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

Only $11.99/month after trial. Cancel anytime.

Combustion, Flames and Explosions of Gases
Combustion, Flames and Explosions of Gases
Combustion, Flames and Explosions of Gases
Ebook1,287 pages9 hours

Combustion, Flames and Explosions of Gases

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Combustion, Flames, and Explosions of Gases, Second Edition focuses on the processes, methodologies, and reactions involved in combustion phenomena. The publication first offers information on theoretical foundations, reaction between hydrogen and oxygen, and reaction between carbon monoxide and oxygen. Discussions focus on the fundamentals of reaction kinetics, elementary and complex reactions in gases, thermal reaction, and combined hydrogen-carbon monoxide-oxygen reaction. The text then elaborates on the reaction between hydrocarbons and oxygen and combustion waves in laminar flow. The manuscript tackles combustion waves in turbulent flow and air entrainment and burning of jets of fuel gases. Topics include effect of turbulence spectrum and turbulent wrinkling on combustion wave propagation; ignition of high-velocity streams by hot solid bodies; burners with primary air entrainment; and description of jet flames. The book then takes a look at detonation waves in gases; emission spectra, ionization, and electric-field effects in flames; and methods of flame photography and pressure recording. The publication is a valuable reference for readers interested in combustion phenomena.
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483258393
Combustion, Flames and Explosions of Gases
Author

Bernard Lewis

Bernard Lewis (born May 31, 1916) was born in London. He is the author of forty-six books on Islam and the Middle East, including Notes on a Century: Reflections of a Middle East Historian; The End of Modern History in the Middle East; and The Crisis of Islam: Holy War and Unholy Terror. He also wrote three major syntheses for general audiences: The Arabs in History; The Middle East and the West; and The Middle East. Lewis is the Cleveland E. Dodge Professor Emeritus at Princeton University.

Read more from Bernard Lewis

Related to Combustion, Flames and Explosions of Gases

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Combustion, Flames and Explosions of Gases

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

    Combustion, Flames and Explosions of Gases - Bernard Lewis

    COMBUSTION, FLAMES and EXPLOSIONS of GASES

    Second Edition

    BERNARD LEWIS, Ph.D., Sc.D. (Cantab.)

    GUENTHER von ELBE, Ph.D. (Berlin)

    Combustion and Explosives Research, Inc., Pittsburgh, Pennsylvania

    Table of Contents

    Cover image

    Title page

    Copyright

    Preface to Second Edition

    Preface to First Edition

    Inside Front Cover

    List of Principal Symbols

    Part I: CHEMISTRY AND KINETICS OF THE REACTIONS BETWEEN GASEOUS FUELS AND OXIDANTS

    Chapter 1: Theoretical Foundations

    Publisher Summary

    1 Elementary and Complex Reactions in Gases

    2 Some Fundamentals of Reaction Kinetics

    3 The Steady-State Reaction Rate. Branched-Chain and Thermal Explosion Limits

    4 Chain Initiation and Chain Breaking at the Wall

    5 Chain Initiation in the Volume and Chain Breaking at the Wall

    Chapter 2: The Reaction between Hydrogen and Oxygen

    Publisher Summary

    1 The Thermal Reaction

    2 Sensitization and Inhibition of the Thermal Reaction by Additives

    Chapter 3: The Reaction between Carbon Monoxide and Oxygen

    Publisher Summary

    1 The Explosion Peninsula

    2 The Water-Catalyzed Reaction

    3 The Combined Hydrogen-Carbon Monoxide-Oxygen Reaction

    4 Sensitization and Inhibition by Additives

    Chapter 4: The Reaction between Hydrocarbons and Oxygen

    Publisher Summary

    1 Methane and Formaldehyde

    2 Ethane, Acetaldehyde, Ethylene, and Acetylene

    3 Propane and Propylene

    4 Cool Flames and Two-Stage Ignition

    5 Discussion of Mechanism of Hydrocarbon Oxidation

    6 Engine Knock

    Part II: FLAME PROPAGATION

    Chapter 5: Combustion Waves in Laminar Flow

    Publisher Summary

    1 Introduction

    2 The Adiabatic Plane Combustion Wave

    3 General Description of Heat Sink and Flow Effects

    4 Principle of Stabilization of Combustion Waves in Laminar Streams

    5 Quenching of Combustion Waves in Divergent Propagation

    6 Measurements of Limits of Flame Stabilization and Quenching

    7 Structure of Laminar Burner Flames

    8 Development of a Flame in a Laminar Flow Field from an Ignition Source

    9 Laminar Flame Propagation in Tubes

    10 Observations on Flames in Tubes; Effect of Vibrations and Gravity

    11 Wrinkling and Disruption of Combustion Waves by Diffusional Stratification of Fuel-Oxidant Mixtures

    12 Limits of Inflammability

    13 Ignition by Electric Sparks

    14 Ignition by Other Sources

    15 Combustion Waves in Closed Vessels

    16 Determination of Burning Velocity

    Chapter 6: Combustion Waves in Turbulent Flow

    Publisher Summary

    1 Description of Turbulent Burner Flames

    2 Concepts of Turbulent Flow

    3 Turbulent Burning Velocity

    4 Effect of Turbulence Spectrum and Turbulent Wrinkling on Combustion Wave Propagation

    5 Flame Stabilization in High-Velocity Flow by Pilot Flames

    6 Flame Stabilization by Bluff Bodies

    7 Ignition of High-Velocity Streams by Hot Solid Bodies

    Chapter 7: Air Entrainment and Burning of Jets of Fuel Gases

    Publisher Summary

    1 Description of Jet Flames

    2 Theory of Laminar Jet Flames

    3 Theory of Turbulent Jet Flames

    4 Burners with Primary Air Entrainment

    Chapter 8: Detonation Waves in Gases

    Publisher Summary

    1 Introductory Remarks

    2 Theory of Shock and Detonation Waves

    3 The Calculation of Detonation Velocity and Comparison with Experiment

    4 Measurements of Detonation Velocity; Limits of Detonability; Pulsating and Spinning Detonations

    5 Transition from Flame to Detonation

    Chapter 9: Emission Spectra, Ionization, and Electric-Field Effects in Flames

    Publisher Summary

    1 Flame Spectra

    2 Ions and Electric-Field Effects

    Chapter 10: Methods of Flame Photography and Pressure Recording

    Publisher Summary

    1 Flame Photography

    2 Measurement of Pressure Rise in Closed Vessel

    Part III: STATE OF THE BURNED GAS

    Chapter 11: Temperature, Pressure, and Volume of the Burned Gas

    Publisher Summary

    1 Thermodynamic Functions of Gases from Band Spectroscopy

    2 Adiabatic Explosions in Closed Vessels

    3 Adiabatic Flames at Constant Pressure

    Chapter 12: Temperature and Radiation of the Burned Gas

    Publisher Summary

    1 Characteristics of Thermal Radiation

    2 Experimental Investigations of Radiation from Flames

    3 Measurement of Temperature of Stationary Nonluminous Flames

    4 Measurement of Temperature and Emissivity of Soot-Forming Flames

    Part IV: PROBLEMS IN TECHNICAL COMBUSTION PROCESSES

    Chapter 13: Industrial Heating

    Publisher Summary

    Chapter 14: Internal Combustion Engines

    Publisher Summary

    1 Engine Cycles

    2 Combustion Process in the Otto Engine

    3 Combustion Process in the Diesel Engine

    4 Combustion in Jet Engines

    5 Combustion of Liquid Propellants in Rockets

    Appendix A: Data for Thermochemical Calculations

    Appendix B: Limits of Inflammability

    Appendix C: Flame Temperatures

    Author Index

    Subject Index

    Copyright

    Copyright ©, 1961, by Academic Press Inc.

    all rights reserved

    no part of this book may be reproduced in any form, by photostat, microfilm, or any other means, without written permission from the publishers.

    ACADEMIC PRESS INC.

    111 Fifth Avenue

    New York 3, New York

    United Kingdom Edition

    Published by

    ACADEMIC PRESS INC. (London) Ltd.

    17 Old Queen Street, London S.W. 1

    Library of Congress Catalog Card Number 60-14267

    printed in the united states of america

    Preface to Second Edition

    The tempo of combustion research has continued unabated during the past decade. Substantial progress has been made in establishing a common understanding of combustion phenomena. However, this process of consolidation of the scientific approach to the subject is not yet complete. Much remains to be done to advance the phenomenological understanding of flame processes so that theoretical correlations and predictions can be made on the basis of secure and realistic models.

    In this new edition particular emphasis has been placed on the modification of combustion wave propagation due to heat loss to the unburned medium and to localized changes of mixture composition by diffusional processes. Both influences occur as a result of divergent propagation which takes place in a flow field with steep velocity gradients or under conditions of flame initiation from a point source. The heat loss effect imposes stability limits of flames in flow fields and minimal values of flame diameter and ignition energy in flame initiation. Employing a new concept of flame stretch, which refers to the growth of flame surface in divergent propagation, it has been possible by similarity procedure to obtain good correlations of the fundamental, measurable flame parameters of burning velocity and wave width with flame stability limits and with spark ignition data. In this way, for example, the flame stability limits on flame holders in high velocity streams have been deduced. Furthermore, it is shown that in all instances of divergent propagation diffusional stratification of mixture composition occurs to a degree depending upon the relative diffusivities of the fuel and oxidant components of the mixture. This stratification produces effects that are predictable, at least qualitatively. Thus, in an over-rich mixture of a heavy hydrocarbon and oxygen, the minimal flame diameter and minimal ignition energy are smaller than predicted from the flame stretch equation for the original mixture because oxygen diffuses into the flame zone more rapidly than fuel. Similar considerations apply to flame stabilization. The concept of flame stretch and diffusional stratification permits an understanding of limits of inflammability.

    Other parts of this edition have not required the substantial revision given the subject of combustion waves. However, some revisions have been made particularly in the discussion of detonation processes, and much new material has been included where it appears to promote a better understanding of the subject.

    BERNARD LEWIS and GUENTHER VON ELBE,     Pittsburgh, Pennsylvania

    February, 1961

    Preface to First Edition

    During the past decade the scope and tempo of combustion research have increased to such extent, and so many new facts and concepts have developed, that preceding treatises appear to be wholly inadequate to meet the present day demands of student and research worker for a modern exposition of the subject. Although the authors have borrowed the title of their former book published in 1938, the present book cannot properly be labeled a second edition inasmuch as the text is entirely new with the exception of a few brief sections dealing with subjects that had already been well explored at that time. The purpose of the new volume remains the same however, namely, to provide the chemist, physicist, and engineer with the scientific basis for understanding combustion phenomena.

    The terms combustion, flame, and explosion became part of the common language long before clear scientific concepts existed, and therefore their usage has remained somewhat arbitrary and flexible. In this book the treatment of the subject matter comprises the theory of reaction chains and the chemical kinetics of reactions between fuel gases and oxygen, the hydrodynamics of combustion waves, detonation waves and jet flames, and the thermodynamics of combustion gases. In this respect we have followed a plan used in the former book, which we believe has proven valuable for delineating the field. Comparison of the state of knowledge 13 years ago with that of today shows a considerable advance in the field of chemical kinetics, a very large number of new facts and concepts in the field of ignition and combustion wave propagation, and substantial progress in the understanding of diffusion flames and detonation waves. On the other hand, the thermodynamic aspects of combustion had already attained a degree of finality many years ago and therefore no significant conceptual advances are to be noted, though the volume of reliable data has increased markedly.

    The reader will find new chemical mechanisms which have been deduced from the latest available evidence and which should be regarded as the present best judgment of the authors. These mechanisms are recommended as a basis for further experimental work and discussion. There already exists a considerable area of agreement among workers and it is hoped that the new analyses will contribute to its enlargement. The systems that have been treated are hydrogen and oxygen, carbon monoxide and oxygen, and hydrocarbons and oxygen. The system best understood is that of hydrogen and oxygen. In the carbon monoxide-oxygen system the reaction mechanism appears to be approaching some degree of clarification although quantitative data are still lacking. Considerable progress has been made in the understanding of the hydrocarbon-oxygen system and an early clarification of the oxidation mechanism, at least for the lower members, appears to be distinctly possible. There is also discernible a skeleton of consistent facts and theories of the oxidation of higher hydrocarbons. Some interlinking and interdependence of the reaction mechanisms of these systems are noted, and as further studies of selected mixtures are made the requirement of interconsistency of reaction mechanisms will become a powerful factor in the final elucidation of elementary chemical reactions.

    In the second part of the book dealing with flame propagation, emphasis shifts from reaction kinetics to hydrodynamics as the branch of science applicable to combustion and detonation waves and the combustion of fuel jets. It is on the subject of ignition and propagation of combustion waves that the greatest progress has been made in recent years. For the first time one understands the relationship between stabilization of combustion waves in burner flames and quenching in channels of critical diameters. An understanding is also provided of the various phenomena of flame propagation in tubes. A particularly significant advance is the integration of the data on spark ignition with the concept of minimum ignition energy which is derived from theoretical considerations of the development of combustion waves from an ignition source. It is shown that the hydrodynamic equations can be simplified to permit solutions which lead to correlation of the quenching distance with burning velocity and other measurable quantities.

    Important advances have been made by the recent theoretical and experimental investigations of Karlovitz on the interaction between combustion waves and turbulent motion in explosive gas mixtures. Very considerable progress is noted in the field of combustion of laminar and turbulent fuel jets since the early work of Burke and Schumann on laminar diffusion flames.

    A theory of burner performance is made possible by combining the theory of air entrainment of gas jets with the theory of flame stabilization. It is expected that this development will find useful application to the problem of fuel interchangeability in the gas industry. A discussion of this subject is included.

    The theory of detonation waves, while already far advanced by the early work of Chapman, Jouguet, and Becker, has been developed further by von Neumann and by Brinkley and Kirkwood. New explanations are suggested for the frequently observed discontinuous and spiral propagation of detonation waves. An insight is gained into the wave structure and of the interaction between shock front and reaction zone.

    Fundamental research should ultimately contribute toward improved understanding and control of technical combustion processes, particularly in engines. To accomplish this it is necessary to analyze the engine process in terms of the fundamental physical and chemical processes that occur in the various phases of starting and operation. At present this has been pursued only to a very small extent. Too often engine studies are confined to observations of the effect of fuel and engineering variables on over-all performance in a manner that excludes the possibility of recognizing the controlling physical and chemical processes. While development of modern engines has been eminently successful, this success has only been made possible by the accumulation of a very large volume of empirical information and by the continual maintenance of large and costly testing facilities. The question may be asked whether timely fundamental research could not have eliminated a large portion of this testing that has been carried on and is still continuing on a worldwide scale, and whether practical developments could not have been materially facilitated by scientific knowledge. In this text some attempt has been made to collect the meager information of fundamental character, first, to show that in principle it appears to be quite feasible to estimate the knock-limited performance of an Otto engine in a rather simple manner, and second, to illustrate that in jet engines a close scrutiny of the flame structure might lead to useful information on performance limits.

    The authors are grateful to their colleagues in the Explosives and Physical Sciences Division of the Bureau of Mines whose interest in and devotion to their researches have materially assisted in providing the experimental basis of the newer concepts of combustion wave propagation.

    BERNARD LEWIS and GUENTHER VON ELBE,     Pittsburgh, Pennsylvania

    May, 1951

    Inside Front Cover

    In the preparation of this new edition the authors acknowledge the support in part by the United States Air Force through the Propulsion Research Division, Air Force Office of Scientific Research of the Office of Aerospace Research, under contract Number AF49(638)-307.

    List of Principal Symbols

    PART I

    PART II

    PART III

    PART IV

    Part I

    CHEMISTRY AND KINETICS OF THE REACTIONS BETWEEN GASEOUS FUELS AND OXIDANTS

    Outline

    Chapter 1: Theoretical Foundations

    Chapter 2: The Reaction between Hydrogen and Oxygen

    Chapter 3: The Reaction between Carbon Monoxide and Oxygen

    Chapter 4: The Reaction between Hydrocarbons and Oxygen

    CHAPTER I

    Theoretical Foundations

    Publisher Summary

    This chapter describes the fundamentals of theoretical foundations in the evaluation of elementary and complex reactions in gases. As the basic information concerning elementary reactions is to be obtained predominantly from experiment, it is the task of the experimenter to discover those external experimental conditions for which the number of controlling elementary reactions is a minimum. Under such conditions, there is a chance of determining the nature of these reactions unambiguously, and, this having been accomplished, to determine quantitatively the reaction probabilities. The number of collisions that occur in a gas per unit time and volume between two kinds of molecules is proportional to the concentration of each. The classical theory of reaction kinematics has been supplanted by a more sophisticated rate theory that is based on the concept of activated collision complexes and their statistics. However, the simple equation remains substantially valid for reactions of atoms and can also be reasonably applied to simple molecules. In a collision of two polyatomic molecules capable of forming a chemical bond, or of a polyatomic molecule with an atom, association may be accomplished without the intervention of a third body, as the energy may be distributed among a number of degrees of freedom and is no longer concentrated in any one bond.

    1 Elementary and Complex Reactions in Gases

    This book deals with systems of gases that are capable of spontaneous chemical reaction with large energy release. In a volume of such gas or gas mixture the reaction may proceed as a wave propagating from a localized source of ignition, or it may occur more or less simultaneously throughout the volume. In this and the immediately following chapters we consider principally the latter mode of reaction.

    The molecules of a gas are in rapid motion and collide frequently with each other. The chance for chemical reaction between two molecules is given at the instant of collision. With ordinary molecules such as hydrogen and oxygen or hydrogen and chlorine this chance is extremely small, as is evident from the fact that at room temperature such mixtures do not react perceptibly during long periods of storage although at atmospheric pressure a molecule suffers about a billion collisions per second. However, if in a mixture of hydrogen and chlorine at room temperatures and ordinary pressures, moderate numbers of atoms of chlorine or hydrogen are introduced, as for example by illumination with light that dissociates Cl2, rapid formation of hydrogen chloride is observed which, under suitable conditions, proceeds with explosive rate.

    The reaction of hydrogen and chlorine furnishes a simple illustration of the subject of this section. Once chlorine atoms are formed in the system, the sequence of events is given by the chain reaction

    H and Cl are the chain carriers in this reaction. Thus, for every chlorine atom liberated large numbers of HCl molecules are formed. The individual reactions of the chain are elementary reactions. The over-all reaction H2 + Cl2 = 2HCl is thus a complex reaction comprising several elementary reactions and the rate of the over-all reaction is governed by the rates of the elementary reactions. The latter rates evidently depend on the frequency of the collisions of species such as Cl and H2 and on the probability of reaction in a collision. From statistical theory of chemical processes it is possible in a few simple cases to calculate a priori rates of elementary reactions.¹ Such instances are infrequent so that the theoretical analysis of elementary reactions does not at present provide the volume of data required for analysis of complex reactions. For the latter purpose one must therefore turn to experiment. The study of systems such as hydrogen and chlorine or hydrogen and oxygen requires, first, the determination of the governing elementary reactions, and second, the quantitative determination of the rate coefficients of these reactions. The rate of the complex reaction can then be calculated.

    Since the basic information concerning elementary reactions is to be obtained predominantly from experiment, it is the task of the experimenter to discover those external experimental conditions for which the number of controlling elementary reactions is a minimum. Under such conditions there is a chance of determining the nature of these reactions unambiguously, and, this having been accomplished, to determine quantitatively the reaction probabilities.

    This method of approach is illustrated in this book by the examples of the reactions between hydrogen and oxygen, carbon monoxide and oxygen, and hydrocarbons and oxygen. These systems are chosen because they have been studied most extensively and thus furnish object lessons of the accomplishments and limitations of chemical kinetic research in the field of combustion. No similar scope of information is provided by other fuel-oxidant systems which have come into use or are being considered, particularly in the field of jet and rocket propulsion.

    In the following sections we present the essential theoretical equipment which is used in this text for the evaluation of experimental data on the various reactions.

    2 Some Fundamentals of Reaction Kinetics

    A BlMOLECULAR REACTIONS

    The number of collisions that occur in a gas per unit time and volume between two kinds of molecules is proportional to the concentration of each. If the molecules are capable of reacting with each other a fraction of the collisions will be successful. Let us suppose that the gas consists of a mixture of hydrogen molecules and free hydroxyl radicals, OH. Since the reaction

    (I)

    is possible, water is formed at the rate

    (1)

    where the brackets represent concentrations, usually in molecules per cc. For the total number of collisions per unit time and volume between molecules 1 and 2 with concentrations N1 and N2 the kinetic theory of gases yields the equation

    (2)

    where m1 and m2 are the molecular weights, σ1,2 is the average diameter of the colliding molecules, R is the gas constant 8.314 × 10⁷ ergs per degree per mole, and T is the absolute temperature. In the classical theory of gas reactions, the value of a rate coefficient such as k, in equation (1) is represented by

    (3)

    According to the physical concept on which this equation is based, successful collisions between two reaction partners result if the two colliding molecules possess in a specific degree of freedom an energy equal to or exceeding the activation energy E (this is the meaning of the Arrhenius factor e−E/RT), and if the geometry of collision, viz., the line of approach of the two reacting species, is favorable to their interaction. The latter condition is expressed by the steric factor, whose value is undefined by theory except that according to the physical model it should be a number smaller than 1.

    The classical theory has been supplanted by a more sophisticated rate theory which is based on the concept of activated collision complexes and their statistics.² However, the simple equation (3) remains substantially valid for reactions of atoms and can also be reasonably applied to simple molecules. Only for larger molecules does the simple collision statistics underlying equation (3) become grossly inadequate, as shown by unrealistically high or low values of the steric factor. Such cases are not dealt with in this text and equation (3) will in general be found adequate for application to experimentally determined rate coefficients.

    B THREE-BODY-COLLISION REACTIONS

    In the special case of two colliding atoms no reaction to form a diatomic molecule occurs. If a molecule were formed at the instant of collision, its lifetime would be very much shorter than the time between collisions—of the order 10−3 to 10−5 as short at ordinary pressures and temperatures. This is because the sum of the energy of the reaction and the relative kinetic energy of the colliding atoms is more than enough to dissociate the molecule. The atoms will, therefore, fly apart again unless some third molecule participates in the collision—a three-body collision—to remove at least the excess energy. In this case a stable diatomic molecule results. In a collision of two polyatomic molecules capable of forming a chemical bond, or of a polyatomic molecule with an atom, association may be accomplished without the intervention of a third body, since the energy may be distributed among a number of degrees of freedom and is no longer concentrated in any one bond.

    As a general rule fairly large systems containing different types of bonds are required to make a direct association possible.³ For example, direct association between two CH3 radicals would be considered not possible. However, there is no theoretical reason to exclude such associations as C2H5 + O2 to form a peroxidic radical C2H5OO.

    For the number of three-body collisions per unit time and volume, Tolman has derived⁴ the following equation:

    (4)

    where N1 N2 and N3 are the concentrations of the colliding species. The equation is based on the assumption that the molecules are rigid elastic spheres and that a collision lasts while the molecules are within some small distances δ of each other. Since δ is an unknown quantity, the equation does not allow a determination of the actual number of three-body collisions. However, it is reasonable to suppose that δ is considerably smaller than the diameter of a molecule. It should therefore be demanded that experimental rates satisfy the condition

    (5)

    A much-studied reaction is the recombination of two hydrogen atoms to form a hydrogen molecule. Observations made by various investigators⁵ have shown conclusively that the reaction occurs in three-body collisions without appreciable activation energy. The rate is thus represented by

    (6)

    where [M] is the concentration of the third-body molecules and the coefficient krecomb. is found from equation (4) by means of the relation

    (7)

    From an analysis of the experimental material, Smallwood⁵ has concluded that krecomb. is approximately 16¹⁶ cm.⁶ mole−2 sec.−1 when H atoms act as third bodies, and more than 50 times smaller than this value when H2 molecules act as third bodies. Dividing krecomb. = 10¹⁶ cm.⁶ mole−2 sec.−1 by the square of the Avogadro number, 6.06 × 10²³, one obtains the rate coefficient in terms of single collisions as corresponds to equation (4). From σH2 = 2.36 × 10−8 cm. and σH = 2.14 × 10−8 cm.,⁶ one calculates δ ∼ 5 × 10−9 cm. for H as third body, and δ < 10−10 cm. for H2 as third body. The former amounts to about a quarter, and the latter to less than 2/100 of the gas-kinetic molecular diameters σ, so that both values are consistent with the above inequality (5).

    An upper bound of δ for H2 as third body is found independently from the fact that in thermal equilibrium the rate of the recombination reaction H + H + H2 = H2 + H2 is equal to the rate of the dissociation reaction H2 + H2 = H + H + H2; viz.,

    (8)

    An upper bound for kdiss. is given by

    (9)

    where ZH2,H2 is the collision frequency of H2 molecules, calculable from equation (2) above, and Ediss. is the dissociation energy of 103,240 cal./mole (see Appendix, p. 680). The ratio of the rate coefficients of the recombination and dissociation reactions is the thermal equilibrium constant K, viz.,

    (10)

    From equations (7) to (10), and using values of K listed in the Appendix, p. 680, one finds δ < 0.6 × 10−10 cm. for H2 as third body, in agreement with the previous value obtained from kinetic data.

    Analogous calculations yield δ < 10−11 cm. for each of the following reactions: O + O + O2 = O2 + O2; N + N + N2 = N2 + N2; and O + O2 + O2 = O3 + O2.

    Farkas and Sachsse⁷ produced H atoms at room temperature in a hydrogen-filled reaction vessel, using a well-known method which consists of adding a trace of mercury vapor to the gas and irradiating the mixture with light of the mercury resonance frequency. In this way excited mercury atoms are formed which in collisions with H2 transfer their energy and dissociate the hydrogen molecules into atoms. The authors measured the stationary H-atom concentration which is established at constant light intensity by observing the rate of conversion of para- to ortho-hydrogen, the conversion rate being a known function of the H-atom concentration. Their measurements did not yield an acceptable value of the coefficient krecomb. because apparently the authors overestimated the effective light flux in their experiments; however, another phase of their work is of special interest. By adding small amounts of O2 to the system, they were able to observe a decrease of the H-atom concentration depending on the O2 concentration. O2 molecules react with H atoms in a three-body collision reaction [given the number (VI) in this text]:

    (VI)

    Using Farkas and Sachsse’s data, von Elbe and Lewis⁸ found the rate coefficient of reaction (VI), with H2 as third body, to be smaller than the rate coefficient of H-atom recombination by a factor of 1:1327. Using Smallwood’s data quoted above, this yields

    or in units of single collisions,

    from which one calculates δ < 0.6 × 10−13 cm.

    Table 1 summarizes the upper-bound values of δ that have been found for the various reactions.

    TABLE 1

    UPPER BOUNDS FOR Δ IN TOLMAN’S THREE-BODY-COLLISION EQUATION

    The table provides a measure of guidance for estimating the magnitude of δ and hence of the rate of various ternary reactions. In particular, it may serve to estimate lower bounds of relaxation times in the adjustment of gaseous systems to thermodynamic equilibrium, where the rate-determining process is frequently the recombination of atoms and free radicals in three-body collisions.

    C MONOMOLECULAR REACTIONS

    A reaction may also occur by the decomposition of a single molecule that has the requisite energy. Such activated molecules are produced by collisions and are destroyed in two ways, by collisional deactivation and by decomposition. At high pressures the first of the latter two processes is predominant and the concentration of activated molecules will correspond to the thermal equilibrium concentration. For this condition the theory leads to a reaction rate that is proportional to e−E/RT times the concentration of reactive species. At pressures so low that collisional deactivation becomes negligible because the lifetime of the active molecule is shorter than the time between collisions, the rate is bimolecular since it depends on the rate of activation.

    D SURFACE REACTIONS

    Besides gas-phase reactions one has to consider reactions occurring at surfaces. A surface may act as a catalyst, facilitating a reaction that takes place in the gas phase with difficulty or not at all. According to accepted views of heterogeneous catalysis, it is required that one or both of the reacting gases be adsorbed on the surface. The reaction rate would then be proportional to the surface concentrations of the adsorbed gases. With glass, the following relation between gas-phase and surface concentrations has been found for a number of gases:

    (11)

    s is the number of adsorbed molecules per unit area, s0 the number of elementary spaces per unit area capable of holding one molecule of adsorbed gas, p the pressure, and b a constant. The equation can be derived theoretically on introducing certain assumptions concerning the state of the adsorbed molecule. If the pressure is very low or the surface forces are small, corresponding to a very small value of b, s becomes equal to s0bp. In the other extreme s becomes s0. If the rate of the surface reaction is slow so as not to disturb seriously the equilibrium between surface and gas-phase concentrations, it is seen to be proportional to the gas-phase concentrations of the reactants raised to a power between zero and one. For example, the rate of the heterogeneous combination of hydrogen and oxygen at about 550°C. in a silica vessel filled with small silica spheres was found to be approximately proportional to the hydrogen pressure and independent of the oxygen pressure,¹⁰ indicating strong adsorption of oxygen and weak adsorption of hydrogen on such a surface.

    The property of a surface to destroy chain carriers plays a very important role in the kinetics of chain reactions. The destruction may occur by a reaction with the material of the wall or by firm adsorption followed by a heterogeneous reaction. Different surfaces possess quite different efficiencies for destroying chain carriers. A surface may be rendered ineffective for chain destruction by poisoning it. For example, water is a strong poison for the destruction of H atoms on a glass surface. Steiner¹¹ found that if 2 to 4% of water vapor is added to a mixture of H atoms and H2 molecules flowing through a glass tube, only approximately 1 in 10⁶ collisions of an H atom with the wall leads to its destruction, that is, the efficiency of destruction is approximately 10−6. Poole¹² reports an efficiency of 1.6 × 10−4 under similar conditions, and on a phosphoric acid-coated glass surface an efficiency of 3.9 × 10−5. In a dry mixture Steiner¹¹ found the efficiency to be several orders of magnitude greater. For Cl atoms the efficiency is about 10−3;¹³ for Br atoms it is about l;¹⁴ for CH3 and C2H5 values of about 10−3 have been reported.¹⁵ The recombination of H and of OH on clean pyrex and pyrex coated with various salts has been studied by Smith¹⁶ over a temperature range from room temperature to more than 500°C. The efficiency varies with the nature of the surface. H recombines strongly on all salts investigated, except KCl. OH recombines strongly on KCl.

    In case the wall of the vessel is a sink for chain carriers the kinetics of the over-all reaction is strongly dependent on the diffusion of chain carriers to the wall. Since the chain carrier concentration is very low, the diffusion coefficient D may be written¹⁷

    (12)

    where λe the velocity of the chain carrier. If the chain carrier diffuses through a single gas, 2, of approximately equal molecular weight, λe equals the ordinary mean free path

    (13)

    where N2 is the number of molecules per cc. of the gas (2). If the chain carrier diffuses through a gas, 3, of much smaller molecular weight, the persistence of velocity of the chain carrier in collisions with the light gas causes λe to be larger than λ. The relation is approximately*

    (14)

    The effective mean free path in a mixture of gases is computed from the equation

    (15)

    The average molecular velocity is given by the equation⁴

    (16)

    3 The Steady-State Reaction Rate. Branched-Chain and Thermal Explosion Limits

    The development of equations for the rate of chain reactions will now be considered. As an example we choose the following simplified system of reactions involving hydrogen and oxygen.

    (I)

    (II)

    (III)

    (IV)

    As we shall see later these reactions represent only a portion of the actual mechanism. The numbering of the reactions follows a precedent established by the authors in earlier publications. Let us suppose that radicals OH are formed spontaneously from neutral molecules H2 and O2 at the rate I per cc. per second. For the purpose of this illustration the mechanism of this initiating reaction need not be specified further. It may be a gas-phase or a surface reaction. It will be assumed that the radical HO2 is sufficiently inert chemically so that it survives a period of diffusion to the wall of the vessel where it is adsorbed and ultimately destroyed by recombination with other HO2 radicals. Thus reaction (VI) is a chain-breaking reaction since it removes the free valence associated with the chain carriers, H, O, and OH.

    The rate of the complex over-all reaction is measured by the rate of formation of water [equation (1)]. It is proportional to the concentration of OH, and the kinetic problem consists in determining this concentration. The time rate of change of the concentrations of chain carriers is determined by the following three equations:

    (17)

    (18)

    (19)

    By eliminating the concentrations [H] and [O], one obtains from equations (17), (18), and (19)

    (20)

    The significance of this equation may be illustrated as follows: An H atom has a choice between reaction paths (II) and (VI), viz.,

    An O atom always reacts via (III), and the H atom emerging from the reaction has the same choice, viz.,

    The primes indicate free valences. Thus, the rate of formation of free valences from hydrogen atoms is proportional to the rate of formation of H atoms multiplied by a factor representing twice the rate of reaction (II) relative to reaction (VI). This factor is 2k2/k6[M]. Furthermore, the rate of formation of free valences from oxygen atoms is proportional to the rate of formation of O atoms multiplied by the factor (1 + 2k2/k6[M]). The left-hand side of equation (20) therefore represents the rate of formation of free valences in the system. As the number of free valences per cc. increases the OH concentration increases. If now 2k2/k6[M] < 1, the rate of formation of free valences decreases and tends to vanish owing to the negative sign of the OH term on the right-hand side. Therefore, after an induction period the chain carrier concentration attains a constant value and the rate of formation of water becomes constant. This statement must be qualified somewhat because as the reaction proceeds hydrogen and oxygen are consumed and therefore [H2] and I decrease with time. However, the establishment of a steady-state chain carrier concentration requires only a minute consumption of reactants; so that except for the initial, usually unmeasurable, induction period, the chain carrier concentration can be considered to be in equilibrium with the reactant concentrations. The OH concentration is then given by

    (21)

    Since reaction (II) leads to self-multiplication of free valences, it is referred to as a chain-branching reaction.

    A generalized form of equation (20) is

    (22)

    Here n refers to the chain carrier concentration, n0 is the rate of the chain-initiating reaction, and β and α are the coefficients of the chain-breaking and the chain-branching reactions, respectively, β and α may be individually composed of various rate coefficients depending on the reaction mechanism, and n may comprise several different chain carriers. the chain carrier concentration, and therefore the reaction rate, increases exponentially with time. Owing to the rapidity of molecular events this condition corresponds to very rapid heat release, that is, explosion. For the condition α < β, dn/dt becomes zero, so that

    (23)

    The condition α = β defines the explosion limit.

    If chain breaking were to occur by the recombination of two chain carriers in the gas phase, the steady-state concentration n would be determined by an equation of the form

    (24)

    from which no explosion limit is obtained, n remaining always finite. In actual explosive gas mixtures such gas-phase recombinations are not found to be controlling reactions. They occur in the later stages of a branched-chain explosion when the chain carrier concentration has become high and the reaction rapidly approaches completion.

    Self-acceleration of the reaction rate to explosion can also occur as a result of heat release. If heat is produced in the reaction vessel at a rate exceeding the rate of heat loss to the surroundings, the temperature rises; and since reaction rate coefficients are exponential functions of temperature, the over-all reaction is auto-accelerated until a large fraction of the reactants is consumed. The condition of thermal explosion is not as simply defined as a branched-chain explosion. The region of non-explosive reaction is defined by the condition that the rate of accumulation of heat in the reaction vessel is zero, i.e.,

    (25)

    Here q is the thermal energy of the contents of the vessel; Q, the heat of reaction per mole of gas decomposed; V, the volume of the vessel; r, the number of moles reacting per unit time and volume; a, the area of the wall of the vessel; K, the heat transfer coefficient; T, the average temperature in the vessel; and T0, the temperature of the wall. The first term on the right side is the rate of heat evolution, and the second term is the rate of heat loss to the surroundings. The rate r is a complicated function which is determined by the reaction mechanism, and generally comprises a factor of the form e−E/RT. If equation (25) has a real root in T, the system will come to equilibrium at this temperature and the reaction is in the steady state subject to consumption of the reactants. But if T0 is increased sufficiently, the equation will have no real root in T, dq/dt will always be positive, the temperature will increase, and the process of self-acceleration of the reaction, i.e., explosion, will set in. The critical temperature or ignition temperature T0 depends on various parameters. Thus, increase of pressure, insofar as it increases the reaction rate r, decreases the ignition temperature; increase of vessel diameter, which increases V more than a, also decreases the ignition temperature. If r is dependent on some power of the pressure, a plot of ignition temperature versus pressure may generally be expected to follow an equation of the form

    (26)

    where A and B are approximately constant, A representing an over-all activation energy divided by the gas constant, and B depending upon vessel factors and other parameters. Equation (26) has been found to agree with many experimental observations but in itself is not very informative. In particular it does not serve as a criterion for distinguishing between thermal and branched-chain explosion limits because the temperature-pressure relations of the latter generally also follow equations of the form (26).

    As mentioned earlier, when the reaction mechanism is unknown it is the problem of the experimenter to discover those external experimental conditions for which the number of controlling elementary reactions is a minimum. To this we may add that it is also necessary to avoid complicating effects such as nonisothermal conditions. That is, the reaction rate should be studied well below the thermal explosion limit. As far as throwing light on the reaction mechanism is concerned, the thermal limit is uninformative. In contrast, a branched-chain explosion limit, if clearly recognizable as such, provides crucial information. An outstanding example is the second explosion limit of hydrogen and oxygen which will be discussed later.

    Solutions of equation (25) for special cases have been given by a number of investigators. We mention the treatment of azomethane and ethyl azide explosions¹⁸ and methyl nitrite explosions¹⁹ which appear to be unimolecular decomposition reactions. Extensive theoretical work on the induction period, calculation of limits, and temperature distribution in the reaction vessel for generalized and specific cases has been done by Frank-Kamenetsky and others.²⁰

    From the foregoing it is clear that the phenomenon of the explosion limit in itself is not a criterion for a chain reaction, branched or unbranched. The reaction mechanism must be determined in every individual case from the sum total of experimental data. In a number of cases of which the hydrogen-chlorine reaction is the classical example, a chain mechanism is compellingly demonstrated by the large quantum yield in the photochemical reaction. Likewise, the phenomenon of the explosion peninsula in the hydrogen-oxygen reaction is unequivocal evidence for a branched-chain mechanism. In other cases, like the reaction between methane and oxygen, the chain character is demonstrated by analysis of numerous observations. It may be remarked that chain reactions are generally of a high and variable order, that is the rate is dependent on some high and variable power of the concentration of the reactants. However, complex reactions of the nonchain type that also possess this feature are possible.²¹ Under certain conditions such complex reactions may be dependent on the diameter, the nature of the material of the vessel and on inert gases. Such dependence is, therefore, no certain criterion for chain reactions.

    The following sections treat the general case of isothermal reaction with chain breaking at the vessel surface. The chain carrier may reach the wall either by diffusion or convection. Of these two modes of transport only the first will be considered, that is the treatment will be confined to quiescent gas mixtures. Moreover, in order to express the rate of change of concentration of chain carriers in an element of volume, it will be necessary to assume some definite geometric shape of the reaction vessel. A spherical vessel will be chosen, although the results will not be very different for other vessels such as cylindrical or plane-parallel.²²

    The wall may serve simultaneously as an initiator as well as a destroyer of chains. The treatment²³ must distinguish between the two cases of chain initiation at the wall and in the gas phase.

    4 Chain Initiation and Chain Breaking at the Wall

    The rate of change of concentration of chain carriers n at any distance ρ from the center of a vessel of radius r is given by the equation

    (27)

    where D is the diffusion coefficient and α the coefficient of the rate of chain branching.

    Only the steady state will be considered, that is,

    (28)

    With this condition the solution of equation (27) is given by

    (29)

    In order to determine the integration constant A, a boundary condition must be introduced. The latter may be found by equating the rate of branching throughout the volume to the rate of surface destruction of chain carriers minus the rate of production of new chain carriers by the surface. The rate of surface destruction of chains is ∈ times the rate at which the chain carriers strike the surface, ∈ being the chain-breaking efficiency of the wall with values ranging from zero to one. From gas kinetic considerations the latter rate is taken to be equal to

    where is the average molecular velocity.

    ∈, , and λ depend on the nature of the diffusing species. If several kinds are involved, the values of these constants become functions of the ratios in which the different chain carriers are present. In the reactions to be considered later, the problem is simplified by the fact that one kind of chain carrier can be considered to predominate.

    is given by

    (30)

    Substituting equation (29) and solving the integral, one obtains

    (31)

    and the rate of production of new chain carriers by the surface is 4πr²m0, where m0 is the rate of chain initiation per unit area, the boundary condition leads to the equation

    (32)

    The first term in the denominator can be expanded by the trigonometric relation

    is a very small number since in all physically interesting cases λ << r cannot exceed π if n is to remain positive [compare .

    times the chain carrier concentration at the wall. The latter concentration is never zero in the present case where the wall itself is the source. If ∈ equals 1, the only chain carriers that are permitted to leave the wall are those generated by the chemical process at the rate m. From the equation for n(wall) obtained by combining equations (29) and (32) and setting ρ = r, for ∈ = 1 is fulfilled only when

    (33)

    The usual derivation of D . In the present treatment, a consistent theory is obtained only by means of equation (33). The difference in the numerical coefficients is attributable to the approximations that have been made in formulating the boundary condition; it does not seem of sufficient importance to warrant a reconsideration of the treatment.

    Substituting equations (32) and (33) into (31), one obtains

    (34)

    as can be verified by cross-multiplying and adding up the terms of the denominator.

    as a function of D, α, r, and ∈ for the assumed model of chain carrier diffusion and reaction in a spherical vessel. The equation is far too complicated for use in a practical situation, where exact numerical calculations are usually unimportant, where the applicability of the model is usually problematical, and where no direct methods exist for determining the values of the coefficients α and ∈. Such situation demands a form of theoretical correlations as exemplified by equation (21) above, in which the effects of the various concentrations and rate coefficients are readily separable. In order to obtain an equation of this form, the denominator of equation (34) must be separated by means of suitable approximations into two terms representing the coefficients of chain breaking and chain branching, respectively. This may be done in the following way.

    that creates the difficulty in reduces in the limit to 2/[(π²D/αr²and for all values of ∈ between zero and 1. The new equation, by means of a simple transformation, may be written in the form

    (35)

    This equation becomes identical with that are somewhat smaller than those calculated from equation (34), but numerical calculations have shown²⁴ that even for extreme conditions this discrepancy is less than 40% and hence insignificant in kinetic studies.

    In equation (35) the numerator represents the rate of chain initiation at the surface, adjusted for the destruction of some of the newly formed chain carriers by the wall (when ∈ is large compared to 2λ/r the numerator becomes quite small indicating that few chain carriers actually escape from the surface); and the two terms in the denominator represent the rate coefficients of chain breaking and chain branching, respectively. For 2λ/∈r << 1, the chain-breaking coefficient²⁵ reduces to π²D/r.

    Physical interpretation of the condition 2λ/∈r << 1 signifies that chain carriers diffusing to the vessel surface linger there for a sufficiently long time so that they are destroyed with high probability by repeated collisions with the wall. It does not in general signify that ∈ itself is close to unity; in fact, in the usual range of values of vessel diameter and mean free path, viz., pressure, quite small values of ∈ are sufficient to render the vessel wall an effective sink for the chain-carrier diffusion. Such being the case, the rate of chain breaking is determined solely by the number of chain carriers reaching the surface per unit time. It is dependent on the diffusion coefficient D but independent of the chain-breaking efficiency ∈. It is proportional to the steepness of the stationary concentration gradient and, in addition, to the surface-to-volume ratio, both of which are inverse functions of r; thus, the factor 1/r, which represents the number of chain carriers impinging per unit time on unit area of the wall; to the chain-breaking efficiency ∈, which determines the fraction of the impinging chain carriers that are destroyed; and to the surface-to-volume ratio, which for a spherical vessel is 3/r, a result which is also obtainable from the original 1.24 in place of the numerical factor 0.75; however, this inaccuracy does not affect the utility of equation (35), considering that values of ∈ are at best known only to the order of magnitude.

    5 Chain Initiation in the Volume and Chain Breaking at the Wall

    If chains are initiated in the volume the condition for the steady state is given by

    (36)

    where n0 is the rate of chain initiation per unit volume. Equation (36) has the solution

    (37)

    Introducing the boundary condition that the rate of chain branching plus the rate of chain initiation throughout the volume is equal to ∈ times the rate at which the chain carriers strike the surface, the integration constant A becomes

    (38)

    is

    (39)

    and ∈ → 1. This leads to

    (40)

    Again, the approximate that are somewhat lower than those calculated from equation (39), but the difference is less than 40% even for extreme conditions.²⁴ The numerator represents the rate of chain initiation in the gas phase and the two terms in the denominator represent the rate coefficients of chain breaking and chain branching, as before. The chain-breaking coefficient is identical with the coefficient appearing in equation (35) above.


    ¹For a summary of the subject, consult H. S. Taylor, High Speed Aerodynamics of Jet Propulsion, Vol. II, pp. 101–159. Princeton Univ. Press, Princeton, N. J., 1956.

    ²H. Eyring, J. Chem. Phys. 3,107 (1935); cf. Taylor.¹

    ³G. E. Kimball, J. Chem. Phys. 5, 310 (1937); cf. L. S. Kassel, ibid. 5, 922 (1937).

    ⁴R. C. Tolman, Statistical Mechanics. Am. Chem. Soc. Monograph, 1927.

    ⁵W. Steiner and Z. Bay, Z. physik. Chem. (Leipzig) B3, 149 (1929); W. Steiner, Trans. Faraday Soc. 31, 623 (1935); J. Amdur and A. L. Robinson, J. Am. Chem. Soc. 55, 1395 (1933); H. M. Smallwood, ibid. 56, 1542 (1934).

    ⁶P. Harteck, Z. physik. Chem. (Leipzig) A139, 98 (1928).

    ⁷L. Farkas and H.

    Enjoying the preview?
    Page 1 of 1