100% found this document useful (1 vote)
156 views

Chemical Reactor Design Mathematical Modeling and Applications

Alarm Management for Process Control

Uploaded by

vrajakisoriDasi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
156 views

Chemical Reactor Design Mathematical Modeling and Applications

Alarm Management for Process Control

Uploaded by

vrajakisoriDasi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 338

Chemical Reactor Design

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.

Cover Library of Congress Card No.:


© zorazhuang/Getty Images applied for

British Library Cataloguing-in-Publication


Data
A catalogue record for this book is
available from the British Library.

Bibliographic information published by


the Deutsche Nationalbibliothek
The Deutsche Nationalbibliothek lists
this publication in the Deutsche
Nationalbibliografie; detailed
bibliographic data are available on the
Internet at <http://dnb.d-nb.de>.

© 2020 Wiley-VCH Verlag GmbH &


Co. KGaA, Boschstr. 12, 69469
Weinheim, Germany

All rights reserved (including those of


translation into other languages). No
part of this book may be reproduced in
any form – by photoprinting,
microfilm, or any other means – nor
transmitted or translated into a
machine language without written
permission from the publishers.
Registered names, trademarks, etc. used
in this book, even when not specifically
marked as such, are not to be
considered unprotected by law.

Print ISBN: 978-3-527-34630-1


ePDF ISBN: 978-3-527-82339-0
ePub ISBN: 978-3-527-82338-3
oBook ISBN: 978-3-527-82337-6

Typesetting SPi Global, Chennai, India


Printing and Binding

Printed on acid-free paper

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

Part I Reactor Analysis, Design, and Scale-up 1

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 Convolution and Deconvolution of Residence Time


Distribution Curves in Reactors 35
2.1 Introduction 35
viii Contents

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

3 Use of Transfer Function for Convolution and Deconvolution of


Complex Reactor Systems 57
3.1 Introduction 57
3.2 Definition and Properties of the Transfer Function 57
3.3 Laplace Transform 58
3.3.1 Laplace Transform of Some Important Functions for Reactor
Characterization 58
3.3.1.1 Ramp Function 59
3.3.1.2 Sinusoidal Function 60
3.3.1.3 Pulse Function 60
3.3.1.4 Other Functions 61
3.4 Use of Laplace Transform in Chemical Reactor Characterization 62
3.4.1 Study of the RTD in the CSTR 62
3.4.2 Study of the RTD in the PFR 65
3.5 Complex Network of Ideal Reactors 66
3.5.1 Systems in Series 67
3.5.2 Systems in Parallel 69
3.5.3 Systems with Recycle 71
3.6 Transfer Function for the Dispersion Model 81
Bibliography 85

4 Partial Differential Equations in Reactor Design 87


4.1 Introduction 87
4.2 Classification of Partial Differential Equations 87
4.3 Approximations by Finite Differences 88
4.3.1 First-order Approximation 88
4.3.2 Approximation of Second Order 89
4.4 Approaching the Problem Using Finite Differences 91
4.4.1 Explicit Method 93
4.4.2 Initial and Boundary Conditions 94
4.4.3 Stability 96
4.4.3.1 Resolution of the Selected Problem and Programming 97
4.5 Other Applications of PDE – Numerical Methods 98
4.5.1 The RTD of a Complex System 98
4.5.1.1 Boundary Conditions for Partial Differential Equations 101
4.5.2 Concentration Profile in a Reactor in Which There Is Flow and
Dispersion 104
Contents ix

4.5.3 Reaction on a Catalytic Flat Wall 106


Bibliography 110

5 Unsteady State Regime Simulation in Reactor Design 111


5.1 Introduction 111
5.2 CSTR Working in Unsteady State 111
5.3 PFR Working in Dynamic Regime (No Dispersion) 113
5.4 PFR Working in Dynamic Regime (with Dispersion) 115
5.5 Multiple Steady States in CSTR with Exothermal Reaction 118
Bibliography 125

6 Scaling and Stability of Chemical Reactors 127


6.1 Introduction 127
6.2 Scaling the Batch Tank Stirred Reactor 129
6.2.1 Temperature Control. Heat Transmission 129
6.2.2 Example of Scaling a Batch and Semi-batch Reactor 130
6.3 Rapid Exothermic Reaction in a Tubular Reactor 136
6.3.1 Study of the Stability of the Process 139
Bibliography 155

7 Forced Unsteady State Operation of Chemical Reactors 157


7.1 Introduction 157
7.2 Objectives and Types of FUSO 158
7.3 Periodic Variation of the Input 159
7.3.1 Modes of Operation 160
7.3.2 Design Strategy 162
7.3.2.1 Choice of the Entrance to Manipulate 162
7.3.2.2 Choice of Handling 162
7.3.2.3 Choice of Mode 162
7.3.3 Periodic Variation of Concentration 163
7.3.4 Periodic Variation of the Flow 164
7.3.5 Periodic Variation in Temperature 165
7.4 Periodic Flow Reversal 165
7.4.1 Operation Design 167
7.5 Operation with Variable Volume (VVO) 168
7.6 Oscillating Pressure 169
Bibliography 170

Part II Catalytic, Multiphase and Biochemical Reactor


Design 173

8 Industrial Catalysis 175


8.1 Introduction 175
8.1.1 Reactors for Solid-Catalyzed Reactions 175
8.1.2 Solid Catalysts (Supports) 178
x Contents

8.1.2.1 Choice of Catalyst Support 180


8.1.2.2 Comparison and Uses of Supports 180
8.1.2.3 Silica, Alumina, and Mixtures 181
8.1.2.4 Zeolites 181
8.1.2.5 Activated Carbons (ACs) 183
8.2 Industrial Preparation of Catalysts 183
8.2.1 Structure of Commercial Catalysts 184
8.2.1.1 Synthesis of Zeolites 184
8.2.1.2 Manufacturing of the Catalytic Support 185
8.2.1.3 Impregnation with Active Metals 185
8.2.2 Key Definitions in Catalysts Performance 186
8.3 Main Catalytic Processes in Industry 188
8.3.1 Acid Catalysis 190
8.3.1.1 Fluid Catalytic Cracking 191
8.3.1.2 Ethylbenzene Production 193
8.3.2 Oxidation Catalysis 194
8.3.2.1 Ethylene Oxide Production from Ethylene 194
8.3.2.2 Acrylic Acid from Acrolein 196
8.3.3 Reduction Catalysis 196
8.3.3.1 Steam Reforming of Alcohols 197
8.3.3.2 Steam Reforming of Hydrocarbons (Methane) 199
8.3.3.3 Methanation: CO/H2 to Methane 199
8.3.4 Environmental Catalysis 200
8.3.4.1 Catalytic Reactions for the Removal of Pollutants in the Exhaust
Gases 201
8.3.4.2 Components of the Cleaning Systems 202
8.3.4.3 Three-Way Catalyst 202
8.3.4.4 SCR Catalyst 203
8.3.4.5 Diesel Oxidation Catalyst (DOC) 204
8.3.4.6 Diesel Particulate Filter (DPF) Catalyst 204
Bibliography 205

9 Catalytic and Multiphase Reactor Design 207


9.1 Introduction 207
9.2 Rate Equation in Catalytic Systems 208
9.2.1 Steps in the Catalytic Reaction 214
9.3 Rate Equation When Chemical Step Is Limiting Reaction Rate 215
9.3.1 Mechanisms of Catalysis 215
9.3.2 Theories About Adsorption 217
9.4 Rate Expression for External Diffusion as a Limiting Step 219
9.5 Reaction Rate When Internal Diffusion Is Slow 221
9.5.1 First-order Kinetics in Flat Particles 222
9.5.2 First-order Kinetics in Other Geometries 225
9.5.3 Limits of Thiele Modulus and Weisz Modulus 229
9.6 Combination of Resistances 236
9.7 Monolithic Catalytic Reactors 237
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 245
Contents xi

9.8.1 Transfer Models 245


9.8.2 Two-Film Theory 247
9.8.2.1 Case A: Instantaneous Reaction 247
9.8.2.2 Analysis of the Controlling Steps: The Hatta Modulus 250
9.8.2.3 Other Cases in Fluid–Fluid Reactions: The General
Rate Equation 251
9.8.3 Gas–Liquid Reactions in Solid Catalysts. General Equation 256
9.8.3.1 Estimation of the Controlling Resistance in Multiphase Systems 258
9.8.3.2 General n-th Order Kinetics 260
9.9 Design of Multiphase Reactors 261
9.9.1 Types of Flow in Multiphase Reactors 261
9.9.2 Design Models for Flow in Multiphase Reactors 262
9.9.3 Case 1: Pure Absorption (No Reaction) 263
9.9.3.1 Situation 1: Gas and Liquid Phases in Plug Flow 263
9.9.3.2 Situation 2: Gas and Liquid Phases Completely Mixed 264
9.9.3.3 Situation 3: Gas Phase in Plug Flow. Liquid Phase Completely
Mixed 264
9.9.4 Case 2: Systems with Reaction 265
9.9.4.1 Situation 1: Gas and Liquid Phases in Plug Flow 265
9.9.4.2 Situation 2. Gas and Liquid as Mixed Flow 268
9.9.4.3 Situation 3. Gas in Plug Flow. Liquid Mixed Flow 269
9.9.5 Case 3. Multiphase Reactors 269
Bibliography 287

10 Biochemical Reactors 289


10.1 Introduction 289
10.2 Enzymatic Catalysis 290
10.2.1 Characteristics of Enzymatic Catalysis. The Active Center 290
10.2.2 Kinetics of Enzymatic Reactions 292
10.2.2.1 Kinetics of Reactions with a Single Substrate. Michaelis–Menten
Equation 292
10.2.2.2 Meaning of the Parameters of the Michaelis Equation 293
10.2.3 Enzymatic Reactions with Inhibition 295
10.2.4 Enzymatic Reactions with More than One Substrate 296
10.2.4.1 Case 1. Enzymatic Reactions with Two Substrates by Formation of a
Ternary Complex 296
10.2.4.2 Case 2. Enzymatic Reactions with Two Substrates Without Formation
of a Complex 300
10.2.4.3 Strategies to Distinguish the Previous Cases 301
10.3 Microbial Kinetics 304
10.3.1 Stoichiometry of the Microbian Growth 304
10.3.2 Stoichiometry of Product Formation 304
10.3.2.1 Yields 305
10.3.2.2 Theoretical Yield Obtained from Stoichiometric Coefficients 305
10.3.3 Cell Growth, Substrate Consumption, and Product Formation 305
10.3.3.1 Kinetics of Growth 306
10.3.3.2 Kinetics of Maintenance 307
xii Contents

10.4 Immobilization of Enzymes and Cells: Mass Transfer Effects 310


10.4.1 Effect of Limitation by Internal Diffusion 313
10.5 Bioreactors 314
10.5.1 Continuous Stirred Tank Bioreactor (CSTB) 316
10.5.1.1 Influence of the Dilution Rate. Calculation of the Bioreactor
Wash 317
10.5.1.2 Cell Recirculation 319
10.5.2 Tubular Fermenters with Flocs 319
10.5.2.1 Tubular Fermenter with Recirculation and Monod Kinetics 320
10.5.3 Fed-batch Bioreactor 320
Bibliography 325

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.

Alicante, Spain Juan A. Conesa


May 2019
xv

Nomenclature

Suggested units are indicated for each variable.

N Amount of substance (mol)


Q Volumetric flow rate (m3 /s)
C Concentration (usually mol/m3 , or kg/m3 )
E (in the RTD context) → residence time distribution function, RTD (–)
F Integral form of the residence time distribution function (–)
tm First moment of the RTD; mean value of time (s)
t Average residence time (=V /Q) (s)
V Volume (m3 )
𝜎 Standard deviation (s)
s (in the RTD context) → skewness (s)
s (in the LT context) → main variable in the Laplace space
𝛿 Dirac delta function (–)
nt Number of tanks in the tanks-in-series model (–)
k Kinetic constant (units depend on the kinetic law)
De Effective diffusion coefficient (m2 /s)
nT Total molar flow (mol/s)
S Section (m2 )
CT Total concentration (mol/m3 )
u Linear velocity (m/s)
L Characteristic length (m)
M Tracer mass (kg or kmol)
Vp Volume with piston flow regime (m3 )
Vd Dead volume (m3 )
XA Molar conversion of reactant A (–)
rA Reaction (or process) rate based on the external catalyst surface
(mol/(s m2 ))
r′ A Reaction (or process) rate based on the weight of catalyst (mol/(s kg))
r′′ A Reaction (or process) rate based on the volume of reacting species
(mol/(s m3 ))
r′′′ A Reaction (or process) rate based on the volume of catalyst particles
(mol/(s m3 ))
C As Concentration of A in the surface (mol/m3 )
DAB Diffusion coefficient (diffusivity) of A in B (m2 /s)
xvi Nomenclature

qs Heat flow in the surface (J/s)


h Heat transfer coefficient (J/K s)
T Temperature (K)
kL Mass transfer coefficient in the liquid phase
(mol/(s m2 )/(mol/m3 ) = m/s)
Qc Flow rate in the cooling zone (m3 /s)
ΔH r Enthalpy of a reaction (kJ/mol)
U Global heat transfer coefficient (J/K s)
E (in the reaction rates context) → activation energy (J/mol)
H(s) Transfer function (–) (Laplace space)
X(s) Stimulus function (–) (Laplace space)
Y (s) Response to stimulus function (–)(Laplace space)
h(t) Time-dependent transfer function (–)
y(t) Time-dependent response to stimulus function (–)
x(t) Time-dependent stimulus function (–)
L{h(t)} Laplace transform of function h(t)
𝛼 Fraction of flow rate (–)
𝛽 Fraction of volume (–)
Cp Concentration of product “P” (mol/m3 )
cp Calorific capacity of the reacting flow (J/K g)
cPc Calorific capacity of the cooling fluid (J/K g)
Nc Dimensionless cooling capacity (–)
N ad Dimensionless adiabatic heat increment (–)
Δ𝜐 Dimensionless temperature difference (–)
kh Kinetic constant at the temperature of the cooling medium (units
depend on the reaction order)
k0 Pre-exponential factor of the kinetic constant (units depend on the
reaction order)
Φ Thiele modulus (–)
We Weisz modulus (–)
m kinetics to diffusion ratio in catalytic reactions (1/m)
KM Michaelis constant (g/l)
KS Monod constant (g/l)
W Weight of catalyst (g)
𝜇 Growth rate per unit of cell (1/s)
𝜙 Tortuosity factor (–)
𝜂 Effectiveness (–)
𝜂e Effectiveness (related to external diffusion) (–)
J0 Bessel function of zero order (–)
J1 Bessel function of first order (–)
𝛼 Fraction of volume of a subsystem (–)
𝛽 Fraction of flow passing through a subsystem (–)
q Heat flux (J/s)
qC Heat flux in cooling media (J/s)
u Linear velocity (m/s)
TC Temperature of the cooling media (K)
Am×n Matrix of convolution
Nomenclature xvii

gA Generation term of a mole balance (mol/s)


nA Molar flow of component “A” (mol/s or mol/(s m2 ))

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

Reactor Analysis, Design, and Scale-up


3

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.

1.2 Residence Time Distribution (RTD) Function


The idea of introducing the RTD in the analysis of the behavior of reactors
occurred thanks to MacMullin and Weber (in 1935), although it was Danckwerts
(later, in 1953) who structured this analysis and defined most of the distributions
of interest.
In an ideal PFR, all the particles (or units) of material that leave the reactor have
remained in it the same time. Analogously, in an ideal (well-mixed) batch reactor,
all particles are in the reactor the same period of time. The time that these units
have remained in the reactor is what we call the residence time of those particles
in that reactor.
The ideal reactor’s PFR and batch are the only ones in which all the portions
of reactants present in the reactor have the same residence time. In all other
reactors, the particles entering the reactor vessel remain inside the reactor for
different periods of time; that is, there exists a RTD inside the reactor.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
4 1 Nonideal Flow

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.

1.2.1 Measurement of the RTD


The experimental measurement of the RTD is done by injecting a tracer into the
reactor at a definite time (t = 0). This tracer is a chemical, molecule, or an inert
atom. The tracer concentration is then measured at the outlet at different times.
The tracer must be inert, easily detectable, with physical properties similar to
those of the substances present in the reaction mixture and easily soluble in it.
In addition, it should not be adsorbed on walls or other surfaces on the reac-
tor. The objective is to reflect, as best as possible, the behavior of the substances
that are flowing through the reactor. The most commonly used tracers are dyes
and radioactive material, while the most commonly used injection methods are
pulse input (“Dirac delta function”) and step input (suddenly increase of the tracer
concentration).

1.2.1.1 Pulse Input


This kind of experiment consists of introducing into the current entering the reac-
tor, quickly and at once, an amount N 0 of tracer. The output concentration is
subsequently measured as a function of time. Concentration-time curve charac-
teristics for the input and exit of an arbitrary reactor can be observed in Figure 1.1.
The concentration-time curve corresponding to the effluent is called curve C in
RTD analysis.
Let us consider a system with a single input and a single output, in which a
tracer is injected in pulse and in which it is transported, exclusively because of
the flow (not because of dispersion), through the system. If a time increment Δt
small enough is chosen so that the concentration of tracer, C(t), which leaves
the system between the instants “t” and “t + Δt” is constant, we can express the
amount of tracer (in moles or grams), ΔM, that leaves the reactor between “t”
and “t + Δt” as
ΔM = C(t) ⋅ Q ⋅ Δt (1.1)
1.2 Residence Time Distribution (RTD) Function 5

Tracer
input

Feed Output
Reactor

Tracer
Tracer detection
concentration Tracer
concentration
Pulse input
Pulse
response
0 Time
0 Time
Tracer Tracer
concentration concentration

Step input Step


response

0 Time 0 Time

Figure 1.1 Measurement of the RTD.

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.

1.2.1.2 Step Input


During a step input experiment, a tracer is added at a steady rate to the reactor
feed to give a steady input concentration of C 0 (see Figure 1.1). The concentration
of the tracer in the effluent is then monitored from the time of adding the tracer
until it reaches a concentration approximating that of C 0 .
As stated before, the RTD curve can be easily obtained by injecting a tracer
in a pulse input, but now we will formulate a more general relationship between
a tracer injection (not necessarily a pulse input) and the corresponding tracer
concentration in the effluent (current leaving the system).
Chapter 2 presents a more general equation establishing the relation of the con-
centration of a tracer leaving a reactor (C out ) and the input concentration (C in ).
This is presented as the convolution integral:
t
′ ′ ′
Cout (t) = Cin (t − t )E(t )dt (1.9)
∫0
In this equation, E(t) is the corresponding RTD for the reactor. The concentra-
tion at the input is evaluated at different times and the resulting signal C out is a
convolution of C in and the RTD in the vessel.
In this way, in the following chapters we are able to calculate the E(t) curve
given any function of concentration input C in and measuring C out . For now, we
continue with more simple techniques, considering that the inlet concentration
is introduced in the form of a step input (Figure 1.1).
Let us consider a system of constant volumetric flow (Q) in which a tracer input
is introduced in the form of a step, with a rate of addition of tracer to the constant
input flow and which begins to occur at time t = 0. Before that moment, it is
considered that no tracer enters with the inlet current. Under these conditions:
Cin (t) = 0 for t < 0
Cin (t) = constant = C0 for t ≥ 0
1.2 Residence Time Distribution (RTD) Function 7

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.

1.2.2 RTD Concept in Heterogeneous Systems


Just as in the homogeneous reactors, the concept of RTD is valid in heteroge-
neous systems, with the particularity that there will exist an RTD related to the
flow of each phase involved in the reaction. For example, let us think about the
flow produced in a catalytic reactor where a gaseous species is introduced on one
side, and, simultaneously, the solid catalyst is introduced on the other. Figure 1.2
presents such a situation.
The experimental techniques to measure these RTD functions are equivalent
to the ones used in one-phase reactors. The application of such measurements
is extremely important in the design of reactors. For example, let us think of a
catalyzed fluidized bed, where gas is flowing at its bottom and circulating up.
The flow of the solid is very similar to that of the ideal CSTR, as the fluidized
bed is supposed to be perfectly mixed due to the fact that the mass transfer is
8 1 Nonideal Flow

Gas out Solid E(t)


output
Gas

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.

1.2.3 Characteristics of RTD


Going back to the use and measurement of RTD, it is important to know some
characteristics. If we consider the age of a reacting particle as the time it has
remained under the conditions in which the reaction is carried out, then E(t)
is related to the age distribution of the effluent. In fact, the E(t) is also called
function of distribution of exit ages.
The fraction of the output current that has remained in the reactor for a period
less than a given value, t, equals the sum over all times of the function E(t)⋅Δt, or
expressed continuously:
t [ ]
Fraction of output current that has remained
E(t) = = F(t)
∫0 in the reactor a time less than “t”
which represents the cumulative RTD function, F(t). Analogously, the integral of
the E(t) function from time “t” to infinity will represent the fraction of the output
current that has remained in the reactor a time higher than “t.”
This function F(t) (cumulative distribution function) can be calculated at var-
ious “t” values using the area of the curve of the representation of E(t) vs. time.
An example of curve F(t) for a tracer step input can be seen in Figure 1.3, where

Figure 1.3 Interpretation of the cumulative


distribution curve, F(t).
1.0

0.8

0.6 Step response

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.

1.2.3.1 Mean Residence Time


Variables described by distribution functions are characterized by their moments.
Correspondingly, function E(t) can also be characterized by its moments. In this
sense, the mean value of the variable is equal to the first moment of the RTD
function, E(t). So, the mean residence time is defined as

∫0 t ⋅ E(t)dt ∞
tm = = t ⋅ E(t)dt (1.12)
∫0

E(t)dt ∫0
When studying the ideal reactors, the spatial time (𝜏 = mass of catalyst/Q) and
the average residence time (t = V ∕Q) are usually used.
We can demonstrate that, regardless of the RTD that exists in a given reactor,
ideal or not ideal, the nominal average residence time, t = V ∕Q is equal to the
mean residence time of the RTD distribution, t m , if the system is of constant den-
sity (where Q is constant). For doing this, let us consider a reactor completely
filled with a colored fluid (for example, red bromine gas); at an instant t = 0,
we start to inject yellow gas to replace the bromine filling the reactor. That is,
the reactor volume V is equal to the volume occupied by the bromine leaving
the reactor. In such a situation, during a certain time dt, the volume of bromine
that will leave the reactor is Q⋅dt, “Q” being the volumetric flow rate (considered
at constant temperature and pressure). In that case, [1 − F(t)] will represent the
fraction of the gas that has remained in the reactor for a longer time. Since only
this red gas has remained in the reactor more than “t,” the volume of bromine,
dV , leaving the reactor at time “dt” is
dV = (Q ⋅ dt)[1 − F(t)] (1.13)
If we add all the bromine that has left the reactor in the period between
0 < t < ∞, we have

V = Q ⋅ [1 − F(t)] ⋅ dt (1.14)
∫0
For constant volumetric flow:

V =Q [1 − F(t)] ⋅ dt
∫0

V
= [1 − F(t)] ⋅ dt (1.15)
Q ∫0
For the integration of the right-hand side, we will use the integration by parts,
indicating that

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

At t = 0, the cumulative function is zero, F(t) = 0, and when t → ∞,


[1 − F(t)] = 0:
1
V
= ∞ ⋅ (1 − 1) − 0 ⋅ (1 − 0) + t ⋅ dF
Q ∫0
In this way, the first term on the right-hand side is zero, and we can write
1
V
=t= t ⋅ dF (1.17)
Q ∫0
On the other hand, since dF = E(t)dt:

t= E(t) ⋅ dt (1.18)
∫0
Now, the right side of this equality is, precisely, the mean residence time, so we
can conclude that spatial time and mean residence time are equal in systems of
constant density, regardless of whether the flow is ideal or not:
t = tm (1.19)
This result is only valid for closed systems (that is, without dispersion). The
volume of the reactor can now be easily determined:
V = Q ⋅ tm (1.20)

1.2.3.2 Second and Third Moments of the RTD


When comparing RTDs, it is usual to use the moments associated with the distri-
bution instead of using the entire distribution. To this end, three are the moments
that are usually used. The first moment is the average residence time, already
defined. The second moment is calculated from the mean and is the variance
which is defined as

𝜎2 = (t − tm )2 ⋅ E(t)dt (1.21)
∫0
This equation corresponds to the square of the standard deviation. The magni-
tude of this moment gives a measure of the dispersion of the distribution, in such
a way that the greater this moment, the greater the dispersion of the distribution
(see Figure 1.4).

Figure 1.4 Effect of the second


E(t) moment of the RTD in the curve
shape. Both distributions are
Low variance centered at the same tm .
High variance

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.

1.3 RTD in Ideal Reactors


1.3.1 RTD of the Batch and PFR Reactors
The RTD of these two types of reactors is the simplest we can consider, as already
mentioned. All the particles (considered very small portions of the fluid, or solid
particles if there is a solid moving in the reactor) that leave these reactors have
stayed the same time inside the reactor. The output is then infinite at a particular
time, and zero otherwise. This will produce an output function similar to that of
an arrow of infinite height and zero width but presenting an area equal to one.
This arrow appears at t = V /Q = t. Mathematically, this arrow is represented by
the Dirac delta function, in such a way that the E(t) for an PFR is
E(t) = 𝛿(t − t) (1.23)
The delta function has the following properties:
𝛿(x) = 0 when x ≠ 0
𝛿(x) = ∞ when x = 0 (1.24)

𝛿(x) dx = 1 (1.25)
∫−∞
b
g(x) ⋅ 𝛿(x − t) dx = g(t) if a < t < b (1.26)
∫a
b
g(x) ⋅ 𝛿(x − t) dx = 0 if t ∉ [a, b] (1.27)
∫a
12 1 Nonideal Flow

Figure 1.6 Dirac delta function, i.e. response of an ideal


PFR to a pulse tracer input when “x” is time and the
δ(x – x0) signal to a tracer is registered. In that case, x 0 would
represent the average residence time.

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.

1.3.2 RTD of an ideal CSTR


The main characteristic of a CSTR is that the concentration of all substances
inside the reactor are identical to those concentrations in the output stream.
Imagine that an amount M0 of inert tracer is injected into a CSTR. Once the
tracer in injected, we can do a material balance for inert tracer, and we obtain
Input − Output = Accumulation
dC
0−Q⋅C =V ⋅ (1.28)
dt
Note that the input is zero (we consider the balance when the tracer has been
already injected, i.e. at t > 0). Obviously, the generation term is also zero as the
tracer is inert.
Since the reactor is perfectly agitated, C is the tracer concentration at the outlet
and also at any point of the reactor. Separating variables and integrating, taking
into account that C = C 0 for t = 0, we obtain
C(t) = C0 ⋅ exp(−t ⋅ Q∕V ) = C0 ⋅ exp(−t∕t) (1.29)
an expression that allows to obtain the concentration of the tracer in the effluent
for any instant “t.”
If we take into account the definition of E(t) and introduce the value of C(t), we
obtain the RTD of an ideal CSTR:
C(t) C0 ⋅ exp(−t∕t)
E(t) = ∞ = ∞
(1.30)
∫0 C(t) ⋅ dt ∫0 C0 ⋅ exp(−t∕t) ⋅ dt
Evaluating the integral of the denominator:

C0 ⋅ exp(−t∕t) ⋅ dt = C0 ⋅ [t ⋅ exp(−t∕t)]∞
∫0 0

= −C0 ⋅ t ⋅ [exp(−∞) − exp(0)] = C0 ⋅ t


1.3 RTD in Ideal Reactors 13

Figure 1.7 Response of a CSTR to a pulse tracer E(t)


input.

t t

Thus, the RTD is obtained for an ideal CSTR (Figure 1.7):


1
E(t) = ⋅ exp(−t∕t) (1.31)
t
We have already seen that the average residence time in a reactor is given by
V /Q or t. This relationship can now be obtained in a simpler way by applying the
definition of the average residence time of an RTD for a CSTR:
∞ ∞ ( )
1 t
tm = t ⋅ E(t) ⋅ dt = t ⋅ ⋅ exp − ⋅ dt = t (1.32)
∫0 ∫0 t t
That is to say, the nominal residence time (spatial time) t = V /Q coincides with
the mean residence time that the material remains in the reactor, t m .
To know the degree of dispersion in the reactor, we will calculate the second
moment, the variance:
∞ ( )
(t − t)2 t 2
𝜎 =
2
⋅ exp − ⋅ dt = t (1.33)
∫0 t t
That is to say, 𝜎= t. Or in other words, the standard deviation of the RTD coin-
cides with the average.

1.3.3 RTD of PFR/CSTR in Series


In agitated tank reactors, there are zones, in the vicinity of the agitator, with a
high degree of agitation and where an ideal CSTR could be a valid model. Nev-
ertheless, depending on the location of the conductions, the reaction mixture
can follow a somewhat tortuous path when entering, or when leaving the area
perfectly agitated, or in both cases. This tortuous path can be modeled as if it
behaved like a PFR. Thus, the tank-type reactor can be modeled as if it were a
CSTR in series with a PFR (Figure 1.8), and the PFR can be before or after the
CSTR. Next, we study the RTD for this type of reactor.
Consider, first, the system constituted by a CSTR followed by a PFR (Figure 1.8).
The mean residence time in the CSTR will be called tt and the mean residence
time in the PFR, t p . If we inject a tracer pulse at the entrance of the CSTR, the con-
centration at the exit of the tank will vary with time according to the expression:
C(t) = C0 ⋅ exp(−t∕t t ) (from Eq.(1.29))
14 1 Nonideal Flow

Feed

CSTR PFR

CA, T
Product

Figure 1.8 Real reactor modeled as a CSTR and a PFR in series.

E(t) Figure 1.9 RTD for a CSTR and a PFR in


series.

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

1.4 Modeling the Reactor with the RTD


On many occasions, the flow inside a reactor is not adjusted either to the com-
plete mixture or to the plug flow, so when trying to use the RTD to predict the
conversion that we are going to obtain, we find that the ideal models so far no
longer serve us. There will, therefore, be a need to model the real reactor with
some type of combination of ideal reactors or introduce new models. In these
models, the adjustable parameter is usually evaluated on the basis of the RTD
analysis obtained with a tracer test. We classify the models according to the num-
ber of adjustable parameters that are extracted from the information provided by
the RTD.

1.4.1 Models with One Parameter: Tanks-in-series and Dispersion


Models
In these models, we use a single parameter to bear in mind the nonideal behavior
of a particular reactor. The parameter is determined by analyzing the measured
RTD in a tracer test, as mentioned before.
For modeling the nonideal behavior of CSTRs, usually a dead zone volume (V d ),
where the reaction does not take place, is used. Also, it is usual to consider the
existence of a part of the fluid that passes through the reactor in short circuit and
therefore does not react. In the case of tubular reactors, there are two models that
usually represent the flow: the tanks-in-series model and the dispersion model.
Both are one-parameter models, being the number of tanks in the first one, nt ,
and the dispersion coefficient, De , in the last one.
For both types of distributions, once the value of the parameter is known, we
will be able to calculate the conversion and or the concentrations at the output of
the reactor.
In the case of a nonideal tubular reactor, usually it is assumed that the fluid
moves in a plug flow through the reactor, so that each atom passes through the
reactor the same time, and that the velocity profile is flat and there is no axial
mixing. Both statements are false, to a greater or lesser extent, in all tubular reac-
tors. Two approaches are often used to compensate for failures in the two ideal
assumptions. In one case, the real reactor is modeled as a series of CSTRs of the
same size. In the other (dispersion model), an axial dispersion is superimposed
on the piston flow.

1.4.1.1 Tanks-in-series Model


In this model, we analyze the RTD of a particular tubular reactor to determine
the number of CSTRs in series that will present an RTD approximately the
same as the actual RTD. Next, we apply the balance equations valid for the ideal
CSTR in order to calculate the conversion. We first consider the case of three
tanks (Figure 1.10), developing the equations for the expression of the RTD, and
then generalize it for “nt ” reactors connected in series. In this way, we obtain an
equation that permits to calculate the number of tanks that best correlates the
data of the actual RTD.
16 1 Nonideal Flow

Feed

CA0

…..

CA1, V1

CAn, Vn

CA2, V2

Figure 1.10 Tanks-in-series model.

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

Figure 1.11 Response to a pulse nt = ∞


tracer input in function of the E(t)
nt = 10
number from tanks according to the
tanks-in-series model. In the figure, t
represents the average residence nt = 4
time of the whole system. nt = 2

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).

1.4.1.2 The Dispersion Model


In this model, it is considered that there is a dispersion of the material overim-
posed to the flow, and that this dispersion is governed by an expression analogous
to that of Fick for diffusion, which is superimposed on the plug flow. Thus, in
addition to the term (u⋅S) due to the flow of the mass of fluid, each component
of the mixture will be transported through any section of the reactor with an
additional rate [De ⋅S⋅(dC/dz)] due to molecular and turbulent diffusion. At first
glance, this simple model could only serve the effects of axial mixing. However,
it can be seen that it serves to compensate also for the effects of radial mixing
and those due to non-flat velocity profiles. These variations in concentration may
be due to different speeds and flow paths, as well as to molecular and turbulent
diffusion.
Let us imagine a pulse injection of tracer to a tubular reactor of section “S.” Dur-
ing the movement of the fluid through the reactor, the pulse widens and becomes
more diluted. The molar flow rate of the tracer (nT ), both by dispersion and con-
vection, is
( )
𝜕C
nT = u ⋅ S ⋅ CT + −De T ⋅ S (1.49)
𝜕z
In this expression, “z” is the spatial dimension where the fluid is moving, De
is the effective dispersion coefficient (m2 /s) and “u” is the superficial velocity.
1.4 Modeling the Reactor with the RTD 19

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

Closed recipient Open recipient


– + – + – + – +
z=0 z=0 z=L z=L z=0 z=0 z=L z=L

De = 0 De>0 De = 0 De>0 (same value of De>0)

Plug flow

z=0 z=L z=0 z=L


Flat profile of velocity Fluctuations due to dispersion

Figure 1.12 Effect of the dispersion on the velocity profile.

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

The injected tracer mass (M0 ) is given by



M0 = u ⋅ S ⋅ CT (0− , t) ⋅ dt (1.62)
∫0
In the closed–closed case, we do not have an analytical expression for the
E curve, but the curve can be calculated by numerical methods and, also, we
can calculate exactly its mean and variance. Bischoff and Levenspiel (in 1963)
found the following relationships for the mean residence time and variance for
this case:
t = tm
( )2 [ ( )]
𝜎 1
= 2 ⋅ Bo − 2 ⋅ Bo2 ⋅ 1 − exp − (1.63)
tm Bo
Experimentally, the dispersion module can be calculated from the values of t m
and 𝜎 2 obtained from the RTD, substituting in the previous expression.

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

In the open–open case, there exists an analytical solution of the differential


equation, with an expression of the E curve that is not much complicated:
( )2
⎡ t ⎤
⎢ 1 − ⎥
1 t
E= √ ⋅ exp ⎢− ⎥ (1.69)
4 ⋅ 𝜋 ⋅ Bo ⋅ t∕t ⎢ 4 ⋅ t ⋅ Bo∕t ⎥
⎣ ⎦
The corresponding average residence time is
tm = t ⋅ (1 + 2 ⋅ Bo) (1.70)
where t is based on the volume (and flow rate) between z = 0 and z = L; that is,
the reactor volume measured with a calibrated apparatus. Note that, as a result of
the previous equations, the average residence time in the open system is longer
than the one corresponding to the closed system. The variance will be
( )2
𝜎
= 2 ⋅ Bo + 8 ⋅ Bo2 (1.71)
tm
In the next two chapters, we discuss the numerical solution to these partial
differential equations systems and we present the problem of calculating the con-
version in systems with dispersion.

1.4.2 Models with Two Parameters


1.4.2.1 Two CSTR with Exchange of Matter
Consider the case that there is a strongly agitated region near the agitator of
an CSTR. Nevertheless, outside this region, the agitation is lower (Figure 1.13).
Both regions have an important exchange of material. We also consider that the
entrance and exit pipes are connected to the zone of greatest agitation. Each zone
will be modeled as a CSTR, both being connected and there is material transfer
between them.
If we propose a molar balance on the tracer, with a pulse injected at t = 0, for
each of the tanks, we obtain
Accumulation = Input − Output (there is no tracer generation)
dCT1
V1 = Q1 CT2 − (Q0 CT1 + Q1 CT1 ) (in the first tank) (1.72)
dt
Feed
Q0

Q1
CSTR CSTR
CA2, V2
CA1, V1 Q1

Product CA1, Q0

Figure 1.13 Real reactor and modeling using two CSTR.


1.4 Modeling the Reactor with the RTD 23

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 }).

1.4.2.2 CSTR with Dead Volume and Short Circuit


In this case, the real CSTR is modeled as the combination of an ideal CSTR of
volume V t , a dead zone of volume V d , and a bypass (short circuit) of volumetric
flow rate Qb (Figure 1.14). Using a tracer injection, we will calculate the parame-
ters of the model V t and Qt , since the total volume and the volumetric flow are
known.
24 1 Nonideal Flow

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

0.4 α = 0.5; β = 0.3 α = 0.5; β = 0.1 Ideal CSTR

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.

1.5 Other Models of Real Reactors Using CSTR and PFR


So far, we have discussed about various reactor models. All of them are based on
physical observations, which, in almost all the stirred tanks, are based on the exis-
tence of a well-mixed zone in the proximity of the agitator, usually represented
by a CSTR. The region out of this well-agitated area can be modeled in various
ways. The simplest form is using a model implying a CSTR connected to a dead
zone; if it is suspected that some of the feed to the reactor may short out, a bypass
current is added.
When the models do not satisfactorily represent the deviations of the ideal flow,
we have to try more complicated models. In these models, it is assumed that the
real reactor is constituted by a series of regions (flow in piston, flow dispersed in
piston, flow in complete mixture, dead volumes) interconnected with each other
in different ways (flow in bypass, with recirculation or cross-flow).
The simple types of these models are shown in Figure 1.16, where the form of
the tracer response, in terms of E or F curves, for different models can be seen.

Example 1.1 Tubular Reactor


A tubular reactor was designed in order to obtain a conversion of 98% and process
0.03 m3 /s. The reaction is a first-order irreversible isomerization. The reactor is
3 m long, with a cross-sectional area of 0.1 m2 . In the newly constructed reactor, a
tracer pulse test gave the following data: t m = 10 s and 𝜎 2 = 65 s2 . What conversion
can be expected in the real reactor?
26 1 Nonideal Flow

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

Vp2 Vp1 t Vp2 Vp1 t


t2 = Q2
t1 = Q1
t2 = Q2
t1 = Q1

(1 + r)Q
E(t) F(t)
Q Vp Q
rQ

Figure 1.16 Combinations of ideal reactors used to model real reactors.


1.5 Other Models of Real Reactors Using CSTR and PFR 27

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

Example 1.2 Reaction in a Complex System


The second-order reaction 2A → B is going to be carried out in a CSTR showing
both a short circuit and a stagnant region. The tracer concentration at the outlet
of this reactor is shown in the table, when a step input with initial concentration
of tracer equivalent to 10 mg/l was used.

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)

ln(CT /(CT0 – CT)) vs. time


5

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

Clearing C A , the concentration at the exit of the system:


Qb CA0 + CAs Qt
CA = (1.88)
Q0
So, finally:
CA (1 − 𝛽)2
= 1 − XA = 𝛽 + (1.89)
CA0 (1 − 𝛽) + 𝛼tk
the expression that allows to calculate the conversion based on the parameters of
the model.
As in Example 1.2, the reaction is second order, we can write:
2
Qt CA0 − Qt CAs − kCAs Vt = 0 (1.90)
Rearranging:
Vt ⋅ k
⋅ CAs
2
+ CAs − CA0 = 0 (1.91)
Qt
Let us be aware that the flow rate and volume in the previous equations are
those in the well-stirred part of the reactor, in such a way that:
Vt 𝛼V 𝛼
= = t
Qt (1 − 𝛽)Q0 (1 − 𝛽)
and clearing, the value of C As results to be

𝛼
−1 + 1 + 4 1−𝛽 tkCA0
CAs = 𝛼
(1.92)
2 1−𝛽 tk

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

Example 1.3 Second-order Reaction in a Series of CSTRs


An irreversible second-order reaction occurs in an isothermal, but not ideal,
CSTR. The reactor volume is 1000 l and the flow velocity of the reagent stream
is 1 l/s. At reactor temperature, k = 0.005 l/(mol s). The concentration of A in the
feed stream is 1 mol/l. The DTR is obtained by a tracer test in this reactor at the
desired feed rate and reaction temperature. Calculate the conversion that can be
obtained with the tanks-in-series model.
RTD data obtained:

t (s) 0 5 10 25 40 70 100 175 250 325


E(t) 9.712 2.206 6.411 0.000 0.000 0.000 0.000 0.000 0.000 9.712
(s−1 ) 18⋅10−6 86⋅10−5 68⋅10−5 1090 4 199 787 286 936 476 181 619 435 719 564 18⋅10−6
t (s) 400 700 1000 2 500 4 000 7 000 10 000 15 000 20 000
E(t) 0.000 0.000 0.000 6.954 4.508 1.200 2.505 6.80 1.60
(s−1 ) 782 749 791 834 627 888 06⋅10− 5 28⋅10−6 36⋅10−8 25⋅10−11 646⋅10−16 549⋅10−20
1.5 Other Models of Real Reactors Using CSTR and PFR 31

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

In a second reactor connected in series with this, we would have


CA0
CA2 = (1.96)
n−1 n−1
(1 + kt 1 CA1 )(1 + kt 2 CA2 )
For nt tanks in series, the product in the denominator would have nt factors
and the intermediate concentrations C A1 , C A2 , … are needed.
In the present example, a value of nt = 2.221 tanks would represent the system,
but it is not possible to do the balance in a part of a reactor. In such situations, the
best solution is to take the nearest integer to nt or take both limits, in the present
case nt = 2 and nt = 3.
Assuming two tanks t i = t∕2, i.e. all tanks of the same volume, we have
t i = 500 s
We can now calculate C A1 using (Eq. (1.95)) with the corresponding values of k
and n. Easily, the equation gives C A1 = 0.461 mol/l. If a second reactor is consid-
ered, (Eq. (1.96)) would give us C A2 = 0.274 mol/l, that is, a conversion of 0.723.
On the other hand, if three tanks are considered, the residence time in each
tank is 333.3 s and the solution C A1 = 0.533 mol/l, C A2 = 0.326 mol/l, and
C A3 = 0.245 mol/l. The final conversion would be 0.754.

Example 1.4 Nonideal CSTR


In an isothermal nonideal CSTR, an irreversible second-order reaction is carried
out. The values of flow rate, volume, and rate constant are known, and the RTDs
has been measured with a tracer. Indicate the equations necessary to determine
the concentration at the output of the reactor as a function of the input flow (or
average residence time), if the reactor is modeled: (a) as three equal CSTRs in
series, or (b) as three equal CSTRs in parallel.

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

Convolution and Deconvolution of Residence Time


Distribution Curves in Reactors

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:

C(u) = f (x) ⊗ g(x) = f (x) ⋅ g(u − x) ⋅ dx


∫space

= g(x) ⊗ f (x) = g(x) ⋅ f (u − x) ⋅ dx (2.1)


∫space
Chemical Reactor Design: Mathematical Modeling and Applications,
First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
36 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors

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

Figure 2.1 Visual example of the convolution of two functions.

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.

2.2.1 Convolution Properties


Convolution operation has the mathematical properties of commutativity, asso-
ciativity, and distributivity, i.e.:
f (x) ⊗ g(x) = g(x) ⊗ f (x) (2.2)

f (x) ⊗ [g(x) ⊗ h(x)] = [f (x) ⊗ g(x)] ⊗ h(x) (2.3)

f (x) ⊗ [g(x) + h(x)] = [f (x) ⊗ g(x)] + [f (x) ⊗ h(x)] (2.4)


It is easy to see a demonstration of the commutative property. First, we start
with the convolution integral written one way:
+∞
C(u) = f (x) ⊗ g(x) = f (x) ⋅ g(u − x) ⋅ dx (2.5)
∫−∞
Now we can substitute the variables, replacing (u − x) with a new variable x* :
x∗ = u − x, dx∗ = −dx (2.6)
−∞
C(u) = − f (u − x∗ ) ⋅ g(x∗ ) ⋅ dx∗ (2.7)
∫+∞
In Eq. (2.7), the sign of the variable of integration has changed, and so we have
to change the signs of the limits of integration. Because these limits are infinite,
the shift of the origin does not modify the magnitude of the limits.
Now we can reverse the order of the limits, which will change the sign of the
equation, and interchange the order of the functions “g” and “f .”
+∞
C(u) = f (u − x∗ ) ⋅ g(x∗ ) ⋅ dx∗ (2.8)
∫−∞
It does not matter if we change the name of the integration variables, so we can
call it “x” again and we have the result we wanted to show.
In this way, the function C(u) will contain the convoluted values of functions
f (x) and g(x) in the whole space. Nevertheless, in chemical reactor characteriza-
tion, the independent variable will be time, so surely the lower limit is fixed at
t = 0. In the same way, sometimes we are required to calculate the convolution
until a given time “t,” so the upper limit of the integration will be this time, in
such a way that, usually:
t
C(u) = f (u − t ′ ) ⋅ g(t ′ ) ⋅ dt ′ (2.9)
∫0
38 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors

Cin Figure 2.2 Input signal to a reactor,


RTD, and output convoluted signal.
E
Cout

2.2.2 Application to a Reactor RTD


Let C in (t) and C out (t) be the concentration of tracer (in mg/l, for example) at
the inlet and outlet of a continuous system (Figure 2.2). When the convolution
theorem is applied to the RTD in a reactor or a combination of reactors, for
example, to calculate the signal exit of a reactor with a given RTD when the input
signal is known, the solution of the system also implies the calculation of the asso-
ciated times to the input and output signals (see Figure CDC-2.2). In the case
shown in the figure, the reactor distribution is E and the signal input is C in , and
we need to convolute both signals to get C out :
Cout = E ⊗ Cin (2.10)

2.2.3 Calculating Convolution Functions


For functions of a discrete variable “x” (i.e. arrays of numbers), the definition of
the convolution operation is


C(u) = f (x) ⋅ g(u − x) (2.11)
u=−∞

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

We can write this in matrix form as


⎛C ⎞ ⎡ f1 0 0 0⎤
⎜ 1⎟ ⎢ ⎥
⎜C2 ⎟ ⎢ f2 f1 0 0⎥
⎜ ⎟ ⎢ ⎥
⎜C3 ⎟ ⎢ f3 f2 f1 0 ⎥ ⎛g1 ⎞
⎜C ⎟ ⎢ f ⎜ ⎟
f3 f2 f1 ⎥ ⎜g2 ⎟
⎜ 4⎟=⎢ 4 ⎥⋅⎜ ⎟ (2.12)
⎜C5 ⎟ ⎢ f5 f4 f3 f2 ⎥ ⎜g3 ⎟
⎜ ⎟ ⎢ ⎥
⎜C6 ⎟ ⎢ 0 f5 f4 f3 ⎥ ⎜⎝g4 ⎟⎠
⎜ ⎟ ⎢ ⎥
⎜C7 ⎟ ⎢ 0 0 f5 f4 ⎥
⎜C ⎟ ⎢ ⎥
⎝ 8 ⎠ ⎣0 0 0 f5 ⎦
or
C =A⋅g (2.13)
with A being the “matrix convolution,” with values of f (x) in the form of columns.
For a general case, when vectors of n and m values are combined, the resulting
matrix convolution is an (n + m − 1) × m matrix and the resulting C is a vector
of (n + m − 1) values:
C( n+m−1) × 1 = [A(n+m−1) × m) ⋅ g(m × 1) ] (2.14)
Let us call “v” to the amount of rows of the C vector, i.e. v = (n + m − 1). Let us
apply these equations to the convolution of an RTD curve, E, and the signal input
to a reactor, C in . As mentioned before, we will need to convolute both signals, in
such a way that:
Cout = E ⊗ Cin (2.15)
If C in has “m” values and E has “n” values, the convolution is given by
⎡E 0 0 ··· 0⎤
⎢ 1 ⎥
⎛Cout,1 ⎞ ⎢E2 E1 0 ··· 0⎥
⎜ ⎟ ⎢ ⎥
⎜Cout,2 ⎟ ⎢E3 E2 E1 ··· 0 ⎥ ⎛ Cin,1 ⎞
⎜C ⎟ ⎢ ⎜ ⎟
⎜ out,3 ⎟ = ⎢ ⋮ ⋮ ⋮ ⋮ ⋮ ⎥⎥ ⎜ Cin,2 ⎟
⋅⎜ ⎟ (2.16)
⎜Cout,4 ⎟ ⎢ 0 En En−1 ··· · · ·⎥ ⎜ ⋮ ⎟
⎜ ⎟ ⎢ ⎥ ⎜
⎜ ⋮ ⎟ ⎢ En ⎥ ⎝Cin,m ⎟⎠
⎜ ⎟ ⎢ ⎥
⎝Cout,v ⎠ ⎢ ⋮ ⎥
⎢0 0 0 0 En ⎥⎦

This is
Cout (v×1) = Cout(n+m−1)×1 = [A(n+m−1)×m ⋅ Cin(m×1) ] (2.17)
In order to calculate the time vector associated to C out , we would need the cor-
responding times associated to C in and E. In general, let us assume that the first
and last element of C in corresponds to times t0,Cin and tm,Cin while the first and
last element of the E vector are t 0,E and t n,E . For the convolution to be possible
40 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors

(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).

Example 2.1 Manual Calculation of Convolution


In an experimental reactor, a tracer is injected with the following distribution:

t (s) 0 1 2 3 4
C (mol/l) 0 8 4 6 0

In the reactor, the RTD is given by

t (s) 2 3 4 5 6
E 0 0.05 0.50 0.35 0

Calculate the evolution of the output signal with time.

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

To calculate the concentration:


Cconvolution1 = Cin1 ⋅ E1
Cconvolution2 = Cin2 ⋅ E1 + Cin1 ⋅ E2
Cconvolution3 = Cin3 ⋅ E1 + Cin2 ⋅ E2 + Cin1 ⋅ E3
Cconvolution4 = Cin4 ⋅ E1 + Cin3 ⋅ E2 + Cin2 ⋅ E3 + Cin1 ⋅ E4
Cconvolution5 = Cin5 ⋅ E1 + Cin4 ⋅ E2 + Cin3 ⋅ E3 + Cin2 ⋅ E4 + Cin1 ⋅ E5
Cconvolution6 = Cin5 ⋅ E2 + Cin4 ⋅ E3 + Cin3 ⋅ E4 + Cin2 ⋅ E5
Cconvolution7 = Cin5 ⋅ E3 + Cin4 ⋅ E4 + Cin3 ⋅ E5
Cconvolution8 = Cin5 ⋅ E4 + Cin4 ⋅ E5
Cconvolution9 = Cin5 ⋅ E5
Tanking the values of C in = tracer input and E = RTD of the reactor, we can
easily calculate:
Cconvolution1 = 0⋅0
Cconvolution2 = 8⋅0 + 0 ⋅ 0.05
Cconvolution3 = 4⋅0 + 8 ⋅ 0.05 + 0 ⋅ 0.5
Cconvolution4 = 6⋅0 + 4 ⋅ 0.05 + 8 ⋅ 0.5 + 0 ⋅ 0.35
Cconvolution5 = 0⋅0 + 6 ⋅ 0.05 + 4 ⋅ 0.5 + 8 ⋅ 0.35 + 0⋅0
Cconvolution6 = 0 ⋅ 0.05 + 6 ⋅ 0.5 + 4 ⋅ 0.35 + 8⋅0
Cconvolution7 = 0 ⋅ 0.5 + 6 ⋅ 0.35 + 4⋅0
Cconvolution8 = 0 ⋅ 0.35 + 6⋅0
Cconvolution9 = 0⋅0
And, finally:

C convolution (mol/l) 0 0 0.4 4.2 5.2 4.4 2.1 0 0

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

From this equation, it is not possible to calculate directly matrix A. If we have


a look at Eq. (2.16), we can check that, for the calculation of n values of vector E,
we have (n + m − 1) equations, so it is an overdetermined system of equations.
This means that there are many solutions for the system and we need to choose
in any way the best solution. In this way, for the resolution of the deconvolution
of two given signals (C in and C out ), it is necessary to pose an optimization by least
squared residuals of the components of matrix A.
For calculating the E curve, the easiest way to do the optimization is to min-
imize an objective function (O.F.) in the form of the sum of squared residuals
between the C out curve (known) and the product [A⋅Cin ]:
O.F. = [Cout(v×1) − A(v×m) ⋅ Cin(m×1) ]2 (2.22)
For that purpose, an initial value of the vector E is taken, and the use of an
optimization method is necessary.
For the calculation of the corresponding time vector for E distribution (decon-
voluted signal), we have
t0,E = t0,Cout − t0,Cin
(initial time of the convolution is the sum of the initial times)
tv,E = tv,Cout − tm,Cin
(final time of the convolution is the sum of the final times)

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).

Example 2.2 Manual Calculation of Deconvolution


In an experiment to determine the E(t) of a certain reactor, the tracer is intro-
duced in the reactor following the equation C in = t ⋅ exp(−t/2). The response of
the nonideal tank reactor is given in the following table:

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

to calculate two values of E(t). Let us choose to use m = 5 points in C in , so we will


calculate six points in E(t). Choosing t 0,in = 0, then:

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

The following figure presents graphically the results.


6

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

2.4 Computer Program Using Matlab (Convolution) ®


Numerical computing tools such as Matlab can facilitate the development of dif-
ferent aspects in engineering. Compared to other computational tools and pro-
gramming languages, such as C++ and Java, Matlab is easier for students to learn,
and it has been widely incorporated in advanced courses in certain engineering
disciplines. Even through Matlab is a commercial software, a freeware package
is available (the software is called “GNU Octave”); and it can be used with small
modifications of the programs, as almost all functions are compatible.
Although Matlab (and GNU Octave) contain specific functions such as
“conv” and “deconv” for convolution and deconvolution, respectively, a simple
use of these is not valid for the calculation of RTD curves (especially the
deconvolution function). Furthermore, those programs are able to correctly
calculate the convolution of two signals, but the corresponding time vectors are
obviated.
For the convolution to be calculated using Matlab, let us assume a system where
C in (n values) and E (m values) and the corresponding time vectors (t0,Cin … tn,Cin )
and (t 0, E … t m, E ) are known. Note that the time increment in both vectors should
be the same.
It is easy to arrange matrix A from the given values in a for-end loop in Matlab,
and then calculate the convolution simply using relation Eq. (2.16). The use of the
“conv” function would give exactly the same result in the major part of the cases.
Calculating the corresponding time vector for C out is simple taking into account
that Δt convolution = Δt, as shown in Eq. (2.20). The time vector goes from the sum
of the initial times (t0,Cin + t0,E ) to the sum of the final times (tn,Cin + tm,E ). Let us
remember that neither Matlab nor GNU Octave calculates this time vector.
The following examples are simple applications of the previous equations. More
complicated input–output signals could be used, but maybe it is easier to under-
stand the method if we use simple examples.

Example 2.3 Convolution of Two Given Signals


Let us have the following data for the input signal to a reactor:

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 the E distribution is given by

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

Program Listing 2.1


Cin=[0 0 8 6 4 5 6 3 1 0 0]; % Input signal to the reactor
tin=[0 1 2 3 4 5 6 7 8 9 10]; % Time vector associated to Cin
Er=[0 0.05 0.5 0.35 0.1 0]; % RTD distribution in the reactor
tr=[0 1 2 3 4 5]; % Time vector associated to E
inct=tr(2)-tr(1); % Time increment for Cin and E
m=length(Er); % Dimension of vector E
n=length(Cin); % Dimension of vector Cin

A=zeros(m,n); % Initializing the matrix

for i=1:n
A(i:(m-1)+i,i)=Er; % Construction of the matrix
convolution
end

Cout=(A*Cin')'; % Calculation of the convoluted curve

% Calculation of time vector for the convoluted curve:

t0convolution=(tin(1)+tr(1)); % Initial time for the


convoluted curve
tendconvolution=(tin(n)+tr(m)); % Final time for the
convoluted curve
tconvolution=t0convolution:inct:tendconvolution;
% Time vector for the convoluted curve

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.

Example 2.4 Convolution of a Given Vector with an Ideal Reactor. Contin-


uously Stirred Tank Reactor
Let us have the following data for the input signal to a continuously stirred tank
reactor (CSTR):

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

8 Figure 2.3 Solution to


Cin Example 2.3. Convolution of
7 RTD two given signals.
Cout

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

Figure 2.4 Solution to 8


Example 2.4. Convolution of a Cin
given signal with the RTD of a 7 RTD
Cout
CSTR.
6

0
0 2 4 6 8 10 12 14 16 18 20
Time (s)

Figure 2.5 Solution to 8


Example 2.5. Convolution of a Cin
given signal with the RTD of a 7 RTD
Cout
PFR.
6

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.

2.5 Computer Program Using MATLAB (Deconvolution)


It is more complicated to calculate the deconvolution of the curves C in and C out
to obtain the E function. In this case, the use of the “deconv” function is not able
48 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors

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:

Program Listing 2.2


function OF=minimal(f0) % f0 is the E curve being optimised
global Cin Cout
n=length(f0); % Dimension of the resulting E vector
v=length(Cout); % Dimension of vector Cout
m=length(Cin); % Dimension of vector Cin
for i=1:m
A(i:i+n-1,i)=f0;% Constructing the convolution matrix
end
OF=sum((Cin'-A'*Cout'). ̂ 2);

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:

Program Listing 2.3


global Cin Cout

Cin=[0 0 8 6 4 5 6 3 1 0 0]; % Input signal to the reactor


tin=[0 1 2 3 4 5 6 7 8 9 10]; % Time vector associated to Cin

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];

inct=tin(2)-tin(1); % Time increment for Cin and Cout

v=length(Cout); % Dimension of vector Cout


m=length(Cin); % Dimension of vector Cin

n=v-m+1; % Dimension of the resulting E vector

f0=ones(1,n)/2; % First assumption

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.

Example 2.6 Deconvolution of Two Given Signals


Let us have the following data for the input signal to a reactor:

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:

Program Listing 2.4


Cin=[0 0 8 6 4 5 6 3 1 0 0]; % Input signal to the reactor
tin=[0 1 2 3 4 5 6 7 8 9 10]; % Time vector associated to Cin

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];

inct=tin(2)-tin(1); % Time increment for Cin and Cout

v=length(Cout); % Dimension of vector Cout


m=length(Cin); % Dimension of vector Cin

n=v-m+1; % Dimension of the resulting E vector

f0=ones(1,n)/2; % First assumption

Erd=fminsearch(@minimal,f0); % Minimization of the O.F.

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

8 Figure 2.6 Solution to


Cin Example 2.6. Deconvolution.
7 Cout
Er

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.

2.6 Convolution of Signals in Reactors Connected


in Series
The convolution and deconvolution of signals is also important in systems where
more than one reactor is combined in series. In such cases, 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 the next chapters, we treat
more complex systems in detail, but let us now treat systems with two or three
reactors.
Figure 2.7 shows a system of three reactors connected in series. The convolution
of the tracer curve with the RTD of each reactor follows the equations:
Cout1 = E1 ⊗ Cin
Cout2 = E2 ⊗ Cout1 = E2 ⊗ E1 ⊗ Cin
Cout3 = E3 ⊗ Cout2 = E3 ⊗ E2 ⊗ E1 ⊗ Cin
In this case, we must do three convolution operations to get the final signal,
with the difficulties that this would have. If C in were a pulse input, it would
2.6 Convolution of Signals in Reactors Connected in Series 51

Pulse Cout3
input
Reactor 1 Reactor 2 Reactor 3
Cout1 Cout2

Figure 2.7 Set of three reactors connected in series.

be eliminated in the convolution. Assuming that the input is a perfect pulse,


we will have
M
Cout1 (t) = E (t) (see Chapter 1)
Q 1
The first convolution, in reactor 2 will be, from Eq. (2.8):
t ∑
t∕Δt ′
Cout2 (t) = Cout1 (t ′ ) E2 (t − t ′ ) dt ′ =
Cout1 (t − nΔt ′ )E2 (nΔt ′ ) Δt ′
∫0 n=1
Note that the integral is taken over time from 0 to time “t.” In the sum, the
number of summands will depend on the number of increments used in the cal-
culation. Cout1 (t − nΔt ′ ) represents the value of C out1 in “n” time intervals with
an increment of Δt ′ . Similarly, E2 (nΔt ′ ) is the RTD function of the second reac-
tor at the different times. Note that the number of increments is t∕Δt ′ for the
integration of the curve until this time “t.”
In the same way, for the signal obtained after the third reactor:

t
′ ′ ′

t∕Δt
′ ′ ′
Cout3 (t) = Cout2 (t ) E3 (t − t ) dt = Cout2 (t − nΔt ) E2 (nΔt ) Δt
∫0 n=1
As we see, the combination of several reactors involves the calculation of the
convolutions of each curve, which makes it very tedious. There are, as we see in
the next chapter, math techniques that allow us to approach the calculation of
more complex systems, and involve the use of integral transforms.

Example 2.7 Pulse Input Tracer


The response to a pulse input tracer is given by the following equations:
Cf = t − 2 when 2 ≤ t ≤ 5
Cf = 3 when 5 ≤ t ≤ 8
Cf = 11 − t when 8 ≤ t ≤ 11
Cf = 0 elsewhere
Calculate the average residence time and the function E(t) in this reactor. If this
signal is fed to a CSTR of t = 1, calculate the output signal. Note that units of time
are not specified.

Pulse input Reactor 1 C


Cf
52 2 Convolution and Deconvolution of Residence Time Distribution Curves in Reactors

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

Example 2.8 Experimental Solid Flow Reactor


An experimental flow reactor, initially designed as a CSTR, is fed with small par-
ticles of solids, and the response to an impulse signal (solid impregnated in a
colored compound) has the function: C(t) = a(10 − t) (with time in minutes).
(a) Represent approximately the function C(t) and calculate E(t) of the solid.
(b) This reactor is followed by a fluidized bed reactor of mean residence time
(of the solid) equal to three minutes. Determine the concentration of tracer
at the output of the system.

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

where we know the value of t2 = 3 min. The convolution of both signals is


obtained at the exit of this second reactor, and then:

E1 (t) ⋅ E2 (t − t ) ⋅ dt
′ ′
Cout =
∫0
In the next chapter, we see a way to solve this equation analytically, but for
now let us use the method mentioned earlier, by taking values of time with
the same increment for both functions. Using a value of Δt = 1 min, we can
construct the following table of data:

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.

Program Listing 2.5


t=0:12; % Values of time (inct=1);
inct=1;

E1=0.02*(10-t) % Values of E for the first reactor (see Table)


E2=1/3*exp(-t/3); % Values of E for the second reactor (see Table)

m=length(E2); % Dimension of vector E2


n=length(E1); % Dimension of vector E1 (in this case equals to m but
not in general

A=zeros(m,n); % Initializing the matrix

for i=1:n
A(i:(m-1)+i,i)=E2; % Construction of the matrix convolution
end

Cout=(A*E1')'; % Calculation of the convoluted curve


Bibliography 55

% Calculation of time vector for the convoluted curve:

t0convolution=(2*t(1)); % Initial time for the convoluted curve


tendconvolution=(2*t(n)); % Final time for the convoluted curve
tconvolution=t0convolution:inct:tendconvolution; % Time vector for
the convoluted curve

plot(t,E1,t,E2,tconvolution,Cout)
legend('E1','E2','Cout')
xlabel('time(s)')

The following figure is obtained in the calculation.


0.35
E1
0.3 E2
Cout

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

Use of Transfer Function for Convolution and


Deconvolution of Complex Reactor Systems

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)

3.2 Definition and Properties of the Transfer Function


The transfer function is defined as the quotient between the Laplace transform of
the output function and the Laplace transform of the input function, under the
assumption that the initial conditions are zero (equal to zero). This is valid for a
linear and time-invariant system.
This condition can be seen as a convolution process, formed by the input exci-
tation convolved with the considered system, resulting in the response within a
time interval. Now, in that sense (that of the convolution), it must be observed
that the transfer function is formed by the deconvolution between the input signal
and the system.
By definition, a transfer function can be determined by the expression:
Response to stimulus Y (s)
H(s) = = (3.1)
Forcing function X(s)
where H(s) is the transfer function, Y (s) is the Laplace transform of the response
signal, and X(s) is the Laplace transform of the input signal.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
58 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

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)

3.3 Laplace Transform


The Laplace transform is a type of integral transform frequently used for the
resolution of ordinary differential equations (ODEs). The Laplace transform of
a defined function f (t) (in differential equations, in mathematical analysis or
in functional analysis …) for all positive numbers t ≥ 0, is the function F(s),
defined by

F(s) = L{f (t)} = e−st ⋅ f (t)dt (3.5)
∫0
The Laplace transform converts a time domain function (where time is an inde-
pendent variable) to the Laplace domain (where “s” is the independent variable).
This transform can also be used, obviously, for other systems where “t” would
represent a real variable. Contrarily, “s” is, in general, a complex variable of the
form s1 + i⋅s2 . Nevertheless, in this book, all the Laplace transforms are in the
real domain.
Let us discuss and explain some of the properties of the Laplace transform.
Table 3.1 shows the most important properties.
One of the most important properties of the Laplace transform is related to
the calculation of the derivatives, very common in reactor design. Note that to
calculate df (t)/dt in the Laplace space, we just need to evaluate “s⋅f (s),” and then
the Laplace transform of an integral can be calculated by “f (s)/s.”

3.3.1 Laplace Transform of Some Important Functions for Reactor


Characterization
Here are some important functions in reactor design and characterization.
3.3 Laplace Transform 59

Table 3.1 Properties of the Laplace transform.

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)

L{f (t)} = s2 L{f (t)} = s2 ⋅ f (s)


′′

L{f n (t)} = sn L{f (t)} = sn ⋅ f (s)


Integral of a function:
{ }
t f (s)
L ∫0 f (t)dt = s
Pure delay theorem. The Laplace transform of a function f (t) delayed “a” units of time is
L {f (t − a)} = e−s ⋅ a ⋅ f (s)
And then:
L {eat f (t)} = f (s − a)
Final value theorem. If the temporary limit exists, the following fulfills:
lim f (t) = lim sf (s)
t→∞ s→0

Therefore:
f (∞) = lim[sf (s)]
s→0

3.3.1.1 Ramp Function

Ramp function

Slope = K

Time
60 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

Suppose f (t) = Kt, applying the Laplace transform:


∞ ∞
L{Kt} = Kte−st dt = K te−st dt
∫0 ∫0
and integrating by parts, we obtain
K
L{Kt} =
s2
3.3.1.2 Sinusoidal Function

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

This function is particularly useful in reactor characterization, as it is generally


used to determine the residence time distribution (RTD). The function is a Dirac
delta, which can be represented by the derivative of a unitary step function:
d
𝛿(t) = U(t)
dt
3.3 Laplace Transform 61

This unitary step can be written as


U(t) = lim (1 − e−at )
a→∞

and using these relations, we can write


{ }
d
L{𝛿(t)} = L lim (1 − e ) = lim L(ae−at )
−at
dt a→∞ a→∞

Taking the Laplace transform:


a
lim L(ae−at ) = lim =1
a→∞ a→∞ s + a

showing that the Laplace transform of a pulse is the unity, as stated before.

3.3.1.4 Other Functions


Table 3.2 shows the most important Laplace transform function generally used.

Table 3.2 Laplace transform of basic functions.

f (t) F(s) = L{f (t)}

Pulse function, Dirac delta 1


c
c
s
1
U(t), unitary step
s
⎧ ⎫
⎪f (t − a) t ≥ a⎪
f (t) = ⎨ ⎬ e−as f (s)
⎪ 0 t<a ⎪
⎩ ⎭
n!
tn
t n+1
1
eat s>a
s−a
n!
t n ⋅ eat s>a
(s − a)n+1
a
sin(at)
s2 + a 2
s
cos(at)
s2 + a 2
a
sinh(at)
s2 − a 2
s
cosh(at)
s2 − a 2
s+a
e−at cos(wt)
(s + a)2 + w2
w
e−at sen(wt)
(s + a)2 + w2

𝜋
t −1/2
s
62 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

3.4 Use of Laplace Transform in Chemical Reactor


Characterization
If we introduce in a reactor a pulse input, as we have seen, X(s) = 1, and then the
transfer function of the reactor is the Laplace transform of the signal at the exit,
Y (s), that equals the RTD in the reactor:
Y (s)
H(s) = ; if X(s) = 1; Y (s) = H(s) (3.6)
X(s)
If we introduce in a reactor a step input:
1
X(s) = (3.7)
s
and then,
H(s) = Y (s) ⋅ s (3.8)
Having in mind that:
[ ]
df (t)
L = s ⋅ f (s) (3.9)
dt
it is fulfilled:
dY (t)
L−1 [s ⋅ Y (s) ] = (3.10)
dt
and then,
t
y(t) = H (t)dt (3.11)
∫0
This means that the exit Y (t) is the integral function of the RTD, as expected
when working with a step input.

3.4.1 Study of the RTD in the CSTR


We will study the different reactors using several input signals and using the con-
cept of transfer function or direct solution with ODEs. This will allow us to later
solve more complex systems.
(1) Step input, direct solution of the ODE:
If we consider the input and system shown in Figure 3.1.
A mass balance in nonsteady state has the following expression (see Chapter 4
for details):
Input + generation = Output + Accumulation
( )
dCout
Q ⋅ Cin + 0 = Q ⋅ Cout + ⋅V (3.12)
dt
Dividing by the flow rate, and rearranging:
( )
dCout
Cin = Cout + t (3.13)
dt
3.4 Use of Laplace Transform in Chemical Reactor Characterization 63

Tracer
concentration Feed

Step input

C A, T
Product
0 Time

Figure 3.1 Step input of tracer to a CSTR.

where, as usual, t = VQ . The former equation has as initial condition: C out = 0


at t = 0. Then we can write
dCout 1
= (Cin − Cout )
dt t
t
ln(Cin − Cout ) + = ln(Cin ) (3.14)
t
and, finally:
Cout
F(t) = = 1 − exp (− t∕t) (3.15)
Cin
This is the “F” curve (integration of the RTD) mentioned in many textbooks
for the continuous stirred tank reactor (CSTR).
(2) Step input, solution using Laplace transform:
If u(t) is the step input function, the mass balance can be written as
In + Generation = Out + Accumulation
( )
dCout
u(t) = Cout + t (3.16)
dt
As for the derivative, the following is fulfilled:
[ ]
dCout (t)
L = s ⋅ Cout (s) (3.17)
dt
and for the step input, the Laplace transform is
Δu
u(s) = (3.18)
s
In the previous equation, Δu is the value of the change in the step input. We
can write
Δu
= Cout (s) + tsCout (s) (3.19)
s
Solving:
Δu Δu Δu
Cout (s) = = − (3.20)
s(1 + ts) s (1 + ts)
64 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

Looking for the inverse of the Laplace transform:


[ ( )]
t
Cout (t) = Δu 1 − exp − (3.21)
t
and, therefore:
Cout (t)
F(t) = = 1 − exp (− t∕t) (3.22)
Δu
As we would expect, Eq. (3.22) and (3.15) are the same.
(3) Pulse input, direct solution with the differential equation:
Let us now think about a pulse input injected into a CSTR. In that case, the
mass balance is
Input + Generation = Output + Accumulation
( )
dCout
0 = Q ⋅ Cout + V (3.23)
dt
But with the initial condition of this ODE as
M
t = 0 → Cin (0) = (3.24)
V
Using Eq. (3.23) and the definition of t, then:
( ) ( )
dCout dt Cout t Cout t
− = → ln =− → = exp −
Cout t Cin (0) t Cin (0) t
For calculating the E(t) function, we must calculate the integral of this output
signal; therefore:
∞ ∞ ( ) [ ( )]∞
Cout t t
dt = exp − dt = t ⋅ exp − =t
∫0 Cin (0) ∫0 t t 0
In this way, we can calculate:
Cout ( )
Cin (0) 1 t
E(t) = ∞ Cout
= exp − (3.25)
∫0 C (0)
dt t t
in

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

3.4.2 Study of the RTD in the PFR


Let us now discuss the mass balance of a tracer in different situations in the plug
flow reactor (PFR). The mass balance in a differential volume of a PFR in non-
steady state is (see Chapter 5):

Input + Generation = Output + Accumulation


( )
dC
Q ⋅ C + 0 = Q ⋅ (C + dC) + ⋅ dV
dt
( )
dC
−Q ⋅ dC = ⋅ dV (3.33)
dt
Using partial derivatives:
( ) ( )
𝛿C 𝛿C
Q⋅ =− (3.34)
𝛿V 𝛿t
If the input has the form of a step, in the time–space coordinates, the corre-
sponding initial values are
}
C(0, t) = u(t)
(3.35)
C(V , 0) = 0
66 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

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)

3.5 Complex Network of Ideal Reactors


Combinations of ideal reactor elements connected in series and parallel are often
used to model the results of tracer tests performed on real reactors. Such systems
are easily analyzed with the Laplace transform (L.T.) technique. Also, the L.T.
technique provides a very convenient means of obtaining the moments of RTD
functions (Chapter 1) as well as aids considerably in obtaining the solution of
partial differential equations (Chapter 4).
We already know how the RTDs and how the transfer functions of the two
ideal reactors (PFR and CSTR) are. Using these RTDs, we can derive the func-
tions corresponding to a compartmentalized general model with delays in time.
This means that given a real flow system or reactor that can be represented by
a set of CSTRs and PFRs in series or in parallel, we will be able to deduce the
curve corresponding to the combination. We will use for this purpose the Laplace
transforms.
In this sense, any network of CSTRs and PFRs consists of the following:
• Elements in parallel
• Series elements
• Mix points
• Points of separation (split)
3.5 Complex Network of Ideal Reactors 67

Before presenting the general mathematical approach, let us see a simple


example that will help us understand more complex systems.

Example 3.1 Combination of Two Reactors in Series


During the analysis of the behavior of a reactor, the curve of the response to a
tracer is determined. In view of the results, it is verified that the reactor behaves
like a PFR connected with a CSTR in series, each one with a different volume,
as shown in the following figure. Determine the transfer function of the entire
system.
V1 = αV V2 = (1– α)V
X(s) Q Q
RFP CSTR
Q Y(s) Z(s)

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.

3.5.1 Systems in Series


For a system of two subsystems in series, such as the one shown in Figure 3.2, we
can write
Y1 (s) = H1 (s) ⋅ X1 (s)
Y2 (s) = H2 (s) ⋅ Y1 (s) (3.40)
and then:
Y2 (s) = H1 (s) ⋅ H2 (s) ⋅ X1 (s) (3.41)
68 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

System System
X1(s) H1(s) Y1(s) H2(s) Y2(s)

Figure 3.2 Series of coupled systems and signals.

This would be equivalent to the convolution operation discussed in Chapter 2.


Nevertheless, as we have seen in the previous example, each transfer function
should be evaluated with the corresponding change in volume (or residence
time). In this way, it is more correct to write
Y2 (s) = H1 (𝛼1 s) ⋅ H2 (𝛼2 s) ⋅ X1 (s) (3.42)
where 𝛼 1 is the ratio of volume in reactor 1, and 𝛼 2 corresponding to reactor 2. If
we look for the transfer function of the whole system, i.e. Y 2 (s)/X 1 (s), we get
H(s) = E(s) = Y2 (s)∕X1 (s) = H1 (𝛼1 s) ⋅ H2 (𝛼2 s) (3.43)
That corresponds to the Laplace transform of the residence time distribution
of the whole system.
In general, for a system of N-subsystems in series:

N
E(s) = Ej (𝛼j s) (3.44)
j=1

where Ej (s) is the L.T. of a pulse introduced in the individual system having a
volume V .

Example 3.2 N-CSTRs in Series


If we want to find the E-curve of N equal-sized CSTRs in series, like the system
shown in Figure 3.3, we will use the RTD of a single stirred tank of volume V and
flow rate Q:
( )
1 t
E(t) = exp − (3.45)
t t
So that the transfer function is
1
E(s) =
1 + ts
1
where t = VQ . For N equal-sized tanks in series, each tank contains 𝛼j = N
fraction
of the volume. Hence:
( )
s 1
Ej (𝛼 s) = Ej = (3.46)
N 1+ t s N

The overall transfer function is obtained by



N
1
E(s) = Ej (𝛼j s) = ( )N (3.47)
j=1 t
1+ N
s
3.5 Complex Network of Ideal Reactors 69

Feed

.....

CA1, T1

CAn, Tn

Product

CA2, T2

Figure 3.3 System of CSTRs connected in series.

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.

3.5.2 Systems in Parallel


Let us consider a system with two branches in parallel, like the one shown in
Figure 3.4.

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

Example 3.3 System Modeled Through Two CSTRs in Parallel


For a system of two reactors, like the one shown in Figure 3.5:
𝛼
t 1 (s) = t
𝛽
1−𝛼
t2 (s) = t (3.56)
1−𝛽
3.5 Complex Network of Ideal Reactors 71

βQ
S αV M
Q Q
(1– β)Q

(1–α)V

Figure 3.5 Two continuous stirred tanks in parallel.

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

3.5.3 Systems with Recycle


Now that we have learned the derivation of the transfer function for subsystems in
parallel and in series, let us face systems with recirculation. The only new “rule” is
the division rule: if a current is divided, each of the outgoing currents has the same
transfer function. Consider a general recirculation system like the one shown in
Figure 3.6.
Flow rate Q flows through a recycle system (the system within the dashed box
is the system with recycle) of total volume V . Internally, at point M flow rate, Q

Figure 3.6 Combination of systems


showing internal or external recycle αV
M Y S
current. Q Q
G1
U (1 + R)Q E

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.

Example 3.4 Recirculation in a Packed Bed Reactor


A reactor packed with solid catalyst is used to crack ethylene to its monomers at
an elevated but constant temperature. A pulse input of C13 -marked tracer is fed,
and the signal schematized in the figure is obtained. The flow rate was 1 l/min
and the reactor volume (not occupied by the solid) is 5 l. Find the following: (a)
the mean residence time of the gas, (b) a model valid for the reactor, and its
E(t) function.
3.5 Complex Network of Ideal Reactors 73

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

We can write, in the recirculation point:


Q ⋅ C0 + RQ ⋅ C2 = (1 + R)Q ⋅ C1
and then:
C0 + RC2
C1 =
1+R
In the reactor itself, we have
C C2 (1 + R) ⋅ C2
G1 = 2 = =
C1 C0 + RC2 C0 + RC2
1+R
74 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

“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).

Example 3.5 Nonideal CSTR


A second-order irreversible reaction (A → Products) is being produced in a tank
reactor of volume V = 25 m3 . The flow rate entering the reactor is Q = 1 m3 /min,
and the system was designed to operate as a CSTR with a high conversion, but
now it is poorly operating at X A = 0.25.
3.5 Complex Network of Ideal Reactors 75

A pulse input of M = 250 g of tracer is used to characterize the reactor. At the


exit, the concentration of tracer is measured, giving the following data:

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

And the impulse response is


𝛽2 𝛽t
E(t) = (1 − 𝛽) ⋅ 𝛿(t) + e− 𝛼t
𝛼t
V
where 𝛿(t) is the Dirac delta function and t = Q
If we have in mind that C(t) = M⋅E(t)
Q
, then:
[ ]
M 𝛽 2 𝛽t
C(t) = (1 − 𝛽)𝛿(t) + e− 𝛼t
Q 𝛼t
76 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

We can plot in a semilog graph:


( )
M𝛽 2 𝛽
ln(C(t)) = ln − t
Q𝛼t 𝛼t
The slope of the graph with the data given provides a value of −0.1177 and the
extrapolation to t = 0 gives a value of 2.1549, this is, exp(2.1549) = 8.62 mg/l.
Consequently, we can calculate the values of 𝛽 = 0.3 and 𝛼 = 0.1. In this way,
the active volume is approx. 0.3 × 25 = 7.5 m3 .
In a CSTR for a second-order reaction, we have
V CA0 XA
t= = 2
Q kCA0 (1 − XA )2
where V and Q are the active volume and flow rate, respectively. It is impor-
tant to note that X A represents the conversion of the reactor, which does not
equal the conversion of the isolated CSTR as there exists a bypass. In our
example, we can write
𝛼V CA0 XAv
=
𝛽Q 2
kCA0 (1 − XAv )2
where XAv represents the conversion just outside the reactor vessel. If a bal-
ance over the mixing point M is done:
(1 − 𝛽)nA0 + 𝛽nA0 (1 − XAv ) = nA0 (1 − XA )
In our case, the system gives a conversion X A = 0.25. Taking these consider-
ations, one can see that:
X
XAv = A = 0.83
𝛽
And from all this, it is possible to calculate kCA0 , resulting in a value of
3.44 min−1 .
1000 ⋅ 0.25
(b) Vnew = = 128 l
3.44 ⋅ (1 − 0.25)2
1000 ⋅ 0.83
(c) Vnew = = 8333 l
3.44 ⋅ (1 − 0.83)2

Example 3.6 Deep Reactor


Co ,Qo

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)

t1 = (1 − 𝛽 − 𝛾) t (the total flow Q0 is flowing through this vessel)


The transfer functions are
G1 = exp(−t1 s)

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

Example 3.7 Square Pulse of Tracer


A square pulse of tracer of magnitude C in /a and of duration “a” units of time is
charged to a CSTR, starting at t = 0. We are interested in simulating the response
of that reactor vs. time for a = 0, 1, and 2.
78 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

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

Example 3.8 Combining Mole Balance and Laplace Transform in a Reactor


A test with a tracer enters a CSTR following the function:
Cin = 0.06 ⋅ t when 0 ≤ t ≤ 5
Cin = 0.06 ⋅ (10 − t) when 5 < t ≤ 10
Cin = 0 when t > 10

Calculate the exit concentration as a function of time.

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

In the interval t > 10:


t ⋅ s ⋅ C(s) + C(s) = 0
and
( )
t − 10
Cout (t) = 0.122 ⋅ exp −
t

Example 3.9 System of Two Plug Flow Reactors


In a system of two reactors, designed as PFRs, certain deficiencies in its behavior
are observed, so that it could be modeled with the following scheme:

Q0 (1 – p)·Q0 Reactor 1: C1 C3 Reactor 2: C4 C5


CSTR PFR
V1 = a·V V2 = b·V
C0 p·(1 – γ)·Q0

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

In the same way, at the final mixing point:


(1 − p𝛾)C4 + p𝛾C2 = C5
We can say that the global transfer function is C 5 /C 0 ; therefore:
(1 − p)G1 + p(1 − 𝛾)G3
Gglobal = p𝛾G3 + (1 − p𝛾)G2
(1 − p𝛾)
= p𝛾G3 + (1 − p)G2 G1 + p(1 − 𝛾)G2 G3
As we know:
1 1
G1 = · · · G2 = exp(−t2 s) · · · G3 =
1 + t1 s 1 + t3 s
So, finally:
1 1 1
Gglobal = p𝛾 + (1 − p) exp(−t2 s) + p(1 − 𝛾) exp(−t2 s)
1 + t3 s 1 + t1 s 1 + t3 s

3.6 Transfer Function for the Dispersion Model


It might be instructive to show how to obtain the solution to the ODE of the
dispersion model (Eq. (1.52)) and the effect of the boundary conditions on that
solution.
For the dispersion model, we have
𝜕 2 CT 𝜕CT 𝜕CT
De −u⋅ = (1.52)
𝜕z2 𝜕z 𝜕t
By taking Laplace transform respect to time:
𝜕 2 C(s) 𝜕C(s)
De ⋅ −u = s ⋅ C(s) (3.62)
𝜕z2 𝜕z
And then:
𝜕 2 C(s) u 𝜕C(s) s
− − ⋅ C(s) = 0 (3.63)
𝜕z 2 De 𝜕z De
This is a homogeneous linear differential equation with constant coefficients, of
the type:
A ⋅ y′′ + B ⋅ y′ + C ⋅ y = 0 (3.64)
whose characteristic equation is
A ⋅ r2 + B ⋅ r + C = 0 (3.65)
In our case:
( ) ( )
u s
r2 + − ⋅r+ − =0 (3.66)
De De
In that equation, (B2 − 4⋅A⋅C) > 0, this is
( )2 ( ) ( )2 ( )
u s u s
− −4⋅1⋅ − = +4⋅ >0 (3.67)
De De De De
82 3 Use of Transfer Function for Convolution and Deconvolution of Complex Reactor Systems

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.

Example 3.10 Laplace Transform Solution to Example 2.8


In Chapter 2 we have treated in Example 2.8 the case of an experimental flow reac-
tor, initially designed as a CSTR, fed with small particles of solids. The response
to an impulse signal (solid impregnated in a colored compound) has the function:
C(t) = a(10 − t) (with time in minutes). This reactor is followed by a fluidized bed
reactor of mean residence time (of the solid) equal to three minutes; determine
the concentration of tracer at the output of the system.
Let us solve the problem using Laplace transforms and transference functions.
Solution
At the exit of the second reactor, we have
Cout = E1 ⊗ E2
In the Laplace space:
{ ( )}
1 t
Cout (s) = E1 (s) ⋅ E2 (s) = L{0.02 ⋅ (10 − t)} ⋅ L ⋅ exp −
t2 t2
Taking the transforms:
( )
0.2 0.02 1 0.2 0.02
Cout (s) = − 2 ⋅ = −
s s 1 + t2 ⋅ s s ⋅ (1 + t2 ⋅ s) s (1 + t2 ⋅ s)
2

0.2 0.02
= −
(s + t2 ⋅ s ) (s + t2 ⋅ s3 )
2 2
3.6 Transfer Function for the Dispersion Model 83

Doing the anti-Laplace transform:


( ( )) [ ( ) ]
t t
Cout (t) = 0.2 ⋅ 1 − exp − − 0.02 ⋅ t2 ⋅ exp − − (t2 − t)
t2 t2
The following figure presents the data calculated for time between 0 and 12
minutes using the equation, which is completely equal to the result presented in
Example 2.8.
0.350

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)

Example 3.11 Complex System of Three Reactors


The behavior of a mixed flow reactor can be considered equivalent to three ele-
ments in parallel, with a fraction of flow for each one. The response of each branch
of the system to a pulse input is given by (time in minutes):
C1 = exp(−5t)
C2 = t ⋅ exp(−t)
C3 = t 2 ⋅ exp(−3t)

(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

and, doing the anti-transform, we get


E(t) = 2 ⋅ (31𝛿(t) + 10𝛿 ′ (t) + 16𝛿 (t) + 6𝛿 (t) + 𝛿 iv (t) − 128e−5t )
′′ ′′′
Bibliography 85

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

Partial Differential Equations in Reactor Design

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.

4.2 Classification of Partial Differential Equations


Consider the equation:
𝜕2 f 𝜕f 𝜕2f
A +B +C 2 =0 (4.1)
𝜕x 2 𝜕x𝜕y 𝜕y
Partial differential equations are classified according to the value of the dis-
criminant D defined as
D = B2 − 4AC
where
⎧D < 0 elliptical

⎨D = 0 parabolic (4.2)
⎪D > 0 hiperbolic

Chemical Reactor Design: Mathematical Modeling and Applications,
First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
88 4 Partial Differential Equations in Reactor Design

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

4.3 Approximations by Finite Differences


The resolution of a differential equation in partial derivatives by finite differences
tries to discretize the equations. In our case, we must obtain a difference approxi-
mation of the first and second derivatives using only values of the function we are
studying. Let us see how we can obtain a first-order and a second-order approx-
imation using a Taylor series development.

4.3.1 First-order Approximation


Developing the function f (x) to first order around the point xi :
( )
df (xi )
f (xi+1 ) = f (xi ) + (xi+1 − xi ) + … (4.3)
dx
where (xi + 1 − xi ) = Δx
Simply by isolating the derivative, we obtain
( )
f (xi+1 ) − f (xi )
f ′ (xi ) = (4.4)
Δx
Equation (4.4) is a first-order forward approximation. Likewise, we could have
obtained a backward approximation by simply using approximations in the points
xi and xi−1 .
To obtain the approximation of the second derivative, we use the previously
obtained result:
f (xi+2 ) − f (xi+1 ) f (xi+1 ) − f (xi )
( ′ (xi+1 ) ′ ) −
′′ f − f (xi ) Δx Δx
f (xi ) = =
Δx Δx
f (xi+2 ) − 2f (xi+1 ) + f (xi )
= (4.5)
Δx2
4.3 Approximations by Finite Differences 89

4.3.2 Approximation of Second Order


To obtain a second-order approximation of the first and second derivatives, let
us now consider the Taylor series approximation of the function f (x):
( ) ( 2 )
𝛿f (xi ) 1 𝛿 f (xi )
f (xi+1 ) = f (xi ) + Δx + Δx2 + · · ·
𝛿x 2! 𝛿x2
( ) ( 2 )
𝛿f (xi ) 1 𝛿 f (xi )
f (xi+2 ) = f (xi ) + 2Δx + (2Δx)2 + · · · (4.6)
𝛿x 2! 𝛿x2
Multiplying by “a” and “b,” respectively, the two previous equations and adding
we get
( )
𝛿f (xi )
a ⋅ f (xi+1 ) + b ⋅ f (xi+2 ) = (a + b)f (xi ) + Δx (a + 2b)
𝛿x
( 2 )
1 𝛿 f (xi )
+ Δx2 (a + 4b) + · · · (4.7)
2! 𝛿x2
These equations must be met for any value of “a” and “b.” Therefore, if we want
to obtain the value of the derivative, we can do that by
a + 2b = 1
Also, since we are interested in obtaining an expression for the derivative where
only function values appear, we can do
a + 4b = 0
Solving the above-given system of equations results in a = 2, b = 0.5.
Thus:
( )
1 3 𝛿f (xi )
2f (xi+1 ) − f (xi+2 ) = f (xi ) + Δx
2 2 𝛿x
and, isolating the derivative:
( )

4f (xi+1 ) − 3f (xi ) − f (xi+2 )
f (xi ) = (4.8)
2Δx
To obtain an approximation of the second derivative, we can follow the same
procedure:
( ) ( 2 )
𝛿f (xi ) 1 𝛿 f (xi )
f (xi+1 ) = f (xi ) + Δx + Δx2
𝛿x 2! 𝛿x2
( 3 )
1 𝛿 f (xi )
+ Δx3 + · · ·
3! 𝛿x3
( ) ( 2 )
𝛿f (xi ) 1 𝛿 f (xi )
f (xi+2 ) = f (xi ) + (2Δx) + (2Δx)2
𝛿x 2! 𝛿x2
( 3 )
1 𝛿 f (xi )
+ (2Δx)3 + · · ·
3! 𝛿x3
90 4 Partial Differential Equations in Reactor Design

(
) ( 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.

Table 4.1 First-order approximations of the


first and second derivative.

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

′′ f (xi ) − 2f (xi−1 ) + f (xi−2 )


f (xi ) =
Δx2
Central differences: They do not exist.
4.4 Approaching the Problem Using Finite Differences 91

Table 4.2 Second-order approximation of the first


and second derivative.

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

4.4 Approaching the Problem Using Finite Differences


To facilitate the interpretation of the methodology, we propose the resolution
scheme by means of an example, which we use throughout this chapter. The
results obtained can be generalized directly to other types of problems.
We consider the diffusion of a species toward the center of a catalyst particle,
which is completely explained in Chapter 9. The problem calculates the concen-
tration distribution of the species along the particle. In the example, the substance
“A” is contained in a gas and it diffuses and reacts inside of a solid catalyst. Before
the contact between the phases (t < 0), the concentration of the species in the gas
phase is known (C As ) and it is zero inside the solid. As time goes on, “A” is trans-
ferred to the inner part solid by a combination of diffusion and reaction of “A”
in the catalyst: A (gas phase) → products (catalyzed reaction). Figure 4.1 shows a
scheme of this situation.
If a mass balance for species “A” is done in a dV = S ⋅ dx, “S” being the surface
of the interface:
Input + Generation = Output + Accumulation

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

At this point, the


derivative is zero

Figure 4.1 Diffusion of a gas inside a plane catalyst particle.

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

4.4.1 Explicit Method


The explicit method, for the parabolic equations, combines a forward difference
in time with a normally centered approximation of order 2 of the spatial
derivatives always using values of the dependent variable at the time instant
before the one considered, which produces an explicit formula. It is possible to
use other differences schemes; we only have to take into account the stability
conditions that are discussed later.
Let us consider an interior point shown in Figure 4.2. Since all the values at the
time “t” are assumed to be known, the spatial derivatives can be approximated
using central differences:
t t
𝜕CA CAi+1 − CAi−1
= (4.17)
𝜕x 2 ⋅ Δx
𝜕CA2 t
CAi+1 − 2 ⋅ CAi
t t
+ CAi−1
= (4.18)
𝜕x2 Δx2
However, the first-order time derivative can be approximated by the forward
difference:
𝜕CA C t+1 − CAi
t
= Ai (4.19)
𝜕t Δt
The resolution procedure consists of substituting these approximations in the
t+1
differential equation that governs the problem and resolving explicitly for CAi .
t
This equation will contain on the right side only terms of CA . The boundary con-
ditions will be used to calculate the limit points and the problem can be solved
forward in time. As considerations, we must point out that the error in the reso-
lution of this problem is proportional to (Δt + Δx2 ).
Applying the procedure to the previous example and using the indices “i” for
the position and “t” for the time:
t+1
CAi t
− CAi t
CAi+1 − 2 ⋅ CAi
t t
+ CAi−1
= DAB − k ⋅ CAi
t
(4.20)
Δt Δx2

Calculation scheme by the explicit method

Temporal
location ‘t’ Known initial condition
Known boundary condition

ti + 1
Δt
ti
ti – 1

xi – 1 xi xi + 1
Δx
Spatial location ‘i”

Figure 4.2 Resolution scheme for the explicit method.


94 4 Partial Differential Equations in Reactor Design

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.

4.4.2 Initial and Boundary Conditions


As can be seen, the problems of differential equations in partial derivatives
require initial and limit conditions. The initial condition provides the value of
the dependent variable over the entire spatial domain for a fixed time (although
conditions at initial time are usually considered, the same methodology can be
used for conditions at any other time).
For the case mentioned earlier, a valid initial condition is

t = 0 → CA = 0, for all points but the one contacting the other phase,
where CA = CAs

In chemical engineering, boundary conditions are usually particularized in


three types: constant value, insulation, and contours exposed to a transport
boundary layer. In the case of the diffusion presented, a good assumption can be
the following:

At the interface (x = 0)∶ CA = CAs (known and constant value), ∀t


𝜕C
At the center of the particle (x = L), by simmetry: A = 0
𝜕x
In this sense, the first condition is of the constant value limit type. This type
of condition is normally used as an approximation to the real situation, in which
there are fluctuations on the constant value, but they are of small magnitude and,
for practical purposes, the fluctuations can be neglected without introducing sig-
nificant errors in the problem.
The boundary condition at the center of the catalyst, x = L, is similar to other
conditions used in chemical engineering and reflects the fact that, in the present
example, the derivative in the center is zero because it represents a minimum in
the concentration of the species reacting.
𝜕C
The boundary condition 𝜕xA = 0 is like those used for “insulating surfaces,” rep-
resenting surfaces that do not allow the transport of a certain magnitude. For
example, thermal insulators prevent the passage of heat, while solid surfaces act
4.4 Approaching the Problem Using Finite Differences 95

as insulators when passing matter. The common feature of a perfect insulator is


that it does not allow passage in a direction perpendicular to it.
Other types of boundary conditions are also present in chemical technology,
such as the transport boundary layer condition, which is usually modeled using
an individual transport coefficient.
For the problem of the example, the initial conditions can be written as
t = 0; CA = 0; for ∀x ≠ 0

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.

4.4.3.1 Resolution of the Selected Problem and Programming


®
For the resolution of the selected examples, we use Matlab programming. Mat-
lab program for this example is the following one.

Program Listing 4.1


% DIFFERENTIAL EQUATION IN PARTIAL DERIVATIVES
% METHOD OF FINITE DIFFERENCES "EXPLICIT"
% NON-STATIONARY CONCENTRATION PROFILE OF A SUBSTANCE ‘A’
% DIFFUSING TO A CATALYST
DAB = 0.01;% cm2 / s
L = 0.3; % cm
k = 1; % l / mol ⋅ s
CAs = 1; % mol / l
N = 50; % Number of values of x to integrate in x
incx = L / (N-1);
inct = 0.9 / (k + 2 * DAB / incx ̂ 2);
C (1) = CAs;
C (2: N) = 0;
x = 0: incx: L;
A = DAB * inct / incx ̂ 2;
plot (x, C)
t=0;

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.

4.5 Other Applications of PDE – Numerical Methods


4.5.1 The RTD of a Complex System
Let us apply the finite difference method to the simulation of the flow in a
complex system. In this example, we have an electrochemical reactor having a
three-dimensional electrode made of carbonaceous matrix inserted between two
4.5 Other Applications of PDE – Numerical Methods 99

Figure 4.4 Scheme of the


model proposed for an
electrochemical reactor with Flow rate = Q
Dynamic zone
a three-dimensional
electrode. Stagnant zone

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

the length coordinates:


𝜕Cd 𝜕 2 Cd 𝜕Cd
𝛽dyn ⋅ S ⋅ = 𝛽dyn ⋅ S ⋅ Dd ⋅ − 𝛽dyn ⋅ S ⋅ ud ⋅
𝜕t 𝜕z 2 𝜕z
− 𝛽tot ⋅ S ⋅ 𝛾 ⋅ (Cd − Cs ) (4.31)
𝜕Cs
𝛽stat ⋅ S ⋅ = −𝛽tot ⋅ S ⋅ 𝛾 ⋅ (Cs − Cd ) (4.32)
𝜕t
where S = cross-sectional area of the reactor (m2 ); Dd = dispersion coefficient
in dynamic phase (m2 /s); ud = velocity in the dynamic phase (m/s) of volume
V = Q/S𝛽 dyn ; and Q = electrolyte flow through volume V (m3 /s).
For the calculation of the RTD due to this pathway, we will solve the system of
Eqs. (4.31) and (4.32) with the appropriate boundary conditions.
As in other examples, we will use the finite differences method for solving the
system of equations. If we develop Eq. (4.32) in terms of this method, we find that:
t+1 t
Cs,i − Cs,i
𝛽stat ⋅ S ⋅ = −𝛽tot ⋅ S ⋅ 𝛾 ⋅ (Cs,i
t t
− Cd,i ) (4.33)
Δt
where the subscript “i” refers to a position in the length dimension, and the super-
script “t” to the positions in time. Rearranging:
𝛽
t+1
Cs,i t
= Cs,i − tot ⋅ Δt ⋅ 𝛾 ⋅ (Cs,i
t t
− Cd,i )
𝛽stat
( )
𝛽 𝛽
= 1 − tot ⋅ Δt ⋅ 𝛾 Cs,i t
+ tot ⋅ Δt ⋅ 𝛾 ⋅ Cd,i
t
(4.34)
𝛽stat 𝛽stat

Using the following groups of constants:


𝛽tot
A=1− ⋅ Δt ⋅ 𝛾
𝛽stat
𝛽tot
B= ⋅ Δt ⋅ 𝛾
𝛽stat
We can rewrite the equation by
t+1
Cs,i = A ⋅ Cs,i
t
+ B ⋅ Cd,i
t
(4.35)
For Eq. (4.31), following the same procedure, it is found that:
t+1 t t t t
Cd,i − Cd,i Cd,i−1 − 2Cd,i + Cd,i+1
𝛽dyn ⋅ = 𝛽dyn ⋅ Dd ⋅
Δt Δz2
t t
Cd,i+1 − Cd,i
− 𝛽dyn ⋅ ud ⋅ − 𝛽tot ⋅ 𝛾 ⋅ (Cd,i
t t
− Cs,i )
Δz
t t t
Cd,i−1 − 2Cd,i + Cd,i+1
t+1
Cd,i = Δt ⋅ Dd ⋅
Δz2
Δt ⋅ ud
− ⋅ (Cd,i+1
t t
− Cd,i )
Δz
𝛽 ⋅ Δt ⋅ 𝛾
− tot ⋅ (Cd,i
t
− Cs,it t
) + Cd,i
𝛽dyn
4.5 Other Applications of PDE – Numerical Methods 101

( )
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.

4.5.1.1 Boundary Conditions for Partial Differential Equations


For a “closed” reactor (i.e. when a portion of fluid that enters the reactor cannot
leave it by the same way), the consideration of flux balances at the entrance and at
102 4 Partial Differential Equations in Reactor Design

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).

Program Listing 4.2


%PDE DIFFERENTIAL EQUATIONS
%ELECTROCHEMICAL REACTOR RESIDENCE TIME DISTRIBUTION

%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

N= 150; % Number of length intervals

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

At =min(At1, At2) % Choosing the minor value of At1 and At2

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.

4.5.2 Concentration Profile in a Reactor in Which There Is Flow


and Dispersion
PDE numerical methods explained before can be applied to many engineering
situations. In this example, a stream of water is circulating along a tube. At a
given moment, we begin to introduce a certain substance at one end, so that,
keeping the flow constant, the concentration varies with the dimensionless time,
according to the following expression:
( )
(t − 0.4)2
C0 = exp −
0.005
where t refers to units of dimensionless time and C is the concentration that
enters the pipe at a given time. The variation of the concentration as a func-
tion of time along the axial direction within the tube can be calculated by the

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.

Program Listing 4.3


% DIFFERENTIAL EQUATION IN PARTIAL DERIVATIVES
% METHOD OF FINITE DIFFERENCES "EXPLICIT"
% "FLOW + DISPERSION IN A PIPE"
M = 0.01;
L = 1;
N = 200; % Number of values of x to make the integration in x
106 4 Partial Differential Equations in Reactor Design

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

4.5.3 Reaction on a Catalytic Flat Wall


In a catalytic reactor, a fluid of concentration “C” and linear velocity “u” is flowing
and reacting between two wide plates, with rate equation −rb = kb ⋅ Cb𝛼 (“b” refers
to the bulk fluid). The walls are coated with a catalyst on which the reaction rate
is first order −rw = k w C w (“w” refers to the wall). Diffusion in the axial direction is
small in comparison with the mass transfer by bulk flow, but it has been observed
as a certain diffusion in the direction from wall to wall. Make a material balance
over an elemental prism of unit width between y and y + dy laterally and z + dz
axially, in order to determine the concentration of the reacting species over the
complete system. Apply the finite differences method and determine the stability
conditions, as well as the boundary and initial conditions of the system.
Apply the solution to the following case: 𝛼 = 1; k b = 1⋅10−4 mol/(l s);
C A0 = 1 mol/l; k w = 1⋅10−6 l/(mol s); 2w = 0.02 m; u = 0.1 m/s; D = 10−3 m2 /s.
Length of the (micro)reactor 0.01 m.
A scheme of the system considered is given in Figure 4.6. Assume that the sep-
aration between the plates is “2w” and then y = 0 at the center and y = w in both
catalytic plates.
Considering the elemental prism, the volume of that prism is dV = 1⋅dy⋅dz. The
current is flowing in the “z” direction, and then the area of the prism in the axial
direction is S = 1⋅dy. Nevertheless, the lateral area of the prism, through which
has been observed diffusion, is S′ = 1⋅dz = dz.
With this premise, we can write

Input = (Convection + diffusion in the′ y′ direction)


𝜕Cb
= u ⋅ Cb ⋅ S − D ⋅ S′ ⋅
𝜕y
4.5 Other Applications of PDE – Numerical Methods 107

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

Figure 4.6 Catalytic flat wall.

𝜕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

We can apply the finite differences method in this way:


n+1 n
𝜕Cb Cb,m − Cb,m
=
𝜕z Δz
108 4 Partial Differential Equations in Reactor Design

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

as the position z = 0 coincides with the wall. Then:


n
Cb,wall−1
n
Cb,wall = kw Δy
1+ D

We can use the later equations to program in Matlab the case required. The
following program can be used.

Program Listing 4.4


clear all
close all

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

N=50; % Number of increment in the y-axis

incy=w/((N-1)); % Increment of the variable y - half of the total


incz=0.95*incy ̂ 2/(2*D/u+kb); % Must be fulfilled
incz<incy ̂ 2/(2*D/u+kb)

y=0:incy:w; % y-values to be used in the calculation


CAv=[]; % Initialization of a matrix that will contain all CA calcu-
lated
% Initial condition (entering the reactor):
z=0;
CA=ones(1,N)*CA0;

zfinal=0.01; % Length of the reactor


gamma=incz*D/(u*incy ̂ 2);
A=gamma;
B=-2*gamma-incz*kb/incy ̂ 2+1;

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

Unsteady State Regime Simulation in Reactor Design

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.

5.2 CSTR Working in Unsteady State


Let us first discuss the system shown in Figure 5.1. During an unsteady state, we
will have
Input = nA0 (mol/time)
Output = nA (mol/time)
Accumulation = V ⋅dC A /dt (volume⋅(mol/volume)/time)
Generation = rA ⋅V ((mol/volume⋅time)⋅volume)
So that, in the mass balance:
Input + Generation = Output + Accumulation
dCA
nA0 + rA ⋅ V = nA + ⋅V (5.1)
dt
If we assume a first-order reaction:
dCA
Q ⋅ CA0 − k ⋅ CA ⋅ V = Q ⋅ CA + ⋅V
dt
dCA C − CA C − CA
= A0 − k ⋅ CA = A0 − k ⋅ CA (5.2)
dt V ∕Q t
This equation is an ordinary differential equation, not in partial derivatives, and
can be solved numerically with no problem using the finite differences method:
CAt+1 − CAt CA0 − CAt
= − k ⋅ CAt (5.3)
Δt t
Chemical Reactor Design: Mathematical Modeling and Applications,
First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
112 5 Unsteady State Regime Simulation in Reactor Design

Feed Figure 5.1 Continuous stirred tank reactor (CSTR).

CA, T

Product

and solving for the concentration:


[ ]
CA0 − CAt
t+1
CA = Δt − k ⋅ CA + CAt
t
(5.4)
t

®
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.

Program Listing 5.1


clear all
tresid=10; % sec, average residence time
k=1; % sec-1, kinetic first-order constant
CA0=1; % Input concentration, mol/m3
t=0;
i=1;
CA(1)=CA0; % Boundary condition at the input
tfinal=8; % sec, final time
Nt=100; % Number of intervals of time
inct=tfinal/Nt; % Sec, increment of time
tv=0:inct:tfinal; % Values of time to be used in the calculation
%CSTR concentrations are the same in all points of the reactor
while t<tfinal
i=i+1
t=t+inct;
CAn(i)=inct*((CA0-CA(i-1))/tresid-k*CA(i-1))+CA(i-1);
End
CA=CAn;
plot(tv,CA(1:Nt+1))
xlabel('Time (sec)')
ylabel('C_A (mol/m ̂ 3)')

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.

5.3 PFR Working in Dynamic Regime (No Dispersion)


As in the PFR, different reaction rates are present, and the mass balance should
be done over a differential volume, dV , as shown in Figure 5.3.
As different variables are used, differential forms are partial differential, and
introducing into the balance:
Input + Generation = Output + Accumulation
𝜕CA
nA + rA ⋅ 𝜕V = (nA + 𝜕nA ) +⋅ 𝜕V (5.5)
𝜕t
Bearing in mind, for a constant-density system:
Q ⋅ 𝜕CA = 𝜕nA (5.6)
which means that:
𝜕CA 𝜕C
= rA − Q A (5.7)
𝜕t 𝜕V
If we assume a first-order kinetics:
𝜕CA 𝜕C
= −kC A − Q A (5.8)
𝜕t 𝜕V
Figure 5.3 The mole balance in dV
a PFR.
CA0 CA
k
nA0 A products nA

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

Using the finite differences approximation, first order:


t+1 t t t
CAi − CAi t
CAi+1 − CAi
= −kCAi −Q (5.9)
Δt ΔV
Rearranging:
( )
Δt Δt t
t+1
CAi = 1 − k ⋅ Δt + Q t
CAi −Q C (5.10)
ΔV ΔV Ai+1
Applying the stability criteria (already mentioned in Chapter 4), the system is
stable (errors do not increase), if:
Δt
Q >0
ΔV
( )
Δt
1 − k ⋅ Δt − Q >0
ΔV
( )
Δt Δt
1 − k ⋅ Δt − Q +Q ≤1
ΔV ΔV
First and third conditions are always fulfilled, and from the second one:
1
Δt < ( ) (5.11)
Q
k + ΔV

In the initial state:


∀V if t = 0 → CA = 0 (5.12)
The boundary conditions of this system are
• At the input (V = 0): C A = C A0 (known concentration)
• At the exit (V = V total ) → C A (V − )=C A (V + ) (closed vessel), so that at this
section:
[ ]
𝜕C
=0 (5.13)
𝜕V Vtotal
The procedure can be applied to different designs. We will apply the method
for the simulation of a 20 m3 reactor where a feed of 1 mol/m3 of reactive is fed,
with a flow rate of 100 m3 /s and k = 17 mol/(m3 s). The Matlab program in this
case is that shown here.

Program Listing 5.2


clear all

CA0=1; % Input concentration mol/m3


Qv0=100; % Volumetric flow rate m3/s
Vtotal=20; % m3 Total volume
k=17; % Reaction Kinetic constant , mol/m3⋅s
N=50; % Number of increment in x (position)

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.

5.4 PFR Working in Dynamic Regime (with Dispersion)


As pointed out in previous section, the mass balance in the PFR system in the
dynamic regime can be expressed as
𝜕CA
nA + rA ⋅ 𝜕V = (nA + 𝜕nA ) + ⋅ 𝜕V (5.14)
𝜕t
But in this case:
Q ⋅ 𝜕CA ≠ 𝜕nA (5.15)
As:
𝜕CA
nA = Q ⋅ CA − De ⋅ S ⋅ (5.16)
𝜕z
116 5 Unsteady State Regime Simulation in Reactor Design

1 1
0.9 0.8

CA at the exit (mol/m3)


0.8 0.6
0.7 0.4
CA (mol/m3)

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

CA at the exit (mol/m3)


0.8
0.03
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
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

Using the finite differences approximation, first order, and assuming a


first-order reaction:
t t t t t t+1 t
CAi+1 − 2CAi + CAi−1 CAi+1 − CAi CAi − CAi
De ⋅ −u⋅ − k ⋅ CAi
t
= (5.19)
Δz2 Δz Δt
and rearranging:
( )
Δt Δt Δt
t+1
CAi = De 2 ⋅ CAi−1
t
+ 1 − 2 2 De + u t
− kΔt CAi
Δz Δz Δz
( )
Δt Δt t
+ D − u CAi+1 (5.20)
Δz2 e Δz

The system is stable (errors do not increase) if:


Δt
De 2 > 0
Δz
( )
Δt Δt
1 − 2 2 De + u − kΔt > 0
Δz Δz
( )
Δt Δt
D −u >0
Δz2 e Δz
[ ( ) ( )]
Δt Δt Δt Δt Δt
De 2 + 1 − 2 2 De + u − kΔt + D − u ≤1
Δz Δz Δz Δz2 e Δz
As in other equations previously discussed, first and last conditions are always
fulfilled, but from the third one:
D
Δz < e (5.21)
u
and from the second one:
1
Δt < ( ) (5.22)
2 u
Δz 2
D e − Δz
+ k

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.

Program Listing 5.3


clear all

CA0=0.01; % Input concentration mol/m3


u=0.01; % Gas Velocity m/s
Ltotal=1; % m
k=0.000017; % Reaction kinetic constant, mol/m3⋅s
De=1e-4; % Difusivity in m2/s
N=150; % Number of increments in Z (position)
incZ=Ltotal/(N-1); % Z increment
if incZ>De/u
118 5 Unsteady State Regime Simulation in Reactor Design

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.

5.5 Multiple Steady States in CSTR with Exothermal


Reaction
Let us assume we are working in a CSTR with a cooling jacket as that shown in
Figure 5.6. Let us assume we are working with an exothermal first-order reaction:
A → B. Let us assume also a negligible heat loss, and constant density. We will use
the symbols included in Figure 5.6.
In the cooling jacket, we will assume that the heat removed is
qc = U ⋅ A ⋅ (T − Tc ) (5.23)
The mass balance of reacting species in the reactor, in the unsteady state, is
dC A
QCA0 − (kC A ) ⋅ V = QCA + V (5.24)
dt
The last term represents the variation of the amount of reactant with time, i.e.
the accumulation term of the mole balance. This is usually assumed to be zero,
5.5 Multiple Steady States in CSTR with Exothermal Reaction 119

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.

Figure 5.6 CSTR with cooling jacked. T Feed, T0


is the temperature of the reactor; T c is
the temperature of the cooling fluid. Q Q (flow rate)
and Qc are the flow rates of reactants
and cooling fluid, respectively.
Cooling
jacket, Tc0

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:

Q0 = 0.01 m3 /s k 0 = 7e11 s−1 T 0 = 300 K


3 −1
Qc0 = 0.01 m /s E/R = 10 500 K ΔH r = −10 000 J/mol
C A0 = 1 mol/m3 U = 40 W/m2 ⋅K cp 𝜌o = 720 J/K
V = 1 m3 A = 2 m2 cpc 𝜌oc = 1000 J/K
V c = 0.01 m3 T c0 = 300 K

Program Listing 5.4


%%% CSTR multiple states

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

% First part: steady state assumption. Looking for the solutions


of the balance equations

QR=[];
QG=[];
Tv=280:430; % Temperature range, in K

% Heat generated and removed (steady state):


for T=Tv
k=k0*exp(-E_R/T);
CA=Qv0*CA0/(Qv0+V*k);
beta=U*A/(rCpR*Qr);
Tc=(TC0+beta*T)/(1+beta);
QG=[QG -DHr*V*k*CA];
QR=[QR -(rCp*Qv0*(T-T0)-U*A*(T-Tc))];
end

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

Table 5.1 Steady-state conditions of the example, without considering


their stability.

Steady state T (K) T c (K) C A (mol/m3 )

Upper 352.9 347.02 0.1066


Intermediate 328.8 325.6 0.5137
Lower 303.8 303.37 0.9360
122 5 Unsteady State Regime Simulation in Reactor Design

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.

Program Listing 5.5


%%% CSTR multiple states
% Second part: non-steady state
% Values of the variables at the input:
CA0=1;
T0=300;
TC0=300;
T=[];
CA=[];
TC=[];
% Initial values of the variables at the reactor == steady state
(in this case, lower steady state)
CAss=0.9360;
Tss=303.8;
TCss=303.3778;
CA(1)=CAss;
T(1)=Tss;
TC(1)=TCss;
t=0;
i=1;
tfinal=200 ; % s
inct=0.1; % s
tv=0:inct:tfinal; % Values of time to be used in the calculation

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

Temperature in the reactor (K)


Temperature in the reactor (K) 350 325

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

Temperature in the reactor (K)


Temperature in the reactor (K)

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:

T0=300-5; % Perturbation of (-5) K in the


temperature at the input
The result of the simulation, starting at the stable lower steady state, is shown
in Figure 5.8d. As we can see, the system evolves to the new steady state that has
been changed as the term “Qremoved ” has changed.
Bibliography 125

Figure 5.9 Small changes in the input Qgenerated


temperature will move the Qremoved and
Qremoved
define new steady states. Qg
Qr′
3′
Qr
3
2′

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

Scaling and Stability of Chemical Reactors

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.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
128 6 Scaling and Stability of Chemical Reactors

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.

Equipment Maximum factor of scale

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

Table 6.2 Standards for scaling based on residence time.

Extrapolation Scaling Precautions/


Data from reactor to large reactor criterion commentary

Batch isotherm, Ci = f (t) Batch isotherm of t1 = t2 – Control heat transfer


greater volume in the large reactor
– Ensure same
conditions of agitation
Batch isothermal with Isothermal tubular t1 = t2 – Heat transfer
constant density, Ci = f (t) reactor of constant – Plug flow deviations
density
Batch isothermal isobaric Isothermal tubular t1 = t2 – Heat transfer
with variable density, Ci = f (t) reactor with variable – Plug flow deviations
density, ΔP ≈ 0
Isothermal PFR, Ci = f (t) Isothermal PFR t1 = t2 – Heat transfer
– Deviations in RTD
CSTR isothermal, Ci = f (t) CSTR isothermal t1 = t2 – Heat transfer
larger size – RTD deviations

that the average residence time is maintained:


t 1 = V1 ∕Q1 = V2 ∕Q2 = t 2
t constancy during scaling is natural for chemical reactors, since any alteration
will change the yields and can also change the selectivity.
In processes where heat transmission is important, it would also be desirable
for the external area of the system to follow the same scale factor, but this is rarely
possible. In fact, it is impossible if you want to maintain geometric similarity.
Table 6.2 shows the rules that we could deduce for the scaling of homogeneous
reactors, based on the different concepts studied in previous topics.
As can be seen in Table 6.2, in the scaling of isothermal reactors, it is impor-
tant to know the heat transfer phenomena. More important will be in adiabatic
reactors, which are abundant in the industry. In this type of reactors, the phenom-
ena associated with energy changes are so important that direct scaling from the
reaction time is very difficult in practice.
Thus, this chapter discusses the study of the phenomena related to the heat
transfer during the process of scaling a chemical reactor.

6.2 Scaling the Batch Tank Stirred Reactor


Batch reactors are frequently used for production of specialty chemicals such as
cosmetics, medicines, etc. The hardest problems occur when we are dealing with
exothermic reactions.

6.2.1 Temperature Control. Heat Transmission


Often, batch reactors operate as shown in Figure 6.2: at first, the temperature of
the heat transfer media (T c ) is high in order to heat the reaction mixture to the
130 6 Scaling and Stability of Chemical Reactors

Case B: Very exothermic


Temperature reaction, “runaway”

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).

6.2.2 Example of Scaling a Batch and Semi-batch Reactor


In general, the histories of temperature, concentration, and mixing conditions
in a large reactor differ significantly from those obtained in a laboratory reactor.
Therefore, it is not surprising that selectivities also change during scaling.
6.2 Scaling the Batch Tank Stirred Reactor 131

We illustrate the influence of heat transfer performance and selectivity during


scaling in a semi-batch batch reactor by means of a simulation using a reaction
in series. This reaction is composed of two irreversible elementary steps, both
exothermic and with first-order kinetics:
k1 k2
A −−→ P −−→ S
A is the reactant, P the desired product, and S an unwanted side product. k 1
and k 2 are the kinetic constants of the reaction (s−1 ). For the batch reactor, the
material and energy balances have the following form:
dCA
= r1 = −k1 ⋅ CA (6.1)
dt
dCP
= r2 − r1 = −k2 ⋅ CP − (−k1 ⋅ CA ) (6.2)
dt
dT A
𝜌 ⋅ CP ⋅ = (−r1 ) ⋅ ΔH1 + (−r2 ) ⋅ ΔH2 + U ⋅ (Tc − T) ⋅ (6.3)
dt V
Remember that in the given expression of the heat energy balance, it is assumed
that enthalpies are not dependent on the reaction temperature, and that the val-
ues of the density and capacity of reactants and products are the same. Thus,
the first term is derived from the enthalpy change of reagents due to change in
temperature, terms with ΔH 1 and ΔH 2 are associated with changing enthalpy
occurring as a result of the reaction, and the last term accounts for the heat trans-
fer through the walls.
The overall coefficient of heat transfer U is evaluated using the appro-
priate expressions. The ODE system (1–3) can be solved using numerical
methods.
The values that we use to simulate the reactor are the following:
k 10 = 5 s−1 , k 20 = 2.5⋅108 s−1 , pre-exponential factors of the constants k 1 and k 2
E1 = 20 kJ/mol, E2 = 90 kJ/mol, activation energies
ΔH 1 = −400 kJ/mol, ΔH 2 = −300 kJ/mol, reaction enthalpies
cp = 3 kJ/(kg K), specific heat
𝜌 = 1000 kg/m3 , density
C A0 = 1500 mol/m3 , C P0 = 0 mol/m3 , initial concentration of A and P
T 0 = 295 K, input temperature of the reaction mixture
Tc = 345 K if 0 < t < 360 s
= 295 K if t > 360 s, temparature of cooling medium
U = 5 kJ/(s m2 K), global coefficient of heat transfer.
We assume the reactor to be cylindrical and use two values of A/V runs corre-
sponding to a reactor 42 cm in diameter and one with 400 cm. The corresponding
quotients are
(A∕V )small = 9.50 m−1 (A∕V )large = 1.00 m−1

®
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

Small reactor Large reactor


410 1600 650 1200
390 1400 600 T
Tc 1000
370 1200 550 CP
800

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.

Program Listing 6.1


k10 = 5;% s -1
k20 = 250000000;% s -1, pre-exponential factors of the
constants k1 and k2
E1 = 20;% kJ mol -1,
E2 = 90;% kJ mol -1, activation energies
DH1 = -400;% kJ mol -1,
DH2 = -300;% kJ mol -1, reaction enthalpies
cp = 3;% kJ kg-1 K-1, specific heat
d = 1000;% kg m-3, density
CA0 = 1500;% mol m-3,
CP0 = 0;% mol m-3, initial concentration of A and P
T0 = 295;% K, input temperature of the reaction mixture
Tc = 345;% K if 0 <t <360 s
% 295 K if t > 360 s, temperature of cooling medium
U = 5;% kJ s -1 m -2 K -1, global coefficient of heat
transfer

A_V=9.5 ;% m-1 small reactor area-to-volume ratio,


should be changed to see the effect

% 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.

A/V (m−1 ) Final C p (mol/m3 ) T max (K)

Small 9.50 1409.1 394.7


Large 1.00 0.0 622.9
Large semi-batch 1.00 308.8 393.7

Program Listing 6.2


k10 = 5;% s -1
k20 = 250000000;% s -1, pre-exponential factors of the
constants k1 and k2
E1 = 20;% kJ mol -1,
E2 = 90;% kJ mol -1, activation energies
DH1 = -400;% kJ mol -1,
DH2 = -300;% kJ mol -1, reaction enthalpies
cp = 3;% kJ kg-1 K-1, specific heat
d = 1000;% kg m-3, density
CA0 = 150;% mol m-3, Current entering the system
CP0 = 0;% mol m-3, initial concentration of A and P
Q0=10/3600; %m3 s-1
T0 = 295;% K, input temperature of the reaction mixture
Tc = 345;% K if 0 <t <3600 s
% 295 K if 3660 <t <5400 s, temperature of cooling medium
U = 5;% kJ s -1 m -2 K -1, global coefficient of heat
transfer

% A_V=9.5 ;% m-1 small reactor area-to-volume ratio


A_V=1 ;% m-1 large reactor area-to-volume ratio

% 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.

6.3 Rapid Exothermic Reaction in a Tubular Reactor


Sometimes a continuous operation is desired, even for small productions. For
example, batch reactors are not suitable for very fast reactions or purposes requir-
ing short response times of the order of seconds or minutes.
We will consider the reaction presented earlier, A → P → S, but now with values
of k 1 and k 2 much higher than before. This means that the required reaction time
will be quite low. A tubular reactor operating as a plug flow reactor (PFR) will be
appropriate in this case. However, the overall heat transfer coefficient, U, must
be larger than that in the previous example due to the better transfer that occurs
as a result of the movement of the liquid stream in the tubular reactor. We will
use a value of 1 kJ/(s m2 K).
Remember that the equations of balances of energy and mass for a PFR are
very similar to those corresponding to a batch reactor, but with the dependent
variables expressed as a function of volume or distance. In the case treated before,
we have
dC
u A = r1 (6.5)
dL
dC
u P = r2 − r1 (6.6)
dL
The energy balance can be deduced by considering the transfer in a differential
volume:
(Input–Output) + (Generation by reaction)
− (Transfer of heat to exterior) = 0
Q ⋅ 𝜌 ⋅ CP ⋅ dT + (−rA ) ⋅ (−ΔHr ) ⋅ S ⋅ dL − U ⋅ (T − Tc ) ⋅ dA = 0
In this expression, S refers to the section of the reactor and “A” is the exterior
area of the tubular reactor, where the heat transfer takes place. Using the relation
Q = S⋅u and dividing by S:
dA
u ⋅ 𝜌 ⋅ CP ⋅ dT + (−rA ) ⋅ (−ΔHr ) ⋅ dL − U ⋅ (T − Tc ) ⋅ =0
S
Dividing by the dL:
dT dA
u ⋅ 𝜌 ⋅ CP ⋅ + (−rA ) ⋅ (−ΔHr ) − U ⋅ (T − Tc ) ⋅ =0
dL S ⋅ dL
dA A
If the tubular reactor has no irregularities, we can consider S⋅dL
= V
, and then,
for the system with two reactions:
dT A
u ⋅ 𝜌 ⋅ CP ⋅ = (−r1 ) ⋅ ΔH1 + (−r2 ) ⋅ ΔH2 + U ⋅ (Tc − T) ⋅ (6.7)
dL V
6.3 Rapid Exothermic Reaction in a Tubular Reactor 137

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.

Program Listing 6.3


k10 = 50;% s -1
k20 = 25000000;% s -1, pre-exponential factors of the
constants k1 and k2
E1 = 20;% kJ mol -1,
E2 = 60;% kJ mol -1, activation energies
DH1 = -300;% kJ mol -1,
DH2 = -250;% kJ mol -1, reaction enthalpies
cp = 4;% kJ kg-1 K-1, specific heat
d = 1000;% kg m-3, density
CA0 = 1000;% mol m-3,
CP0 = 0;% mol m-3, initial concentration of A and P
T0 = 295;% K, input temperature of the reaction mixture
Tc = 345;% K if 0 <t <3600 s
% 295 K if 3660 <t <5400 s, temperature of cooling medium
U = 1;% kJ s -1 m -2 K -1, global coefficient of heat
transfer
u=2.5;% m/s Linear velocity of the flow
Diameter=0.1; % m It can be changed to visualize the effect
A_V=4*3.1416/Diameter ;% m-1 area-to-volume ratio

% Conditions at the input:


CA(1)=CA0;
CP(1)=CP0;
T(1)=T0;
L(1)=0;
N=10000; % Number of length increments
Lfinal=250; % m, length of the reactor
incL=Lfinal/(N-1);

% Integration by finite increments:


for i=2:N
L(i)=L(i-1)+incL;
if L(i)<25
Tc=345;
else
Tc=295;
end
k1=k10*exp(-E1/(0.008314*T(i-1)));
138 6 Scaling and Stability of Chemical Reactors

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)

0.05 80 250 96.2 85.1


0.067 60 155 88.7 76.2
0.1 40 127 82.3 71.1
0.2 20 115 77.3 68.9
(Adiabatic) (0) 110 74.5 66.9

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.

6.3.1 Study of the Stability of the Process


Chemical reactors, especially those in which one exothermic reaction is going on,
are often difficult to control. This is the case when large temperature changes are
produced due to variations in one or more operational parameters. These zones
of operation are known as “parametrically sensitive areas” and “areas of runaway.”
In the last ones, a small change the initial input conditions produces results in the
system not returning to the initial state, whereas in the parametrically sensitive
areas, the system is able to return to the initial conditions.
Operating the reactor in an unstable region can produce a defective product,
a not well-controlled temperature, rapid deterioration of the catalyst, unwanted
side reactions, etc.
140 6 Scaling and Stability of Chemical Reactors

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

than N ad . As the conversion approaches unity, d(Δ𝜐)


dXA
becomes a growing negative
number, so Δ𝜐 will tend to zero. So, except in the case where Δ𝜐 is always negative,
there must be a point for each solution of (Eq. (6.20)) where there is a Δ𝜐max . The
value of the maximum temperature difference that will occur is what is called the
“hot spot” of the reactor.
We do note that this is true only if taken as independent variable the conversion
of the reactive A, that is, eliminating the differential length (dL) of reactor, as has
been done previously. If the treatment is done using dL, it would not necessarily
be a maximum. This is what is called a “safe approach,” criticized by some authors
for not being very real.
Figure 6.7 shows, as an example, temperature profiles obtained for the case
N c /N ad = 2, for various values of N ad and assuming that the initial temperature
matches the temperature of the refrigerant (that is Δ𝜐 = 0) and a single curve
obtained in the case Δ𝜐0 = 2.
142 6 Scaling and Stability of Chemical Reactors

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

5 Figure 6.9 System behavior for


Nc /Nad = 5 and Nad = 8.
4
Δυ0 = 3
2.52 3
Δυ

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

The corresponding value of the conversion is obtained by substituting in


Eq. (6.21):
Nc
XA,mm = 1 − (6.24)
e ⋅ Nad
Other studies use the temperature T mm as the reference temperature instead
of the temperature of the cooling medium, T c . Thus, the condition Δ𝜐mm ≤ 1 is
converted into
E
2
(Tmm − Tc ) ≤ 1 (6.25)
RTmm
That is equivalent to
R 2
T − Tmm + Tc ≤ 0
E mm
The application of Eq. (6.25) gives greater values of T mm . Solving the
second-order degree equation (6.25) you get to
√( )
RTmm ⎡ 4RTc ⎤⎥
≤ 0.5 ⎢1 − 1− (6.26)
E ⎢ E ⎥
⎣ ⎦
This is what is called the first stability criterion, and it provides the conditions
inside the reactor that produce a risk situation. But what matters is knowing
6.3 Rapid Exothermic Reaction in a Tubular Reactor 145

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.

Example 6.1 Batch Reactor


In a batch tank stirred reactor in which the reaction A → P → R occurs and in
which the desired product is “P,” the following data is valid:
k 10 = 0.5 s−1 , k 20 = 25 000 s−1 , pre-exponential factors of the constants k 1 and k 2
E1 = 20 kJ/mol, E2 = 60 kJ/mol, activation energies
ΔH 1 = −300 kJ/mol, ΔH 2 = −250 kJ/mol, reaction enthalpies
C A0 = 1000 mol/m3 , C P0 = 0, initial concentration of A and P
U = 0.5 kJ/(s m2 K), global coefficient of heat transfer.
Calculate the cooling temperature profiles necessary to maintain the reactor at
300 K throughout the reaction, for a cylindrical reactor of 42 cm in diameter and
another of 400 cm.

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)

270 Small reactor


Large reactor
250
230
210
190
170
150
0 0.2 0.4 0.6 0.8 1
Molar conversion

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.

Example 6.2 Jacketed Piston Flow Reactor


In a jacketed piston flow reactor, the first-order reaction A → R is carried out, in
which the following constants are present:
k 0 = 1⋅109 s−1
U = 150 W/(m2 K)
Diameter = 250 mm
E/R = 15 300 K
𝜌⋅C P = 1300 J/(m3 K)
ΔH r = −1800 kJ/mol
(a) If it is intended that at the entrance the food is introduced at the temperature
of the refrigerant, which is 300 ∘ C, what temperature in the reactor makes
the system thermally unstable? What concentration of maximum at the inlet
makes the reactor stable?
(b) Maintaining the concentration calculated before, and estimating that the
maximum stability temperature is 800 K, what value of the heat transfer
coefficient makes it possible? What if it was 900 K?
6.3 Rapid Exothermic Reaction in a Tubular Reactor 147

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

Example 6.3 Using Matlab to Calculate Stability Conditions


In a tubular reactor system, the A → P reaction is occurring with the following
constants:
k = 3.94⋅1012 ⋅exp (−11 400/T) s−1
ΔT ad = 146 ∘ C
4U/(dt⋅𝜌⋅C P ) = 0.2 s−1
The inlet temperature of the reagents is, in any case, 340 K.
(a) Calculate the temperature-conversion profiles for various refrigerant tem-
peratures between 335 and 350 K. Determine the maximum temperature that
will be obtained in the reactor at each of these temperatures.
(b) Determine the variation of N c /N ad with the cooling temperature.
(c) Determine the Δv-conversion curves for different values of the quotient
N c /N ad .
(d) Calculate the curve Δvmax -conversion for this system.
(e) If the temperature of the refrigerant is given by the conditions of the equip-
ment and is 335 K, determine what will happen when changing to a system
whose diameter is four times higher.
(f ) Using the “quiver” function of Matlab, draw the relative size and the direc-
tion of the derivatives of the representation Δv-conversion, in the case that
N c /N ad = e.

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.

Program Listing 6.4


clear all
close all
% Data given in the example:
To=340; %K
dTad=146; %K
%4U/dtdCp=0.2; %s-1

dv=[]; % Vector for saving values of dv


T=[];% Vector for saving values of T
DVmax=[]; % Vector for saving the maximum values of dv
NcdivNad=[]; Vector for saving values of Nc/Nad
XAmaxima=[]; Vector for saving values of XA in the
maximum value of dv
incXA=0.001; % Increment of the conversion among two
consecutive points
6.3 Rapid Exothermic Reaction in a Tubular Reactor 149

for Tc=335:350 % Values of Tc


dv=[];
dv(1)=11400/Tc∧ 2*(To-Tc); % Initial value of dv
T(1)=((dv(1)*Tc∧ 2)/11400)+Tc;
XA=0;
i=1;

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

Figure 6.10 Solution of Example 6.4.

The results are shown in Figure 6.10a–e.


Two different situations can be distinguished in the previous plots. In the first
case are those lines that represent controlled systems, in which by increasing the
conversion, the system is able to keep the temperature low (T c ≤ 337 K). The sec-
ond case is represented by those lines that for conversion values close to 1, a
runaway is produced, since there is a loss of temperature control in the reactor,
which continues to increase without control (335 < T c < 350 K).
The maxima of the Δ𝜐 − X curve can be clearly observed (Figure 6.10c) until
T c > 337 K, where the curves have lost their maximum and the runaway is
produced.
6.3 Rapid Exothermic Reaction in a Tubular Reactor 151

(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.

Program Listing 6.5


% Stability of a tubular reactor
Nad=24;
Nc=exp(1)*Nad;
% Creating the mesh
v=linspace(0.1,5,20);
ta=linspace(0.05,0.8,20);
[V,Ta]=meshgrid(v,ta);

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

Example 6.4 Change in Initial Concentration


The liquid-phase sulfuric acid catalyzed hydration of 2,3-epoxypropanol
(ΔH r = − 88 200 J/mol) has been studied in a batch reactor by Heiszwolf and
Fortuin. The reaction takes place by a first-order kinetics reaction dependent on
the concentration of the alcohol. The kinetic constant is defined by
( )
8809 −1
k = 8.25 ⋅ 109 ⋅ exp − s
T
The authors show that the system is stable when the coolant temperature is
306 K starting at a concentration of alcohol that produced a dimensionless adia-
batic temperature rise of 13.1, but it presents a high increase in temperature when
this temperature rise is 15.9. In all runs, the heat transfer capacity was maintained
constant at a value of 23.7. The density of the fluid is 1000 kg/m3 . The overall heat
transfer capacity is 50 W/K and the overall heat capacity of the mix and vessel is
3500 J/(kg K).
(a) Reproduce the system behavior and check if the limits in ΔT ad produce any
change according to the reasoning presented before.
(b) The system will be scaled up to a plug flow tubular reactor with a diameter
of 5 cm. Study the behavior of the continuous system and predict one stable
and one nonstable initial concentration of the alcohol.

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

If the balance is multiplied by the total volume, we have


dT
m ⋅ CP ⋅ = (−r1 ) ⋅ V ⋅ ΔH1 + U ⋅ (Tc − T) ⋅ A
dt
For this system, we can define modified dimensionless values, N’ad and N’c ,
equivalent to those used in the plug flow reactor:
U ⋅A E ⋅ ΔTad
N ′c = N ′ad =
kc ⋅ m ⋅ CP R ⋅ Tc2
In the same way as in the plug flow reactor, we can check that Eq. (6.20) is
valid in a batch reactor using these last dimensionless values:
d(Δ𝜐) Δ𝜐 ⋅ exp(−Δ𝜐)
= N ′ad − N ′c
dXA (1 − XA )
The values of dimensionless adiabatic temperature rise given in the state-
ment, refer to the N ′ c and N ′ ad we have just defined, so we have
N ′c = 23.7

Nc
N ′ad1 = 13.1 ′
= 1.81 (Experimentally System stable)
Nad1

Nc
N ′ad2 = 15.9 ′
= 1.49 (Experimentally Runaway)
Nad2

Nc
Both situations present a value of ′
Nad
much lower than “e,” so a runaway situ-
ation is probably found, depending on the specific conditions. Let us graphically
see how the system will perform. For that, a program very similar to that shown
in Program Listing 6.6 will be used:

Program Listing 6.6


T0=306; % Input temperature (K)
Tc=306;
Nc=23.7;

for Nad=[13.1 15.9] % Values of Nad to be used


XA(1)=0; % Initial conversion
dv(1)=(T0-Tc)*11600/(Tc∧ 2); % Initial value of the
dimensionless temperature difference
% Integration of the equation by using finite increments:
for i=2:99 % We will use 100 increments in XA
D=Nad-Nc*dv(i-1)*exp(-dv(i-1))/(1-XA(i-1)); % Value of
the derivative d(Dv)/dXA
XA(i)=XA(i-1)+1/100; % The increment of conversion is
1/100
dv(i)=D*(XA(i)-XA(i-1))+dv(i-1); %Corresponding value of
dimensionless temperature difference
end
plot(XA,dv,'.')
hold on
154 6 Scaling and Stability of Chemical Reactors

end
axis([0 1 0 12])
legend('Situation 1','Situation 2')
xlabel('X_A')
ylabel('Dimensionless temperature difference')

As a result of this calculation, the following plot is obtained


12
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

As we can check, situation 2 is much more conducive to a runaway situation.


If the initial temperature of the mix is somewhat higher than that of the cooling
media (Δ𝜐 > 0), there will probably be a quick increase in temperature making a
runaway.
(b) In the proposed tubular reactor, we have A/V = 2/R = 2/(0.05/2) = 80 m−1 .
Let us calculate the adiabatic and cooling numbers from the data:
CA0 ⋅ (−ΔHr ) CA0 mol∕l ⋅ 88 200 J∕mol
ΔTad = =
𝜌 ⋅ CP 1000 kg∕m3 ⋅ 3500 J∕(m3 K)
= 0.0252 CA0 (in K when CA0 in mol∕l)
Assuming that the global transfer coefficient is constant, the dimensionless
numbers are
A 50
U⋅ V A
J∕(s m2 K) ⋅ 80 m−1
Nc = =
kc ⋅ 𝜌 ⋅ CP 2.56 ⋅ 10−3 s−1 ⋅ 3500 J∕(kg K) ⋅ 1000 kg∕m3
0.4464
= with A in m2
A
E ⋅ ΔTad 8809 ⋅ 0.0252 CA0
Nad = = = 0.002 37 CA0 with CA0 in mol∕l
R ⋅ Tc2
3062
N
In that situation, c = 188.3∕(CA0 ⋅ A) with A in m2 and CA0 in mol∕l
Nad
Bibliography 155

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

Barkelew, C.H. (1959). Stability of chemical reactors. Chemical Engineering Progress


55: 37–46.
Heiszwolf, J.J. and Fortuin, J.M.H. (1996). Runaway behaviour and parametric
sensitivity of a batch reactor – an experimental study. Chemical Engineering
Science 51 (11): 3095–3100.
Jiang, J., Jiang, J., Pan, Y. et al. (2011). Investigation on thermal runaway in batch
reactors by parametric sensitivity analysis. Chemical Engineering and Technology
34 (9): 1521–1528.
Morbidelli, M. and Varma, A. (1982). Parametric sensitivity and runaway in tubular
reactors. AIChE Journal 28: 705–713.
Moulijn, J.A., Makkee, M., and van Diepen, A. (2013). Chemical Process Technology,
2e. Wiley.
Van Welsenaere, R. and Froment, G.F. (1970). Parametric sensitivity and runaway in
fixed bed catalytic reactors. Chemical Engineering Science 25: 1503–1516.
Vogel, H. (2000). Process development, Principles of Chemical Reaction
Engineering and Plant Design. In: Ullmann’s Encyclopedia of Industrial
Chemistry, vol. B4 (eds. B. Elvers, S. Hawkins and G. Shulz), 437–475. Weinheim,
Germany: Wiley-VCH Verlag GmbH & Co. KGaA.
Westerterp, K.R., Van Swaaij, W.P.M., and Beenackers, A. (1984). Chemical Reaction
Design and Operation, 2e. Chichester: Wiley.
157

Forced Unsteady State Operation of Chemical Reactors

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.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
158 7 Forced Unsteady State Operation of Chemical Reactors

FUSO product Final product

Storage tank

Figure 7.1 Effect of regulation in unsteady operations.

However, despite designing to work in a steady state, the chemical engineer


must face situations in which the process is not stable over time, as, for example,
in start-up processes and stops or regeneration of catalysts, which are usually
problematic, by disturbing the work routine of the plant, so generally we tend to
avoid them as much as possible. Why then consider work in an unsteady regime if
it seems to be clearly disadvantageous? This chapter tries to answer this question,
considering the most common ways of operating in a FUSO regime.
In the first place, as already indicated, it is necessary to specify that when speak-
ing of operation in FUSO, it does not refer to a batch reactor, in which the design
variable is precisely time, but to the intentional disturbance of the steady state in
a continuous reactor. The regeneration of catalysts has points in common with
this type of operation, but it is not considered in this study since, except in rare
cases, the periods between regenerations take days at the least, and often months
and even years, by what is the normal operation in the steady state. What we are
dealing with here are continually transitory processes. In addition, the regenera-
tion tries to return the initial performance to the process, while the objective of
the FUSO is to improve that performance with respect to the steady regime.
However, we must not forget that we are still looking for a homogeneous pro-
duction over time, so any type of steady-state disturbance cannot be used. It is
intended as a periodic variation, so that the average production in time is con-
stant, and that is why the FUSO is also known as periodic operation or cyclic
operation. In addition, it will be necessary to introduce storage tanks that collect
the different fractions and provide an approximately constant output, whether it
is the final output of the process or whether it will pass to other units or devices
(separation, purification, another reactor, etc.). Figure 7.1 schematizes the objec-
tive sought with the FUSO.

7.2 Objectives and Types of FUSO


Obviously, if it is proposed to use the FUSO as a way of working, some improve-
ment in the process has to be produced with respect to the steady regime. For
example, in catalytic systems, the phenomena of adsorption, desorption, diffu-
sion, etc. always play a role. Essentially, for a catalytic reaction to take place, these
phenomena are dynamic, and this aspect could be exploited in a more efficient
way in a periodic reactor. From another point of view, the steady state is simply
7.3 Periodic Variation of the Input 159

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).

7.3 Periodic Variation of the Input


The terms periodic operation, cyclic operation, forced periodicity, modulation,
etc. are used interchangeably and refer to those operations in which one or more
of the inputs to a chemical reactor vary with time, but in such a way that this
variation is repeated causing a periodic signal.
Figure 7.2 shows a system in which there is the flow of reagents A and B period-
ically between two values, so that a pulse string representing the variation of the
concentration of reactants in the stream is generated. This example corresponds
to a well-studied system, but there are other variables or input conditions that
can also be varied, such as reactor temperature, total flow velocity, and pressure
or volume of the reactor.
160 7 Forced Unsteady State Operation of Chemical Reactors

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.

7.3.1 Modes of Operation


Many types of cycles can be carried out in periodic operations. On the one hand,
there are systems with more or less equal cycle divisions (with s = 0.5, the most
studied), and on the other are those where one of the parts of the cycle is very
short, that is, the “pulse mode.”
7.3 Periodic Variation of the Input 161

Figure 7.3 Comparison of the


input signals of a steady- state
operation with another carried
out in an unsteady state. Source:
Time-average rate
Silveston et al. (1995). Adapted

Reactor rate
with permission of Elsevier.

Instantaneous cycle-
invariant rate

A1
Mole fraction of reactants in feed

A2

Cycle
period
τ

Mean mole
fraction

Cycling operation

The classification of the modes is usually done according to the characteris-


tic response of the system or the relaxation time (t c ) of the catalytic reaction or
reactor that is being forced periodically. At one extreme are the cycles with long
periods (𝜏 ≫ t c ), in this case the reactor behaves as if it were operating in a steady
state, and, in this way, it is called “quasi-steady mode.” At the other extreme, of
short periods (t c ≫ 𝜏), two situations can appear:
(i) If there is a good and rapid mixture, then the system behaves as if it were in
a steady state, and the average value of the forced variable is obtained at the
input. This situation is reached in reactors with recirculation or in which the
mixing level is very high (special tanks).
(ii) If the degree of mixing is negligible, or the section is small, the “relaxed
steady-state mode” is reached.
Between both extremes, there is a region where 𝜏 ≅ t c , in which two modes can
also be given: on the one hand, there is the “pulse operation,” when the offset at
the inputs is zero, and on the other hand is the “standard cyclic mode,” when
the offset is 180∘ (while reagent A increases, B decreases). The “pulse mode”
is interesting in practice, since it can cause a “chromatographic” effect to be
162 7 Forced Unsteady State Operation of Chemical Reactors

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.

7.3.2 Design Strategy


The choice of which entry is the one that interests to vary, the way to vary it, and
the structure of the cycle that must be carried out are part of the strategy to be
followed in order to develop an efficient periodic operation. The steps to get it
established are listed here.

7.3.2.1 Choice of the Entrance to Manipulate


One input to a reactor is a variable that can be manipulated, and most inputs can
be forced periodically. In the same system, one can force more than one entry
simultaneously. The possible inputs to be manipulated are flow rate (it has only
been studied by simulation; it does not seem very useful for single-phase systems,
but it does seem useful for biphasic and three-phase systems); composition of the
feed (they are the most studied systems); pressure (systems little studied because
the effects are scarce); feeding temperature (they are also well-studied systems by
simulation as well as by experimentation); and wall temperature (i.e. heat trans-
mission condition, the temperature of the furnace, or a cooling or heating fluid
is varied, but high thermal inertia makes it difficult that the periodic cycles affect
the ordinary functioning of the system).

7.3.2.2 Choice of Handling


If a variation of the concentration is intended to be carried out, at least two com-
ponents are required, which may be both reactants, or a reagent and an inert
component. In this case, two options can be followed: that both components vary
simultaneously, or that the concentration of one of them remains constant and
that of the other varies periodically.
If three components are considered, the possible combinations are more
numerous, and, therefore, there are more possibilities to establish the periodicity
in the system: varying three, two, or only one component.
For the cases in which the periodic variations are in phase, as the composition
of the feed varies, the input flow varies. To avoid this, a new component must be
introduced, an inert one, whose flow will also vary periodically with time (thus,
the number of options is also extended).

7.3.2.3 Choice of Mode


There will be different possibilities: choice between periodic mode or operation
by pulses, relaxed or quasi-steady state, exchange between pure reactants or a
mixture of them, or between an inert and a mixture of reagents, etc.
Some cases are distinguished from others according to the variables of
periodicity that are modified. The difference between periodic and pulsed
strategy lies in the value of the split(s), and in the same way, the difference in the
exchange between mixtures and pure reactants is in the amplitude or that of the
quasi-steady states and the relaxed state, which are distinguished by the period
7.3 Periodic Variation of the Input 163

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).

7.3.3 Periodic Variation of Concentration


This is undoubtedly the most studied FUSO together with the periodic reversal
of the flow, although it has not yet reached the industrial level. In this mode of
operation, the input that varies is the reactive concentration.
For example, if the catalytic oxidation of CO is carried out with O2 , in a
steady-state operation the CO adsorbed on the catalyst surface predominates,
with O2 adsorption being the rate-limiting step due to the absence of empty
pores in the catalyst. However, if a periodic operation is used by cyclically
introducing CO and O2 , the distribution of the adsorbed species is substantially
varied, and the reaction rate is increased, although the total concentration of
adsorbed CO varies very little. The explanation for this phenomenon can be
found using the model of the islands (Figure 7.4).
In the steady state, large sets of adsorbed CO (islands) are formed, and if the
reaction rate is proportional to the perimeter of these islands, it would be con-
venient to have a smaller size, increasing the number of islands and the total
perimeter. This effect is achieved thanks to the periodic variation of the concen-
tration. In addition, another favorable situation is reached, and that is of a new
species being formed on the most reactive surface (CO mixed with surface O),
which leads to a higher reaction rate. This is summarized in Figure 7.4, where
part (a) represents the steady-state situation and (b) the periodic operation.

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

Significant increases in selectivity can also be achieved. For example, it is pos-


sible to achieve the partial oxidation of certain hydrocarbons in the absence of air
using redox catalysts. The catalyst provides oxygen to the hydrocarbon by form-
ing the products of partial oxidation, and in a second step and in the absence of
hydrocarbon, the catalyst can be oxidized again by air to return the initial activ-
ity. If both reagents are not in direct contact, and one is introduced first and then
the other, the selectivity of the process improves significantly. Even if the conver-
sion decreases, production (which is what interests the engineer) may increase.
Studies on the oxidative dehydrogenation of propane to obtain propene show
increases in production of more than 30%.

7.3.4 Periodic Variation of the Flow


This way of causing the system to work in an unsteady state is applied to sys-
tems in which the presence of more than one phase is necessary, as is the case of
three-phase fixed-bed reactors (trickle-bed), where coexists the gas, liquid, and
solid catalyst.
Cyclically varying the flow rate of the fluid phase in a catalytic reactor can
modify the hydrodynamics in the catalyst bed and improve its operation. This
variation can cause an increase in the degree of wetting of the catalyst, since
the film of liquid that surrounds each catalyst particle is much thinner than that
which would result in a steady-state operation. Thus, an increase in the reaction
rate can be achieved.
In addition to influencing the mass transfer from the system, the modulation
of the fluid affects in a very direct way the contact time of the liquid, the catalyst
particles, and the interfacial holes, so it alters the permanence of the liquid in the
catalytic bed, thus influencing the residence time of the reagents.
The temperature of the catalyst will also be affected, since a permanent temper-
ature oscillation will be created between the catalyst particles, which can cause
an increase in the average reaction time.
Anyway, the most important effect of the periodic variation of the input flow to
the reactor will be that produced on the mass transfer between the liquid and the
solid, since they will be provided new contact channels previously unavailable.
The industrial implementation of this type of systems is difficult. The decrease
of the liquid flow can cause a bad distribution of the temperature and the for-
mation of hot spots in the system. The fact of periodically shutting off the flow
of liquid supposes a strong risk for very exothermic reactions, although it can be
applied for less exothermic reactions, such as the oxidation of SO2 . This process
for the regenerative elimination of sulfur dioxide from flue gases is characterized
by the following reaction:
1
SO2 + O2 + H2 O → H2 SO4
2
where water is used both to remove the reaction products in the form of sulfuric
acid and to absorb heat. The resistance to transport offered by the aqueous phase
decreases the reaction rate, but this phase is essential for the development of the
reaction, so it cannot be eliminated.
7.4 Periodic Flow Reversal 165

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.

7.3.5 Periodic Variation in Temperature


Another input that can be changed periodically for FUSO is the feed temperature.
The investigations in this sense are not very advanced, mainly due to the difficulty
in the control. These studies are limited to the presentation of some satisfactory
data that improve the steady process, but do not go further.
For reversible exothermic processes, oscillating temperatures in the catalytic
bed open an opportunity to increase the conversion, without having to resort to
elimination of heat in the bed. This option has been investigated for many pro-
cesses, for example, in the synthesis of ammonia.

7.4 Periodic Flow Reversal


Let us have an exothermic reaction carried out in an adiabatic fixed-bed reactor,
with the catalyst, preheated, initially at a high temperature. There are several pos-
sibilities of behavior, but one of them is as follows (Figure 7.5): when the relatively
cold feed current is introduced, the temperature of the bed logically decreases
while that of the feed increases; and during this period there is basically no reac-
tion. As soon as the ignition point is reached, the temperature rises sharply and
the reaction begins to elapse. The next fraction of food entering the reactor finds
the cold catalyst in the first zone of the bed due to the previous cooling by the
reactant stream, then it reaches the hot zone, which is cooled by this new cur-
rent at low temperature until, a little more advanced in the reactor, the point of
ignition is reached again. The result is a temperature front that moves as time
passes along the reactor, which logically causes a reaction front; it is observed
that as time passes, in more reactor length there is no appreciable reaction. This
is observed in Figure 7.5.
This behavior has two important characteristics. The first is that the maximum
temperature of the front is much higher than that predicted by the adiabatic
behavior of the system and it does not depend only on the kinetics of the reaction,
but on the transport processes in the bed as well. The second refers to the rate of
propagation of the front, considerably lower than the rate of mass transfer, which
carries reagents and products. These two effects are a consequence of the time
scale and the inertia commented on earlier and make the carrying out of catalytic
166 7 Forced Unsteady State Operation of Chemical Reactors

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

Length Length Length Length

Figure 7.6 Development of a temperature profile during operation in “reverse flow.”

processes in this regime of temperature fronts interesting (for example, in front


of an adiabatic operation). However, after a certain time, the temperature front
would “come out” of the reactor, and then the bed would have to be heated again,
which reduces the possible interest in this technique.
A periodic inversion of flow (periodic flow reversal) has been proposed to avoid
this phenomenon (Figure 7.6). The idea is that before the temperature front leaves
the reactor, the food stream is inverted, that is, it is introduced through the other
end of the reactor. With this, it is possible to generate a new temperature front
that moves in the opposite direction. After a certain time, a typical temperature
profile is developed, as can be seen in Figure 7.6, which can be experimentally
achieved with a simple set of valves. In addition, with the bell-shaped profile
obtained, a lower temperature at the outlet of the reactor is achieved, which
favors the equilibrium of the exothermic reaction. It could be said that this way
of operating forces the reaction to go “beyond the balance.” These profiles, which
are practically impossible to achieve continuously, are very close to the optimal
behavior of the system.
Obviously, an analogous reasoning could be done with an endothermic reaction
generating inverted temperature profiles compared to the previous ones, but this
would no longer be an advantage to displace the equilibrium, but a problem.
7.4 Periodic Flow Reversal 167

In the example described, implemented industrially in some countries (or on


which studies are being carried out in a large pilot plant, especially in Russia),
what is achieved is to combine the reactor and the heat exchangers in one unit,
in such a way that the heat of the reaction is used, saving on the use of additional
equipment and passing the reactor from several stages to one, and saving also
on necessary catalyst. In addition, the reaction occurs mainly in the center of the
reactor, so it is not necessary that the entire bed is catalytic, being able to put at
the beginning and at the end of the reactor an inert filler that fulfills its tasks as a
heat transmitter, and yet is cheaper.
Figure 7.7 shows a simple way to get reverse flow in a reactor with a combi-
nation of two three-way valves. The flow direction is controlled by opening and
closing simultaneously both valves, as indicated in the figure.

7.4.1 Operation Design


The design of a periodic operation of this type must be carried out through exper-
imentation for each particular system. However, unlike other FUSOs, such as the
periodic variation of concentration at the entrance, there are not many parame-
ters to be determined, the main design parameter being the half-cycle time, which
must simply be checked to be less than the time that it would take the gener-
ated front to leave the reactor. The main problem of the operation with periodic
inversion of the flow is that at the beginning of each half cycle there are large vari-
ations in the reactor output concentrations. To solve this, several methods have
been proposed, among which it is important to collect all the output fractions in
a storage tank that provides a constant output (Figure 7.1); however, this would

Figure 7.7 Reverse flow in a Three-way valve


reactor with a simple
combination of two three-way
valves.

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.

7.5 Operation with Variable Volume (VVO)


From the point of view of the reaction mixture, a chemical process can operate
at an intermediate point between the two ideal mixing ends, that is, between
the total mixture (ideal perfect stirred) and minimum mixing (ideal piston flow).
Operating in an unsteady regime can also lead to this intermediate mixing
situation.
Manipulating the type of mixture is equivalent to varying the residence time,
and there is no doubt that this is a very important parameter in all types of
reactors; therefore, the possibility of acting on it is very interesting. This can be
achieved by the so-called VVO, which is a convenient method for manipulating
the residence time in a stirred tank reactor. This mode of operation can convert
a batch process into a continuous one without losing many of the benefits of its
batch nature while at the same time reducing its typical disadvantages.
The mode of operation of a VVO reactor is outlined in Figure 7.8. We can
observe up to four different periods in each cycle: the feeding period, the “batch”
period (in which the volume in the reactor is kept at the maximum value), the
discharge period, and the “rest” period (in that the reactor volume is at its min-
imum value). The duration of the individual periods, as well as the amplitude of
the oscillations, are parameters or conditions to be optimized.
VVO is widely applied in biotechnological processes, where microorganisms
are used as catalysts. This specific application of VVO reactors is known as
“fed-batch” reactors (see Chapter 10). In this equipment, the crop starts as it does
in a discontinuous reactor, but at a later stage, when the period of exponential
growth has been exceeded, more food is introduced into the reactor, without
extracting material simultaneously (which is the difference of a continuous
fermentation). The system continues feeding food with a certain flow rate, which
is reduced with time in a predefined way.

Figure 7.8 Mode of


τ
Vmax Batch operation of a variable
volume operation.
Feed
Reactor
volume
Discharge
Vmin
Rest

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.

7.6 Oscillating Pressure


In reactors that work with oscillating pressure, the pressure varies periodically,
and the rest of the conditions and/or variables will do so accordingly. In these
cases, the equilibrium of the system can be favored because the products of the
reaction are eliminated, for example, in a reaction catalyzed by a solid, by des-
orption.
A simple system operating with oscillating pressure is seen in Figure 7.9. In this
case, a tubular reactor contains a mixture of catalyst and adsorbent. The gaseous
reactant is pumped into the packed bed at high pressure during a part of the
operation cycle. An additional volume (“surge volume”) is placed between the
reaction bed and the outlet stream to allow reflux (purging) and perhaps higher
product purities.
As can be supposed, there are several modes of variation of the pressure, in the
form of sinusoidal, triangular, square, and “bang–bang” or “on–off” signal. The

Figure 7.9 System for a Feed


periodic separation reactor. Exit

Scape

Tubular reactor packed with Extra volume


catalyst and sorbent (“Surge volume”)
170 7 Forced Unsteady State Operation of Chemical Reactors

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

Stankiewicz, A. and Kuczynski, M. (1995). An industrial view on the dynamic


operation of chemical converters. Chemical Engineering and Processing: Process
Intensification 34: 367–377.
Sterman, L.E. and Ydstie, B.E. (1991). Periodic forcing of the CSTR. An application
of the generalized II-criterion. AIChE Journal 37: 986–997.
Vanden Bussche, K.M. and Froment, G.F. (1996). The STAR configuration for
methanol synthesis in reversed flow reactors. Canadian Journal of Chemical
Engineering 74: 729–734.
Xiao, W.D., Wang, H., and Yuan, W.K. (1999). Practical studies of the commercial
flow-reversed SO2 converter. Chemical Engineering Science 54: 4645–4652.
Zhou, X., Barshad, Y., and Gulari, E. (1986). Transient response and rate
enhancement through forced concentration cycling. Chemical Engineering
Science 41: 1277–1284.
Zwijnenburg, A., Stankiewicz, A., and Mouljin, J.A. (1998). Dynamic operation of
chemical reactors: friend or foe? Chemical Engineering Progress 94: 39–47.
173

Part II

Catalytic, Multiphase and Biochemical Reactor Design


175

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.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
176 8 Industrial Catalysis

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

Figure 8.1 Trickle-bed reactor.


Gas in

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

8.1.2 Solid Catalysts (Supports)


One important aspect in the catalytic reactor design is the nature and properties
of the catalyst itself. It is important then to know the most important required
properties of the solid catalysts and the industrial way to get these properties.
This would give the engineer a background knowledge that will improve his/her
ability to do a good design.
Generally speaking, the chemical nature of the material used for catalyst devel-
opment determines whether a catalyst’s applicability may be conditioned by its
physical properties.
Figure 8.2 shows a scheme of the main requirements for an industrial catalyst,
which is to get a material with high activity and selectivity for the reaction of
interest. This material can exist as a simple catalyst, but usually it is necessary to
support the active catalyst on other materials that can have or not have catalytic
activity.
The external surface of a catalyst is generally too small for the required prop-
erties of surface-to-volume ratio, so the catalyst should be porous. It is difficult
to achieve at the same time high porosity and high mechanical strength, so the
necessity of using catalyst supports is justified.
Supports are usually essential to prepare porous and nonporous catalysts. Sup-
port is generally an inert material where the active component is dispersed. Mate-
rials used for supports must have a high mechanical strength, high surface area,
high pore volume, etc. and the selection of the material is done as a function of the
operation parameters and the reaction system. Besides the mechanical strength,
the supports provide the following other properties:

Catalyst requirements Figure 8.2 Catalyst key


points.

Resistance to
poisoning Intrinsic activity
(specially to and selectivity
coke)
Appropriate
distribution of
Stable structure pores (and pore
sizes)

Reproducible Adequate thermal


behavior properties

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

Diameter Diameter Diameter Diameter


Uniform Outer egg shell Inner egg yolk Middle egg white

Figure 8.3 Typical distribution profiles of active phase in a spherical support.

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.

8.1.2.1 Choice of Catalyst Support


The choice of the support depends on several factors. The first requirement is that
it must be stable under the reaction conditions. Many materials are not resis-
tant to high temperatures. An example is the transformation of highly porous
γ-alumina into α-alumina at temperatures exceeding 1400 K. Activated carbon
has a high thermal stability, but it is oxidized in an oxygen-containing environ-
ment. Acidic or basic solutions can cause the dissolution of oxidic supports. In
addition, catalysts requiring regeneration must be based on supports that are sta-
ble under regeneration conditions. For example, hydrocracking and FCC catalysts
must be stable in an oxidative environment at high temperature.
The texture of the catalyst support (average diameter of the pores, shape, and
connection among pores, etc.) influences the mass and heat transfer character-
istics. In the case of pores of very low diameter (too narrow) or very long ones,
the selectivity will be reduced in partial oxidation processes, because reactions
completing the oxidation are favored. In this respect, zeolites are a special case
due to their pores of molecular dimensions.
The interaction between the support and the active phase also is critical. One
of the functions of the support is to stabilize the active phase against sintering, as
mentioned, and a strong interaction favors this stability, but care must be taken
because the chemical characteristics of the active phase could be also modified.
Obviously, the cost of the support must be as low as possible. Other properties
of interest are density, heat capacity, and size/shape.

8.1.2.2 Comparison and Uses of Supports


In Table 8.1 we find the main characteristics of some materials used as catalyst
supports. Silica and alumina are the most common supports. Alumina is available
in many forms, of which a high surface area (γ-Al2 O3 ) and a low surface area
(α-Al2 O3 ) are shown.
8.1 Introduction 181

Table 8.1 Characteristics of some support materials.

Surface area Pore volume Mean pore size


Material (m2 /g) (cm3 /g) (nm)

γ-Al2 O3 180–200 0.6 3.5


α-Al2 O3 3–5 — —
SiO2 150–400 0.5–1.2 Bimodal (5–50 and >50)
SiO2 –Al2 O3 150–160 0.2–0.7 10
TiO2 (anatase) 50 0.34 9
C (activated carbon) 200–1500 0.1–2 1 (often bimodal)
Zeolites 400–700 0.2–0.6 0.4–1

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.

8.1.2.3 Silica, Alumina, and Mixtures


Silica and alumina are both prepared by a method that starts with precipita-
tion from a concentrated solution of the metal salt. The resulting hydrogel, after
purification, can be formulated into different shapes, such as powders, pellets,
extrudates, or simple spheres.
Silica has no catalytic activity of its own. It is used as a support for Ni, Pd,
or Pt in gas-phase hydrogenation reactions, and also as a support for vanadium
pentoxide (V2 O5 ) in the oxidation of SO2 .
γ-Alumina is used as support in the catalysts for exhaust-gas treatment,
hydrotreating catalysts, and some hydrogenation catalysts. This material itself
has catalytic activity, and is used as catalyst for alcohol dehydration and in the
conversion of H2 S into the elements.
Silica-alumina is a mixed oxide that can be produced by impregnation of porous
silica with an Al2+ solution. This support contains Lewis and Brønsted acid sites
(this is explained later), constituting a remarkable strong acid, more acidic than
silica or alumina alone.

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

Sodalite unit (24 tetrahedral units)

Zeolite super-structure (Faujacite type)

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

Table 8.2 Use of zeolite in a modern oil refinery.

Oil fraction Catalytic process

Gas Aromatization and alkylation


Light ends Isomerization
Naphta Hydrotreating and reforming
Middle distillates Hydrotreating and dewaxing
Luboils Dewaxing
Flashed distillates Catalytic cracking and hydrocracking
Residue Hydrodesulfurization and hydroconversion

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.

8.1.2.5 Activated Carbons (ACs)


ACs are carbons with high specific surface area. Carbon is extensively used in
industrial adsorption applications, but also in catalytic chemistry, due to the
exceptional properties it presents, as they possess the following:
– High stability both in the presence of acids and bases,
– Very high adsorption capacity of a variety of molecules, especially organic
molecules,
– Wide variety of textures properties,
– Extraordinary resistance to high temperatures,
– Not much erosion on pumps.
Other properties of the ACs make difficult their application in oxidizing envi-
ronments, as they are easily burnt. Also, the AC can present a weak mechanical
strength, making it difficult to avoid the formation of fines, which can make sep-
aration difficult.
Many biomass feedstocks can be used for preparing ACs. Wood, coal, and lig-
nite are the most common. Usually, the manufacturing process consists of pyrol-
ysis (heating in inert atmosphere) followed by activation (with steam or oxygen)
to increase the porosity and surface area.

8.2 Industrial Preparation of Catalysts


The industrial manufacture of the catalytic converters is characterized by its
secret end, not only by the manufacturer but also by the user. The catalyst market
184 8 Industrial Catalysis

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.

8.2.1 Structure of Commercial Catalysts


As mentioned, during the design of the catalysts, special attention is required
to understand the factors that can limit the rate of the process. In the design of
environmental catalysts, the diffusion phenomena are minimized by synthetizing
them in a process divided into three steps.
In this way, many catalysts have a monolithic ceramic cylindrical structure
(cordierite) that has a three-dimensional scale. The cordierite comprising the
monolith cannot be used to support the catalysts due to its low surface ratio
since it is not capable of efficiently dispersing the metal particles. In this sense,
a stable material of high surface (coating or “washcoat”) is adsorbed in the
cordierite structure. Then, the metal active particles are dispersed in the coating
(Figure 8.5, see also Figure 9.8). Some examples of materials used as coating are
Al2 O3 , SiO2 , TiO2 , CeO2 , ZrO2 , V2 O5 , La2 O3 , and zeolites.

8.2.1.1 Synthesis of Zeolites


Zeolites are produced in the “hydrothermal process.” This is the common name
of the process, which involves water and heating of different sodium solutions.
Usually, sodium hydroxide, sodium aluminate, and sodium silicate are used, and
an “aqueous glass” is formed. This forms a gel that is then converted into the
crystalline zeolite. The structure and shape of the zeolite will depend on the com-
position, temperature, reaction time, and pH of the medium. A scheme of the
process is given in Figure 8.6.
By the type of reagent used in the synthesis, the sodium salts of the zeolites are
actually obtained, not the acid form. To convert them into the active form, they

Washcoat Metallic active sites


Cordierite

Figure 8.5 Catalyst structure and active components.

Sodium GEL Amonium salt NH3 (g) +


aluminate + (sodic salt of the zeolite Protonic form 450 °C
350 °C
Sodium silicate + of the Amonium salt
Sodium hidroxide zeolite) H2O + Lewis acidic
form of zeolite

Figure 8.6 Scheme of the production of zeolites.


8.2 Industrial Preparation of Catalysts 185

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.

8.2.1.2 Manufacturing of the Catalytic Support


The conventional supports of the catalysts for refining of petroleum have as raw
material the aluminum oxide, Al2 O3 , usually in the 𝛾 type. A plant for preparing
alumina supports has the following basic parts:

(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).

8.2.1.3 Impregnation with Active Metals


Active metals for catalytic activity are obtained from different salts that are dis-
persed in the surface of the supports. From the metals most used in the manu-
facturing of catalysts, the following can be highlighted.

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

Molybdenum The salt that is used in most industrial preparations is ammonium


heptamolybdate ((NH4 )6 Mo7 O24 ), which is soluble in water, although it can be
expensive.

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.

8.2.2 Key Definitions in Catalysts Performance


It is important to be familiar with some characteristics of a catalyst, related to
the ability to increase the rate of a reaction. First of all, the activity of a catalyst
refers to its ability to accelerate the overall conversion of a reactant to products.
Mathematically, activity can be considered as the quotient between the reaction
rate in a precise moment and the reaction rate with a fresh catalyst (usually, t = 0):
(−rA )t
a= (8.1a)
(−rA )t=0
In the studies arising from the deactivation of a catalyst, i.e. its loss of activity,
the reaction rate at any moment is given by
(−rA )t = (−rA )t=0 ⋅ a (8.1b)
while the loss of activity is modeled through equations of the type:
da
= f (C, a) (8.2)
dt
where C is be the concentration of the species producing the deactivation.
On the other hand, the residence time in a catalytic reactor can be calculated
as the volume of the catalyst bed divided by the reagent flow rate, that is
catalyst volume
Residence time = (8.3)
(volume reagents∕time)
While this concept is equivalent to the reaction time in a batch reactor, it does
not give a measure of the ability of the catalyst to convert reactants into prod-
ucts. For example, if we have 100 mL of a catalyst A capable of converting 100%
ethane flowing at 100 ml/min, it will have the same residence time as a catalyst
B that can convert 100% of a current that enters only 10% ethane in the catalyst
bed. However, catalyst B is only able to convert 1/10 of the amount that catalyst
A converts.
A better measure of the conversion of reactants in a particular catalyzed reac-
tion is the “space velocity.” In that concept, the conversion is measured per unit
time and catalyst, so it indicates the rate at which a given catalyst is able to convert
reactants into products per unit of time. This spatial velocity can be expressed
with a volumetric base (which could be “liquid hourly space velocity,” LHSV, or
“gaseous hourly space velocity,” [GHSV] that will be given in vol/vol-h) or in a
mass basis (weight/weight-h, WWH). Both parameters are usually expressed in
inverse units of time.
8.2 Industrial Preparation of Catalysts 187

Table 8.3 Examples of space velocity.

Process WWH or LHSV (h−1 ) GHSV (h−1 )

Chemicals products manufacturing 0.01–1 10–1 000


Refinery 1–10 1 000–10 000
Exhaust gas elimination 10–100 10 000–100 000

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

where SD is the density of sites, which is measured by spectroscopic or chemical


methods.
TR is a valid definition of reaction rate. Indeed, it is the most fundamental
one, as it reflects the frequency at which molecules are reacting on the catalyst.
Turnover rates, as other reaction rates, must be defined at a definite temperature
188 8 Industrial Catalysis

and concentration of reactants. To be acceptable, TRs should be measured under


conditions where only chemical effects are observed, i.e. in the absence of heat
and mass transport limitations. This, as will be seen in Chapter 9, is generally got
at low temperature and low conversion.
TRs have units of moIecules/site-second or s−1 . Values of TRs for commercial
catalysts are typically in the range of about 10−3 –101 s−1 , while TR values for
enzymes are in the range of 102 –104 s−1 .

8.3 Main Catalytic Processes in Industry


Some important processes catalyzed in industry are summarized in Table 8.4,
with special interest in the oxidation/reduction reactions. In this section, we
batch them into four groups of reactions.

Table 8.4 Important industrial catalyzed reactions with the corresponding catalyst.

Main Type of
Reactants products/coproducts Type of catalyst reactiona)

Ammonia/air NO Pt–Rh gauze SO


Ammonia/air N2 Pd-V2 O5 -WO3 / SO
TiO2 –SiO2
H2 S/air S8 Fe2 O3 /SiO2 SO
Methane/O2 /NH3 HCN Pt–Rh gauze SO
CH4 or (CH2 )x/O2 Syn gas (CO/H2 ) Supported Rh or Ni SO
Methanol/air Formaldehyde Ag, Fe–Mo–oxide SO
Ethylene/O2 Ethylene oxide Ag(K,CI)/a-AI2 O3 SO
Ethylene/air or 1,2-Dichoroethane Cu/CI/O/K-yAl2 O3 SO
O2 /HCl
Ethanol/O2 Acetaldehyde Ag, Cu SO
Ethylene/acetic Vinyl acetate Pd/Au/K-aAI2 O3 SO
acid/air
Ethylene/O2 Acetaldehyde Pd/CI/acetate SO
supported
Propene/air Acrolein Bi/Mo/Fe/Co/ SO
K-oxide
Propene/air/NH3 Acrylonitrile Bi/Mo/Fe/Co/ SO
K-oxide
Propene/ Epoxide/alcohol Mo-complexes or SO
hydroperoxide silica supported Ti
Acrolein/air Acrylic acid V/Mo/W-oxide SO
n-Butane/air Maleic anhydride V-P-oxide SO
n-Butane/O2 or air Acetic acid Co, Mn salts SO
n-Butane/air Butenes/butadiene Bi-Mo-P-oxide SO
F-butyl alcohol Methacrolein Bi-Mo-Fe-Co- SO
K-oxide
8.3 Main Catalytic Processes in Industry 189

Table 8.4 (Continued)

Main Type of
Reactants products/coproducts Type of catalyst reactiona)

Isobutene/air Methacrolein Bi-Mo-Fe-Co- SO


K-oxide
Methacrolein/air Methacrylic acid V/Mo/W-oxide SO
Cyclohexane/air Cyclohexanone Co salts SO
Cyclohexanone/ Adipic acid Co salts SO
HNO3
Cyclohexanone/ Cyclohexanone Ti-silicalite SO
H2 O2 /NH3 oxime
Benzene/air Maleic anhydride V-Mo-oxide SO
o-Xylene/air Phthalic anhydride V-W-P-Cs-TiO2 SO
p-Xylene/air Terephthalic acid Co/Mn/Br SO
Cumene/O2 or air Phenol/acetone Co-Mn salts SO
Naphthalene/air Phthalic anhydride V-K-oxide/SiO2 SO
Alkanes Alkenes Cr/Al2 O3 , Cr/ZrO2 , DH
Pt/Sn/Al2 O3 ,
Mo/Al2 O3
n-Butane Isobutene Cr/AI2 O3 , DH
Pl/Sn/Al2 O3 , alkali
promoted
Pt/Sn/Al2 O3
Ethylbenzene Styrene Fe2 O,/K2 O/Cr2 O3 DH
Cyclohexanes Benzenes Ag, Cu gauze, Pt/C, DH
Pd/C
Alcohols/ Ketones Cu, Cu chromite. DH
aldehydes Cu/K/SiO2
Cyclohexanol Cyclohexanone Cu-ZnO/AI2 Oj , DH
Cu/MgO,
Cu/K/SiO2 , Ru/C
Acids Alcohols Re, Ru H
Aliphatic Alcohols Ru H
aldehydes
Aromatic Alcohols Pd H
aldehydes
Aromatic Hydrocarbons Pd H
aldehydes
Unsaturated Unsaturated Os, Re, Pt H
aldehydes alcohols
Mono-alkenes Alkanes Pd H
Alkadienes Alkenes Pd H
Cyclohexenes Cyclohexanes Pt, Rh H
Cyclohexenes Cyclohexane + Ru, Rh H
benzene
(Continued)
190 8 Industrial Catalysis

Table 8.4 (Continued)

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

a) Type of reactions: SO, Selective oxidation; DH, dehydrogenation; H, hydrogenation.

8.3.1 Acid Catalysis


In the classic Brønsted definition, acids are sources of protons, but acids can be
also contemplated as substances accepting electrons (in the definition of Lewis
acids), i.e. acids are electron-poor materials. The elements of the periodic table
used in acid catalysis are those showing acid characteristics, as is the case of
hydrogen (in the form of H+ ), transition metals with high oxidation states (Si, Al)
, and elements of the upper right part of the periodic table. Also, halogens are
used due to their very high electronegativity and the high strength of the conju-
gate acid. The elements that are the most used in industrial acid catalysts are H
and Si/Al/O such as silica, alumina, and a wide variety of materials containing
these compounds in the form of zeolites or clays.
Brønsted and Lewis acid catalysts are used. Protonic acids (Brønsted) react with
substances by donation of a proton, while Lewis acids do not possess protons,
8.3 Main Catalytic Processes in Industry 191

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.

8.3.1.1 Fluid Catalytic Cracking


Let us see in more detail the FCC process. During the FCC, the heavier frac-
tions of petrol are converted into products of lower boiling point and higher
added value. The cracking reaction takes place in a reactor (“riser”) at 500 ∘ C
approximately and a residence time of two to four seconds. The gasoil is pre-
heated and contacts the catalyst coming from the regenerator (see Figure 8.7) in
the bottom part of the riser. The products and the catalyst are separated in the
unit called “stripper”; the deactivated catalyst is then passed to the regenerator.
In that equipment, the combustion of the coke deposited in the surface of the
catalyst is produced. This process is done at 700 ∘ C approximately in the pres-
ence of air and steam. Later, this catalyst is turned back to the riser, continuing
the cycle.
The catalyst used in the FCC process has a quite complicated formulation.
Among others, the catalyst contains alumina, clays, and silica in the support, and
platinum, antimony, and sodium in the active phase, although the main compo-
nent is a zeolite. Each of these active points intervenes in different reactions. For
example, zeolites are active in the cracking of heavy fractions, and platinum is
192 8 Industrial Catalysis

Figure 8.7 Scheme of the FCC


Products
process.

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

8.3.1.2 Ethylbenzene Production


Ethylbenzene (C6 H5 CH2 CH3 ) is an alkyl aromatic compound with a single ring.
It is also known as phenyl ethane and ethylbenzol. This compound is mainly
(>90%) used as an intermediate for the manufacture of styrene (C6 H5 CH=CH2 ).
It is one of the most important large-volume chemicals and is the monomer of
polystyrene.
The main method of obtaining ethylbenzene is from the alkylation of benzene
and ethylene. The technology and the advances in catalysts have allowed develop-
ing a method respectful of the environment and with safe conditions. It is based
on an alkylation reactor, where the reaction takes place, a transalkylation reac-
tor, to recover the aromatic compounds with more than one ethyl radical, and a
purification unit to separate the desired product (Figure 8.8).
The alkylation unit consists of the supply lines of the two necessary reagents and
the auxiliary equipment needed. The main element of this zone is a multitubular
reactor, which is where the reaction takes place.
Benzene enters via line 1 (Figure 8.8) and pump B-1 compresses it to the reactor
pressure. Exchangers E-1 and E-2 heat it up to a temperature of 235 ∘ C. Line 2 is
that of ethylene and works in a similar way. Compressor C-1 brings ethylene up
to 40 bars and exchanger E-1 decreases its temperature until it is the same as that
of benzene to enter the reactor. Lines 9 and 10 run through the cooling water of
the multitube reactor to control the temperature.

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.

8.3.2 Oxidation Catalysis


The formation of partial oxidation products in a preferential manner was not
commercialized until selective oxidation was done in a realistic manner, in
the late 1960s, when appropriate catalysts were discovered. These catalyst are
constituted by elements that should present a cycle between several oxidation
states, offering a path of low activation energy to form partial oxidation products.
In this way, the elements present in the oxidation catalysts must present multiple
oxidation states, among which the transition metals stand out, and of course
the oxygen. The most important will be V, Mn, Cr, Co, Fe, Ni, Nb, Cu, Mo,
Ru, Ru, Pd, Rh, Ag (transition metals) and O, Sb, S, Te, Sn, Bi as other active
elements.
The design of a catalyst for selective oxidation is based on many parameters.
Among them, the selection of the metal that will be part of the active center
is obviously one of the most important. The isolation of active sites is another
key concept in the design of the catalyst. A strong oxidant such as Cu2 O3 is
very active, but not very selective for the oxidation of olefins. Nevertheless, the
selectivity can be increased if some part of the copper oxide is reduced using a
reductant such as H2 . Also to be borne in mind is that the surface of an oxidation
catalyst is usually chemically different from the internal zone.
Table 8.4 shows the most important processes in this industrial oxidation catal-
ysis. We describe in detail two of these processes.

8.3.2.1 Ethylene Oxide Production from Ethylene


Ethylene oxide (C2 H4 O) is an organic compound that belongs to the family of
cyclic ethers. At room temperature, ethylene oxide is a flammable, toxic, pale gas
that readily dissolves in water.
This compound is an important intermediate in obtaining ethylene glycol, this
being its main use. A small amount (less than 1% of production) is used in hos-
pitals for sterilization of medical equipment and supplies that cannot be steril-
ized by thermal processes. The world production of ethylene oxide accounted to
approximately 20 million tons in the past years.
For a long time, the only method to obtain ethylene oxide was the so-called
Würtz synthesis, which consisted of a two-stage process (with ethylene chloro-
hydrin as an intermediate product). In 1931, the French chemist Theodore Lefort
managed to develop a silver catalyst able to directly oxidize the ethylene. This
process was much cheaper on an industrial scale than previously used; therefore,
8.3 Main Catalytic Processes in Industry 195

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

Figure 8.9 Ethylene recirculation diagram.


196 8 Industrial Catalysis

8.3.2.2 Acrylic Acid from Acrolein


Acrylic acid production in the industrial scale has been studied for many years.
Different processes have been tested in the past, involving the use of acetylene
(HC=CH) or ketene (CH2 =C=O) as reactants. Also, the hydrolysis of acryloni-
trile was introduced as a valid method, but usually it occurs as an excessive
polymerization, producing big amounts of ammonium bisulfate as by-product,
causing a disposal problem.
A promising route to produce acrylic acid is the partial oxidation of propy-
lene in vapor phase. It is generally carried out in a two-step process, implicating
acrolein as an intermediate. The general reaction scheme is the following one:

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

This is a complex reaction scheme, in which there is a precursor, acrolein,


formed from the reaction R1, and by-products of carbon oxide and water by the
combustion of both acrolein and acrylic acid. Acetic acid, formaldehyde, and
acetaldehyde are also produced in the process, although the amount produced
depends to a large extent on the conditions under which the reaction is carried
out. Both the oxidation of propylene to acrolein and the oxidation of this inter-
mediate to acrylic acid are partial oxidation reactions, carried out in catalytic
reactors.
The second reaction of the previous scheme, the production of acrylic acid by
oxidation of acrolein, is usually carried out using a catalyst containing antimony,
nickel, and molybdenum in the form of oxides.
The commercial applications of acrylic acid and its esters are numerous. For
example, they are used as binders for paints, and as adhesives. The process of
production of paper, leather, and textiles also has a final treatment with these
chemicals.
The temperature of the process is usually in the range 285–360 ∘ C, with a clear
increase of the conversion (0.15–0.45) with temperature. Nevertheless, the selec-
tivity of the reaction decreases as temperature increases, so a low-medium tem-
perature is adopted. At 285 ∘ C, the conversion is 0.2 approx. for a space time of
two seconds, and the selectivity is increased at low catalyst particle size, with an
optimum of 92% of selectivity with a particle diameter of 0.8 mm.

8.3.3 Reduction Catalysis


Fuels and other chemicals are produced using catalysts that operate in a reducing
atmosphere. Hydrogen, the simplest and cheapest reducing species, is included
as a reactant or product in these reactions.
8.3 Main Catalytic Processes in Industry 197

Table 8.5 Catalytic processes in the syngas production.

Reaction Catalyst used

Reforming with methane vapor: conversion of CH4 and H2 O to NiO/SiO2


CO/H2
Methanation: conversion of CO/H2 to methane NiO/SiO2
Methanol synthesis: conversion of CO/H2 to methanol Zn/Co/SiO2
Methanol decomposition: conversion of CH3 OH to CO/H2 Co/SiO2
Fischer–Tropsch reaction: conversion of CO/H2 to hydrocarbons Co/SiO2 Fe/SiO2
in the range of gasolines
Steam reforming of hydrocarbons: conversion of hydrocarbons Co/SiO2
and water into CO/H2
Oxygenated synthesis: conversion of CO/H2 into oxygenated Rh/SiO2
hydrocarbons of higher molecular weight

The catalysts employed in hydrogenation catalysis or reduction are typically


transition metals in their low oxidation states. Usually, noble metals such as plat-
inum, palladium, and rhodium are used. Also, basic metals (Fe, Co, Ni, Zn) are
active in some cases. In some processes, water is used as reactant and the H of
this molecule is used as reducing agent.
The synthesis gas, a mixture of H2 and CO, is an important raw material for the
synthesis of numerous chemicals and fuels. Some of the products are as impor-
tant as methanol, methane, higher hydrocarbons, and oxygenates. The synthesis
gas can be obtained by coal gasification, and then represents a source of chemical
products apart from petroleum.
The proportion of CO in the synthesis gas can be modified by converting it to
H2 through the water-shift reaction, so that the reactions involved in the synthesis
gas would be those indicated in Table 8.5, together with the most commonly used
catalysts. Some of them are commented on in detail.

8.3.3.1 Steam Reforming of Alcohols


The complete process for energy production using a biomass source may be
divided into four stages, namely: bioethanol production, hydrogen production,
hydrogen purification, and power generation. Usually, these stages are treated
separately, and scarce information is available on the process integration.
Obviously, the first stage (Figure 8.10) depends largely on the type of biomass
used and also on the purification steps. The main products are ethanol and
water. In the second stage (H2 production from bioethanol), usually the steam
reforming of ethanol (SRE) is run in order to increase the hydrogen yield.
The SRE process is an endothermic process, mainly represented by the reaction:
CH3 CH2 OH + 3H2 O (+ heat) → 2CO2 + 6H2
This equation actually represents a very wide network of reactions that will
depend on the catalyst used, and also obviously on the reaction conditions. The
final product is a complex mix of by-products. Among them, the main are carbon
198 8 Industrial Catalysis

(2) H2 production (4) Power generation


H2O
H2
H2 stream
CO free of CO
Biomass SRE CO2
Fermentation CH4 CO removal Adsorber Fuel cell
Condenser

Solids and heavy aliphatics Air


CO2
H2O
(1) Bioethanol production CH4
(3) H2 purification

Figure 8.10 Complete process for energy production from biomass.

monoxide, ethane, methane, and ethylene, and, in some cases, acetaldehyde


(CH3 CHO).
One of the most important reactions being produced in this process is the
water–gas shift reaction (WGSR):
CO + H2 O ⇌ CO2 + H2 (+ small amount of heat)
This WGSR favors CO2 formation with increasing H2 and decreasing CO con-
centration in the outlet stream.
In the power generation stage (Figure 8.10), fuel cells (FCs) are usually used,
most of them demanding a very low (<100 ppm) CO concentration. If this is the
case, the by-product formation should be minimized, and a previous H2 purifica-
tion stage is usually needed. Actually, CO removal from post-reformed streams is
currently one of the biggest problems to connect H2 -producing systems to com-
mercially available FCs devices.
SRE catalysts are usually constituted by a support where base metals such as
Ni, Cu, Al, Zn, and Co, and also noble metals such as Pt, Rh, Pd, Ru, and Ir are
dispersed. The ethanol decomposition route depends on the metal present, but
the catalyst support can also influence the reaction. Furthermore, a low metal
loading is usually used (typically, less than 1%), due to the high costs. This pro-
motes the role of the support on the reaction mechanism. A support avoiding
a high oxygen storage is usually more effective, properties that are conferred by
supports that can be reduced, such as CeO2 , ZrO2 , and La2 O3 .
Conventionally, CO removal by conversion through WGSR produces an outlet
CO content close to 2000–5000 ppm, too high for FC use. A second stage may be
needed, which ensures a low CO (<100 ppm).
For reducing the amount of CO in the mix, two different processes are com-
mon: (i) preferential oxidation of CO (CO-PROX, Eq. (8.5)) and (ii) CO metha-
nation (Eq. (8.6)):
2CO + O2 → 2CO2 Preferential oxidation of CO, CO − PROX (8.5)
CO + 3H2 → CH4 + H2 O CO methanation (8.6)
Nevertheless, high H2 loss is common during the final stage because of the
presence of secondary reactions such as
CO2 + 4H2 → CH4 + 2H2 O CO2 methanation (8.7)
2H2 + O2 → 2H2 O H2 oxidation (8.8)
8.3 Main Catalytic Processes in Industry 199

8.3.3.2 Steam Reforming of Hydrocarbons (Methane)


Actually, bioethanol reforming is not the main source of hydrogen. Nowadays,
more than 95% of the hydrogen produced in the world is produced by reforming
of natural gas (methane) in large central plants. In the industrial H2 production,
natural gas containing methane (CH4 ) is used as a precursor of the hydrogen, by
treating it with steam at high temperature (700–1000 ∘ C). Methane (and other
hydrocarbons) reacts under high pressure in the presence of a catalyst to produce
the hydrogen and CO in the WGSR.
A final process can be included in the reaction network, called “pressure-swing
adsorption,” where CO2 and other by-products are removed from the gas stream,
producing hydrogen as pure as possible. It is also possible to obtain a CO/H2 mix
by partial oxidation of hydrocarbons, with a limited amount of oxygen. This is
an exothermic process that is usually much faster than WGSR, but produces less
hydrogen per unit of the input fuel than is obtained by steam reforming of the
same fuel.

8.3.3.3 Methanation: CO/H2 to Methane


Methanation is the chemical process where carbon monoxide and/or carbon
dioxide are converted to methane. The reaction is the opposite to that of the
WGSR, and is based on the following exothermic reaction:
CO2 + 4H2 → CH4 + 2H2 O ΔH < 0
Carbon monoxide conversion to methane is also possible (CO methanation), by
CO + 3H2 ↔ CH4 + H2 O ΔH < 0
Generally, CO2 is used for methane production, and few plants use CO instead.
The former reactions are exothermic, so they are favored at low temperatures.
The temperature control (efficient heat removal) is one of the greatest challenges
involved in methanation, closely linked to reactor design. Also, the equilibrium
is displaced to the products under conditions of high pressures.
The simplest option for a catalytic reactor is probably the adiabatic fixed bed.
The reactor consists of a bed of catalytic pellets, and its temperature is increased
along the reactor due to the heat production. The reaction rate is quite high and
a low amount of catalyst is usually needed. As temperature is increased, the CO2
conversion decreases, and the catalyst can be deactivated via sintering. For that
reason, the reactor feed gas is usually diluted, to limit in any way the temperature
increment.
Nickel-based catalysts are considered the best option for methanation. Nickel
oxide (NiO) exhibits a very high activity and selectivity for producing methane.
This is mainly due to the ability of NiO to undergo a reduction process. Another
reason for the high activity of NiO is that it usually presents a bimodal pore
structure, i.e. two sizes of pores, which is optimum for the adsorption of both
reactants (CO2 and H2 ). Other more expensive metals such as Rh, Ru, Pd, and
Pt also present high activity in the methanation of CO or CO2 , but usually their
high prices do not make them competitive.
Nevertheless, it has been proved that the addition of Pd and Pt considerably
enhanced the behavior of Ni/Mg(Al)O catalysts, allowing decreasing the tem-
perature and increasing the hydrogen consumption.
200 8 Industrial Catalysis

Also, the selection of support is important. As already commented on, it may


influence both the activity and selectivity of the reactions. Particularly important
is the fact that the addition of alumina eliminates the interference that introduces
the presence of H2 S, thus increasing methanation activity. Therefore, Al2 O3 is
generally an inexpensive and convenient support.

8.3.4 Environmental Catalysis


The reduction of environmental pollutants is one of the most important objec-
tives achieved by the chemistry of catalysis. The United States, Europe, and
Japan have required in the past years a drastic reduction of pollutants emission.
Especially important are the depletion of emissions of particulates (mainly
carbonaceous particulates), NOx, hydrocarbons, and CO. This applies to
both stationary and mobile sources, but the most important problems are in
nonstationary sources (vehicles), which will be treated with greater importance.
For example, the simultaneous reduction of carbon and NOx in diesel com-
bustion gas was a particularly complicated technical problem since the required
reduction levels were more than 90% compared to historical levels. Almost all the
techniques that produce this reduction call for the presence of catalysts.
Catalysts for the conversion of exhaust gases fall into one of these four cate-
gories:

Noble metals Basic metals Alkaline earth metals Rare earth metals

Pt, Pd, Rh V, Fe, Co, Ni, Cu Mg, Ca, Sr, Ba La, Ce


(alumina supported) (silica supported)

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.

Table 8.6 Typical composition of the exhaust gases of diesel and


gasoline engines.

Component Gasoline Diesel

NOx 1 050 ppm 270 ppm


Soot/PM 60 mg/km 1 300 mg/km
CO 6 800 ppm 350 ppm
Hydrocarbons (HC) 750 ppm 350 ppm
H2 2 300 ppm 120 ppm
O2 5 100 ppm 60 000 ppm
8.3 Main Catalytic Processes in Industry 201

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

NOx, CO, NOx, NH3


PM, CO, HC

NOx, PM NOx, NH3

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

NOx, CO, NOx


PM, CO, HC

Figure 8.11 Some catalytic processes present in the gas exhaust of diesel and gasoline
engines.

8.3.4.2 Components of the Cleaning Systems


An outline is shown in Figure 8.11.

8.3.4.3 Three-Way Catalyst


The catalyst used for the simultaneous elimination of NO, CO, and HC has the
following main components:
* Alumina of the 𝛾 type (γ-Al2 O3 ) with high surface. This component acts as a
support of the noble catalytic metals such as Cerium. The noble metals act as
oxygen buffers, storing O2 when it reaches a certain critical level.
* LaO2 (lanthanum oxide), which stabilizes the γ-Al2 O3 and prevents Rh from
being eliminated from the support.
* Platinum (Pt), which is the oxidation catalyst of CO, HC, and NO2 .
* Rhodium (Rh), which is the catalyst for the decomposition of NOx.
In some compositions, Pa (palladium) is used for replacing a part of the Rh or
Pt. The emission of CO, HC, and NOx in the vehicle engines mainly depend on the
ratio air:fuel, and also on other factors as the precise moment of the ignition. The
general behavior of the emissions with the ratio air:fuel is shown in Figure 8.12.
For obtaining a high conversion of all pollutants simultaneously, a stoichiometric
air:fuel ratio (marked in figure by S*, close to 14.6) should be maintained.
The main reactions being produced in the system are as follows:
2NO + 2CO → N2 + 2CO2 (R1)
Cm Hn + (m + n∕4)O2 → mCO2 + n∕2 H2 O (R2)
2CO + O2 → 2CO2 (R3)
8.3 Main Catalytic Processes in Industry 203

Figure 8.12 Variation of the 100


conversion of the three-way catalyst
reactions as a function of air/fuel ratio.

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.

8.3.4.4 SCR Catalyst


SCR is the process in which ammonia is introduced as a reducer for the con-
version of NO to N2 . As mentioned before, the ammonia is produced in situ by
hydrolysis of urea. Three types of catalysts are used in this process:
• Supported basic metal: V/TiO2 or W/TiO2 , or mixtures of both metals. Vana-
dium catalyzes the reduction, anatase (titanium oxide) provides high surface
area, and tungsten oxide stabilizes anatase, which is necessary to have a high
conversion. A good control of the ratio between ammonia and NOx is always
required. If the ratio is too high, ammonia, which is also a pollutant, will be
emitted and this will increase the cost of the process. Contrarily, at low values
of the ratio, NOx conversion will not be sufficient. The optimal molar ratio is
1 : 1 = NH3 :NO.
• Supported noble metal: Pt/Al2 O3 . Platinum acts as a catalyst for the reduction
of NO with ammonia. Alumina provide high surface area.
• A zeolite, which catalyzes the reduction of NO by NH3 . It may contain
Fe and Cu.
204 8 Industrial Catalysis

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.

8.3.4.5 Diesel Oxidation Catalyst (DOC)


The DOC catalyst uses a Pt system supported in alumina to convert CO into CO2 .
The oxidation of HC and soot also occurs. The DOC also has a NOx absorber
system to reduce NOx in exhaust gases. This process combines components that
catalyze the reduction of NOx with the capture of NO. Barium oxide (BaO) is
used as a storage medium. In poor fuel conditions (excess air), NO is initially
oxidized to NO2 , followed by reaction with barium oxide to form barium nitrate:
NO + 1/2 O2 → NO2
BaO + NO2 + 1/2 O2 → Ba(NO3 )2
For a diesel engine, the exhaust contains an excess of air so that most of the time
the system is in this state, until all the available BaO is transformed into Ba(NO3 )2 .
At this stage, BaO must be regenerated in fuel-rich conditions by decomposition
of the nitrate, forming NOx and BaO:
Ba(NO3 )2 → BaO + 2NO + 3∕2 O2
Ba(NO3 )2 → BaO + 2NO2 + 1∕2 O2
These two reactions are promoted by the presence of ceria in the catalyst (in
the form of oxide CeO) and also of rhodium. Also, in fuel-rich conditions, the
NOx will be reduced with CO to form N2 and CO2 :
NO + CO → 1/2 N2 + CO2
This last reaction is produced in the Al2 O3 support. The fuel-rich condition
only happens during a very small fraction of time and returns to fuel-poor con-
ditions, working this system in a forced unsteady state operation (FUSO). It will
change between controlled conditions “lean” and “rich”; it is known as modula-
tion lean-rich “poor-rich” and is necessary for the system to work.
These catalysts have a structure similar to that of the diesel particulate filter
(DPF), but the channels are all open.

8.3.4.6 Diesel Particulate Filter (DPF) Catalyst


The DPF system catches particulate matter (PM), formed mainly by carbonaceous
particles, in a ceramic filter where a Pt/Rh/MgO catalyst is supported. The pres-
ence of platinum catalyzes the oxidation of the trapped carbon (soot) to CO2 , in
addition to other hydrocarbons or CO present. Rh inhibits the oxidation of SO2
to sulfate salts at the outlet, which would contribute to PM. Magnesium works as
a promoter for the processes catalyzed by the platinum.
In this device, the end of every two adjacent cells in the monolith is closed,
which forces the gas to pass through the walls of each cell. In the process, the walls
Bibliography 205

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

Afandizadeh, S. and Foumeny, E. (2001). Design of packed bed reactors: guides to


catalyst shape, size, and loading selection. Applied Thermal Engineering 21 (6):
669–682.
Anderson, J.R. and Boudart, M. (eds.) (1981–1985). Catalysis: Science and
Technology, vol. 8. Berlin: Springer-Verlag.
Burrington, J.D. (2016). Industrial Catalysis: Chemistry and Mechanism. Imperial
College Press.
206 8 Industrial Catalysis

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

Catalytic and Multiphase Reactor Design

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.

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
208 9 Catalytic and Multiphase Reactor Design

Energy → With no catalyst, the intermediate


species have more energy

Initial state

The energy barrier diminishes in


the presence of the catalysts, thus Products
increasing the reaction rate

Course of the reaction →

Figure 9.1 Representation of the action of a catalyst.

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.

9.2 Rate Equation in Catalytic Systems


The rate of a chemical process is defined by the following expression:
gA
rA = (9.1)
(extensive property)
where g A is the generation term (mol/time) of a mole balance in the system.
The extensive property is additive for subsystems, and usually for homogeneous
systems, the total volume is used in such a way that the reaction rate is given, for
example, in mol/(s mR 3 ), the subscript “R” referring to the whole reactor.
For heterogeneous systems, the total volume is not always important. For
example, imagine that we have a very big reactor, but a very small amount of
catalyst inside it. If the reaction needs the catalyst to occur, the reaction rate will
be quite low. In fact, for a catalytic reaction, the reaction rate expressed in many
ways can be found. Table 9.1 shows some of the usual forms of expressing the
reaction rate in heterogeneous systems. In the table, “S” is the external surface of
the catalyst, “W ” is the weight of the catalyst particles, “V ” is the volume of the
reactive mixture, and “V p ” is the volume of the catalyst particles.
Indeed, from the expressions in Table 9.1, by clearing g A , the following equiva-
lence can be demonstrated:
gA = rA S = rA′ W = rA′′ V = rA′′′ Vp (9.6)
9.2 Rate Equation in Catalytic Systems 209

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

External surface of (−rA ), kmol/(s m2 cat ) k, m3 L /(m2 cat s) (−rA ) = g A /S (9.2)


catalyst
Weight of catalyst (−rA′ ), kmol/(s kgcat ) k ′ , m3 L /(kgcat s) (−r′ A ) = g A /W (9.3)
3 3 3
Volume of reactor (−rA′′ ), kmol/(s m R ) ′′
k , m L /(m R s) (−r ′′
A) = g A /V (9.4)
mix
Volume of catalyst (−rA′′′ ), kmol/(s m3 cat ) k ′′′ , m3 L /(m3 cat s) (−r′′′ A ) = g A /V p (9.5)

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.

Example 9.1 Second-Order Catalytic Reaction


The second-order decomposition reaction: A → B + 2C is being carried out in
a tubular reactor packed with a catalyst. Component A enters pure at a surface
velocity of 3 m/s, at a temperature of 250 ∘ C, and a pressure of 500 kPa. The rate
constant of the chemical reaction at this temperature is
k ′ = 0.05 m6 ∕(mol kgcat s) (second order reaction)
Calculate the variation of the conversion of A with the length of the reactor for
a total length of 0.1 m in:
(a) Isothermal conditions at 250 ∘ C.
(b) Adiabatic conditions (initial temperature 250 ∘ C)
Data:
ΔC p = 0
k ′ (at 400 ∘ C) = 2 m6 /(mol kgcat s)
ΔH r = −800 J/mol
Catalyst density = 2000 kg/m3
Porosity of the bed = 0.4
Reagent heat capacity A = 15 J/(mol K)

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

Therefore, under isothermal conditions:


2
k ′ CA0 (1 − XA )2
−nA0 dXA = − dW
(1 + 2XA )2
and, reorganizing:
2
dXA kCA0 (1 − XA )2
=
dW nA0 (1 + 2XA )2
The mass of the catalyst is related to the volume of reactor by
( )
kgcat
W (kgcat ) = Vtotal (m3total ) ⋅ 𝜌cat
m3cat
( )
m3cat
× (1 − Porosity of the bed)
m3total
And the volume relates to the length of the reactor (x) by V = S ⋅ x, S being the
section of the bed. On the other hand, Q0 = u0 ⋅ S and nA0 = C A0 ⋅ Q0 . Finally,
we obtain
dXA 𝜌 ⋅ (1 − Porosity of the bed) ⋅ k ′ ⋅ CA0 (1 − XA )2
= cat
dx u0 ⋅ (1 + 2XA )2
9.2 Rate Equation in Catalytic Systems 211

It is easy to apply a finite difference approximation to solve this differential


equation, and we will get
XAi+1 − XAi 𝜌cat ⋅ (1 − Porosity of the bed) ⋅ k ′ ⋅ CA0 (1 − XAi )2
=
Δx u0 ⋅ (1 + 2XAi )2

®
With initial condition X A = 0, we will use Matlab to solve the equation, as
shown in Program listing 9.1:

Program Listing 9.1


clear all

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

XA(1)=0; % Initial value of conversion


L(1)=0;; % Initial value of length
incx=0.0001; % Increment of length
n=0.1/incx; % Number of points

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

In the previous equation, k should be calculated at the temperature of the


differential of length. Applying the finite differences method:
T i+1 − T i ΔHR 𝜌cat ⋅ (1 − Porosity of the bed) ⋅ k (T i ) ⋅ CA0 (1 − XAi )2
=− ⋅ ( i )2
Δx CpA
u0 ⋅ TT ⋅ (1 + 2XAi )2
0
9.2 Rate Equation in Catalytic Systems 213

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.

Program Listing 9.2


clear all

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

T(1)=250+273; % Initial temperature, K


XA(1)=0; % Initial value of conversion
L(1)=0; % Initial value of length, m
incx=0.0001; % Increment of length, m
n=0.1/incx; % Number of points
k=k0*exp(-E_R/T(1)); % m6/mol ⋅ kg ⋅ s
ftemp=(T(1)/523)∧ 2;

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)

9.2.1 Steps in the Catalytic Reaction


Porosity, like density, is an important property of solid catalysts that is used in
calculating and modeling pore diffusional and film mass transfer resistances.
Figure 9.2 shows that the volume of a typical catalytic solid is partitioned into
solids, pores, and voids. The pore volume is more typically distributed among
pores of different sizes, with most of the volume residing in the largest pores.
In porous catalysts, the reaction occurs throughout the solid–fluid interface
both on the external surface and inside the pores of the particle. In order to
deduce the expressions of reaction rates, it is necessary to take into account the
different steps (both physical and chemical) that take place in the reaction process
and that may offer resistance to it, since one or the other of these predominates
steps will have different expressions of rate. Figure 9.2 shows a diagram of the
steps that a reagent A must undergo in order to reach the surface.
The step or steps with the lowest rate that limit the value of the reaction rate
are known as controlling or limiting steps. The different steps are the following:

Section of a catalyst pellet


Gas film containing an ideal pore

Bulk fluid Pore volume Solid volume

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.”

9.3 Rate Equation When Chemical Step Is Limiting


Reaction Rate
9.3.1 Mechanisms of Catalysis
The exact nature of the reasons why the intermediate complex, adsorbed on the
surface of the catalyst, is easily formed is a research topic. Structural require-
ments can be foreseen for the formation and breaking of links, and this has led
to the formulation of an important series of theories. As it has been proved, the
transfer of electrons is involved in the formation of the bonds of the intermediate
compound, and therefore the electronic nature of the catalysts is undoubtedly
important. What we give is not a satisfactory explanation, since it does not exist,
of the different mechanisms, but some examples are proposed with the corre-
sponding explanations presented by the researchers in the bibliography.
For example, acidic catalysts, such as silica and alumina, can apparently act as
Lewis acids (electron acceptor) or Brönsted acids (proton donor). The reaction
mechanism for cracking a hydrocarbon would be as follows:

Catalyst
Hydrocarbon Carbonium ion + 1e–
Catalyst

Olefin (alkene) + Carbonium ion

Catalyst

Olefin + Carbonium ion


216 9 Catalytic and Multiphase Reactor Design

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):

A+L R+L S+L

A·L R·L S·L

This scheme could correspond to the reaction A → S, where the concentration


of R in the fluid would be very small, since the equilibrium of adsorption is very
displaced toward species R⋅L.
9.3 Rate Equation When Chemical Step Is Limiting Reaction Rate 217

9.3.2 Theories About Adsorption


The classical Langmuir theory of adsorption, which has justified a multitude of
experimental data, explains adsorption as a phenomenon in equilibrium:
A+L ↔ A⋅L
If C A⋅L is the concentration of A⋅L (kmol/m2 ), C A is the gas-phase concentra-
tion (kmol/m3 ), C L is the concentration of active centers (kmol/m2 ), when the
equilibrium is reached, it will be fulfilled that:
CA⋅L
KA = equilibrium constant of chemical adsoprtion = (9.7)
CA ⋅ CL
The values of the rates of the direct and inverse process are given by
ra (adsorption rate) = ka CA CL (9.8)
where
ka = ka0 ⋅ exp(−Ea ∕RT) (9.9)
rd (desorption rate) = kd CA⋅L (9.10)
where
kd = kd0 ⋅ exp(−Ed ∕RT) (9.11)
In equilibrium, it will be fulfilled that:
ka CA CL = kd CA.L
CA⋅L k
= a = KA (9.12)
CA ⋅ CL kd
where K A is the equilibrium constant.
We can relate C AL to C t in the following way:
Ct = CL + CA⋅L = CL + KA CA CL = CL (1 + KA CA ) (9.13)
2
where C t is the concentration of active centers (active centers/m ). Then, it is
Ct
CL = (9.14)
1 + KA CA
and the portion of occupied active centers will be
CA⋅L KA CA
= (9.15)
Ct 1 + KA CA
The representation of C A⋅L /C t as a function of C A has the form shown in
Figure 9.3.
From the Langmuir model, expressions of the reaction rate can be obtained as a
function of the concentration of reactants (and products, in reversible reactions).
This can be seen in a simple example: In many catalytic hydrocarbon oxidation
reactions, it has been found that the slow steps, which control the process, are
the chemisorption of the O2 and/or the superficial chemical reaction between
the chemisorbed O2 with the nonadsorbed hydrocarbon in the gas phase. We
will deduce the kinetic equation for both cases. The mechanism of the reaction
is as follows:
218 9 Catalytic and Multiphase Reactor Design

Figure 9.3 Variation of the proportion


1 of active centers bonding to the
ligand.
CAl/Ct
Slope tends to zero

Slope ∼
−1
0
CA

1. Bonding of oxygen to the surface (chemisorption):


O2 + L → O2−L
2. Surface chemical reaction:
O2−L + RH → Compound − L
3. Desorption reaction:
Compound − L → Compound + L
If the controlling process is the chemisorption of O2 , the reaction rate will be
given by step 1 (the others will be very fast), and will be
r = k CO2 CL (9.16)
On the other hand, it is true that the total “concentration” of the active sites is
Ct = CL + CL-O2 (9.17)
As the surface chemical reaction is fast, CL-O2 = 0 and then C L = C t . Therefore:
r = kC t CO2 = k ′ CO2 (9.18)
that is, the reaction rate is directly proportional to the concentration of oxygen
and we will observe a pseudo first-order reaction.
If the controlling step is the superficial chemical reaction, step 1 will be in equi-
librium (constant K ad ), step 2 will be slow and irreversible, and step 3 will be fast
and irreversible. In this case, the rate will be given by
r = k CO2 ⋅L CRH (9.19)
and
Ct = CL + CO2 ⋅L (9.20)
Remembering that:
CO2 ⋅L
Kad = (9.21)
CO2 ⋅ CL
9.4 Rate Expression for External Diffusion as a Limiting Step 219

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.

9.4 Rate Expression for External Diffusion as a Limiting


Step
When the resistance to external diffusion is higher than all other resistances, the
diffusion of the reactant from the bulk of the fluid to the surface of the catalyst
is the limiting step of the reaction rate. The reagent flow that reaches the catalyst
surface is governed by the material transport coefficient k c between the fluid and
the solid. The rate in this case would be given by
g
(−rA ) = − A = kC ⋅ (CAg − CAs ) (kmol∕(s m2 )) (9.25)
Sex
where k C = individual mass transfer coefficient (kmol/(s m2 )/kmol/m3 ),
Sex = external surface of the particle (m2 ), C Ag = concentration of reagent A in
the fluid (kmol/m3 ), C AS = concentration on the surface (kmol/m3 ).
For cases where all other steps are rapid, including the external diffusion of the
products, C AS is the equilibrium concentration of reagent A with the reagents; if
the other steps are very fast, including the chemical reaction, and if the reaction
is irreversible, then C AS is zero.
The step of external diffusion of a reagent is usually the dominant in the reac-
tions that take place at high temperature and with very small catalyst particles.
In these cases, the rate of the overall reaction will be the same for all solid–fluid
reactions, both for catalyzed reactions and for those that are not.
220 9 Catalytic and Multiphase Reactor Design

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 .

Example 9.2 Cracking of Cumene


Cumene is being cracked into benzene and propylene in a fixed bed of zeolite
particles at 362 ∘ C and atmospheric pressure, in the presence of a large excess of
nitrogen. In a section of the reactor, where the partial pressure of the cumene is
0.0689 atm, the reaction rate observed is 0.153 kmol/(kgcat h).
Calculate the % difference in concentration of the reagent between the medium
and the surface of the particle.
Some properties of the reactive mixture and operating conditions are as fol-
lows:
• Average molecular weight of 34.37 kg/kmol,
• Density 0.66 kg/m3 ,
• Viscosity 0.094 kg/(m h),
• Reynolds number 0.052,
• Diffusivity of the reagent through the mixture 0.096 m2 /h,
• Ratio external surface of catalyst/mass of catalyst = 45 m2 cat /kgcat,
• Mass flow rate of the mixture 56.47 kg/(m2 h).
9.5 Reaction Rate When Internal Diffusion Is Slow 221

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%.

9.5 Reaction Rate When Internal Diffusion Is Slow


As mentioned before, the internal diffusion is a step with some special charac-
teristics. First of all, if there exists a high resistance to the diffusion in the pores,
the reaction can still be produced in the exterior surface of the catalyst, so the
effect of the internal diffusion is a kind of a correction of the rate. Mathemati-
cally, this correction is complicated, since the diffusion of the reactant through
the solid is not related in a simple way with the other steps. The general expres-
sion of diffusion of a component A through a gaseous mixture A + B is given by
Fick’s law:
nA = xA (nA + nB ) − c DAB ∇xA (kmol A∕(m2 s)) (9.31)
In this equation, X A is the mass fraction and the molar flux density, nA , depends
on two terms, the first of which is a consequence of the general movement and
the second of diffusion. When there is no global movement, that is, if there is
222 9 Catalytic and Multiphase Reactor Design

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).

9.5.1 First-order Kinetics in Flat Particles


First of all, we will consider the case of a molecule of reagent A, in a reaction
system A → B, which is diffusing through a flat porous pellet through both sides
(Figure 9.4). The chemical reaction step will be assumed of first order.
As it was treated in Chapter 4, we consider the diffusion of a species toward
the center of a catalyst particle. Doing a mass balance in stationary regime for
reagent A in the dV = S dx of the considered pellet, we obtain
Output + accumulation = generation + input (kmol A∕s)

[nA ]x+dx ⋅ S + 0 = rA′′′ ⋅ S ⋅ dx + [nA ]x ⋅ S (9.35)


2
nA being the flow density of A (kmol A/(s m )) and rA′′′
the reaction rate of A (kmol
A/(s m3 cat )), which is to be assumed to be of first order; so it is deduced that:
dnA
= rA′′′ = −k ′′′ ⋅ CA (9.36)
dx
9.5 Reaction Rate When Internal Diffusion Is Slow 223

Figure 9.4 Porous catalyst A x=0


pellet with reagent A entry S dx
on both sides.
x=L
A A

x=0

2L

A A

and from here:


( )
dC
d −De dxA
= −k ′′′ ⋅ CA (9.37)
dx
where k ′′′ is the constant of the first-order reaction (m3 fluid /(m3 cat s)) (sometimes
abbreviated as s−1 ) and De is the diffusivity through the gas mixture (as already
indicated, with units of m3 fluid /(mcat s) ≈ m2 /s).
Therefore, in this particular case:
d2 CA
−De = −k ′′′ ⋅ CA (9.38)
dx2
or
d2 CA
De + k ′′′ ⋅ CA = 0 (9.39)
dx2
The general solution of this differential equation is
CA = M1 ⋅ emx + M2 ⋅ e−mx (9.40)
where the value of m is given by
( ′′′ )0.5
k
m= (9.41)
De
This module “m,” as explained earlier, has the units of 1/mcat . The values M1
and M2 are constants that can be determined by the boundary conditions. One
of these conditions is that on the surface of the pellet, the concentration of A is
equal to that which exists in the fluid. The other assumes that the variation of the
concentration in the middle of the porous pellet is zero (as treated in Chapter 4).
(1) CA = CAs for x = 0 (9.42)
dCA
(2) = 0 for x = L (9.43)
dx
If we take into account the symmetry of the pellet, condition 2 is nothing more
than indicating that in the center the concentration of A will present a minimum.
Condition 1 is also logical.
224 9 Catalytic and Multiphase Reactor Design

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

Figure 9.5 Variation of the concentration of A vs. distance x.

From here, we get


CAs e−mL CAs emL
M1 = M 2 = (9.44)
emL + e−mL emL + e−mL
and, therefore:
CA em(L−x) + e−m(L−x) cosh[m(L − x)]
= = (9.45)
CAs emL + e−mL cosh[mL]
In view of this equation, it can be deduced that there is a decrease in the concen-
tration of reagent as it progresses toward the interior of the pellet. This decrease
depends on a dimensionless parameter “mL,” which is defined as Thiele’s mod-
ule 𝜙. Figure 9.5 shows different graphs C A /C As vs. x/L for different values of the
Thiele’s module 𝜙.
It can be observed that the greater the Thiele module, that is, the faster the
reaction to diffusion, the steeper the decrease in the concentration of A. However,
for small values of the Thiele module, the diffusion through the pellet is very fast
and the concentration of A does not vary appreciably.
To determine the decrease in the rate of the reaction in the case in which the
diffusion through the porous pellet is important, a new magnitude is introduced,
the efficiency factor, which is expressed by
actual average generation (kmol A∕s)
𝜂=
average generation without internal diffusion effect (kmol A∕s)
actual average reaction rate (kmol A∕(s m3 cat))
=
average reaction rate without internal
diffusion effect (kmol A∕(s m3 cat))
9.5 Reaction Rate When Internal Diffusion Is Slow 225

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.

9.5.2 First-order Kinetics in Other Geometries


The problem posed for the porous flat pellet with transport of A through both
sides can be raised in other cases: porous pellet permeable to reagents and
226 9 Catalytic and Multiphase Reactor Design

r + dr Figure 9.6 Porous cylindrical catalyst pellet with


r reagent A entry on the exterior surface.
nA

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

(nA ⋅ S)]r+dr − (nA ⋅ S)]r + 0 = r′′′ A ⋅ dV = −k ′′′ ⋅ CA ⋅ dV (9.53)


where:
S =2⋅𝜋⋅r⋅h
and
dV = d(𝜋 ⋅ r2 ⋅ h) = 2 ⋅ 𝜋 ⋅ r ⋅ dr
Therefore, substituting the expression of “S” and dividing Eq. (9.53) by (2 ⋅ 𝜋),
we get
(nA ⋅ r)]r+dr − (nA ⋅ r)]r = d(NA ⋅ r) = −k ′′′ ⋅ CA ⋅ r ⋅ dr (9.54)
Rearranging:
d(nA ⋅ r)
= −k ′′′ ⋅ CA ⋅ r (9.55)
dr
Introducing Fick’s law:
( )
dC
d −De ⋅ drA ⋅ r
= −k ′′′ ⋅ CA ⋅ r (9.56)
dr
and then:
( )
dC
d drA ⋅ r k ′′′ ⋅ CA
1
⋅ = (9.57)
r dr De
and, operating:
d2 CA 1 dCA k ′′′ ⋅ CA
+ ⋅ − =0 (9.58)
dr2 r dr De
9.5 Reaction Rate When Internal Diffusion Is Slow 227

This second-order differential equation is of the type:


y′ (x)
y′′ (x) + + q ⋅ y(x) = 0
x
This is an ordinary differential equation with a complex function for the con-
centration of reactant, resulting:
( √ ′′′ )
k
CA (r) J 0 R⋅ De
⋅r
= ( √ ′′′ ) (9.59)
CAs J0 R ⋅ kD
e

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

where C Ae is the equilibrium concentration of the reaction, being 0 when treating


an irreversible reaction. This generalized modulus is if the reaction is first order
and reversible with rapid diffusion of the products and, therefore, with constant
concentration inside the particle, converted to

k ′′′
Φ=m⋅L=L⋅ (9.68)
De ⋅ XAe
X Ae being the molar conversion of A in equilibrium with the reaction products (in
the case where the diffusion of the products is also important, the corresponding
equations must be presented and resolved numerically).
For irreversible reactions of nth order, the corresponding expression is
[ ]1
k ′′′ ⋅ CAs
n−1
(n + 1) 2
m⋅L=L⋅ (9.69)
2 ⋅ De
where C As the concentration of C A on the outer surface of the particle.

9.5.3 Limits of Thiele Modulus and Weisz Modulus


As commented on before, a low value of the Thiele modulus indicates a sys-
tem with no diffusion problems, i.e. a system where the effectiveness factor is
unity. A value of Φ = mL < 0.4 is generally accepted to represent a system where
the resistance to internal diffusion is negligible. Another important module giv-
ing an idea of the rate of diffusion vs. chemical reaction is the Weisz modulus,
defined by
(−r′′′ A )Observed ⋅ L2
We = (9.70)
De ⋅ CAs
In the case where mL < 0.4 → 𝜂 ≈ 0.95–1 → 𝜂(mL)2 ≪ 1, we will have
(−rA )Observed = k ⋅ C As and the Weisz modulus is
k ′′′ ⋅ CAs ⋅ L2
We = < 0.15 (9.71)
De ⋅ CAs
On the other hand, in the case where the diffusion in the pores is important,
CAs
we have mL > 4, and 𝜂 = 1/(mL) so (−r′′′ A )Observed = k ′′′ ⋅ m⋅L and then:
k ′′′ ⋅ CAs ⋅ L2 L
We = = k ′′′ ⋅ >4 (9.72)
m ⋅ L ⋅ De ⋅ CAs De
230 9 Catalytic and Multiphase Reactor Design

Example 9.3 Diesel Cracking


A diesel is cracked at 630 ∘ C and 1 atm by passing vaporized feed through a
packed bed of silica-alumina spherical particles with radius 0.088 cm. For a feed
of 60 cm3 liquid diesel/(cm3 catalytic reactor h), the conversion of the diesel is
50%. The density of the diesel is 0.869 g/cm3 , the average molecular weight is
255 g/mol, the overall density of the filler bed (catalyst mass/filling volume) is
0.7 g/cm3 , the catalyst density is 0.95 g/cm3 , the effective diffusivity of the diesel
in the catalyst is 8⋅10−4 cm2 /s. The average concentration of diesel in the volume
considered is 0.6⋅10−5 mol/cm3 . Admitting a first-order reaction:
(a) Calculate the average reaction rate (generation/volume catalytic reactor).
(b) Analyze if there is influence of internal diffusion.
(c) Calculate the efficiency factor.
(d) Estimate the value of the kinetic constant.

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

cm3 diesel 𝜌 g∕cm3


(−r′′ A ) = 60 3
⋅ XA ⋅ diesel
cm reactor h MWdiesel g∕mol
3
cm 0.869 g∕cm3 mol
(−r′′ A ) = 60 3 diesel ⋅ 0.5 ⋅ = 0.1022
cmreactor h 255 g∕mol cm3reactor h

(b) For this purpose, we will calculate the Weisz modulus:


(−r′′′ A ) ⋅ L2
We =
De ⋅ CA
For a spherical particle, we have that L = R/3, and then:
cm3reactor mol 1 cmreactor
3
(−rA′′′ ) = (−rA′′ )(1 − 𝜀) = 0.1022 ⋅
cm3cat 3
cmreactor h 0.7 gcat
gcat mol
× 0.95 = 0.1387
cm3cat cm3cat h
( )2
mol 1h 0.088
(−r ′′′
⋅ (R∕3) 2 0.1387 ⋅ ⋅ cm2
A) cm3cat h 3600 s 3
We = = = 6.2
De ⋅ CA 8 ⋅ 10−4 cm2
⋅ 0.6 ⋅ 10−5 mol
s cm3

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

Example 9.4 Fixed-Bed Reactor Design


In a stirred tank experimental reactor with 10 g of 1.2 mm spherical catalyst par-
ticles and a feed of 5 l/s of pure A at 1 atm and 333 ∘ C, a conversion of 70% is
obtained for the first-order reaction:
A → R ΔHr = 0
We want to design a fixed-bed reactor (considered ideal piston flow) of 1.4 cm
particles. Calculate the amount of catalyst needed in this bed. Data: Solid den-
sity = 2000 kg/m3 De = 10−6 m3 /(mcat s)

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

But in the calculation of “m” the correct units must be used:



√ √
√ 1.166 ⋅ 10−3 m3 ⋅ 103 g ⋅ 2000 kgcat

k ′′′ √ sg kg m3cat
m= =√ m 3
= 48 305 m−1
De −6
10 m s
cat

and L = 4.6⋅10−3 m, so m⋅L = 225.4 and 𝜂 = 4.43 ⋅ 10−3


We can now calculate the weight of the catalyst required:
XAfinal
Q Q
W= ⋅ dXA = − ′ ln(1 − XAfinal ) = 116.4 g
∫0 (k ′ ⋅ (1 − XA ) ⋅ 𝜂) (k ⋅ 𝜂)

Example 9.5 Investigating the Controlling Step


In an experiment to investigate the controlling step of a chemical reaction of
decomposition of a substance “A,” with a catalyst, we found that:
✓ Spherical particles of dp = 2.4 mm of diameter
✓ Effective diffusion coefficient, De = 5⋅10−5 m3 /(h mcat )
✓ Mass transfer coefficient, k C = 300 m3 /(h m2 cat )
✓ First-order reaction, with C A = 20 mol/m3 (at 1 atm and 336 ∘ C)
✓ Observed reaction rate, (−rA )obs = 0.042 mol/(h m2 cat )
Do the needed calculations to investigate which factors are affecting the reac-
tion rate.

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:

−r′′′ A ⋅ L2 105 mol A


h m3cat
⋅ (1.2 ⋅ 10−4 )2 m2cat
We = = = 0.0162 ≪ 0.15
De ⋅ CAs 3
5 ⋅ 10−5 h mm ⋅ 20 mol A
m3
cat
9.5 Reaction Rate When Internal Diffusion Is Slow 233

and there are no internal diffusion problems.


As the observed reaction rate is much lower than that of the external diffusion,
it is only possible that the chemical reaction is dominating the process. In this
way, the lowest rate is that of the chemical reaction and equals 0.042 mol
h⋅m2
A
.
c

Example 9.6 Catalytic Reactor with Spherical Particles


A 50 ml volume reactor is completely filled with spherical catalyst particles 3 mm
in diameter. The porosity of the bed is 0.6. The reactor can be considered as an
ideal piston flow. The decomposition of pure A follows the kinetics of first order:
(−rA ) = k ⋅ CA … (mol∕(s m2cat ))
where C A = concentration of A (mol/m3 ), k = 3⋅10−6 m3 /(s m−2 cat )
For the catalyst, the following data is available: density = 1500 kg/m3 ; specific
surface = 400 m2 /g; diffusion coefficient of reactant = 2⋅10−6 m2 /s
(a) Indicate the expression to calculate the efficiency factor and the design
equation of an isothermal piston flow reactor with a certain amount of this
catalyst.
(b) In the given situation, the conversion of A to the reactor is 0.4. Calculate
the efficiency factor of the catalyst and the volumetric flow rate that passes
through the reactor.
(c) The reaction tube is subjected to a vibration system, causing the density of
the bed to increase, decreasing the porosity to 0.52. Calculate the conversion
that will result.
(d) The 3 mm particles are replaced by the same catalyst mass, but now with a
diameter of 0.03 mm. The porosity of this bed is also 0.52. What is the new
conversion?

Solution
(a) The design equation for a PFR is given by

+(k ′ ⋅ CA0 (1 − XA ) ⋅ 𝜂) ⋅ dW = +Q ⋅ CA0 ⋅ dXA


and integrating:
XAf
Q dXA Q
W= =− ′ ⋅ ln(1 − XAf )
k′⋅ 𝜂 ∫0 (1 − XA ) k ⋅𝜂
Note that the kinetic constant in the mass balance done must be expressed
in units of flow rate/(mass of catalyst), i.e. m3 /(s g). The effectiveness factor
can be calculated using the following expression:
tanh(m ⋅ L)
𝜂=
m⋅L

k ′′′ R
m⋅L= ⋅
De 3

In that expression, the kinetic constant should be given in units of (time)−1 .


234 9 Catalytic and Multiphase Reactor Design

(b) We know that the conversion is X A = 0.4 and:


m3 m2cat g
k ′′′ = 3 ⋅ 10−6 2
⋅ 400 ⋅ 1500 ⋅ 103 cat = 1800 s−1
s mcat gcat m3cat

1800 s−1 1.5 ⋅ 10−3 (m)
m⋅L= ⋅ = 15
2 ⋅ 10−6 (m2 ∕s) 3
tanh(15)
𝜂= = 0.0667
15
We need to know the amount of catalyst in the reactor bed:
m3bed m3cat kgcat
W = 50 mlbed ⋅ 10−6 ⋅ (1 − 0.6) ⋅ 1500 = 0.03 kgcat = 30 gcat
mlbed m3bed m3cat
Using the data:
m3 m2cat m3
k ′ = 3 ⋅ 10−6 2
⋅ 400 = 12 ⋅ 10−4
s mcat gcat s gcat
Q
W =− ⋅ ln(1 − XAf ) = 30 g
k′ ⋅𝜂
Q (m3 ∕s)
=− ⋅ ln(1 − 0.4)
12 ⋅ 10−4 m3 s−1 g−1 ⋅ 0.0667
Resulting Q = 4.7⋅10−3 m3 /s.
(c) In this case, the lower porosity will produce an increase in the void volume,
but the amount of catalyst will be the same, so the conversion does not vary.
(d) If the particles had a diameter of 0.03 mm:
L = 0.5 ⋅ 10−6 m → m ⋅ L = 0.15 → 𝜂 = 0.992 → 30
4.7 ⋅ 10−3
=− ⋅ ln(1 − XA )
12 ⋅ 10−4 ⋅ 0.992
From that equation, X A = 0.9994

Example 9.7 Reaction with Internal Diffusion Resistance


In the design of the reactor presented in Example 9.1, it has been found that the
internal diffusion stage plays an important role in the reaction rate under certain
conditions.
The system consists of spherical particles 6 mm in diameter with an internal
surface area of 400 m2 /g, in this case presenting reactant A with an effective dif-
fusivity of 2.66⋅10−8 m2 /s.
Calculate the variation of the conversion of A with the length of the reactor for
a total length of 0.1 m under isothermal conditions. To carry out the design, it will
be necessary to determine the variation of the Thiele module and the efficiency
factor of the catalyst along the bed.

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 )

Program Listing 9.3


clear all
clear all

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

De=2.66e-8;% diffusivity, m2/s


dp=6e-3;% diameter of particles, m
XA(1)=0; % Initial value of conversion
L(1)=0; % Initial value of length
incx=0.0001; % Increment of length
n=0.1/incx; % Number of points
CA=Ca0;
mL=(dp/6)*(k.*CA*3/(2*De)).∧ 0.5;
effectivity=tanh(mL)/mL;

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)

9.6 Combination of Resistances


In the general case, all steps in the chemical reaction process can affect the reac-
tion rate, so let us see what happens if there is no step controlling the rate.
The flow of reactant through the external layer could be expressed using
Eq. (9.25):
g
(−rA ) = − A = kC ⋅ (CAg − CAs ) (kmol∕(s m2 )) (9.25)
Sex
But in the general case, C As does not equal zero. At the same time, the diffusion
thought the pores and chemical reaction rate is given by:
(−rA ) = kCAs 𝜂 (9.73)
Obviously, there is no accumulation of reactant species A in the external sur-
face, so these two rates must be equal. Therefore:
(−rA ) = kC ⋅ (CAg − CAs ) = kCAs 𝜂 (9.74)
9.7 Monolithic Catalytic Reactors 237

and doing a combination of resistances, we get


(CAg − CAs ) CAs CAg
(−rA ) = 1
= 1
= 1 1
(9.75)
kC k𝜂 kC
+ k𝜂

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).

9.7 Monolithic Catalytic Reactors


As stated in Chapter 8, many industrial operations take place with catalysts sup-
ported on ceramic materials, or similar. This is a quite different situation com-
pared to those mentioned in previous sections, as the “characteristic length” of
these catalytic structures is not easily visualized.
The cordierite that forms the monolith cannot be used to support the catalysts
due to its low surface ratio and because it is not able to efficiently disperse the
metal particles. In this sense, the first step of the process is to adsorb a thin layer
(called “coating” or “washcoat”) of a stable material of high surface, on which
the metal particles will be dispersed. A diagram of the structure is shown in
Figure 9.8. A good approximation is achieved when half the width of the coating
is used as the characteristic length of the catalyst washcoat.

Example 9.8 Pore Diffusional Resistance For a Monolith Catalyst


Elimination of SO2 by reduction is an important process in industrial systems.
It has been reported as a first-order reaction kinetics, i.e. r = k⋅CSO2 for a 5%
V2 O5 /TiO2 catalyst. The value of the kinetic constant is 0.566 cm3 /(gcat s) at 500 K
and the activation energy found was 110 kJ/mol. Calculate the effectiveness fac-
tor and reaction rate at the inlet of a commercial unit containing an extruded
cellular monolith consisting of 5% V2 O5 /TiO2 with a channel wall thickness of
1.35 mm and open frontal area of 64% under typical commercial operating inlet
conditions of 400 ppm of SO2 , 350 ∘ C and 1 atm. Assume that a catalyst in the
square monolith walls can be treated as a flat plate. Additional data: particle den-
sity = 1.48 g/cm3 ; De = 0.07 cm2 /s. Assume that the film mass transfer resistance
is negligible.

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

Active catalyst component

2·L

Alumina washcoat

Monolith substrate

Figure 9.8 Monolithic catalyst structure.

The equivalent length can be calculated:


Vcat length ⋅ width ⋅ thickness thickness
= =
Scat 2 ⋅ (length ⋅ width) 2
where the factor of 2 accounts for each wall having two sides. The Thiele modulus
is then:
√ ( )
k ′′′′ thickness
Φ=m⋅L= ⋅
De 2

√ ( )
√ 105 cm3 ⋅ 1.48 gcat3
√ 1.35 ⋅ 10−1 cm
= √ gcat s cm
⋅ = 3.18
2
0.07 cm s
2

And the effectiveness factor is 𝜂 = tanh(mL)


mL
= tanh(3.18)
3.18
= 0.31
For calculating the reaction rate:
pSO2 mol
CSO2 = = 4.82 ⋅ 10−9 3
RT cm
mol

(−rSO2 ) = k ′ ⋅ CSO2 ⋅ 𝜂 = 105 ⋅ 4.82 ⋅ 10−9 ⋅ 0.31 = 2.58 ⋅ 10−7
gcat s

Example 9.9 De-NOX via SCR


A monolithic reactor with vanadium oxide catalyst dispersed in a bee-type hon-
eycomb will be designed. The reactor will be used for the elimination (minimum
90%) of the nitrogen oxides of a stream having the following characteristics:
Flow rate = 30 000 Nm3 ∕h
NOx content = 1000 mg∕Nm3 (95 wt%of these are NO, the rest NO2 )
A current of 24 wt% ammonia will be fed to reduce the NOx by the catalytic
reactions:
4NO + 4NH3 + O2 → 4N2 + 6H2 O
2NO2 + 4NH3 + O2 → 3N2 + 6H2 O
As for the kinetics of the reaction, it has been shown to be of order 1 with
respect to the NO component and of zero order with respect to the ammonia
9.7 Monolithic Catalytic Reactors 239

component. The rate constant follows the equation:


ln(k ′′′ ) = (−3266.8∕T) + 8.2208 … with k ′′′ in s−1 and T in K.
Negligible effects of diffusion to the catalyst surface are observed if the surface
rate is greater than 6 m/s. It will be assumed that the reactor operates under an
isothermal regime at a temperature of 623 K. Catalysts supported on panels with
the following characteristics are available:
Dimensions: 0.15 × 0.15 × 1.35 m
Fraction of holes: 0.7
Ceramic honeycomb thickness: 5.5⋅10−4 m
Relative dose of catalyst: 5.88%
Diffusivity of NO in the bed: 2.3⋅10−6 m2 /s
Determine the dimensions of the bed needed to achieve the desired conversion.
Simulate the evolution of the NOx concentration with the length of the reactor.
Calculate the flow of ammonia needed for that conversion.

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

Q·CNOin·(1 – (XNO + dXNO))

nNO + (−r′′′ NO ) ⋅ dV = (nNO + dnNO )


Q ⋅ CNOin ⋅ dXNO = k ′′′ ⋅ CNOin ⋅ (1 − XNO ) ⋅ dV
240 9 Catalytic and Multiphase Reactor Design

1 Figure 9.9 Solution to


Example 9.7.
0.8
Molar conversion

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3
Volume (m3)

Eliminating the concentration and integrating:


XNO
Q
V = dX
∫0 k ′′′ ⋅ 𝜂 ⋅ (1 − XNO ) NO
The value of Q is known, Q = 30 000 Nm3 /h = 62 720 m3 /h (at 623 K) =
17.42 m3 /s
Remember that “Nm3 ” stands for “normal cubic meter”, i.e. measured at 0 ∘ C
and 1 atm. The conversion to the corresponding temperature is done by multi-
plying by T/T normal , in K.
The numerical or analytical integration of the design equation gives the evo-
lution of the conversion with the volume (see Figure 9.9). For a design value of
X NO = 0.9 the needed volume is 2.61 m3 . This represents the volume of the reactor
containing the corresponding catalyst.
The dimension of each cell is 0.15 × 0.15 × 1.35 m = 0.030 375 m3 , but not all
the volume is active, as one of the data given is the fraction of holes. In this way,
the active volume of each cell is 0.030 375*(1 − 0.7) m3 and the number of cells
can be calculated by
1 cell m3
2.61 active m3 ⋅0.7 active m3
Number of cells = = 122.51
0.030 375 cell m3
If cells are
√ disposed in a square, the number of cells in each row or column
would be 122.51 ≈ 11. A scheme can be found in the following figure, and it
contains 121 cells:

11 cells
11 cells

1.35 m
9.7 Monolithic Catalytic Reactors 241

Once we have calculated the needed cell distribution, it is important to check if


external diffusion can play a role in the reaction. For this purpose, we can calcu-
late the rate of the gases through the cells. A flow rate of 17.42 m3 /s in a section
area of 121 × (0.15 × 0.15) = 2.7225 m2 . Nevertheless, only 70% of this area is free
for flowing (fraction of holes), so must have a lineal velocity of
Q 17.42 m
v= = = 9.13
Section ⋅ 0.7 1.905 s
That is enough according to the requirements. Note that the free area is not
calculated subtracting the thickness of the ceramic to the dimension of the cell,
and doing the square, as the honeycomb also can contain the monolith substrate
as shown in Figure 9.8.
Finally, the calculation of the ammonia feed rate is to be calculated. In the
input, we have 31.66 mmol NO/Nm3 and 1.08 mmol NO2 /Nm3 , in the propor-
tions that the statement indicates. For reacting with these amounts, we will need
(31.66 + 2*1.08) = 33.84 mmol NH3 /Nm3 . Having in mind the total flow, this rep-
resents 282 mmol NH3 /s, i.e. 4.794 mg/s. Bearing in mind the 24 wt% concentra-
tion of this solution, the feed flow needed is approximately 20 g/s.

Example 9.10 Oxidation of Ethylene in a Fluidized Bed


A fluidized-bed reactor will be designed for producing 100 kg/h of ethylene oxide,
via the oxidation of ethylene. We will oxidize directly the gas over a silver catalyst.
The following conditions will be used:
✓ Temperature of the bed = 549 K
✓ Pressure of the reactor = 9.65 atm
✓ The velocity of the gases will be eight times umf (minimum velocity of fluidiza-
tion).
✓ The gas entering the reactor has a 6% of oxygen
✓ The conversion of the oxygen must be below 40% in order to minimize side
reactions.
Two parallel reactions occur in the reactor. On the one hand, the oxidation
of ethylene to produce ethylene oxide, and on the other the decomposition of
ethylene into final oxidation products:
kJ
O2 + 2 C2 H4 → 2C2 H4 O ΔH1 = −2.10 ⋅ 105
kmol
1 2 2 kJ
O2 + C2 H4 → CO2 + H2 O ΔH2 = −4.73 ⋅ 105
3 3 3 kmol
Both reactions are first order with respect to the oxygen, and the kinetic con-
stants are
m3 m3
k ′′′ 1 = 0.255 3 fluid k ′′′ 2 = 0.361 3 fluid
mcatalyst s mcatalyst s
The catalyst to be used in the reactor will contain a 5% in weight of silver, and an
alumina monolith will be the main component. The catalyst will be dispersed into
this support in the form of spheres of 75⋅10−6 m of radius. The effective diffusion
coefficient of ethylene in the oxygen is 4.64⋅10−6 m2 /s. The density of the catalyst
is 1800 kg/m3 .
242 9 Catalytic and Multiphase Reactor Design

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

where S1 refers to the selectivity of the reaction 1. As the temperature is constant


in this design, we can check that the selectivity is also constant:
k ′′′ 1 CA k ′′′ 0.255
S1 = ′′′
= ′′′1 = = 0.706
k2 CA k2 0.361
For X Afinal = 0.4, nA0 = nBfinal /(2⋅X Afinal ⋅S1 )=0.6313/(0.8⋅0.706)=1.117 mol/s.
Let us use Matlab to solve this example. A valid program to calculate all needed
is shown (Program listing 9.4).

Program Listing 9.4


%Data
T=549; %K
P=9.65; %atm
Q=100; %kg/h
k1=0.255; %m3fluid/m3cat*s
k2=0.361; %m3fluid/m3cat*s
DH1=-2.1e5; %kJ/kmol
DH2=-4.73e5; %kJ/kmol
R=(75e-6)/2; %m
De= 4.64e-6; %m2/s
denscat=1800; %kg/m3
Cox=0.06;
Rg=0.082/1000; %m3*atm/(K*mol)
MW=44; %g/mol

%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

Wcat(1)=0; % weight of catalyst


nA(1)=nA0;
nB(1)=0;
nC(1)=0;

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

at u0 ≫ umf generally the model with rapid bubbles would be acceptable.


Estimation of the external mass transfer would be necessary too.

9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed)


Industrial application of fluid–fluid reactions is very extended, and almost 90% of
them are catalyzed reactions, where three different phases (two fluids and a solid
catalyst) are present.
Some of the reasons why fluid–fluid reactions are present in the industry are
as follows:
(a) The reaction product of the immiscible reagents is a substance of interest
(production reactions).
(b) To facilitate the separation of a component contained in a fluid (purification
reactions).
(c) Obtain a more favorable distribution of the product in multiple reactions in
a homogeneous phase than could be obtained using a single phase.
There are many examples of applications of this type of reactions in the indus-
try. Among the purification reactions are the elimination of CO2 in mixtures
CO + H2 with potassium carbonate or ethanolamines, and the elimination
of H2 S and CO in cracking hydrocarbon gas reactions by ethanolamines or
sodium hydroxide. Production reactions include the reaction between CO2 and
an aqueous ammonia stream to give ammonium carbonate, air oxidations of
acetaldehydes to give the corresponding acids, chlorination of benzene and other
hydrocarbons, absorption of NO2 in water to give nitric acid, the absorption of
SO3 in H2 SO4 to give oleum, and many others.
Also important is the reactor design for catalysis involving gaseous reactants.
This design requires attention to efficient transfer of reactants from the gas to
the liquid phase; mass transfer rates are increased by maximizing interfacial area
by sparging (creating small bubbles) and increasing the mass transfer coefficient
through higher flow rates and/or stirring, which decreases film thickness.

9.8.1 Transfer Models


For the transfer of component “A” followed by reaction with other species “B,”
there are three resistances in series: that of the gaseous layer, that of the liquid
layer, and that of the bulk of the liquid phase. The reaction, in general, can take
place in the liquid layer or in the middle of the liquid phase. Rapid reactions, for
example, will occur in a narrow area within the liquid layer, and, on the contrary,
slow reactions will take place within the fluid.
When only two fluid phases are present, and there is no need of a catalyst,
different steps must be produced for the reaction to take place. Like most hetero-
geneous reactions, the following steps are present:
1. Diffusion of the reagents towards each other,
2. Chemical reaction itself,
3. Diffusion of the products outside the reaction zone.
246 9 Catalytic and Multiphase Reactor Design

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

Figure 9.10 Double-layer


(two-film) model. INTERFACE

Phase I Phase II

Component A transfer

Component B transfer

Turbulent Laminar flow Laminar flow Turbulent


flow Phase I film Phase II film flow

9.8.2 Two-Film Theory


Next, we will study by applying the two-film theory to some cases that are pre-
sented according to the relative values of the rate of the chemical reaction and the
rate of diffusion. Figure 9.11 shows the most frequent cases, depending on the rel-
ative rate of diffusion and reaction. Case A corresponds to a situation where the
chemical step is much faster than the diffusion, and assumes that the concen-
tration of reactant A in the liquid phase decreases to zero inside the liquid film.
Case B is similar, but the concentration of reactant B is high enough to displace
the reaction plane to the interface. Case C represents a situation where the chem-
ical step rate is similar to that of diffusion, and then widening the reaction zone,
which is not a plane but a 3D zone. Finally, case D represents the situation where
the chemical step is much slower than diffusion, so a constant value of concentra-
tion of reactant A inside the liquid phase can be observed. Note that the reaction
is displaced to the bulk liquid as the chemical step rate decreases.

9.8.2.1 Case A: Instantaneous Reaction


This case represents more than 90% of the industrial cases in which this type of
reactions occurs, since it is usually a diffusion control. The global rate of reaction
is given by the rate of diffusion since this is the slowest step. This means that for
component A (phase I) to react with component B (phase II), first A must diffuse
and find B, so that the reaction plane is located within phase II but very close
to the interface, and between the liquid that contains A and the liquid that con-
tains B. Figure 9.11a shows a diagram of this case, applied in a case of gas–liquid
reaction.
In this case, when varying the concentration of A or the concentration of B, the
plane moves because what is being done is varying the corresponding diffusion
rate. Since the limiting step of rate is diffusion, the rate per unit area of interface
248 9 Catalytic and Multiphase Reactor Design

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

pAi = HA·CAi pAi


(Case A) (Case B)
x
x0
(a) x0 (b)

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

Analogously for B, you get:


DBl
kBl = (9.80)
x0
The mass transport coefficients k Al and k Bl have been defined for transport
without a chemical reaction (therefore, referred to as the flow along the entire film
of thickness x0 ), so that in this case the relationship k Ag (pA − pAi ) = k Al (C Ai − C A )
is fulfilled. When there is a chemical reaction, the expression is modified as pre-
viously described (Eq. (9.78)).
From Eqs. (9.79) and (9.80), it follows that:
kAl D
= Al (9.81)
kBl DBl
Since the values of x, x0 , PAi , and C Ai are not known, an expression of (−rA )
must be found where these variables do not appear. Therefore, it can be written
that:
kAl CAi x0 k C x
(−rA ) = kAg (pA − HA ⋅ CAi ) = = Bl B 0 (9.82)
x b(x0 − x)
Combining the last two terms:
( ( ) )
k
kAl CAi + bBl CB ⋅ x0 ( )
kBl
(−rA ) = = kAl CAi + CB (9.83)
x + x0 − x b
Bearing in mind that the first term can be expressed as
(pA − HA CAi ) (pA kAl ∕HA − KAl CAi )
(−rA ) = 1
= kAl
(9.84)
kAg HA kAg

and combining Eqs. (9.83) and (9.84):


pA kAl kBl CB
HA
+ b
(−rA ) = kAl
(9.85)
1+ HA kAg

Finally, the expression of the rate will be


( ) ( )
pA DBl CB pA k CB
HA
+ DAl
⋅ b HA
+ kBl ⋅ b
Al
(−rA ) = 1 1
= 1
(9.86)
H ⋅k
+k H ⋅k
+ k1
A Ag Al A Ag Al

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

This equation results from imposing the condition k Bl C Bl b ≫ k Ag pA in (9.78),


whereby pAi becomes zero.
A special case that occurs when the resistance of the gas phase is negligible,
which means a coefficient k Ag = ∞ and pA = pAi . The equation obtained in this
case is
( )
pAi D C
+ DBl ⋅ bB ( )
HA DBl CB

Al
(−rA ) = = k C 1 + (9.88)
1
+ 1
Al Ai
b ⋅ DAl CAi
HA ⋅kAg kAl

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.

9.8.2.2 Analysis of the Controlling Steps: The Hatta Modulus


A dimensionless module has been defined to analyze the effect of the diffusion
steps in the reaction. The parameter MH was developed by Shirôji Hatta in 1932,
and serves to deduce when a reaction is slow or fast. For this, only the gas–liquid
interface is considered, and it is assumed that the resistance of the gas phase is
negligible. The conversion parameter in the film has been defined as
maximum transfer in the film
MH2 = (9.92)
maximum transfer by difussion in the film
The expression of MH2 can be transformed into
k ⋅ DAl ⋅ CB
MH2 = 2
(9.93)
kAl
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 251

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.

9.8.2.3 Other Cases in Fluid–Fluid Reactions: The General Rate Equation


As discussed before, there are many possibilities regarding the importance of
the different steps in the whole process. The final form of the rate equation will
depend on the relative values of the rate constants k Ag , k Al , and k, the concentra-
tion of reactants (pA and C B ), and the Henry (or phase equilibrium) constant. For-
tunately, a general expression has been proposed that considers all these aspects:
pA
(−rA′′ ) = 1 HA HA
(9.94)
k ⋅a
+ k ⋅a⋅E
+ (k ⋅C )⋅f
′′
Ag Al B L

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

Reaction occurs Reaction occurs 200


into bulk liquid into liquid film

100 100

50
Ei
E

20

10 10

2
1
0.1 1 10 100 1000
MH

Figure 9.12 General rate equation for different values of MH and E i .

Obviously, f L is not necessary if we know the value of the kinetic constant in


m3L ∕(mol s) or similar.
The absorption of A is greater when there is a reaction, so for the same concen-
trations in two different conditions, the E factor of increase is defined as
[ ]
actual adsorption rate of component A
E=
adsorption rate of A without chemical reaction at the same CAi ,CA ,CBi ,CB
(9.95)
Van Krevelens and Hoftijzer deducted numerically a good correlation for the
coefficient E as a function of MH and Ei :

E −E
MH2 Ei −1
(√ )
i
E= (9.96)
2 Ei −E
tanh MH E −1
i

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.

Example 9.11 Reaction Rate with No Catalyst


CO2 must be separated from air by contact with water in countercurrent at 25 ∘ C.
The individual volumetric coefficients for the transfer of CO2 in the gas and liquid
phases are 80 mol/(h l atm) and 25 h−1 , respectively. The solubility ratio is given
by Henry’s law with a value of H = 30 (atm l)/mol:
(a) Derive an expression for the calculation of the absorption rate.
(b) If a 2 M solution of NaOH is used, the reaction CO2 + OH− is instantaneous
and there is a large excess of NaOH relative to CO2 that can be absorbed,
indicate the most appropriate form to express the reaction rate.
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 253

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

Example 9.12 Mass Transfer with Instantaneous Reaction


A fluid–fluid reaction is developed with two reagents A (initially in phase I) and B
(initially in phase II). The reaction will take place in phase II. The concentrations
of A and B in its phases are 1 kg/m3 , and the stoichiometry is 1 mol of A reacts with
2 mol of B and takes place instantly in phase II. The diffusivities of both reagents
in phase II can be considered to be similar. Deduct the controlling stages of the
process in the following cases:

Case (A) k Ag (through phase I): 20 m/s


H A = 3 m3 phase II/m3 phase I
k Al (through phase II) = 5 m/s
Case (B) k Ag (through phase I): 20 m/s
H A = 10 m3 phase II/m3 phase I
k Al (through phase II) = 500 m/s

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

But we have two liquid phases; therefore:


CA, phase I
(−rA′′ ) = 1 HA HA
kA, phase I ⋅a
+ kA,phase II ⋅a⋅E
+ (k ′′ ⋅CB )⋅fphase II

In this case, the chemical reaction is instantaneous, so:


CA, phase I
(−rA′′ ) = 1 HA
k ⋅a
+k ⋅a⋅E
+0
A, phase I A,phase II

and we should compare the two terms in the denominator.


(a) As the reaction is instantaneous, E ≅ Ei , and then:
( )
DBl ⋅ CB,phase II
E = 1+
b ⋅ DAl ⋅ CAi, phase II
The concentration of “A” at the interface in phase II can be calculated using
the value of H A :
(′ A′ transfered from I to interface) = (′ A′ transfered from interface to II)

kA,phase I (CA,phase I − HA ⋅ CAi,phase II ) = kA,phase II (CAi,phase II − 0)


From this expression, we can calculate C Ai, phase II = 0.308 kg/m3 and then
E = … = 2.62
The values of the terms in the denominator are then:
1 1 s
= = 0.05
kAg ⋅ a 20 m
HA 3 s
= = 0.22
kAl ⋅ a ⋅ E 5 ⋅ 2.62 m
So, in this case the controlling step is the diffusion through phase II.
(b) In this case also, E ≅ Ei , but the value results to be 18.48. When comparing
1
k ⋅a
= 0.05 ms and:
Ag

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.

Example 9.13 Membrane Reactor


In a 1 m3 membrane reactor, the transport of oxygen from a gaseous (air) to a liq-
uid phase is taking place at 1 atm. To increase the transfer rate, an organometallic
molecule (RH) is added to the liquid, with a concentration of 5 mol/m3 . In a sim-
plified way, we have
O2 + RH → (RH ⋅ O2 ) r = k ⋅ CO2 ⋅ CRH
Estimate the rate at which oxygen is consumed in this system, referred to its
volume (mol O2 s−1 m−3 reactor ) (Figure 9.13).
9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 255

Membrane reactor RH·O2

Air, po2 = 0.21 atm

Liquid Air, po2 < 0.21 atm


RH

Figure 9.13 Membrane reactor.

Data:

HO2 = 0.937 atm m3 /mol DO2 ,L = 7.1⋅10−6 m2 /s


2
kO2 ,g = 0.03 mol/(s m atm) DRH,L = 8.3⋅10−6 m2 /s
′′ 3 3
k = 1.8⋅10 m /(mol s) Total membrane thickness = 5⋅10−3 m
Membrane surface = 7 m2

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.

9.8.3 Gas–Liquid Reactions in Solid Catalysts. General Equation


As mentioned, many fluid–fluid reactions take place over a solid catalyst. These
multiphase reactions are of the type:
disolution in the catalyst
A (g −−−−−−−→ l) + bB (l) −−−−−−−−−→ products (9.97)
The gaseous reagent must first be dissolved in the liquid, then the reagent must
be diffused or moved to the surface of the catalyst for the reaction to take place.
Thus, in the expression of rate, the resistance to transfer through the gas–liquid
interface and toward the surface of the solid intervenes.
These types of reactions can be carried out in different ways. Packed contact
devices use large catalytic solid particles, sludge reactors (or suspended solids,
also called slurry type) use very fine suspended catalysts, while fluidized beds
use both sizes, depending on the flow rates.
For the chemical reaction step, the following reaction rates will be considered:
⟩ r ′′′
−rA′′′ = kA′′′ CA CB −rA′′′ = − Bb mol A
m3cat s
where k ′′′
(9.98)
−rB′′′ = kB′′′ CA CB k ′′′ = B mol3 B
A b mcat s

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

Volume element containing


liquid (L), solid (S), and gas (G)

HA = pA/CA, Henry’s law constant


V = reactor volume
ai = (surface of the interface gas–liquid interface)/V
ac = (external surface of the particles)/V
fs = (particle volume)/V, i.e. solid fraction
fl = (liquid volume)/V, i.e. liquid fraction

Figure 9.14 Nomenclature used.


9.8 Fluid–Fluid Reactions (Catalyzed and Noncatalyzed) 257

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

kg = mass transfer coefficient in the gas film (mol/(Pa m2 s))


kL = mass transfer coefficient in the liquid film (m3L/(m2 s))
kC = mass transfer coefficient within the solid film (m3L/(m2cat s))
DAe, DBe = effective diffusion coefficients (m3L/(mcat s))

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

Combining these quotients, the following general expression of rate can be


written, with reference to the reactor volume:
pAg
(−rA′′ ) = 1 HA HA HA
(9.100)
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

where 𝜂 A and 𝜂 B are efficiency factors.


Equation (9.100) or Eq. (9.101) would give the reaction rate. Both expressions
(−r′′ )
are related by (−rA′′ ) = bB and the terms in parentheses express the rate constant
of pseudo first order. Unfortunately, even with all known system parameters (k, a,
f , etc.), these expressions cannot be solved without a trial and error procedure,
because C B in Eq. (9.100) and C A in Eq. (9.101) are not known. However, generally
one works in one of two extreme cases in which simplifications can be made,
which are very useful.
258 9 Catalytic and Multiphase Reactor Design

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

9.8.3.1 Estimation of the Controlling Resistance in Multiphase Systems


Let us work a little more with the rate equation. From Eq. (9.99), you can easily
reach:
pAg 1 HA HA HA
= + + + (9.102)
′′
−rA kAg ai kAl ⋅ ai kAc ⋅ ac (k ′′′ ⋅ C ) ⋅ 𝜂 ⋅ f
A B A s

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

• A zero slope in this representation (RC + RL ) vs. dp , would indicate a process


controlled by the superficial chemical reaction.
• A slope 1 indicates limited reaction by internal diffusion.
• A slope of 1.5–2.0 indicates a limitation due to external diffusion.

Example 9.14 Catalyst in a Gas–Liquid Reaction


In a 1 m3 reactor, a gas–liquid reaction is carried out on a solid catalyst. The
reagent of the liquid phase is in excess of that of the gas. The following data is
available:
Mass of Partial pressure Reaction
catalyst (g) of the gas (atm) rate (mol/(h m3 R ))
1 0.1 1
2 0.15 2
3 0.2 3
4 0.2 4

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

0.10 y = 0.06x + 0.0417


–pA/rA ((atm h m3R)/mol).

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

chemical reaction. If we also had data at different particle diameters, we would


distinghish bewteen these two contributions, in the way that representing
(RS + RR ) vs. dp we will obtain a zero slope if the chemical reaction is the
controlling step, and different values for internal and external diffusion control.

9.8.3.2 General n-th Order Kinetics


The rate expressions used so far have been developed assuming a first-order reac-
tion with respect to A and first order with respect to B. But, how are more general
forms of rate treated? For example:
in the catalyst
A(g) + bB(l) −−−−−−−−−→ products (−rA′′ ) = (−rB′′ )∕b = kA′′ CAn CBm
To be able to combine the chemical step with the steps of transfer of mass in a
simple way, we must replace the uncomfortable previous expression of rate with
a first-order approximation, in the following way:
n−1 m
(−rA′′ ) = (−rB′′ )∕b = kA′′ CAn CBm ⇒ (−rA′′ ) = (kA′′ C A C B )CA
with the average values calculated in the places where the reaction occurs.
This approach is not completely satisfactory, but it is the best thing to do. Thus,
instead of the general rate equation, the rate equation to be used in all design
expressions will be
pAg
(−rA′′ ) = 1 HA HA H
(9.100b)
k ⋅a
+ k ⋅a + k ⋅a + ′′′ n−1 Am
Ag i Al i Ac c (kA C A C B )⋅𝜂A ⋅fs

Example 9.15 n-th Order Reaction Rate


A solution of acetone (C B0 = 1 M) with a flow rate of 1 l/s, and pure hydrogen
at a total pressure of 1 atm and a flow rate of 0.04 m3 /s are fed by the base of
a column of 0.1 m2 of section, packed with porous nickel catalyst (dp = 5 mm,
density = 4500 kg/m3 , load of solid = 0.6 m3 cat/m3 reactor, effective diffusiv-
ity = 8⋅10−10 m3 L /(mcat s)), and maintained at 14 ∘ C. Under these conditions,
acetone reacts with hydrogen to give propanol:
in the catalyst
H2 (g → l) + CH3 COOCH3 (l) −−−−−−−−−→ CH3 CHOHCH3
If H2 is designed by “A” and acetone by “B,” the rate is given by
( ) ( )1∕2
3
mol 1∕2 m L mol
−rA = k ′ CA CB0 and k ′ = 2.351 ⋅ 10−3
kgcat s kgcat s m3L
What will be rate of that process?
Data:
HA = 36 845 Pa m3 L ∕mol kAc ⋅ ac = 0.05 m3 L ∕(m3 R s)
( )
1 1 1
+ = (s ⋅ mL 3 )∕mR 3
HA ⋅ kAg ⋅ ai kAl ⋅ ai 0.02
9.9 Design of Multiphase Reactors 261

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

and consider the term in parenthesis a pseudo first-order kinetic constant.


pAg
(−rA′′ ) = 1 HA HA H
k ⋅a
+ k ⋅a + k ⋅a + ( − 1 A)
Ag i Al i Ac c k⋅CA 2 ⋅𝜂A ⋅fs

Also, for n-th order kinetics:


[ ]1
k ⋅ CAs
n−1
(n + 1) 2
m⋅L=L⋅
2 ⋅ De
( ) 1

⎡ 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

9.9 Design of Multiphase Reactors


In this section, we describe several types of multiphase reactors and we see the
basic equations for their design.

9.9.1 Types of Flow in Multiphase Reactors


The reactions between the components of a gas and a liquid are carried out in a
great variety of equipment, which often have a confusing name. This wide variety
of systems must fulfill some basic conditions: efficient contact between gas and
liquid (and, in many cases, a solid catalyst), limitation in pressure drop, easy heat
extraction, and low construction and maintenance costs. Depending on whether
the greater resistance to the transfer of mass occurs in the liquid or in the gas,
the multiphase reactors operate with the dispersed liquid phase and the continu-
ous gas, or vice versa. The choice of parallel or countercurrent circulation of gas
and liquid depends on the availability of driving forces for mass and heat transfers
and reaction.
262 9 Catalytic and Multiphase Reactor Design

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.

9.9.2 Design Models for Flow in Multiphase Reactors


This section shows some general models that can be used for the design of any
of these reactors. The models are based on the plug flow or stirred tank of one or
both phases. A distinction is made between the processes taking place without
the presence of a reaction (absorption processes) and those where a reactant is
present in each phase.
In the reasonings that follow, the following reaction scheme is assumed:
A (gas) + bB (liquid) → products (R or S)
where component A must diffuse to the phase where B is present, and react.
Before continuing, the nomenclature to be followed is presented.
a = Area of interface per unit volume of the tower
f = Fraction in volume of the phase in which the reaction takes place
U = Carrier or component inert in a phase, therefore it is neither reactant, nor
product, nor component that diffuses
T = Subscript referred to total moles in the liquid phase
p moles of A
YA = p A = moles of inerts
in the gas
U
C moles of A
XA = C A = moles of inerts
in the liquid
U
′ ′
G , L = Molar flow of inert gas and liquid, usually per unit of normal section to
the tower (kmol/(s m2 )).
PT = total pressure
C T = total concentration
Using this nomenclature (see also Figure 9.16), it is fulfilled that:
p
G′ = G ⋅ U ; G′ ⋅ PT = G ⋅ pU (9.106)
PT
C
L′ = L ⋅ U ; L′ ⋅ CT = L ⋅ CU (9.107)
CT
PT = pA + pB + · · · + pU (9.108)
CT = CA + CB + · · · + CU (9.109)
( )
pA p dp − p dp dp
dYA = d = U A 2 A U ≈ A (9.110)
pU pU pU
( )
CA CU dCA − CA dCU dCA
dXA = d = 2
≈ (9.111)
CU CU CU
The amount of component A accompanying the gas for a given composition, in
mol/(s m2 ), can be calculated by G′ ⋅Y A or using the relation G⋅pA /PT .
9.9 Design of Multiphase Reactors 263

Figure 9.16 Scheme of a multiphase system with the Liquid (L′)


nomenclature used. CA1 or CAin
pA1 or pAout
CB1 or CBin
Section 1

Section 2

pA2 or pAin CA2 or CAout


Gas (G′) CB2 or CBout

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.

9.9.3 Case 1: Pure Absorption (No Reaction)


9.9.3.1 Situation 1: Gas and Liquid Phases in Plug Flow
To develop the design equation, the rate equation will be combined with the
material balance. Thus, for the steady state in countercurrent operations like the
one in Figure 9.16, we have a volume differential element:
(A lost by gas) = (A gained by the liquid) = (A transferred)
= (−r′′ A ) dV (rA ′′ ) in kmol∕(m3 equipment s) (9.112)

(kmol∕s) G′ dY A = L′ dX A = (−rA′′ )dV (9.113)


The first two terms can be expressed in different ways:
G′ dY A = L′ dX A
dpA dC
G′ = L′ A (9.114)
pU CU
and the rate of “reaction,” if there is only mass transfer, is
(−rA′′ ) = (−rA ) ⋅ a = kAg ⋅ a ⋅ (pA − pAi ) = kAl ⋅ a ⋅ (CAi − CA ) (9.115)
2
(−rA ) being in kmol/(m i s). In this way, the balance in a differential volume is
G′ dY A = L′ dX A = (−rA ) ⋅ a ⋅ dV
Integrating for the whole tower gives
YA2 XA2
G′ dYA L′ dXA
V = = (9.116)
a ∫YA1 (−rA ) a ∫XA1 (−rA )
264 9 Catalytic and Multiphase Reactor Design

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 )
∗ ∗

where pressures and concentrations refer to equilibrium, that is


p∗A = HA ⋅ CA
CA∗ = pA ∕HA

9.9.3.2 Situation 2: Gas and Liquid Phases Completely Mixed


This is the simplest situation from the point of view of the calculation, since the
concentrations of A and B are uniform and differential equations are not involved,
as there is only one value of the reaction rate throughout the reactor. The conti-
nuity equation for A (molar balance of A), for the whole volume can be written as
(A lost by gas) = (A gained by the liquid) = (A transferred) = (−r′′ A ) V
(9.112)
Note that the previous equation is the same as Eq. (9.112), but in this case we
can obviate the differential volume as there is only one value of the reaction rate
in the whole equipment.
Considering diluted systems, it can be written that:
G′ L′
(pA2 − pA1 ) = (C − CA1 ) = (−rA2′′
)⋅V (9.119)
pT CT A2
′′
(−rA2 ) being the reaction rate in the final conditions of the reaction, which are
the conditions present at the reactor. In this equation, (−rA2′′
) = (−rA2 ) ⋅ a.

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.

9.9.4 Case 2: Systems with Reaction


In this section, it is assumed that the reaction that is being produced is
A(G → L) + bB(L) → products (L)
We will assume that this reaction is fast enough so that there is no unreacted
A within the liquid phase or the quantity is negligible for the mass balance.

9.9.4.1 Situation 1: Gas and Liquid Phases in Plug Flow


The design equation is obtained by combining the material balance with the rate
equation. Thus, for a differential volume element:
(A lost by the gas) = (1∕b) (B lost by the liquid)
= (A disappeared by reaction)
which is the same as
L′ dXB
G′ dYA = − = (−rA′′ ) ⋅ dV (9.125)
b
( ) ( )
pA L′ CB
G′ d =− d = (−rA ) ⋅ a ⋅ dV (9.126)
pU b CU
The value of the reaction rate can be calculated from the equations studied in
the previous sections, i.e. the general equation would be used for fluid–fluid reac-
tions (refer Section 9.8.2.3), and in particular systems it could be simplified. In
the case of a fluid–fluid reaction taking place in a solid catalyst, the rate reaction
equation would consider the stages mentioned in Section 9.8.3.
If the first and third terms of Eq. (9.126) are ordered and integrated:
YA2
G′ dYA
V = (9.127)
a ∫YA1 (−rA )
266 9 Catalytic and Multiphase Reactor Design

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

Example 9.17 CO2 Elimination


CO2 will be removed from an air stream at 1 atm by absorption in a 0.5 M solu-
tion of NaOH at 20 ∘ C in a 1 m diameter column packed with 25 mm rings. The
air enters at a feed rate of 0.02 kmol/s and contains 0.1 mol% CO2 that must be
reduced to 0.005 mol% at the exit. NaOH is supplied to the column at such speed
that its concentration does not vary appreciably when going through the reactor.
Calculate the height of the packed required. Is a packed bed the best system for
this operation? Data:
✓ Second-order speed constant for the reaction CO2 + OH → HCO− 3
k 2 = 9.5⋅103 m3 /(kmol s). In a NaOH solution, this reaction is followed
by the instantaneous reaction HCO3 − + OH− → CO3 2− + H2 O, leading to a
global reaction: CO2 + 2NaOH → Na2 CO3 + H2 O
✓ For 0.5 M NaOH at 20 ∘ C: CO2 diffusivity: 1.8⋅10−9 m2 /s. NaOH diffusiv-
ity = 1.7 times that of CO2 .
✓ CO2 solubility: Henry’s law constant = 25 bar m3 /kmol
✓ Effective interfacial area of the packing: 280 m2 i /m3 R
✓ Coefficients of transfer of mass in the film: Liquid: k Al = 1.2⋅10−4 m/s; Gas:
k AG = 0.056 kmol/(m3 s bar)

Solution

0.005% mol CO2


0.5 M NaOH

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

Solving for E = 24.37.


The value of the reaction rate can now be calculated, using Eq. (9.94) and
results:
pA kmol
(−rA′′ ) = 1 HA HA
= … = 0.0206 ⋅ pA 3
+ k ⋅a⋅E + (k⋅C )⋅𝜀 m s
k ⋅a Ag Al B L

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.

9.9.4.2 Situation 2. Gas and Liquid as Mixed Flow


Since the composition is the same at any point of the tower/reactor, only one value
of the reaction rate is found, and the balance can be made in the whole system:
(kmol A lost by gas) = (1∕b) (kmol B lost by the liquid) = (kmol A missing)
9.9 Design of Multiphase Reactors 269

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)).

9.9.4.3 Situation 3. Gas in Plug Flow. Liquid Mixed Flow


As in the previous case with no reaction, there exist in the system several values
of reaction rate, as one of the phases is in plug flow. This indicates that differential
balances are needed, although the conditions of the liquid phase will be the same
in all points.
In this way, in a differential volume of gas:
(kmol A lost by gas) = (kmol A missing by reaction)

G′ dYA = −rA′′ ]exit conditions ⋅ dV (9.131)


From a balance around the reactor, you get:
L′
G′ (YAin − YAout ) =
(X − XBout ) (9.132)
b Bin
Integrating Eq. (9.131) and using Eq. (9.132) is
YAin
G′ dYA
V = (9.133)
a ∫YAout (−rA )
For diluted systems, these equations can be simplified in the usual way. The
value of (−rA ) is relatable to pA .

9.9.5 Case 3. Multiphase Reactors


Until this point, we have discussed the form of the kinetic treatment of reactors
in which a reaction occurs between a liquid and a gas with no presence of a solid
catalyst. When treating with multiphase systems, where a component initially in
the gas phase is transferred to the liquid, and then reacts over the surface of a
catalyst, the mass and rate balance equations do not vary, not depending on the
type of reaction being designed. In this sense, Eqs (9.125–9.133) are valid for the
design of fluid–fluid catalyst reactors.
Nevertheless, the reaction rate expressions must consider the presence of the
catalysts, and then the general equation studied in Section 9.8.3 is used. On many
occasions, these equations are only valid in the cases of excess of A or excess of
B, so this should be approximately met in the system to be designed.
Figure 9.17 shows a scheme of the different cases and situations, which can be
useful for the design of multiphase reactors.

Example 9.18 Ammonia Absorption


In a packed column for the absorption of ammonia with sulfuric acid, the ammo-
nia partial pressure at the inlet is 0.05 atm in a gas that enters at atmospheric
270 9 Catalytic and Multiphase Reactor Design

Multiphase reactor design

(Case 1). Pure adsorption (Case 2). Reactive systems

(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

Liquid (L′) Mole balance: Liquid (L′) Mole balance:


L′
G′(YAin – YAout) = L′(XAout – XAin) G′(YA2 – YA1) = (X – XB2)
pA1 or pAout CA1 or CAin pA1 or pAout CB1 or CBin
b B1
Adsorption rate: Reaction rate:
(–r″A) = kAG.(pA–pAi) ·a = kAL·(CAi–CA) ·a = (–r″A) = General equations (with/
pA2 or pAin CA2 or CAout = kAG·(PA–PA*) ·a = KAL·(CA*–CA) ·a pA2 or pAin CB2 or CBout without) catalyst

Gas (G′) Gas (G′)


L′
(Situation 1). G and L in PF G′dYA = L′dXA = (–r″A)·dV (Situation 1). G and L in PF G′ dYA = dXB = (–r″A)·dV
–b
(Situation 2). G and L mixed flow (Situation 2). G and L mixed flow

G′(YA2 – YA1) = L′(XA2 – XA1) = –r″A2 ·V ( ) G′(YA2 – YA1) =


L′
(X – XB1) = –r″A2 ·V
b B2
( )
(Situation 3). G in PF, L mixed flow (Situation 3). G in PF, L mixed flow

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

PA2 = 0.05 atm


CB2 ?
G = 45 kmol/h
9.9 Design of Multiphase Reactors 271

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:

G′ (YA2 − YA1 ) = (G2 yA2 − G1 yA1 ) ≈ G(yA2 − yA1 )


and, similarly, for the liquid:
L′ L
(X − XB2 ) ≈ (CB1 − CB2 )
b B1 b
In the system, we have b = 0.5 and then we can calculate the concentration at
the exit:
L
G(yA2 − yA1 ) = (C − CB2 )
b B1
3

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

As the chemical step is instantaneous (k → infinity), the reaction will be


governed(by the diffusion
) of the reactants. When the reaction is instantaneous,
DBl ⋅CB
E = Ei = 1 + b⋅D ⋅C , and then previous equation can be transformed into:
Al A

pA
(−rA′′ ) = 1 H
+ ( A )
D ⋅C
kAg ⋅a kAl ⋅a⋅ 1+ b⋅DBl ⋅CB
Al A

This equation is expressed in terms of the volume of the reactor. Instead of


this, the one needed in the ( present) case is that based on the interface surface, i.e.
m2cat
(−rA ) skmol
m2
= (−r ′′ kmol
)
A s m3
⋅ a m3
, and then
cat R R

pA
(−rA ) = 1 HA
+ ( )
D ⋅C
kAg kAl ⋅ 1+ b⋅DBl ⋅CB
Al A
272 9 Catalytic and Multiphase Reactor Design

Using the data given:


pA
(−rA ) = 1 0.0033
+ ( )
0.35 C ⋅0.0033
0.05⋅ 1+ B0.5⋅p
A

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.

yA C B (kmol/m3 ) (−rA ) (kmol/(s m2 i )) G′ /(−rA )

0.01 0.6 0.003 13069.9


0.015 0.5875 0.005 8728.8
0.02 0.575 0.007 6553.4
0.025 0.5625 0.009 5246.3
0.03 0.55 0.010 4374.0
0.035 0.5375 0.012 3750.5
0.04 0.525 0.014 3282.6
0.045 0.5125 0.015 2918.5
0.05 0.5 0.017 2627.2

Finally, Si = 215.5 m2 .

Example 9.19 Trickle-Bed Reactor


A trickle-bed-type reactor will be designed to eliminate 75% of the thiophene
present in a hydrocarbon mixture containing 0.012 kmol/m3 of thiophene. The
bed of catalyst particles will operate at a temperature of 200 ∘ C and at a pressure
of 40 bar. Under these conditions, it can be assumed that the reaction rate is first
order with respect to dissolved hydrogen and independent of the concentration
of thiophene. The superficial velocity of the liquid in the reactor will be 0.05 m/s.
Determine the height of the bed needed to provide a thiophene conversion
of 0.75. The necessary data are given here:
* Reaction rate constant ... k = 1.1⋅10−4 m3 cat /(kgcat s)
* Density of the catalyst ... 𝜌cat = 960 kg/m3
* Effectiveness factor = 0.98
* Volumetric transfer coefficients for hydrogen:
in the liquid film: k Al ⋅ ai = 0.030 s−1
in the solid film: k C ⋅ aC = 0.5 s−1
* Value of Henry’s coefficient: H A = 1940 bar m3 /kmol
9.9 Design of Multiphase Reactors 273

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

The reaction rate is then:


pAg
(−rA′′ ) = = 1.14 ⋅ 10−5 ⋅ pAg
87 293
with the rate in kmol/(s m3 R ) if the pressure is in bar.
In this example, the gas is pure hydrogen, so its partial pressure equals the total
pressure, and the reaction rate is constant in this particular case:
kmol
(−rA′′ ) = 1.14 ⋅ 10−5 ⋅ 40 = 4.58 ⋅ 10−4
s m3R
For a conversion of 0.75, we have
CBout = 0.012 kmol∕m3 ⋅ (1 − 0.75) = 0.003 kmol∕m3
And we can write
L′ ⋅ CBin − L′ ⋅ CBout = (−r′′ A ) ⋅ V
L′ ⋅ 0.012 − L′ ⋅ 0.003 = 4.58 ⋅ 10−4 ⋅ V i.e.V ∕L′ = 19.56 s
As the lineal velocity is 0.05 m/s, we can write
Height = 0.05 m∕s 19.56 s = 0.978 m

Example 9.20 Sulfur Oxide Removal


A column with a packing of active carbon that constitutes a bed is placed in the
laboratory. The column is used for the removal of a stream of SO2 with a content
274 9 Catalytic and Multiphase Reactor Design

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

For the SO2 (A), the rate of the process would be


(−rA ) = −kAG (pAg − pAi ) = k ⋅ CO2 s ⋅ 𝜂
where the first equality is giving the rate of the diffusion through the gas–solid
interface, and the second refers to the reaction + pore diffusion, and so the
effectiveness factor appears. This effectiveness factor is given by the ability
of the oxygen (species appearing in the chemical reaction rate) to diffuse
through the catalyst particles.
For the SO3 (C), we can write
(−rC ) = kCl (CCs − CCc ) = k ⋅ CO2 s ⋅ 𝜂
So, globally:
(−rA ) = kAG (pAg − pAi ) = k ⋅ CO2 s ⋅ 𝜂 = −kCl (CCs − CCc )
(b) Qualitatively, the evolution of the concentrations in the reactor would be
those shown in the next scheme.

SO2

O2

H2SO4
SO3

Height

(c) A mass balance in a volume differential can be written as


(O2 lost by the gas) = (−r′′ A ) ⋅ dV
G ⋅ CO2 in ⋅ dX O2 = (−rA ) ⋅ dV

Example 9.21 Reaction in a Reactor of Known Volume


In a reactor of known volume, a gas–liquid reaction is carried out on a solid
catalyst. In the reactor, the gas and liquid phases behave according to the com-
plete mixing pattern. If you know the quantity and composition of the liquid
foods (C Ain , C Bin , L′ ) and gas (pAin , G′ , pBin = 0), propose a method to deter-
mine the output compositions. What happens if the two fluids follow the piston
flow pattern?

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

In this case, we know the values of G′ , L′ , Y Ain (that is proportional to C Ain ),


and V . We can calculate the values of the exit conditions following the next cal-
culation:

Calculate XBout using terms (I)


Assume a value of YAout
and (II) of the mass balance

Calculate (–rA) at the values of YAout and XBout

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 .

Example 9.22 Fluid–Fluid Reaction Over Catalyst


In a certain chemical plant, a reversible isomerization in fluid phase is carried out
reversible
A ←−→ B
on a solid catalyst in a packed-bed tubular reactor. If the reaction is so rapid that
the mass transfer between the surface of the catalyst and the bulk fluid limits the
9.9 Design of Multiphase Reactors 277

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

In this reversible system, we can write


(−r′′ A ) = k1 CAs − k2 CBs
where the concentrations are evaluated at the surface of the catalyst.
Using the relationship K = k 1 /k 2 , we have
( )
C
(−r′′ A ) = k1 CAs − Bs
K
On the other hand, the mass transfer will have the following expressions:
Moles of A transferred
For the component A → = kA (CA − CAs )
time ⋅ surface
Moles of B transferred
For the component B → = kB (CBs − CB )
time ⋅ surface
From all three equations, which must be equal, we can eliminate the unknowns
C As and C Bs , and then:
[( ) ]
1 1 1 C
′′
(−r A ) + + = CA − B
kA k1 K ⋅ kB K
As the system is limited by mass transfer, we can do k 1 ≫ k A , and then:
[ ]
1 1 C
′′
(−r A ) + = CA − B
kA K ⋅ kB K
So, finally:
( )
CB
kB CA − K
(−r′′ A ) = kB
1
K
+ kA
278 9 Catalytic and Multiphase Reactor Design

Example 9.23 Catalytic Reaction with Expansion of the Gas


In a catalytic system, a cracking reaction is taking place:
A → 4R
With the conditions indicated in the figure, calculate the weight of the catalyst
needed for the conversion of the reactant.
First-order kinetic constant: k′ = 0.1 l/h
Isobaric
Isothermal
CA0 No diluent in the feed
Q0 = 20 000 l/h Assume plug flog

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.

Example 9.24 Second-Order Isomerization


An isomerization and second-order reaction proceeds at atmospheric pressure
in a packed bed with spherical alumina pellets containing metal oxides deposited
by impregnation. The process is carried out in the gas phase, at atmospheric
pressure, in the absence of inert (only the two isomers) at 600 K. The effective
diffusivity of a compound into the other isomer is 2.98⋅10−7 m2 /s, and the
9.9 Design of Multiphase Reactors 279

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.

Example 9.25 Isomerization in a Fluidized Bed Reactor


An isomerization reaction is going to take place in a fluidized-bed-type reactor.
In the absence of diffusional effects, it has been deduced that the rate constant
of the process is 0.15 (kmol A/(kgcat h))/((kmol A)/m3 gas ). The catalyst density is
12 g/cm3 . The flow rate of the fluid is 10 m3 /h. Calculate the following:
(a) In the absence of diffusional effects, both external and internal, the mass of
catalyst needed to obtain a reagent conversion of 90%, for different cases of
initial composition of A in the feed mixture: 10%, 20%, and 30% molar.
(b) If the fluidization was not very effective, and the individual external transfer
coefficient is estimated to be 5⋅10−4 (kmol A/(h m2 cat )/(kmol A/m3 fluid ), cal-
culate the mass of catalyst of diameter equal to 0.1 mm if you want to reach
the 90% conversion, for different cases of initial composition of A in the feed
mixture: 10%, 20%, and 30% molar.
(c) With a catalyst of radius 10 mm and being the effective diffusivity of A of
2⋅10−6 m2 /s, calculate the mass of catalyst necessary to achieve the degree of
conversion of 90%, for different cases of initial composition of A in the feed
mixture: 10%, 20%, and 30% molar.
280 9 Catalytic and Multiphase Reactor Design

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

The mass balance is similar to the former question, so:


( )
XA
dXA kc ⋅ ac
= ⋅W
∫0 (1 − XA ) Q
and doing the corresponding calculations, we have W = 921 kg regardless of
the concentration of reagent.
(c) In that case, we probably will have problems in the internal diffusion. In this
situation, the calculation of the effective factor is needed:
[ ′′′ ] 1 ( )[ ]1
tanh(mL) k 2
1 ⋅ 10−2 0.15 2
𝜂= with m ⋅ L = L ⋅ = = 0.9128
mL De 3 2 ⋅ 10−6
The efficiency factor is then 𝜂 = 0.7914
And the mass of catalyst needed can be calculated as
XA ( )
dXA k⋅𝜂
= ⋅W
∫0 (1 − XA ) Q
W = 153.31/0.7914 = 193.7 kg.
9.9 Design of Multiphase Reactors 281

Example 9.26 Sucrose Acid Treatment


The following first-order reaction has been studied:
C12 H22 O11 (sucrose) + H2 O (acid medium)
⇒ C6 H12 O6 (glucose) + C6 H12 O6 (fructose)
From the experimental results obtained with spherical particles of resin
at 50 ∘ C, it was deduced that the effective diffusivity was on the order of
1.8⋅10−11 m2 /s and the first-order kinetic constant is 0.0193 (kmol sucrose/
(s m3 particle ))/((kmol sucrose/(m3 disol )). Calculate: (a) the critical size from which
the internal diffusion begins to have an effect and (b) the critical size from which
internal diffusion plays a dominant role.

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

Example 9.27 Chlorine Removal


A chemical plant is removing traces of chlorine from a waste gas stream by pass-
ing it through a solid granular absorbent in a packed tubular bed. A conversion of
63.2% is currently being achieved, but it is believed that a higher removal could be
achieved if the flow were increased by a factor of 4, the diameter of the particles
was reduced by a factor of 3, and the length of the tube packed increased by 50%.
What percentage of chlorine would be eliminated with the proposed scheme?
Consider that the chlorine transferred to the absorbent is completely eliminated
by a practically instantaneous chemical reaction. Consider that the filling are per-
fect spheres. In the bed, you can apply the Thoenes and Kramers equation:
( ) ( ) ( )1∕3
kc ⋅ dp 𝜌 ⋅ u ⋅ dp 1∕2 𝜇
= ⋅
DAB (1 − 𝜀) 𝜇 (1 − 𝜀) 𝜌 DAB
Solution
Solving for k C :
[( )0.5 ( )0.33 ] ( )0.5
𝜌 𝜇 u
kc = ⋅ ⋅ DAB (1 − 𝜀) ⋅
𝜇 ⋅ (1 − 𝜀) 𝜌 ⋅ DAB dp
( )0.5
u
The term in the square bracket is a constant, so k C is proportional to dp
, in
such a way that:
( )0.5
u
kC = (constant) ⋅
dp
In this way, if we call k C1 and k C2 the values of k C corresponding to the first and
second situation, respectively, we can write
( )
u1 0.5
kC1 = (constant) ⋅
dp1
( )
4u1 0.5 √
kC2 = (constant) ⋅ d = 12 ⋅ kC1
p1∕3

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:

nA0 ⋅ dXA = kC ⋅ (CA0 (1 − XA )) ⋅ 𝜌cat ⋅ Sbed ⋅ (1 − 𝜙) ⋅ dL


XA
nA0 ⋅ dXA XA
u0 ⋅ Sbed ⋅ dXA
= = kC ⋅ 𝜌cat ⋅ Sbed ⋅ (1 − 𝜀) ⋅ L
∫0 (CA0 (1 − XA )) ∫ 0 (1 − XA )
Finally, we have
( )
1 k ⋅ 𝜌 ⋅ (1 − 𝜙) ⋅ L
ln = C cat
1 − XA u0
Comparing both situations:
( )
1 kC1 ⋅𝜌cat ⋅(1−𝜙)⋅L1 kC1 ⋅L1 L1
ln 1−X u01 u01 u01
( ) = k ⋅𝜌 ⋅(1−𝜙)⋅L =
A1
√ = √ = 0.7698
1 C2 cat 2 12⋅kC1 ⋅L2 12⋅1.5⋅L1
ln 1−X u02
A2 u02 4⋅u01

Taking the given value X A1 = 0.632 we can calculate: X A2 = 0.7129.

Example 9.28 Fluid–Fluid Reaction in a Tower


Absorption of a gaseous compound A is occurring in a liquid containing a com-
pound B which reacts according to
A (g → L) + B (L) → R (L)
The rate equation of the chemical stage is: (−rA ) = kC A C B and the reaction
occurs in a packed bed in which:
k Ag = 0.001 mol/(h m2 )
k Al = 100 m3 L /(m2 i h)
a = 100 m2 i /m3 R
f L = 0.01 m3 L /m3 R
DAl = DAB = 10−6 m2 /h
k = 10 m3 L /(mol h)
H A = 105 (Pa m3 L )/mol
Check how the reaction rate varies, as well as the influence of each resistance
throughout the reactor, for a reactor in which the liquid is fed at a concentration of
100 mol/m3 and is countercurrent to a gas that passes through it as a piston flow,
making the design so that the partial pressure of component A decreases from
100 to 1 Pa. Previous data obtained in a pilot plant indicates that the optimum
ratio between G′ and L′ fulfills:
G′ ∕PT
= 1 mol∕(m3 Pa)
L′ ∕CT

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

where the increment factor is given by



E −E
MH2 Ei −1
(√ )
i
E= (9.96)
2 Ei −E
tanh MH E −1
( i )
DBl ⋅ CB
Ei = 1 + (9.90)
b ⋅ DAl ⋅ CAi
The concentration of the gas in the interface can be calculated according to
Henry’s law:
pAi = HA CAi (9.77)
and the Hatta modulus is
k ⋅ DAl ⋅ CB
MH2 = 2
(9.93)
kAl
A molar balance of the reactant in the reactor gives
L′
G′ (YAin − YAout ) =(X − XBout ) (9.132)
b Bin
Bearing in mind that Y A ≈ pA /PT and X A ≈ C B /C T :
G′
PT
CBout = CBin − L′
(pAin − pAout )
b⋅CT
G′ ∕PT
From the data, we know that L′ ∕CT
= 1, and b = 1, so, in the present example:
CBout = CBin − (pAin − pAout )
Note that in this problem we cannot calculate the volume of the system,
because we do not know the value of G′ to use Eq. (9.133).
We will assume constant values of (−rA′′ ) at small sections of the reactor, cor-
responding each one to a value of pA . For the calculation of (−rA′′ ), a nonlinear
equation given in Eq. (9.96) must be solved at each value of pA . A Matlab program
will be used in the calculation, as shown in Program listing 9.5.

Program Listing 9.5


clear all
clear all
close all
pA0=100; % Pa
CB0=100; % mol/m3
9.9 Design of Multiphase Reactors 285

kAg = 0.001;% mol/(h ⋅ m2)


kAl = 100;% m3L/(m2sup ⋅ h)
a = 100;% m2sup/m3reactor
fL = 0.01;% m3L/m3reactor
DAB = 10e-6;% m2/ h
k = 100000; % m3liq/(mol ⋅ h)
HA = 10e5; %Pa ⋅ m3liq/mol

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))’)

The following plots are generated:

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

Carberry, J.J. (1976). Chemical and Catalytic Reaction Engineering. McGraw-Hill.


Conesa Ferrer, J.A. and Font Montesinos, R. (2002). Reactores Heterogeneos.
Universidad de Alicante.
Coulson, J.M. and Richardson, J.F. (1979). Chemical Engineering III. Chemical
Reactor Design, 2e. Oxford: Pergamon Press.
Gates, B.C. (1992). Catalytic Chemistry, 1992. Wiley.
Doraiswamy, L.K. and Sharma, M.M. (1984). Heterogeneous Reactions: Analysis,
Examples, and Reactor Design. Wiley.
Fogler, H.S. (1998). Elements of Chemical Reaction Engineering, 3e. Upper Saddle
River, New Jersey: Prentice Hall.
Jackson, R. (1977). Transport in Porous Catalyst. Elsevier.
Levenspiel, O. (1999). Chemical Reaction Engineering, 3rde. New York, NJ: Wiley.
289

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

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
290 10 Biochemical Reactors

(by adaptation of the microorganisms to the new environmental conditions)


can take place.
6. The distribution of products can be strongly conditioned by the environ-
mental operating conditions since, in many cases, a microorganism can
derive its metabolism in different directions, depending on the pH, dissolved
oxygen, etc.
7. Reaction enthalpies are low, so the design of heat exchange equipment should
not be a major problem.
8. The activation energy of the reaction is important so, especially in some tem-
perature ranges, the reaction rate is strongly dependent on the reaction tem-
perature. On the other hand, there are areas where the deactivation is also
very dependent on the temperature.
9. The operating temperature and pressure are moderate.
10. Good solid/liquid/gas contact is necessary to achieve effective action.
11. Both the gas supply and the disposal of the gas produced can be problematic.
12. The rheological properties of the fluid may vary during the fermentation
process. Furthermore, sometimes the fluids present a non-Newtonian
behavior.
13. Fermentation operations often require separation or, at least, management
of solid mass.
However, the efficiency of the fermentation processes is usually limited by dif-
ferent causes, among which they have a greater importance:
1. Low productivity due to a discontinuous operation,
2. Substrate inhibition problems,
3. Inhibition problems by product,
4. Low cellular or enzymatic concentration in the bioreactor,
5. Limitations for mass transfer.

10.2 Enzymatic Catalysis


From the structural point of view, enzymes are proteins, and from the kinetic
point of view, they are catalysts. The detailed knowledge of the catalytic function
of an enzyme requires the maximum knowledge of the structure of the enzy-
matic molecule. This is not only to know which amino acids and in what num-
ber are present in a given molecule but also the sequence of the amino acids in
the polypeptide chain. And, finally, the spatial conformation, which has a great
importance on the catalytic action.
Amino acids are not the only constituents of proteins (enzymes), as many
enzymes contain other organic or inorganic groups. These parts of the conjugated
proteins are called prosthetic groups.

10.2.1 Characteristics of Enzymatic Catalysis. The Active Center


Enzymatic catalysis involves the formation of a complex between the reagent
(substrate) and the enzyme, in a process of equilibrium. This complex (ES) is
10.2 Enzymatic Catalysis 291

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

Figure 10.1 Biochemical route in catalyzed enzymatic reactions.


292 10 Biochemical Reactors

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.

10.2.2 Kinetics of Enzymatic Reactions


In this section, we study the expressions of the reaction rate for enzymatic reac-
tions. The simplest reactions, consisting of the conversion of a single substrate
into a single product, for which the formation of an enzyme–substrate complex
is postulated, as discussed, is taken up first. Next, other reaction mechanisms that
will lead to more complex kinetic expressions are presented.

10.2.2.1 Kinetics of Reactions with a Single Substrate. Michaelis–Menten


Equation
Michaelis and Menten proposed a model to relate the reaction rate to the sub-
strate concentration. Assuming the existence of a single central complex, the
reaction scheme would be
k1 k
−−−−−−−−→
E+S← −− ES −−−−→ E + P
2

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

Substituting in the product balance equation:


dP k E Sf
=𝛽 2 0 f (10.9)
dt KM + S
The second hypothesis that is introduced is to suppose that the molar concen-
tration of the enzyme is much smaller than that of the substrate (E0 ≪ S0 ), so it
can be considered that the concentration of free substrate is approximately equal
to that of the substrate not converted Sf or S:
dP kES
=𝛽 2 0 (10.10)
dt KM + S
The reaction rate r is based on
dS 1 dP
r=− = (10.11)
dt 𝛽 dt
Finally, the Michaelis–Menten equation is obtained:
k2 E0 S
r= (10.12)
KM + S
This expression is usually written by grouping the term k 2 ⋅E0 as the maximum
reaction rate rm . The Michaelis–Menten equation predicts the first-order kinetic
behavior for low substrate concentrations and zero order for high concentrations.

10.2.2.2 Meaning of the Parameters of the Michaelis Equation


The Michaelis constant (K M ) corresponds to the substrate concentration at which
the reaction velocity is half the maximum velocity (rm = k 2 ⋅E0 ). Furthermore, for
small values of k 2 , K M represents the constant of the equilibrium of dissociation
of the E⋅S complex (K M = k −1 /k 1 ), in such a way that a high value of K M indicates
294 10 Biochemical Reactors

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.

Example 10.1 Metabolism of Ethanol


A patient in a hospital presents an alcohol content in the blood of 0.2 mg/dl, and
affirms that he had just taken a cup of white wine three days ago. It is reasonable
that this cup represents 250 ml of wine, what would produce an initial content
of alcohol in blood of 25 mg/dl. The patient’s liver only metabolizes ethanol at a
maximum rate of 1 mg/(dl h), and it is known that the enzyme involved in this
metabolism presents a value of K M of 10 mg/dl. Assuming a Michaelis–Menten
kinetics, try to find out if the patient is lying or, on the contrary, telling the truth.

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)

10.2.3 Enzymatic Reactions with Inhibition


Inhibitors are those substances that cause an enzymatic reaction to proceed more
slowly. This effect can occur in several ways:
(a) By means of a substrate analog that binds (reversibly or irreversibly) to the
active center, reducing the enzymatic activity.
(b) By binding an inhibitor to another part of the enzyme molecule, causing
conformational changes that reduce the ability of the enzyme to bind to the
substrate.
The kinetics of the reactions with inhibition depends on the mechanism of the
process, distinguishing between reversible or irreversible inhibition depending
on whether the original activity can be recovered or not:
(a) Irreversible inhibition. The inhibitor combines with the enzyme to give a sta-
ble but inactive complex.

E+S ES E + P
+
I EI (inactive)

(b) Reversible inhibition.

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

concentration. In fact, there are also situations of intermediate inhibition, which


can be explained by the general inhibition equation:
rm ⋅ S
r= ( ) ( ) (10.14)
KM 1 + k1 + S 1 + k1
i1 i2

in which k i1 and k i2 are the equilibrium constants of dissociation of El and EIS,


respectively.
(c) Inhibition by the substrate. In some cases of enzymatic reactions with a sin-
gle substrate, a behavior different from the hyperbolic law proposed by the
Michaelis equation is observed, so that the reaction velocity presents a max-
imum with the substrate concentration.
Several mechanisms have been proposed to explain the behavior of inhibition
by the substrate: (i) the formation of a second enzyme–substrate complex with
two substrate molecules bound to the enzyme, (ii) the binding of a fraction of
substrate molecules to the enzyme in a place other than the active center.
From any of the proposed mechanisms, the same kinetic equation is obtained:
rm ⋅ S
r= (10.15)
KM + S + K ′ S 2

10.2.4 Enzymatic Reactions with More than One Substrate


Many enzymes catalyze reactions between two or more substrates to give two or
more products. A particular case of this type of reactions are those catalyzed by
enzymes that use cofactors (for example, NAD+ in oxidoreductases), where the
cofactor plays the role of a second substrate.
Various mechanisms have been proposed to explain this type of reactions, most
of which can be classified into two types:
(a) All substrates bind to the enzyme before catalysis (in the case of two sub-
strates, it is called the mechanism of formation of a ternary complex).
(b) A product is released from a first binary complex before binding to the second
substrate (for example, in group transfer reactions).
Restricting the discussion to enzymatic reactions with two substrates, in
any of the indicated cases, the treatment by means of the hypothesis of the
pseudo-stationary state of the intermediates leads to kinetic equations that can
be summarized in the following general form:
rm ⋅ A ⋅ B
r= ′ (10.16)
KA ⋅ KB + KB A + KA B + A ⋅ B
where A and B are the molar concentrations of the two substrates involved, A and
B. The particular expression corresponding to the values of the constants varies
according to the mechanism that takes place.

10.2.4.1 Case 1. Enzymatic Reactions with Two Substrates by Formation of a


Ternary Complex
In this case, the ternary complex can be formed independently of the binding
order of the substrates (random) or only in a certain order (ordered mechanism).
10.2 Enzymatic Catalysis 297

(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

And, bearing in mind that the observed reaction rate is


r = k3 ⋅ [EAB] (10.22)
we can finally write
[A][B]
r = k3 ⋅ [EAB] = k3 ⋅ E0 ⋅ AB AB BA
KA KM
+ [A]KM + [B]KM + [A][B]
[A][B]
= rmax ⋅ AB AB BA
(10.23)
KA KM + [A]KM + [B]KM + [A][B]
In the present case, using the previous definitions, k cat = k 3, if the runs are per-
formed with a constant concentration of “A,” but not saturating the environment,
we can observe that the following is fulfilled:
[B]

r = rmax ⋅ (10.24)
K ′ + [B]
where
[A]

rmax = k3 ⋅ E0 ⋅ BA
(10.25)
KM + [A]
(KA + [A])
AB
K ′ = KM ⋅ BA
(10.26)
KM + [A]
If we represent the reaction rates vs. concentration of the substrate, we will

obtain graphs similar to those shown in Figure 10.2. Plotting the value of rmax
BA
obtained at different rates, [A]we can obtain the value of KM as the value of
concentration giving half of the maximum reaction rate. On the other hand, at

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

Using the Cleland’ symbols, we can say:

A B X Y

E EA EAB EXY EY E

Using Michaelis–Menten kinetics for the formation of the complexes, we can


then write
[E] ⋅ [A]
KA = (10.17)
[EA]
AB [EA] ⋅ [B]
KM = (10.19)
[EAB]
E0 = [E] + [EA] + [EAB] (10.28)
From the previous equations, we can deduce that:
[A][B]
r = k3 ⋅ [EAB] = k3 ⋅ E0 ⋅ AB AB
KA KM + [A]KM + [A][B]
[A][B]
= rmax ⋅ AB AB
(10.29)
KA KM + [A]KM + [A][B]
In the present case, using the previous definitions, k cat = k 3 , if the runs are per-
formed with a constant concentration of “A,” but not saturating the environment,
we can observe that the following is fulfilled:
[B]
r = rmax ⋅ (10.30)
K ′ + [B]
where
( )
KA

K = + 1 ⋅ KM
AB
(10.31)
[A]
Now we can construct the graphs shown in Figure 10.3 to calculate the values
of the constants.
300 10 Biochemical Reactors

K′ Figure 10.3 K ′ vs. concentration of


substrate in a bisusbtrate ordered
mechanism.
Slope K′ vs. 1/[A]
equals KA . KMAB

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.

10.2.4.2 Case 2. Enzymatic Reactions with Two Substrates Without Formation


of a Complex
The mechanism is called ping-pong, forming a complex of the enzyme with a
substrate (EA), which releases the product X, remaining as EA (for example, an
acyl-enzyme) that is attacked by the second substrate (for example, transferring
the acyl group).
k1 k2
−−−−−−−−→
E+A← −− EA −−−−→ EA + X

k−1
k3 k4
EA′ + B −−−−→ EA′ B −−−−→ E + Y
This is
A X B Y

E EA EA′ EA′B E

In that case:
r = k4 ⋅ [EB] (10.35)
and k cat = k 4 .
10.2 Enzymatic Catalysis 301

Using the Michaelis–Menten constants:

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

where rmax = k 4 ⋅ E0 . For [A] the constant we can see that:


[A]

r = rmax ⋅ (10.40)
K ′ + [A]
with:
[A]

rmax = rmax ⋅ ( ) (10.41)
k4 A
k2
KM + [A]

and
[A]
B
K ′ = KM ⋅( ) (10.42)
k4 A
k2
KM + [A]

10.2.4.3 Strategies to Distinguish the Previous Cases


Representing the value of the observed constant K ′ vs. the concentration of both
substrates, we can easily distinguish the previous cases. Figure 10.4 presents a
scheme of the expected graphs for the different mechanisms.

Example 10.2 Calculating Kinetic Law


In different experiments trying to calculate the kinetic law of the reduction of
pyruvate (A) and hydroxybutyrate (B) with lactate dehydrogenase (LDH, EC
1.1.1.27), we find the following reaction rates at the different concentrations:

K′ K′

Random Random

Ping-pong Ping-pong

Ordered Ordered

[A] [B]

Figure 10.4 Exemplification of K ′ vs. concentration for the different mechanisms.


302 10 Biochemical Reactors

[A] (g/l) [B] (g/l) r (g/(l s)) [A] (g/l) [B] (g/l) r (g/(l s))

0.05 0.1 0.34 0.5 0.1 1.35


0.05 1 1.93 0.5 1 3.34
0.05 1.5 2.34 0.5 1.5 3.57
0.05 2 2.61 0.5 2 3.66
0.05 2.5 2.82 0.5 2.5 3.76
0.05 3 2.94 0.5 3 3.81
0.1 0.1 0.58 2 0.1 1.78
0.1 1 2.51 2 1 3.60
0.1 1.5 2.90 2 1.5 3.76
0.1 2 3.11 2 2 3.77
0.1 2.5 3.28 2 2.5 3.87
0.1 3 3.37 2 3 3.85

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

2.00 [A] = 0.05


(A) = 0.1
1.50 (A) = 0.5
(A) = 2
1.00

0.50

0.00
0 0.5 1 1.5 2 2.5 3 3.5
[B] (g/l)

Considering Eqs. (10.24) or (10.30) or (10.40), we see that, in any case, we


can do:
1 K′ 1 1
= ⋅ +
r rmax [B] rmax
10.2 Enzymatic Catalysis 303

Using the data in the statement, the following is obtained:


3.5
[A] = 0.05
(A) = 0.1
3
(A) = 0.5
(A) = 2
2.5
Y = 0.273x + 0.2463

2
1/r (l s/g)

1.5 Y = 0.1473x + 0.2477

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

[A] (g/L) k′ (g/l) 1/rmax ((l s)/g)

0.05 1.10 0.246


0.1 0.60 0.248
0.5 0.20 0.248
2 0.13 0.247

The value of rmax is then 4 g/(l s) approximately. Plotting the values of K ′ , we


obtain
1.20

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)

This is typical of an ordered mechanism, as mentioned before, with


( )
KA

K = + 1 ⋅ KM AB
[A]
304 10 Biochemical Reactors

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 Microbial Kinetics


A cell growth process involves the consumption of the energy supplying sub-
strates and the raw mass needed for the synthesis of cellular mass and other
metabolic products. From a macroscopic viewpoint, this requires that the
microenvironment contains all the elements necessary for the formation of
additional cellular mass and that, in addition to this, the energy consumed by
substrates must be greater than the energy of the cells or products. It is to be
fulfilled, in addition to this, that all elements supplied as nutrients are present in
a manner consistent with enzymatic mechanism of the cell.
Cell growth obeys the laws of conservation of matter. The atoms of carbon,
nitrogen, oxygen, and other elements are reordered in the metabolic processes
so that the total amount incorporated matches the amount eliminated from
the environment. Furthermore, the amount of some of the metabolic products
formed or heat generated by cell growth is often proportional to the consumed
amount of any of the substrates or the formed amount of a particular product.
All these properties make possible the approach of energy and mass balances in
processes of cell growth, expressed generally as
Source of C + N + source O2 + minerals + specific nutrients
→ cell mass + products + CO2 + H2 O
The stoichiometric treatment of the cell growth processes, consumption of sub-
strates, and production of substances and heat is done in Section 10.3.1. In the
analysis in Section 10.3.2, we study the kinetics to which these processes take
place.

10.3.1 Stoichiometry of the Microbian Growth


From a quantitative point of view, the required amounts of nutrients can be
determined from the stoichiometry of the growth and product formation. For
example, the reaction aerobically consuming glucose to synthetize cellular
matter is
C6 H12 O6 + aO2 + bNH3 → cCHx Oy Nz + dCO2 + eH2 O
In order to calculate the stoichiometric coefficients, it is necessary to know the
elemental composition of the microorganism (CHx Oy Nz ). This composition can
be determined experimentally by an elemental analysis, obtaining the propor-
tions of the various elements.

10.3.2 Stoichiometry of Product Formation


In a growth medium, a large variety of products that accumulate intracellularly
or are released into the medium can be produced.
10.3 Microbial Kinetics 305

If the main product is a consequence of primary metabolism, a chemical


reaction similar to the former one can be written, introducing the product indi-
cated as CHv Ow (if the product contains nitrogen, it should be also considered):
aCHm On + bO2 + cNH3 → dCHx Oy Nz + eCO2 + f H2 O + gCHv Ow

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.

10.3.2.2 Theoretical Yield Obtained from Stoichiometric Coefficients


As the biomass–substrate yield is defined, yields can be defined for other pairs of
parameters of the process, as indicated in Table 10.1.
Using the definitions of the yields, we can relate the rate of the formation or
consumption of different species, so that, once having characterized the behavior
of one, the other can be calculated:
( ) ( ) ( )
1 1 1
rS = − ⋅ rX = − ⋅ rP = − ⋅ rC (10.44)
YX YP YC∕S
S S

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.

10.3.3 Cell Growth, Substrate Consumption, and Product Formation


As mentioned in the introduction of this chapter, the process of cell growth is very
complex, since it is the result of the interaction between a population of cells and
their environment.
In a batch reactor, the cell growth occurs within the reactor. But in a definite
moment, as a consequence of the absence of regeneration of the culture medium,
Table 10.1 Yield definitions.

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

Figure 10.5 Phases of the cell growth.


Stationary phase

Death phase
Log (X)

Exponential growth
phase

Lag phase

Time

the growth stops when it finds a limitation (consumption of an essential nutrient,


accumulation of a toxic product for metabolism, and limitation by oxygen are the
more frequent). Figure 10.5 shows the typical phases of batch cell growth.
Firstly, there is a lag phase, in which a substantial increase in the number of
cells does not occur, and that is required for adaptation to the medium and cul-
ture conditions. Then the exponential growth phase begins, in which growth is
produced at a specific rate constant. In this phase, it can be considered that the
growth is balanced, in the sense that there are no limitations and the metabolic
and physiological behavior of the cells therein is basically constant. The equation
describing the increased cell concentration is in this case:
dX
=𝜇⋅X (10.45)
dt
with a certain concentration of cells at the end of the lag phase (t lag ), which
depends on the inoculum used (X 0 ).
X = X0 at t = tlag
The integrated form of these equations is
( )
X
ln = 𝜇(t − tlag ) (10.46)
X0
X = X0 e𝜇(t−tlag )
The usual way to determine the value of μ from the experimental data of batch
growth is a semilog representation of X vs. time.

10.3.3.1 Kinetics of Growth


As mentioned, although the cell growth is a complex phenomenon, often we can
obtain a reasonably good overall description using relatively simple equations.
Among them, the most usual is the Monod equation, which describes the cell
growth as a function of the availability of a limiting substrate, and that can be
expressed as
Substrate (S) + Cells (X) → More cells (X) + Product (P)
dX S⋅X
rX = = 𝜇m (10.47)
dt KS + S
10.3 Microbial Kinetics 307

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.

10.3.3.2 Kinetics of Maintenance


Although the Monod kinetics is one of the most widely used and provides a rea-
sonable description of cell growth in the function of the substrate available for
most cases, these equations do not include the death phase. This can be changed
with the addition or an extra term that reflects the consumption of cellular mass
to produce energy for maintenance:
S
𝜇 = 𝜇m − kd (10.50)
KS + S
The constant k d is known as the coefficient of endogenous breathing. At high
growth rates, the breathing virtually does not affect the formation of biomass;
however, in an environment with low levels of nutrients, it becomes significant
and must be included in the kinetics expressions. In any case, the rate of cell death
can be calculated as
rd = kd ⋅ X (10.51)
308 10 Biochemical Reactors

Example 10.3 Bacterial Growth


A strain of Papallacta bacteria is going to be used as precursor of an enzy-
matic broth, which studies show can be used for the degradation of cellulosic
compounds and hydrocarbons. The chosen microorganism is a thermophilic
bacterium isolated from the thermal sources of Papallacta that, according to
elemental analysis, fits the formula C 4,17 H 7,21 O1,79 N. The microbiological reports
indicate that the source of nitrogen for growth is nitrate, but it is also possible to
grow on ammonium, and it also indicates that the bacteria can grow in an aerobic
or anaerobic medium (facultative bacteria, which indicates that the electron
acceptor may or may not be oxygen) and use as a source of carbon (electron
donor) monosaccharides of 5 or 6 carbon atoms, which could well come from
cane juice or molasses. In a 1-l laboratory batch reactor, the kinetics of bacterial
growth is determined using glucose as substrate. The following data is obtained:

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

Determine the values of K S and 𝜇m of the Monod equation.

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)

0 0.262 0.52 0.4094 1.98


5 0.452 0.76 0.2318 1.68
10 0.266 0.52 0.0632 1.95
15 0.288 0.52 0.0520 1.81
20 0.504 0.72 0.0722 1.43
25 0.16 0.3 0.0168 1.88
30 0.34 0.72 0.0330 2.12
35 0.14 0.34 0.0117 2.43
40 0.08 0.24 0.0063 3.00
45 0.08 0.22 0.0061 2.75
50 0.04 0.12 0.0030 3.00

As we can see, the rate of production of cells is continuously increasing, in


such a way that a phenomenon similar to autocatalysis is found. The amount of
cells produced by a gram of substrate is also continuously increasing. Following
Monod kinetics, we have that (from Eq. (10.48)):
1 1 K 1
= + S ⋅
𝜇 𝜇m 𝜇m S
So, a graph 1/𝜇 vs. 1/S will give the desired parameters. In the following plot,
we can see the corresponding data graphically:
400

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.

10.4 Immobilization of Enzymes and Cells: Mass


Transfer Effects
The immobilization of a biocatalyst, whether it is an enzyme or a cell, consists
of its location in a defined region of space, while maintaining a desired catalytic
activity and, in the specific case of the cells, very possibly also its viability. A gen-
eral characteristic of any system with immobilized biocatalysts is that the trans-
port of substrates and products is produced by diffusional mechanisms.
The main methods of immobilization of biocatalysts can be divided into five
large groups, depending on the mechanism on which they are based:
1. Adsorption: The immobilization is produced by interaction of the ionic type
or weak attraction forces on the surface of the support.
10.4 Immobilization of Enzymes and Cells: Mass Transfer Effects 311

2. Covalent link: The immobilization is also performed on the surface of a


support, but the interaction between the biocatalyst and the support is by
covalent bonding.
3. Cross-links and self-immobilization: In this group, there is no support itself.
The particles of the immobilized biocatalyst are achieved by direct interac-
tion, either by covalent binding or, in the case of some cells, by flocculation
processes.
4. Entrapment: Formation of a three-dimensional structure, in the presence of
the biocatalyst, which is trapped uniformly inside it.
5. Systems with membranes (microencapsulation, preformed membranes):
In both cases, the biocatalyst is immobilized inside a space limited by a
membrane. In the first case, the microcapsules are produced in the presence
of the biocatalyst, which is incorporated in its interior. In the second case, the
biocatalyst is introduced into a system with preformed membranes.

The immobilization of biocatalysts presents a series of advantages:

– 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

Enzymes entrapped in a matrix

Figure 10.6 Scheme of immobilization of enzymes in different structures.


312 10 Biochemical Reactors

In the case of immobilized enzymes, it is common to observe an increase in


their stability and resistance to denaturation and proteolysis when they are immo-
bilized. In the case of cells, the effects of accidental contamination of the process
are usually reduced, given the population of contaminating cells that are in sus-
pension. And at a much lower concentration than the population of immobilized
cells, it can be eliminated much more easily from the reactor.
On the other hand, the use of immobilized biocatalysts also presents some
drawbacks:

– In the first place, it is necessary to consider that the activity of a cell or an


enzyme can be directly affected by the conditions under which the immobi-
lization process is carried out, losing part of its activity or its viability.
– The rate of diffusion of substrates and products within the system of immo-
bilized biocatalysts can limit their catalytic activity and their efficiency, when
this rate is slower than the rate of the transformation that is being carried out.
– Finally, from the point of view of the overall process, it must be borne in
mind that the immobilization means the incorporation of a new stage to it,
and therefore an increase in its complexity, which has to be compensated for
by clear productivity increases, improvements in the separation between the
product and the biocatalyst, and longer operation times.

The decision to use enzymes or cells as the main groups of immobilized


biocatalysts depends fundamentally on the type of conversion that is to be
carried out. When it is a conversion involving a single stage, it is normally more
appropriate to use purified enzymes, since their cost can be reasonable and they
have sufficiently good activities and stabilities. In addition, these preparations
are simple and well-defined systems. In those more complex transformations,
in which multiple stages are required, involving a large number of enzymes and
their cofactors, the use of whole cells is indicated. In this case, it is not necessary
to carry out a preliminary extraction and purification of the enzymes, which
is often expensive. On the other hand, the cells have other enzymes and active
metabolic routes in addition to those that intervene in the transformation of
interest, which can cause the appearance of unwanted by-products, or even the
modification or degradation of the desired final product.
An important aspect is the flexibility of cells as biocatalysts. Different systems
can be prepared from cells, depending on whether the transformation involves
only one enzyme (in this case, dead cells can be used permeabilized as if they were
enzymes, but without the need for purification) or multiple stages, for which it is
necessary to use viable cells.
An intermediate situation between the two cases described, although much
less frequent, is the immobilization of a multienzyme system, interesting in those
cases in which a reduced number of enzymes are involved.
Regarding the main types of bioreactors in which the immobilized biocat-
alysts are used, which can be similar to heterogeneous catalysis reactors, the
same type of reactors are used: fixed-bed or fluidized-bed reactors, reactors
with membranes (which, as mentioned, constitute at the same time a form of
immobilization), and, in a smaller proportion, the agitated reactors, which can
10.4 Immobilization of Enzymes and Cells: Mass Transfer Effects 313

be used in very specific cases, since agitation usually severely affects the physical
integrity of most particles of immobilized biocatalysts.

10.4.1 Effect of Limitation by Internal Diffusion


The influence of the external mass transfer on the global reaction rate is classically
expressed using the concept of efficiency factor, similar to that used for internal
mass transport in catalytic reactors (see Chapter 9). The physical meaning of this
factor is the quotient between the observed reaction rate (therefore, the one that
takes place) and the reaction rate that would be had in the absence of limitations
by external matter transfer, that is, if the concentration on the surface was the
same as that of the fluid phase:
observed reaction rate
𝜂e = (10.52)
rate with no limitation by external mass tranfer
Figure 10.6 shows an scheme of the diffusion steps ocurring in an immo-
bilized enzymatic system. The process will be produced at the rate, using the
Michaelis–Menten equation:
Ss
r = kC (Sf − Ss ) = rm (10.53)
KM + Ss
The rates of diffusion and reaction should be the same. So, reordering the
equations, we can do:
(Sf − Ss ) Ss Sf
r= 1
= (K +S )
= 1 (KM +Ss )
(10.54)
k
M
r
s
k
+ r
C m C m

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

Using the following definitions:


𝜉 = Ss ∕Sf and 𝜅 = Ss ∕KM
and solving for the efficiency factor:
( )
𝜅𝜉
1+𝜅𝜉
𝜂e = ( ) (10.56)
𝜅
1+𝜅

In the case of gel-containing biocatalysts or where the biocatalyst has been


immobilized in the pores of a support, it is not expected that the reaction will
only occur on the surface. Similarly, it is expected that substrate consumption by
a microbial sheet or a floccule will occur at a certain depth in the microbial mass.
This situation is more complicated than in the case of surface immobilization;
in this case, the transport and the reaction occur in parallel. By analogy with
314 10 Biochemical Reactors

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

1. Stirred Tank: Can Be Batch, Fed-batch (with Discontinuous Feed), and


Continuous (CSTR or CSTB).
Batch: it is the most traditional and the most widely used on an industrial scale.
The necessary operating time can range from hours to several weeks depending
on the conversion and the operating conditions. During the operation it is impor-
tant to avoid contamination, maintain agitation, and control pH and temperature.
They operate with low cell density, especially in the initial period, and the feeding
of high concentrations of substrate must be avoided to prevent inhibitions by the
substrate or by the product. It is used mostly in the food, pharmaceutical, and
biotechnology industry, in general, since it is easy to reach and maintain aseptic
conditions during the operation. This fact is important when trying to maintain
processes that use very nutrient-rich media. Typical examples are the production
of acetic acid, vitamin C and C12 , baking yeast, penicillin, and anaerobic sludge
digestion (microorganisms produced in continuous wastewater treatments).
The disadvantages that batch reactors present are typical of this type of systems:
loss of efficiency due to start-ups and shutdowns, lack of product homogeneity
between loads, and difficulty in implementing energy integration schemes.
In general, the design variables are as follows:
1. Dimensions of the reactor,
2. Operation time,
3. Initial concentrations,
4. Volume of microbial mass per unit of reactor volume,
5. Power and aeration,
6. Heat transfer surface.
Fed-batch: the substrate is fed in successive loads and no product is removed,
thus varying the volume of reaction. By varying the speed and concentration of
the feed, the concentration of one or more nutrients or substrate in the culture
medium is controlled or altered. This increases overall productivity. This strategy
is applied in the production of bread yeast and antibiotics.
Continuous CSTR (or CSTB): this continuous stirred tank reactor (or bioreac-
tor) is also called the chemostat. This reactor is similar to the batch reactor but
allows the entry and exit of a flow. However, the chemostat mainly operates in
a steady state, which implies uniform conditions not only from the geometric
point of view but also with respect to time. The control of the variables of opera-
tion is much easier, but the feasible values of the variables are limited; if the flow is
too high, it can happen that the production of microorganisms is lower than the
number of those that are dragged by the output current and, consequently, if the
situation persists it happens that no microorganisms remain inside the reactor. It
is said that they have been washed (washout) from the fermenter. The situation
can be lightened by adding microorganisms similar to those of the reactor in the
feed. The natural source of these microorganisms is obviously the output current.
Thus, the effluent is passed through a settler or a centrifuge, and a concentrated
solution of microorganisms is separated, which is largely recycled to the reactor.
This procedure not only prevents the washing of cells but also increases produc-
tion as it increases the residence time of the microorganisms in the reactor.
316 10 Biochemical Reactors

2. Tubular Bioreactor (TB)


It should be noted for this bioreactor that the operation is not feasible when a
sterile medium is fed, given that it is an autocatalytic process. This drawback is
solved with a recirculation or with immobilized microorganisms. According to
the disposition of the microorganism, they are divided into the following:
• TB with flocs in suspension: A floccule is an aggregate of microorganisms
with a size an order of magnitude higher than that of the microorganism.
The mechanism controlling this size is not very well known but can act on
the agitation and flocculating agents such as aluminum chloride and calcium.
A fixed-bed operation is not viable. The viable operation involves the entrain-
ment of microorganisms and, consequently, a necessary feeding of them.
• Tubular film bioreactor: The microorganisms are arranged in the form of a
film that grows on the surface of an inert filler. The thickness of the film is
controlled by removing excess microbial production normally by mechanical
methods. It does not present dragging problems. It is widely used in the treat-
ment of wastewater and is called percolating or drip filter. It is an inert bed
on which the microbial film is formed and in which the liquid circulates in a
downward flow by gravity, while the oxygen usually flows upward. Its height
varies between 2 and 15 m. The control of the thickness of the film is carried
out by the organisms themselves, since by increasing oxygen diffusion prob-
lems are created, so that the microorganisms cannot metabolize the substrate,
lose adherence, and eventually become detached.

10.5.1 Continuous Stirred Tank Bioreactor (CSTB)


Given the reactor scheme shown in Figure 10.7, the first step in the analysis or
design, is to do the mass balances. It is first necessary to indicate that the cell
balance is carried out on the so-called viable cells, which are those capable of
dividing and forming a colony in the culture medium. So, you have
d(V ⋅ X)
Viable cells ∶ = Q0 X0 − QX + V (rX − rd ) (10.62)
dt
rd being given by Eq. (10.57). In this kind of reactors, usually dV /dt = 0 (as
Q = Q0 ), and in the steady state (dX/dt = 0) we have
0 = Q0 (X0 − X) + V 𝜇 X − V kd X (10.63)

Feed: Q0, X0, S0, P0 Figure 10.7 Scheme of a continuous


stirred tank bioreactor (CSTB).

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

10.5.1.1 Influence of the Dilution Rate. Calculation of the Bioreactor Wash


For small flows and fixed volume, the value of D is also small and tends to
zero when it does flow, Q0 → 0; D → 0. In consequence of Eq. (10.65), it follows
that S also tends to zero, S → 0, which means that the substrate has been
consumed practically since the microorganisms have had enough time to do
it. When the flow increases and, therefore, D increases, it is observed that the
substrate concentration increases first slightly linearly and then abruptly when
D approaches 𝜇m .
From Eq. (10.66) it follows that when D → 𝜇m , S → ∞, which does not make
sense, since S ≤ S0 . At most, the concentration of substrate can be equal to the
initial input. In parallel, the concentration of biomass first decreases slowly and
then abruptly tends to zero when S → S0 .
That is, before D = 𝜇m , S = S0 , and X = 0 have already been produced, which
means that the cells have disappeared from the bioreactor, and the washing of
the same has taken place. The value of the dilution rate for which such a limit
situation is called “washout dilution rate” and is determined from Eq. (10.67) by
doing S = S0 :
S0
Dmax = Dwashout = 𝜇m (10.69)
KS + S0
318 10 Biochemical Reactors

S0 Figure 10.8 Variation of substrate, biomass, and cell


X, S or D·X production concentrations with the dilution rate.

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.

Example 10.4 Optima Microbial Growth with Monod Kinetics


The microbial growth of Escherichia coli in glucose follows the kinetics of Monod,
with K S = 4 g/m3 and maximum velocity of 1.33 g/(m3 h) cells. The value of Y X/S
is 0.1 g/g. Determine the flow rate of glucose feed (concentration S0 = 60 g/m3 )
to a continuous stirred tank fermenter of 1 m3 that would give a maximum rate
of glucose consumption and cell production. Determine these values.

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

The rate of the microbial growth at this dilution rate is


Sopt
ropt = 𝜇m X
KS + Sopt opt
KS 0.325 ⋅ 4
Sopt = Dopt ⋅ = = 1.33 g∕m3
𝜇m − Dopt 1.3 − 0.325
Xopt = (S0 − Sopt ) ⋅ YX∕S = (60 − 1.33) ⋅ 0.1 = 5.86 g∕m3
1.33
ropt = 1.3 5.86 = 1.9 g∕(h m3 )
4 + 1.33

10.5.1.2 Cell Recirculation


A non-sterile feed is easily achieved by recirculating part of the previous product
by separation and concentration of microorganisms. A schematic of this proce-
dure is shown in Figure 10.9.
A balance of microorganisms to the bioreactor leads to
0 = Qr X0 − (Q0 + Qr )X1 + rX V = Qr X0 − (Q0 + Qr )X1 + 𝜇X1 V (10.70)
Using the definition of D = Q0 /V , and renaming 𝛼 = Qr /Q0 and 𝛽 = X 0 /X 1 :
𝜇
D= (10.71)
1 − 𝛼(𝛽 − 1)
whereas the microorganisms in the recirculation stream are more concentrated
than in the output current of bioreactor results 𝛽 > 1. Consequently, now the
dilution rate is higher than the specific speed of growth and, therefore, with the
same rate of growth, the use of cell recirculation allows to treat more food per
unit of time and volume than when there is no recirculation. Comparing with
Eq. (10.65), we have
Dwithout recirculation
Dwith recirculation = (10.72)
1 − 𝛼(𝛽 − 1)

10.5.2 Tubular Fermenters with Flocs


As in the case of ordinary reactors, an overall balance for the entire reactor is not
valid as the conditions there are not uniform with the position in the reactor. In

Feed: Q0, X = 0, S0, P0


Product: Q0, S1, X

Recycle:
Qr, X0
Centrifuge
Volume = V

Qr + Q0, X1, S1, P1

Figure 10.9 Continuous agitated tank fermenter with recirculation of microorganisms.


320 10 Biochemical Reactors

dV Figure 10.10 Scheme of a


tubular bioreactor.
Q, P0, S0, X0 P, S, X

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.

10.5.2.1 Tubular Fermenter with Recirculation and Monod Kinetics


Considering the substrate balance in the mixing point of the inlet of the reactor
(see Figure 10.11):
Q ⋅ S0 + Qr ⋅ S = (Q + Qr ) ⋅ Sin (10.75)
Sin being the concentration of substrate entering the reactor. Expressing
(Q + Qr ) = Q⋅(1 + R), we have, from Eq. (10.74):
S S
1 V dS (KS + S)
= = (1 + R) = (1 + R) dS (10.76)
D Q ∫Sin YS∕X ⋅ rX ∫Sin YS∕X ⋅ 𝜇max ⋅ S ⋅ X

10.5.3 Fed-batch Bioreactor


Fed-batch culture consists of a biological process where one or more substrates
are fed during operation, but the products remain in the reactor until the end of
the run. From the reactor-type point of view, it is a semi-batch reactor system. In

dV
P, S, X

Q, P0, S0, X0 Q

Qr R = Qr/Q

Figure 10.11 Scheme of a tubular bioreactor with recycle.


10.5 Bioreactors 321

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)

Figure 10.12 Solution of Example 10.5.

and then, using the previous equation, we can write


( )
X t+1 − X t Q ⋅ (X0 − X) Q0 ⋅ (X0 − X t )
= 0 + rX X t+1 t
= X + Δt + rX
Δt V V
In the same way:
( )
t+1 t
Q0 ⋅ P
P = P + Δt − + YP∕X rX
V
( )
Q0 rX
St+1 t
= S + Δt (S − S) −
V 0 YX∕S
Using an approach similar to previous programs, a valid Matlab calculation is
given. The results are shown in Figure 10.12.

Program Listing 10.1


clear all
close all
mumax = 0.20; % h-1 Maximum Growth Rate
Ks = 1.00; % g/l Monod Constant
Yxs = 0.5; % g/g Cell yield
Ypx = 0.2; % g/g Product yield
S0 = 10.0; % g/l Feed Substrate concentration
Q=1; % L/h Input Flowrate

V0=1; % L Initial Volume


10.5 Bioreactors 323

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.

Example 10.6 CSTR Bioreactor with Recirculation of Cells


The reaction of the previous example is done in a CSTR bioreactor with recircu-
lation of cells. Determine the variation of the exit concentrations with the recycle
relation, assuming that the concentration of cells in the recycling current is half
of the concentration at the exit of the reactor. A volume of 10 l of reactor should
be used. Repeat the simulation for a fixed value of recycle ratio and varying the
concentration of cells at the exit of the reactor.
Solution
A scheme of such a system can be found in Figure 10.9. A balance of microorgan-
isms to the bioreactor leads to
0 = Qr X0 − (Q0 + Qr )X1 + rX V = Qr X0 − (Q0 + Qr )X1 + 𝜇X1 V (10.70)
324 10 Biochemical Reactors

Using the definition of D = Q0 /V , and renaming 𝛼 = Qr /Q0 and 𝛽 = X 0 /X 1 :


D = 𝜇∕(1 − 𝛼(𝛽 − 1)) (10.71)
In this example we have: X 0 = 0.05 g/l (note that in this case this concentration
represents that of the recirculating current, Figure 10.9), S0 = 10 g/l, V = 10 l (con-
stant in this case), Q0 = 1 l/h, P0 = 0, 𝜇m = 0.20 h−1 ; K S = 1.00 g/l; Y X/S = 0.5 g/g;
Y P/X = 0.2. With this data, we can calculate D = (1 l/h)/10 l = 0.1 h−1 .
The fraction of cells that are being recirculated can be calculated from
Qr ⋅ X0 𝛼 ⋅ X0 𝛼⋅𝛽
Fraction of recirculated cells = = =
(Qr + Q0 ) ⋅ X1 (𝛼 + 1) ⋅ X1 (𝛼 + 1)
(10.80)
A mass balance in the centrifuge gives
(Qr + Q0 ) ⋅ X1 = Q0 ⋅ X + Qr ⋅ X0 (10.81)
and then:
X = X1 ⋅ [1 − 𝛼(𝛽 − 1)] (10.82)
From a global mass balance in the system:
𝜇 ⋅X ⋅S
0 = −Q0 X + rX V = −D ⋅ X + m 1 1 (10.83)
KS + S1
After substituting X by Eq. (10.82), we can get the relation giving the substrate
concentration at the exit and in the recirculating current:
D ⋅ [1 − 𝛼(𝛽 − 1)] ⋅ KS
S1 = (10.84)
𝜇m − D ⋅ [1 − 𝛼(𝛽 − 1)]
There are many ways to work in this type of system. One can change the value
of the relation X 1 /X 0 = 𝛽, and the system evolves accordingly. Let us take the
value of 𝛽 = 0.5 mentioned in the statement; in that case, we can use the previous
equation to solve, and the following plot is obtained:
3.5 0.16

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

Doran, P.M. (2013). Bioprocess Engineering Principles. Elsevier/Academic Press.


Lee, K., Comolli, N., Punzi, V. et al. (2014). Survey: Matlab & Mathcad education in
biochemical engineering. Chemical Engineering Education 48 (1): 59–61.
Levenspiel, O. (1999). Chemical Reaction Engineering, 3e. New York, NY: Wiley.
327

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

Chemical Reactor Design: Mathematical Modeling and Applications,


First Edition. Juan A. Conesa.
© 2020 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2020 by Wiley-VCH Verlag GmbH & Co. KGaA.
328 Index

continuous stirred tank reactor (CSTR) e


(contd.) entrained-flow reactors 175, 176
dead volume and short circuit environmental catalysis
23–25 cleaning systems 202
exchange of matter 22–23 DOC 204
exothermal reaction 118–125 DPF system 204–205
experimental solid flow reactor exhaust gases 200
53–54 SCR 203–204
non-ideal 32 three-way catalyst 202–203
PFR in series 13–14 enzymatic catalysis
RTD 4, 12–13 characteristics of 291
RTD Laplace transform 62 inhibitors 295–296
second-order reaction in a series metabolism of ethanol 294–295
30–32 Michaelis constant 293
tanks-in-series model 16 Michaelis–Menten equation 293
unsteady state 111–113 substrates 296–304
conventional scaling process 128 enzyme-substrate complex 291, 292,
296
convolution
ethylbenzene production 193–194
calculation of 38–41
ethylene oxide 194–195, 241, 242, 244
definition 35
exothermal first-order reaction 118
RTD reactor 38
explicit method 93–94
of signals
external diffusion 215, 219–221, 232,
experimental solid flow reactor
233, 237, 239, 241, 242, 259, 260,
53–55
280
properties 37–38
extruder 185
pulse input tracer 51–52
three reactors connected in series f
50, 51 fed-batch bioreactor 320–325
two functions 36 fermentation 168, 289, 290
cordierite structure 184 first order approximation 88, 90, 95,
Cumene 189, 191, 220 105, 122, 260
cyclic operation 158, 159 fixed-bed reactor 164, 165, 175, 176,
231–232
d fluid catalytic cracking (FCC) 176,
Danckwerts boundary conditions 20, 191–192
27, 102 fluid-fluid reactions
deconvolution gas/liquid reactions 256–261
calculation of 42–43 in industry 245
of signals 50 transfer models 245–247
deep stirred reactor 77 two-film theory
De-NOx via SCR 238–241 Hatta modulus 250–251
diesel cracking 230–231 instantaneous reaction 247–250
diesel oxidation catalyst (DOC) 157, rate reaction 251–256
201, 204 fluidized-bed reactor 176
diesel particulate filter (DPF) 204–205 forced unsteady state operation (FUSO)
dispersion model 15–18, 27, 81–84 objective 158–159
Index 329

oscillating pressure 169–170 j


periodic flow-reversal 166 jacketed piston flow reactor 146–147
periodic operation/cyclic operation
158 k
periodic variation of kinetic law 301–304
concentration 163–164
design strategy 162–163 l
flow rate 164 lanthanum oxide (LaO2 ) 202
flow variation of reagents 159, Laplace transform (L.T.)
160 basic functions 61
input signals and process rate example of 82–83
160, 161 properties of 58, 59
modes of operation 160–162 pulse function 60–61
temperatures 165 ramp function 59–60
variables 160 RTD
types of 159 in CSTR 62–65
VVO reactor 168 PFR 65–66
sinusoidal function 60
g technique 66
𝛾-alumina 181 Lewis acid 185, 190, 191
gas-liquid reaction 247, 256–261, 275 liquid-liquid reaction 266
gas-solid reactors 175
m
h mass transfer effect 310–314
Hatta modulus 250–251, 255 matrix convolution 39, 45, 54
Henry’s law 246, 248, 252, 284 membrane reactor 254–256
hydrothermal process 184 metal active particles 184
methanation 197–200
i Michaelis complex 291
industrial catalysis Michaelis constant 293
acid catalysts 190–194 Michaelis–Menten equation 292–293,
commercial catalysts 313
active metals 185–186 microbial growth 318–319
catalytic support manufacturing microbial kinetics
185 cell growth 305–310
zeolites synthesis 184–185 product formation 305–310
environmental pollutants 200 stoichiometry of growth 304–305
oxidation products 194 modified Bessel function 227
oxidation/reduction reactions 188 modulation lean-rich “poor-rich”
reactions catalyzed, solids 175–177 204
solid catalysts 178–183 mole balance, PFR 113
internal diffusion 313 molybdenum 186, 196, 216
first-order kinetics Monod equation 306–308, 317
flat particles 222–225 Monod kinetics 307, 309, 310,
porous pellet 225 318–320
Thiele and Weisz modulus 229–236 monolithic catalytic reactors
irreversible inhibition 295 De-NOx via SCR 238–241
330 Index

monolithic catalytic reactors (contd.) p


fluidized bed reactor 241, 242, 244 parametrically sensitive areas 139
pore diffusional resistance 237–238 partial differential equations (PDE)
structure 238 classification of 87–88
multiphase reactors finite differences
design models, flow explicit method 93–94
ammonia absorption 269–272 first order approximation 88
cases and situations 269–284 gas phase 91
initial and boundary condition
catalytic system 278
94–96
chlorine removal 282–283
second-order approximation
fluid-fluid reaction over catalyst
89–91
276–277
stability 96–98
fluid-fluid reaction, tower
numerical methods
283–284
catalytic flat wall reaction 107
isomerization in fluidized bed
flow and dispersion 104–106
reactor 279–280
RTD of complex system 98–104
nomenclature 263
particulate matter (PM) 201, 204
pure absorption 263–265
Peclet–Bodenstein module 19
reactor of known volume
periodic operation 158–160, 162, 163,
275–276
167, 170
second order isomerization platinum (Pt) 191, 197, 202–204
278–279 plug flow reactor (PFR)
sucrose acid treatment 281 CSTR in series 13–14
sulfur oxides removal 273–275 dynamic regime
systems with reaction 265–268 no dispersion 113–15
trickle-bed type reactor 272–273 real reactors 25–32
flow types 261–262 with dispersion 115–118
RTD 4, 11–12, 14
n system of two reactors 80–81
Nerst’s law 248 tubular reactor 136–155
nickel 185, 186, 196, 199 pre-drying oven 185
non-ideal CSTR 32, 74–76 protonic acids 190
non-ideal flow pulse function 60–61
CSTR and PFR pulse input tracer 51–52
complex system 28–30 pulse modes 160, 161
second-order reaction 30–32 pure types of inhibition 295
tubular reactor 25–28
types of 26 q
RTD 3–11 quasi-steady mode 161

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

steam reforming of alcohols 197 sifting machine 185


syngas production 197 silica 179–181, 190, 191, 195, 215, 230
residence time distribution (RTD) sinusoidal function 60
characteristics of slurry reactors 176
cumulative distribution curve, F(t) solid catalysts
8 activated carbons 183
mean residence time 9–10 catalyst support 180
second and third moments 10–11 characteristics of 180
convolution 38–41 requirements 178
CSTR with dead volume and short silica and alumina 181
circuit 23–25 spherical pellet 179, 180
dispersion model 18 zeolites 181–183
experimental measurement of space velocity 186, 187
in heterogeneous systems 7–8 spherical catalyst 231, 233
pulse input 4 square pulse of tracer 77–78
step input 6–7 steam reforming
ideal reactors alcohols 197–198
batch and PFR reactors 11–12 of hydrocarbons 199
CSTR 12–13 sucrose acid treatment 281
PFR/CSTR in series 13–14 sulfur oxides removal 273–275
Laplace transform
CSTR 62–65 t
PFR 65–66 tanks-in-series model 15–18, 30, 31,
tanks-in-series model 15–18 69
two CSTR with exchange of matter tank-type reactor 13
22–23 theory of adsorption 217–219
reversible inhibition 295 thermal control 130
rhodium (Rh) 197, 202, 204 Thiele modulus 228–236
time-dependent functions 57
s transfer function
safe approach 141 CSTRs and PFRs
scaling of chemical reactor parallel systems 69–71
batch tank stirred reactor recycle system 71–81
example of 130 series systems 67–69
temperature control heat definition 57
transmission 130 dispersion model 81–84
definition of 127 Laplace transform 58–61
maximum factor, equipments 127 non-ideal chemical reactors 57
production rate and the time intervals pulse input signal 58
127 transfer models 245–247
rules 129 trickle-bed reactor 177, 272–273
second-order approximation 88–91, tubular bioreactor (TB) 316, 320
95 tubular reactor 25
selective catalytic reduction (SCR) isothermal operation 138
176, 201, 203–204, 238–241 optimum length, conversion and
semi-batch reactor 130–136, 320 performance for 138, 139
Sherwood module 220 PFR 136
332 Index

tubular reactor (contd.) instantaneous reaction 247–250


piston flow 137 rate reaction 251–252
stability process
areas of runaway 139 u
batch tank stirred reactor 145 unsteady state, CSTR 111–113
energy balance 141
v
initial concentration changes
variable volume (VVO) 157, 159,
152–155
168–169
jacketed piston flow reactor
146–147 w
Matlab 148–152 washcoat 184, 237
parametrically sensitive areas 139 Weisz modulus 229–236
temperature profiles 141, 142 wet mixer 185
van Welsenare and Froment Würtz synthesis 194
method 145
two-film theory z
Hatta modulus 250–251 zeolites 179–184, 190, 191, 216

You might also like