Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Catalytic Kinetics: Chemistry and Engineering
Catalytic Kinetics: Chemistry and Engineering
Catalytic Kinetics: Chemistry and Engineering
Ebook1,510 pages9 hours

Catalytic Kinetics: Chemistry and Engineering

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Catalytic Kinetics: Chemistry and Engineering, Second Edition offers a unified view that homogeneous, heterogeneous, and enzymatic catalysis form the cornerstone of practical catalysis.

The book has an integrated, cross-disciplinary approach to kinetics and transport phenomena in catalysis, but still recognizes the fundamental differences between different types of catalysis. In addition, the book focuses on a quantitative chemical understanding and links the mathematical approach to kinetics with chemistry.

A diverse group of catalysts is covered, including catalysis by acids, organometallic complexes, solid inorganic materials, and enzymes, and this fully updated second edition contains a new chapter on the concepts of cascade catalysis. Finally, expanded content in this edition provides more in-depth discussion, including topics such as organocatalysis, enzymatic kinetics, nonlinear dynamics, solvent effects, nanokinetics, and kinetic isotope effects.

  • Fully revised and expanded, providing the latest developments in catalytic kinetics
  • Bridges the gaps that exist between hetero-, homo- and enzymatic-catalysis
  • Provides necessary tools and new concepts for researchers already working in the field of catalytic kinetics
  • Written by internationally-renowned experts in the field
  • Examples and exercises following each chapter make it suitable as an advanced course book
LanguageEnglish
Release dateJun 4, 2016
ISBN9780444634634
Catalytic Kinetics: Chemistry and Engineering
Author

Dmitry Yu Murzin

Dmitry Murzin is Professor of Chemical Technology at Åbo Akademi University in Turku, Finland, and for the past 10 years he has served as Professor at the Finland University's Laboratory of Industrial Chemistry and Reaction Engineering. Previously, he headed the chemicals division of BASF Corporation in Moscow. He also spent seven years as a researcher in the Department of Catalysis at the Karpov Institute of Physical Chemistry, Moscow. From 2009–2013 Murzin served as Vice President of the European Federation of Catalysis Societies. Murzin is author or co-author of more than 650 journal articles and book chapters.

Related to Catalytic Kinetics

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Catalytic Kinetics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Catalytic Kinetics - Dmitry Yu Murzin

    Catalytic Kinetics

    Chemistry and Engineering

    Second Edition

    Dmitry Yu. Murzin

    Tapio Salmi

    Table of Contents

    Cover image

    Title page

    Copyright

    About the Authors

    Dmitry Yu. Murzin

    Tapio Salmi

    Preface

    Chapter 1: Setting the Scene

    Abstract

    1.1 History

    1.2 Catalysis

    1.3 Formal Kinetics

    1.4 Acquisition of Kinetic Data

    1.5 Kinetics and Thermodynamics

    1.6 Examples and Exercises

    Chapter 2: Catalysis

    Abstract

    2.1 Homogeneous Catalysis

    2.2 Heterogeneous Catalysis

    2.3 Organocatalysis

    2.4 Examples and Exercises

    Chapter 3: Elementary Reactions

    Abstract

    3.1 Reaction Rate Theory

    3.2 Elementary Reactions in Solutions

    3.3 Reaction Mechanism

    3.4 Quasi-equilibrium Approximation

    3.5 Relationship Between Thermodynamics and Kinetics

    3.6 Transition State Theory of Surface Reactions

    3.7 Rates of Reactions on Nonideal Surfaces

    3.8 Deterministic and Stochastic Models

    3.9 Microkinetic Modeling

    3.10 Compensation Effect

    3.11 Isotope Effects

    3.12 Examples and Exercises

    Chapter 4: Complex Reactions

    Abstract

    4.1 Steady State Kinetics of Complex Reactions

    4.2 Basic Routes of Complex Reactions

    4.3 Single-Route Steady-State Reaction

    4.4 Topological Analysis of Complex Reactions

    4.5 Electrical Analogy of Reaction Networks

    4.6 Thermodynamic Consistency of Rate Constants for Complex Networks

    4.7 Kinetic Aspects of Selectivity

    4.8 Parallel Reactions: Kinetic Coupling

    4.9 Reduction of Complexity

    4.10 Polynomial Kinetics

    4.11 Examples and Exercises

    Chapter 5: Homogeneous Catalytic Kinetics

    Abstract

    5.1 Homogeneous Acid-Base Catalysis

    5.2 Nucleophilic Catalysis

    5.3 Catalysis by Metal Ions

    5.4 Catalysis by Organometallic Complexes

    5.5 Organocatalysis

    5.6 Polymerization Catalysis

    5.7 Examples and Exercises

    Chapter 6: Enzymatic Kinetics

    Abstract

    6.1 Enzymatic Catalysis

    6.2 Cooperative Kinetics

    6.3 Inhibition

    6.4 Effects of pH

    6.5 Single Molecule Enzymology

    6.6 Enantioselectivity in Enzyme Catalyzed Reactions

    6.7 Generalized Rate Laws for Enzymatic Reactions

    6.8 Heterogeneous Systems/Immobilized Enzymes

    6.9 Examples and Exercises

    Chapter 7: Heterogeneous Catalytic Kinetics

    Abstract

    7.1 Reactions on Ideal Surfaces

    7.2 Reactions on Nonideal Surfaces

    7.3 Selectivity

    7.4 Polyatomic Nature of Reactants and Coverage-Dependent Adsorption Mode

    7.5 Solvent Effects

    7.6 Ionic Species

    7.7 Transfer of Labeled Atoms in Heterogeneous Catalytic Reactions

    7.8 Electrocatalytic Kinetics

    7.9 Photocatalytic Kinetics

    7.10 Nanokinetics

    7.11 Examples and Exercises

    Chapter 8: Kinetics of Catalytic Reactions With Multiple/Multifunctional Catalysts

    Abstract

    8.1 General

    8.2 Combined Catalytic and Noncatalytic Reactions

    8.3 Multiple Catalysts of the Same Type

    8.4 Multiple Catalysts of Different Types

    8.5 Examples and Exercises

    Chapter 9: Dynamic Catalysis

    Abstract

    9.1 Transient Kinetics

    9.2 Relaxation Methods

    9.3 Temperature-Programmed Desorption

    9.4 Oscillations

    9.5 Dynamic Catalyst Changes

    9.6 Examples and Exercises

    Chapter 10: Mass Transfer and Catalytic Reactions

    Abstract

    10.1 Catalytic Multi-Phase Systems

    10.2 Simultaneous Reaction and Diffusion in Fluid Films and in Porous Materials

    10.3 Liquid-Liquid Diffusion, Phase-Transfer Catalysis

    10.4 Catalytic Two-Phase Systems

    10.5 Mass Transfer and Enzymatic Kinetics

    10.6 External Mass Transfer

    10.7 Internal Diffusion and Selectivity

    10.8 Internal Diffusion and Deactivation

    10.9 Elucidation of the Impact of Mass Transfer

    10.10 Three-Phase Systems

    10.11 Examples and Exercises

    Chapter 11: Kinetic Modeling

    Abstract

    11.1 Basic Principles

    11.2 Heuristic Design of Experiments

    11.3 Parameter Estimation: Classical Methods

    11.4 Parameter Estimation: Regression

    11.5 Numerical Strategies

    11.6 Analysis of Parameters

    11.7 Model Discrimination

    11.8 Software

    11.9 Case Studies

    Index

    Copyright

    Elsevier

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, USA

    © 2016, 2005 Elsevier B.V. All rights reserved.

    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.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

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

    ISBN: 978-0-444-63753-6

    For information on all Elsevier publications visit our website at https://www.elsevier.com/

    Publisher: John Fedor

    Acquisitions Editor: Anita Koch

    Editorial Project Manager: Amy Clark

    Production Project Manager: Anitha Sivaraj

    Cover Designer: Maria Inês Cruz

    Typeset by SPi Global, India

    About the Authors

    Tapio Salmi and Dmitry Yu. Murzin

    Dmitry Yu. Murzin

    Education: 1986: MSc (Chemical Engineering) with honors, Mendeleev University of Chemical Technology, Moscow; 1989: PhD, Karpov Physico-Chemical Institute, Moscow; 1999: Doctor of Chemical Sciences, Karpov Physico-Chemical Institute.

    Career: 1986–1992: Karpov Physico-Chemical Institute; 1992–1994: postdoctoral experience (Université Louis Pasteur, Strasbourg, France; Åbo Akademi University, Turku, Finland); 1995–2000: with BASF; since 2000: Professor in Chemical Technology, Åbo Akademi University, Turku, Finland.

    Editorial board member of several journals in catalysis and chemical engineering field.

    Vice-president of the European Federation of Catalysis Societies (2009–2013).

    An author of textbooks (Engineering Catalysis, DeGruyter, 2013; and Chemical Reaction Technology, DeGruyter, 2015) and author and co-author of over 650 journal articles on kinetics and catalysis.

    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.

    Preface

    Diversity of catalysts, for example, catalysis by mineral acids, complexes of transition metals, organic molecules, enzymes, and alumosilicates, resulted in fragmentation of the discipline. Separate courses are given based on textbooks covering separately, despite all similarities, homogeneous, heterogeneous, and enzymatic catalysis. More general books on physical chemistry, while discussing some aspects of catalytic kinetics, cannot go into details to the extent that is often needed.

    Although in the chemical industry, the main focus is currently on heterogeneous catalysis, other fields (homogeneous, enzymatic, photochemical, electrochemical catalysis) should not be overlooked, as very often the industrial aim is to produce a certain chemical in the most optimal way, but not with a particular type of catalyst. The industrial importance of catalysis also dictates a necessity to have a book devoted to the fundamental aspects of catalytic kinetics, as well as catalytic engineering and, more specifically, transport phenomena.

    Such an integrated approach to kinetics and transport phenomena in catalysis, while still recognizing the fundamental differences among different types of catalysis, was applied in the first edition of this book, published in 2005.

    The authors have received a number of encouraging comments as well as critique about what was missing and suggestions for how the book could be improved. We were also very happy to learn that due to a positive response, Elsevier decided to proceed with the second edition of this book on catalytic kinetics. The authors carefully discussed the material to be included in the revised version. Material for chapters 10.1–10.2, 10.4, and 11.3–11.6 was prepared by Tapio Salmi, while the remaining text was written by Dmitry Yu. Murzin, who also oversaw the second edition and implementation of all changes.

    In this edition, the content was expanded to cover in more depth several areas, such as organocatalysis; enzymatic kinetics; nonlinear dynamics; solvent effects; nanokinetics (structure sensitivity); kinetic isotope effects; and polynomial kinetics, to name a few. In addition, a separate chapter on cascade catalysis has been written.

    Each and every chapter in the second edition also contains examples and exercises, which facilitates understanding and mastering of the underlying concepts.

    In order to stay more focused on the core of catalytic kinetics, a section on gas-liquid mass transfer, which was present in the first edition, was not included.

    In addition to these changes, unfortunate misprints were corrected and general editing of the text, introducing new figures and examples, was performed.

    Similar to the previous edition, the authors limited themselves to the main monographs, 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.

    Dmitry Yu. Murzin, Turku/Åbo

    November 2015

    Chapter 1

    Setting the Scene

    Abstract

    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 heterogeneous 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, physics and biology to work together to achieve the goals set by mankind. There is a direct link between chemical kinetics and catalysis; according to the very definition of catalysis, it is a kinetic process. Despite the fundamental differences between elementary steps in catalytic process on surfaces with enzymes or homogeneous organometallic complexes, there are striking similarities in terms of chemical kinetics. A general overview of different types of catalysis is presented and discussed in this chapter. Basic definitions related to formal kinetics and acquisition of kinetic data are introduced. Different kinds of chemical reactors along with respective mass balances are presented. Finally, a relationship between kinetics and thermodynamics is discussed.

    Keywords

    Catalysis; Enzymes; Homogeneous; Heterogeneous; Formal kinetics; Reactors; Thermodynamics

    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 seminal 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 affinity 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 importance, 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.

    Jöns Jakob Berzelius

    Wilhelm Ostwald and Svante Arrhenius

    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 catalysis 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 kinetics, 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 heterogeneous 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, 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 advancements 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 ammonia 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.

    Fig. 1.1 Catalytic cycle.

    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.

    Fig. 1.2 Potential energy diagram.

    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).

    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:

       (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 theory, in a slightly simplified form

       (1.2)

    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 constant, 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 catalyst can change kinetics but not thermodynamics of a reaction. If a process is thermodynamically 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.

    Fig. 1.4 Concentration versus time dependences for a reversible reaction.

    Furthermore, the ratio of rate constants in the forward and reverse direction for catalytic and non-catalytic reactions is the same:

       (1.3)

    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 catalyst and a reactant should not be too strong. Alternatively, if it is too weak, then the catalytic cycle could not proceed.

    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).

    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 spectroscopy of catalysts and catalyst models, quantum-chemical calculations for reactants, intermediates and products, calculation of the thermodynamics of reactants, intermediates 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 following 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 necessary amount of the catalyst to achieve the specified values of the selectivity of the process 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.

    Very often, the rates of chemical transformations are affected by the rates of other processes, 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).

    Fig. 1.6 Worldwide catalyst market.

    Fig. 1.7 Catalytic treatment of NO x in (A) mobile and (B) stationary sources.

    A comparison between homogeneous and heterogeneous catalysts from the viewpoint of a homogeneous catalysis expert is presented below:

    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:

    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).

    Fig. 1.8 A ligand for Rh catalyzed hydroformylation.

    Fig. 1.9 Images of heterogeneous catalysts.

    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 conditions and displaying very high specificity and capability of regulation.

    Fig. 1.10 A schematic view on an enzyme structure. (From https://upload.wikimedia.org/wikipedia/commons/b/b4/Protein_ACE2_PDB_1r42.png).

    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 product yield.

    Fig. 1.11 Specificity of enzyme catalysis. (From http://csls-text3.c.u-tokyo.ac.jp/large_fig/fig04_05.html. Copyright © 2011 Division of Life Sciences, Komaba Organization for Educational Excellence, College of Arts and Sciences, The University of Tokyo).

    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. 1.5. This in turn means that the kinetics of either heterogeneous or homogeneous catalytic reactions can be very similar, and in fact it is.

    1.3 Formal Kinetics

    Chemical kinetics as a discipline concerns the rates (the velocities) of chemical reactions and deals with experimental measurements of the velocities in batch, semibatch or continuous reactors.

    Interpretation of the experimental data is currently done using the laws of physical chemistry.

    One of the fathers of chemical kinetics, Louis Jacques Thénard, discovered hydrogen peroxide and measured its decomposition rates. He demonstrated for the first time that rates of chemical reactions varied with the concentrations of the reactants. In later study, Ludwig Ferdinand Wilhelmy investigated the inversion of cane sugar in the presence of acids and developed a rate equation, which was the first attempt to interpret the temperature dependence of the rate constant. Unfortunately, this work remained in oblivion until 1884. In 1865, the rate laws combined with mass balances for a batch reactor were proposed by Augustus George Vernon Harcourt and William Esson, giving a mathematical expression for concentration versus t for first order, second order and consecutive reactions, representing a major breakthrough for modern chemical kinetics.

    Following the footsteps of the great scientists of the 19th century, let us try to consider reaction rates for a chemical reaction described by the following equation:

       (1.4)

    where A and B are reactants, C and D products, and a, b, c, and d are stoichiometric coefficients. 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:

       (1.5)

    The following stoichiometric coefficients hold:

       (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 = dni/νi, where νi is the stoichiometric number of any reaction entity i (reactant or product) and ni is the corresponding amount in moles.

    Thus, dξ/dt is an extensive property, which is measured in moles and cannot be considered 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:

       (1.7)

    where V stands for the reactor volume. If the reactor volume is constant, then the reaction rate is simply:

       (1.8)

    where i is the reactant or product with corresponding stoichiometric coefficient νi, and Ci is the concentration of component i. For a reaction:

       (1.9)

    the rate of consumption of reactant A is then:

       (1.10)

    where nA is the number of moles of A in the reactor and [A] is the concentration of A. Similarly, for the reaction:

       (1.11)

    2 mol of NOCl disappear for every 1 mol Cl2 formed; therefore, the rate is defined as:

       (1.12)

    For a heterogeneous reaction occurring on the surface S of a catalyst, the following expression holds:

       (1.13)

    which, if the rate is uniform across the surface, can be further simplified to:

       (1.14)

    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:

       (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:

       (1.16)

    the rate expression based on the formal kinetics is:

       (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 elementary steps:

       (1.18)

    If the rate of the overall process is determined by the first step a, then the rate is defined as:

       (1.19)

    and the overall order is just two.

    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.

    Fig. 1.12 Representation of reaction kinetics of different orders.

    In practice, reaction orders can be fractional, indicating a complex reaction mechanism. 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:

       (1.20)

    are determined using logarithmic plots:

       (1.21)

    with the reaction order corresponding to the slope (Fig. 1.13).

    Fig. 1.13 Determination of reaction orders.

    1.4 Acquisition of Kinetic Data

    Kinetic data for a chemical reaction are gathered in different type of reactors, and we will briefly mention some requirements for chemical reactors from the viewpoint of kinetic analysis. A high precision of the data is needed as large deviations in the values of the experimentally measured rates will be a serious obstacle for quantitative considerations. Reproducibility of rate measurements over a broad range of parameters is also of importance. Another necessary feature is the possibility to reach a goal of obtaining the maximum amount of kinetic information in minimum time. Analysis of products as well as reactor layout should preferably be as easy as possible.

    Essential features for catalytic reactions are the readiness in reduction/activation of heterogeneous catalysts and a possibility to utilize them in the needed geometrical form. Despite the strict definition of catalysis, which states that the catalyst does not change during the catalytic reactions, some activity deterioration takes place; therefore, measurements of catalytic kinetics should always monitor the catalyst activity.

    Different types of reactors are applied in practice (Fig. 1.14). Stirred tank reactors (STR), very often applied for homogeneous, enzymatic and multiphase heterogeneous catalytic reactions, can be operated batch-wise (batch reactor, BR), semi-batch-wise (semi-batch reactor, SBR) or continuously (continuous stirred tank reactor, CSTR).

    Fig. 1.14 Different types of stirred tank reactors.

    Alternatively, tubular reactors with plug flow (piston flow) (PFR) are used and operate in continuous mode (Fig. 1.15).

    Fig. 1.15 Tubular reactors.

    1.4.1 Batch Reactors

    The batch mode of operation (Fig. 1.16) brings several advantages, as it allows us to monitor the progress of the reaction over time and thus acquires the whole kinetic curve in one experiment.

    Fig. 1.16 Approach to kinetic analysis in batch reactors.

    The high precision and wide range of parameters afforded by this operation mode made batch reactors very popular for kinetic studies, especially in the field of fine and pharmaceutical chemicals. Another advantage is the possibility to utilize heterogeneous catalysts of different geometrical shapes.

    Such reactors can be made either of glass (Fig. 1.17A) or stainless steel to sustain high pressures (Fig. 1.17B) and can be applied in a parallel mode (Fig. 1.17C).

    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).

    At the same time, activity control presents a challenge for heterogeneous catalytic reactions and will be discussed further in Chapter 9. Moreover, catalyst pretreatment (reduction) and regeneration are not straightforward. Quantitative treatment is not easy and will be briefly discussed below.

    For an infinitesimal volume element ΔV in Fig. 1.18, the mass balance could be written in a form:

    leading to a following equation in terms of moles:

       (1.22)

    are the mole fluxes, ηA is the catalyst effectiveness factor (Chapter 10) taking into account mass transfer.

    Fig. 1.18 A volume element.

    For a batch reactor it holds that IN = 0, OUT = 0; therefore:

       (1.23)

    If the volume is constant, one gets:

       (1.24)

    From Eqs. (1.23) and (1.24):

       (1.25)

    making use of the relationship between concentration and conversion α:

       (1.26)

    Assuming that the catalyst effectiveness factor ηA is equal to 1 and taking into account boundary conditions (t = 0, α = 0), we arrive at:

       (1.27)

    In fact, treatment of heterogeneous, homogeneous and enzymatic reactions is basically the same with the only difference in the expressions of reaction rates, which reflect different reaction mechanisms. Some specific cases will be discussed in Chapters 5–7. Here we present few examples:

    Inserting the expression of reaction rate in Eq. (1.25) for a reaction A ⇒ P, which occurs over a catalyst:

       (1.28)

    and assuming that n = 1 and ηA = 1, we arrive at:

       (1.29)

    which after integration gives an expression for reaction time:

       (1.30)

    For the first order, reaction k has units of s− 1 (the rates are given in mol l− 1 s− 1).

    For a second order, reaction with ηA = 1 instead of Eq. (1.30) one arrives at:

       (1.31)

    with k in l/mol s, and for a zero order reaction, units of k are mol/l s and:

       (1.32)

    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 complicated 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.

    Fig. 1.19 Stirred tank reactors in (A) batch and (B) continuous mode.

    For a perfectly mixed CSTR at steady state, it holds that there is no accumulation:

       (1.33)

    therefore:

       (1.34)

    Substituting concentration and volumetric flows for molar flows:

       (1.35)

    and assuming that volumetric flows in and out are equal (eg, density is constant):

       (1.36)

    and introducing residence time as the ratio of reactor volume to volumetric flow rate:

       (1.37)

    we arrive at:

       (1.38)

    which can be further transformed to:

       (1.39)

    In the case of a first order reaction, the expression for the residence time becomes:

       (1.40)

    1.4.3 Plug Flow Reactors

    An example of a tubular reactor is presented in Fig. 1.20.

    Fig. 1.20 A tubular reactor.

    For heterogeneous catalysis, the catalyst is packed in such reactors, which are easy to design and control, as the gases or liquids pass through the reactor and are analyzed. Such reactors are efficient for catalyst screening, especially when they are arranged in a parallel mode (Fig. 1.21).

    Fig. 1.21 Multitubular reactor.

    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).

    Fig. 1.22 Catalyst deactivation with time-on-stream at different conditions.

    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:

       (1.41)

    For larger catalyst particles, some gas bypass close to the reactor walls is possible. When the reactants move along the reactor, the composition of the mixture changes and axial diffusion can become prominent. To avoid the possible impact of axial diffusion in kinetic experiments, the level of conversion should be below 10–15%.

    In general, non-isothermal behavior of a tubular reactor should also be taken into account, since so-called isothermal reactors are seldom isothermal (Fig. 1.23); however, for the sake of simplicity it will not be considered below.

    Fig. 1.23 Temperature profile in a tubular reactor.

    The mass balance for a plug flow reactor (Fig. 1.23) is then given by:

       (1.42)

    where

       (1.43)

    Replacing the amount in moles through concentration and volume and considering an infinitely small volume:

       (1.44)

    we arrive at:

       (1.45)

    Further replacements at the constant volumetric flow:

       (1.46)

       (1.47)

    lead to:

       (1.48)

    , which means that the concentration is independent of time; therefore:

       (1.49)

    and

       (1.50)

    which could be transformed to:

       (1.51)

    For the first order reaction, finally, we obtain analogously to a batch reactor:

       (1.52)

    .

    For packed bed reactors (Fig. 1.24), the balance is exactly the same as for homogeneous PFR, except that the rate in Eq. (1.48) is replaced by rAρB, where ρB is the catalyst bulk density, giving:

       (1.53)

    Fig. 1.24 Molar flows in a packed bed reactor.

    , the design equation is:

       (1.54)

    1.4.4 Gradient-Free Recycle Reactors

    Such reactors (applied for gas and liquid systems) are stirred flow or recirculation reactors, characterized ideally by very small concentration and temperature gradients within the catalyst region.

    For a recirculation reactor with an external mixing device (Fig. 1.25), the pressure in the system is constant and gas (liquid) composition does not change with time or in space, and each single pass of reactants through the bed results in very small conversion, as the circulation rate is very high. The mixture of gases is fed through a pipe 1 to the reactor 4 and then to the outlet 2 for further analysis (GC). Pump 3 provided the necessary circulation of the mixture.

    Fig. 1.25 A recirculation reactor with an external mixing device.

    The system is then free from concentration gradients along the catalyst bed, concentration gradients due to axial dispersion and temperature gradients. The treatment of data is simplified as the rate is extracted directly in the differential form:

       (1.55)

    Efficient circulation pumps are needed and the inventory of reactants should be sufficient enough. The relaxation time (eg, time to reach steady state) can be significant. These disadvantages are, however, compensated by the easy mathematical treatment of experimental data.

    This reactor system can operate in a closed mode as well. Then, the entire kinetic curve can be measured in one experiment. At the same time, by-products can accumulate in the cycle.

    The concept of a circulating flow reactor was further developed in the Buss reactor technology (Fig. 1.26). Large quantities of reaction gas are introduced via a mixer to create a well dispersed mixture. This mixture is rapidly circulated by a special pump at high gas/liquid ratios throughout the volume of the loop and permits the maximum possible mass transfer rates. A heat exchanger in the external loop allows for independent optimization of heat transfer. For continuous operation, the product is separated by an in-line, cross-flow filter, which retains the suspended solid catalyst within the loop. Such a system can operate in batch, semi-continuous and continuous mode.

    Fig. 1.26 The Buss loop reactor. (From http://www.buss-t.com/up/files/PDFs_RT/Brochure_RT_Advanced_Loop_Reactor_operating_principle.pdf © 2009 Buss ChemTech AG).

    Some other types of gradientless reactors with stirred flow are the Berty reactor, which circulates gas through a stationary catalyst bed (Fig. 1.27) and a spinning basket reactor (Fig. 1.28). In the latter case, the solid catalyst is retained in a spinning, woven wire mesh basket to allow gas-liquid circulation with low pressure drop. The large catalyst particles required for this operation mode often lead to conducting the reaction in the region of external diffusion.

    Fig. 1.27 Berty gradientless reactors. (From http://www.autoclaveengineers.com/products/catalytic_reactors/berty_stationary_basket/images/L_bertyStatBasketCatReac.gif 2015 © Parker Autoclave Engineers).

    Fig. 1.28 Rotating basket. (From http://www.parrinst.com/products/stirred-reactors/options-accessories/catalyst-baskets/. © Parr Instrument Company).

    By utilizing structured catalysts, these problems can be avoided. Active carbon cloths, for instance, could be used in combination with a propeller stirrer, which serves at the same time as a catalyst holder (Fig. 1.29) and is coupled to a Rushton turbine for effective gas distribution.

    Fig. 1.29 Integrated gadget of a combined stirrer and catalyst holder.

    1.5 Kinetics and Thermodynamics

    The rates of forward and reverse reactions are related through thermodynamics. While thermodynamics tells us the final product distribution for a given set of reactants concentrations, pressures and temperature, kinetics describes how fast the products will be generated. Expressing the rates of the forward reaction:

       (1.56)

    and the reverse:

       (1.57)

    it follows for the ratio of rates, that:

       (1.58)

    The equilibrium constant is defined similar to Eq. (1.3):

       (1.59)

    leading to the De Donder equation:

       (1.60)

    where A is the reaction affinity, expressed by:

       (1.61)

    The affinity A is defined as the derivative of the Gibbs free energy with respect to the extent of the reaction.

       (1.62)

    At equilibrium, the Gibbs free energy is at its minimum value and:

       (1.63)

    where ν is the stoichiometric coefficient for species and μ is the chemical potential of this species. According to De Donder formulation, the relationship between the rates of forward and reverse reactions is expressed by the following equation:

       (1.64)

    or:

       (1.65)

    where r is the overall rate, r+ is the rate of the forward reaction, φ(P) is a function of only reactants concentrations, and K is the equilibrium constant, which is the ratio of rate constants of the forward and reverse reactions. As an example, we will consider the following one-route elementary reaction A + B ⇔ C + F. The rate of this reaction is given by:

       (1.66)

    where PA is the partial pressure (in case of gas-phase reactions) or concentration (for liquid-phase reactions). Taking into account Eq. (1.59), one gets:

       (1.67)

    In a more general case, Boreskov showed the rate expression of a nonlinear single-route reaction mechanism with a limiting step is:

       (1.68)

    are functions of reactants and product concentration respectively, m is the Boreskov molecularity or the reciprocal value of the Horiuti stoichiometric number of the rate-limiting step.

    Georgy Boreskov

    For the overall reaction, the temperature dependence of the reaction constants is expressed by the Arrhenius equation:

       (1.69)

    which is an approximation, since the pre-exponential factor, A, depends on temperature. Such dependence is rather weak compared to the exponential term; therefore, it is often neglected.

    More rigorous treatment of the temperature dependence will be presented in a chapter on transition state theory. Taking the natural log of both sides of the Eq. (1.69)

       (1.70)

    A plot of ln k versus 1/T (Fig. 1.30) gives a straight line with a slope of − Ea/R and an intercept of ln A.

    Fig. 1.30 Determination of the activation energy.

    The activation energy is thus represented by a following equation:

       (1.71)

    Transition state theory is the theory behind the Arrhenius equation and should be applied to elementary reactions only. In the majority of catalytic reactions, the reaction mechanism is more complex and contains several steps with different temperature dependences.

    For example, the hydrogenation of o-xylene (Fig. 1.31) is thermodynamically preferred up to 460 K (Fig. 1.32).

    Fig. 1.31 Reaction network for o -xylene hydrogenation.

    Fig. 1.32 Thermodynamics of o -xylene hydrogenation.

    At the same time, the rate clearly passes through a maximum (Fig. 1.33).

    Fig. 1.33 Kinetics of o -xylene hydrogenation.

    The apparent activation energy defined as:

       (1.72)

    is negative, demonstrating an interplay between adsorption of xylene on the surface (diminishing with T) and reaction (increase with temperature).

    1.6 Examples and Exercises

    1.6.1 The noncatalytic reaction occurs at 500°C and has an activation energy of 50 kJ/mol. By using a catalyst, the activation energy can be reduced to 35 kJ/mol. What is the difference in the reaction rates at this temperature?

    Solution:

    1.6.2 For isomerization of substituted benzene on a Pt foil, the reaction order was equal to unity above 500°C, while below this temperature, a zero order was obtained. The following values of rate constants were obtained:

    Explain temperature dependence. Take into account Hüttig temperature and the value of activation energy for thermal isomerization (250 kJ/mol).

    Solution:

    After plotting the data in Arrhenius coordinates, the changes in the activation energies could be seen. The low temperature domain corresponds to the kinetic regime, while in the high temperature region, the thermal isomerization is dominating. This is also reflected in the changes of reaction order. The Hüttig temperature of Pt is ca. 590°C, when surface atoms become mobile.

    1.6.3 Imagine that you are running propane dehydrogenation on supported Pd catalyst in a fixed bed reactor. The catalyst volume is 10 cm³, the density is 1.2 g/cm³ and catalyst surface area is 100 m²/g. The inlet feed of the reactant is 5 mol/h, while at the outlet, the flow of propane is 4.5 mol/h. Calculate conversion, the rate per catalyst volume, weight and catalyst surface.

    Solution:

    Expression for the reaction rate, where A is the reactant:

    Analogously values of the reaction rate per weight:

    and surface can be obtained:

    1.6.4 Lilja et al. [1] investigated esterification of propionic acid with methanol over a heterogeneous fiber catalyst. Calculate (a) activation energy based on initial rates (10 min); (b) Gibbs energy of the reaction at 60°C based on experimentally recorded equilibrium data at equimolar initial ratios of reactants; (c) Reaction enthalpy at 60°C, assuming that the entropy change during the reaction is negligible. The temperature dependence is presented below:

    Esterification kinetics of propionic acid with methanol (the initial molar ratio 1:1) as a function of temperature, 55°C (O), 60°C (×) and 63°C (+).

    Solution:

    (a) First reaction rates are calculated based on differences in concentrations after 10 min (Δct). Activation energy can be obtained from a plot in Arrhenius coordinates (ln k vs. 1/T). Note that temperature should be in K.

    The experimentally recorded values of equilibrium c/c.

    .

    .

    1.6.5 Prove that for two simultaneous reactions A → B (k1), A → C (k2), the ratio [cB]/[cC] = k1/k2 when the products are not present initially.

    Solution:

    . Dividing the equations by each other, separating the variables and integrating between the corresponding limits:

    .

    1.6.6 Tolune chlorination was performed at 20°C in the presence of a catalyst SnCl4. The initial pressure of Cl2 was always 100 mmHg. The reaction vessel (50 mL) contained always 30 mL of toluene. Experimental data at different catalyst concentration are given below. Calculate the rate constant assuming the first order in tolune.

    Solution:

    The concentration of toluene

    Enjoying the preview?
    Page 1 of 1