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

Only $11.99/month after trial. Cancel anytime.

Some Problems of Chemical Kinetics and Reactivity: Volume 1
Some Problems of Chemical Kinetics and Reactivity: Volume 1
Some Problems of Chemical Kinetics and Reactivity: Volume 1
Ebook559 pages6 hours

Some Problems of Chemical Kinetics and Reactivity: Volume 1

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Some Problems of Chemical Kinetics and Reactivity, Volume 1 consists of calculations on radical and radical chain reactions. The subject bond dissociation energies are fully discussed. The concept of uniradical reactivities is comprehensively explained. Isomerizations are a class of radical reactions in which the free valency takes another bond in the same radical. The text provides sample of experiments on the subject. The book contains a section on polar factors in organic reactions. Polymerizations are another concept covered in the book. Subjects such as the reactions of biradicals, the start and end of a chain reaction, and ions of variable valency are explained. A separate chapter of the book focuses on the kinetics of chain reactions. The cracking of hydrocarbons such as the alkane is analyzed in detail. The oxidation of hydrocarbons is another topic explained in the book. The text will provide excellent insight for chemists, students, and researchers in the field of chemistry.
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483194950
Some Problems of Chemical Kinetics and Reactivity: Volume 1

Related to Some Problems of Chemical Kinetics and Reactivity

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Some Problems of Chemical Kinetics and Reactivity

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

    Some Problems of Chemical Kinetics and Reactivity - N. N. Semenov

    SOME PROBLEMS OF CHEMICAL KINETICS AND REACTIVITY

    VOLUME 1

    Academician N.N. SEMENOV

    Institute of Chemical Physics, Academy of Sciences of the U.S.S.R.

    PERGAMON PRESS

    Table of Contents

    Cover image

    Title page

    Copyright

    INTRODUCTION

    PART I: RADICAL REACTIONS (PROPAGATION AND BRANCHING OF CHAIN REACTIONS)

    Chapter 1: UNIRADICAL REACTIVITIES

    Publisher Summary

    1. RADICAL REACTIONS: BASIC TYPES

    2. EXPERIMENTAL DETERMINATION OF ACTIVATION ENERGIES

    3. BOND DISSOCIATION ENERGIES

    4 THE CORRELATION OF ACTIVATION ENERGIES WITH HEATS OF REACTION

    5 RADICAL AND MOLECULAR ACTIVITIES

    6. EMPIRICAL FORMULAE FOR COMPUTING BOND-ENERGIES IN ORGANIC MOLECULES

    7. ADDITION AT MULTIPLE BONDS

    8. RADICAL DECOMPOSITION AND THE ENERGIES OF π-BONDS

    9. RADICAL ISOMERIZATIONS

    10. TRANSFER REACTIONS

    11. POLAR FACTORS IN ORGANIC REACTIONS

    12. POLAR FACTORS IN POLYMERIZATION *

    Chapter 2: ALTERNATIVE UNIVALENT FREE-RADICAL REACTIONS

    Publisher Summary

    1 THE OCCURRENCE OF ALTERNATIVE RADICAL REACTIONS

    2. THE INFLUENCE OF TEMPERATURE AND PRESSURE ON ALTERNATIVE RADICAL REACTIONS

    3 INTERMEDIARY AND TERMINAL PRODUCTS IN CHAIN REACTIONS

    Chapter 3: THE REACTIONS OF BIRADICALS

    Publisher Summary

    1 THE TRANSITION OF ATOMS TO THE ACTIVE-VALENCE STATE

    2 REACTIVITY IN THE MOLECULES O2, S2 AND Se2

    3 DIVALENT CARBON

    4 COMPLEX BIRADICALS

    5 PRODUCTION AND REACTIVITY OF > CH2 AND —O—

    6 THE ROLE OF BIRADICALS IN CHAIN REACTIONS

    PART II: INITIATION AND TERMINATION OF CHAIN REACTIONS

    Chapter 4: MOLECULAR DISSOCIATION AND RADICAL COMBINATION

    Publisher Summary

    1. HOMOGENEOUS INITIATION OF CHAINS

    2. HOMOGENEOUS RECOMBINATION OF RADICALS

    3. DEPENDENCE OF OVERALL REACTION KINETICS ON CHAIN-TERMINATION MECHANISM

    4. SPECIAL ASPECTS OF RADICAL GENERATION AND RECOMBINATION IN LIQUIDS

    5. THE EFFECTS OF INHIBITORS ON THE RATES OF CHAIN REACTIONS

    Chapter 5: IONS OF VARIABLE VALENCY AS CHAIN INITIATORS

    Publisher Summary

    1. FORMATION OF RADICALS AND IONIC RADICALS BY ELECTRON TRANSFER

    2. THERMODYNAMIC AND KINETIC ASPECTS OF FREE RADICAL FORMATION

    3. IONS OF VARIABLE VALENCY AS CHAIN INITIATORS

    Chapter 6: WALL INITIATION AND RETARDATION OF CHAIN REACTIONS

    Publisher Summary

    1. GENERATION AND REMOVAL OF FREE RADICALS AT THE VESSEL WALLS [1]

    2. THE METHOD OF DIFFERENTIAL CALORIMETRY

    3. FREE VALENCIES IN THE WALLS

    4. RADICAL PRODUCTION AT ACTIVE SURFACES

    5. HETEROGENEOUS PRODUCTION OF RADICALS WHEN GASEOUS ADMIXTURES ARE PRESENT

    6. HETEROGENEOUS MOLECULAR REACTIONS GIVING FREE RADICALS

    7. APPLICATION OF THE ABOVE CONCEPTS TO HETEROGENEOUS CATALYSIS

    8. HETEROGENEOUS CATALYSIS IN BIOLOGY

    PART III: KINETICS OF CHAIN REACTIONS

    Chapter 7: CHAIN REACTIONS COMPETING WITH REACTIONS BETWEEN SATURATED MOLECULES

    Publisher Summary

    1. ENERGY DEPENDENCE OF CHAIN REACTIONS AND MOLECULAR REACTIONS

    2. THE RELATIVE RATES OF DIRECT MOLECULAR AND CHAIN REACTIONS

    3. ALKYL BROMIDE DECOMPOSITION

    4. ALKYL CHLORIDE DECOMPOSITION

    5. THE CRACKING OF HYDROCARBONS

    6. FACTORS INFLUENCING THE CHAIN LENGTH

    7. THE ELEMENTARY ACT OF DECOMPOSITION

    8. FREE RADICAL PRODUCTION BY REACTION BETWEEN STABLE MOLECULES

    9. PARTICULAR EXAMPLES

    THE ACTIVATED COMPLEX

    THE QUANTUM-MECHANICAL CALCULATION OF ACTIVATION ENERGIES

    NAME INDEX

    SUBJECT INDEX

    Copyright

    PERGAMON PRESS LTD.

    4 &5 Fitzroy Square, London W.1

    PERGAMON PRESS INC.

    122 East 55 th Street, New York 22, N.Y.

    P.O. Box 47715 Los Angeles, California

    PERGAMON PRESS S.A.R.L.

    24 Rue des Écoles, Paris Ve

    Copyright

    ©

    1958

    Pergamon Press Ltd.

    Library of Congress No. 58–10936

    Printed in Great Britain by the Pitman Press, Bath

    INTRODUCTION

    N. SEMENOV

    This book was first published in 1954 by the U.S.S.R. Academy of Sciences, its history being as follows: The Chemistry Division of the Academy had arranged an All-Union Conference on Chemical Kinetics and Reactivity, to be held in 1955, and I was requested to present the opening paper; while working on this I came to the conclusion that it would be of value to produce a small book in which the published data and my personal views on the topics to be discussed could be presented. The book appeared in 1954, and a collection of 69 papers presented at the Conference by various Soviet workers was published by the Academy in the first half of 1955; the Conference was held on 20–25 June, 1955, being a general discussion on my book and on the papers in the collection. A number of important topics, touching on the development of studies in chemical kinetics and reactivity, were considered.

    The book was produced only in small numbers, and the edition was soon sold out; I was requested by the Academy to prepare a second edition. Immediately afterwards I was requested to grant my permission for the book to be translated into English and German, by Pergamon Press and Princeton University, and by the Academy of Sciences of the German Democratic Republic, respectively. The book has been substantially enlarged for the second edition, fresh data appearing in the world literature or obtained at the Institute of Chemical Physics being incorporated. The book consists of various views and calculations developed by myself and certain colleagues at the Institute of Chemical Physics during the last 10–15 years from our experiments and from the published data on chemical kinetics. The book therefore presents a particular viewpoint, does not cover all work done on reaction mechanisms, and does not pretend to be a treatise or work of reference. The main topics considered are radical and radical-chain reactions. More than 20 years have passed since my previous book Chain Reactions appeared; meanwhile studies in chemical kinetics and chain reactions have developed enormously, vast numbers of papers have appeared, but it still seems as though the main problems are as yet not completely solved; the situation is very different from that in the physics of nuclear reactions, particularly chain fission, which was discovered in 1939 and since then has been so completely elucidated. The unsatisfactory situation in relation to reaction mechanisms has arisen, in my opinion, because the world’s scientists have been occupied with isolated observations on particular reactions and have not undertaken a general all-round study. Such isolated observations are now quite inadequate in most cases, and can even be useless in developing the theory, either in relation to particular types of reaction, or to general problems in kinetics and reactivity. To undertake isolated studies unrelated to general problems is like determining particular points on a surface of which we wish to define the form; if the surface is complicated, knowledge about a few points tells us nothing about the form, and only enables us to reject unfounded hypotheses. Only the organized efforts of physical, organic and inorganic chemists directed to elucidating the general problems of mechanism, in relation to basic types of reaction and related reactivity problems, can lead to a decisive advance in chemistry.

    Here, as in all other problems, from the fundamental political problems of the struggle for peace onwards, the united forces of the world’s scientists are required. I shall be happy if this book is of interest to chemists, even if only slightly so, and if it directs attention to the solution of problems in chemical kinetics and reactivity.

    I record my thanks to S. S. Polyak, A. B. Nalbandyan, N. S. Enikolopyan, D. G. Knorre, A. E. Shilov and several others of my colleagues at the Institute of Chemical Physics, and also to A. M. Pravednikov, who have all participated in writing the new chapters and in supplementing the previous chapters for the second edition.

    PART I

    RADICAL REACTIONS (PROPAGATION AND BRANCHING OF CHAIN REACTIONS)

    Outline

    Chapter 1: UNIRADICAL REACTIVITIES

    Chapter 2: ALTERNATIVE UNIVALENT FREE-RADICAL REACTIONS

    Chapter 3: THE REACTIONS OF BIRADICALS

    Chapter 1

    UNIRADICAL REACTIVITIES

    Publisher Summary

    This chapter discusses the types of radical reactions and the radical and molecular activities. The reactions of univalent free radicals may be divided into three basic classes: (1) substitution, (2) addition at multiple bonds or reaction with atoms that have unpaired electrons, and (3) isomerization. It is very difficult to measure the activation energies of elementary free radical reactions. Free radicals are unstable, and they disappear rapidly by combination either on the walls of the vessel or by a homogeneous reaction. Consequently, it is very difficult to maintain a constant concentration of free radicals and to measure it especially in the gas phase. Free radicals are commonly obtained from electrical discharges in gases by the action of ultraviolet light or by pyrolysis of compounds that decompose to give free radicals. Direct quantitative determination of the free radical concentration is usually extremely difficult, particularly as the concentrations normally encountered are very small. Therefore, the concentrations and rates of reaction are usually deduced from the rates of formation of stable intermediates or of final products.

    1. RADICAL REACTIONS: BASIC TYPES

    The reactions of univalent free radicals may be divided into three basic classes:

    (1) Substitution:

    where R1, R2, and R3 are atoms or radicals. R2 is a single atom in the majority of substitution reactions which have so far been studied.

    (2) Addition at multiple bonds, or reaction with atoms which have unpaired electrons:

    (a) 

    (b) 

    CO+·l→O=·–Cl

    A particular case of (a:

    (2′) Decomposition reactions (i.e. the inverse of addition) :

    (a

    (b

    ·OCl→CO+·l

    (3) Isomerization, e.g.:

    H2·−CH2−CH3→H3C−·H−CH3

    All these elementary reactions, except those of types 2(b) and 2′(b), involve breaking one of the bonds of the initial molecules, and result in the formation of a new bond.

    The rates of all these processes are given by the formulae:

    (1)

    (unimolecular decomposition; isomerization), and

    (2)

    (bimolecular reactions; e.g. of radicals with molecules, either by substitution or addition).

    The rate constant is a numerical expression of the reactivity of a molecule when its various bonds are subjected to attack by a radical.

    The value of ε —the activation energy of the elementary process—indicates the minimum energy which the reacting system must possess in order that the reaction may take place. (The system may be a molecule or radical in unimolecular processes, or both the reacting particles in bimolecular processes.)

    The state of the system may be represented graphically, taking the reaction co-ordinate as abcissa, and the potential energy of the reacting system as ordinate. Then the activation energy will be dependent on the height of the potential barrier that has to be surmounted during the reaction.

    If, as is shown in Fig. 1, the reaction is endothermic, then the activation energy ε will be composed of two terms: firstly, ε0, which is strictly the height of the potential barrier; and, secondly, the difference in energy between the final and initial states, i.e. the absolute value of the heat of reaction, |q|:

    Fig. 1

    (3)

    If the process is exothermic (such as the reverse reaction to that ofFig. 1), then the activation energy will simply be equal to the height of the potential barrier ε0. The difference in activation energy between the forward and back reactions is equal to the heat of the forward reaction. The factor a1 is about 10¹³ sec−1 in many unimolecular reactions; this is of the order of molecular vibrational frequencies.

    The values of a1 for molecules and radicals are frequently assumed not to differ widely. In support of this assumption, we have the direct experiments of Butler and Polanyi [1], on the initial decomposition of iodides into two free radicals, and also the results of Ladacki and Szwarc [2] on the decomposition of a range of organic compounds; in these reactions, a1 was found to be about 10¹³ sec−1.

    In bimolecular reactions a2 is often close to the collisional frequency z (between molecules and radicals), i.e. about 10−10 cm³/sec (or 10¹⁴ cm³/mole-sec, depending on the unit of concentration used).

    In fact, however, not every collision between a radical and a molecule leads to reaction, even if the energy of the colliding particles is sufficient; consequently, a2 = fz, where f is the so-called steric factor (f < 1 normally). In most reactions 0·1 < f < 1. Latterly, it has been shown (mainly by Steacie et al. [3, 4, 5]) that f falls to 10−3−10−4 in some reactions of hydrocarbon radicals and hydrogen atoms with alkanes. The problem of the f-values in these reactions cannot yet be considered as definitely decided, however. Extensive experimental data on free radicals were presented at the Faraday Society Discussion in September, 1952. The results of different workers were sometimes inconsistent; e.g. according to Berlie and Le Roy [6], ε is 6·8 kcal *, and f = 4·8 × 10−3, in the reaction between atomic hydrogen and ethane. For the same reaction (D + C2H6) Darwent and Roberts [7] found ε = 9 kcal, and f = 0·6. We propose to use the latter values, since we can see no special theoretical reason for the occurrence of very small f-values in substitution reactions.†

    Addition reactions (radicals or atoms) at double bonds would seem always to have low f-values (10−3−10−5). Numerous experimental results, both direct and indirect, bear this out. Theoretical reasons may be advanced for the small f-values in reactions of this type; these are based on the statistical theory of the activated complex. It is found that, in this case, the activated complex, unlike in substitution reactions, does not decompose and its entropy is lower than that of the initial state. The smallness of f thus follows by analogy with corresponding data for the equilibrium constant in addition reactions: the equilibrium constant can be expressed with a degree of accuracy which is sufficient for this purpose, in the formml: K = A exp (− Q /RTC2H5Br, A is about 10²⁷, being equal to the ratio a1/a2 (the a-values for decomposition and addition respectively): A = 10²⁷ = a1/a2. In many reactions, a 10¹³, so a10−14; consequently, f 10−4.

    By analogy with molecular additions, the steric factors in radical additions are supposed small.

    1 atm, unimolecular reactions (decomposition or isomerization) proceed considerably more rapidly than bimolecular reactions (substitution or addition), temperature and activation energy being equal. The rates of unimolecular reactions are given by:

    where c is the number of molecules per cubic centimetre; c being about 10¹⁹ at atmospheric pressure. The rates of bimolecular reactions are given by:

    Then, if

    i.e. under the same conditions, unimolecular reactions are about 10⁴ times more rapid than bimolecular ones.

    2. EXPERIMENTAL DETERMINATION OF ACTIVATION ENERGIES

    The quantity exp (– ε /RT) in the expression for the rate constant will now be examined. The magnitude of this factor is determined by the value of ε. The latter quantity is, as we have seen above, equal to the potential barrier height ε0 in exothermic reactions. When radicals react with molecules, or when radicals decompose, ε0 < 10 kcal usually; most frequently, it lies in the range 3–6 kcal. Strongly endothermic radical reactions may have much larger activation energies. For instance, the decomposition of alkyl radicals to olefin + H (e.g. CH3−CH2 → C2H4 + H) may involve an activation energy of about 40 kcal. These reactions are endothermic, the energy absorbed in the forward reaction being about 38 kcal. Now, according to (3), ε0 = ε – |q|, so ε0 is not great in these reactions, being only about 2 kcal. The reverse (exothermic) reaction—addition of a hydrogen atom at the double bond in olefins—has been found by direct experiment to have a very small activation energy, about 2–4 kcal [7, 9, 10].

    The endothermic reaction Br + H2 → HBr + H has an ε of 17·6 kcal: since we have, from (3), ε = ε0 + |q| and q = − 16·4 kcal, ε0 = 1·2 kcal.

    Table 1 gives values of ε0 determined for exothermic reactions of hydrogen atoms with various molecules. In Tables 2, 3, 4, and 5 the activation energies which have been determined for some other radical substitution reactions are given, together with the calculated values for the heats of reaction q.

    Table 1

    Heat of reaction and activation energy for some elementary reactions involving hydrogen atoms

    *q is an expression of the thermochemical value of the reaction: for exothermic reactions it will therefore be positive.

    Table 2

    Heat of reaction and activation energy for some elementary reactions with hydroxyl radicals

    Table 3

    Heat of reaction and activation energy for some elementary radical reactions

    Table 4

    Heat of reaction and activation energy for some elementary reactions with hydrogen or halogen atoms

    *Indicates that the value given was derived by indirect methods. Unreliable values are indicated by parentheses.

    Table 5

    *Heat of reaction and activation energy for some elementary reactions of sodium atoms

    *Note that the reactions of sodium with bromine derivatives of hydrocarbons usually proceed rapidly; in Fig. 2, page 27, the experimental points usually lie in the vicinity of the lower limit of the straight line 1. (The equations for these lines are given by ε0=A −α |q|.) Data for reactions of sodium with iodides always lie considerably below the limits of the band defined by the lines 1 and 3.

    It is very difficult to measure the activation energies of elementary free radical reactions; much more so than for reactions between molecules. Free radicals are unstable, and disappear rapidly by combination, either on the walls of the vessel, or by homogeneous reaction. In consequence, it is very difficult to maintain a constant concentration of free radicals, and to measure it especially in the gas phase.

    Free radicals are commonly obtained from electrical discharges in gases, by the action of ultra-violet light, or by pyrolysis of compounds that decompose to give free radicals (such as peroxides). Direct quantitative determination of the free radical concentration is usually extremely difficult, particularly as the concentrations normally encountered are very small indeed. For this reason, the concentrations and rates of reaction are usually deduced from the rates of formation of stable intermediates, or of final products. In addition to the primary reaction between the radical and molecule in question, secondary reactions (between the radicals resulting from the primary reaction and other molecules) also take place.

    These difficulties make determination of the activation energies of radical reactions so troublesome that it has so far only been undertaken in a limited number of cases—principally for the simplest radicals. For some of these, the activation energies are known with a fair degree of precision (0·5−1 kcal). In other cases, various authors give differing results for the same reaction, the discrepancies being as large as 2–3 kcal. Let us take, for example, the results of Avramenko and Lorentso [15–17] on the activation energies when OH radicals react with various molecules. They used the line spectra of OH, observed in absorption, to determine [OH] under steady-state conditions. (This method was developed by Kondratiev [23] and his school.) The basic principle of the method is that the rate constant may be deduced from the decrease in [OH] which takes place when the substance under investigation is introduced. The OH radicals are produced by passing a steady stream of water-vapour through a discharge. The greater the rate at which the molecules introduced react with the OH radicals, the greater will be the reduction in [OH]. [OH] is determined with the compound both present and absent. Since this method of observing the changes in [OH] is largely specific to OH, it is not subject to interference from other atoms (such as oxygen or hydrogen) which may be present.

    To determine the rate constant of the reaction, use is made of the formula that makes allowance for the homogeneous and heterogeneous reactions of OH. The rate constant for the recombination reaction has been given by Kondratiev and Ziskin [25], and also by Smith [26]. At each temperature, the rate constant for the reaction between OH and the substance in question was determined, and the activation energy deduced from the temperature coefficient of the rate constant.

    In this way, the absolute values of the rate constants and the energies of activation were obtained for the reactions of free OH with hydrogen and CO [17], with CH4, C2H6, C2H4 and C2H2 [15], and with HCHO and CH3CHO [16].

    The applicability of the method is limited by the need to work at low pressures when the radicals are produced in discharges. Photochemical methods are free from this disadvantage, but the radical concentration is small when light of ordinary intensities is used. The production of very high concentrations of radicals within extremely short periods of time was demonstrated by Norrish and Porter [27, 28], using flash photolysis. The use of very high light intensities (10²¹ quanta in 1 msec) enabled them to obtain large photochemical effects in a time comparable with the lives of radicals or of excited states of molecules.

    The general arrangements used in flash photolysis are as follows : a quartz reaction vessel is illuminated by a discharge of very high instantaneous power (10,000 joules dissipated in 0·1−3 msec), giving about 10²¹ quanta per msec. The intensity is sufficiently high to dissociate the gas almost completely into atoms and radicals. Subsequent to the exciting flash, absorption spectra are taken at intervals of 10–100 msec, to investigate the kinetics of the radical reactions. The absorption spectra of a number of radicals were discovered by flash photolysis; those of C3, NH2, C, CH, CN, CS, SO, and ClO being amongst them [29]. The most detailed studies were those of Porter and Wright [30] on the reactions of ClO. This radical was formed by combination of chlorine atoms with oxygen, apparently according to the following scheme :

    (1) 

    (2) 

    The amount of ClO formed did not depend on T in the range 25–300° C.

    The ClO radicals then react:

    (3) 

    Porter and Writght showed that (3) was of second order, with a rate constant:

    Such a small steric factor (10–6) is unusual in radical reactions.

    Using flash photolysis, Christie et al. [31] studied the influence of different inert gases on the recombination of iodine atoms; in these experiments, the iodine was about 80 per cent dissociated. They found that I2 is extremely effective in causing iodine atoms to recombine (see Chapter 4).

    Flash photolysis enables us to investigate the constants of elementary reactions, the kinetics of complex reactions, and to discover new radicals and the modes of formation of excited molecular states.

    3. BOND DISSOCIATION ENERGIES

    The difficulties encountered in determining heats of reaction for radicals are considerably greater than for molecules. Direct thermochemical determination is practically impossible in the vast majority of cases. The heat of reaction, q, is therefore calculated as the difference between the energy liberated in the formation of the new bonds in the products, and the dissociation energies of the bonds in the initial molecules. For instance, q in H+Cl2→HCl+Cl+q, is equal to QH–Cl–QCl—Cl, where Q is the dissociation energy of the corresponding bond. The Q-values for the Cl–Cl and H–Cl bonds are well known, being derived from spectroscopic and thermodynamic data, and are 57 kcal and 102 kcal respectively. Consequently, q is 102−57=45 kcal. Unfortunately, such direct methods of determining the Q-values of bonds are only applicable in the simplest molecules. Even in a molecule as simple as methane, it is difficult to deduce the Q-values of the C–H bonds. It is not permissible to use mean values for the bond energies in this case (all that can be deduced from the heat of combustion). It is well known that the energy absorbed by the abstraction of the first hydrogen atom differs substantially from the mean bond dissociation energy for all four hydrogen atoms. QC−H for the first hydrogen atom must be determined by direct experiment. A number of different methods have been applied (photobromination of methane, pyrolysis of methyl iodide, electron impact, etc.), and it has been shown that QCH3—H is 101±2 kcal; i.e. it is larger than the mean energy of the four C–H bonds in CH4.*

    Q0), so that the activation energy† determined from the pyrolysis of CH3Br is directly equal to QC–Br. Experiments on these lines gave [33] QCH3–Br=67 kcal. Thus we have:

    CH3Br→CH3+Br−67 kcal

    Since QBr−Br is 45·2 kcal, this may be written as:

    From published tables [34]‡ of the heats of formation of compounds from their elements, we find ΔHCH3Br=−8·10 kcal; for bromine vapour ΔHThen,−8·10=x+3·80−44·4, where xHCH3. ΔHCH3=32·5 kcal. Knowing this, it is easy to calculate QCH3−H. ΔHCH4 is −17·89 kcal, and ΔHH is 51.90 kcal. Then:

    QCH3−=ΔHCH3+ΔHH−ΔHCH4=102·3 kcal

    **

    And so the energy required to abstract the first hydrogen atom from methane is 102 kcal.

    Knowing ΔHCH3, QC–C in ethane may be evaluated, from ΔHC2H6. The dissociation energies of the C–Br bonds in CH3Br, CH2Br2, CHBr3 and CBr4 have been determined from decomposition experiments [33], and from these we may deduce the ΔH values for CH2Br, CHBr2 and CBr3, in a similar fashion; and thus arrive at QC–H in the corresponding bromides without recourse to direct experiment.

    Consider, for instance, the reaction CH3Br → CH2Br+H. ΔHCH3Br and ΔHH are known from the tables. Since ΔHCH2Br is computed from experimental data on the decomposition of CH2Br2, we may also compute QC–H for the first hydrogen atom in CH3Br.

    The examples given suffice to show that, by determining the dissociation energy of a given bond in a compound, we are able to deduce the energies of the other bonds in the compound (and even the bond dissociation energies in some other compounds) from the thermochemical values for the heats of formation of compounds from their elements.

    The strengths of C–I bonds have been obtained by the pyrolysis of organic iodides. By methods analogous to that given above, the C–H bond energies in various hydrocarbons were calculated. Polanyi [1] obtained 102·5 kcal for QCH3−H in this way.

    The most exact determinations of the C−H bond energies in CH4, C2H6, and (CH3)4C are those by Kistiakowsky, et al. [22, 35–37] from their photobromination. The activation energies of the forward and reverse reactions (ε1 and ε1′) were determined.

    (1) RH+Br→R+HBr+q

    (1′) R+HBr→RH+Br

    q in (1) was equal to the difference between the activation energies: q=ε1−ε1′. On the other hand, q=QRH−QHBr. Consequently,

    ε1−ε1′=QR−H−QH−Br

    QR−H=QH−Br+ε1−ε1′

    QH−Br is taken from the tables. For CH3Br, ε1=18 kcal, and ε1′=2 kcal. Hence q=18−2=16 kcal, at 453° K, and 15 kcal at 0° K. QH–Br=85·8 kcal, so QCH3−H=85·8+15 ≈ 101 kcal.

    Bond dissociation energies are determined by electron impact methods [38, 39], by pyrolysis in a stream of toluene (Szwarc [40]), and by the use of metallic mirrors (Rice and Dooley [41]) spectroscopically and recently by ionic impact [42].

    Molecular dissociation due to collision with electrons is observed in electron impact studies. The basic principles of the method were established by Frank [43]. Kondratiev [44] was the first to use the mass spectrometer for this purpose. The changes that occur as the electron energy is varied are followed by mass-spectrometric methods; so-called ionization curves are drawn up, i.e. curves showing the ion current for a particular species as a function of the electron energy. The electron energy at which the spectrometer begins to register that ion current is called the appearance potential. When the ion is formed without dissociation, the appearance potential is equal to the ionization potential. When an ion R1+ is formed by dissociation of the molecule R1−R2 via the process :

    the appearance potential may be expressed in the form :

    where AR1+ is the appearance potential for R1+, QR1–R2 is the dissociation energy of the bond R1–R2, IR1 is the ionization potential of the fragment R1, K is the kinetic energy carried off by R1+ and R2, and W is the excitation energy of R1+ and the fragment R2. Determination of K is always difficult, and demands special experiments; its value is always inaccurate. W can be determined only in rare, very simple cases [38]. Therefore, the value of QR1–R2 is never very accurate, even if IR1 is known from other data.

    Q may be determined from the appearance potentials in two different processes, e.g.:

    q in C2H6+CH4 → C3H8+2H+q being known from thermochemical data, QCH3−H may be deduced if we neglect possible differences in K and W in (1) and (2).

    This is possible if different pairs of processes of types (1) and (2) give consistent results. The best accuracy attained was ∼ 2 kcal.

    Data on the dissociation energies of molecules and ions can sometimes be obtained by the ionic impact method suggested by Talrose and Frankevitch [42] in 1956, dissociation energies being evaluated by comparing various ion-molecule reactions occurring in the mass spectrometer ion source at 10−3–10⁴ mm Hg. This method was used for reactions of the type

    accompanied by proton or H atom transfer. Such reactions have no activation energy [45] (see paragraph 11) and the mere fact that the mass-spectrometer records secondary ions shows that the heat of reaction is positive or 0±2 kcal

    where PM is the affinity of M for the proton. If secondary MH+from the RH++M collision is not found Q < 0. Two such processes have been shown to

    Enjoying the preview?
    Page 1 of 1