1*Fef1-
k2*Fef2)));
Wcat(i)=Wcat(i-1)+dWcat*incXA;
nA(i)=nA0*(1-XA(i));
nB(i)=nA0*XA(i)*S1*2;
nC(i)=(2/3)*nA0*(1-S1)*XA(i);
end
figure(1)
plot(Wcat,XA)
xlabel(’Weight of catalyst, kg’)
ylabel(’Molar conversion of Oxygen’)
figure(2)
plot(Wcat,nA,’o’,Wcat,nB,’-.’,Wcat,nC)
xlabel(’Weight of catalyst, kg’)
ylabel(’Molar flow of the different compounds, mol/s’)
legend(’Oxygen’,’Ethylene oxide’,’Carbon dioxide’)
This program uses the finite difference method (first order) to do the integra-
tion of the ordinary differential equation. In the program, we can choose the
number of increments, and the corresponding increment of molar conversion
is calculated, and used. The following figures are generated in this program:
0.45 1.2
Oxygen
Molar conversion of oxygen
0.4
Molar flow od the different
Ethylene oxide
1 Carbon dioxide
0.35
compounds (mol/s)
0.3 0.8
0.25
0.6
0.2
0.15 0.4
0.1
0.2
0.05
0 0
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Weight of catalyst (kg) Weight of catalyst (kg)
The total amount of catalyst needed in this example is W = 81.2 kg. If the
reaction were to take place in a fluidized bed, bubbles must be considered,
and modification of the previous calculation will be necessary. In this sense,
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 245
Since the substances are in two distinct phases, they must be contacted before
reacting, and the overall expression of the reaction rate must consider both the
rate of diffusion of the reagents and the rate of the chemical reaction.
The reaction will proceed in one phase or in another depending fundamen-
tally on the solubility of the reactive components in each phase, since solubility
determines the displacement of the reactants between the two phases.
Consider that reagent A is initially in phase I, and that reagent B is in phase II.
Reagent A is more soluble in phase II, than reagent B in phase I, so the reaction
will take place mainly in phase II. In this sense, the reaction we will work with is
A (phase I → phase II) + bB (phase II) → rR (solid, liquid or gas)
where “b” is the stoichiometric coefficient of reactant B, and the reaction rate
(−rA ) in phase II can have any of the expressions developed and proposed for
homogeneous reactions. We will assume that the chemical reaction occurs with
first-order kinetics with respect to both reagents, so:
( )
mol
(−rA′′ ) = k ′′ ⋅ CA ⋅ CB (9.76)
s m3L
the units of the constant k′′ will be m3 L /(mol s).
To facilitate the nomenclature, phase I will be considered as a gas and phase II as
liquid. If the system were a liquid–liquid one, it would be necessary to replace the
nomenclature used by the appropriate one. The solubility of reagent A in phase II
is given by the expression (Henry’s law if phase I is a gas and phase II is a liquid):
pAi = HA ⋅ CAi (9.77)
The partial pressure of component A in phase I will be designated pA and C A is
the molar concentration of A (kg/m3 or kmol/m3 ) in phase II. The units of the sol-
ubility ratio H A will depend on the concentration units. The subscript “i” refers to
the fact that the equilibrium relation is fulfilled at the interface. For liquid–liquid
systems, it would be necessary to substitute pA for concentrations and the ratio
of solubility for the adequate one.
Several models have been proposed to describe the phenomenon that occurs
when the two fluid phases come into contact. The most used model is the
two-film theory proposed by Lewis–Whitman that proposes the existence of
different zones along the total volume of reaction that acts as a set of resistance
in series with the movement of the reagents. An outline of this model is observed
in Figure 9.10, considering the two fluids moving parallel to the interface.
The different areas in the outline are as follows:
– Resistance to the transfer of reagent A through phase I, equivalent to the resis-
tance that would oppose by molecular diffusion that of a layer adhered to the
interface;
– Resistance through the interface (often negligible);
– Resistance to the transfer of reagent A from the interface to the reaction zone,
equivalent to that which would oppose a film where the transport regime is
molecular diffusion in zone II;
– Resistance to the transfer of reagent B through phase II to the reaction zone.
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 247
Phase I Phase II
Component A transfer
Component B transfer
Bulk gas Gas film Liquid film Bulk liquid Bulk gas Gas film Liquid film Bulk liquid
Reaction plane
Reaction plane
pA CAi pA
CB CB
Bulk gas Gas film Liquid film Bulk liquid Bulk gas Gas film Liquid film Bulk liquid
Reaction zone
pA CAi pA
CB CB
CA
pAi
(Case C) (Case D)
x0
(c) (d)
Figure 9.11 Concentration of components being transferred between phases. The interface is
indicated by a red line. The boundaries between the well-mixed and the interfacial films are
indicated by dotted lines.
(m2 i ) must be used, which is the factor that influences the diffusion, and not the
volume of reaction. The expression of the rate equation is
( )
kmol A r D dC D (C − 0)
−rA 2
= − B = kAg (pA − pAi ) = Al A = Al Ai
s mi b dx x
DBl dCB D (C − 0) 1
= = Bl B ⋅ (9.78)
b ⋅ dx x0 − x b
where k Ag is the individual mass transfer coefficient of reagent A in phase I (for
example, in kmol/(s m2 i atm)), and DAl and DBl are the molecular diffusivities of
reagents A and B respectively in phase II (in m2 /s).
The relationship between pAi and C Ai is given by Henry’s law, whether it is
between pressures and concentrations for gas–liquid reactions, or Nerst’s law if
it is a relationship between concentrations for fluid–fluid reactions, as shown in
Eq. (9.77).
In the (9.78), the diffusion coefficients have been used according to the double-
layer model. The relationship between the diffusion coefficient and the individ-
ual coefficient of mass transfer in the same phase, according to the theory of the
film, is
D
kAl = Al (9.79)
x0
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 249
Keep in mind that in the previous expressions, the solution is direct if the value
of H A does not depend on the concentrations. Otherwise, it must be solved by
trial and error. A very important aspect that should not be ignored is the reference
to the fact that Eq. (9.86) is valid as long as C B becomes zero within the film, that
is, it is in case A of Figure 9.11. If it is not, when in case B of Figure 9.11, the
reaction rate is calculated as
(−rA ) = kAg ⋅ pA (9.87)
250 9 Catalytic and Multiphase Reactor Design
If there were no reaction, i.e. in the case of a pure absorption system, we would
have x0 = x in the scheme of Figure 9.11a, with which, from Eq. (9.82), the rate of
the overall process would be given by
(−rA ) = kAl (CAi − 0) (9.89)
Therefore, the rate increase factor is defined by the instantaneous reaction Ei ,
which acts by producing a greater concentration gradient, which is the driving
force of diffusion, such as
( )
reaction rate with instantaneous reaction DBl ⋅ CB
Ei = = 1+
mass transfer rate b ⋅ DAl ⋅ CAi
(9.90)
So the rate could be expressed as
(−rA ) = kAl CAi Ei (9.91)
If we compare this expression with Eq. (9.82), we verify that the equality
Ei = x/x0 is met, that is, the factor Ei gives us an idea of how the diffusion layer
is enlarged by including the reaction.
In addition to these cases, different situations can be found. Intermediate cases,
apart from that shown in Figure 9.11 can be found, depending on the relative rate
of diffusion and rate.
For liquids, the value of diffusivities are on the order of 105 cm2 /s, and x0 can
take values of 10−2 cm.
If MH2 is much greater than unity, it means that the transfer is very fast, which
in turn means that the chemical reaction rate is very high and that the overall rate
will depend only on the diffusion, so the reaction is produced entirely in the film
and the surface of it is the limiting factor of the rate. From this development, it is
deduced that if it wants to favor the reaction, it must be carried out by means of
a contact device that develops a high interface area. In these cases, the agitation
is an important factor when increasing the contact area by homogenizing the
mixture, e.g. the columns of rain or dishes.
On the contrary, if MH2 is very small, it means that the chemical reaction is very
slow, and that diffusion is not the limiting factor. Therefore, to favor the reaction,
all that is needed is a large volume of reaction, e.g. bubble columns.
More precisely, it has been found that:
• If MH2 is greater than 4, the reaction takes place in the film and cases A or B
considered in Figure 9.11 would be taken.
• If the value of MH2 is between 0.0004 and 4, the situation corresponds to the
intermediate case C.
• If MH2 is much less than 0.0004, the reaction is infinitely slow and corresponds
to the case D. The reaction would occur in the liquid core although the main
resistance can be anywhere, i.e. in the gas film or in the liquid, but in any case,
is negligible.
where the rate (−rA′′ ) is based on the reactor volume (mol/(mR 3 s)), and the value
of “a” is the ratio (surface of the interface)/(reactor volume) (mi 2 /mR 3 ). f L rep-
resents the liquid fraction of the reactor, i.e. the amount of liquid in the reactor
(m3 L /m3 R ).
The first summing of the denominator represents the resistance to the transfer
in the gaseous layer. In this, k Ag has units of mol/(s mi 2 atm) (assuming a gaseous
phase) and a of mi 2 /mR 3 , so the summing has the units of ((atm s mR 3 )/mol).
The second addition takes into account the resistance in the liquid phase and
the third, the chemical reaction. This reaction is supposed to be first or pseudo
first order, according to the equation: (−rA′′ ) = (k ′′ CB )CA . The units of the product
(k ′′ C B ) ⋅ f L are m3R ∕(mol s) ⋅ (mol∕m3L ) ⋅ m3L ∕m3R = 1∕s, or other equivalent units.
252 9 Catalytic and Multiphase Reactor Design
1000
Reaction zone approaches the interface 1000
100 100
50
Ei
E
20
10 10
2
1
0.1 1 10 100 1000
MH
Note that this equation is nonlinear and must be solved by iterative methods.
In Figure 9.12, the equation of a graphic form is shown.
Solution
( )
mol CO2
(−rA ) = kAg ⋅ a ⋅ (pA − pAi ) = kAg ⋅ a ⋅ (pA − HA ⋅ CAi )
hl
( )
mol CO2
(−rA ) = kAl ⋅ a ⋅ (CAi − CA )
hl
Both rates must be equal in the stationary regime, so we can derive the equation
for the C Ai :
kAg ⋅ pA + kAl ⋅ CA
CAi =
HA ⋅ kAg + kAl
Then it can be used for estimating the reaction rate:
( ) ( )
mol CO2 kAg ⋅ pA + kAl ⋅ CA p − HA CA
(−rA ) = kAl ⋅ a ⋅ − CA = A1
hl HA ⋅ kAg + kAl H
+ A kAl ⋅a kAg ⋅a
pA −30CA
(a) So, in the first case: (−rA ) = 1
+ 30
= 0.8247 (pA − 30CA )
80 25
(b) Now, C A = 0 and then: (−rA ) = 0.8247pA
Solution
Note that in this case, two liquid phases are involved and the units of the coeffi-
cients can change. For example, the H A represents the equilibrium in the inter-
face, so:
CA,phase I m3phase II
HA = … in
CA, phase II m3phase I
From the general rate equation:
pA
(−rA′′ ) = 1 HA H
k ⋅a
+ k ⋅a⋅E + (k ′′ ⋅CA )⋅f
Ag Al B L
254 9 Catalytic and Multiphase Reactor Design
HA 3 s
= = 0.00 108
kAL ⋅ a ⋅ E 500 ⋅ 18.48 m
So, in this case the controlling step is the diffusion through phase I.
Data:
Solution
We will assume that the membrane is a liquid film. The rate of the process will be
given by the general rate equation (Eq. (9.94)). In the example, we have
pO2 = 0.21 atm (air)
a (m2interface ∕m3system ) = 7∕1 = 7 m−1
DO2 ,L 7.1 ⋅ 10−6 m
kO2 ,L = = = 1.42 ⋅ 10−3
film thickness 5 ⋅ 10−3 s
Using these data, we can estimate the value of the Hatta modulus and then the
increment factor “E”:
√
DO2 ,L ⋅ k ′′ √
7.1 ⋅ 10−6 ⋅ 1.8 ⋅ 103
MH = = = 79.6
kO2 ,L 1.42 ⋅ 10−3
DRH,L ⋅ CRH ⋅ HO2 8.3 ⋅ 10−6 ⋅ 5 ⋅ 0.937
Ei = 1 + =1+ = 27.08
b ⋅ DO2 ,L ⋅ pO2 1 ⋅ 7.1 ⋅ 10−6 ⋅ 0.21
These values of the parameters suggest that the reaction is being produced in a
regime where both diffusion and reaction have similar rates. Using the relation:
√
E −E
MH2 Ei −1
(√ )
i
E= (9.96)
E −E
tanh MH2 Ei −1
i
We can check that E = 24.6 and then calculate the reaction rate (Eq. (9.94)):
pA
(−rA′′ ) = 1 HA H
k ⋅a
+ k ⋅a⋅E + (k ′′ ⋅CA )⋅𝜀
Ag AL B L
pA 0.21
(−rA′′ ) = 1 HA HA
= 1 0.937 0.937
+ + + 1.42⋅10−3 ⋅7⋅24.6
+ 1.8⋅103 ⋅5
kAg ⋅a kAL ⋅a⋅E (k⋅CB )⋅𝜀L 0.03⋅7
mol
= 0.024
s m3reactor
256 9 Catalytic and Multiphase Reactor Design
Actually, the third adding in the denominator is very small compared to the
other two, so the reaction is quite fast compared to diffusion. The result indi-
cates that, bearing in mind the volume of the reactor, 0.024 mol of oxygen is being
transferred per second in the whole system if the conditions are maintained along
the length.
The development of the rate equation will be based on the two-film theory,
using the nomenclature presented in Figure 9.14.
The resistances are shown graphically in Figure 9.15. The expressions of the
rate of each of the consecutive steps reflected in this figure can be expressed as
(mol/(m3 R s))
– Transfer through the gaseous film: k Ag ai (pA − H A C Ai )
– Transfer through the liquid film: k Al ai (C Ai − C Al )
– Transfer through the solid layer: k Ac ac (C Al − C As )
– Surface chemical reaction: (kA′′′ CB )𝜂A fs CAs
gas–liquid interface
Porous catalyst
particle
kC DAe, DBe
kg kL
Centre of the particle
CBl
pAg
CAi
CAl CBs CB
pAi
Bulk liquid
Bulk gas CAs
(no resistance) CA
Figure 9.15 Schematic showing the resistances involved in the gas–liquid reaction on the
surface of a solid catalyst. Source: Levenspiel (1999). Adapted with permission of John
Wiley & Sons.
If they are put in the form of resistances, and equaling, you get
( ) pA
− CAi CAi − CAl CAl − CAs CAs
mol HA
(−rA′′ ) = = = = (9.99)
s m3R 1 1 1 1
HA kAg ai kAl ai kAc ac (kA′′′ C B )𝜂A fs
For component B, the only resistances are the solid layer and the reaction, so:
( )
′′ mol CBl
(−rB ) 3
= 1 1
(9.101)
s mR +
k ⋅a ′′′
Bc c (kB ⋅C A )⋅𝜂B ⋅fs
In systems with pure liquid B and a slightly soluble gas A, it can be assumed
that the concentration of B has the same value at any point, so:
CBs = C Bwithin the particle = CBl
with C B constant, the reaction becomes of first global order (first order with
respect to A) and the previous expressions are reduced to a directly solvable
expression, i.e. Eq. (9.100) with known (kA′′′ ⋅ C B ).
On the contrary, in systems with diluted liquid B reagent and a very soluble gas
p
A, and at high pressure, it can be assumed that CAl = HAg throughout the reactor.
A
The rate becomes first order with respect to B and is reduced to
( )
′′ mol CBl
(−rB ) 3
= 1 1
(9.101)
s mR + ( pAg
)
k ⋅a Bc
′′′
c kB ⋅ H ⋅𝜂B ⋅fs
A
The first term of this equation is what is called gas resistance RG and the second
is the resistance of the liquid RL . Considering that for spheres:
ac = dp ∕6
the resistance of the solid layer is defined as
HA
Rs = (9.103)
kAc ⋅ dp ∕6
and the resistance of the chemical reaction is taken as
HA
RR = (9.104)
(kA′′′ ⋅ C B ) ⋅ 𝜂A
So that Eq. (9.100) remains as
pAg
(−rA′′ ) = R +R
RG + RL + s f R
s
Reordering:
pAg R + RR
= RG + RL + s (9.105)
(−rA′′ ) fs
In this way, a representation of the quotient between the pressure and the reac-
tion rate vs. 1/fs can give an idea of the controlling resistances: on the one hand,
the ordinate at the origin is equal to the sum of the resistances of the liquid and
gas, and the slope is the sum of RS and RR . The representation of this slope in front
of the values of dp will give an idea whether the chemical reaction or the gaseous
layer is the controlling one, because of the following:
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 259
What can be said about the controlling resistances of the process? Indicate how
they could estimate, and, if possible, determine their relative value.
Solution
Using Eq. (9.102):
pAg 1 HA HA HA
= + + +
−rA′′
kAg ai kAl ⋅ ai kAc ⋅ ac (k ′′′ ⋅ C ) ⋅ 𝜂 ⋅ f
A B A s
In this case, the value of f s (m3 solid/m3 reactor) is related to the mass of solid
used in each run. We can calculate −pAg /rA ((atm h m3 R )/mol) and 1/f s (1/g).
When plotting the data, we obtain the following:
0.12
0.08
0.06
0.04
0.02
0
0 0.2 0.4 0.6 0.8 1 1.2
1/fs (g–1)
The slope of the plot is 0.06 and represents (RS + RR ). The value at the
origin (0.0417) is related to (RG + RL ). It is clear that, in this case, (RS + RR )/
(RG + RL ) = 1.44, so the main resistance is either in the solid layer or in the
260 9 Catalytic and Multiphase Reactor Design
Solution
This is a complicated case, because all the equations mentioned are valid only for
first-order kinetics, but in this case, it is even difficult to do an approximation.
The only thing we can do is
1
( )
− 12
(−rA ) = k ⋅ CA = k ⋅ CA
′ ′ 2 ′
⋅ CA
⎡ 1
1 ⎤2
−3 ⎢ 2.35 ⋅ 10 ⋅ 2.75− 2
−3
+1 ⎥
5 ⋅ 10 2
3 ⎢⎢ ⎥ = 64.4
=
2 ⋅ De ⎥
⎣ ⎦
And then, 𝜂 A = 0.0155
In this example, pA = 1 atm in all points of the reactor, because the gas is
pure hydrogen that does not vary its partial pressure. From all data given in the
example, a reaction rate can be calculated:
mol
(−rA′′ ) = 0.0317
mR 3 s
Many textbooks classify the different types of equipment for gas–liquid reac-
tions. The classification is based on geometrical aspects, the presence or absence
of solid catalysts, and flow directions. A description of each type can be consulted
in classical textbooks such as “Chemical Reaction Engineering,” by O. Levenspiel.
Section 2
The design of the equipment is addressed first when there is no reaction, and
thus the reaction is considered. All the equations we present are based on coun-
tercurrent flow. In the case of parallel flow, a change of L to −L is required, but
the equations are valid.
The design of the tower can be done numerically or graphically. For diluted sys-
tems, C A ≪ C T and pA ≪ 𝜋, so L′ ≅ L and G′ ≅ G, and then the differential balance
can be written as
G′ L′
⋅ dpA = ⋅ dCA = (−rA ) ⋅ a ⋅ dV
pT CT
The mass transfer rate can also be calculated using global mass transfer coeffi-
cients, with driving force (pA − p∗A ) and (CA∗ − CA ), using the equations:
1 1 H
= + A
KAg kAg kAl
1 1 1
= + (9.117)
KAl kAl HA ⋅ kAg
Then, the mode balance will give
pA2 CA2
G′ dpA L′ dCA
V = =
∫ ∫
pT ⋅ kAg ⋅ a pA1 (pA − pAi ) CT ⋅ kAl ⋅ a CA1 (CAi − CA )
and also:
p C
G′ A2
dpA L′ A2
dCA
V = = (9.118)
∫ ∫
pT ⋅ KAg ⋅ a pA1 (pA − pA ) CT ⋅ KAl ⋅ a CA1 (CA − CA )
∗ ∗
9.9.3.3 Situation 3: Gas Phase in Plug Flow. Liquid Phase Completely Mixed
Here the gas contacts the liquid in the final conditions, but different rates exist in
the reactor, because for the gas, differential elements are necessary. For a differ-
ential element:
G′ dYA = −rA2
′′
]liquid at the exit conditions ⋅ dV (9.120)
9.9 Design of Multiphase Reactors 265
Integrating (9.120):
YAout
G′ dYA
V = (9.121)
∫
a YAin (−rA )
and for diluted gases, these equations can be written as
pAout
G′ dpA
V = (9.122)
pT ⋅ a ∫pAin (−rA )
A mass balance throughout the reactor provides:
G′ (YAin − YAout ) = L′ (XAout − XAin ) (9.123)
and then:
G′ L′
(pAin − pAout ) = (C − CAin ) (9.124)
pT CT Aout
In the integration, we must consider that pA varies along the column or reactor,
but C A does not.
and also we can work with the second and third terms of Eq. (9.125):
XB1
L′ dXA
V = (9.128)
∫
a ⋅ b XB2 (−rB )
The subscripts 1 and 2 refer to the extreme sections of the column (entry and
exit). A global mass balance is also fulfilled:
L′
G′ (YA2 − YA1 ) = (X − XB2 ) (9.129)
b B1
These equations are simplified in the case of diluted systems.
Note that these equations can be also applied from a fixed section of the tower
(point “1” in the equations) to an intermediate point, in such a way that:
L′
G′ (YA − YA1 ) = (X − XB )
b B1
Example 9.16 Liquid–Liquid Transfer with Reaction
A liquid–liquid reaction is carried out in a packed tower. The heaviest liquid is
fed at the top with a flow rate of 10 m3 /min and has a concentration of reagent A
of 1 kmol/m3 , while the lighter fluid is fed from the bottom with a flow of 20 m3 /s
and with a B concentration of 2 kmol/m3 . According to the reaction scheme,
1 mol of A reacts with 3 mol of B. Calculate the height of the fill tower to get that
react 90% of A, admitting an average value for the reaction rate is 5⋅10−5 kmol
B/(s m2 interface ), and knowing that the diameter of the tower is 30 cm and the area
ratio of Interface/reactor volume is 1200 m2 interface /m3 .
Solution
In this example, A + 3B → P and we want X A = 0.9, the reaction rate is
(−rB ) 5 ⋅ 10−5 kmol A
(−rA ) = = = 1.67 ⋅ 10−5
3 3 s m2interface
A change in the units is necessary:
kmol A m2interface kmol A
(−rA′′ ) = (−rA ) ⋅ a = 1.67 ⋅ 10−5 ⋅ 1200 = 0.02
s m2interface m3 s m3
In the system, both phases behave as piston flow. If we do a mole balance in a
differential volume:
(moles of A reacted in a differential volume) = (−rA′′ ) ⋅ dV
nA0 ⋅ dXA
nA0 ⋅ dXA = (−rA′′ ) ⋅ dV i.e. = dV
(−rA′′ )
where nA0 is the mole feed rate of component A:
kmol A m3 1 min kmol
nA0 = 1 3
⋅ 10 ⋅ = 0.164
m min 60 s s
9.9 Design of Multiphase Reactors 267
Kmol A
Considering that (−rA′′ ) is approximately constant and equal to 0.02 s m3
, we
can easily calculate the required volume:
XAfinal
nA0 ⋅ dXA
V = = 7.5 m3
∫0 (−rA′′ )
The required variable is the height of the tower, so: h = 7.5/(𝜋⋅(0.3/2)2 ) =
106.1 m
Solution
Section 1
Diameter = 1 m
Section 2
Air + CO2
0.02 kmol/s
0.1% mol CO2
268 9 Catalytic and Multiphase Reactor Design
First of all, let us calculate the values of MH and Ei , just to have an idea of the
controlling steps of the reaction.
√
(k ⋅ CB ) ⋅ DAl
MH = = … = 24.4
kAl
For the calculation of Ei , let us assume C Ai = C A * (concentration in equilibrium
with the maximum value of pA ); in any case, the actual value of C Ai will be lower.
( )
DBl ⋅ CB
Ei = 1 +
b ⋅ DAl ⋅ CAi
pAi 0.001 ⋅ 1 atm kmol
CAi = = atm m3
= 4 ⋅ 10−5 3
HA 25 m
kmol
( )
D ⋅C
With these values: Ei = 1 + b⋅DBl ⋅CB = … = 10 626
Al Ai
(in any case, as C Ai may be lower, the value of Ei will be higher). This indicates
that the reaction is done with both chemical reaction and diffusion interfering in
the reaction rate. Let us calculate the “E” value:
√
E −E
MH2 Ei −1
(√ )
i
E=
2 Ei −E
tanh MH E −1
i
where pA must be in atm. As we are in a case where both liquid and gas are flowing
in a piston flow pattern, we can write:
L′ dXB
G′ dYA = = (−rA′′ ) ⋅ dV
b
where Y i refers to the mole fraction of the species “i” in its phase. As the concen-
p C
trations are very low, YA ≈ PA and YB ≈ CB . Operating we can write:
T T
YA2
G ′ dYA
V =
∫
0.0206 YA1 YA ⋅ pT
We have G′ =0.02 kmol/s, and operating we find V = 2.89 m3 . For a column of
1 m in diameter (Section = 𝜋(1/2)2 = 0.785 m2 ), this means a height of 3.69 m.
Or in another way:
L′
G′ (YAin − YAout ) =(X − XBout ) = −rA′′ ]at the exit conditions ⋅ V (9.130)
b Bin
The calculation of V is direct: evaluate the reaction rate from the known compo-
sitions and solve (Eq. (9.130)).
(A lost by the gas) = (A gain by the liquid) = (–r″A)·Vdesign (A lost by the gas) = 1/b(B lost by the liquid) = (–r″A)·Vdesign
G′dYA = (–r″A)liquid in exit conditions ·dV G′dYA = (–r″A)liquid in exit conditions ·dV
Figure 9.17 Summary of cases and situations studied for the design of multiphase reactors.
pressure. The partial pressure at the exit must be 0.01 atm. The total gas flow is
45 kmol/h. The liquid enters the column at its top and flows in countercurrent
with the gas at 9 m3 /h. The input concentration of sulfuric acid is 0.6 kmol/m3 .
Consider the isothermal operation at 25 ∘ C. Determine the sulfuric acid outlet
concentration and the required interfacial area. The reaction is second order,
instantaneous, and irreversible:
2NH3 + H2 SO4 → (NH4 )2 SO4
kmol atm ⋅ m3
Data ∶ kG = 0.35 H = 0.0033
h atm m2 kmol
m
kL = 0.05 DAl = DBl .
h
Solution
Let us first draw a scheme of the system variables.
3
PA1 = 0.01 atm L = 9 m /h
CB1 = 0.6 kmol/m3
The reaction is of the type 2A + B = C, and the mole balance around all the
reaction system is
L′
G′ (YA2 − YA1 ) =(X − XB2 )
b B1
As we are working in a diluted system, the previous equation can be approxi-
mated by this:
kmol 9 mh kmol
45 (0.05 − 0.01) = (0.6 − CB2 ) 3
h 0.5 m
Then: C B2 = 0.5001 kmol/m3 . Note that this equation can be applied to any
point in the column, in such a way that a value of yA is related to the corresponding
value of C B with this equation.
Now, let us calculate the necessary reactor interface. The rate of the process
will be given by
pA
(−rA′′ ) = 1 HA HA
kAg ⋅a
+ kAl ⋅a⋅E
+ (k⋅CB )⋅fL
pA
(−rA′′ ) = 1 H
+ ( A )
D ⋅C
kAg ⋅a kAl ⋅a⋅ 1+ b⋅DBl ⋅CB
Al A
pA
(−rA ) = 1 HA
+ ( )
D ⋅C
kAg kAl ⋅ 1+ b⋅DBl ⋅CB
Al A
272 9 Catalytic and Multiphase Reactor Design
As we see, the reaction rate is a function of the pair C B –pA that are related with
the mass balance. To solve the design, we need to solve the rate equation that, in
the case of L and G in plug flow, is
G′ dYA = (−rA ) ⋅ dSi
YAout
G′ ⋅ dYA
Si =
∫YAin (−rA )
Si being the interface surface. The design can be finally calculated by means of
the following table, for calculating numerically the integral.
Finally, Si = 215.5 m2 .
Solution
Let us name hydrogen by “A” and thiophene by “B.” In this case, we have
A + B → products (−r′′ A )
= k ⋅ CA … With the reaction rate given in Kmol∕(s m3 R )
From the general equation:
pAg
(−rA′′ ) = 1 HA HA HA
kAg ⋅ai
+ kAl ⋅ai
+ kAc ⋅ac
+
(kA′′′ ⋅C B )⋅𝜂A ⋅fs
We can calculate all terms. The first sum in the denominator is zero, as the
transfer in the gas phase is quick. The second and third can be easily calculated
by
3
HA 1940 bar⋅m
kmol bar m3 s
= = 64667
kAL ⋅ ai 0.03 s−1 kmol
3
HA 1940 bar m
kmol bar m3 s
= = 3880
kAc ⋅ ac 0.5 s−1 kmol
and the last term is
3
HA 1940 bar m
kmol bar m3 s
= = 18746
m3cat kmol
(kA′′′ ⋅ C B ) ⋅ 𝜂A ⋅ fs 1.1 ⋅ 10−4
kg
⋅ 960 m3cat ⋅ 0.98
kgcat ⋅s cat
of 5% in air. To do this, the gas is introduced through the upper part of the column
next to a stream of water, creating a percolating bed in which plug flow can be con-
sidered for both phases. The elimination of SO2 is produced by oxidation to SO3
catalyzed by the active carbon surface. SO3 combines with water to form H2 SO4
that dissolves in water. In previous tests, it was found that the intrinsic reaction
rate is first order with respect to oxygen and independent of the concentration of
SO2 .
(a) Deduce the expression for the calculation of the rate of the process from fun-
damental equations of mass transfer or chemical reaction, making an outline
of the reaction stages and explaining all the terms involved in the equation.
(b) Represent qualitatively how each one of the variables of the equation evolves
within the height of the column.
(c) Write the mass balance of oxygen in a volume differential of the column.
Solution
The reaction to be considered is
SO2 (g) + O2 (g) → SO3 (g)
That is followed by the combination with water:
SO2 (g) + H2 O(l) → H2 SO4 (g)
The reaction is first order with respect to oxygen, so:
rSO2 = −k ⋅ CO2
Let us call A to the SO2 , B to the water, and C to the SO3 .
(a) Let us first make a drawing of the possible concentration profiles. For the
first reaction to occur, available catalyst is needed. The concentration pro-
files would be like those presented in the next figure, although the relative
position of the absolute values of the different species could be different. For
convenience, one part of the particle represents the gas–solid phenomena
and the other part the liquid–solid reaction, but obviously all particles are
reacting at the same time to produce SO3 and H2 SO4 .
Gas-solid interface
Porous catalyst
particle
kc DAe, DO2e
kg
Bulk liquid
CCc
pAg
CAi H2SO4
CCs
pO g
2
CO2s CCi
pAi CO c
2
Bulk gas These
pO i CAs
2 values
CAc are zero
9.9 Design of Multiphase Reactors 275
SO2
O2
H2SO4
SO3
Height
Solution
Assuming a mixed flow pattern, the mass balance can be expressed as
L′
G′ (YAin − YAout ) = (X − XBout ) = −r′′ A ]final conditions ⋅ V
b Bin
(I) (II) (III)
276 9 Catalytic and Multiphase Reactor Design
NO
Is term (I) equal to term (III)?
YES
END
In the case the flow of the fluids follows a plug flow pattern, the mass balance
can be only done in a differential volume; therefore:
G′ dYA = (−r′′ A ) ⋅ dV
From this equation:
YAout
dYA
G′ =V (*)
∫YAin (−r′′ A )
And, on the other hand, we can write the balance over the complete system:
L′
G′ (YAin − YAout ) = (X − XBout ) (**)
b Bin
The procedure to calculate the exit conditions would be as follows:
1. Take values of Y A from Y Ain to Y Aout .
2. Calculate X Bout using Eq. (**).
3. Integrate Eq. (*) using the tabulated values.
4. If the calculated volume equals the given one, then stop the calculation.
5. If the calculated volume is not equal to V , repeat the calculation with a new
value of Y Aout .
rate, show that the kinetics is described in terms of the volumetric concentrations
C A and C B by
( )
C
kB CA − KB
(−r′′ A ) = k
1
K
+ kB
A
where −rA ′′ = moles of A that react per unit area of catalyst per unit time, k A ,
k B = mass transfer coefficients of A and B, and K = equilibrium constant of the
reaction
Solution
Assuming the following reaction rate constants:
k1
−−−−−−−−→
A← −− B
k2
XA = 0.5
Solution
The kinetics in this system are
(1 − XA )
(−r′ A ) = k ′ CA = k ′ CA0
(1 + 𝜀A XA )
As we can see, the denominator accounts for the expansion of the gas through
the reactor assuming constant pressure. The expansion coefficient, 𝜀A , can be cal-
culated from the stoichiometric coefficients of the reaction, and is 𝜀A = 4 − 1 = 3.
The performance equation for a plug flow reactor is
(−r′ A ) ⋅ dW = nA0 ⋅ dXA
nA0 being the moles of “A” at the input of the system per unit time, and dW the
weight of the catalyst in the volume differential. And then:
XA
nA0 ⋅ dX A XA
dX A
W= = nA0
∫0 ′
(−rA ) ∫0 k ′ C (1−XA )
A0 (1+𝜀 X )
A A
n XA
(1 + 𝜀A XA ) ⋅ dX A
= ′ A0
k ⋅ CA0 ∫0 (1 − XA )
Q XA
(1 + 𝜀A XA ) ⋅ dX A 20 000
0.5
(1 + 3XA ) ⋅ dX A
= ′0 =
k ∫0 (1 − XA ) 0.1 ∫0 (1 − XA )
This integral can be done numerically, following the procedure shown in pre-
vious chapters, and finally we get W = 255 kg of catalyst.
molecular weight is 60. The pellets have the following properties: radius 9 mm;
density 1.2 g/cm3 ; internal surface 100 m2 /g; and internal porosity 0.60. The
rate constant for this process is 3.86 m6 fluid /(m3 particle kmol s). Calculate the
efficiency factor and the particle size for which the influence of internal diffusion
is negligible.
Solution
The Thiele modulus can be estimated from
[ ]1
k ′′′ ⋅ CAs
n−1
(n + 1) 2
m⋅L=L⋅ (9.69)
2 ⋅ De
In the present case we have
CAs = PA ∕RT = 1∕(0.082 600) = 2.02 mol∕l = 0.0202 mol∕m3
In this case the order is 2, then:
[ ]1 ( )[ ]1
3 ⋅ k ′′′ ⋅ CAs 2 9 ⋅ 10−3 3 ⋅ 3.86 ⋅ 0.0202 2
m⋅L=L⋅ = = 266
2 ⋅ De 3 2 ⋅ 2.98 ⋅ 10−7
The efficiency factor is then (Eq. (9.50)):
tanh(m ⋅ L) tanh(2.66)
𝜂= = = 0.37
m⋅L 2.66
On the other hand, if we want to avoid internal diffusion effects, we must have
a Thiele modulus higher than 0.4, i.e.:
[ ]1 [ ]1
3 ⋅ k ′′′ ⋅ CAs 2 ( R ) 3 ⋅ 3.86 ⋅ 0.0202 2
0.4 < L ⋅ =
2 ⋅ De 3 2 ⋅ 2.98 ⋅ 10−7
The minimum value of radius accomplishing the inequality is 2.08 mm.
Solution
(a) In the fluidized bed, a good approximation is considering the solid as a com-
pletely mixed flow, but the gas flow is particularly due to the formation of
bubbles. Nevertheless, a good consideration can be to assume a plug flow in
the gas phase, so:
nA0 ⋅ dXA = (−rA′ ) ⋅ dW
( )
nA0 ⋅ (1 − XA )
nA0 ⋅ dXA = k ⋅ ′
⋅ dW
Q
( ′)
dXA k
= ⋅ dW
(1 − XA ) Q
XA ( )
dXA k′
= ⋅W
∫0 (1 − XA ) Q
The value of k ′ represents the kinetic constant, coinciding with the value cal-
culated under chemical control. On the other hand, at the end of the reactor
we have X A = 0.9 and the value of W can be calculated, being: W = 153.31 kg
The mass of catalyst needed in this system is not dependent on the concen-
tration of reagent, as a first-order reaction is considered.
(b) In the case of a system with a clear effect of external diffusion, we can consider
(−rA′′ ) = kc (CAg − CAS ) ⋅ ac ≈ kc ⋅ CAg ⋅ ac
The value of k C given in the data should be corrected with the surface/mass
ratio of the catalyst. Assuming spherical particles:
S S 4𝜋R2 3
ac = = = 4 = = 50 m2 ∕kg
W V ⋅𝜌 𝜋R3 ⋅𝜌 R
3
Solution
The critical sizes will be calculated using the limits:
– Negligible internal diffusion effect → m⋅L < 0.4
– Important internal diffusion effects → m⋅L > 4
As stated before, this is equivalent to
−r′′′ A L2
Negligible internal diffusion effect → We = < 0.15 (9.134)
De CAs
−r′′′ A L2
Important internal diffusion effects → We = >4 (9.135)
De CAs
(a) In the case where the diffusion effects are negligible, we have
−r′′′ A L2 (k ′′′ ⋅ CAS )L2 k ′′′ ⋅ L2
We = ≅ = < 0.15
De CAs De CAs De
From this relation:
√ √
0.15 ⋅ De 0.15 ⋅ 1.6 ⋅ 10−11
L< = = 1.18 ⋅ 10−5 m
k ′′′ 0.0193
For spherical particles, L = R/3, so R < 3.54⋅10−5 m = 0.0354 mm
(b) In diffusion effects are much important, we will have
k ⋅ CAs
(−r′′′ A ) = k ′′′ ⋅ CAs ⋅ 𝜂 ≅
m⋅L
In this way:
k ′′′ ⋅CAs
−r′′′ A L2 m⋅L
⋅ L2
We = = >4
De CAs De CAs
and then:
√ √
4 ⋅ De ⋅ m ⋅ L 4 ⋅ 1.8 ⋅ 10−11 ⋅ 4
L> ′′′
= = 1.22 ⋅ 10−4 m
k 0.0193
So, R > 3.66⋅10−4 m = 0.366 mm
282 9 Catalytic and Multiphase Reactor Design
Let us now do a molar balance in the reactor bed, considered as plug flow:
dnA = (−r′ A ) ⋅ dW
W being the weight of catalyst needed and (−r′ A ) the reaction rate based on the
solid catalyst weight. We will relate the amount of catalyst with the length of the
reactor, as one of the changes suggested is to increase reactor length. It is easy to
see that:
W = catalyst weight = 𝜌cat ⋅ Sbed ⋅ L ⋅ (1 − 𝜀)
Sbed being the cross-sectional area of the bed, and (1 − 𝜀) is the porosity of the
catalyst. Considering a first-order reaction, we can write
dnA dnA nA0 ⋅ dXA
= = = (−r′ A )
dM 𝜌cat ⋅ Sbed ⋅ (1 − 𝜀) ⋅ dL 𝜌cat ⋅ Sbed ⋅ (1 − 𝜀) ⋅ dL
= kC ⋅ (CAg − CAs ) = kC ⋅ (CAg )
9.9 Design of Multiphase Reactors 283
In the last equality, C AS = 0 as the reaction is quick and it is not possible to have
a reactant in the particle. Reordering and integrating through the whole bed:
Solution
The objective is to obtain the reaction rate throughout the reactor when the gas
pressure varies from 100 to 1 Pa. For this, the equation based on the sum of
284 9 Catalytic and Multiphase Reactor Design
the resistances corresponding to the gaseous layer, to the liquid layer and to the
chemical reaction is used:
pA
(−rA′′ ) = 1 HA HA
(9.94)
k ⋅a
+ k ⋅a⋅E
+ (k ⋅C )⋅f
′′
Ag Al B L
CBv=[];
rAv=[];
pAv=100:-1:1;
R1v=[];
R2v=[];
R3v=[];
for pA=pAv
CB=CB0-(pA0-pA);
MH=(k*DAB*CB)∧ 0.5/kAl;
Ei=1+CB/(pA/HA);
% Calculation of the ’E’
dif=10; % Value needed for starting calculation
E=1; % Value needed for starting calculation
while dif>0.0001
value= sqrt(MH∧ 2*(Ei-E)/(Ei-1));
Enew = value/tanh(value);
dif=(E-Enew)∧ 2;
E=Enew;
end
% Calculation of the reaction rate
R1=1/(kAg*a);
R2=HA/(kAl*a*E);
R3=HA/(k*CB*fL);
rA=pA/(R1+R2+R3);
CBv=[CBv CB];
rAv=[rAv rA];
R1v=[R1v R1/(R1+R2+R3)]; % Relative value of R1
compared to the total
R2v=[R2v R2/(R1+R2+R3)]; % R2
R3v=[R3v R3/(R1+R2+R3)]; % R3
end
figure(1)
plot(pAv,R1v,’--’,pAv,R2v,’*’,pAv,R3v)
legend(’R_g_a_s’,’R_l_i_q_u_i_d’,’R_r_e_a_c_t_i_o_n’)
xlabel(’P_A (Pa)’)
ylabel(’Relative resistance’)
286 9 Catalytic and Multiphase Reactor Design
figure(2)
plot(pAv,rAv)
xlabel(’P_A’)
ylabel(’Reaction rate (mol/(h ⋅ m∧ 3))’)
0.9
0.8
0.7
Reaction rate (mol/(h m3))
0.6
0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60 70 80 90 100
PA
1
Rgas
0.9
Rliquid
0.8 Rreaction
0.7
Relative resistance
0.6
0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60 70 80 90 100
PA (Pa)
Bibliography 287
Bibliography
10
Biochemical Reactors
10.1 Introduction
The term “fermentation”, in its strict sense, refers exclusively to the production of
alcohol from sugar. Etymologically, it means bubbling or boiling action and was
effectively used for the first time in the production of wine. At present, the mean-
ing of the word is broader: fermentation is a molecular transformation in which
an organic matter is converted into a product by the direct action of microorgan-
isms or by the action of enzymes.
The aim of the fermentation can be the production of microorganisms (for
example, the production of bread yeast), the disappearance of the substrate (water
purification), the production of a product, or a combination of the previous objec-
tives. In the cases in which the microbial mass produced is a residual product, it
is used as a feed for animals, due to its high protein content.
Biotechnology is important in the production of several substances, and in
the pharmaceutical, chemical, and food industries, as well as in the application
of microorganisms in the treatment of contaminated wastewater. In these pro-
cesses, the microorganisms transform certain unwanted products into harmless
matter.
The biological reactions, under the point of view of kinetics and thermodynam-
ics, present a series of own characteristics that differentiate them substantially
from other processes of chemical transformation. We can remark on some of
them:
1. They are generally slow processes, compared to chemical reactions. For this
reason, its time constants are usually measured in hours, and even days.
2. The fermentation processes have an autocatalytic character, since the
microorganisms (product) act as catalysts of the process.
3. The specific activity of biocatalysts is usually low compared to that of chem-
ical catalysts. They are extraordinarily dependent on environmental condi-
tions (T, pH, oxygen, etc.).
4. They may present problems of inhibition by product and/or substrate
depending on the operating conditions and work concentrations.
5. The nature of the biocatalyst can be significantly modified over a relatively
long process. Thus, both a loss of activity (due to problems of inhibition,
deactivation of extracellular enzymes, etc.) and an activation thereof
usually called the Michaelis complex. The formation of the product is according
to the following reaction scheme:
E+S ⇔E⋅S →E⋅P ⇔E+P
The existence of the E⋅S complex has been verified by different techniques such
as X-ray crystallography, spectroscopy and electronic spin resonance.
The substrate binds to a specific region of the enzyme, called the active site,
in each catalytic cycle, and the catalysis occurs only at that site. The formation
of the complex is due, in many cases, to the action of weak forces of attraction
(ionic effects, hydrogen bridges, and hydrophobic attractions between nonpolar
groups), although there are cases in which a covalent bond is involved.
A fundamental characteristic of enzymes as catalysts is their specificity, which
differentiates them from synthetic catalysts. Most catalysts used in the chemical
industry are nonspecific, that is, they catalyze similar reactions involving different
types of reagents. Although some enzymes are nonspecific, most catalyze a single
reaction for certain substrates. A simple model has been proposed to explain this
specificity depending on the existence of the active site, and the formation of the
enzyme–substrate complex, which is illustrated in the Figure 10.1.
This model states that the active site is the geometric complement of the sub-
strate and only the substrates that have the appropriate complementary form can
form the Michaelis complex. This hypothesis has been confirmed from the knowl-
edge of the tertiary structure of the enzymes.
The degree of specificity in the catalytic action of enzymes is a consequence
of their biological action in the cell. The specificity can be absolute: there are
enzymes that only catalyze a reaction from a single substrate. It is evident that
the enzymes that catalyze a certain reaction in the cellular metabolism have to
discriminate the substrates on which they act. Other enzymes (for example, pro-
teases), have as a function the hydrolysis of proteins breaking the peptide bonds
and their specificity is much lower, acting on a relatively wide range of proteins,
although with preferences in the recognition of certain amino acids on both sides
of the protein peptide bond. Many enzymes are stereospecific (recognize only
one molecular configuration) and regiospecific (they recognize a group among
others of the same molecule), which is one of its main advantages over chemical
catalysts.
Another characteristic of the enzymes that differentiate them from the usual
chemical catalysts is the frequent need for cofactors. A cofactor is a non-protein
Enzyme−substrate Products
Substrate complex
Enzyme
substance that is combined with a protein that does not act as a catalyst alone
(apoenzyme) to form a catalytically active complex. In biochemistry, when the
apoenzyme is linked to the cofactor it is called a holoenzyme, although in this
topic it is simply called an enzyme. There are two large groups of cofactors:
(a) Metal ions, such as Fe2+ , Fe3+ , Mn2+ , Zn2+ , etc.,
(b) Organic molecules called coenzymes (two of the most common coenzymes
are NAD or FAD).
Finally, when the cofactor is irreversibly bound to the enzyme, the name pros-
thetic group is used, as discussed earlier.
k−1
In fact, this scheme of reactions can be written more adequately in the following
way:
k1 k
−−−−−−−−→
E f + Sf ← −− ES −−−−→ E + 𝛽P
2 f
k−1
According to this mechanism, the free enzyme Ef and the free substrate Sf
are reversibly bound to form the enzyme–substrate complex ES, which leads,
irreversibly, to the obtaining of 𝛽 moles of the product P. In a batch reactor
of constant volume, the equations of balance of the different species are the
following:
dEf
= k−1 (ES) + k2 (ES) − k1 Ef Sf (10.1)
dt
dSf
= k−1 (ES) − k1 Ef Sf (10.2)
dt
d(ES)
= −k−1 (ES) − k2 (ES) + k1 Ef Sf (10.3)
dt
dP
= 𝛽k2 (ES) (10.4)
dt
10.2 Enzymatic Catalysis 293
where E, S, (ES), and P represent the molar concentrations of the different species.
From a mass balance, the initial concentrations of enzyme and substrate can be
expressed as follows:
E0 = Ef + (ES) (10.5)
P
S = Sf + (ES) + (10.6)
𝛽
Applying the hypothesis that the complex (ES) is in a pseudo stationary state
(d(ES)/dt = 0), from Eq. (10.1):
k1 Ef Sf = k−1 (ES) + k2 (ES)
E f Sf k + k2
= −1 = KM (10.7)
(ES) k1
K M is known as the Michaelis constant. Substituting Ef for its value of the ini-
tial concentration of E, we can obtain an expression for the concentration of the
enzyme complex-substrate:
(E0 − (ES)) ⋅ Sf E0
KM = i.e. (ES) = KM
(10.8)
(ES) 1+ Sf
a low affinity of the E⋅S complex and then it is easily dissociated to Ef and Sf . The
contrary is true for low values of K M .
In the Michaelis–Menten model, the maximum reaction rate per molecule of
enzyme is rm /E0 . On the other hand, k cat is defined as the turnover number, the
maximum number of substrate molecules that can be converted into product per
unit of time and per active site of the enzyme. If the enzyme has a single active
center, k cat matches with k 2 ; for an enzyme containing “n” active centers, it can
be calculated as rm /(n⋅E0 ).
The quotient k cat /K M can be interpreted from the Michaelis equation, written
for low substrate concentrations such as
k
r = cat ⋅ E0 ⋅ S (10.13)
KM
This quotient represents a pseudo second-order rate constant. Although it is
not an actual rate constant, the value of this quotient is used to determine the
affinity of the enzyme for different substrates, in the case of two substrates in
competition.
Solution
In this situation, the metabolism follows the equation:
dS S mg S
r= = rmax =1 ( )
dt KM + S dl h S + 10 mg
dl
where the substrate S is the alcohol. The initial amount is 25 mg/dl and we must
check if the level goes down to 0.2 mg/dl in three days (72 h).
We can use a simple finite differences method, in such a way that:
St − St+1 mg St
=1 ( )
Δt dl h St + 10 mg
dl
Δt ⋅ 1 ⋅ St
St+1 = St − t
(S + 10)
Choosing an adequate value of time increment (for example, 0.5 h), we will have
®
enough precision to give a response. This can be done both in Matlab or using
a spreadsheet. Using the last method, we get the following figure, where we can
check the patient is right and his level of alcohol in the blood corresponds with
that expected at t = 72 h.
10.2 Enzymatic Catalysis 295
Metabolism of alcohol
0.8 30
0.7
r (mg/(dl h)) S (mg/dl) 25
0.6
20
Rate (mg/(dl h))
Alcohol (mg/dl)
0.5
0.4 15
0.3
10
0.2
5
0.1
0.203 mg/dl
0 0
0 10 20 30 40 50 60 70 80
Time (h)
E+S ES E + P
+
I EI (inactive)
ESI (inactive)
ki2 I
+
E+S ES E+P
+ ki1
I EI (inactive)
The so-called pure types of inhibition have been considered, with K i as the
dissociation constant of the El or EIS complex, with “I” being the inhibitor
296 10 Biochemical Reactors
(a) Random mechanism: The substrates A and B form a complex with the
enzyme. Either of these can be combined with the other substrate to
form the EAB ternary complex, which leads to the X and Y products and
regenerates the enzyme:
k1 k2 ⎤
−−−−−−−−→
E+A← −−−−−−−−→
−− EA + B ← −− EAB ⎥ k
k−1 k−2
⎥ −−−−
3
→E + X + Y
k1′
−−−−−−−−→
k2′
−−−−−−−−→ ⎥
E+B← −− EB + A ← −− EAB ⎥
′
k−1 ′
k−2 ⎦
Working with a multisubstrate, reactions can be quite complicated when treat-
ing to represent the schemes of the reactions. Throughout history, several sys-
tems have been proposed, but the most successful has been the one proposed by
Cleland. The input of substrates and the output of products are represented by
vertical arrows indicating the entry direction or output of the reaction. Using the
Cleland’s symbols, the random mechanism can be written as
A B X Y
E EA EAB EXY EY E
EB EBA EYX
EX
B A Y X
The starting conditions for deriving the rate equation with two or more sub-
strates are the same as we used for the derivation of the rate equation in a simple
system: (i) we will assume the formation of a rapid equilibrium between the E
and the substrates (A and B); (ii) the equation of conservation tells us that (E0 )
equals the sum of all the chemical species in which the enzyme; and (iii) we will
assume the existence of steady-state conditions. Using Michaelis–Menten kinet-
ics for the formation of the complexes, we can then write
[E] ⋅ [A]
KA = (10.17)
[EA]
[E] ⋅ [B]
KB = (10.18)
[EB]
AB [EA] ⋅ [B]
KM = (10.19)
[EAB]
BA [EB] ⋅ [A]
KM = (10.20)
[EAB]
E0 = [E] + [EA] + [EB] + [EAB] (10.21)
AB
In that equations, KM is the concentration of substrate “B” that produces
BA
half of the maximum rate when the substrate “A” is in excess, and KM is the
298 10 Biochemical Reactors
concentration of substrate “A” that produces half of the maximum rate when
substrate “B” is in excess. Note that the following is true:
KA K BA
= MAB
KB KM
From the previous equations, we can deduce that:
[E][A] [E][B] [E][A][B]
E0 = [E] + + +
KA KB KA ⋅ KM
AB
r′max K′
rmax AB
KM
r′max/2
KA
BA
KM [A] [A]
Figure 10.2 Reaction rate and value of K ′ obtained vs. concentration of one of the substrates.
10.2 Enzymatic Catalysis 299
[A] = 0, the graph K ′ vs. [A] gives us the K A , and the value of K ′ at [A] very high
AB
is KM .
The same is true if the substrate with constant concentration is “B” (obviously
changing “A” by “B” in the expressions). Furthermore, if the concentration of sub-
strate “A” is saturating (i.e. the reaction rate does not increase with an ulterior
increase of [A]), we can see that:
[B]
r = rmax ⋅ BA (10.27)
KM + [B]
(b) Ordered mechanism: A substrate (A) initially binds with the enzyme forming
the EA complex. The second substrate only joins the EA complex to lead to
the EAB ternary complex:
k1 k2 k3
−−−−−−−−→
E+A← −−−−−−−−→
−− EA + B ← −− EAB −−−−→ E + X + Y
k−1 k−2
A B X Y
E EA EAB EXY EY E
AB
KM
[A]
Nevertheless, in that case, the equations are not true for the other substrate
concentration constant. If the substrate with constant concentration is “B”:
[A]
′
r = rmax ⋅ ′ (10.32)
K + [A]
where
[B]
′
rmax = k3 ⋅ E0 ⋅ AB (10.33)
KM + [B]
K
AB
K ′ = KM ⋅ BA A (10.34)
KM + [B]
Graphs similar to the other cases can be constructed to calculate the corre-
sponding values of the constants.
E EA EA′ EA′B E
In that case:
r = k4 ⋅ [EB] (10.35)
and k cat = k 4 .
10.2 Enzymatic Catalysis 301
A k−1 + k2 [E][A]
KM = = (10.36)
k1 [EA]
k + k ′
4 [EA ][B]
A
KM = −3 = (10.37)
k3 [EA′ B]
E0 = [E] + [EA] + [EA′ ] + [EA′ B] (10.38)
And solving for the rate equation, we find
[A][B]
r = rmax ⋅ ( ) ( ) (10.39)
k4 A B k4
k2
KM [B] + [A]KM + [A][B] 1 + k2
and
[A]
B
K ′ = KM ⋅( ) (10.42)
k4 A
k2
KM + [A]
K′ K′
Random Random
Ping-pong Ping-pong
Ordered Ordered
[A] [B]
[A] (g/l) [B] (g/l) r (g/(l s)) [A] (g/l) [B] (g/l) r (g/(l s))
Considering that all runs were performed at the same initial concentration of
enzyme, determine the mechanism of the transformation and the corresponding
kinetic constants.
Solution
First of all, let us show in a graph the measured values of “r” at the different runs.
As the runs are grouped in different concentrations of pyruvate (A), let us use the
following plot:
4.50
4.00
3.50
3.00
Reaction rate (g/s l)
2.50
0.50
0.00
0 0.5 1 1.5 2 2.5 3 3.5
[B] (g/l)
2
1/r (l s/g)
1 Y = 0.0491x + 0.2479
0.5
Y = 0.0313x + 0.2471
0
0 2 4 6 8 10 12
1/[B] (l/g)
From these lines, we can obtain the best values of K ′ and rmax , for each of the
concentrations of “A” used, and we obtain
1.00
0.80
K′ (g/l)
0.60
0.40
0.20
0.00
0 0.5 1 1.5 2 2.5
[A] (g/l)
AB
At the highest concentration [A], we see that KM = 0.13 g∕l and the slope of
′ AB
the plot K vs. 1/[A] at the initial stages ([A] close to zero) is KM ⋅ KA = 0.051 so
0.051
KA = 0.13 = 0.39 g/l.
10.3.2.1 Yields
The limiting substrate concept gives the possibility to define process yields.
From experimental data of the amount of substrate consumed and the biomass
obtained, it can be seen that there is a relationship between them that, in
many cases, is a direct proportionality. The biomass–substrate yield is defined
as the quotient between the increase in cell mass obtained and the substrate
consumption (usually, a carbon source):
ΔX
YS∕X = (10.43)
ΔS
The yield has units derived from those used for the measurement of the biomass
and the substrate (g dry weight of biomass/g of substrate, g/mol, etc.). Overall
yields, calculated from experimental data growth, are dependent on the carbon
source used and the operation conditions, and may vary throughout the process.
where r represents the rate of disappearance of the substrate and rx , rp , and rc are
the rate of formation of biomass, product, and CO2 , respectively.
Symbol Definition
Y X/S Molar growth rate: gram of dry biomass per mole of substrate consumed
Y X/O gram of dry biomass per gram of oxygen consumed or per mole of oxygen consumed
Y P/S gram or mole of product per gram or mole of substrate consumed
Y C/S mole of CO2 produced per mole of substrate consumed
306 10 Biochemical Reactors
Death phase
Log (X)
Exponential growth
phase
Lag phase
Time
where rX is the growth rate of the cells, 𝜇m is the maximum specific rate of growth,
and K S is the Monod constant. It is quite common to express the equation in the
function of the specific growth rate:
1 dX S
𝜇= ⋅ = 𝜇m (10.48)
X dt KS + S
in which 𝜇m is the maximum value that the growth rate can reach, when S ≫ K S
and the concentrations of the rest of nutrients have not changed significantly. K S
is the value of the concentration of the limiting nutrient, the specific growth rate
of which is half of the maximum.
The Monod equation is very simple, although it is not always possible to obtain
a good representation of the data of the growth of a microorganism. Therefore,
other models have been developed, but we focus on the Monod that is more
general.
The generalized Monod equation proposed by Han and Levenspiel tries to
cover most of the situations, in particular the inhibition by the substrate, the
product, or the cells. Its form is
( )
C n S
𝜇 = 𝜇m 1 − ∗i ( )m (10.49a)
Ci C
KS 1 − C ∗i +S
i
S
𝜇 = kobs (10.49b)
kSobs + S
C i is the concentration of the inhibitor, C i * is the critical concentration of the
inhibitor to completely stop the process (in this case, cell growth). n and m are
constants, usually related to the toxic power of the inhibitor. Substituting in
equation or generalized C i by S, P, or X, it could model the inhibition by the
substrate, product, or the cells themselves. As can be seen, when C i ≪ C i * , the
equations are reduced to that of Monod.
Concentration Concentration
Time (h) biomass (g/l) glucose (g/l)
0 0.64 30
5 1.95 27.4
10 4.21 23.6
15 5.54 21
20 6.98 18.4
25 9.5 14.8
30 10.3 13.3
35 12 9.7
40 12.7 8
45 13.1 6.8
50 13.5 5.7
55 13.7 5.1
Solution
From the data presented, we can calculate the rate of biomass production and
substrate (glucose) consumption using the finite differences approximation, i.e.:
dX X t+1 − X t dS St+1 − St
= and =
dt Δt dt Δt
Although this is usually valid for small values of Δt, in the data we have only
time increments of five or more hours, so we have no choice. Using the approxi-
mation, we can calculate the first two columns of the following table. The last two
columns are calculated as indicated in the table:
10.3 Microbial Kinetics 309
rX /rS (g cells
rX = (1/V)⋅dX/dt rS = −(1/V)⋅dS/dt 𝝁 = rX /X produced/g
Time (h) (g/(h l)) (g/(h l)) (h−/ ) substrate)
350
300
y = 2.017E + 03x – 8.539E + 01
250 R2 = 8.974E – 01
200
1/s (l/g)
150
100
50
0
0 0.05 0.1 0.15 0.2
–50
1/μ (h)
310 10 Biochemical Reactors
We can see that the data seems to fit the Monod kinetics, as a straight
line is more or less obtained, but from the slope and ordinate (fitted using
least-squares approximation), a negative value of the maximum growth rate is
found. This is obviously not possible, and a better approximation is needed. From
the data in the former table, the maximum value of 𝜇 is approximately 0.41 h−1 ,
which we will use as the value of 𝜇m . Making the line obtained to pass through
the point (0, 0.41), the slope changes slightly and becomes 1225: therefore:
KS
= 1225 (g h)∕l
𝜇m
So we can take a value of K S = 1225 × 0.41 = 501.5 g/l. With these values of the
constants, let us finally see the fit by representing the experimental and calculated
values of the growth rate vs. concentration of glucose.
0.50
Experimental
0.40 Monod kinetics
0.30
rX/X (h−1)
0.20
0.10
0.00
0 5 10 15 20 25 30 35
S (g/l)
We see that the approximation gives values of the growth rate in the order
of those experimental, but the model is not good enough to represent this
bioreaction.
– It allows the continuous use of the biocatalyst, since it is easily retained inside
the reactor, while maintaining a continuous flow of liquid in and out. This
means that in the case of cells, the continuous reactors can operate with flow
rates higher than those corresponding to the wash limit observed when oper-
ating with free cells. In the case of batch reactions, the immobilization favors
the reuse of the biocatalyst.
– Another important aspect is that with the immobilization the biocatalyst con-
centration can be significantly increased, with respect to a process in which it
is in suspension. With this it is deduced that in systems operating continuously,
the immobilized biocatalysts allow to increase sensibly the productivity of the
corresponding bioreactors, since they allow to operate at a higher concen-
tration of catalyst (and, therefore, greater conversion of substrate), and with
higher flow rates.
External diffusion
Internal diffusion
Sf
Ss
Enzymes entrapped
in droplets
be used in very specific cases, since agitation usually severely affects the physical
integrity of most particles of immobilized biocatalysts.
where the subscripts refer to the fluid (f ) and the surface of the particle(s).
In Eq. (10.52), the rate with no limitation by external mass transfer equals
Sf
rm K +S , so the efficiency factor is
M f
Sf
1 (KM +Ss )
kC
+ rm
𝜂e = Sf
(10.55)
rm K
M +Sf
heterogeneous catalysis, the substrate flow is related to the reaction rate with
the use of an efficiency factor. In general, the substrate flow will be given by
dX S
rX = = 𝜂 ⋅ rm (10.57)
dt KS + S
where rm is the maximum reaction rate per volume of immobilized catalyst.
The simplest case to consider is that of a layer of microbes or enzymes that are
immobilized uniformly in a block of support material with infinite area but finite
depth. Since it is assumed that the reaction rate is given by a Michaelis–Menten
kinetics, a material balance for any point of the block gives
d2 S S
De = rm (10.58)
dx2 KM + S
where De is the effective diffusion for the substrate in the block and “x” is the
distance measured from the surface of the block. This equation can be made
dimensionless by adding 𝜅 = S/K M and Z = x/L, where L is the thickness of the
block. With this previous equation, the system is defined by
d2 𝜅 𝜅
= Φ2 (10.59)
dZ2 1+𝜅
where Φ is the Thiele module defined by
√
rm
Φ=L (10.60)
De ⋅ KM
for a system of immobilized enzymes. For a microbial layer, the equation defining
the system is similar to (10.58), but substituting rm by 𝜇m and K M by K S . Finally,
we can get
√
𝜇m ⋅ X
Φ=L (10.61)
YX∕S ⋅ De ⋅ KS
for a microbial layer.
Equation (10.59) can be solved by numerical integration using the usual bound-
ary conditions:
𝜅 = 𝜅s when Z = 0
d𝜅
= 0 when Z = 1
dZ
to obtain the curves’ rate position or concentrationposition. The aspect of these
profiles is similar to those discussed in Chapter 9; for internal diffusion resistance,
see Figure 9.5. The family of curves obtained shows that the global reaction rate
decreases when the value of the Thiele module increases. The rate is controlled
by the reaction kinetics at low values of the Thiele module.
10.5 Bioreactors
Bioreactors or fermenters are classified according to the terminology used in the
design of chemical reactors in:
10.5 Bioreactors 315
Gas exit
Volume = V
Product: Q, X, S, P
Oxygen
10.5 Bioreactors 317
When working in the exponential growth (k d = 0), and with sterile feed (X 0 = 0),
we can write
S
Q0 X = V 𝜇 X = V 𝜇 m X (10.64)
KS + S
Usually, the term Q0 /V is called “dilution rate” (D) and is equivalent to the
inverse of the residence time. Using this definition:
S r
D = 𝜇m =𝜇= X (10.65)
KS + S X
Therefore, D = 𝜇 in the steady state, so the specific rate of growth can be con-
trolled by manipulating the dilution rate.
Introducing the Monod equation in the balance:
S
D = 𝜇m
KS + S
KS
S =D⋅ (10.66)
𝜇m − D
For the substrate, we can say
X0 − X −X
YX∕S = = (10.67)
S0 − S S0 − S
X
S = S0 +
YX∕S
and, equally, for the product:
P0 − P
YP∕S = (10.68)
S0 − S
D·X
D
Doptima D
washout
Figure 10.8 reproduces the situation described where the sensitivity of the con-
centrations varies with the dilution rate. On the other hand, in the case of a
non-sterile feed, X 0 ≠ 0, no washing of the reactor takes place since when D → ∞,
X = X0.
Solution
In this case, we are looking for the maximum in Figure 10.8. So, in that point:
d(D ⋅ X)
=0
dD
and we can write
KS
S =D⋅ (10.66)
𝜇m − D
X = (S0 − S) ⋅ YX∕S
and then:
( ( ) )
KS
d(D ⋅ (S0 − S) ⋅ YX∕S ) d D ⋅ S0 − D ⋅ 𝜇m −D
⋅ YX∕S
= =0
dD dD
Deriving this expression:
( )
𝜇m
2
⋅ KS
S0 + KS − Dopt ⋅ ⋅ YX∕S = 0
(𝜇m − Dopt )2
Rearranging and clearing for the optimum dilution rate:
( √ ) ( √ )
KS 4
Dopt = 𝜇m ⋅ 1 − = 1.3 ⋅ 1 − = 0.325 s−1
KS + S0 4 + 60
10.5 Bioreactors 319
Recycle:
Qr, X0
Centrifuge
Volume = V
the ideal tubular flow-through fermenter (Figure 10.10), the substrate concentra-
tion decreases from entry to exit, and that of cells and products increase.
The most important factor to consider is the need for a non-sterile feed since,
otherwise, microorganisms cannot exist inside. In a steady state, conditions vary
with position but not with time. So, in an infinitesimal volume of bioreactor, the
conditions and the rate are practically constant, and the balance for biomass in
this volume is expressed as follows:
0 = Q ⋅ X − Q(X + dX) + rX ⋅ dV
Q ⋅ dX = rX ⋅ dV (10.73)
Integrating between V = 0, X = X 0 , V = V , S = S and the design equation is
obtained
X S
1 V dX dS
= = = (10.74)
D Q ∫X0 rX ∫S0 YS∕X ⋅ rX
which will allow us to analyze and design the ideal tubular fermenter. It is required
to integrate, of course, make explicit the reaction rate as a function of the corre-
sponding concentration.
dV
P, S, X
Q, P0, S0, X0 Q
Qr R = Qr/Q
some cases, all the nutrients are fed into the bioreactor. An alternative description
of the method is that of a culture in which a base medium supports initial cell
culture and a feed medium is added to prevent nutrient depletion.
The advantage of the fed-batch culture is that one can control the concentration
of the fed substrate in the culture liquid at arbitrarily desired levels (in many cases,
at low levels).
Generally speaking, fed-batch culture is superior to conventional batch culture
when controlling concentrations of a nutrient (or nutrients) affect the yield or
productivity of the desired metabolite.
Assuming an input of Q0 (flow rate), X 0 (concentration of cells) equal to zero,
S0 (concentration of substrate), and a V 0 (initial volume of the reactor), the mass
balances are
d(XV ) dV dX
=X +V = V ⋅ rX (Cells), (Eq. 10.77)
dt dt dt
d(PV ) dV dP
=P +V = V ⋅ rP (Product), (Eq. 10.78)
dt dt dt
d(SV ) dV dS
=S +V = Q0 S0 − V ⋅ rS (Substrate), (Eq. 10.79)
dt dt dt
One aspect of the fed-batch model is that volume is not constant; therefore,
the cell, product and substrate are subject to a dilution effect. Mathematically, the
chain rule of differential calculus is applied to bear in mind the variation of the
volume, Q0 = dV/dt and then, rearranging:
dX Q ⋅X
=− 0 + rX (10.77)
dt V
dP Q ⋅P
=− 0 + rP (10.78)
dt V
dS Q0
= (S − S) − rS (10.79)
dt V 0
Example 10.5 Fed-batch Reactor Simulation
A fed-batch reactor is being set up in order to produce a product “P” in a cell cul-
ture. The following parameters were chosen: input cell concentration = 0.05 g/l,
input substrate concentration = 10 g/l, initial volume = 1 l, and flow rate input
1 l/h. The feed has no any concentration of the desired product. Other data is
available: 𝜇m = 0.20 h−1 ; K S = 1.00 g/l; Y X/S = 0.5 g/g; and Y P/X = 0.2. Simulate the
behavior of this reactor calculating the evolution of all concentrations and the
volume with time.
Solution
The present design is similar to that just mentioned, but the feed of the reactor is
not sterile and has a known concentration of cells X 0 = 0.05 g/l. So, in this case,
Eq. (10.77) is not valid, and the balance of cells is
d(XV )
= QX0 + V ⋅ rX
dt
Using the chain rule for derivation and bearing in mind that dV /dt = Q, we can
see that:
dX
V = Q(X0 − X) + V ⋅ rX
dt
322 10 Biochemical Reactors
10
Cells
Substrate
Product
8
Concentration (g/l)
0
0 10 20 30 40 50
Time (h)
V(1)=V0;
X(1) = 0; % g/l Initial concent. of cells in the tank
X0=0.05; % g/l concentration of cells in the input
current
S(1) = 0; % g/l
P(1) =0; % g/l
t=0; % h
tfinal=50; % h
inct=0.1; % h
i=1;
tv=0:inct:tfinal; % Values of time used in the
calculation
while t<tfinal
i=i+1;
t=t+inct;
rx=mumax*S(i-1)*X(i-1)/(Ks+S(i-1));
rp=rx*Ypx;
rs=rx/Yxs;
X(i)=X(i-1)+inct*(Q*(X0-X(i-1))/V(i-1)+rx);
P(i)=P(i-1)+inct*(-Q*P(i-1)/V(i-1)+rp);
S(i)=S(i-1)+inct*(Q*(S0-S(i-1))/V(i-1)-rs);
V(i)=V(i-1)+Q*inct;
end
plot(tv,X,tv,S,'*',tv,P)
legend('Cells','Substrate','Product')
xlabel('Time (h)')
ylabel('Concentration (g/l)')
As we can see in the figure, the consumption of substrate is too high when time
approaches to 40 hours, so the best operation time is less than 40 hours, when
the product concentration does not increase anymore and the concentration of
cells behave in a similar way. In that case, the volume of the reactor needed will
be approximately 40 l.
3 0.14
0.12
2.5
0.1
2
S1 (g/l)
X (g/l)
0.08
1.5
S1 0.06
1 X
0.04
0.5 0.02
0 0
0 0.2 0.4 0.6 0.8 1
Recirculating ratio (α)
Bibliography 325
Note that it is possible to have a value of 𝛼 higher than unity, as the Qr can be
higher than Q0 , but this usually is not useful and, furthermore, is expensive due to
the high costs of impulsion of the currents. In the extreme case that 𝛼 is infinity,
the whole system will behave as a CSTR and it will be not possible to observe
differences between X 1 and X 0 .
If, on the other hand, the fixed parameter is 𝛼 = 0.5 (as an example), by varying
the value of 𝛽 we will get a different concentration X 1 , if X 0 is fixed. Then, using
the same equations, we can simulate the behavior of that system, and, in this way,
we will find the next figure.
3.5 0.16
3 0.14
0.12
2.5
0.1
2
S1 (g/l)
X (g/l)
0.08
1.5
0.06
1
S1 X 0.04
0.5 0.02
0 0
0 0.2 0.4 0.6 0.8 1
Ratio β = X0/X1
Bibliography
Index
a c
acid catalysts calcination oven 185
ethylbenzene production 193–194 catalysts
FCC 191–192 characteristics of 207
Lewis acid 191 definition 207
acrylic acid production 196 external diffusion 219
activated carbons (AC) 180, 181, fluid-fluid reactions 245–261
183 internal diffusion 221–236
active metals 185–186 monolithic catalytic reactors
alumina 179–182, 184, 185, 190, 191, 237–245
195, 200, 203, 204, 215, 241, rate equation
278 heterogeneous systems 208
ammonia absorption 269–272 mechanism of 215–216
aqueous glass 184 porosity and void volumes 214
second-order decomposition
b reaction 209
bacterial growth 308–310 theory of adsorption 217–219
batch tank stirred reactor resistance combination 236–237
example of 130–136 catalyst support 178, 180, 198
temperature control heat cell growth 304–310
transmission 130 cell recirculation 319–320
biocatalysts immobilization 310–313 chemical engineering boundary
biological reactions 289 conditions 94
biomass-substrate yield 305 Chemical Reaction Engineering 87,
bioreactors 262
batch reactors 315 chlorine removal 282–283
CSTB 315–319 cobalt 186
CSTR 315 CO2 elimination 245, 267–268
fed-batch 315 continuous stirred tank bioreactor
fed-batch culture 321–323 (CSTB) 316–317
tubular bioreactor 316 continuous stirred tank reactor (CSTR)
tubular fermenters with flocs 63, 111, 112, 315
319–320 bioreactor 323–325
biotechnology 289, 315 complex system 28–30
o r
oscillating pressure 157, 159, 169–170 ramp function 59–60
oxidation catalysis reactor of known volume 275–276
acrylic acid production 196 recipient dispersion module 19
ethylene oxide 194–195 reduction catalysis
reduction catalysis 197 methanation 199–200
Index 331