[FREE PDF sample] Catalytic Kinetics. Chemistry and Engineering 2nd Edition Dmitry Yu Murzin ebooks
[FREE PDF sample] Catalytic Kinetics. Chemistry and Engineering 2nd Edition Dmitry Yu Murzin ebooks
[FREE PDF sample] Catalytic Kinetics. Chemistry and Engineering 2nd Edition Dmitry Yu Murzin ebooks
com
https://textbookfull.com/product/catalytic-
kinetics-chemistry-and-engineering-2nd-edition-
dmitry-yu-murzin/
https://textbookfull.com/product/physical-chemistry-thermodynamics-
statistical-thermodynamics-and-kinetics-fourth-edition-engel/
textbookfull.com
https://textbookfull.com/product/engineering-chemistry-2nd-edition-o-
g-palanna/
textbookfull.com
https://textbookfull.com/product/english-humour-for-beginners-george-
mikes/
textbookfull.com
https://textbookfull.com/product/cruel-prince-royal-hearts-academy-
book-one-a-jade/
textbookfull.com
https://textbookfull.com/product/software-for-exascale-computing-
sppexa-2016-2019-hans-joachim-bungartz/
textbookfull.com
https://textbookfull.com/product/antimicrobials-in-agriculture-1st-
edition-arti-gupta-ram-prasad/
textbookfull.com
https://textbookfull.com/product/contemporary-topics-in-analytical-
and-clinical-chemistry-volume-3-1st-edition-david-hercules/
textbookfull.com
Historical Turning Points in Spanish Economic Growth and
Development, 1808–2008 Concha Betrán
https://textbookfull.com/product/historical-turning-points-in-spanish-
economic-growth-and-development-1808-2008-concha-betran/
textbookfull.com
CATALYTIC KINETICS
CATALYTIC KINETICS
Chemistry and Engineering
Second Edition
No part of this publication may be reproduced or transmitted in any form or by any means, electronic
or mechanical, including photocopying, recording, or any information storage and retrieval system,
without permission in writing from the publisher. Details on how to seek permission, further
information about the Publisher’s permissions policies and our arrangements with organizations such
as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our
website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing. As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical treatment
may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating
and using any information, methods, compounds, or experiments described herein. In using such
information or methods they should be mindful of their own safety and the safety of others, including
parties for whom they have a professional responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume
any liability for any injury and/or damage to persons or property as a matter of products liability,
negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas
contained in the material herein.
ISBN: 978-0-444-63753-6
TAPIO SALMI
Education: 1980: MSc (Chemical Engineering), Åbo Akademi University, Turku,
Finland; 1986: Doctor of Technology, Åbo Akademi University.
Career: 1985 Researcher at Technical University of Denmark; 1990–1998: Associate
Professor; since 1998: Professor in Chemical Reaction Engineering, Åbo Akademi
University, Turku, Finland; Academy Professor, Academy of Finland.
A co-author of Chemical Reaction Engineering and Reactor Technology (CRC Press, 2011)
and author and co-author of over 500 journal articles on various aspects of chemical
reaction engineering. Member of the European Federation of Chemical Engineering
Working Party on Chemical Engineering.
ix
PREFACE
xi
xii Preface
Similar to the previous edition, the authors limited themselves to the main mono-
graphs, review articles, and key references without extensive coverage of original
literature.
It is the pleasure of the authors to acknowledge many colleagues from academia and
industry who, through the years, shared their knowledge and expertise in fundamental
and applied aspects of catalytic kinetics. The late Professor M.I. Temkin deserves a special
mention as a role model of an exemplary person and scientist for the author of these words.
The authors hope that the second edition will be useful for students and researchers
working in the field of catalytic kinetics.
1.1. HISTORY
All processes occur over a time ranging from femtosecond to billions of years. The same
holds for chemical and biochemical transformations. Kinetics (derived from the Greek
word κινητιχοζ meaning dissolution) is a science which investigates rates of processes.
Chemical kinetics is the study of reaction rates.
However complex a process is, it can be, in principle, divided into a number of
elementary processes which can be studied separately.
Chemical kinetics emerged as a branch of physical chemistry in the 1880s, with sem-
inal works of Harcourt and Esson demonstrating the dependence of reaction rates on the
concentrations of reactants. It was German scientist K. Wenzel who stated that the affin-
ity of solid materials towards a solvent is inversely proportional to dissolution time.
100 years after that Guldberg and Waage (Norway) formulated a law, which was later
coined the “law of mass action,” meaning that the reaction “forces” are proportional
to the product of the concentrations of the reactants.
When the rate of a certain process is measured, especially if it is of practical impor-
tance, a curious mind is always eager to know if it is possible to accelerate its velocity.
Moreover, one could even imagine a situation that for a system demonstrating complete
inertness, the introduction of a foreign substance could enhance the rate dramatically.
Conversion of starch to sugars in the presence of acids, combustion of hydrogen over
platinum, decomposition of hydrogen peroxide in alkaline and water solutions in the
presence of metals, etc., were critically summarized by Swedish scientist J.J. Berzelius
in 1836, who proposed the existence of a certain body, which “effecting the (chemical)
changes does not take part in the reaction and remains unaltered through the reaction.”
He called this unknown force catalytic force and defined catalysis as decomposition of
bodies by this force.
This new concept was immediately criticized by Liebig, as this notion was putting
catalysis somewhat outside other chemical disciplines. A catalyst was later defined by
Ostwald as a compound, which increases the rate of a chemical reaction, but which
is not consumed by the reaction. This definition allows for the possibility that small
amounts of the catalyst are lost in the reaction or that the catalytic activity is slowly lost.
1.2. CATALYSIS
Already from these definitions, it is clear that there is a direct link between chemical
kinetics and catalysis; according to the very definition of catalysis, it is a kinetic process.
There are different views, however, on the interrelation between kinetics and catalysis.
While some authors state that catalysis is a part of kinetics, others treat kinetics as a part of
a broader phenomenon of catalysis.
Despite the fact that catalysis is a kinetic phenomenon, there are many issues in catal-
ysis which are not related to kinetics. Mechanisms of catalytic reactions, elementary
reactions, surface reactivity, adsorption of reactants on the solid surfaces, synthesis
and structure of solid materials, enzymes or organometallic complexes, not to
mention engineering aspects of catalysis, are obviously outside the scope of chemical
kinetics.
Some discrepancy exists whether chemical kinetics include also the mechanisms of
reactions. In fact, if reaction mechanisms are included in the definition of catalytic kinet-
ics, it will be an unnecessary generalization, as catalysis should cover mechanisms.
Catalysis is of crucial importance for the chemical industry. The number of catalysts
applied in industry is very large, and catalysts come in many different forms, from het-
erogeneous catalysts in the form of porous solids to homogeneous catalysts dissolved in
the liquid reaction mixture to biological catalysts in the form of enzymes. Catalysis is a
multidisciplinary field requiring efforts of specialists in different fields of chemistry,
Setting the Scene 3
physics and biology to work together to achieve the goals set by mankind. Knowledge of
inorganic, organometallic, organic chemistry, materials and surface science, solid state
physics, spectroscopy, reaction engineering and enzymology is required for the advance-
ments of the discipline of catalysis.
Despite the fundamental differences between elementary steps in catalytic process on
surfaces, there are striking similarities also in terms of chemical kinetics with enzymes or
homogeneous organometallic complexes. Although superficially it is difficult to find
something in common between the reaction of nitrogen and hydrogen forming ammo-
nia on a surface of iron, D-fructose 6-phosphate reaction with ATP involving an enzyme
phosphofructokinase or ozone decomposition in the atmosphere in the presence of NOx,
all these transformations require that the catalyst forms bonds with the reacting molecules.
Such a complex then reacts to products, leaving the catalyst unaltered and ready for taking
part in a next catalytic cycle.
Fig. 1.1 is an example of a catalytic reaction between two molecules A and B with the
involvement of a catalyst.
In order to understand how a catalyst can accelerate a reaction, a potential energy
diagram should be considered.
Fig. 1.2 represents a concept for a non-catalytic reaction. Arrhenius suggested that
reactions should overcome a certain barrier before a reaction can proceed.
The change in the Gibbs free energy between the reactants and the products ΔG does
not change in case of a catalytic reaction; however, the catalyst provides an alternative
path for the reaction (Fig. 1.3).
START A
t
uc
od
Pr
Catalyst Reagent A
on to catalyst
Product
leaves
Catalyst
Catalyst
A
Product
Reaction B
occurs on catalyst Catalyst Reagent B
on to catalyst
B A
X‡
ΔG‡
G
A+B
Δ Greaction
P+Q
Reaction coordinate
Fig. 1.2 Potential energy diagram.
X‡
Uncatalyzed ‡
ΔΔGcat
(the reduction
in ΔG‡ by the
catalyst)
G
Catalyzed
A+B
P+Q
A + B GP + Q
Reaction coordinate
Fig. 1.3 Potential energy diagram for catalytic reactions.
In general, reaction rates increase with increasing temperature. Kooij and van’t Hoff
(1893) proposed an equation for the temperature dependence of reaction rates:
Ea
k ¼ AT m e =RT (1.1)
where A is pre-exponential factor and activation energy, Ea, is related to the potential
energy barrier. This equation, which could be derived on the basis of transition sate the-
ory, in a slightly simplified form
Ea
k ¼ ko e =RT (1.2)
Visit https://textbookfull.com
now to explore a rich
collection of eBooks, textbook
and enjoy exciting offers!
Setting the Scene 5
Concentration
A
Time
Fig. 1.4 Concentration versus time dependences for a reversible reaction.
was applied by Arrhenius and is referred to as the Arrhenius law. It is immediately clear
from Eq. (1.2) that a decrease in activation energy will lead to an increase of the rate con-
stant, thus the reaction rate (a discussion on the relationship between the rate and rate
constant will be given below).
At the same time, the catalyst (heterogeneous, homogeneous, or enzymatic) affects
only the rate of the reaction; it changes neither the thermodynamics of the reaction
(Gibbs energy) nor the equilibrium composition. An important conclusion is that a cat-
alyst can change kinetics but not thermodynamics of a reaction. If a process is thermo-
dynamically unfavorable, there is no need to apply any modern and fancy methods (high
throughput screening and the like) to find such catalyst.
The dashed line in Fig. 1.4 demonstrates the equilibrium that cannot be overcome for
a given set of parameters.
Furthermore, the ratio of rate constants in the forward and reverse direction for cat-
alytic and non-catalytic reactions is the same:
k kcat ½P eq
¼ ¼ ¼K (1.3)
k0 kcat 0 ½Aeq
It also implies that if a catalyst is active in enhancing a rate of the forward reaction, it will
do the same with a reverse reaction.
Fig. 1.3 is somewhat simplified, as it does not take into account possible bonding of
the catalyst and reactant. In order for a catalyst to be effective, the energy barrier between
the catalyst-substrate and activated complex must be less than that between substrate and
activated complex in the un-catalyzed reaction. The binding of substrate to an enzyme
lowers the free energy of the catalyst substrate complex relative to the substrate (Fig. 1.5).
This is a general feature of catalysis and is relevant for heterogeneous, homogeneous, and
enzymatic catalysis. If the energy is lowered too much without a greater lowering of the
activation energy, then catalysis would not take place, meaning that bonding between a
6 Catalytic Kinetics
Transition
state
Energy +
ES+
Substrates
ES
Products
Binding Catalysis EP
Reaction coordinate
Fig. 1.5 Potential energy diagram of an enzymatic reaction. (From https://upload.wikimedia.org/
wikipedia/commons/thumb/c/c0/Enzyme_catalysis_energy_levels_2.svg/1280px-Enzyme_catalysis_
energy_levels_2.svg.png?1450038342767).
catalyst and a reactant should not be too strong. Alternatively, if it is too weak, then the
catalytic cycle could not proceed.
Chemical kinetics as a discipline addresses how the reaction rates depend on reactant
concentration, temperature, nature of catalysts, pH, and solvent, to name a few reaction
parameters.
Chemical kinetics, together with other means of studying catalytic reactions like spec-
troscopy of catalysts and catalyst models, quantum-chemical calculations for reactants,
intermediates and products, calculation of the thermodynamics of reactants, intermedi-
ates and products from measured spectra and quantum-chemical calculations, form the
modern basis for understanding catalysis.
Kinetic investigations are one of the ways to reveal reaction mechanisms. The follow-
ing problems can be solved using the kinetic model:
• choosing the catalyst and comparing the selectivity and activity of catalysts and their
performance under optimum conditions for each catalyst;
• the determination of the optimum sizes and structure of catalyst grains and the neces-
sary amount of the catalyst to achieve the specified values of the selectivity of the pro-
cess and conversion of the starting products;
• the determination of the composition of all byproducts formed during the process;
• the determination of the stability of steady states and parametric sensitivity; that is, the
influence of deviations of all parameters on the steady-state regime and the behavior of
the reactor under unsteady state conditions;
• the study of the dynamics of the process and deciding if the process should be carried
out under unsteady-state conditions;
• the study of the influence of mass and heat transfer processes on the chemical reaction
rate and the determination of the kinetic region of the process;
• selection of the reactor type and structure of the contact unit providing the best
approximations to the optimum conditions.
Setting the Scene 7
Very often, the rates of chemical transformations are affected by the rates of other pro-
cesses, such as heat and mass transfer. The process should be treated as a part of kinetics.
The gas/liquid mass transfer in multiphase heterogeneous and homogeneous catalytic
reactions could be treated in a similar way. The mathematical framework for modeling
diffusion inside solid catalyst particles of supported metal catalysts or immobilized
enzymes does not differ that much, but proper care should be taken of the reaction
kinetics.
The immense importance of catalysis in chemical industry is manifested by the fact
that roughly 85–90% of all chemical products have seen a catalyst during the course
of production.
Fig. 1.6 demonstrates applications of catalysis in industry. In the last few years, there is
an increase of catalytic applications also for non-chemical industries, including treatment
of exhaust gases from cars and other mobile sources, as well as power plants (Fig. 1.7).
Chemical Chemical
1997 2003
Environmental Environmental
Billion US$ 7.4 9.0
(A) (B)
Fig. 1.7 Catalytic treatment of NOx in (A) mobile and (B) stationary sources.
8 Catalytic Kinetics
Homogeneous Heterogeneous
Activity High Variable
Selectivity High Variable
Conditions of reaction Mild Harsh
Lifetime of catalyst Variable Long
Sensitivity to deactivation Low High
Problems due to diffusion None Difficult to solve
Recycling of catalyst Usually difficult Can easily be done
Steric and electronic properties Easily changed No variation possible
Mechanism Realistic models exist Not obvious
The topics addressed above will be discussed in more detail in the subsequent chapters.
A great variety of homogeneous catalysts are known, including metal complexes and
ions, Brønsted and Lewis acids, and enzymes. Homogeneous transition metals are used in
several industrial processes, a few of them are given below:
World capacity
Process (million t/a) Catalyst Temperature (K) Pressure (bar)
Acetaldehyde 2.5 Pd/Cu 375–405 3–8
Acetic acid 4.0 Rh 425–475 30–60
Oxo-alcohols 7 Co or Rh 335–470 200/30
Dimethyl terephthalate 3.3 Co 415–445 4–8
Terephthalic acid 9.4 Co 450–505 15–30
Metal complexes can have a very sophisticated structure with a variety of ligands. An
example of such ligands for Rh catalyzed hydroformylation is given below (Fig. 1.8)
along with some images of heterogeneous catalysts (Fig. 1.9).
Enzymes represent a special type of homogeneous catalyst. They are large proteins
(Fig. 1.10) capable of increasing the reaction rates by a factor of 106 at mild reaction con-
ditions and displaying very high specificity and capability of regulation.
Specificity (Fig. 1.11) is controlled by the enzyme structure; more precisely, a unique
fit of substrate with the enzyme that controls the selectivity for the substrate and the prod-
uct yield.
Superficially, there is not that much in common between a large protein and a Pt/
Al2O3 heterogeneous catalyst. At the same time, the chemical reactions which occur with
both types of catalysts involve certain active sites; for example, regions where catalysis
occurs. Whatever the specific reaction, it can be schematically represented by
Fig. 1.8 A ligand for Rh catalyzed hydroformylation.
Active center
of the enzyme
(His - 57)
(Asp - 102)
O (Ser - 195)
– H N N H O
O
O
R1 N C
H R2
Fig. 1.5. This in turn means that the kinetics of either heterogeneous or homogeneous
catalytic reactions can be very similar, and in fact it is.
where A and B are reactants, C and D products, and a, b, c, and d are stoichiometric coef-
ficients. An equation for a chemical reaction is written in such a way that all the molecules
participating in the reaction are balanced.
Very often in chemical reaction engineering, the stoichiometric coefficient νi is
defined as the amount of product produced after one run of the reaction. It implies that
the stoichiometric coefficient is positive for a product and negative for a reactant.
Thus, for the reaction:
A + B)C (1.5)
The following stoichiometric coefficients hold:
vA ¼ 1, vB ¼ 1, vC ¼ + 1 (1.6)
An extensive quantity describing the progress of a chemical reaction equal to the number
of chemical transformations (the total number of reaction runs) divided by Avogadro’s
number (it is essentially the amount of chemical transformations) is called the extent of
reaction. The change in the extent of reaction is given by dξ ¼ dni/νi, where νi is the
stoichiometric number of any reaction entity i (reactant or product) and ni is the correspond-
ing amount in moles.
Thus, dξ/dt is an extensive property, which is measured in moles and cannot be con-
sidered a reaction rate, as it is proportional to the size of the reactor.
In general, for a homogeneous reaction for which both the reaction rate changes with
time and is not uniform over a volume of a reactor, the reaction rate is:
@2ξ
r¼ (1.7)
@t@V
where V stands for the reactor volume. If the reactor volume is constant, then the reac-
tion rate is simply:
1 dCi
ri ¼ (1.8)
νi dt
where i is the reactant or product with corresponding stoichiometric coefficient νi, and Ci
is the concentration of component i. For a reaction:
A + B)C (1.9)
the rate of consumption of reactant A is then:
1 dnA dCA d ½A
rA ¼ ¼ ¼ (1.10)
νA Vdt dt dt
where nA is the number of moles of A in the reactor and [A] is the concentration of A.
Similarly, for the reaction:
2NOCl ðgÞ ) 2 NOðgÞ + 1 Cl2 ðgÞ (1.11)
12 Catalytic Kinetics
2 mol of NOCl disappear for every 1 mol Cl2 formed; therefore, the rate is defined as:
1 d ½NOCl 1 d ½NO d½Cl2
rate ¼ ¼ + ¼ + (1.12)
2 dt 2 dt dt
For a heterogeneous reaction occurring on the surface S of a catalyst, the following
expression holds:
@2ξ
r¼ (1.13)
@t@S
which, if the rate is uniform across the surface, can be further simplified to:
1 @ξ
r¼ (1.14)
S @t
Rate laws express how the rate depends on concentration. If a reaction follows Eq. (1.4),
the law of mass action could be applied leading to a following equation:
r ¼ r + r ¼ k + ½Aa ½Bb k ½C c ½Dd (1.15)
where k+ and k are reaction rate constants and stoichiometric coefficients appear as the
powers (reaction orders towards particular components).
In reality, the chemical equation (1.4) does not tell us how reactants become products;
it is a summary of the overall process. In fact, it is molecularity; for example, the number
of species that must collide to produce the reaction which determines the form of a rate
equation. Reactions whose rate laws can be written from their molecularity are called
elementary. The kinetics of the elementary step depend only on the number of reactant
molecules in that step.
For the reaction:
2NO2 ðgÞ + Cl2 ðgÞ ) 2 NO2 Cl ðgÞ (1.16)
the rate expression based on the formal kinetics is:
r ¼ k ½NO2 2 ½Cl2 (1.17)
with the overall order defined as the sum of orders to each reactant being equal to 3.
However, the reaction mechanism is more complicated and consists of several elemen-
tary steps:
ðaÞ NO2 ðgÞ + Cl2 ðgÞ ) NO2 ClðgÞ + ClðgÞ
(1.18)
ðbÞ NO2 ðgÞ + ClðgÞ ) NO2 ClðgÞ
If the rate of the overall process is determined by the first step a, then the rate is defined as:
r ¼ k ½NO2 ½Cl2 (1.19)
and the overall order is just two.
Setting the Scene 13
For elementary reactions, the reaction orders have orders that are integers, which
are usually equal to one or two (Fig. 1.12), and occasionally three for trimolecular
reactions.
In practice, reaction orders can be fractional, indicating a complex reaction mecha-
nism. The majority of this book is devoted to catalytic reaction mechanisms, which are
typical examples of complex reactions.
Reaction orders for a reaction A ) P described by a following equation for
the rate:
dcA
¼ rA ¼ kcAn (1.20)
dt
are determined using logarithmic plots:
dcA
ln ¼ ln k + n lncA (1.21)
dt
slope = n
0
0 ln c
Fig. 1.13 Determination of reaction orders.
14 Catalytic Kinetics
Fig. 1.17 Batch reactors: (A) glass, (B) high pressure, and (C) in parallel mode. (From http://
www.parrinst.com/wp-content/uploads/2011/03/MRS-5000_burettes.jpg. © Parr Instrument Company).
16 Catalytic Kinetics
V
Flow
Flow
nA ṅA
ṅ0A
For an infinitesimal volume element ΔV in Fig. 1.18, the mass balance could be writ-
ten in a form:
IN + GENERATION ¼ OUT + ACCUMULATION
leading to a following equation in terms of moles:
dnA
n_0A + ηA rA ΔV ¼ n_A + (1.22)
dt
where n_0A and n_A are the mole fluxes, ηA is the catalyst effectiveness factor (Chapter 10)
taking into account mass transfer.
For a batch reactor it holds that IN ¼ 0, OUT ¼ 0; therefore:
dnA
ηA rA V ¼ (1.23)
dt
If the volume is constant, one gets:
dnA dðcA V Þ dcA
¼ ¼V (1.24)
dt dt dt
From Eqs. (1.23) and (1.24):
dcA
¼ ηA rA (1.25)
dt
making use of the relationship between concentration and conversion α:
CA ¼ CAo ð1 αÞ (1.26)
Assuming that the catalyst effectiveness factor ηA is equal to 1 and taking into account
boundary conditions (t ¼ 0, α ¼ 0), we arrive at:
ðα
dα
t ¼ CAo (1.27)
rA
0
Inserting the expression of reaction rate in Eq. (1.25) for a reaction A ) P, which
occurs over a catalyst:
dCA
¼ ηkCAn (1.28)
dt
and assuming that n ¼ 1 and ηA ¼ 1, we arrive at:
dCA
¼ kCA0 ð1 αÞ (1.29)
dt
which after integration gives an expression for reaction time:
ðα
1 dα 1
t¼ ¼ ln ð1 αÞ (1.30)
k 1α k
0
For the first order, reaction k has units of s1 (the rates are given in mol l1 s1).
For a second order, reaction with ηA ¼ 1 instead of Eq. (1.30) one arrives at:
1 1 1
t¼ (1.31)
k CA0 CA
with k in l/mol s, and for a zero order reaction, units of k are mol/l s and:
1 0
t¼ CA CA (1.32)
k
Eqs. (1.30)–(1.32) can be applied for batch reactors independent of the type of catalysis
that is operative, if reactions could be described by zero, first or second order. More com-
plicated cases for Langmuir kinetics or Michaelis-Menten kinetics will be considered
further.
1.4.2 CSTR
Examples of continuous stirred tank reactors are presented in Fig. 1.14 (right). Such a
system can be applied for both homogeneous and heterogeneous reactions. Fig. 1.19
illustrates the differences between batch and CSTR reactors.
For a perfectly mixed CSTR at steady state, it holds that there is no accumulation:
dnA
¼0 (1.33)
dt
therefore:
n_0A n_A + ηA rA V ¼ 0 (1.34)
18 Catalytic Kinetics
Substrate
Product
Immobilized Immobilized
catalyst
(A) (B) catalyst
Fig. 1.19 Stirred tank reactors in (A) batch and (B) continuous mode.
reactors are efficient for catalyst screening, especially when they are arranged in a parallel
mode (Fig. 1.21).
The apparent drawback is that one experiment leads to only one data point. On the
other hand, catalyst deactivation with time on stream could be easily seen (Fig. 1.22).
Quantitative treatment of plug flow reactors is somewhat cumbersome; therefore,
several assumptions are usually made. The fluid composition is considered to be uniform
along the reactor cross section (ie, there is no radial dispersion). This is valid only when:
dcat 1
< (1.41)
dreact 30