You are on page 1of 373

Thermal Decomposition and Combustion

of Explosives and Propellants


Thermal Decomposition and
Combustion of Explosives
and Propellants

G.B. Manelis

G.M. Nazin

Yu.I. Rubtsov

and

V.A. Strunin

Taylor &. Francis


Taylor & Francis Group

LONDON AND NEW YORK


First published 2003 by Taylor & Francis
11 New Fetter Lane, London EC4P 4EE

Simultaneously published in the USA and Canada


by Taylor & Francis Inc,
29 West 35th Street, New York, NY 10001

Taylor & Francis is an imprint of the Taylor & Francis Group

© 2003 Taylor & Francis

This book has been produced from camera-ready copy supplied


by the authors
Printed and bound in Great Britain by
MPG Books Ltd, Bodmin

All rights reserved. No part of this book may be reprinted or reproduced or


utilised in any form or by any electronic, mechanical, or other means, now
known or hereafter invented, including photocopying and recording, or in any
information storage or retrieval system, without permission in writing from
the publishers.

Every effort has been made to ensure that the advice and information in this
book is true and accurate at the time of going to press. However, neither the
publisher nor the authors can accept any legal responsibility or liability for
any errors or omissions that may be made. In the case of drug administration,
any medical procedure or the use of technical equipment mentioned within
this book, you are strongly advised to consult the manufacturer's guidelines.

British Library Cataloguing in Publication Data


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

Library of Congress Cataloging in Publication Data


A catalogue record for this book has been requested

ISBN 0-415-29984-5
CONTENTS

PREFACE ix

Chapter 1 INTRODUCTION. KINETIC PROBLEMS OF 1


DECOMPOSITION OF EXPLOSIVES AND
PROPELLANTS
REFERENCES 7
Chapter 2 GENERAL REGULARITIES OF T H E R M A L 9
DECOMPOSITION OF HIGH-ENERGY
COMPOUNDS
2.1 Gas phase reactions 10
2.2 L i q u i d phase reactions 12
2.3 Decomposition in the solid state 16
REFERENCES 32
Chapter 3 ALIPHATIC NITROCOMPOUNDS 35
3.1 Gas phase decomposition by the molecular mechanism 36
Mononitroalkanes. Homogeneous stages of 36
decomposition
The pre-exponential factor 38
The polar character of the reaction of HN0 2 38
elimination
a-Halogen-nitroalkanes 39
a-Nitroolefins 41
3.2 Gas phase decomposition by the radical mechanism 42
The chemical mechanism 42
The activation energy of decomposition and C — / V 47
bond strength in poly nitrocompounds
Regularities of variation of the pre-exponential factor 50
3.3 Competition o f the radical and molecular mechanisms 51
Nitro-nitrite rearrangement 51
3.4 Reactions in the condensed state 53
REFERENCES 56
Chapter 4 AROMATIC NITROCOMPOUNDS 61
4.1 Gas phase decomposition by the radical mechanism 61
4.2 Molecular mechanism o f decomposition 65
4.3 Decomposition in the liquid state 66
4.4 Decomposition in the solid state 69
REFERENCES 71

v
vi Contents

Chapter 5 SECONDARY NITROAMINES 73


5.1 First stage o f decomposition 73
5.2 Secondary reactions 75
5.3 Kinetic data 79
REFERENCES 90
Chapter 6 ORGANIC AZIDES 95
6.1 Aliphatic azides 96
6.2 Aromatic and heterocyclic azides 99
REFERENCES 102
Chapter 7 ORGANIC DIFLUOROAMINO COMPOUNDS 105
REFERENCES 111
Chapter 8 HETEROCYCLIC COMPOUNDS 113
8.1 Thermal decomposition o f furazanes and furoxanes 113
Chemical mechanism of decomposition 113
Kinetic data 115
8.2 Stability o f tetrazoles 119
/, 5-Disubstituted tetrazoles 119
2,5-Substituted tetrazoles 121
Tetrazole and 5-substituted tetrazoles 121
REFERENCES 123
Chapter 9 NITROESTERS 127
REFERENCES 132
Chapter 10 COMPOUNDS WITH MIXED FUNCTIONS 135
REFERENCES 139
Chapter 11 GENERAL REGULARITIES OF T H E R M A L 141
DECOMPOSITION O F ONIUM SALTS. NITRIC
AND P E R C H L O R I C ACIDS, DINITRAMIDE
11.1 General regularities 141
11.2 Nitric acid 146
11.3 Perchloric acid 153
11.4 Dinitramide 159
REFERENCES 172
Chapter 12 AMMONIUM NITRATE 175
12.1 General regularities 175
12.2 Influence o f water, acid, and ammonia on the 177
decomposition rate
12.3 The heat and macrokinetic regularities o f 179
decomposition
12.4 Thermal decomposition below the melting temperature 181
12.5 Influence o f additives on the rate o f thermal 183
decomposition
REFERENCES 187
Chapter 13 AMMONIUM PERCHLORATE 189
13.1 Kinetics o f the low-temperature decomposition 189
13.2 Kinetics o f the high-temperature decomposition 194
13.3 Influence o f excessive acids and bases 196
Contents vii

13.4 Topochemical peculiarities o f thermal decomposition 198


13.5 Influence o f the preliminary irradiation and additives 204
13.6 Thermal decomposition o f composite AP-based 210
systems
REFERENCES 219
Chapter 14 A M M O N I U M DINITRAM1DE 221
14.1 Decomposition in the melt 221
14.2 Decomposition in the solid phase 227
REFERENCES 230
Chapter 15 T H E SALTS OF HYDRAZINIUM, 231
H Y D R O X Y L A M M O N I U M , AND NITRONIUM
15.1 Hydrazinium nitrate and chloride 231
15.2 Hydrazinium iodide 234
15.3 Hydrazinium azide 235
15.4 Hydrazinium perchlorate and diperchlorate 236
15.5 Hydroxylammonium sulphate and phosphate 238
15.6 Hydroxylammonium chloride 240
15.7 Hydroxylammonium perchlorate and its 241
hydroxylamine complex
15.8 Hydroxylammonium nitrate 243
15.9 Nitronium perchlorate 245
REFERENCES 248
Chapter 16 M E T A L P E R C H L O R A T E S AND NITRATES, 251
M E T A L SALTS O F DINITRAMIDE
16.1 Metal perchlorates 251
16.2 Metal nitrates 255
16.3 Metal salts o f dinitramide. Decomposition o f potassium 257
salt in the liquid state
16.4 Decomposition o f potassium salt in the solid phase 260
16.5 Dinitramide salts o f other metals 263
REFERENCES 266
Chapter 17 BASIC A S P E C T S O F T H E C O M B U S T I O N 267
MECHANISM
Chapter 18 COMBUSTION OF PURE SUBSTANCES: 271
R E A C T I O N S IN T H E C O N D E N S E D P H A S E
18.1 The model with dispersion o f the solid substance 271
18.2 The model with foaming o f the reacting substance 273
18.3 The model with evaporation or sublimation 275
Ammonium perchlorate (AP), NH C10 4 4 280
Hydroxylammonium perchlorate (HAP),NH 0HC10 3 4 282
Ammonium nitrate (AN), NH N0 4 3 285
Hydrazinium nitrate (HN), N H N0
2 5 3 286
Hydrazinium chloride (HQ, N H Cl 2 5 288
Ammonium dinitramide (ADN), NH [N(N0 )] 4 2 2 290
REFERENCES 294
viii Contents

Chapter 19 COMBUSTION O F PURE SUBSTANCES: 297


R E A C T I O N S IN T H E G A S P H A S E
REFERENCES 304
Chapter 20 COMBUSTION O F PURE SUBSTANCES: 307
R E A C T I O N S IN T H E C O N D E N S E D A N D G A S
PHASES
REFERENCES 314
Chapter 21 COMBUSTION O FCONDENSED COMPOSITE 315
SYSTEMS
21.1 Quasihomogeneous composites 315
21.2 Layered systems 321
Chemical arc "
il
321
A "sandwich type" system 327
21.3 Composite propellants 339
REFERENCES 345
A U T H O R INDEX 347
S U B J E C T INDEX 349
PREFACE

The study o f thermal decomposition, thermal explosion, the combustion


process o f high-energy substances, as well as o f explosives and composites,
powders and propellants prepared on their basis, is o f obvious practical
importance. A t the same time, investigations carried out in this direction are
of considerable fundamental importance. The most important problem o f
the relation o f the molecular structure to the reactivity has been solved.
Great progress has been achieved in the chemistry o f solids and in the
theory o f solid phase reactions, as well as in the theory o f thermal explosion
and combustion o f condensed substances.
The works o f N . N . Semenov and his followers, Y u . B . Khariton, Y a . B .
Z e l ' d o v i c h , A . F. Belyaev, P. F. Pokhil, A . Y a . A p i n and other workers o f
the Institute o f Problems o f Chemical Physics, provide the basis for these
studies.
Starting in the late fifties, studies in the field o f explosives were
conducted most actively in the Institute o f Chemical Physics (ICP),
Academy o f Sciences o f the U S S R . Later, extensive studies o f powders and
rocket propellants were initiated. From 1960, these studies were carried out
mostly at the Branch o f ICP in Chernogolovka, which was later
reorganized into the Institute o f Chemical Physics in Chernogolovka,
Russian Academy o f Sciences.
Over the years, a large team o f scientists engaged in this exciting field o f
sciences was formed.
A complex approach to the problem, the revealing and consideration o f
the fundamental problems are the characteristic features o f the scientific
school o f I C P . This approach clearly manifested itself in the studies
presented in this book.
The quantitative kinetics o f chemical reactions in combination with an
accurate study o f the chemical mechanism provided the basis for these
investigations. M u c h attention was given to the physical processes, which
proceed simultaneously with the chemical transformation: phase transitions,
the variation o f the actual structure o f solids, and the heat and mass
exchange. Such an approach required the application, and sometimes the
development, o f a new experimental technique, obtaining quantitative and
objective information in a form, that would allow the results to be employed
in theoretical studies to predict the behavior o f the reacted systems under
various conditions. The results obtained made it possible to predict
quantitatively the stability o f the products and to carry out technological
calculations which have been successfully employed in the calculation o f
thermal explosion and ignition, in the development o f the quantitative
theory o f combustion o f high-energy substances and the systems prepared
on their basis.

IX
X Preface

The book presents the results o f work carried out in the I C P in the past
ten years. This book is a translation o f the Russian edition (1996), which
was revised and updated. The chapters on the combustion and
decomposition o f dinitramide salts are new. The studies o f these compounds
were started in 1970, but comprehensive publications appeared only in the
late 1990s.
This monograph does not pretend to be a complete presentation o f
numerous studies carried out in many countries. However, the leading
position o f I C P in this field o f science, the original approach o f the authors
to the problems considered, the essence o f which is the successive
consideration o f the problems o f stability, safety, and combustion from a
unique kinetic point o f view, a combination o f the fundamental approach
with the quantitative solution o f applied problems, and a successive
consideration o f major points make this book useful for a wide audience
engaged both in the field o f chemical kinetics, synthesis, and studies o f
high-energy compounds, the theory o f combustion and explosion, and in the
field o f explosives, powders, and solid rocket propellants.

F. I. Dubovitskii, Corresponding Member, Russian Academy o f Sciences

Chernogolovka, June 1998


Chapter 1

INTRODUCTION. KINETIC PROBLEMS


OF DECOMPOSITION OF EXPLOSIVES
AND PROPELLANTS

A quantitative kinetic study o f the decomposition processes o f explosives


and propellants, started by the pioneer works o f Robertson (1909) [1.1],
Farmer (1920) [1.2], Hinshelwood (1921) [1.3], Garner (1931) [1.4], and
Sapozhnikov and Roginskii (1930, 1931) [1.5, 1.6], was intensively
developed after World War II in many countries all over the world and
became progressively a very wide and multi-branched scientific direction.
The Semenov Institute o f Chemical Physics o f the Russian Academy o f
Sciences founded by N . N . Semenov is one o f the first-rate centers o f such
studies. Studies o f the thermal decomposition o f high-energy compounds
were conducted at this Institute from the very beginning of its
establishment. Extensive systematic studies in this direction were carried
out at the Semenov Institute o f Chemical Physics under the general
guidance o f F. I. Dubovitskii from the mid-1950s in M o s c o w and from the
beginning o f the sixties at the Branch o f the Semenov Institute o f Chemical
Physics in Chernogolovka, which was later reorganized into the Institute o f
Problems o f Chemical Physics in Chernogolovka o f the Russian A c a d e m y
o f Sciences. This monograph, written by researchers o f this Institute,
presents mainly the results o f studies that have been carried out at the
Institute o f Problems o f Chemical Physics in Chernogolovka o f the Russian
Academy o f Sciences over the past 36 years.
The development o f studies on thermal decomposition o f high-energy
compounds requires the solution o f a number o f fundamental problems o f
chemical kinetics and chemical physics.
A l l high-energy substances are thermodynamically unstable and the
possibility o f their existence is caused by kinetic reasons. The rate constant
of decomposition in the gas phase, when the influence o f the medium,
crystal lattice, interaction o f molecules with one another, etc., are
eliminated, is the most accurate characteristic o f molecule stability.

1
2 Kinetic Problems of Decomposition

During the decomposition o f high-energy compounds in the gas phase,


nearly all known decomposition mechanisms are realized: monomolecular
decay with the formation o f free radicals; elimination; rearrangements
followed by decomposition; opening o f cycles, etc.
Since high-energy substances are, as a rule, complex multifunctional
compounds, it is important to determine the center o f the reaction and to
establish the relation o f the molecule structure with their reactivity. The
competition o f various transformation pathways is, first o f all, determined
by the structure o f decomposing molecules, but, sometimes, variation o f the
decomposition mechanism may also occur on variation o f the conditions
under which the process occurs, for instance, temperature.
In many cases, the rate and characteristic features o f the reaction o f
thermal decomposition are preserved on transition from the gas to the liquid
phase. A t the same time, the possibility o f realization o f new pathways o f
decomposition arises in the liquid phase, which, first o f a l l , are connected
with the acid-base processes, with occurring ionic reactions, with the
chelation processes, etc. During decomposition o f explosives, chain-radical
and even chain-branching reactions, in which the acid-base processes play
an important role, have been observed.
The processes in the solid phase have a number of essential
peculiarities, which required further development o f the general theory o f
reactions in the solid phase.
The increase in volume during formation o f an activated complex in the
crystal lattice gives rise to a noticeable variation o f the reaction rate
constant in the solid phase in comparison with the liquid phase due to the
necessity o f lattice deformation. Various lattice imperfections, primarily
dislocations, as well as the variation o f the crystal state at the melting
temperature ("pre-melting") play an important role. Variation o f the crystal
lattice state as a result o f chemical reaction and formation of the
decomposition products leads to the localization process, to the appearance
o f auto-oscillating regimes o f reaction propagation in the crystal lattice, etc.
A l l these result in variations o f the characteristic features o f the kinetics,
absolute reaction rates, and activation energies. Under certain conditions,
the temperature dependence o f the transformation rate is not governed by
the Arrhenius equation and is o f a complex character.
Macroscopic characteristics of decomposition of high-energy
substances also have a number o f peculiarities. The increase in the volume
o f reactants during decomposition changes the formal regularities o f the
time evolution o f the process because the products o f decomposition are
gases. The appearance o f the liquid phase during decomposition o f solid
substances (a progressive dissolution o f the parent substance in the
condensed products o f the reaction) also changes the observable formal
regularities and temperature dependence o f the reaction rate.
A catalytic influence o f the volatile products o f decomposition on the
reaction rate in the liquid or solid phase results in the appearance o f the
Thermal Decomposition and Combustion 3

dependence o f the rates o f transformation on the value o f the free volume o f


the reactor and on the conditions o f elimination o f gaseous and volatile
reaction products. In this case, one also can observe critical phenomena o f
nonthermal character.
A l o n g with the solution o f fundamental problems, the study o f thermal
decomposition and combustion o f high-energy compounds is stimulated by
technical problems, related to the production and use o f explosives and
powders. There were many applied problems, which have been solved in
these studies. One o f the main characteristics o f explosives, powders or
their components is thermal stability. Its level has to guarantee the safety o f
production o f products containing explosives and preservation o f their
physical, chemical and ballistic properties for a long time. The
compatibility o f the components is also an important property o f explosives
and propellants.
The stability o f a substance or composition under given conditions may
be determined as the time during which the main characteristics (energetic,
ballistic, explosive, physical, chemical, etc.) and chemical composition
remain the same or their variation does not exceed some predetermined
values.
In this case, an important problem is to determine the most sensitive
characteristic and possible limits o f its variation, as well as its relation with
the degree o f decomposition and, sometimes, with the rate o f thermal
decomposition. In general, the solution o f this problem is a very
complicated issue, which is far beyond the framework o f this book.
It is necessary to note only a few major aspects. In most cases (solid
propellants and powders, charges o f explosives), the variation o f the
physical-chemical properties or even the charge continuity loss are the
processes which determine stability. In turn, they depend on the processes
in the polymer binder and are strongly activated by stresses in the product.
Strains and stresses in the charge are determined by the product design, the
temperature and gas emission as a result o f thermal decomposition o f high-
energy components. In addition, the thermal decomposition of the
components o f composition may lead to the formation o f pores, which
changes the physical-chemical properties o f the explosive composition or
powder. A l s o , it may change the ballistic characteristics, stimulate the
transition o f combustion in detonation, and cause an increase o f sensitivity
to mechanical impacts. Intermediate and final products of thermal
decomposition may oxidize the polymer binder, which also results in a
strong variation o f the physical-chemical characteristics.
The processes occurring in the polymer binders and the influence o f
stresses on these processes are an area o f special interest [1.7], and w i l l not
be considered in this monograph. However, in order to determine the
stability o f explosives and powders, one should take these processes into
account together with the influence o f thermal decomposition o f high-
energy components on them.
4 Kinetic Problems of Decomposition

For solid propellants, as well as for molten and molded blasting


charges, the criterion of stability with respect to the degree of
decomposition may be determined in the most general form as the limitation
o f the possible degree o f transformation during the time o f treatment and
storage o f the composition by the value o f 0.02%.
For such a degree o f transformation, all gaseous products w i l l be
dissolved in the condensed phase. The appearance o f an internal pressure,
the formation o f pores, and swelling o f the product are excluded. In the
calculation o f this, probably the most rigid criterion o f stability, it was taken
into account that in a complete decomposition o f explosive usually no more
than 100 c m 7 g o f such gases as N , C O , and N O are formed, and their
2

solubility in organic substances is approximately 2 vol % under the pressure


o f 1 atm. In specific cases, the requirements o f thermal stability may be
changed based on accurate data on the composition o f gases, their solubility
in the fuel bulk and the strength characteristics o f the material. Depending
on the design and service conditions o f blasting charges, this criterion o f
stability for the same composition may vary by a factor o f 100.
It is convenient to use the value o f T (20 ), 002
o
that is the time o f
approaching the degree o f transformation o f 0.02% at 2 0 ° C , to compare the
stability o f various substances and compositions.
The completeness and accuracy o f the stability test, first o f all, depend
on the thermal stability o f the substance itself. Usually, the warranty storage
time o f explosives is 4 0 - 6 0 years, and 15-20 years for rocket propellants
because o f the fast modification o f rockets. For R D X - t y p e substances,
trinitrotoluene ( T N T ) , tetryl, and pentaerythritol tetranitrate ( P E T N ) , which
have an activation energy o f decomposition higher than 40 kcal/mol, the
parameter r 2 ( 2 0 ° ) is as large as hundreds o f thousands o f years. For these
00

substances, the problem o f stability arises only in the case o f their


employment at high temperatures or in mixtures with reactive components.
However, many compounds, especially high-power explosives or high-
energy oxidizers, have a considerably lower stability. If the activation
energy o f decomposition is lower than 37 kcal/mol, very accurate
measurements o f stability are necessary, and, in the case when the
Arrhenius parameters E < 34 kcal/mol and A > 1 0 15
s , the substance can be
_1

placed in the class o f low stable substances.


If the decomposition rate is determined not only by the first
monomolecular stage, but also by the secondary autocatalytic or chain
reactions, then, in principle, chemical stabilization o f the system is possible
by means o f an appropriate inhibitor. Stabilization is the most difficult
problem o f the theory o f stability. Its realization depends not only on
success in understanding the nature o f acceleration reactions, but also on the
possibility o f the choice o f inhibitors, which have to meet a lot o f
requirements imposed on a component o f a solid rocket propellant.
To plan the synthesis of new explosives and preparation of
compositions it is extremely important to predict the stability and
Thermal Decomposition and Combustion 5

compatibility o f a new substance. The development o f such a prediction is


based on an array o f experimental data on the dependence o f stability on the
structure. The correlation dependences o f chemical kinetics, the analog
method, the quantum-mechanical calculations and semi-empirical relations
for estimation o f the activation energy, and the calculation o f the pre-
exponential factor from the R R K M theory are also employed. A t present,
the possibilities o f prediction may be estimated as rather good, and the
efficiency o f predictions w i l l increase as new data appears, especially on the
decomposition reactions o f complex compounds with hybrid functions.
The study o f the kinetics o f thermal decomposition and determination
of the stability are usually carried out with the use o f small charges o f the
substance studied under conditions o f a good heat exchange, when the
process may be considered as an isothermal one.
However, when actual products are manufactured, especially on a large
scale, or when the charges are used under elevated temperatures and the rate
of heat release during thermal decomposition becomes equal to or larger
than the rate o f heat removal from the system, self-ignition (thermal
explosion) occurs [1.8]. The phenomenon o f thermal explosion of
explosives is well studied. The theory is developed, which describes this
phenomenon quantitatively and makes possible engineering calculations
[1.9]. It should be noted that the complex mechanism of thermal
decomposition reactions leads to a situation where the rate o f gas separation
frequently does not coincide with the rate o f heat release. T o determine the
conditions o f the emergency o f heat explosion (the critical charge diameter,
critical temperature), as well as the induction period, quantitative data on
the kinetics o f heat release are required (the kinetic law, the values o f the
rate constants and activation energies, the thermal effect). The possibility o f
thermal explosion is usually underestimated, because under temperatures
slightly below the critical ones, and during the induction period, the reaction
of thermal decomposition is hardly visible, and the heating o f the system is
also small. For autocatalytic reactions, the induction period o f heat
explosion, especially near the critical conditions, may be rather long and
take several days, months, and sometimes even years (this is connected with
the characteristic features o f the kinetics o f thermal decomposition o f
explosives and rocket propellants such as strong self-acceleration, high and
various values o f activation energies, high thermal effect, etc.).
In fact, the process o f combustion o f explosives, powders, and solid
rocket propellants is also the process o f thermal decomposition. However,
the conditions under which thermal decomposition occurs on combustion
differ sufficiently from those under which the kinetics and decomposition
mechanism are usually studied. The process o f thermal decomposition on
combustion occurs both in the condensed (solid and liquid) and gas phase,
and is accompanied by heat transfer, diffusion, melting, evaporation, etc.
A wide temperature range and high magnitudes o f the maximal
temperatures, which are realized on combustion, as well as phase transitions
6 Kinetic Problems of Decomposition

on combustion (solid-liquid-gas) lead to a situation where the role o f


different stages o f thermal decomposition may be changed due to various
values o f activation energies, as well as to the competition and mutual
influence o f the reactions in the gas and condensed phases. In this
connection, attempts to obtain a direct correlation o f the rate and
combustion regularities with the macroscopic parameters o f the reactions o f
thermal decomposition obtained under rather low temperatures are not
correct and cannot provide the correct result.
Understanding o f the mechanism o f combustion o f explosives,
powders, and solid rocket propellants requires data on the reaction
mechanism o f thermal decomposition, as well as quantitative kinetic data,
which make it possible to describe the process o f thermal decomposition in
a wide temperature range. In a general case, it is necessary to have
information about the reactions o f decomposition in various aggregate
states.
Actually, the combustion is always a multistage process. First o f a l l ,
this is connected with the fact that the formation o f final products from the
parent substances proceeds through a number o f consecutive and parallel
reactions. In addition, the physical processes, which essentially influence
the rate and regularities o f heat release (absorption), also proceed
simultaneously: melting, evaporation (sublimation), and dispersion. A
spatial separation o f zones o f chemical transformation, which in a stationary
process move with the same velocities, may proceed because various
processes have various temperature dependences. In this case, it is not
necessary that the limiting stage turned out to be inside the zone o f maximal
combustion temperature. Then, the study o f the combustion mechanism and
understanding o f the main regularities o f the combustion front propagation
(the combustion rate, its dependence on pressure and initial temperature),
first o f all, require the determination o f the limiting stage, its localization in
space (the gas or condensed phases), as well as the temperature at which the
limiting chemical reaction proceeds.
In most theoretical studies o f combustion o f condensed systems, either
the maximal flame temperature determined from thermodynamics, or the
experimentally measured temperature in those zones where the existence o f
a limiting stage is supposed (for instance, the temperature o f a burning
surface) is used. In a self-consistent model, the determining temperature
should be found as a result o f calculation, in which the kinetics o f
corresponding chemical reactions has to be considered together with not
only heat transfer, but also regularities o f such physical processes as
evaporation, sublimation, and dispersion.
Such an approach requires a sufficiently detailed study o f the kinetics
o f chemical reactions proceeding on combustion, the dependence of
regularities o f transformation not only on temperature, but also on the phase
state together with the quantitative characteristics o f phase transitions.
A complex consideration o f the combustion process makes it possible
Thermal Decomposition and Combustion 7

not only to describe formally the observable regularities, but also to


separate the determining parameters by changing their values (for instance,
introducing catalysts and inhibitors); thus, one may vary the rate and
regularities o f combustion.

REFERENCES

1.1. R. Robertson, J. Chem. Soc, 95: 1241-1248 (1909).


1.2. R. C . Farmer, Ibid., 117:806-818, 1432-1445 (1920).
1.3. C . N . Hinshelwood, Ibid, 119-120: 421-434 (1921).
1.4. W . E. Garner, Ibid, 720-722 (1934).
1.5. A . V . Sapozhnikov and M . Borisova, Journal of Russian
Physical-Chemical Society, 36: 836-841 (1904).
1.6. S. Z . Roginskii, Russian Journal of Physical Chemistry, 3: 419-424
(1932).
1.7. G . B . Manelis and E. V . Polianchik, Sow Sci. Rev. B. Chem.,
5(5): 61-101 (1991).
1.8. D . A . Frank-Kamenetskii, Diffusion and Heat Transfer in Chemical
Kinetics, M . : Nauka, 1967, 491 p.
1.9. A . G . Merzhanov and V . G . Abramov, Prop. Explos. Pyrotechn.,
6: 130-148 (1981).
Chapter 2

GENERAL REGULARITIES OF
THERMAL DECOMPOSITION OF HIGH-
ENERGY COMPOUNDS

The character o f the study o f the decomposition reaction depends on the


problem formulated. T o determine the stability, compatibility and estimated
heat explosion it is sufficient to carry out the formal kinetic measurements,
which may be time-consuming but are always feasible. T o consider the
problem o f stabilization and stability prediction, more complicated studies
are required such as the determination o f the detailed reaction mechanism,
the influence o f the structure and state o f aggregation on the rate and
decomposition mechanism. Thus, the general problem of stability
incorporates both an in-depth study o f particular real explosives being used
in practice, and, even to a larger extent, the study o f the classical tructure~
tC

reactivity" problem with the use o f numerous model compounds, analogs,


and structural fragments o f real explosives.
Systematic studies o f the general regularities o f decomposition have
been conducted with substances o f the following classes: aliphatic and
aromatic nitrocompounds, secondary nitramines, nitroesters, difluoro-
amines, organic azides, furazans, furoxans, derivatives o f tetrazole, onium
salts o f chloric and nitric acids, and several other types o f substances.
Some results, especially those obtained in large reaction series, are
beyond the framework o f studies o f explosives themselves and touch on
certain general problems studied in chemical kinetics. The development o f
the theory o f reactions in the solid phase, determination o f the energy o f
bond breaking in complex molecules, experimental study o f the factors
influencing the value o f the pre-exponential factor o f monomolecular
reactions, certain new reactions of molecules and radicals, kinetic
regularities o f complex processes with equilibrium stages, new formal
kinetic methods o f reaction description, and critical phenomena of a
diffusion nature are among these problems. Considerable attention has been
given to these and some other general problems.

9
10 General Regularities of Thermal Decomposition

In studies o f decomposition reactions, the manometric and calorimetric


methods were widely used, which are rather sensitive and the most simple.
In particular, the manometric method makes it possible to observe the
reaction starting from the conversion level o f 0.01%, to measure the rate
constants under static conditions within the range o f 10~— 10 4 1 0
s" , and to
1

conduct a reaction at any arbitrary values o f the coefficient o f charge o f the


vessel filled by the substance (m/V, g/cm ) up to m/V = p. Simple equipment
J

makes possible an automatic recording o f the kinetic curve under pressures


of 10 —760 Torr or 1-100 atm. Up-to-date calorimeters have a sensitivity
2

o f 10~ W and permit one to measure directly the reaction rates, which
7

correspond to a rate o f less than 0.1% per year i f the total heat o f the
process exceeds 100 cal/g. The methods o f thermogravimetry and
differential thermal analysis were used as auxiliary ones. Analysis o f the
products and parent substances was used to check the correctness o f these
methods and to obtain data concerning the reaction mechanism. For this
purpose, we also used some other kinetic techniques: high-vacuum
pyrolysis, the radioactive tracer labeling method, traps o f intermediate
products, a strong dilution by the inert solvent, introduction o f inhibitors
and promotors o f the reactions. A combination o f several methods and
attraction o f indirect data (thermochemical calculations and correlation
relationships), as a rule, makes it possible to avoid errors and to prove the
reaction mechanism.

2.1 Gas Phase Reactions


The data on decomposition o f vapors o f explosives are widely used to
interpret the combustion processes. However, the majority o f works on the
gas phase decomposition are intended to determine the reaction mechanism
under the simplest conditions without the influence o f the environment.
During decomposition o f explosives, all types o f homolytic monomolecular
reactions can be observed: the breaking o f a simple bond to form free
radicals (polynitroalkanes and nitramines) or nitrenes (azides),
rearrangements (nitrocompound -> nitronic acid), a- or /^-elimination
through 4-, 5-, or 6-membered cycles to form unsaturated compounds
(mononitroalkanes), step-by-step or simultaneous ring opening (furoxanes,
tetrazoles), which often compete with each other and are complicated by
chain and heterogeneous processes.
Due to the possibility o f varying the conditions under which the
reaction occurs in the gas phase (pressure, temperature, contact time,
inhibitor, and surface-to-volume ratio) within a wide range, it is possible to
separate the first monomolecular stage o f decomposition from a complex
process and to determine its kinetic characteristics.
Sometimes, the monomolecular character o f decomposition is preserved
over the whole reaction. These cases can be easily recognized by an
Thermal Decomposition and Combustion 11

absolute independence o f the rate and composition o f the reaction products


on pressure. Only particular reactions of elimination, simultaneous
decomposition o f cycles, and, also, radical decomposition, i f the latter are
unstable and disappear, for instance, due to intermolecular oxidation or i f
along with their formation efficient inhibitors ( N O , N 0 ) o f chain processes 2

are formed, proceed according to a pure monomolecular mechanism. The


reactions o f decomposition o f polynitroalkanes, azides, and furoxanes are
among them. In the majority o f cases, the secondary reactions affect the
parent substance causing the chain decomposition [CH N0 ,
3 2 sec-
nitramines, A r N 0 , and R N ( N 0 ) ] . A decrease o f the initial pressure
2 2 2

facilitates the linear termination on the wall and, as a rule, makes it possible
to eliminate the chain reactions.
If, for any particular reasons, decomposition in the liquid phase
proceeds faster than in vapors, a heterogeneous reaction on the vessel
surface becomes noticeable against the background of gas phase
decomposition. For large surface-to-volume (S/V) ratios, the contribution o f
this reaction may be very large. Heterogeneous reactions are typical o f
aromatic nitrocompounds and difluoroamines. The role o f heterogeneous
processes decreases with the temperature increase and with the decrease o f
the initial pressure p and the S/V ratio. In the case o f combination o f the
0

chain and heterogeneous reactions, the dependence o f the observable rate


constant k ohs on S/V passes through the minimum that is indicative o f such a
combination.
In the case o f substances that have a tendency to chain or heterogeneous
decomposition, the method o f high-vacuum pyrolysis provides valuable
data on the products o f the primary monomolecular stage. Under the
conditions o f this method (pressure p < 1 0 ° Torr, contact time is less than
1 0 ° s), bimolecular reactions do not proceed. Nevertheless, even in this
case, it is necessary to carry out special studies to eliminate a catalytic role
of the wall.
The reactions o f decomposition proceeding through the bond breaking
are reversible ones. However, i f the reaction reversibility is neglected, this
leads to a very insignificant underestimation o f the rate constant o f the first
stage. The reaction o f disproportionation o f radicals is the main competitive
reaction for the recombination reaction. Thus, during decomposition o f
nitrocompound R N 0 2 we w i l l have:

1 3
RN0 2 ^—- R + N0 , R + N0 2 2 — - RO + NO,
2

where k Jk: 2
=
2, i.e., disproportionation proceeds faster than recombination
by a factor o f two. In addition, due to the low stationary concentration o f
radicals, their reactions with the molecules o f the medium and, under low
pressures, termination on the wall, may also be faster than recombination.
12 General Regularities of Thermal Decomposition

In the case o f small molecules (for which the number o f C , N , and O


atoms is less than 7), the monomolecular reactions fall into the pressure
nonequilibrium region at P 0 > 1 Torr. For instance, for nitromethane, the
high-pressure limit is attained at 400 Torr, for nitroethane at 80 Torr, but for
nitrobenzene, the transition pressure does not exceed 1 Torr.
In practice, it is sometimes difficult to take into account all the factors
that influence the rate. So, quite often, the data o f different authors and
interpretations o f the mechanism can differ considerably. For instance, for
the gas phase decomposition o f nitromethane, the following values o f the
activation energies are reported: 61 [2.1], 53.5 [2.2], 50 [2.3], and 36.5 [2.4]
kcal/mol and, correspondingly, the first stage o f decomposition is treated as
O atom detachment [2.1], C — N bond breaking [2.2, 2.4], and HN0 2

elimination [2.3]. Actually, the chain decay is realized with a small chain
length and the activation energy o f the first stage, which is as large as 55
kcal/mol, coincides with the dissociation energy o f the C — N bond.

2.2 Liquid Phase Reactions


O n transition from the gas phase to the liquid state, an essential variation o f
the decomposition rates does not occur only in the case o f monomolecular
homolytic reactions such as bond breaking or concerted reactions o f cycle
opening. The only factor that can exhibit itself in these cases is the cell
effect, which can enhance the recombination o f primary radicals. The
decrease o f the initial decomposition rate o f liquid nitroesters by a factor o f
4 - 5 in comparison with the gas phase is explained by this effect. In the case
o f C-nitrocompounds, as has already been mentioned, disproportionation
successfully competes with recombination. Nevertheless, even for this type
of substance, in the case of a pure monomolecular decomposition
(polynitroalkanes), the rate in the liquid state is lower than in vapor by 2 0 %
[2.5].
In the majority o f cases, on decomposition o f liquid explosives, the
non-specific and specific medium effects exhibit themselves. A l l reactions,
proceeding in the gas phase through the polar transition states, for instance
HN0 2 elimination from mononitroalkanes or H F from difluoroamino-
compounds, should be accelerated in polar solvents and in the melts o f the
compounds. For such quasiheterolytic gas reactions in the liquid phase, new
efficient decomposition channels are frequently opened, which is connected
with ionization. For mononitroalkanes, this is isomerization into nitronic
acids and for difluoroamines, ionization along the N—F bond.
Nitromethane in the liquid state at 180°C decomposes at the same rate as in
vapors at 4 4 0 ° C . The decomposition rate o f difluoroamines in the liquid
phase is higher by a factor o f 10 —10 than in the gas phase.
2 5

In the case o f decomposition o f aromatic nitrocompounds, the influence


o f the medium is great. In this case, the role o f the secondary chain
Thermal Decomposition and Combustion 13

processes increases due to the growth o f substance concentration in the


liquid phase.
M a n y aromatic nitrocompounds, nitramines, and nitroesters at room
temperature represent the associated liquids. Nevertheless, there is no
evidence that this factor may influence the decomposition rate. It is possible
that under elevated temperatures used in thermal decomposition studies,
self-association simply disappears. In order to influence the decomposition
rate through the formation o f complexes, the partner should possess
stronger donor properties than nitrocompounds themselves. For instance,
hexanitroethane, which is the strongest acceptor in the series o f aliphatic
nitrocompounds and has the highest sensitivity to the influence o f donor
compounds, exhibits an increase o f the decomposition rate under the action
o f toluene, naphtalene, and diphenylamine, i.e., the substance with such n-
or A7-donor properties which nitrocompounds themselves do not possess. For
weaker acceptors, for instance, 1,1,1-trinitroethane, the influence o f these
donors is not pronounced. The initial decomposition rate of sym-
trinitrotoluene increases as anthracene and aniline are added, but not
nitrobenzene or dinitrotoluene. Thus, self-association in a pure substance
does not influence the decomposition but chelation may strongly influence
compatibility with other compounds.
Autocatalytic processes play an important role in decomposition o f
virtually all explosives in the liquid state. Most often, they are o f an acid-
base nature. Thus, nitroesters are catalyzed by nitric acid, tetryl—picric acid,
and the decomposition of difluoroamines is catalyzed by HF.
Decomposition o f polynitroalkanes is accelerated by bases. Nitramines are
sensitive to the action o f acids and bases. Aromatic nitrocompounds
decompose under the action o f electron donors to form charge transfer
complexes with them. Nitric oxides, which represent efficient oxidizers, are
formed during decomposition o f explosives. Therefore, the presence o f such
groups as O H , N H , and C H O in the molecules o f explosives, which can be
2

easily oxidized, w i l l always accelerate decomposition.


The nature o f autocatalysis also predetermines the methods of
inhibition. Nitroesters are stabilized by weak bases o f diphenylamine type,
nitrocompounds - by acids. Cases when a substance is compatible with
neither acids nor bases present difficulties. For example, for this reason,
methods to stabilize the secondary nitramines and difluoroaminocompounds
have not yet been found.
The liquid phase decomposition o f explosives may rarely be completely
described by the equation o f autocatalysis o f the first order. One o f the
reasons is the manifestation o f the burning-out effect (the decrease o f the
liquid phase volume) at autocatalysis by condensed products.
The equation o f autocatalitic reaction with account teken o f the
burning-out effect for the case o f catalysis by condensed products takes the
following form [2.6]:
14 General Regularities of Thermal Decomposition

^ = A ( l - 7 ) + A: C (l-77),
1 7 2 jk
(2.2.1)
dt

where 77 is the reaction degree, C = a(\- k ju)/(\-/jrf) is the concentration o f


catalyst, (\-ju)rj is the amount o f catalyst, a is its fraction in a condensed
residue, jli = ( K i n i t - K )/K
fin init . Integration o f (2.2.1) gives:

/7o] * <* 0 - / 0 ' =


2

(2.2.2)
1
In 1+—^—^-77 - ( l - / i ) l n ( l - 7 7 ) ,
1 - n

where % = k] /k a (\- ju).


2

Fig. 2.2.2

Fig. 2.2.1. Dependence of the reaction rate w drj/dt on the degree of


transformation according to eq. (2.2.1) at rj - 0.1. 0

Fig. 2.2.2. Dependence of the specific reaction rate w speciflc = [\/(\—n)]drj/dt on the
degree of transformation: (/) the first order reaction, (2) the first order autocatalysis,
and (3) calculation by eq. (2.2.1).

The dependences o f the total (w) and specific ( w speci f ) reaction rates vs. the
lc

degree o f transformation calculated by eq. (2.2.1) are presented in Figs.


2.2.1 and 2.2.2. The values o f the maximal rate and the degree o f
transformation n , when this maximal rate is attained, strongly depend on
max

the value o f /u. A t large //, 7 / max is close to unity, i.e., the rate maximum is
located at the end o f decomposition.
Thermal Decomposition and Combustion 15

Fig. 2.2.3. Dependence of the specific reaction rate on the degree of transformation
during thermal decomposition of R D X . The points are the experimental data [2.7J.
The solid curve is the result of calculation by equation (2.2.1) for the following
parameters: k = 4 • 10 min '. k a(\ - / / ) = 58.1 10 ' min" ,// = 0.90, 7'= 238°C.
x
3
2
1

The experimental data obtained during decomposition o f a number o f


substances in the liquid phase are in good agreement with eq. (2.2.1). The
dependence o f the specific rate on the degree o f transformation during
thermal decomposition o f R D X at 2 3 8 ° C is presented in F i g . 2.2.3
(according to the data [2.7]). Coincidence o f the calculated curve with the
experimental points demonstrates the validity o f eq. (2.2.1) for this reaction.
A t the high rate o f thermal decomposition in the condensed state, self-
heating and heat explosion o f the system are possible. Under isothermal
conditions, the experiment is always restricted by the substance mass. The
critical conditions o f heat explosion may be estimated by the classical
formula o f Frank-Kamenetskii for a sphere:

^critical Ae = 3.32. (2.2.3)


RT JA {

For an R D X - t y p e substance with the following parameters, Q = 600 cal/g, p


= 1.9 g/cm , A = 10
3 3
cal/(cm • s • K ) , A = 10 15
s ', E = 42 kcal/mol and the
rate constant value o f 1 0 ° s 1
( 2 3 7 ° C ) , a particle with the mass o f 175 mg
w i l l ignite. Under static conditions, one can measure the rate constant with
the value no larger than 10~ s 4 1
( 2 1 0 ° C ) . In this connection, one should use
only small samples with a mass o f 1 g without the danger o f explosion.
16 General Regularities of Thermal Decomposition

Nevertheless, even in this case, further reduction o f the sample mass is


required in order to avoid heating.

2.3 Decomposition in the Solid State

A s a rule, the crystal lattice strongly inhibits the reaction, and, thus, the
decomposition in the solid state proceeds noticeably more slowly than in the
liquid and gas phases. Due to this difference in the rates, several effects
appear, which disturb and mask the solid phase process. Decomposition
through the gas phase, the sub-melting o f a substance on impurities and
reaction products, the effects o f pre-melting, and phase transitions are
among them.
Decomposition through the gas phase gives a sufficient contribution to
the reaction rate at large ratios o f the rate constants k /k \ g?i so ld and in
experiments with a small m/V ratio. One can recognize the gas phase
reaction and separate the rates w g a s and w | S0 ld from the dependence o f the
observable rate constant £ bserv 0
o n
V/m, which takes the form:

*observ.=W - + W WS SQM , (2.3.1)


" m

where J f gas =£ g a s • P^ox ' MV/RT. When expression (2.3.1) is used for a
quantitative determination o f W &as and lV sM it is necessary to ensure that
the equilibrium pressure o f the substance vapors in the course o f the
reaction is settled over the whole volume o f the vessel. The most reliable
method to eliminate the gas reaction is the decomposition at large m/V
ratios chosen so that the substance fraction in vapors would not exceed
0.01% o f the sample mass. This value results from the empirical
relationship /% /£ oiid < 10 forany arbitrary reactions. If P
as S
4
vapor = 1 Torr, the
required value is m/V = 0.01 g / c m \
V e r y often, decomposition o f solid compounds proceeds through the
liquid phase, which is formed as a result o f the melting o f impurities,
decomposition products, or their eutectics with the substance. One can
reveal and take into account the liquid phase reaction by mathematical
processing o f the kinetic curves [2.8].
If in the solid and liquid phases the reaction proceeds according to the
equation o f the first order, then, for reaction B —» A the dependence o f the
rate on the degree o f transformation takes the form:

^ = A,+A 7 , /II 7 (2.3.2)


Thermal Decomposition and Combustion 17

where k = a(\ - ju)(k - k\) - k\\ rj = (B - B)/B ; k\ and k are the rate
m 2 0 0 2

constants in the solid and liquid phases, correspondingly; a = C \ /C = B u] Aliq

Z?ii //4|i
q qis the value constructed under the constant temperature and
independent o f the degree o f transformation up to the moment o f
disappearance o f the solid phase: /u = (V - K )/K is the volume uut flIV init

variation at complete decomposition.


At a( \ - ju)(k - k\) > k\ and k > k the total reaction rate w i l l increase
2 2 h

proportionally to the degree o f transformation up to the moment o f


complete dissolution o f the parent substance in the reaction products. After
this, the reaction w i l l proceed according to the ordinary first-order equation

drj
It

The solid phase disappears at the degree o f transformation rj which is equal


to

B =0
5oM = B (\-r )-B a
0 I
t
0 i\ rj =
a (!-//) + !

At this instant, k\ + k rj = k (l - n) .
m 2

Upon integrating eq. (2.3.2), we obtain:

kj = In (2.3.3)
V , v
i J

The kinetic curves calculated by eq. (2.3.3) have a pronounced


S-shape for the case

a(\-p)(k -k )>k .
x 2 x

The dependence o f the rate on the degree o f transformation has a


distinctive form (2.3.1), calculated by eq. (2.3.2) and the classical equation
of the first order reaction. In this case, the maximal rate is approached at the
point o f a complete dissolution at n = 7].
In the cases when condensed decomposition products can serve as
catalysts, the expression for the total reaction rate has to be written in the
following form:
18 General Regularities of Thermal Decomposition

dB_ \__dA
~dt • n dt

= \ so\
k B
+ 2 \\q
k B
+ k
3 \\q A ^>
B C
v

where C i iA q is the concentration o f the catalyst in the liquid phase; C A H q

aA/V ; a is the catalyst fraction in the condensed reaction products:


V]q

a
^I.q = B Viq + A = A(\ + a); C A

1+a

Thus, the reaction rate, up to the moment o f complete dissolution in the case
o f catalysis by the final condensed products, w i l l be determined by the
following expression:

dr]
~dt 1+a
(2.3.4)

+a

where k is the value, which depends only on temperature and remains


m

constant during the decomposition process at t = const. Consequently, in


the case o f catalysis by the final condensed products, we also obtain the
expression, which is similar to eq. (2.3.2) for the reactions proceeding
according to the first order law:

(2.3.4a)
dt

Equation (2.3.4a) differs from (2.3.2) by the value o f k^. After complete
dissolution, which also comes at

a{\-ju) + l

the decomposition process is governed by the following equation

(2.3.5)
at 1-/^7
Thermal Decomposition and Combustion 19

obtained for the case o f autocatalytic decomposition o f liquid explosives


under conditions o f alternating volume o f the condensed phase.
The dependence o f the reaction rate on the degree o f transformation
calculated by eqs. (2.3.4) and (2.3.5) is presented in Figs. 2.3.1 and 2.3.2.
One can see from these figures that the maximal reaction rate may be at the
point o f complete dissolution o f the parent substance or later, when the
reaction occurs in the absence o f the solid phase.

Fig. 2.3.1 Fig. 2.3.2

Fig. 2.3.1. Dependence of the reaction rate on the degree of transformation


calculated by eq. (2.3.2) at k\ = 1, k = 100, ^ = 0.9, and a is equal to the following
2

value: (/) 20, (2) 10, and (3) 5. The straight line (4) corresponds to the equation of
the first order reaction.

Fig. 2.3.2. Dependence of the reaction rate on the degree of transformation


calculated by eq. (2.3.4) at k = 1, k = 20. k^a =100, and ju = 0.9, and a is equal to
] 2

the following value: (/) 20. (2) 5. (3) 2, and {4) 1. The curve (5) corresponds to eq.
(2.3.5) of the reaction in the liquid phase.

The influence o f impurities, even chemically inactive ones, which may


increase the decomposition rate o f explosives, transferring a certain part o f
it into the liquid phase (in the case when the eutectic temperature o f the
system "explosive + inert additive" is lower than the experimental
temperature), is o f considerable interest. In this case, the decomposition rate
w i l l be equal to the sum o f the rates in the solid and liquid phases (in the
solid and liquid phases the monomolecular reactions proceed):

W =
\ B^ so\
k C v
^ C ^ / H q = Msol +^2 iiq-5

If T is the amount o f an inert additive, than we will have


20 General Regularities of Thermal Decomposition

v
hq = A
+ Uq+ >
B T
A= B (\-ju)rj,
0

B
\\q = B (\-p)anQ + Ta, £sol = B - B n-
0 Q B (\
0 -p)an-Ta,

where a = B /(A Viq + 7) (the solubility o f the substance B in T and in the


mixtures A + T is constant).
The reaction rate as the function o f the degree o f transformation may be
written in the following way:

^ = * i + (* " 2) ~
2
k
+ [0 - / ' ) a(k 2 - *,) - k ]77.
x (2.3.6)

Fig. 2.3.3. Dependence of the reaction rate on the degree of transformation


calculated by eq. (2.3.4) at A, = 1. k = 100. // 2 0.9. and a= 10. The ratio T/B for 0

the straight lines (I)-(4) is equal to 0, 0.01. 0.025. and 0.050, correspondingly. Line
(5) corresponds to the equation of the first order reaction.

Thus, in this case the decomposition rate is also directly proportional to the
degree o f transformation up to the moment o f complete dissolution, after
which it decreases according to the first order law, and the initial rate is
directly proportional to the amount o f additive or impurity and to the
dissolubility o f the parent substance in them (Fig. 2.3.3).
In those cases when solubility o f the explosive is different in the inert
additive and in the reaction products, one can easily obtain the
corresponding dependence o f solubility on the system composition and use
it to derive eq. (2.3.6).
A t the point o f a complete dissolution o f the parent substance in the
reaction products, the solid phase disappears and the reaction law is
Thermal Decomposition and Combustion 21

changed. It is known that for ideal systems the following relationship is


obeyed:

X(T T)
lruV
X
melt (2.3.7)
RTT,
melt

where N is the molar fraction o f a dissolved substance, X is the melting heat,


is the melting temperature, T is the experimental temperature.
Consequently, the degree o f transformation, at which complete dissolution
takes place, decreases with the temperature increase according to the
following law:

1 - exp
^(Tmelt ~ T)
N RTT
me\i
•juN f
- JLI exp
RTT melt
X
J

It should be noted that in real systems more complicated situations can


be observed, which are connected with a complex mechanism o f the
chemical reaction o f decomposition. However, in all cases it is necessary to
consider the possibility o f a progressive dissolution o f the parent explosive
in the reaction products, which is accompanied by a corresponding variation
of the kinetic laws o f the reaction.
In the presence o f impurities, the observable rate constant o f the initial
stage o f decomposition is determined by the following equation [2.9]:

P 0 A liq f AE (AQ
* |(1 + exp exp — ^ - 1 (2.3.8)
#0 ^sol v I RT
S0
R T

where B and P are the initial amounts o f the substance and impurity, AE =
0 0

£soi ~ £iiq> A and E are the Arrhenius equation parameters, AQ = A(T md{ -
TyTmeit- The sub-melting effect distorts the Arrhenius dependence o f the
rate constant in the vicinity o f the melting point due to the temperature
dependence o f solubility, which has the form o f relation (2.3.7). A n
accurate purification o f the substance from impurities leads to the reduction
o f the temperature range where the deviation from the Arrhenius law is
observed (Fig. 2.3.4).
Besides melting o f impurities and reaction products, in the vicinity o f
r i ,
m e t the effect o f pre-melting may become apparent, which is not
connected with the presence o f impurities. For this effect, Ubbelohde
introduced the new term "homophase pre-melting" in contrast to "two-
phase pre-melting" which is caused by impurities. The vibrational pre-
22 General Regularities of Thermal Decomposition

melting, fluctuation homophase pre-melting, and homophase pre-melting


caused by formation o f defects are known. A l l these types o f pre-melting
may influence the rate o f chemical reaction because it is highly sensitive to
the mobility o f molecules. The temperature range near r m e h , in which such
influence may exhibit itself, apparently does not exceed several degrees.
The impurity effects usually mask the homophase pre-melting. However,
the case is known (the decomposition of carefully purified
hexanitrodiphenylsulphide with the r i = 2 3 4 ° C [2.9]) when the Arrhenius
m e t

dependence for A | S0 id is obeyed up to the temperature T = T meh - 1°C.

Fig. 2.3.4 Fig. 2.3.5

Fig. 2.3.4. Dependence log k-\/T for the reaction of PETN decomposition in the
liquid phase (/), in the solid phase (2) (the sample was purified by recrystallization
according to the data [2.10]), and in the solid phase {3) (the sample was purified by
sublimation [2.11]).

Fig. 2.3.5. The influence of phase transition on H M X decomposition: (/) £-phase,


(2) a-phase, and (3) /?-phase.

The phase transition leads to a step-wise change o f the decomposition rate.


In the high-temperature phase, the reaction proceeds at a higher rate than in
the low-temperature phase, but slower than in the liquid state. Near the
temperature o f the phase transition, one can observe the same anomalies o f
temperature dependence o f the rate constants as those observed in the
vicinity o f the melting temperature (Fig. 2.3.5) [2.13].
Molecular crystals o f organic substances are distinguished from ionic
ones by a smaller energy o f the crystal lattice and, in principle,
decomposition can occur in them not only on the surface and imperfections,
but also in the whole crystal volume. There are several models, which can
Thermal Decomposition and Combustion 23

describe an irreversible monomolecular reaction in the ideal crystal lattice


[2.13-2.16]. In these models, the volumetric characteristics o f the reaction
and crystal are used and they are based on the assumption that in order for
the reaction to proceed it is necessary to create near the reacting site o f the
molecule a certain cavity (which has the meaning o f the effective activation
volume in the solid phase AV ) S
U
where a transition state o f the reaction is
formed, which is free from the forces o f intermolecular interaction.
If one considers a crystal as a continuous elastic medium, then the
volume variation is equal to the derivative o f the free energy with respect to
the pressure {c{AG)/dP)j = AV$, the pressure can be estimated from the
value o f the compressibility coefficient j3 as P = \/j3 - AV/V, and variation
of the free energy o f formation o f the activated complex w i l l be expressed
by the following relation:

AG;=AG 0 + (AV:) /2fJV ,


2
0 (2.3.9)

where V is the molar volume o f the substance.


0

Thus, the variation o f the activation energy AE = £ i - £ so g a s is

A£ = (AK S
#
) 2
lip V ,
0 (2.3.10)

and the inhibitory effect o f the lattice may be expressed as the ratio
^gas^sol = e x p ( A £ / / ? r ) . (2.3.11)

The activation volume in the solid phase, neither by its shape nor by its
value, is equal to the actual activation volume o f the reaction, which is
measured in the solutions. A s follows from [2.15], the value o f AV s
n
for the
reactions o f bond breaking as a first approximation can be determined as the
increment o f a cell volume occupied by the molecule during the expansion
of its walls for the value o f elongation o f the breaking bond, i.e., 10-15%
from the equilibrium bond length. In practice, for all bonds this value is
close to 0.2 A . The expansion o f the cell by 0.2 A w i l l result in the release
of the molecule from the forces o f intermolecular adherence and makes
possible the reactions connected with the rotation o f the molecule as a
whole or with the rotation o f its parts, or with the formation o f cyclic
transition states (rearrangement, isomerization, and elimination). Thus, for
reactions o f various types, a common approach is possible to estimate the
AV$ value based on the V magnitude.
0

However, a difficulty appears when the AV S


U
value is determined for
large and, especially, long-chain molecules. In these cases, one can suppose
that the reaction can occur i f not the whole molecule but only the fragment
24 General Regularities of Thermal Decomposition

o f it containing the reaction site acquires mobility. The uncertainty in the


AV S
U
value estimate gives no way o f using formula (2.3.10) in the general
case to determine the lattice inhibitory effect, though the estimate
calculations are possible and easily accomplished. The fact that this formula
predicts the dependence o f the lattice inhibitory effect on the elastic
properties o f the crystal, and, consequently, on other characteristics (i.e.,
r i , the sublimation enthalpy) connected with the crystal lattice energy is
m e t

also an advantage o f this formula.


The known values o f the lattice inhibitory effect for 18 compounds are
presented in Table 2.3.1. One can see a regular increase o f the lattice
inhibitory effect as the crystal volumetric compressibility decreases and the
melting temperature o f the substance increases. In this connection, the
tendency predicted by a theoretical model appears: at the same T , larger meh

molecules have a smaller lattice inhibitory effect than compact molecules,


which have a small mass.

Table 2.3.1. Inhibitory effect of the crystal lattice and the physical properties of a
compound [2.13]

Compound *,/*s* W ° C /MO , 1 0


J' , 0

Pa" 1
cnr'mor 1

[FC(N0 )CH ] NN0 2 2 2 2 4 86 3.53 173.9


[F NC(N0 ) CH 1 NN0
2 2 2 2 2 2 4 102 - 198
C(N0 ) C(N0 ) 2 3 2 3 6 140 5.5 169.4
PhN=N-NHPh 6 101 - -
[C(N0 ) CH ] NN0 2 3 2 2 2 10 95.5 1.05 199
l,3,5-Trinitro-2,4,6- 28 131 - -
triazidobenzene
N (CH NN0 ) CH N3
3 2 2 4 2
74 177 - -
HOOCCH COOH 2 90 135 - -
[(N0 ) CCH N(N0 )CH ]
2 3 2 2 2 2 92 180 1.06 254.5
2,4,6-(N0 ) C H N(N0 )CH 2 3 6 2 2 3 100 130 1.2 164.5
HOOC-COOH 122 189 - -
1,4-Dinitro-1,4-diazacyclohexane 210 213 1.31 107.3
[CH C(N0 ) CH ] NN0
3 2 2 2 2 2 230 177 0.968 189.5
C ( C H O N 0 ) (PETN) 2 2 4
360 142 - 178.5
l,l,3,5,5,7-Hexanitro-3,7- 500 250 0.925 203.2
diazacyclooctane
1,4,6,9-Tetranitro-1,4,6,9- 1000 236 - -
tetraazadecalin
l,3,5-Trinitro-l,3,5- 1400 201 0.809 122
triazacyclohexane (RDX)
1,3,5,7-Tetranitro-1,3,5,7- 7800 277 0.547 155.8
tetraaza-cyclooctane ( H M X )

At the temperature T= T meh - 20°C.


Thermal Decomposition and Combustion 25

In the case o f large experimental values o f the lattice inhibitory effect, one
may determine from the value o f AE = E s - E\ whether the reaction
proceeds in the bulk or on nonequilibrium crystal defects. In the first case,
AE > 0; in the second, AE = 0. If k\ k < 100, an expected difference AE for
s

the actual lattice inhibitory effect is small, and it is difficult to conclude,


from the comparison o f the £ i and £ so g a s values, where the reaction is
localized. A s follows from the experimental measurements, in all processes
in which the E and E\ values have been accurately measured, they are equal
s

to (2-4 kcal/mol) within the limits o f experimental errors. Within the


framework o f the models considered, an unambiguous conclusion follows
from this fact: for the cases when k s > 100, the reaction on the
nonequilibrium imperfections is more probable than decomposition in the
bulk (otherwise, AE would have a noticeably larger value). The maximal
experimentally observed value o f the lattice inhibitory effect (10 ) is limited 4

by the reaction surface exposure. L o w values o f the lattice inhibitory effect


indicate that the reaction occurs in the bulk o f the lattice. It is difficult to
perceive that 15-20% of substance is localized at nonequilibrium
imperfections o f a carefully cleaned crystal. A t least, the decomposition o f
all ductile crystals and substances with the melting temperature r m e i t <
100°C proceeds as a homogeneous bulk process with a slight inhibition in
comparison with the gas or liquid phases.
A m o n g the mechanisms considered, there are some other mechanisms
in the solid phase which may lead to an increase o f the rate o f chemical
reactions.
In the case o f particular structures, the existence o f the ordered lattice
can facilitate "collective" interactions, creating "intermediate products".
The recombination process o f ionic radicals N H 3
+
formed in the N H C 1 0 4 4

matrix under radiation treatment [2.17] can be considered as an example o f


such interactions. Recombination proceeds according to the second order
and its rate is noticeably higher and the activation energy lower than for
diffusion o f ions o f the same size in this lattice. The process occurs due to
the transfer o f a free valence by the jump o f hydrogen atoms along the chain
of N H 4
+
ions. Introduction o f impurity K ions, which may replace some o f
+

the N H 4
f
ions and, thus, interrupt the chain o f transformations, reduces the
rate constant by a factor o f 25 due to the increase o f the activation energy.
The influence o f imperfections o f the crystal lattice o f different kinds
on the rate o f chemical reactions in the solid phase has been repeatedly
studied and discussed. It was clearly demonstrated that the occurrence o f
lattice imperfections results in the increase o f the process rate.
Much attention was given to an extremely important type of
imperfections: dislocations, whose occurrence determines a considerable
part o f the mechanical and physical properties o f a crystal. The energy o f
formation o f dislocations, whose value in the ionic lattice can approach 20
kcal/mol, made it possible to suppose that chemical reactions on the
26 General Regularities of Thermal Decomposition

dislocations should also proceed at noticeably larger rates than in the ideal
part o f the crystal.
It was shown experimentally by the example o f the reaction o f
ammonium perchlorate thermal decomposition that the reaction occurs
predominantly on the dislocations and starts in those places o f the crystal
where the dislocations have already been presented before the beginning o f
this process [2.18, 2.19]. This reaction is considered in detail in Chapter 13.
A similar picture was obtained in the microscopic study o f ammonium
bichromate decomposition. In this case, the formation o f solid products o f
decomposition proceeds on the dislocations, exhibiting the network o f
interacting dislocations in the reacting crystal.
In those cases when a nonvolatile product is formed, the tendency o f
imperfections to aggregation accompanied by the energy gain results in the
mechanism w h i c h for thermal decomposition o f sodium azide can be
presented in the following form:

Na +N 3 Na + N 3

Na + Na + N
2 3
N a -f N 3 3

2N 3 3N 2

Na + Na + N
n 3 Na n + 1 +N 3

A distinctive addition accompanied by bond breaking (in this case,


accompanied by electron transfer) proceeds, and each next act is more
advantageous because, as in the case o f condensation, the energy gain o f
metal sodium formation is finally attained. This increase o f the rate o f
elementary processes occurs as long as particles o f metal sodium grow to a
noticeable size, for which the contribution o f surface energy to the energy
o f particle formation becomes negligibly small. After this, the reaction rate
per unit o f interface area becomes constant.
Variation o f the sodium concentration, formed in the course o f sodium
azide decomposition, determined by the electron paramagnetic resonance
( E P R ) method for a low degree o f transformation (1(T % ) , is shown in F i g .
3

2.3.6. It is important that such a significant self-acceleration at such small


degrees o f transformation proceeds due to an abrupt decrease o f the
activation energy from 50 kcal/mol (for initiation o f the process) up to 30
kcal/mol (for the stage o f reaction development) at the stage o f metal
sodium precipitation and interface appearance.
The mechanism o f reaction development at high degrees o f
transformation, when precipitation o f the reaction product into the separate
phase has already happened, is determined to a large extent exactly by this
fact.
Thermal Decomposition and Combustion 27

no

50 75 fOO 125
t,min

Fig. 2.3.6. The time dependence of the concentration of sodium atoms during
thermal decomposition of sodium azide (550°K): (/) £ a c l l = 50 kcal/mol, (2) £ a c t 2 =
28 kcal/mol.

The simplest ideal case can easily be analyzed by the example o f a


reversible reaction o f the first order:

A B A B ^ A A + B G A S ,

A B B + A ^ = ^ (A + l ) + B B - the first type


A A A A

BAB ^ B G A S ,

ABAB ^ AAB + B B, A the second type

A B —— A A ; B A B — —
A B G A S -

Subscripts A , A B , and "gas" correspond to components A , B , and A B in the


corresponding phase.
The reaction o f the first type is the reaction o f addition accompanied by
a simultaneous breaking o f the A B bond. The reaction o f the second type is
a preliminary decomposition o f the A B substance in the parent lattice with
the following precipitation o f the solid component A in the separate phase.
It is natural that the reaction o f the first type occurs easily, with smaller
activation energy, and the process is localized on the "parent substance-
reaction products" interface.
In this case, the velocity o f the interface displacement can be described
by the modified P o l y a n y i - W i g n e r equation:
28 General Regularities of Thermal Decomposition

a = d- [k -exp(-E'/RT)-k -Qxp(-E"/RT)
0 0 • C g a s ] =

= d-k -exp(-E'/RT)
0 (\-kP),

where C g a s = aP, a = C exp(X a&s / RT),

^0 Qlgas
f
E" -E' + A^
exp = K is the equilibrium constant.
RT
*0

M o s t likely, a great number o f reactions proceed according to mechanisms


o f this type. A s a result, the solid product is formed. Sodium azide
decomposition as an example o f this type has already been mentioned
above.
However, the possibility exists that the reaction proceeds in the bulk
(the second type) [2.20, 2.21].
In this case, the velocity o f the interface displacement will be
determined by the rates o f the direct and reverse reactions, by the diffusion
rate towards the interface o f the decomposition products, and by the
boundary conditions (the equilibrium solubility o f the product in the parent
lattice):

D ^ - + u— + k - k l n C C =0 9

dx" dx

dC_
x = 0, C = C , D 0

dx
x = oo, C = C x = £, / k C , 2 S0] C s o l = aP,

where D is the diffusion coefficient, C is the concentration o f substance A , x


is the distance from the interface "parent substance ( A B ) - r e a c t i o n product
( A ) " , u is the velocity o f interface displacement, k\ is the rate constant o f
the direct decomposition reaction, k 2 is the rate constant o f the reverse
reaction, C g a s is the concentration o f the dissolved substance B , and C i , is s o d

the concentration o f substance A in the solid phase A .


The solution o f this equation gives the value o f the velocity and the
distribution o f concentrations, as well as the characteristic width o f a
chemical reaction zone:

Cm —Cn
C sol • c , OJ

at C « Cc i, we w i l l have
0
Thermal Decomposition and Combustion 29

Cqq-C 0 JM> = H )
C , s o y C \k aP'
sol 2

c = Coo - (^oo - Q ) ) P ( ~ * / e x

where ^ = £ C a / £
2 0 p L^^DCJk[\

Thus, the reaction occurs within a more or less narrow zone in the
vicinity o f the interface. The overall process can be developed irreversibly
up to the end due to the precipitation o f the product into the separate phase.
The reversibility may be observed only for intermediate products. In
accordance with the analysis, in contrast to the surface reaction, the reaction
in the bulk within a moving zone should be characterized by the noticeably
overestimated pre-exponential factor in the observable regularities and
larger activation energies. This is connected with the fact that, according to
the zone mechanism, the overall rate involves a larger amount o f substance,
in spite o f the smaller values o f the rate constants and larger activation
energies.
Sometimes, the experimental data on the kinetics o f decomposition
reactions can only be interpreted based on the concepts presented above.
Thus, for instance, in complete agreement with theory, abnormally large
values o f the pre-exponential factor are observed upon dehydration o f
crystal hydrates.
A c c o r d i n g to [2.22], for dehydration o f C a C 0 • 6 H 0 , we w i l l have k 3 2 0

= 1 0 at E = 38-40 kcal/mol, and dehydration o f K S 0 • C r ( S 0 ) •


34
2 4 2 4 3

2 4 H 0 proceeds with the following parameters: k = 1.2 10 , E = 31.2


2 0
25

kcal/mol [2.23], k = (3-5)-10 , and E = 30 kcal/mol [2.24]. Similar


0
28

regularities are also observed during decomposition o f solid solutions


[2.25].
The criterion o f transition from the surface mechanism to the zone
mechanism can easily be obtained. In the simplest case, we w i l l have:

(i) u' = dk -exp(-E'/RT);


0

(2) », = [ l - ( C / C ] A : o i e x p ( - £ / / ? r ) L ;
0 0 0 1

( E -E ~A\
D 2

~ 2RT ,
30 General Regularities of Thermal Decomposition

A s a rule, £ j > E, £D > E + A.


2

Usually, the relationship between the activation energies o f direct and


reverse reactions and diffusion is such that the rate o f the zone reaction
increases faster with the increase o f the temperature and zone width than the
surface reaction rate. Consequently, one may expect that in the range o f
high temperatures, the zone mechanism will occur more frequently.
Apparently, the reaction development in the solid phase in the form o f
propagation o f a more or less narrow reaction zone is a rather common
phenomenon and it can be realized in the most unexpected situations.
Precipitation o f the reaction product in the separate phase leads to one
more extremely important circumstance: the appearance o f the stressed-
deformed mode. In most cases, the action o f artificially created stresses was
studied. However, such stresses practically always appear by themselves
during the reaction proceeding in the solid phase due to the difference in the
densities o f the parent substances and the reaction products.
A peculiar mechanism o f the zone reaction development under the
action o f stresses, which appeared due to the product precipitation, was
found and studied in collaboration with Raevskii [2.18, 2.19, 2.26, 2.27] in
a large series o f works on the decomposition o f ammonium perchlorate. A s
already mentioned, the reaction proceeds predominantly on dislocations.
The products evolved generate stresses and deformations around the
growing site that, in turn, cause the motion o f dislocations along the slip
planes and, due to their interaction with each other and other defects, result
in the dislocation multiplication that leads to an increase o f the reaction rate.
One can observe a peculiar chain reaction.

i i I I 1
6J 89 SS X-tf'^jum

Fig. 2.3.7. Variation of the number density of nuclei (n) in the reaction development
zone of ammonium perchlorate decomposition. The displacement of the reaction
zone (the growth of the centers of decomposition) along the direction (010) is
plotted along the X-axis: (7) 196, (2) 212, and (3) 225 C.
Thermal Decomposition and Combustion 31

A n approximate description o f such a process can be carried out on


consideration o f simultaneously proceeding reactions o f generation o f
dislocations and chemical transformation by itself.
In the stationary case (a plane front), we w i l l have:

- n(dnldx) = k e x p ( - cr \
{ u iim /a)(n + n^-k^n',
- u(drj/dx) = k n\ 2

at x = oo; n - 0; 77 = 0;
*
at x - 0; 77 = 77 .

Since n/rj « 1, at cr \ /au {im = ax, we w i l l have:

for k^k ,2 u = (kja){njrf) [ 1+ (*,/2£ ) + (£, /3£ )] , 2


2
2

for^>^ , 2 u = (k /a) y [l/ln(krf / k n j]


2 0 ,

where w is the velocity o f the reaction front propagation, n is the


concentration o f dislocations, k\ is the rate constant o f generation o f
dislocations, cris the stress in the crystal, cr i u tim is the ultimate stress value at
which the generation o f dislocations proceeds, k 2 is the rate constant o f
chemical reaction on dislocations, and n is the concentration o f dislocations
0

at infinity (for simplicity, this value was assumed to be constant; actually, it


is necessary to consider the propagation o f dislocations from the reaction
zone).
It should be noted that the reaction rate is determined to a large extent
by the parameters responsible for generation o f dislocations, and the actual
rate constant value slightly influences the rate o f chemical transformation.
The same is true for the temperature dependence o f the process rate, which
in extreme cases follows the Arrhenius law but with the parameters, in
which the values prevail, determining the temperature dependence o f the
rate o f generation o f dislocations.
It is typical that the pre-exponential factor in the expression for the
dependence o f the velocity o f the reaction front propagation on the
temperature turns out to be noticeably smaller than the normal one, in
complete agreement with the experiment, in which the value k = 1 0 0
8
s' 1

was observed (Chapter 13).


The concentration o f dislocations and the value o f the degree o f
transformation depend exponentially on the distance to the reaction front.
A t k\ < k , n
2 {x) = (kx/k^riQe ° ; rj = ijez
(xJ
ax
\ % = x a/cr .uhm

A s follows from these relationships, the width o f the chemical reaction


zone increases as the temperature increases, but is noticeably weaker than
the velocity o f the reaction front propagation (Fig. 2.3.7). It was
experimentally shown that the rate of ammonium perchlorate
32 General Regularities of Thermal Decomposition

decomposition is determined by the activation energy o f 30 kcal/mol. In


accordance with theory, the width o f the chemical reaction zone increases
with the activation energy o f 10 kcal/mol.

REFERENCES

2.1 H . Taylor, V . Vesselovsky, J. Phys. Chem., 39: 1095-1101 (1935).


2.2 T . Cottrell, T . Reid, T . Graham, Trans. Faraday Soc, 47: 584
(1951).
2.3. J. Hillenbrandt, M . L . Kilpatrik, J. Chem. Phys., 19: 381 (1951).
2.4. C . Frejacquens, C. R. Acad. Sci., 231: 1061-1062 (1950).
2.5. G . M . N a z i n , G . B . Manelis, F . I. Dubovitskii, Izv. Akad. Nauk SSSR,
Ser. Khim., 11:2628 (1968).
2.6. F. I. Dubovitskii, G . B . Manelis, A . G . Merzhanov, Dokl Akad. Nauk
SSSR, 121: 668-670 (1958).
2.7. A . J. B . Robertson, Trans. Faraday, 44: 81-90 (1948).
2.8. G . B . Manelis, F . I. Dubovitskii, Dokl. Akad. Nauk SSSR, 126: 8 1 3 -
816 (1959).
2.9. Y u . Y a . M a k s i m o v , Zh. Fiz. Khim., 41: 1193-1198 (1967).
2.10. K . K . Andreev, B . N . K a i d y m o v , Zh. Fiz. Khim., 25: 2776-2787
(1961).
2.11. Y u . M . Burov, G . M . N a z i n , VII All-Union Conf. on Kinetics and
Reaction Mechanism in Solids, Chernogolovka: Joint Institute o f
Chemical Physics, A c a d e m y o f Sciences o f the U S S R , p. 43 (1978).
2.12. A . R . Ubbelohde, Melting and Crystal Structure, M . : M i r , 419 p.
(1969).
2.13. Y u . M . Burov, G . M . N a z i n , G . B . Manelis, Russian Chem. Bull.,
48(7): 1250-1254(1985).
2.14. G . B . Manelis, In: The Problems o f Kinetics o f Elementary Chemical
Reactions, M . : N a u k a , 93-106 (1973).
2.15. L . D . Zusman, A . B . Gel'man, Zh. Strut. Khim., 21(1): 7 2 - 7 6
(1980).
2.16. S. B o n , Chemistry of Solid State, V . M . Garner E d . , M . : Izd. Inostr.
Lit., 335-353 (1961).
2.17. F . I. Dubovitskii, N . Y a . Buben, G . B . Manelis, Radiospectroscopy of
Solids: Proc. of the Meeting (Krasnoyarsk, 1964), S. A . AFtshuler
E d , M . : Atomizdat, 450 p. (1967).
2.18. A . V . Raevskii, G . B . Manelis, Dokl. Akad. Nauk SSSR, 151: 8 8 6 -
889 (1963).
2.19. A . V . Raevskii, G . B . Manelis, V . V . Boldyrev, L . V . Votinova,
Dokl. Akad. Nauk SSSR, 160: 1136-1137 (1965).
2.20. G . B . Manelis, V . A . Strunin, Dokl. Akad. Nauk SSSR, 187: 362-363
(1969).
2.21. G . B . Manelis, V . A . Strunin, Zh. Fiz. Khim., 46: 1987-1990 (1972).
Thermal Decomposition and Combustion 33

2.22. B . Topley, J. Hume, Proc. Roy. Soc, 120: 211-221 (1928).


2.23. J. A . Cooper, W . E . Garner, Proc. Roy. Soc, 174: 487-503 (1940).
2.24. M . M . T. Anous, R. S. Bradley, J. C o l v i n , J. Chem. Soc, 3348-3354
(1951).
2.25. B. Y a . Lyubov, Kinetic Theory of Phase Transitions, M.:
Metallurgiya, 263 p. (1969).
2.26. G . B . Manelis, Y u . I. Rubtsov, A . V . Raevskii, Fiz. Goreniya i
Vzryva,6: 3-11 (1970).
2.27. A . V . Raevskii, G . B . Manelis, Proc III All-Union Symp. on
Combustion and Explosion, M . : Nauka, 748-751 (1972).
Chapter 3

ALIPHATIC NITROCOMPOUNDS

During decomposition o f aliphatic nitrocompounds in the gas phase the


following two mechanisms mainly occur: the molecular mechanism, which
consists in H N 0 2 elimination through the 5-membered transition state,

*i —Ok
CH C NO?
O

(3.1)
tf —Ok \ /
: O C = C + HN0 2

C--Cy y
/ \

and the radical mechanism, i.e., C — N bond breaking with the formation o f
free radicals [3.1]

RN0 -> R + N0 .
2 2 (3.2)

Mononitroalkanes, having an H atom in a-position, a-halogen


nitrocompounds o f R C H C H X N 0 2 2 ( X = Hal) type, and many other ct-
nitroolefines are decomposed at moderate temperatures (up to 3 0 0 ° C ) by
the first mechanism. /zewe-Dinitroalkanes, 1,1,1-trinitroalkanes, a,a-
dihalogenated RCX N0 2 2 due to a reduced C — N bond strength are
decomposed by the second mechanism. A t sufficiently high temperatures,

35
36 Aliphatic Nitrocompounds

reaction (3.2) predominates in all cases because its rate constant has a larger
pre-exponential factor as compared with reaction (3.1).
In the condensed state, the decomposition o f aliphatic nitrocompounds
has a number o f peculiarities connected with the manifestation o f polar and
ionic processes.

3.1 Gas Phase Decomposition by the Molecular


Mechanism

Mononitroalkanes. Homogeneous stages of decomposition

The gas phase decomposition o f mononitroalkanes is complicated by a


heterogeneous reaction on the surface [3.2, 3.3]. The rate dependence on
S/V o f the vessel is considerable in the case o f primary nitroalkanes [3.2]
and is practically absent for ternary nitroalkanes [3.3].

Table 3.1.1. Characteristics of molecular decomposition of nitroalkanes [3.4]

Compound j °£ E, \og(A), w ,
re] D(C-N) Q, kcal/ s , kcal/
0

kcal/ s" 1
570 K kcal/ mol* mof mol
mol

Nitroethane 320-380 45.0 12.24 1.0 57.0 18.4 26.6


1-Nitropropane 320-380 44.9 12.15 1.1 55.9 16.4 28.5
1-Nitrobutane 320-380 43.4 11.79 1.5 54.4 - -
1-Nitropentane 320-380 43.3 11.82 1.7 - - -
1-Nitrohexane 320-380 43.3 11.81 1.8 - - -
2-Methylnitro- 320-380 43.6 11.65 0.9 - - -
propane
3-Methylnitro- 320-380 43.9 11.99 1.6 - - -
butane
2-Nitropropane 280-340 43.5 12.72 11.7 58.4 19.5 24
2-Nitrohexane 280-340 43.7 12.93 15.7 - - -
2-Nitro-3- 280-320 42.8 12.68 - - - -
methylbutane
2-Nitro-3,3- 280-320 42.3 12.74 36.2 - - -
dimethylbutane
2-Nitro-2- 250-320 42.8 13.63 170.0 55.1 22.2 20.6
methylpropane
2-Nitro-2- 250-300 41.3 13.30 312.3 - - -
methylbutane
2-Nitro-3- 250-280 40.7 13.33 363.0 - - -
methylpropane
2-Nitro-2,3- 230-280 40.7 13.43 721.0 - - -
dimethylbutane

* Thermochemical data for calculation of D and O are taken from [3.5]


Thermal Decomposition and Combustion 37

Correspondingly, a noticeable scatter in the literature data, which is caused


by the neglect o f heterogeneous reactions, is observed only in the case o f
primary nitroalkanes [3.1]. Elimination o f the heterogeneous reaction is
approached by the calculation with the use o f the observable rate constant
dependence (k ) ohs on S/V. The characteristics o f homogeneous gas phase
reactions are presented in Table 3.1.1.
During decomposition o f nitroalkanes, corresponding olefins are
formed [3.4, 3.6, 3.7]. In the course o f the reaction, they are oxidized, but
the extent to which these reactions proceed is small [3.3]. In all cases, the
final gas yield is close to 2.5 p , i.e., this value is the same as for the
0

expected initial process

CH- NCb C = C HNCb

,C=C +
0 . 5 H O + 0.5NO + 0 . 5 N O
2 2

Table 3.1.2. Relative yield (/) of alkene-1 during decomposition of nitrocompounds


[3.4]

Compound r,°c % Compound T°C Y %

Exp. Calc. Exp. Calc.

2-Nitrobutane 250 55 60 2-Nitro-2- 190 77 75


methylbutane
330 56 60 250 77 75

2-Nitrohexane 300 58 60 3-Nitro-3- 250 43 42.8


methylpentane

2-Nitro-3- 300 78 75 2-Nitro-2,3- 250 95 85.7


methylbutane dimethylbutane

During decomposition o f di- and r r / - R N 0 , the yield o f alkene-1 and


2

alkene-2 is close to the statistical one and practically does not depend on
temperature (Table 3.1.2). Only in the case o f 2-nitro-2,3-dimethylbutane is
a steric barrier for the formation o f alkene-2 observed.
38 Aliphatic Nitrocompounds

The pre-exponential factor

The increasing o f the pre-exponential factor A , to a large extent, causes that


of the rate on transition from f ? - R N 0 to f n - R N 0 . If we take into account
2 2

the variation o f the reaction statistical factor cr, which is equal to the
doubled number o f hydrogen atoms in /^-positions, then, within the series n-
RN0 , 2 <i/-RN0 , and
2 fr*/-RN0 ,2 the values o f log(v4/cr) remain nearly
constant and are equal to 11.34, 11.80, and 12.35, correspondingly. This
difference is explained by a step-like growth o f the internal rotation barrier
^intem around the C — N bond at ot-alkylation (from K i n t e r n < 1 kcal/mol for n-
RN0 2 up to K i n t e r n > 10 kcal/mol for / n ' - R N 0 ) . A certain influence on the
2

variation o f the A value is also exerted by the increase o f the internal


rotation barrier around the C — C bond. The loss o f this partly hindered
rotation causes a reduced value o f A/a in / r / - R N 0 (in comparison with the 2

"normal" A = 1 0 13 5
s ). The details o f the calculation o f the value A in the
_1

reactions o f elimination from the primary nitroalkanes are presented in


[3.8].

The polar character of the reaction of HNO2 elimination

Reaction (3.1) is endothermic and its activation energy is E = Q + £Q, where


Q is the reaction thermal effect and s is the activation energy o f the reverse
Q

process o f H N 0 2 addition to olefin.


A s can be seen from Table 3.1.1, the values o f Q and s 0 are varied
under the influence o f alkyl substitutes in the opposite directions so that the
value o f E remains nearly constant.
a-Methylation influences the rate o f reaction (3.1) relatively weakly
[3.9], and, by this feature, H N 0 2 elimination reaction can be attributed to
the weak quasi-heterolytic processes (in terms of MacCall [3.10]).
However, by another feature o f these reactions, in accordance with w h i c h
the elimination rate directly depends on the extent o f charge separation in
the transition state, reaction (3.1) can be attributed to highly polar ones,
because it exceeds by its rate all other reactions o f this type. Undoubtedly, a
high electronegativity o f the N 0 - g r o u p and a large dipole moment o f
2

nitroalkane molecule ( « 3.5 D), w h i c h is nearly twice as large as that for


halogen alkanes ( » 2.0 D ) , facilitates the formation o f the strongly polar
transition state. A low polarizability o f the C — N bond and, simultaneously,
a high electronegativity o f the N 0 - g r o u p results in the fact that already in
2

nitroethane this bond is polarized almost as highly as possible and, at the


further variation o f the structure o f the alkyl group, the state o f this bond
and the charge on the nitrogroup practically remain constant, and,
consequently, the decomposition rate should also vary only slightly. One
can estimate the extent o f variation o f C — N bond polarization at a-
methylation from the relative increase o f the dipole moment Aju at the
Thermal Decomposition and Combustion 39

transition from C H X to / r / - C H X . For instance, for X = CI, A / / = 20%, and


3 4 9

for X = N O its value is only 5%. Similarly, the value o f A/u is also small in
the series of alcohols, amines, esters, and isocyanates [3.9], and,
correspondingly, at a-methylation the rate only slightly changes in the case
o f these compounds.

Table 3.1.3. Absolute variation of Ae in the reactions of H X addition to olefins


Q

[3-9]

Olefin Orien- As , kcal/mol


0

tation
Theory Experiment for H X

HCl HBr HI H0 2 CH,OH HN0 2

C H
2 4
_ 0 0 0 0 0 0 0
C H
3 6
M -4.2 -6.0 -4.1 -4.7 -6.2 - -A 2
AM 0.4 -0.6 -0.2 - -0.2 - 0.3
/so-C H 4 8
M -5.9 -8.4 -7.5 - - - -5.7
AM -0.5 -0.8 -0.8 - - - -0.9
M -12.0 -14.8 -11.1 -10.2 -10.3 -10.5 -10.0

Note: M denotes addition according to the Markovnikov rule, A M denotes addition


contrary to the Markovnikov rule.

B y the degree o f variation o f the value o f <? (As ) in a combination reaction


0 0

with olefins o f different structure, F T N 0 is no different from other H X


2

(Table 3.1.3), and this process corresponds quantitatively to the theoretical


calculations o f four-center reactions by the model o f polar transition state,
which is formed from two closely converged semi-ionic pairs [3.11]. Using
the values o f s for nitroethane, As for olefins [3.12], and additive schemes
0 0

of calculation o f the enthalpies o f formation o f nitrocompounds [3.5], one


can estimate the values o f Q and E o f the reactions o f elimination from
many other nitroalkanes. The absence o f a predominate direction o f
elimination from ^/-nitroalkanes is explained by the mutual compensation
o f Q and s (see Table 3.1.1).
Q

a-Halogen-nitroalkanes

a-Halogenation o f nitroalkanes results in a decrease o f the E value and


increase o f the rate (Tables 3.1.1 and 3.1.4). However, this reduction is
small (1-4 kcal/mol) and only in the case o f C H C H C 1 N 0 2 5 2 does its value
approach 5 kcal/mol. O n the whole, the activation energy o f molecular
decomposition o f a-halogen-nitroalkanes, as in the case o f nitroalkanes, lies
within the range o f 40-45 kcal/mol. B y their influence on the rate o f
40 Aliphatic Nitrocompounds

decomposition, all halogens can be ordered in the following series: CI > B r


> F. Such an order o f influence o f halogens can be explained within the
framework o f the concepts o f the polar character o f the transition state.
Induction influence o f halogen atoms, which leads to a decrease o f the
electron density on the N 0 group, has to increase the rate o f decomposition
2

in the series: B r , CI, F. O n the contrary, the possibility o f halogens


exhibiting a positive mesomeric effect, which can be expressed in the
stabilization o f the transition state or in electron transfer towards the N 0 2

group in the ground state

results in the reduction o f the decomposition rate in this series. It is


precisely the superposition o f these two effects that leads to an observable
order o f the influence o f halogens. The same order is also observed at
decomposition of a-polyhalogenalkanes [3.10] and a-halogen-
nitrocompounds by the radical mechanism [3.14].

Table 3.1.4. Characteristics of the reactions of H N 0 2 elimination from halogen-


nitroalkanes [3.13]

Compound T, °C log/* (s" )


1
Wrei*
kcal/mol

(CH )CHF(N0 )3 2 300-340 44.0 12.62 5.4

(C H )CHC1(N0 )
2 5 2 250-300 39.6 12.55 174.0

(C H )CHBr(N0 )
2 5 2 280-320 41.5 12.48 31.6

(CH ) CC1(N0 )
3 2 2
240-300 42.2 13.62 24.5

(CH )(C2H5)CC1(N0 )
3 2 250-280 41.0 13.37 36.3

(CH ) CBr(N0 )
3 2 2 250-290 42.2 13.40 14.8

(CH ) CHCC1(N0 )CH


3 2 2 3 230-270 39.6 13.19 40.0

(CH ) CHCBr(N0 )CH


3 2 2 3 230-270 39.6 13.19 40.0

*w | is the
re ratio of the decomposition rate of halogen-nitroalkane to the
decomposition rate of a similar unsubstituted nitroalkane at the temperature
corresponding to the rate constant k- 10~ for nitroalkane. 4
Thermal Decomposition and Combustion 41

a-Nitro olefins

a-Nitroolefins are close to their saturated analogs by the rate and kinetic
parameters o f the decomposition reactions (Table 3.1.5). /?-Alkylation
influences the rate and activation energy more strongly than a-alkylation.
Actually, the value o f \og(A/o) for decomposition o f nitroalkenes (11.88) is
equal to that for decomposition o f ^/-nitroalkanes (11.80). If the cyclic
transition state is formed, in the case o f nitroalkenes, the internal rotation
around the C — N bond becomes impossible, and in the case o f nitroalkanes
this is true for the rotation around the C — N and C — C bonds. Hence, it
follows from this fact that the potential barrier o f the internal rotation o f
nitrogroup in nitroalkenes is higher than in ^/-nitroalkanes.
Decomposition o f unsaturated nitrocompounds can also proceed by
other mechanisms, but also o f the molecular type. A s shown by the example
o f 2-methyl-3-nitropropen [3.15], at forced c/s-positioning o f N 0 2 and C H 3

groups, an easy decomposition occurs through isomerization

( C H ) C = CHN0 — H C = | C —CHNOOH
3 2 2 2 products.

CH 3

For this compound we w i l l have:

£ = 10 1 , 8
exp(-36200/i?r),s- . 1

In [3.16], the composition o f the decomposition products of


P h C H = C ( C H ) — N 0 was measured at 6 0 0 ° C and 10~ Torr. The main
3 2
3

products are P h C H O and C H C N . From this, one can conclude that the
3

reaction proceeds by the mechanism o f intramolecular oxidation

H CH 3

PhCH=C(CH )N0 — 3 2 P h — C — C — - P h C H O + CH3CN + O .

O - - N - 0

However, a further study o f this reaction is needed to confirm this


mechanism. First o f all, it is necessary to determine its kinetic parameters,
the influence o f P 0 on the composition o f the products and the rate o f
decomposition, and to study the analogs o f nitrostyrene.
42 Aliphatic Nitrocompounds

Table 3.1.5. Characteristics of the molecular decomposition of nitroalkenes [3.13]

Compound T, °C £, kJ/mol log^(s- ) w


!
2

CH =CHN0
2 2 340-400 191.9 12.10 1.0 1.3
CH CH=CHN0
3 2 307-360 176.8 11.82 9.2 8.1
C H CH=CHN0
2 5 2 290-340 176.0 11.92 13.8 8.3
C H CH=CHN0
6 5 2 290-350 176.0 12.04 18.2 —
C H C(N0 )=CH
2 5 2 2 323-380 187.3 12.16 2.6 0.25

* w, is the relative rate at 350°C.


* w is the ratio of the decomposition rate of nitroalkene to the rate of decomposition
2

of its saturated analog at the temperature, corresponding to the magnitude of r - 10~ 4

s- .
1

3.2 Gas Phase Decomposition by the Radical


Mechanism

The chemical mechanism

The decomposition reactions, the first stage o f which is the dissociation o f


the C — N bond, are complicated by chain processes only in the case o f such
mononitrocompounds as nitromethane [3.17-3.20] and chloropicrin [3.21].
Heterogeneous decomposition on the surface is also o f importance for
nitromethane [3.17].

Kinetic scheme 3.2.1

CH N0
3 2 —— CH + N0
!
3 2

CH3NO2 + C H — ^ C H + C H N 0
3 4 2 2

CH + N 0
3 2 - ^ CH3O + N O

CH3NO2 + C H 3 O — CH3OH + C H N O 2

CH N0 —-
2 2 CH 0 + NO
2

CH3 + N O — ^ CH3NO
7
CH3NO HCN + H 0 2
Thermal Decomposition and Combustion 43

Table 3.2.1. Dependence of the yield of the decomposition products of 1,1,1-


trinitroethane (mole fraction) on the degree of transformation at 170°C [3.26]

Compound Degree of transformation, %

2 10 20 37 48 86 100*

CH 3 N0 2 0.09 0.10 0.13 0.15 0.09 0.16 0.11


CH 3 CN 0.01 0.02 0.02 0.03 0.05 0.02 no
CH 3 OH 0.02 0.04 0.02 0.05 0.04 0.04 0.01
CH 3 ON0 2 0.34 0.24 0.22 0.13 0.12 0.12 0.47
CH 3 COOCH 3 no 0.001 0.004 0.008 0.005 0.005 no
CH 3 COOH - 0.10 0.13 0.34 0.42 0.52 -
CH 3 NO 0.20 0.04 0.02 no no no no
CH 3 ONO - traces traces - - - 0.01

* In excess of N 0 2

A minimal scheme describing the decomposition o f methane at all


temperatures up to ignition [3.22-3.24] consists o f 26 reactions, seven o f
which are presented in kinetic scheme 3.2.1. These are kinetically
important, and explain the formation o f the main intermediate products
( C H 0 , N O , N 0 ) and the final products ( C H , C H O H , H C N , H 0 ) at
2 2 4 3 2

moderate temperatures. Polynitrocompounds, starting from heme-


dinitroalkanes, decompose according to the monomolecular mechanism. A s
a rule, the surface, pressure, and impurities do not influence the rate. Nitric
oxide and dioxide, forming at decomposition, inhibit all free radical
processes. The surface reaction is typical only o f substances o f di- and tri-
nitromethane type, which have an H atom at exposition [3.25]. The typical
secondary reactions o f transformation o f polynitrocompounds are presented
in kinetic scheme 3.2.2. The scheme is based on the composition o f the
products o f decomposition o f 1,1,1 -trinitroethane [3.26] (Table 3.2.2).

Table 3.2.2. Kinetic characteristics of the radical decomposition of aliphatic


nitrocompounds

Nos. Compound T, °C £, kcal/ \og(A/a) AS\ e.u. 2


Refs.
mol

1 CH N0 3 2 360-390 55.1 14.34 4.2 [3.17]


620-1200 58.5 16.21 12.72 [3.20]
2 C H N0
2 5 2 490-555 55.6 15.90 15.9 [3.31]
3 f?-C H N0 3 7 2 600-900 58.5 15.36 8.8 [3.32]
4 /so-C H N0 3 7 2 600-900 54.0 15.38 8.9 [3.32]
5 PhCH N0 2 2 380-430 42.1 13.75 1.4 see 3

6 C1 CN0 3 2 140-170 37.4 15.35 8.8 [3.21]


44 Aliphatic Nitrocompounds

7 C1 C(N0 )
2 2 2 110-150 34.3 14.96 7.0 [3.33]
8 C1C(N0 ) 2 3 120-165 36.4 15.27 8.4 [3.33]
9 BrC(N0 ) 2 3 110-160 36.2 15.58 10.0 [3.33]
10 IC(N0 ) 2 3 113-150 34.4 14.77 6.1 [3.33]
11 FC(N0 ) 2 3 178-236 41.9 14.92 6.8 [3-34]
(42.5) (15.20) (8.8) -
12 HFC(N0 ) 2 2 244-264 48.5 16.38 13.5 [3-25]
(47.5) (15.98) (11.4) -
13 F C(N0 )
2 2 2 235-270 47.4 15.88 9.8 [3.34,
3.35]
14 FC1C(N0 ) 2 2 170-214 41.5 15.44 9.0 [3.34,
3.35]
15 FBrC(N0 ) 2 2 170-210 39.5 15.00 7.2 [3.34,
3.35]
16 FIC(N0 ) 2 2 160-197 39.7 15.40 9.0 [3.34,
3.35]
17 HC(N0 ) 2 3 180-200 42.4 15.46 9.3 [3 25]
18 C(N0 ) 2 4 86-177 38.2 15.70 10.4 [3.36]
19 CH CH(N0 )
3 2 2 197-260 47.1 16.44 13.8 [3.37]
20 C H CH(N0 )
2 5 2 2 215-265 48.0 16.60 14.5 [3.37]
21 C H CH(N0 )
3 7 2 2 220-265 48.1 16.70 15.0 [3.37]
22 CH C(N0 ) CH
3 2 2 3 175-210 50.5 18.20 21.8 [3.38]
(47.5) (16.81) (15.5)
23 CH CF(N0 )
3 2 2 214-262 47.7 16.70 15.0 [3.34]
24 CH CBr(N0 )
3 2 2 160-190 40.5 15.84 11.0 [3.14]
25 CH CC1 N03 2 2 190-235 42.6 15.22 8.2 [3-14]
26 C H CC1 N0
2 5 2 2 200-258 42.7 15.15 7.9 [3.14]
27 CH C(N0 )3 2 3 160-210 43.2 16.70 15.0 [3-39]
28 C H C(N0 )
2 5 2 2 160-200 42.3 16.32 13.5 [3.39]
29 C H C(N0 )
3 7 2 3 160-200 43.6 17.22 17.3 [3.39]
(42.5) (16.72) (15.0)
30 FC(N0 ) C(N0 ) F 2 2 2 2 107-180 42.2 17.4 18.2 [3.34]
31 FC(N0 ) C(N0 ) 2 2 2 3 88-140 36.5 16.60 14.5 [3-34]
32 (N0 ) CC(N0 )2 3 2 3 90-135 35.8 16.52 14.1 [3.36]
33 [(N0 ) CCH ] 2 3 2 2 120-140 35.4 15.08 7.5 [3.40]
CHN0 2

34 [(N0 ) CCH ] 2 3 2 2 120-140 35.1 15.42 9.1 [3.40]


CFN0 2

35 [(N0 ) CCH ] 2 3 2 2 120-140 34.9 15.43 9.1 [3.40]


CC1N0 2

36 [(N0 ) CCH ] 2 3 2 2 120-140 34.2 15.15 7.7 [3.40]


BrNO?
37 [F(N0 ) CCH ] 2 2 2 2 150-200 46.0 15.84 11.0 [3.40]
38 [F(N0 ) CCH ] 2 2 2 2 180-220 41.2 15.00 7.1 [3.40]
CHN0 2

39 [F(N0 ) CCH ] 2 2 2 2 160-200 39.5 14.77 6.1 [3-40]


CFN0 2

40 [F(N0 ) CCH ] 2 2 2 2 160-200 39.5 15.48 9.4 [3.40]


CC1N0 2
Thermal Decomposition and Combustion 45

41 [F(N0 ) CCH ] 2 2 2 2 160-200 38.7 15.12 7.7 [3.40]


CBrN0 2

42 [F(NO,) CCH ], 2 2 130-160 36.3 15.22 8.2 [3.40]


C(N0 ) 2 2

43 CH SC(N0 )
3 2 3 50-80 30.8 15.72 10.5 [3.41]
44 PhSC(N0 ) 2 3 80-100 46.30 15.52 9.6 [3.41]
45 2,4-(N0 ) C H SC 2 2 6 3 75-90 31.8 15.72 10.5 [3.41]
(N0 ) 2 3

46 CH SC(N0 ) CH
3 2 2 3 120-150 38.5 17.45 18.4 [3.41]
47 PhSC(N0 ) CH 2 2 3 120-150 38.2 17.70 19.5 [3.41]
48 2,4-(N0 ) C H SC 2 2 6 3 120-135 37.5 17.25 18.2 [3.41]
(N0 ) CH 2 2 3

49 CH S0 C
3 2 160-190 39.3 14.85 6.5 [3.41]
(N0 ) CH 2 2 3

50 PhS0 C (N0 ) CH
2 2 2 3 170-195 38.9 14.70 5.8 [3.41]
51 C1CH 0CH C 2 2 210-256 46.8 16.50 14.0 [3.42]
(N0 ) F 2 2

52 0 NCH CH C
2 2 2 200-235 46.3 16.55 14.3 [3.42]
(N0 ) F 2 2

Note: The data corrected by equation (3.2.1) are presented in parenthesis.


1
The reaction statistical factor which is equal to the number of the equivalent
nitrogroups.
2
Calculated for 200°C from the relationship A = o(ekT/h)exp{A^/R).
3
The data of the authors of this book.

A l l stages o f this scheme, accept the fourth reaction, correspond to the


commonly accepted concepts o f free radical reactions in N 0 and N O 2

environment. The fourth reaction is a new one. It is observed only for


compounds, which have at least two nitrogroups near one carbon atom. The
nitrile yield increases as N O is added. Decomposition o f R C C 1 ( N 0 ) gives 2

R C N but not C 1 C N (see [3.14]). O n addition o f I to C ( N 0 ) and 2 2 4

C 1 C ( N 0 ) , in the first case, I C N is formed, and, in the second case, only


2 3

C 1 C N [3.27], which corresponds to the essence o f the fourth reaction o f


kinetic scheme 3.2.2.
In the case o f compounds o f hexanitroethane type with vicinal location
o f nitrogroups, the decomposition products are olefins. Thus, in [3.28], it
was shown that during pulsed vacuum pyrolysis o f hexanitroethane,
tetranitroethylene is formed with a high yield. Since the kinetic parameters
o f decomposition o f this compound correspond to N 0 detachment but not 2

N 0 elimination, it has to be accepted that olefin is formed as a result o f


2 4

the secondary reaction o f N 0 2 detachment from the (NO - , ) C C ( N O )


3 2

radical. Detachment o f two N O ? molecules during decomposition o f 1,2-


difluoro-tetranitroethane in the gas phase at low pressure [3.29] can be
explained similarly. In [3.30], the reactions o f 1-adamantane radical
46 Aliphatic Nitrocompounds

transformation were studied. This radical is formed as a result o f radical


decomposition o f 1-nitroadamantane at 6 0 0 ° C .

Kinetic scheme 3.2.2

CH C(N0 ) 3 ^ CH C(N0 ) + N 0
2 3 3 2 2 2

CH C(N0 )
3 ^ CH CO + NO + N 0
2 2 3 2

CH C(N0 ) + N 0 — CH3CO + NO + 2N0


3 2 2 2 2

CH C(N0 ) + NO
3 2 2 CH C(NO)(N0 ) —
3 2 2 CH CN +
3 N0 + N0
2 3

CH CO 3 CH + CO 3

CH CO+ N 0
3 2 CH COO + NO 3

CH COO—-CH + C 0
3 3 2

CH COO + RH — ^ CH3COOH + R
3

CH COO + C H
3 3 ^ CH3COOCH3
CH 3 + N0 2 ^CH N0 3 2

CH + N 0 — - CH 0 + NO
3 2 3

CH + NO + M — - CH NO
3 3 HHCN + H0 2

CH NO + 2N0
3 CH + N + N 0 3 2 3

CH NO + N 0
3 2 CH3NO2 + NO
C H + RH
3 CH3OH + R

CH 0 + N 0 — -
3 2 CH ON0 3 2

CH 0 + N 0 — - C H 0 + HN0
3 2 2 2

C H 0 + NO3 ^CH ONO 3

CH 0 + NO 3 -HNO + CH 0 2

2HNO -H 0 + N 0
2 2

N 0 + CH 0
2 2 CO + C 0 + H 0 2 2

The final reaction products are shown are shown in bold.


Thermal Decomposition and Combustion 47

The activation energy of decomposition and C—N bond


strength in polynitrocompounds

The data on the radical decomposition o f nitrocompounds are presented in


Table 3.2.2. The activation energy o f decomposition is determined by the
bond strength C — N , D(C—N) = E — RT aver and varies within a wide
range. Undoubtedly, the value o f D ( C — N ) is influenced by the induction,
steric, and resonant effects o f substitutes. The latter stabilizes the transition
state with a plane radical, which remains after a nitrogroup detachment. A n
abrupt decrease o f the D(C—N) value during the transition from C H N 0 3 2 to
PhCH N0 2 2 is explained by the mesomeric effect. Probably, a very strong
influence o f the sulfur atom in a-position on the D(C—N) value
(compounds 43—50 in Table 3.2.2) is also connected with the same effect.
The steric effect o f voluminous ^-substitutes can be recognized by the
decrease o f activation energy in the series: H < CI < B r < I in methane
derivatives (compounds 1, 6—10, 14—16, and 24—26 in Table 3.2.2).
In [8], the influence o f all a-substitutes on D ( C — N ) is attributed only
to the steric effect, which looks improbable. It is known that in
polynitrocompounds a mutual polarization o f nitrogroups proceeds. In
addition, CI, B r , and N 0 substitutes influence nearly to the same extent the
2

strength o f both C — H and C — F bonds, which are not subjected to steric


hindrances [3.43]. In [3.44], it was shown that, first o f all, the reaction rate
is influenced by the electron charges on a reactive site. In the series o f
similar compounds without a-steric effects (compounds 1, 2, 17, 22, 31,
and 32 in Table 3.2.2) the induction influence is clearly exhibited, which
takes the form:

D ( C - A O = 220-15.2o-\ (3.2.1)

The value o f <r for nitrosubstitutes can be calculated from the data
[3.45, 3.46]. Thus, for C ( N 0 ) C ( N 0 ) , C ( N 0 ) , and C H C ( N 0 ) , the
2 3 2 2 2 3 3 2 2

values o f <j are 4.65, 4.54, and 2.93, correspondingly.


It should also be noted that, in [3.47], a new method o f estimation o f
the activation energy o f radical decomposition reaction o f C-nitro-
compounds was proposed, based on the following empirical correlation:

E = 46.73 + 2 . 5 2 ( A / / C N - 70) + 0 . 2 2 ( A / / C N - 70) , 2

where A / / C N is a simplified energetic characteristic o f the C — N bond


[3.48], which is calculated by the M I N D O / 3 method. This procedure is not
time-consuming.
In addition to the ordinary effects considered, two other regularities
follow from the analysis o f the structure influence on the D ( C — N ) value.
48 Aliphatic Nitrocompounds

The first implies that F atom in a-position affects the bond not as an
electronegative substitute, but as an absolutely neutral H atom or alkyl
group. If the replacement o f the N 0 - g r o u p with CI, B r , or I reduces the 2

D(C—N) value by 0.5-1.0 kcal/mol, successively, the replacement o f N 0 2

with F results in an increase o f the D(C—N) value by 5 kcal/mol. A n


increased stability o f a-fluoronitrocompounds is connected with this fact.
A specific behavior o f fluorine as a-substitute is typical not only o f the
C — N bond, but also o f C — H and C — H a l bonds. The equivalence o f H and
F atoms as ^-substitutes was first noted in [3.49] and can be examined with
the use o f data presented in [3.44].
The second regularity concerns the manifestation o f the field effect (F-
effect) in the case o f pentane derivatives (series 33-35 and 38-42 in Table
3.2.2), which are highly oversaturated by nitrogroups. In the main
conformation o f these compounds [3.50], which is fixed under the influence
o f steric stresses, one o f the oxygen atoms o f the y -nitrogroup approaches
the end carbon atom within a distance as large as the van der Waals radius.
This gives rise to the weakening o f the C — N 0 2 bond due to stabilization o f
the radical, which is formed as a result o f NO? detachment, by the electron
density o f the oxygen atom o f the ^-nitrogroup.

Table 3.2.3. Energies of bond breaking and enthalpies of formation of nitroalkyl


radicals

R D(R—N0 ),
2 A///(R'), kcal/mof
kcal/mol
from D(C—N) from [3.5]

CH N0 2 2 47.0 24.7 31.6


CH(N0 ) 2 2 41.5 33.2 37.5
CH CHN0 3 2 46.2 14.8 18.2
C H CHN0
2 5 2 47.1 12.3 12.9
(CH ) CN0 3 2 2 47.5 8.5 9.8
CH C(N0 )
3 2 2 42.3 21.7 21.8
C H C(N0 )
2 5 2 2 41.7 15.7 14.9
(N0 ) CC(N0 )
2 3 2 2 35.1 66.0 64.7
F(CN0 ) 2 2 41.6 -10.2 -
F CN0
2 2 46.4 -58.0 -
FC(N0 ) C(N0 ) 2 2 2 2 35.8 7.2 -
C1C(N0 ) 2 2 35.6 38.2 -
BrC(N0 ) 2 2 35.4 28.8 -
C(N0 ) 2 3 37.4 47.9 49.8

The values of A//° for R N 0 compounds, required to calculate A//° (R), are taken
r 2 f

from [3.5, 3.51, and 3.52].

The effect increases as the volume o f the second /-substitute increases ( H <
F < C l < Br).
Thermal Decomposition and Combustion 49

Based on the D ( C — N ) values obtained by the kinetic method (from


Table 3.2.2), one can determine the enthalpies o f formation o f free
nitroalkyl radicals (Table 3.2.3) and then the D(C—H), £>(C—C), or D(C—
Hal) values (Table 3.2.4) in nitrocompounds.

Table 3.2.4. The energies of bond breaking in nitrocompounds

Bond D, kcal/mol Bond D, kcal/mol

from from from from


Table [3.12] Table [3.5]
3.2.2 3.2.2

H—CH N0 2 2
89.7 _ C Ht—CH N0
3 2 2 77.1 84.0
H—CH(N0 )CH 2 3 90.8 94.3 CH —CH(N0 )CH
3 2 3 82.2 84.6
H—CH(N0 )C H 2 2 5 94.2 94.9 C H,—CH(N0 )CH
2 2 3 78.7 85.3
H—CH(N0 )C H 2 3 7 91.5 92.4 CH -CH(N0 )C H
3 2 2 5 85.2 84.8
H—C(N0 )(CH ) 2 3 2 93.2 94.6 CH —CH(N0 )(CH )
3 2 3 2
85.2 84.8
H—CH(N0 ) 2 2 99.5 103.8 CH —CH(N0 )
3 2 2
91.7 94.9
H—C(N0 ) CH 2 2 3 97.8 98.3 C H —CH(N0 )
2 S 2 2 85.0 91.1
H—C(N0 ) C H 2 2 2 S
94.5 95.6 C Ht— CH(N0 )
3 2 2
85.3 89.6
H—C(N0 ) C H 2 2 3 7 96.7 95.0 C H —CH(N0 )
4 9 2 2
83.2 87.7
H—C(N0 ) 2 3 100.1 102.0 CH —C(N0 ) CH
3 2 2 3 86.7 85.8
F—C(N0 ) 2 3 101.7 - CH —C(N0 )
3 2 3 94.5 95.4
F—C(N0 ) F 2 2 113.3 - C H —C(N0 )
2 5 2 3
91.3 98.3
F—C(N0 ) CH 2 2 3 105.6 - C H —C(N0 )
4 9 2 3 88.9 90.8
F—C(N0 ) C(N0 ) 2 2 2 3 116.2 - 0 NCH —CH N0
2 2 2 2
72.3 86.1
F — C ( N 0 ) F C ( N 0 ) F 118.1
2 2 2 - (CH ) (N0 )C—3 2 2
70.2 73.1
C(N0 )(CH ) 2 3 2

F—C(N0 ) C(N0 )F 2 2 2
106.8 — (CH )(N0 ) C—3 2 2 63.6 63.8
C(N0 ) (CH ) 2 2 3

CI—C(N0 ) 2 3 93.1 — (CH )(N0 ) C- 3 2 2


61.5 63.5
C(N0 ) 7 3

Br—C(N0 ) 2 3 52.8 - (N0 ) C(N0 ) C—CH


2 3 2 2 3 91.9 89.6
I-C(N0 ) 2 3 34.6 - (N0 ) C-C(N0 )
2 3 2 3 58.9 62.7
CH —CH (N0 )
3 2 2 83.1 88.6 F(N0 ) C—C(N0 ) F 2 2 2 2
63.8 -
C H —CH (N0 )
2 5 2 2
79.6 86.5 F(N0 ) C-C(N0 ) 2 2 2 3 58.3 -

The numerical values o f D differed somewhat from the data presented in


[5], calculated by a pure thermochemical method based on the regularities
o f variation o f enthalpies o f the formation o f compounds, and, thus, they
supplement each other. Note that, as the number o f nitrogroups increases,
the D ( C — H ) value also increases (due to polarization o f the C — H bond).
A s the number o f electronegative substitutes near one carbon atom
increases, the D ( C — H ) value increases, and, i f they are arranged
symmetrically, its value sharply decreases.
50 Aliphatic Nitrocompounds

Regularities of variation of the pre-exponential factor

The pre-exponential factor approaches large magnitudes for R C — N 0 3 2

bond breaking due to the interference o f several factors [3.53]. The


following factors are among the most important:
(1) Reduction o f frequencies o f deformation vibrations, associated with the
C — N bond. The frequency o f vibration, which passes to an out-of-
plane vibration o f the R C radical, decreases by a factor o f two as
3

judged from the secondary kinetic deuterium isotopic effect [3.54]. The
frequencies o f two pendular oscillations o f the R C group and similar 3

vibrations o f the N 0 group are also reduced twice.


2

(2) Release o f internal rotation with respect to the C — N bond. The


maximal contribution o f this factor in AS* is 3 e. u. (0.6 • \og(A)).
(3) Reduction o f frequencies o f the torsional vibrations (up to a free
rotation) o f the R group adjacent to the reactive site. In the case o f
nitrocompounds the alkyl and N 0 groups can manifest themselves as
2

R groups. The effect o f adjacent groups can be very large, in principle,


but it has never been observed earlier.
(4) In unsaturated and aliphatic-aromatic compounds, the negative effect o f
adjacent groups is possible due to the coupling o f the double bond with
a free electron on the reactive site. Thus, in the reaction

C = C ( 1 ) — C ( 2 ) — N 0 — C ~-C(l)~*C(2)
\ + N0 .
2 2

the C ( l ) — C ( 2 ) bond in the transition state w i l l have the order o f 1.5.


Because o f this, the frequency o f a torsional vibration increases and
AS^ w i l l decrease.
In the case o f nitromethane, for which the barrier o f internal rotation
intern is very small [3.55], factors (2)-(4) are absent and a small growth o f
AS* in comparison with the ordinary value AS* = 0 is caused only by the
first factor. The decrease o f the frequencies o f asymmetrical deformation
vibration (1482 c m ) and the four pendular vibrations (1087, 1100, 605,
-1

and 477 c m ) [3.56] by a factor o f 2-2.5 gives rise to the growth o f AS* by
-1

6 e. u. which corresponds to the experimental value.


A detailed calculation o f the pre-exponential factor o f the
decomposition reactions o f C H N 0 , H C ( N 0 ) , F C ( N 0 ) , H F C ( N 0 ) ,
3 2 2 3 2 2 2 2 2

and C H C F ( N 0 ) with the help o f R R K M theory, was carried out by the


3 2 2

quantum-chemical method M P D P in [3.57]. The best agreement with the


experimental data was obtained within the framework o f the model o f a
semi-rigid activated complex with the critical bond length C — N o f 2.45 A .
Introduction o f halogen atoms or N 0 group in the molecule o f 2

nitromethane results in the growth o f K and an increase o f AS? by the


i n t e m
Thermal Decomposition and Combustion 51

additional 2 e. u. In the case o f compounds 6-9, 14-16 in Table 3.2.2, the


^imem value approaches 10 kcal/mol.
A positive effect o f the adjacent groups exhibits itself insignificantly for
N 0 - g r o u p (11, 12, 17, and 18 in Table 3.2.2) and very strongly for the
2

alkyl substitutes (19-23, 27-29 in Table 3.2.2). The data o f N M R studies


testify that the rotation around the C — C bond in 1,1,1-trinitroethane is
sufficiently hindered [3.58]. During decomposition o f this compound, a /?-
deuterium isotopic effect is observed [3.54], which indicates the variation o f
frequencies o f the C H group vibrations in the transition state.
3

A positive effect o f adjacent N 0 " and alkyl-groups is absent in the


2

compounds, which have large CI and B r atoms (compounds 24-26 in Table


3.2.2) in a-position. They hinder the rotation o f the groups not only in the
initial tetrahedral form, but also in a plane structure o f the R C radical. A s 3

has already been mentioned, the steric factors also influence the D(C—N)
value in these compounds.
O w i n g to a negative effect o f the adjacent group, the AS* value o f
PhCH N0 2 2 is smaller than in the case o f C H N 0 . 3 2

3.3 Competition of the Radical and Molecular


Mechanisms
The quantitative regularities described above make it possible to predict the
characteristics o f different decomposition pathways and to determine the
temperature r e q u i i , when the rates o f H N 0 2 elimination and C — N bond
breaking will be the same. Radical decomposition will prevail at
temperatures above T equi \. The data for typical compounds are presented in
Table 3.3.1. Approximate estimates o f r e q u , i for particular mono- and d i -
nitrocompounds are performed in [3.59].

Nitro-nitrite rearrangement

For some years, the possibility o f nitro-nitrite rearrangement has actively


been discussed as an alternative to radical decomposition in addition to
C — N bond breaking, H N 0 2 elimination, and two cases [3.15, 3.16] o f other
intramolecular processes for gas phase transformation o f nitroalkanes. Thus,
in studies o f nitromethane decomposition in a molecular beam [3.60] under
laser heating (1078 c m ) , one could observe C H 0 formation, which could
-1
2

be the product o f nitro-nitrite rearrangement. The rate o f nitro-nitrite


rearrangement was 6 0 % o f the rate o f C — N bond breaking and its
activation energy (55.6 kcal/mol) was found to be close to the energy o f
C — N bond breaking. The first theoretical calculation by the M I N D O / 3
method [3.61] gave the value o f £ a c t i v = 45 kcal/mol for the activation
energy o f nitro-nitrite rearrangement. However, later, more comprehensive
52 Aliphatic Nitrocompounds

and rigorous calculations [3.62] did not confirm this result and, apparently,
led to the reliable conclusion that £ o f nitro-nitrite rearrangement is a c t i v

higher by 10 kcal/mol than the D(C—N) value.

Table 3.3.1. The competition of radical and molecular decomposition mechanisms

Compound Radical decomposition Molecular decomposition T eQUl\ °C

Estimate Experiment Estimate Experiment

E, log/4 E, log/4 E, log/4 E, log/4


kcal/mol (s ) kcal/mol (s ) kcal/mol (s ) kcal/mol (s )
_1 _1 _1 _l

CH CH N0 3 2 2
_ 55.6 15.9 45.0 12.2 340
(CH ) CHN0 3 2 2 58.0 15.0 - — — - 43.5 12.7 1000
(CH ) CN0 3 3 2 55.0 15.2 - - - - 42.8 13.6 1500
CH CH(N0 ) 3 2 2 - - 47.1 16.7 42.0 13.0 - - 80
C H CH(N0 )
2 5 2 2 - - 48.0 16.9 42.0 12.8 - - 50
C H CH-
2 5 48.0 15.0 - - - - 39.6 12.5 550
CI(N0 ) 2

CH3CHFNO2 57.0 14.8 - - - - 44.0 12.6 960


CH CF(N0 ) 3 2 2
- - 47.7 17.0 41.0 13.0 - - 100
C H C(N0 )
2 5 2 3 - - 42.3 16.8 40.0 13.5 - - -130
CH CC1 N0 3 2 2 - - 42.7 15.1 39.0 13.2 - - 160
CH CBr N0 3 2 2 42.5 15.2 - - 40.0 13.2 - - 10
CH CF N0 3 2 2 57.0 14.8 - - 42.0 12.2 - - 1000
CH =CHN0 2 2 65.0 14.5 - - - - 45.8 12.1 1600

In principle, nitro-nitrite rearrangement can be realized not as a


kinetically independent process but as a reaction, accompanying C — N
bond breaking. It is possible in cases where the decomposition proceeds
through the free transition state [3.63], in which separated fragments R and
N0 2 are a considerable distance apart and are engaged in free three-
dimensional rotation. In this state the recombination o f R and N 0 2 can
proceed through the oxygen atom with simultaneous N O detachment. The
free transition state is typical o f strong bond breaking, and in the case o f
nitrocompounds, it takes place in decomposition o f nitrobenzene and its
derivatives [D(C—N) > 70 kcal/mol], for which nitro-nitrite rearrangement
is observed [3.64]. In the case of aliphatic nitrocompounds, the
decomposition always proceeds through the semi-rigid transition state (as
demonstrated by the recombination rate constants o f alkyl radicals with
N0 2 [3.65]), and nitro-nitrite rearrangement is improbable for them even as
a secondary process, accompanying C — N bond breaking. It is possible to
solve the question of the presence of nitro-nitrite rearrangement
experimentally from the study o f C H N 0 3 2 decomposition in the presence o f
N 1 8
0 2 (in the environment o f hydrogen donors R H ) from the dependence o f
the C H 3 0 H / C H 3
1 8
O H ratio on the N , 8
0 2 concentration.
Thermal Decomposition and Combustion 53

3.4 Reactions in the Condensed State


Mononitro- and halogenated nitroalkanes, which have the hydrogen atom in
a-position, dinitrocompounds with -CH(N0 )2 2 end groups, and the
nitroform decompose in the liquid state considerably faster than in vapors,
apparently, due to isomerization in nitronic acids or ionization and
autoprotolysis. For instance, the liquid nitroform decomposes at a
noticeable rate at 8 0 ° C , whereas in vapors, the reaction starts only at 180°C.
The liquid phase reactions are subjected to a very strong influence o f
autocatalysis by N O and N 0 . Their stabilization is also possible with the
2

help o f acidation, which prevents the formation o f aci-forms and ionization.


The influence o f an ac/-form on the decomposition and detonation
ability o f liquid nitrocompounds was clearly demonstrated in [3.66] by
examples o f nitromethane and nitroethane. Reduction o f the critical
diameter o f detonation under the influence o f small additives o f bases
(dimethylenetriamine and pyridine), high pressure, which shifts the
equilibrium in the direction o f ac/-form formation (the molar volume o f the
ac/-form is smaller by 15% than that for the ox<9-form), and UV-radiation,
which also causes isomerization, are typical o f these compounds. A
considerable isotopic effect o f C D N 0 3 2 decomposition is observed at a
pressure o f 10 kbar [3.67].
For liquid nitromethane, the formal kinetic regularities were determined
and the following expression for the rate was obtained in the case o f the
maximal filling o f the vessel by the substance:

A n even lower activation energy (32.5 kcal/mol) was observed for the first
order reaction o f nitromethane decomposition in the liquid phase at
pressures o f 4 0 - 5 0 kbar and temperatures o f the order o f 140°C [3.68]. In
the same work, it was found that at low pressures and temperatures below
120°C, the nitromethane decomposition is governed by the second order
equation.
T o explain the reduction o f the activation energy o f decomposition in
the liquid state in comparison with the gas phase, the following bimolecular
mechanism o f decomposition was proposed in [3.68, 3.69], which was
justified by quantum-chemical calculations:

2CH N03 2 -> ( C H N 0 ) ->


3 2 2 transition state - »

-» C H N O + H O C H N 0 .
3 2 2
54 Aliphatic Nitrocompounds

The enthalpies o f formation o f the dimer and transition state, calculated by


the M I N D O / 3 method in [3.69], were found to be 21.1 and 21.4 kcal/mol,
correspondingly. Thus, the calculated value o f activation energy (42.5
kcal/mol) coincided with the value measured experimentally. However, it is
possible that this is only an accidental coincidence. The reduction o f
activation energy in the liquid phase can also be easily explained by
nitromethane decomposition through the a d - f o r m and autoprotolysis, but
not through the oxygen transfer. Note that the theoretical calculation carried
out in [3.68] refers to free molecules and in no way takes into account the
specific character o f the condensed state. In the gas phase, the bimolecular
interaction is absent. Apparently, one can clear up the question o f the
mechanism o f nitromethane decomposition by means o f a detailed study o f
the decomposition kinetics in solutions. It is desirable to have available
information about transformations o f nitronic acids to identify the reactions
connected with the isomerization o f nitroalkanes. Unfortunately, the
kinetics and mechanism o f these reactions have not been studied until now.
It is known that these substances are unstable [3.70, 3.71]. For nitroalkanes,
which have two hydrogen atoms in a-position, the following water
detachment mechanism was proposed in [3.8], based on quantum-chemical
estimates:

H C ( N 0 ) -> H ( N 0 ) C = N O O H -> H 0 + 0 N O N
2 2 2 2 2 2
+
-> O",

CH N0 3 2 -> H O N O O H - » H 0 + H O N
2 2
+
- » O".

For nitronic acids o f the A r C H = N O O H type, it is supposed that the


transformation into A r C H O and A r C H = N O H is possible [3.72].
Apparently, in the gas phase at high pressures and temperatures o f 4 0 0 -
8 6 0 ° C , nitromethane decomposition proceeds through the aci-form, i.e.,
under supercritical conditions. The activation volume is equal to (15-25)
crr^mor i f the vapor density is 0.6 g/cm\ i.e., has the same value as in the
1

case o f isomerization.

Compound Gas yield Compound Gas yield


volume, volume,
cm g~'3
cm g 3 -1

(N0 ) CCH CH
2 3 2 3 0.5 (CN)(N0 ) CCH CH(OAc) 2 2 2 2 » 25

(N0 ) CCH OCH CH


2 3 2 2 3 > 25 [(N0 )iCCH NH] C=0
2 2 2 0.8

(CN)(N0 ) CCH OCH


2 2 2 3 » 2 5 [ ( N 0 ) C C H ( C H ) N H ] C = 0 > 25 (5h, 90°C)
2 3 3 2

(CN)(N0 ) CCH CH OAc


2 2 2 2 > 25 (N0 ) CCH(OAc)CH
2 3 3 19 (20 min,
100°C)
Thermal Decomposition and Combustion 55

The ionic mechanism o f decomposition was observed in [3.74, 3.75] for


compounds containing the trinitromethyl group in structures o f the
R O C H — C ( N 0 ) 3 type. The analysis o f the values o f volumes o f gas
2 2

releasing during 48 h at 100°C, presented in [3.74], shows that such


compounds differ sharply from their R C H C H C ( N 0 ) analogs by their 2 2 2 3

stability.
For them, the following reaction was proposed:

( N 0 ) C C H O C H C H -> ( N 0 ) C ~ + C H O C H C H .
2 3 2 2 3 2 3 2 2 3
+

The branching on the /^-carbon atom also stabilizes the positive charge and
facilitates heterolytic C — C bond breaking.
One can expect that for polynitrocompounds, which have far removed
substitutes and do not have the hydrogen atom in a-position, decomposition
in the liquid phase w i l l proceed at the same rates as in vapors. Direct
measurements were carried out for C ( N 0 ) , I C ( N 0 ) , M e C ( N 0 ) , and 2 4 2 3 2 3

some other compounds [3.76]. The data on decomposition of


nitrocompounds in the melt and solutions are presented in Table 3.4.1. The
decomposition o f all compounds presented in this table does not depend on
the solvent nature and is no different from decomposition in the gas phase.
O n l y in the case o f hexanitroethane does dependence o f the rate on the
solvent nature manifest itself. One can suppose that, being the strongest
acceptor, hexanitroethane is capable o f experiencing the influence o f even
weak donors due to the formation o f complexes with charge transfer.
In [3.80], an attempt was made to determine directly, using IR-
spectroscopy, the influence of solvation and other intermolecular
interactions on the activation energy o f monomolecular reactions o f bond
dissociation.

Table 3.4.1. The kinetic characteristics of decomposition of nitrocompounds in the


melt and solutions

Compound Environment r, °C E, log/1 Refs.


kcal/
mol

C(N0 ) 2 4 Melt 80-110 37.4 15.50 0.3 [3.76]


CC1 4
85-150 39.7 16.70 0.3 [3.77]
Freon-1 14V 120-135 41.3 17.17 0.3 [3.77]
Freon-1 13 100-135 39.9 16.80 0.3 [3.77]
(N0 ) CCH CH -
2 3 2 2 Melt 120-185 41.4 16.50 - [3.78]
OCH C(N0 ) 2 2 3 TNT 2
150-185 41.5 16.50 - [3.78]
(N0 ) CCH CH -
2 3 2 2 Melt 120-175 41.4 16.60 - [3.78]
COOCH C(N0 ) 2 2 3 TNT 2
140-200 41.4 16.65 - [3-78]
CH C(N0 )
3 2 3 Melt 100-140 42.6 16.93 1.3 [3.79]
C(N0 ) -C(N0 ) 2 3 2 3 CCI4 70-120 37.8 18.6 1.2 [3.79]
56 Aliphatic Nitrocompounds

Heptane 70-120 28.5 13.0 1.4 [3.79]


Cyclohexane 70-120 21.0 9.1 6.7 [3.79]
[F(N0 ) C- 2 2 TNT 2
190-240 46.7 16.5 - [3.78]
CH 0] CH2 2 2

[F(N0 ) CC- 2 2 Melt 140-200 47.0 16.88 - [3.78]


H CH COO-
2 2

CH C(N0 ) F
2 2 2

C1C(N0 ) 2 3 CC1 4
85-146 38.2 16.50 0.45 [3-77]
BrC(N0 ) 2 3
CC1 4 85-131 33.8 16.60 0.85 [3-77]
IC(N0 ) 2 3 Melt 85-120 33.8 15.50 0.1 [3-76]
C1C(N0 ) - 2 2 Freon-113 45-85 26.5 16.60 - [3-79]
C(N0 ) C1 2 2

1
The ratio of the decomposition rate in the liquid phase to the decomposition rate in
vapors at 85°C.
2
2,4,6-Trinitrotoluene.

The method is based on the use o f the following relationship:

E =
4(jn)'

where a>o is the vibration frequency o f the bond, X\\ * s t n e


anharmonicity
coefficient o f the vibration, in which this bond is involved with sufficiently
large weight (the valence vibration o f the C — N bond with co = 920 c i r f in 0
1

nitromethane). In the experiment with nitromethane, the following E values


were obtained: (1) 50 kcal/mol for the liquid state, (2) 46 kcal/mol for the
solution in C C 1 , and (3) 90 kcal/mol for the solid phase (experiments at -
4

1 7 3 ° C ) . Obviously, this method needs more rigorous validation and proof.


Apparently, the influence o f complex formation, the mechanism o f
decomposition o f nitronic acids and the kinetic study o f the heterolytic
reactions proposed are by no means all o f the problems which have to be
studied in order to understand the essence o f the liquid phase decomposition
o f nitrocompounds. N e w aspects can appear in studies o f compounds with
hybrid functions in which d i - and trinitromethyl-groups are the most
reactive parts o f the molecules. Decomposition o f such compounds is
described in Chapter 10.

REFERENCES

3.1. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Usp. Khim., 37:


1443_1461 (1968).
3.2. V . V . Dubikhin, G . M . N a z i n , G . B . Manelis, Izv. Akad. Nauk SSSR,
Ser. Khim., 7: 1412-1415 (1971).
Thermal Decomposition and Combustion 57

3.3. V . V . Dubikhin, D . N . Sokolov, G . M . N a z i n , G . B . Manelis,, Izv.


Akad. Nauk SSSR, Ser. Khim., 7: 1416-1419 (1971).
3.4. V . V . Dubikhin, G . M . N a z i n , G . B . Manelis, Izv. Akad. Nauk SSSR,
Ser. Khim., 7: 1554-1555 (1971).
3.5. Yu. A. Lebedev, E. A. Miroshnichenko, Y u . K . Knobel',
Thermochemistry of Nitrocompounds, M . : Nauka, 168p. (1970).
3.6. D . J. Woddington, M . A . Warns, J. Phys. Chem., 75: 2427-2430
(1971).
3.7. C . D . Weis, G . R. N e w k o m e , J. Org. Chem., 55: 5801-5802 (1990).
3.8. G . M . Khrapkovskiy, Y u . I. Rubtsov, V . A . Rafeev, et al., Proc. of
XVII Int. Pyrotechnics Seminar Combined with II Beijing Int. Symp.
on Pyrotechnics and Explosives, Beijing, 1: 424^130 (1991).
3.9. V . V . Dubikhin, G . M . N a z i n , Izv. Akad. Nauk SSSR, Ser. Khim., 6:
1345-1350 (1974).
3.10. A . M a c C a l l , Quasiheterolytic Reactions in Theoretical Organic
Chemistry, R. K h . Freidlina E d . , M . : Izd. Inostr. Lit., 289p. (1963).
3.11. S. W . Benson, G . R. Haugen, J. Phys. Chem., 70: 3336-3341 (1966).
3.12. G . R. Haugen, S. W . Benson, Int. J. Chem. Kinet., 2: 235-255
(1970).
3.13. V . V . Dubikhin, G . M . N a z i n , G . B . Manelis, Izv. Akad. Nauk SSSR,
Ser. Khim., 4: 923-925 (1974).
3.14. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR,
Ser. Khim., 6: 1239-1243 (1971).
3.15. V . G . Matveev, V . V . Dubikhin, G . M . N a z i n , Kinet i Katai, 17:
280-285 (1976).
3.16. T. H . Kinstle, J. G . Stam, J. Org. Chem., 35: 1771-1774 (1970).
3.17. V . V . Dubikhin, G . M . N a z i n , Izv. Akad. Nauk SSSR, Ser. Khim., 6:
1339-1342 (1971).
3.18. G . G . Crawforth, D . J. Waddington, Trans. Faraday Soc, 65: 1334—
1349 (1969).
3.19. A . A . Borisov, S. M . Kogarko, G . I. Skachkov, Kinet. i Katai, 7:
589-596 (1966).
3.20. K . Glenzer, J. Troe, Helv. Chim. Acta, 55: 2884-2893 (1972).
3.21. V . V . Dubikhin, G . M . N a z i n , G . B . Manelis, Izv. Akad. Nauk SSSR,
Ser. Khim., 6: 1338-1339 (1971).
3.22. A . R. Perche, J. C . Tricot, M . L u c q u i m , J. Chem. Res. (S), 116-117
(1979).
3.23. A . R. Perche, J. C . Tricot, M . L u c q u i m , J. Chem. Res. (S), 304-305
(1979).
3.24. A . R. Perche, M . L u c q u i m , J. Chem. Res. (S), 306-307 (1979).
3.25. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR,
Ser. Khim., 5: 1035-1039 (1969).
3.26. G . M . N a z i n , G . N . Nechiporenko., D . N . Sokolov, G . B . Manelis,
Izv. Akad. Nauk SSSR, Ser. Khim., 2: 315-322 (1968).
58 Aliphatic Nitrocompounds

3.27. G . M . N a z i n , G . B . Manelis, G . N . Nechiporenko, F. I. Dubovitskii,


Combust. Flame, 12: 102-106(1968).
3.28. K . B a u m , D . Tzeng, J. Org. Chem., 50: 2736-2739 (1985).
3.29. K . B a u m , T. G . Archibald, D . Tzeng, et al., J. Org. Chem., 56: 5 3 7 -
539 (1991).
3.30. A . I. Feinstein, E . K . Fields, J. Org. Chem., 36: 996-998 (1971).
3.31. K . Glanzer, J. Troe, Helv. Chim. Acta, 56: 577-585 (1973).
3.32. K . Glanzer, J. Troe, Helv. Chim. Acta, 56: 1691-1698 (1973).
3.33. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Dokl. Akad. Nauk
SSSR, 177: 1387-1389 (1967).
3.34. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR
Ser. Khim., 11: 2631-2693 (1968).
3.35. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR
Ser. Khim., 12: 2801-2803 (1968).
3.36. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR,
177: 1128-1130(1967).
3.37. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR
Ser. Khim., 11: 2629-2630 (1968).
3.38. J. M . Flournoy, J. Chem. Phys., 36: 1107-1108 (1962).
3.39. G . M . N a z i n , G . B . Manelis, F. 1. Dubovitskii, Izv. Akad. Nauk SSSR
Ser. Khim., 2: 389-391 (1968).
3.40. V . N . Grebennikov, V . G . Prokudin, G . M . N a z i n , Izv. Akad. Nauk
SSSR Ser. Khim., 12: 2822-2824 (1984).
3.41. V . N . Grebennikov, V . I. Erashko, G . M . N a z i n , et al., Izv. Akad.
Nauk SSSR Ser. Khim. , 2 : 3 1 0 - 3 1 2 (1977).
3.42. V . N . Grebennikov, G . M . N a z i n , Izv. Akad. Nauk SSSR Ser. Khim.,
7: 1635-1637 (1978).
3.43. L . V . Gurvich, G . V . Karachevtsev, V . N . Kondrat'ev, Y u . A .
Lebedev, V . A . Medvedev, V . K . Potapov, Y u . S. Khodeev,
Energies of Chemical Bond Breaking. Ionization Potentials and
Electron Affinity: Reference Book, M . : Nauka, 35 l p . (1974).
3.44. Y u . Y u . Nikishev, I. Sh. Saifullin, I. F. Falyakhov, Kinet. i Katai,
32: 749-751 (1991).
3.45. V . I. Slovetskii, S. A . Shevelev, A . A . Fainzil'berg, Izv. Akad. Nauk
SSSR, Ser. Khim., 5: 989-995 (1962).
3.46. V . A . P a l ' m , Usp. Khim., 30: 1069-1123 (1961).
3.47. V . A . Shlyapochnikov, K . I. Rezchikova, V . N . Solkan, S. V .
Zakharkinskaya, Izv. Akad. Nauk SSSR, Ser. Khim., 7: 1420-1421
(1980).
3.48. V . A . Shlyapochnikov, K . I. Rezchikova, V . N . Solkan, S. V .
Zakharkinskaya, Izv. Akad. Nauk SSSR, Ser. Khim., 10: 2399-2401
(1982).
3.49. V . I. Vedeneev, A . P. P u r m a l \ Zh. Fiz. Khim., 32: 1472-1475
(1958).
Thermal Decomposition and Combustion 59

3.50. N . G . Y u n d a , I. Y u . K o z y r e v a , G . V . Lagodzinskaya, G . B . Manelis,


Izv. Akad. Nauk SSSR, Ser. Khim., 8: 1772-1780 (1980).
3.51. V . I. Pepekin, Y u . A . Lebedev, A . Y a . A p i n , Dokl. Akad. Nauk
SSSR, 208: 153-155 (1973).
3.52. G . A . Carpenter, M . F. Zimmer, E . E . Boroody, R. A . Robb, J.
Chem. Eng. Data, 15. 553-556 (1970).
3.53. G . M . N a z i n , G . B . Manelis, Izv. Akad. Nauk SSSR, Ser. Khim., 4:
811-816 (1972).
3.54. G . M . N a z i n , G . B . Manelis, F. I. Dubovitskii, Izv. Akad. Nauk SSSR,
Ser. Khim., 11: 2627-2628 (1969).
3.55. D . Stall, E . Westram, G . Zinke, Chemical Thermodynamics of
Organic Compounds, M . : M i r , 340-342 (1971).
3.56. D . C . Smith, G . Y . Ran, J. R. Nielsen, J. Chem. Phys., 18: 706-712
(1950).
3.57. K . I. Rezchikova, V . N . Solkan, S. L . Kuznetsov, Izv. Akad. Nauk
SSSR, Ser. Khim., 2: 339-344 (1990).
3.58. A. V. Kessenikh, S. A. Shevelev, A. A. Fainzil'berg,
Radiospectroscopic and Quantum-Mechanical Methods in Structural
Studies, L . A . Blumenfel'd E d . , M . : Nauka, 194-195 (1965).
3.59. R. Shaw, Intern. J. Chem. Kinet., 5: 261-269 (1973).
3.60. A . M . Wodtke, E . J. Hintsa, Y . T. Lee, J. Phys. Chem., 90: 3 5 4 9 -
3558 (1986).
3.61. M . J. S. Deuar, P. Pitchie, J. Alster, J. Org. Chem., 50: 1031-1036
(1985).
3.62. M . L . M c K e e , J. Phys. Chem., 93: 7365-7369 (1989).
3.63. B . M . Rice, D . L . Thompson, J. Chem. Phys., 93: 7986-8000 (1990).
3.64. B . L . Korsunskii, G . M . N a z i n , V . R. Stepanov, A . A . Fedotov,
Kinet. i Katai., 34: 775-777 (1993).
3.65. G . M . N a z i n , Usp. Khim., 41:1537-1565 (1972).
3.66. R. Engelke, W . L . Earl, C . M . Rohlfing, Intern. J. Chem. Kinet., 18:
1205-1214 (1986).
3.67. R. Shaw, P. S. Decarli, D . S. Ross, et al., Combust, and Flame, 35:
2 3 7 - 2 4 7 (1979).
3.68. G . M . Khrapkovskii, P. N . Stolyarov, V . P. Dorozhkin, et al., Khim.
Fiz., 9: 648-657 (1990).
3.69. G . J. Peirmarini, S. B l o c k , P. J. M i l l e r , J. Phys. Chem., 93: 457-462
(1989).
3.70. E . B . Hadge, J. Amer. Chem. Soc, 73: 2341-2342 (1951).
3.71. A . T. Nielssen, Chemistry of Nitro- and Nitroso-Compounds, G.
Foier and S. S. N o v i k o v Eds., M . : M i r , 1: 284-285 (1972).
3.72. C . F. A l l e n , G . P. Happ, Canad. J. Chem., 42: 650-654 (1964).
3.73. K . R. Brower, J. Org. Chem., 53: 3776-3779 (1988).
3.74. H . G . A d o l p h , Combust, and Flame, 70: 343-347 (1987).
3.75. T. B . B r i l l , R. Subramanian, Combust. Flame, 80: 150-156 (1990).
60 Aliphatic Nitrocompounds

3.76. G . M . N a z i n , G . B . Manelis, F . I. Dubovitskii, Izv. Akad. Nauk SSSR,


Ser. Khim., 11: 2628-2629 (1968).
3.77. H . P. Marshall, F. G . Borgardt, P. N o b l e Jr., J. Phys. Chem., 69: 2 5 -
29(1965).
3.78. V . G . Matveev, L . D . Nazina, G . M . N a z i n , D . A . Nesterenko, L . T.
Eremenko, Izv. Rus. Akad. Nauk, (1998), 12: 2455.
3.79. H . P. Marshall, F. G . Borgardt, P. N o b l e Jr., J. Phys. Chem., 72:
1513-1516 (1968).
3.80. N . V . Chukanov, B . L . Korsunskii, F. I. Dubovitskii, O. V . Anan'ina,
Dokl. Akad. NaukSSSR, 277: 1181-1184 (1984).
Chapter 4

AROMATIC NITROCOMPOUNDS

Decomposition o f aromatic nitrocompounds in the gas phase is a complex


process, in which heterogeneous and secondary chain-radical reactions,
w h i c h affect the parent substance, play a substantial role [4.1, 4.2].
However, under certain conditions, namely, in vessels with the ratio S/V < 3
cm - 1
and P < 0 5 Torr or at higher pressures but in the presence o f inhibitors
(I , N O , N 0 , and P h N O ) , decomposition proceeds homogeneously
2 2 and
follows a monomolecular mechanism. In this case, the rate of
decomposition is determined by the first elementary reaction stage (Figs.
4.1.1,4.1.2, and 4.1.3).

4.1 Gas Phase Decomposition by the Radical


Mechanism
Nitrobenzene and its derivatives, which do not contain O H , N H , C H , and 2 3

N0 2 groups in orr/zo-positon, decompose by C — N bond breaking [4.3].


ort/?o-Derivatives under temperatures above 7 0 0 ° C also have the same
mechanism [4.4]. The characteristics o f a radical decomposition are
presented in Table 4.1.1.

61
62 Aromatic Nitrocompounds

The influence o f para- and weta-substitutes on the activation energy,


i.e., on the C — N bond strength, is insignificant. Such substitutes as N H , 2

CI, B r , and I w h i c h are capable o f a direct resonance coupling with the


reactive site, being in the para-position, increase the D ( C — N ) value. F r o m
the meta-position, their influence is considerably weaker. T h e ortho-
substitudes remove the N 0 - g r o u p from the ring plane, reducing the D(C—
2

N ) value by up to 60 kcal/mol, which is typical o f mononitroalkanes.

Table 4.1.1. The Arrhenius parameters of A r N 0 2 decomposition by the radical


mechanism [4.3]

Compound r,°c £, kcal/mol log/1 (s~ ) l

Nitrobenzene 410-480 69.7 17.3


p-Nitrotoluene 380-450 65.9 16.7
m-Nitrotoluene 400-470 68.0 16.9
p-Nitroaniline 440-510 79.7 17.5
m-Nitroaniline 430-500 72.0 17.5
p-Dinitrobenzene 420-470 68.6 17.1
m-Dinitrobenzene 420-480 68.0 16.9
p-Chloronitrobenzene 420-490 71.7 17.4
m-Chloronitrobenzene 420-470 70.2 17.4
o-Chloronitrobenzene 400-470 67.3 17.0
p-Iodonitrobenzene 440-500 71.6 17.5
p-Bromonitrobenzene 430-500 72.3 17.5
m-Bromonitrobenzene 430-500 71.0 17.4
o-Nitrodiphenyl 390-460 62.9 16.0
m-Trinitrophenol 430-500 71.2 17.4
s/m-Trinitrobenzene 380-470 67.3 17.2

The rate constants exhibit sensitivity to the induction (oi) and resonance
(OR°) effects o f substitutes. The influence o f the latter is described by the
following correlation equations [4.3]:

l o g ( F / * ) =-1.286 ( a
0 I X -(J 1 N 0 2 ) + 1.563 « x -< N O 2 ) +0.55A^

and

l o g ( r 7 * ) =-0.507 (a -a )
0 [X m0i + 0.687 « x -a° )m2 + 0.55Aa , R

where k relates to nitrobenzene ( X = H ) , Aa


0 is a measure o f the polar
R
+

coupling "donor-acceptor" [4.5]. The ratios p/Zpf = 1.214 and PR /p\ = - 1 m

1.353 show that sensitivity to the resonance effect is higher than to the
induction one, and these effects have opposite directions. The reaction is o f
electrophilic character.
Thermal Decomposition and Combustion 63

S/V, cm 1

Fig. 4.1.1 Fig. 4.1.2

Fig. 4.1.1. Pressure dependence of the rate constant of nitrobenzene decomposition


at 450°C for the following S/V (cm" ) values: (1) 0.6, (2) 3.0, and (3) 0.6
1

(nitrobenzene/I = 1:1, P = 100 Torr); (4) 0.6 and (5) 3.0 (decomposition inhibited
2

by I , N O , and PhNO).
2

Fig. 4.1.2. Dependence of the rate constant of nitrobenzene decomposition on the


S/V ratio for the following initial pressures /? (Torr) at T= 450°C: (/) 70, (2) 40, (3)
0

20, and (4) inhibited decomposition, nitrobenzene/I =1:1.


2

Fig. 4.1.3. Pressure dependence of the rate constant of p-nitrotoluene decomposition


at 430°C for the following S/V (cm" ) values: (1) 0.6, (2) 3.0, (3) 18-20 (additives: T
1

is toluene, E is ethylene, B is benzonitrile); (4) inhibited decomposition ((I) NO, (II)


N 0 (S/V = 3.0 cm" ), (III) NO, N 0 , and PhNO (S/V=0.6\
2
1
2 cm' ))
1

In [6], it was shown that the activation energy o f radical decomposition


correlates with the C — N bond length and ionization potential o f the
nitrocompound. From these relationships it was revealed that the
64 Aromatic Nitrocompounds

experimental data on E for m- and p-nitrotoluenes are underestimated by


several kcal/mol.
D u r i n g decomposition o f nitrobenzene and a series o f its derivatives
with E « 70 kcal/mol, the free transition state is realized. Thus, the rate
constant has a high pre-exponential factor, and i f ortho-substitutes are
absent, C — N bond breaking is accompanied by nitro-nitrite rearrangement
[4.7, 4.8]. Transformation o f the primary radicals by the ordinary reactions,
w h i c h is shown, for instance, for nitrobenzene in kinetic scheme 4.1.1, can
easily explain the composition o f decomposition products at early stages. In
the case o f nitrobenzene, such substances as benzene, phenol, and biphenyl
are among these products [4.9, 4.10].
In addition to recombination and substitution reactions, radicals (Ph,
P h O , P h N 0 , and others) enter the reaction o f insertion into the ring, w h i c h
2

results in the formation o f a large number o f new products, in particular,


terphenyl and dibenzofuran. The consumption o f radicals proceeds mainly
on P h N O with the formation o f stable nitroxyl radicals. The radicals are
capable of various transformations. By the reactions of insertion-
detachment, they give amines, azoxycompounds, and polymer products,
whose formation is also observed during decomposition o f nitrobezene.

Kinetic scheme 4.1.1

PhN0 2 Ph + N 0 2

P h N 0 — P h O + NO
2

P h N 0 + P h - ^ — PhO + PhNO
2

PhN0 + P h O — P h O H + P h N 0
2 2

PhN0 + P h — C H
2 6 6 + Ph N 0 2

P h N 0 + Ph N 0 — ^
2 2 PhNO + ' 0 C H N 0 6 4 2

PhN0 + 0 C H N 0 A P h N 0 + H 0 C H N 0
2 6 4 2 6 4 2

Ph + N O - ^ — PhNO

PhNO + 2 N 0 — Ph + N + N 0 2 3

Ph + Ph Ph— Ph

PhO + P h - ^ — PhOPh

The mechanism o f cycle opening and oxidation with the formation o f


CO and C0 2 remains unclear. In [4.11], it was found that during
Thermal Decomposition and Combustion 65

nitrobenzene decomposition the yields o f C O + C 0 and N are the same.


2 2

The authors supposed that the ring oxidation was a consequence o f reaction
(8) o f kinetic scheme 4.1.1. In accordance with the mechanism o f this
reaction [4.12], in addition to N formation, N 0 is also formed, through
2 3

which, probably, carbon oxidation proceeds.

4.2 Molecular Mechanism of Decomposition

ort/zo-Derivatives o f nitrobenzene, containing N H , O H , and C H groups as


2 3

substitutes, are decomposed through the coordinated isomerization in


nitronic acids. For instance,

Me

Apparently, isomerization is the limiting stage. The validations o f this


mechanism are as follows: the low E and A values (Table 4.2.1), isotopic
effects o f D O - , D C - , and D N-substitutes o f nitrobenzene [4.13, 4.14],
3 2

quantum-chemical estimates o f the possible pathways o f o-nitrotoluene and


trinitrotoluol decomposition [4.15] and also the experimental observations
o f the decomposition products o f nitronic acids [4.16, 4.17].
66 Aromatic Nitrocompounds

The characteristics o f pure homogeneous monomolecular reactions are


presented in Table 4.2.1. Judging from the kinetic parameters, o-dinitro-
benzene decomposition also proceeds by a molecular mechanism. In [4.18],
by the method o f decomposition at very low pressures, the formation o f
benzofuroxan from odinitrobenzene was registered, i.e., decomposition
without N 0 2 detachment was observed. Probably, at the first stage, the O
atom is detached with the formation o f the following compound as an
intermediate particle:
.0

The possible existence o f this compound was mentioned in [4.19].


Decomposition o f nitronic acids, which appeared at the first stage, results in
the formation o f free radicals. Thus, all features o f a radical-chain process
are typical o f the decomposition o f oderivatives and the rate of
decomposition goes up as P increases. However, in this case, the series o f
0

stability is not changed. In [4.20], at P 0 > 100 Torr, the following


experimental regularities were observed. The decomposition rate increases
i f the C H - g r o u p in onitrotoluene is replaced by C H
3 2 5 and iso-C H
3 7 groups
(which corresponds to the decrease o f the C — H bond energy) and decreases
at the transition to trinitromesethylene. Apparently, the removal o f the N 0 - 2

group from the ring plane hinders isomerization. The stability o f o-tetra-
butyltrinitrotoluene, whose decomposition occurs by the radical mechanism,
is even higher [4.20]. Unfortunately, the quantitative characteristics o f these
reactions cannot be extended beyond the framework o f experimental
conditions.

4.3 Decomposition in the Liquid State


In the condensed state, the studies o f decomposition o f the following
compounds were performed: s/m-trinitrobenzene (TNB), 2,4,6-trinitro-
toluene (TNT), 2,4,6-trinitroethylbenzene (TNEB), 2,4,6-trinitro-
mesethylene (TNMes), 2,4,6-trinitroaniline (TNA), 2,4,6-trinitrophenyl
( T N P h ) , 2,4,6-trinitroanisole ( T N A n ) , 2,4,6-trinitrocresol ( T N C r ) , trinitro-
carbazole ( T N C a r b ) , trinitrophenylenediamine ( T N P h D A ) , diaminotrinitro-
benzene (DATNB), triaminotrinitrobenzene (TATNB), trinitrophloro-
glucinol (TNPhG), hexanitrodiphenyl (HNDPh), hexanitroazobenzene
(HNAB), hexanitrooxanilide (HNOA), hexanitrodibenzyl (HNDB),
hexanitrodiphenylsulphide (HNDPhSulfide), /-sulfone (HNDSulfon),
2,4,2,4-tetranitrodiphenylamine (TNDPhA), hexa-nitrobenzophenone
( H N B P h ) , hexanitrostilbene (HNSt), l,3,7,9-tetra-nitrophenothiazine-5,5-
Thermal Decomposition and Combustion 67

dioxide (TNPhThD), diaminohexanitrodiphenyl (DAHND), tetranitro-


2,3,5,6-dibenzo-l,3a,4,6a-tetraazapentalen (Takot-1), tertanitro-2,3,4,5-
dibenzo-1,3a,6,6a-tetraazapentalen (Takot-2), /n's-picrylaminotriazine
( T P A M T ) , 2,4,6,2",4',6',4",6"-octanitro-/w-terphenyl ( O N T ) , and tripicryl-
s-triazine ( T P T ) .
In the liquid state, all nitrocompounds o f aromatic series without
exception are decomposed noticeably faster than in the gas phase, though
by their absolute rate they remain rather stable compounds. The activation
energy for the initial rate decreases by 10-25 kcal/mol and the pre-
exponential factor decreases to 10 °— 1 0 1 12
s" . In the case o f trinitrobenzene,
1

the observable rate constant A bserv is 10~ s 0


4 _1
at 4 1 5 ° C in the gas phase and
at 3 5 0 ° C in the melt. In the case o f trinitrotoluene, these temperatures are
350 and 2 1 5 ° C , correspondingly.
The hypothesis about the role o f molecule association [4.20] and the
presence o f bimolecular reactions [4.20, 4.21] (with an unclear mechanism)
in the liquid phase have been proposed to explain the reason for the positive
effect o f the environment. The latter conclusion was made on the basis o f
the dependence o f £ bserv
0
o n
ArN0 2 concentration in a solvent such as
hexachlorbenzene [4.21]. The hypothesis about the role o f the association is
not confirmed because large k\ /k 1( gas ratios are observed not only for
compounds o f a trinitrotoluene type, for which the anomalous value o f the
Truton constant points to association [4.20, 4.21], but also for the
compounds o f a trinitrobenzene type, which have the ordinary Truton's
constant. The direct reaction between two A r N 0 2 molecules is also
improbable, though this reaction cannot be excluded based on quantum-
chemical calculations [4.22]. In the gas phase (for which this calculation
was carried out), this reaction is not observed.
On the basis o f the data on the dependence o f £ bserv
0
o n t n e

nitrocompound concentration, the observable rate constant o f the initial


stage can be divided into two terms: the rate constants o f the first and
second orders [4.21]

^observ =
k +
k [trinitrotoluene] . 0

However, it was found that in the gas phase, k is also noticeably larger than
k . The increase o f &bserv
0 0
m t n e
liquid phase can be naturally explained
within the framework o f the chain mechanism, which occurs in the gas
phase and has to be realized even more efficiently in the liquid phase (due
to the concentration growth). The formation o f stable nitroxyl radicals
during decomposition o f such compounds as trinitrotoluol, picric acid, and
trinitroaniline confirms the chain mechanism [4.23-4.26]. The initial stage
of decomposition can be considered as the chain unbranched process [4.27]
with the rate, which decreases on dilution insignificantly due to the
participation o f a solvent in the decomposition process. Apparently, there
68 Aromatic Nitrocompounds

are no solvents w h i c h would be inert to radicals o f A r and A r O type. The


transition o f a free valence on the solvent and its further reaction with
ArN0 2 provides the superiority o f the observable rate constant in diluted
solutions over £ gas . The data on the decomposition parameters in the liquid
state and in a 0.3% solution o f hexachlorbenzene are presented in Tables
4.3.1 and 4.3.2.
In spite o f the fact that in the liquid phase the decomposition is not an
elementary process, the regularities o f the rate variation on the structure
correspond to the expected ones for the first reaction stage: the rate
increases as the number o f nitrogroups in a molecule increases, the ortho-
substitutes can be ordered by their efficiency in the following series: O H <
NH 2 < CH . 3

Table 4.3.1. The kinetic parameters of the first (k ) and second (k ) order reactions
1 ,r

at thermal decomposition of aromatic nitrocompounds in the liquid state [4.21]

Compound T°C \0 k, (s )at


7 _,
10 *", 7
E, log/*' log/*",
250°C l/mol/s at kcal/mol kcal/mol l/mol/s
250°C

TNB 250-310 1.0 _ 43.0 10.9 _ _


TNT 220-260 2.8 12.6 46.5 12.9 35.4 9.9
TNEB 250 98.0 4.9 - - - -
TNCr 250 2.7 6.0 - - - -
TNMes 250 0.3 1.1 - - - -
TNA 240-280 3.5 6.0 41.8 11.0 32.0 7.2
- -
TNPhDA
HNDPh
290-320
270-310
1.2
2.2 -
44.2
55.5
11.6
16.8 - -
-
TNDPhA 250-300 1.4 - 48.7 14.5 - -
TNCarb 300-340 0.04 0.2 56.9 15.3 47.9 12.4
HNDB 230-280 36.3 40.7 36.6 9.9 45.4 13.4
HNDPh- 250-300 2.8 38.0 50.9 14.7 33.6 8.7
Sulfide

Decomposition of aromatic nitrocompounds is accompanied by


autocatalysis. Catalytic agents are substances o f a basic nature. For instance,
in the case o f trinitrotoluene, the following compounds can be mentioned
among such agents: dinitroanthranyl, aromatic amines o f the A r N H type, 2

w h i c h are formed in reactions o f nitroxyl radicals with activated benzene


compounds, and polymer nitroxyl radicals. For polyfunctional
nitrocompounds, decomposition through several consecutively catalytic
stages is typical. Thus, the equation o f autocatalysis o f the first order
describes only small portions o f the kinetic curves [4.27]. Autocatalysis is
strongly pronounced for such substances as trinitrotoluene and
trinitroanisole, and only slightly manifests itself in the case o f trinitrocresol,
trinitroresorcin, and unsubstituted nitrocompounds. If the ratio o f the rate
Thermal Decomposition and Combustion 69

constants /t catai //: observ is o f the order o f 100, the contribution o f catalytic
reactions becomes noticeable at the earlier stages o f transformation. In these
cases, a pronounced stabilizing effect is possible from the substances
destroying or bonding nitroxyl radicals and weak aromatic amines.

Table 4.3.2. The decomposition characteristics of aromatic nitrocompounds in 0.3


mol % solutions in hexachlorbenzene [4.21]

Compound r,°c k, s" (300°C)


1
E, kcal/mol \ogA (s- )
1

TNT 215-250 1.4 • io- 5


46.5 12.9
TNA 240-280 5.0 • 10~ 6
41.8 11.0
HNDB 230-280 1.6 • io- s
36.6 12.9
HNAB 240-270 6.3 • io- 5
37.5 10.1
HNDPh- 250-300 2.0 • IO" 5
50.9 14.7
Sulfide
HNDPh 275-312 2.0 • io- 5
35.5 16.5
HNBPh 260-292 6.3 •IO" 5
41.6 11.7
DATNB 290-320 1.1 • io- 5
44.2 11.0
HNSt 250-300 1.4 • 10~ 5
45.2 12.9
HNDPh- 250-300 3.2 • io- 4
45.0 13.7
Sulfide
TNPhThD 290-330 5.0 • io- 7
62.1 17.4
TPT 295-334 5.0 •10~ 5
50.1 14.8
ONT 292-334 3.1 • io~ 5
50.5 14.7
Takot-1 290-330 7.9 • 10~~ 6
45.7 12.3
Takot-2 290-330 5.0 • io- 7
62.1 17.4

Unfortunately, because o f the complexity o f the process, there are no


sufficiently proven and detailed chemical schemes o f A r N 0 decomposition 2

in the liquid state. The methods o f stabilization o f these compounds have


not been developed.

4.4 Decomposition in the Solid State


Decomposition o f aromatic nitrocompounds in the solid state is described in
several works. These works are reviewed in [4.28]. The characteristic
feature o f the solid phase decomposition is the low inhibitory effect o f the
crystal lattice, which can be estimated from the ratio o f the initial rates in
the liquid and solid phases. In the case o f H N A B and H N B P h , the
decomposition rates in the solid and liquid phases are described by the
unique Arrhenius dependence. Apparently, at the decomposition
temperatures, the crystals o f these compounds exist in the plastic state. The
point character o f decomposition on the crystal defects at the surface was
revealed for H N D P h - s u l f i d e , H N O A , T N P h T h D , H N S t , and T A T N B from
70 Aromatic Nitrocompounds

the dependence o f the rate on the crystal size and surface wetting.
Nevertheless, even in these cases, the inhibitory effect is small and, e.g., for
T A T N B at the melting point its value is 2.8. The closeness o f the w !iq and
w \ rate values, when the reaction is localized in the solid phase, makes it
so

possible to conclude that the solid phase accelerates the decomposition.


This can be the consequence o f the chain reaction mechanism or the
concentrating o f catalysts on crystal defects.
The characteristics o f the initial decomposition stages of solid
substances are presented in Table 4.4.1. These data are indicative o f the
relative stability o f aromatic nitrocompounds as explosive substances. In
contrast to the relative rates, the Arrhenius parameters presented in Table
4.4.1 are not accurate (as demonstrated by the compensating effect, revealed
in [4.28], which is a ghost effect) and data extrapolation beyond the limits
o f the temperature ranges used is not reliable.
The reaction development in time is characterized by autocatalysis,
topochemical acceleration, and submelting. The ratio w m a x /w o b s e r v o f the
rates in the solid phase is higher than in the liquid phase. Cases are k n o w n
where liquid products are not formed at all ( T A T N B , H N S t , T N P h T h D , and
Takot-1), but the high rate o f acceleration is preserved. The highest rate
increase ( w m a x /w o b s e r v = 300) is observed in the case o f T N P h T h D .
High-melting aromatic nitrocompounds possess a considerable
inhibitory effect o f the crystal lattice and are the champions among
explosives because o f their stability. The most stable substances are O N T ,
TPT, and TNPhThD. The temperature at which & rv
obse
=
10" , 8
s"1

(decomposition o f 0.2% for 6 h) for these substances is close to 2 5 0 ° C . For


liquid T N B , this temperature is 2 2 0 ° C .
A m o n g explosives, aromatic nitrocompounds have the longest history
o f examination. However, the mechanism o f their decomposition in the
liquid and solid states has been poorly investigated. There are also
noticeable gaps in our knowledge o f the gas phase reactions. The secondary
processes have not been studied. For example, the reactions in which the
benzene ring is decomposed and oxidized are unknown. There is no
agreement among authors regarding the mechanism o f the first stage o f the
decomposition reaction in the gas phase for polynitrocompounds with
orr/20-arrangement o f nitrogroups. The data on the composition o f products
and approximate calculations not supported by kinetic measurements, are
most frequently used to make a judgment about the mechanism o f A r N 0 2

decomposition. This situation results in new hypotheses regarding the


mechanism, for instance, about the decomposition of the
trinitrotriaminobenzene molecule in two parts along the C — C bonds [4.30],
or about a direct reaction A r N 0 2 + R H -> A r N O O H + R [4.31], but it does
not give a reliable solution o f the problem. The shortcomings and the main
problems o f studies o f A r N 0 2 decomposition are considered in [4.32] in
detail.
Thermal Decomposition and Combustion 71

Table 4.4.1. Thermal decomposition characteristics of solid aromatic


nitrocompounds [4.28, 4.29]

Compo­ T melt' °C
1
^ r,°c E, kcal/ log// k, s"
1
^lic/^sol /^solutior/^-sol
und mol (250°C) (Trndt) (300°C)

DATNB 289 220-270 47.2 13.2 3.6 IO 7


6 _
TATNB no melt 284-320 41.2 11.6 1.3 • 10~ 6
- 2.8
TNPhG 167 145-165 54.2 22.0 0.2 10 6
35 -
HNDPh 240 215-230 50.0 - 1.0 TO" 7
4.5 3.2
HNDPh- 234 220-233 50.0 14.9 1.0 IO" 7
3.0 -
Sulfide
HND- 334 220-270 26.0 5.2 2.2 IO" 6
- 16.9
Sulfon
HNAB 220 190-220 29.1 6.8 4.5 - IO" 6
1.0 1.0
HNSt 314 260-300 43.9 12.0 4.8 IO" 7
- 15.6
HNOA 332 230-295 51.5 16.0 3.2 IO" 6
2.0 -
TNPh­ 345 300-350 48.0 11.7 4.8 IO" 9
- 2.0
ThD
Takot-1 380 310-350 56.0 15.4 4.0 10" 8
- 18.0
Takot-2 380 310-350 45.4 11.7 5.6 IO" 8
- 3.2
TPAMT 304 250-290 61.4 19.0 2.5 IO" 8
- -
ONT 360 300-350 67.3 18.5 1.9 IO" 9
- 50.8
TPT 355 300-350 64.4 18.2 1.9 IO" 9
- 125.0

REFERENCES

4.1. V . G . Matveev, G . M . N a z i n , Izv. Akad. Nauk, Ser. Khim., 14(2): 7 4 -


78 (1975).
4.2. V . G . Matveev, V . V . Dubikhin, G . M . N a z i n , Izv. Akad. Nauk, Ser.
Khim., 2: 280-285 (1976).
4.3. V . G . Matveev, V . V . Dubikhin, G . M . N a z i n , Izv. Akad. Nauk, Ser.
Khim., 4 : 7 8 3 - 7 8 6 (1978).
4.4. A . C . Gonzalez, C . W . Larson, D . E . M c M i l l a n , D . M . Golden, J.
Phys. Chem., 89: 4809-4814 (1985).
4.5. V . A . Palm, Usp. K h i m . , 30: 1069-1123 (1961).
4.6. G . M . Khrapovskii, E . A . Ermakova, V . A . Rafeev, Izv. Rus. Akad.
Nauk, Ser. Khim., 11, p. 2118-2122 (1994).
4.7. W . Tsang, D . Robaugh, W . G . Mallard, J. Phys. Chem., 90: 5 3 6 8 -
5373 (1986).
4.8. B . L . Korsunskii, V . R. Stepanov, A . A . Fedotov, G . M . N a z i n ,
Kinet. Katai, 34: 775-777 (1993).
4.9. R. E . Fields, S. Meyerson, J. Amer. Chem. Soc., 89: 3224-3228
(1967).
4.10. E . M c C a r t h y , K . O ' B r i e n , J. Org. Chem., 45: 2086-2088 (1980).
72 Aromatic Nitrocompounds

4.11. V . A . Koroban, Y u . Y a . M a k s i m o v , Kinet. Katai, 31: 775-780


(1990).
4.12. J. Brown, J. Amer. Chem. Soc., 7: 2480-2488 (1957).
4.13. V . G . Matveev, V . V . D u b i k h i n , G . M . N a z i n , Izv. Akad. Nauk, Ser.
Khim., 2:474-477(1978).
4.14. S. A . Shackelford, J. W . Beckmann, J. S. W i l k e , J. Org. Chem., 26:
4201-4206(1977).
4.15. A . G . Turner, L . P. Davis, J. Amer. Chem. Soc, 106: 5447-5451
(1984).
4.16. Y . E . He, J. P. C u i , W . C . M a l l a r d , W . Tsang, J. Amer. Chem. Soc,
110: 3754-3759 (1988).
4.17. C . F. H . A l l e n , G . P. Happ, Canad. J. Chem., 42: 650-654 (1964).
4.18. E . K . Fields, S. Meyerson, J. Org. Chem., 37: 3861-3866 (1972).
4.19. F. B . M a l l o r y , K . E . Schneller, C . S. W o o d , J. Org. Chem., 26:
3312-3316 (1961).
4.20. Y u . Y a . M a k s i m o v , V . F. Sopranovich, Combustion and Explosion:
Proc. of IV All-Union Symp. on Combustion and Explosion, M.:
N a u k a , 619-623 (1977).
4.21. Y u . Y a . M a k s i m o v , E . N . Kogut, Zh. Fiz. Khim., 52: 1400-1404
(1978) .
4.22. L . Peide, K . K . Brower, Proc. of XVIII Int. Pyrotechnics Seminar
Combined with II Beijing Int. Symp. of Pyrotechnics and Explosives,
Beijing, 1 : 4 7 3 - 4 7 7 ( 1 9 9 1 ) .
4.23. E . Janzen, J. Amer. Chem. Soc, 87: 3531-3532 (1965).
4.24. T. M . M c K i n n e y , L . F. Warren, I. B . Goldberg, J. T. Swansen, J.
Phys. Chem., 90: 1008-1101 (1986).
4.25. N . I. Boguslavskaya, G . F. Mart'yanova, O. E . Y a k i m c h e n k o , et al.,
Dokl. Akad. Nauk SSSR, 220: 617--620 (1973).
4.26. R. M . Guidre, L . P. Davis, Thermochem. Acta, 32: 1-18 (1995).
4.27. V . V . D u b i k h i n , V . G . Matveev, G . M . N a z i n , Izv. Akad. Nauk, Ser.
Khim., 266-21 \ (1995).
4.28. Y u . Y a . M a k s i m o v , E . N . Kogut, Izv. Vyssh. Uchebn. Zaved., Khim.
Khim. Tekhnol, 20: 349-356 (1977).
4.29. Y u . Y a . M a k s i m o v , T. A . Kuchina, E . N . Kogut, Chemical Physics of
Condensed Explosive Systems, Trudy MkhTI; M . , Issue 104, 2 8 - 3 0
(1979) .
4.30. M . Farber, R. D . Srivastava, Comb. Flame, 42: 165-171 (1981).
4.31. T. M . M c K i n n e y , L . F. Warren, and I. B . Goldberg, J. Phys. Chem.,
90: 1008-1011 (1986).
4.32. B . B r i l l , K . J. James, Chem. Rev., 93: 2667-2692 (1993).
Chapter 5

SECONDARY NITRAMINES

5.1 First Stage of Decomposition


Thermal decomposition o f secondary nitramines always starts from N —
N0 2 bond breaking. Note that the most important explosives such as
hexogen ( R D X ) and octogen ( H M X ) belong to the secondary nitramines.
However, from time to time other mechanisms o f the first decomposition
stage are proposed. Thus, for the simplest substance o f this class,
dimethylnitramine ( D M N A ) , based on the composition o f the products o f
laser-induced pyrolysis ( C H 0 , N O , N 0 ) in a molecular beam at 900 K , it
2 2

was concluded [5.1] that nitro-nitrite rearrangement occurs together with


N — N bond breaking:

(CH ) NN0
3 2 2
(CH ) NONO.
3 2

These data were not later confirmed experimentally in [5.2]. Apparently, in


[5.1], photolysis products were observed. In addition, this rearrangement is
not in agreement with quantum-chemical calculations [5.3], which also
confirm the primary N — N bond breaking. M a n y different mechanisms o f
the first stage o f decomposition were proposed for cyclic nitramines (see
review [5.4]): cycle expansion to eliminate N 0 and C H 0
2 2

heterolytic breaking o f the C — N bond

73
74 Secondary Nitramines

homolytic breaking o f the C — N bond, H N 0 2 elimination through the 4-


and 5-member transition state, and CH NN0
2 2 decomposition into
fragments. A l l o f these mechanisms were rejected by R. Shaw and F.
Walker in [5.4] for energetic reasons. The assignment o f the products o f
secondary reactions to the first stage o f decomposition was the source o f
erroneous assumptions.
The reaction o f ring reduction was found to be a wrong hypothesis:

This hypothesis was postulated in [5.5, 5.6] upon studies o f R D X and H M X


by the method o f high-vacuum pyrolysis. It was shown in [5.7] that
particles with the ratio m/e = 148 (i.e., [ C H N N 0 ] ) were formed from the
2 2 2

molecules o f nitramines as a result o f electron impact, but not as a result o f


thermal decomposition (in [5.5], these processes were not separated). O n
the contrary, the fragments o f the primary decomposition products o f R D X
[5.7]

N 0 2

indicate that the cycles are preserved at the first decomposition stages.
For solid H M X , on the basis o f the structural fact o f a close contact
(less than the van der Waals radius) o f a nitrogroup o f one molecule with a
Thermal Decomposition and Combustion 75

carbon atom o f the other molecule, the following bimolecular reaction o f


direct hydrogen atom detachment was proposed [5.8]:

NN0 + H C NNOOH + HC
2 2

The high yield o f nitro-derivatives o f R D X during decomposition in the


solutions and isotopic effects o f a solvent [5.9] and deuterium-containing
HMX [5.10] was also the reason for the introduction o f the following
reaction:

NN0 2 + RH NNO + OH+R.

where R H is the molecule o f a solvent or nitramine itself.


The most popular is the opinion that the primary decomposition o f the
secondary nitramines both in the gas phase and in the condensed state
proceeds, at least partly, through HN0 2 elimination [5.11-5.13]. The
occurrence o f a deuterium kinetic isotopic effect, which was observed
during D M N A - d , R D X - d , and H M X - d decomposition [see 5.13], is its
6 6 8

main evidence. However, this isotopic effect can be, to a large extent,
caused by the secondary /^-deuterium isotopic effect, which accompanies
the N — N 0 2 bond breaking, and, in addition, can be connected with
bimolecular reactions between the parent substance and decomposition
products (especially in the solutions).
The common source o f misinterpretations is the assignment o f the
products o f the secondary reactions to the first decomposition stage.
However, kinetic processes, which would be faster than N — N bond
breaking, have never been observed. The constancy o f the activation energy
for the initial stage, which lies in the range 38-40 kcal/mol for the aliphatic
series o f nitramines, testifies that the mechanism is the same for various
substances and aggregate states.

5.2 Secondary Reactions


Taking into account the kinetic data, information regarding the composition
o f the products, and the results o f quantum-mechanical calculations [5.2,
5.3, 5.11-5.17], the following detailed kinetic scheme of DMNA
decomposition can be proposed:
76 Secondary Nitramines

(CH ) NN0 —L^(CH ) N' + N 0


3 2 2 3 2 2

(CH ) N + N 0 ^ ( C H ) N O + NO
3 2 2 3 2

(CH ) N + NO
3 2 (CH ) NNO 3 2

(CH ) N3 2 C H = NCH + H 2 3

(CH ) N —2—
3 2 CH3NHCH2

(CH ) N3 2 C H + CH NO 3 3

CH NO — C H
3 3 + NO
CH NO
3 C H = NOH — H 0 + HCN
2 2

CH NO+2NO-^N0
3 3 + N + CH 2 3

CH NO+N0 -^CH N0 +
3 2 3 2 NO
CH + N 0
3 2 CH N0 3 2

CH + N 0
3 2 C H 0 + NO 3

CH 0+ N 0
3 2 CH 0+ H N 0 2 2

HN0 2 H 0 + NO + N 0
2 2

(CH ) N + ( C H ) N N 0 ^ (CH ) NO + (CH ) NNO


3 2 3 2 2 3 2 3 2

(CH ) NO+ ( C H ) N N 0 - ( C H ) N O H
3 2 3 2 2
M
3 2 +'CH N(N0 )CH 2 2 3

H 0 + CH ^ N C H
2 2 3 CH 0+ N 0 + C H 2 2 3

(CH ) NN0 + H(CH )3 2 2 3 (CH ) NNO + OH + (CH OH)


3 2 3

(CH ) NN0 + OH(CH OH) I*L CH N(N0 )CH + H 0(CH OH)


3 2 2 3 2 2 3 2 3

C H 0 + N^O+CH 2 3

( C H ) N N O + N 0 - ^ — ( C H ) N N 0 + NO
3 2 2 3 2 2

It is also believed that interaction o f the ( C H ) N radical and N 0 with the 3 2 2

formation o f two C H N O molecules can proceed in one stage [5.3]: 3

(CH ) N+N0 3 2 2 2CH NO. 3

Reaction (4) w i l l prevail at low pressures. O n the contrary, reaction (15) can
exhibit itself at a high pressure o f D M N A vapors and at early stages, for the
low concentrations o f N O and N 0 2 in the system. The appearance o f H
atoms and C H radicals can result in short chain processes (due to inhibition
3

by N O and N 0 ) . The decomposition reactions o f cyclic nitramines proceed


2

similarly. The opening o f cycles by reaction o f type 2 (2a)


- C H \ . 2 , r C H
2' - J ° r CH NO
2

> i + NO > N
-
L H NO
2

-CH 2 / L
C H 2 ^ C 2
Thermal Decomposition and Combustion 77

with the following N O detachment results in the structures

yo 2 NO 2

*CH 2 N CH 2 N CH 2 ^

which can be easily split into C H 0 and N 0 , i.e., into the main products o f
2 2

R D X and H M X decomposition. Reaction (20) can proceed only at early


stages for the low concentration o f C H 0 . This compound can bind N 0
2 2 in
a very fast chain reaction o f oxidation

CH 0 + N0
2 2 — CO + C 0 + NO + H 0 .
2 2

The cycle can also be opened in other reactions, for instance, at intra- or
intermolecular oxidation after the appearance o f a free valence on the
carbon atom. The most probable processes proposed by M . Schroeder [5.11]
are the following:

N + NCH 2

N0 2

and
H

N 0
2

The derivatives o f triazine are formed by reaction o f type (4) from R D X


[5.7]. The inhibitory influence o f N O on the gas phase decomposition o f
RDX and l,3-dinitro-l,3-diazacyclopentane (DNI) testifies to the
occurrence o f the chain reactions o f (17-18) type during decomposition o f
cyclic nitramines [5.18].
78 Secondary Nitramines

In hydrocarbon solvents o f R H , due to the transfer o f the free valence


from H , > N ' , and > N — O ' on the solvent molecules, the radicals R appear,
and the formation o f nitrosocompounds as intermediate products is
intensified:

N0 2 + R > N N O + RO
L
C H ^ L
C H /

Thus, the following chain process can proceed:

> N — N 0 2 + RO* > N — N0 2 + ROH


L
C H 2 ^ / L
CH 2

rCHO RH

^CH 2

RO' + R H — R + ROH .

Due to this process, the kinetic isotopic effect o f the solvent is observed.
Picric acid and C H , C H O N 0 , and C H O N O compounds, formed through
4 3 2 3

the C H radical, were observed in the products in the case o f tetryl. For
3

tetryl and its analogs, the following scheme o f decomposition was proposed
[5.19]:

ArN(CH )N0 3 2 ArNCH 3 + N0 2

ArNCH + N 0 3 2 ^ ArN(CH )0 + NO 3

ArN(CH )NO 3 ArON=NCH 3

ArON=NCH 3 — - ArO + N +C H 2 3

ArN(CH )0 — - ArNH + C H 0
3 2

ArNH + NO ^ ArNHO

ArNHO ^ ArN=NOH ArOH +N 2

CH 0 + N0
2 2 ^ CO + C 0 + H 0 +NO.
2 2

M a n y secondary reactions result in the formation o f a large number o f final


decomposition products [5.15, 5.16, 5.20]. Thus, in the case o f R D X , the
following compounds are formed besides N and all types o f nitrogen and 2

carbon oxides: HCN, HCOOH, HOCH NHCOH, 2 HCOOCH3,


Thermal Decomposition and Combustion 79

H C O O N H C H O , N H , 1,2-dihydro-^w-triazine, polymers ( C H N 0 C H ) ,
3 2 2 2 n

and about 10 unidentified compounds. Some o f them catalyze the


decomposition by the mechanism o f chain oxidation-reduction and acid-
base catalysis. The role o f autocatalysis during decomposition o f the
secondary nitramines is usually small because o f a l o w efficiency o f
catalysts and their low concentration.
The secondary decomposition reactions are important for detonation
and explosion because o f their exothermicity. The R D X decomposition can
be easily divided into two stages by the methods o f fast heating [5.21, 5.22]
( 1 0 0 ° C per second): the first endothermic stage and the second exothermic
stage caused by oxidation o f organic particles by N 0 and N O .
2

5.3 Kinetic Data


All o f the kinetic data available on decomposition o f the secondary
nitramines and some o f their nitrosoanalogs are presented in Table 5.3.1. In
spite o f the data scattering caused by imperfection o f the methods, which
cannot separate the initial and catalytic stages or phase states o f the
substance, and also incorrect solution o f the inverse kinetic problem in the
methods o f differential scanning calorimetry, thermogravimetric analysis,
and differential thermal analysis, as well as neglect o f ghost factors
influencing the decomposition o f solid substances, one can observe a series
o f entirely definite regularities concerning the influence o f the structure on
the stability o f nitramines. The main regularity implies that the N — N 0 2

bond is a conservative one and is weakly susceptible to the induction


influence o f substitutes. Correspondingly, the decomposition rates in the
series o f nitramines vary within comparatively small limits. Thus, according
to data 18.1-18.7 o f Table 5.3.1, in [5.56], the dependence o f the rate
constant on the induction effect o f substitutes in R * N ( N 0 ) R was revealed
2
2

to be very weak:

logA: = 0.495 (a' +a' -434)


Rl Rl .

In tetryl, a benzene ring reduces the D ( C — N ) value only by 5 kcal/mol in


comparison with alkyl substitutes. A t the same time, for the C — C , C — H ,
and C — H a l bonds the value o f this reduction is as large as 15-20 kcal/mol.
The geometric structure o f the nitramine group is o f the utmost importance.
Each o f the substances, D N I and R D X , has two pyramidal nitramine
groups, and the rate o f their decomposition in vapors (with inhibitors) is one
order o f magnitude higher than in the case o f D M N A , diethylnitramine,
R D X , and other compounds, having only plane nitramine groups. In [5.18],
it was revealed that the £>(N—N) value in plane nitramine groups was 40.5
kcal/mol and in pyramidal groups, 37.7 kcal/mol. The average value o f the
80 Secondary Nitramines

pre-exponential factor for the reaction o f the N — N 0 2 bond breaking is 10


s~\ Thereby, in the gas phase, the rate o f monomolecular decomposition o f
nitramines is determined by the number and structure o f nitrogroups.
Contribution o f the chain processes to the observable rate constant is
possible. Heterogeneous reactions during decomposition o f vapors o f
nitramines have not been observed. Particular inhibition o f the
decomposition is possible due to the cage effect. A s a rule, the liquid phase
reactions occur with a relatively weak autocatalysis. In contrast to
dimethylnitrosoamine, which is noticeably more stable than D M N A , cyclic
N-nitrosocompounds are less stable than corresponding nitramines, and they
are formed only as intermediate products.
In the solid state, the decomposition rates are considerably lower than
those in the liquid phase or in vapors. The larger the enthalpy o f
sublimation, the melting temperature, and molecular weight o f the
substance, the higher the inhibition o f the reaction in the crystal lattice. The
ultimate values o f the ratio k /k \ gas so = 10 —10 were obtained for w e l l -
3 4

purified samples o f / 2 - H M X and R D X . In these cases, the reaction is


localized on the surface and crystal imperfections, and the activation energy
£ i
so coincides with the value o f £ g a s . The gas phase decomposition,
interaction with the gas phase products, submelting, and, especially, the
purity o f a sample strongly influence the decomposition o f solid R D X and
H M X , as well as the other nitramines. The scatter o f the experimental data
on k \ and, correspondingly, a wide range o f the observable values o f the
so

k /k \ ratio are explained by the influence o f these factors.


gas so

Table 5.3.1. The kinetic characteristics of decomposition reactions of the secondary


nitramines

Nos. Compound Medium Method T, °C E, \ogA k\0 , s~ 5 l


Refs.
kcal/mol (s- ) (200°C) 1

1.1 Dimethyl- Gas MM* - 46.2 15.8 0.3 [5-4]


nitramine
1.2 Dimethyl- Gas MM* 180- 40.6 14.1 0.9 [5.11]
nitramine 240
1.3 Dimethyl- Gas MM* 40.8 14.1 1.8 [5.23]
nitramine
1.4 Dimethyl- Gas + M M * 180- 38.9 13.7 0.6 [5.24]
nitramine CH 0 2 240
1.5 Dimethyl- Solution LC* 200- 46.7 17.8 17.9 [5.13]
nitramine in iso- 300
octane
1.6 Dimethyl- Melt LC* 200- 38.0 13.2 4.8 [5.13]
nitramine 280
2.1 Diethyl- Gas M M * 180- 41.6 15.1 1.3 [5.25]
nitramine 240
2.2 Diethyl- Solution LC* 200- 45.1 16.6 2.6 [5-13]
Thermal Decomposition and Combustion 81

nitramine in iso- 300


octane
3.1 Dipropyl- Gas MM* 200- 42.1 14.8 2.2 [5.25]
nitramine 240
3.2 Dipropyl- Solution LC* 220- 48.0 17.6 2.9 [5-13]
nitramine in iso- 260
octane
4.1 Di-iso- Solution LC* 200- 42.6 15.9 18.4 [5-13]
propyl- in iso- 300
nitramine octane
4.2 Di-iso- Solution LC* 220- 43.9 16.1 9.2 [5-13]
propyl- in acetone 250
nitramine
5.0 N-Nitro- Solution MM* 210- 38.9 13.6 4.2 [5-26]
morpholine in acetone 245
6.1 N-Nitro- Solution LC* 220- 48.2 16.6 0.2 [5.13]
pipyridine in benzene 300
6.2 N-Nitro- Solution LC* 240- 47.5 16.3 0.3 [5.13]
pipyridine in acetone 290
7.0 N-Nitro- Solution LC* 240- 51.5 18.4 0.4 [5-13]
pyrolidine in acetone 290
8.1 N,N- Gas MM* 200- 38.2 13.6 8.9 [5-18]
Dinitro- 240
pipyrazine
8.2 N,N- Gas + N O MM* 220- 36.1 12.0 0.2 [5.18]
Dinitro- 250
pipyrazine
8.3 N,N- Solution MM* 225- 37.1 12.0 0.8 [5.26]
Dinitro- in nitro­ 245
pipyrazine benzene
8.4 N,N- Solution MM* 230- 45.8 15.7 0.4 [5-27]
Dinitro- in trinitro­ 245
pipyrazine benzene
8.5 N,N- Solution LC* 240- 52.6 19.1 0.6 [5.13]
Dinitro- in acetone 290
pipyrazine
8.6 N,N- Melt MM* 216- 47.4 17.3 2.5 [5.26]
Dinitro- 250
pipyrazine
9.1 1,3-Dinitro- Gas MM* 170- 35.0 13.5 213 [5-18]
1,3-diaza- 200
cyclopen-
tane
9.2 1,3-Dinitro- Gas + N O MM* 170- 40.4 15.6 87 [5.18]
1,3-diaza- 200
cyclopen-
tane
9.3 1,3-Dinitro- Solution M M * 170- 37.2 13.6 26 [5.26]
1,3-diaza- in m-di- 210
cyclopen- nitro-
tane benzene
9.4 1,3-Dinitro- Solution LC* 200- 47.8 19.1 132 [5.13]
82 Secondary Nitramines

1,3-diaza- in acetone 240


cyclopen-
tane
9.5 1,3-Dinitro Solution M M * 120- 37.9 14.0 29 [5.27]
1,3-diaza- in trinitro­ 210
cyclopen- benzene
tane
9.6 1,3-Dinitro- Melt M M * 135- 37.1 13.9 57 [5.26]
1,3-diaza- 200
cyclopen-
tane
9.7 1,3-Dinitro- Melt LC* 200- 47.2 18.7 77 [5.13]
1,3-diaza- 250
cyclopen-
tane
10.1 1,3-Dinitro- Solution LC 200- 49.1 18.0 2.0 [5.13]
1,3-diaza- in acetone 280
cyclo-
hexane
10.2 1,3-Dinitro- Solution LC 200- 40.0 14.6 13.0 [5.13]
1,3-diaza- in ethanol 300
cyclo-
hexane
10.3 1,3-Dinitro- Melt LC* 240- 41.9 15.4 11.0 [5.13]
1,3-diaza- 260
cyclo-
hexane
11.1 1,3,3- Solution LC 189- 43.6 16.4 20.0 [5.13]
Trinitro- in acetone 263
azetidine
11.2 1,3,3- 220- 43.7 16.3 12.0 [5.13]
Melt LC*
Trinitro- 250
azetidine
12.1 Trans- 204- 50.1 18.9 6.0 [5.28]
Solid DSC*
1,4,5,8- 234
tetranitro
1,4,5,8-
tetra-aza-
decalin
13.1 RDX Gas DSC* 34.1 13.5 550.0 [5.29]
Gas M M * 170- 30.0 11.7 690.0 [5.30]
13.2 RDX 190
Gas M M * 170- 35.0 13.5 213.0 [5.18]
13.3 RDX 200
Gas + N O M M * 170- 40.4 15.6 85.0 [5.18]
13.4 RDX 200
Gas + N O M M * 40.4 16.0 220.0 [5.31]
13.5 RDX Solution M M * 195- 41.0 15.5 40.0 [5.29]
13.6 RDX in trinitro­ 280
toluene
13.7 RDX Solution M M * 201- 41.5 15.4 33.0 [5.29]
in dicyclo- 280
Thermal Decomposition and Combustion 83

hexyl-
phthala-
te
13.8 RDX Solution MM* 160- 39.7 14.3 9.0 [5.32]
in meta- 200
dinitro-
benzene
13.9 RDX Solution MM* 166- 36.9 13.9 71.0 [5.33]
in trinitro­ 184
toluene
13.10 RDX Solution MM* — 37.8 14.5 110.0 [5.33]
in benzo-
phenone
13.11 RDX Solution DSC* — 48.0 — — [5.34]
in poly-
phenyl
ether
13.12 RDX Solution LC* 200- 45.4 17.9 78.0 [5.13]
in acetone 240
13.13 RDX Solution LC* 206- 38.5 14.1 23.0 [5.13]
in benzene 256
13.14 RDX Melt LC* 200- 37.8 14.3 54.0 [5.13]
250
13.15 RDX Melt MM* 213- 47.5 18.5 35.0 [5.29]
299
13.16 RDX Melt DSC* - 67.4 - - [5.34]
13.17 RDX Melt GLC* 48.7 19.2 53.0 [5.35]
13.18 RDX Melt DSC* 210- 45.2 — — [5.36]
261
13.19 RDX Melt DSC* - 43.1 16.4 31.0 [5.37]
13.20 RDX Melt IC* 204- 41.0 — — [5.38]
225
13.21 RDX Melt DSC* - 42.0 - - [5.38]
13.22 RDX Melt DSC* 204- 50.1 19.6 28.0 [5.39]
240
13.23 RDX Melt TGA* 205- 47.8 18.7 41.0 [5.40]
220
13.24 RDX Solid M M * 150- 50.9 18.6 12.0 [5.41]
197
13.25 RDX Solid M M * 150- 52.0 19.1 12.0 [5.32]
197
13.26 RDX Solid M M * 140- 39.8 11.2 0.01 [5.42]
190
13.27 RDX Solid M M * 178- 45.5 17.7 48.0 [5.43]
200
13.28 RDX Solid M M * 170- 43.0- _ — [5.44]
198 63.0
13.29 RDX Solid + M M * 170- 44.0 — - [5.45]
CH 0
2 197
13.30 RDX Solid IM* 52.4 - [5.46]
13.31 RDX Solid MM* - 41.5 14.5 2.0 [5.47]
84 Secondary Nitramines

13.32 RDX Solid MM* 130- 39.5 11.7 0.03 [5.30]


180
14.1 HMX Gas M M * 171- 52.9 20.2 2630 [5.30]
215
14.2 HMX Gas M M * 230- 32.0 13.2 6.0 [5.29]
250
14.3 HMX Gas M M * 205- 39.5 14.2 9.0 [5.18]
280
14.4 HMX Gas MM* - 38.0 12.5 0.9 [5.31]
14.5 HMX Solution M M * 171- 44.9 16.0 1.5 [5.32]
in meta- 215
dinitro-
benzene
14.6 HMX Solution LC* 189- 50.2 18.9 5.7 [5-13]
in acetone 289
14.7 HMX Melt M M * 271- 52.7 19.7 2.3 [5.29]
314
14.8 HMX Melt DSC* 271- 51.3 18.8 1.2 [5.43]
285
14.9 HMX Melt DSC* - 44.0 - - [5.43]
14.10 HMX Solid M M * 176- 37.9 11.2 0.04 [5.32]
230
14.11 HMX Solid D T A 245- 59.0- - - [5.49]
280 67.0
14.12 HMX Solid DSC* - 180 - - [5.36]
14.13 HMX Solid M M * 150- 36.0 10.8 0.15 [5.47]
170
14.14 HMX Solid M M * 180- 41.0 12.8 0.10 [5.30]
210
14.15 HMX Solid DTA* 180- 41.2 14.3 2.0 [5-50]
280
14.16 HMX Solid DTA* 160- 27.0 10.0 170.0 [5.50]
250
14.17 HMX Solid DSC* 251- 42.0 - - [5.38]
263
14.18 HMX Solid M M * 261- 49.0 17.8 0.55 [5.51]
276
14.19 HMX Solid M M * 130- 37.9 9.2 IO' 3
[5.52]
180
14.20 HMX Solid LC* 230- 52.9 18.4 9.0 [5.13]
270
15.0 1,4-Dinitro- Solid DSC* 225- 52.0 20.9 27.0 [5.53]
tetrahydro- 245
imidazo-
[4,5-d]-
imidazole-
2,3(1H,3H>
dionine
16.0 1,5- Solid DSC* 250- 42.6 15.3 4.0 [5.53]
Diacetyl- 270
3,7-dinitro-
1,3,5,7-
Thermal Decomposition and Combustion 85

tetraaza-
cyclo-
octane
17.0 1,5- Solid - - 45.9 16.6 4.0 [5.54]
Methane-
3,7-dinitro-
1,3,5,7-
tetraaza-
cyclo-
octane
18.0 R'N- Solution
(N0 )R 2in di- 2

butyl-
phtha-
late
18.1 R' = C H Solution 3 M M * 200- 40.4 13.7 1.3 [5.55]
in di- 230
R = C H C - butyl-
2
2

H C O O C H phtha-
2 3

late
18.2 R'=CH Solution 3 M M * 190- 41.9 14.4 1.2 [5.57]
indi- 230
R =CH
2
butyl- 2

-CH -CN 2phtha-


Iate
18.3 R' = C H Solution 3 M M * 190- 40.9 13.9 1.0 [5.55]
indi- 230
R = C H 0 - butyl-
2
2

CH 3 phtha-
late
18.4 R'=CH Solution 3 M M * 200- 40.8 14.0 1.5 [5.55]
in di- 230
R = C H C - butyl-
2
2

H COOH
2 phtha-
late
18.5 R'=CH Solution 3 M M * 180- 40.1 14.7 16.0 [5.56]
in di- 220
R = C H C - butyl-
2
2

H COOH
2 phtha-
late
18.6 R'=CH Solution 2 M M * 160- 42.9 17.3 2.7 [5.57]
-CH CN 2indi- 180
butyl-
R =FcCH **phtha-
2
2

late
18.7 R ' = C H C - Solution 2 M M * 180- 36.8 12.9 0.01 [5.57]
H C-
2 in di- 200
-(N0 ) - butyl-
2 2

-CH 3phtha-
R=
2
late
FcCH ** 2

18.8 R = R ^Solution
1 2
M M * 210- 41.3 13.9 0.1 [5.55]
86 Secondary Nitramines

CH CH 2 r in di- 250
-CN butyl-
phtha-
late
18.9 Solution M M * 200- 40.8 14.1 1.6 [5.55]
CH CH -
2 2 in di- 240
COOH butyl-
phtha-
late
18.10 R'=R = 2
Solution MM* 190- 38.5 13.5 0.05 [5.57]
FcCH 2 in di- 200
butyl-
phtha-
late
18.11 R'=R = 2
Solution M M * 210- 40.7 13.7 0.8 [5.57]
CH CH -
2 2 in di- 250
COOCH 3
butyl-
phtha-
late
18.12 R'=R = 2
Solution MM* 145- 39.7 15.7 230.0 [5.58]
CH C-
2 in meta- 170
-(N0 ) - 2 2 dinitro-
-CH 2 benzene
19.1 Tetryl Melt MM* 211- 38.4 15.4 460.0 [5.59]
260
19.2 Tetryl Melt TGA* 132- 34.9 12.9 59.0 [5.60]
164
19.3 Tetryl Melt IC* 131- 36.0 13.8 148.0 [5.61]
155
19.4 Tetryl Melt M M * 140- 35.2 13.5 174.0 [5.59]
160
19.5 Tetryl Melt M M * 140- 40.0 16.0 330.0 [5.62]
165
19.6 Tetryl Melt M M * \50- 35.9 13.5 166.0 [5.55]
175
19.7 Tetryl Melt M M * 130- 60.9 27.3 10 3
[5.63]
139
19.8 Tetryl Melt M M * 140- 55.5 14.5 10 3
[5.64]
150
19.9 Tetryl Melt M M * 150- 35.0 - - [5.64]
175
19.10 Tetryl Melt DSC* - 54.9 - - [5.65]
19.11 Tetryl Melt DSC* 170- 58.0 - - [5.66]
174
19.12 Tetryl Solution DSC* - 33.5 - - [5.34]
in poly-
phenyl-
ether
19.13 Tetryl Solid DSC* - 52.0 12.5 10 3
[5.63]
19.14 Tetryl Solid DSC* - 36.6 12.7 62 [5.63]
19.15 Tetryl Solid DSC* - 54.9 - - [5.34]
Thermal Decomposition and Combustion 87

20.0 o \ /r
2
2
Solution
° YV °
2 N 2 i n m e t a
-
dinitro-

r
1
benzene
20.1 R = OCH
1
Solution 3 M M * 145- 32.1 12.4 323.0 [5.55]
R = CH
2
in meta- 3 165
dinitro-
benzene
20.2 R = CH
1
Solution 3 M M * 150- 34.2 13.4 416.0 [5.55]
R = CH
2
in meta- 3 170
dinitro-
benzene
20.3 R' = C1 Solution M M * 145- 33.7 13.1 346.0 [5.55]
R = CH
2
in meta- 3 165
dinitro-
benzene
20.4 R = Br
1
Solution M M * 155- 33.8 13.4 616.0 [5.55]
R = CH
2
in meta- 3 175
dinitro-
benzene
20.5 R'=N0 Solution 2 M M * 144- 33.3 13.9 3330.0 [5.55]
R =C H
2
in meta-
2 5 165
dinitro-
benzene
20.6 R'=N0 Solution 2 M M * 135- 32.7 13.3 1548.0 [5.55]
R = A?-C H in
2
meta- 3 7 155
dinitro-
benzene
20.7 R'=N0 Solution 2 M M * 130- 33.5 14.2 4786.0 [5.55]
R = /-C H in
2
meta-3 7 150
dinitro-
benzene
21.1 1,3,3- Solution LC* 189- 43.6 16.4 20.0 [5.13]
Trinitro- in acetone 269
azetidine
21.2 1,3,3- Melt L C * 220- 43.7 16.3 12.0 [5.13]
Trinitro- 250
azetidine
22.1 Hexanitro- Solution LC* 146- 42.4 17.6 100.0 [5.13]
hexa-aza- in acetone 226
isowurtzi-
tane
* M M is the manometric method, L C is the liquid chromatography method, IC is the
isothermal calorimetry method, 1M is the ignition method, D S C is the differential
scanning calorimetry method, T G A is the termogravimetric analysis method, G L C is
the gas-liquid chromatography method, D T A is the differential thermal analysis
method
88 Secondary Nitramines

Nitramines, in particular cyclic nitarmines with high melting


temperatures, and especially octogen, which has three modifications with
various decomposition rates, are convenient models to study the general
regularities of monomolecular reaction in the solid phase, both
homogeneous and localized ones.
In a number o f works, the decomposition o f R D X as a commercially
important explosive has been studied under the influence o f the following
physical factors: shock, electric field, and vibration. The pressure influence
on the reaction in the solid phase is o f importance for kinetics.
The decomposition o f J3-UMX was studied in [5.67] by the FTIR
method at pressures o f 3.6-6.5 G P a and temperatures o f 2 5 5 - 3 1 0 ° C . A s
follows from the rate law, up to 5.5 G P a the reaction proceeds according to
the monomolecular mechanism (the first order), and above 5.5 G P a —
according to the bimolecular mechanism.
The pressure inhibits the decomposition, and the activation volume has
the same constant value o f 4.1 c m / m o l at all temperatures. A t the same
3

time, the activation energy decreases as the pressure increases from 120
kcal/mol at 3.6 G P a to 37.5 kcal/mol at 6.5 G P a . Thus, the entropy o f
activation varies from 0.6 to 0.047 e. u. ( 2 9 0 ° C ) .
Apparently, the abnormally large value o f £ a c t j at a pressure o f 3.6 G P a
v

is connected with the effect o f substance submelting, which disappears at


higher pressures. In [5.68], the variation o f £ a c t i v with pressure growth up to
5 G P a was not observed.
The essential factor is the change o f the mechanism at P > 5.5 G P a . In
[5.67], it was proposed that the most probable reason for it was the reaction
o f H M X with the liquated nitric oxides and other products. A t P > 5.5 G P a ,
the value o f AG n
o f the reaction tends to zero and this fact explains the
reason for H M X explosion at high pressures.
If one alkyl group in the molecule o f secondary nitramines is replaced
by an electronegative substitute, for instance, F or N 0 , a considerable 2

reduction o f the N — N 0 2 bond strength takes place and the compounds


become unstable.
Thus, N-nitro-N-fluoroamines and R N F N 0 2 rapidly decompose at
temperatures o f 5 0 - 8 0 ° C ; they possess low enthalpies (23-30 kcal/mol) and
positive entropies o f activation (1-10 e. u.) [5.69, 5.70]. The decomposition
mechanism o f these compounds has not been determined.
The decomposition o f N,N-dinitramines R N ( N 0 ) was studied in more 2 2

detail [5.71, 5.72]. For these compounds, the rates o f monomolecular


reactions in the gas phase were determined (Table 5.3.2).
In addition, for dinitromethylnitramine, a complete quantitative analysis
o f the products was carried out, and 1 5
N taggant was found to be distributed
in the products. Initially the taggant was located in the N 0 - g r o u p o f the 2

parent molecule [5.72].


Thermal Decomposition and Combustion 89

Table 5.3.2. The rate constants and activation parameters of thermal decomposition
ofRN(N0 ) 2 2 [5.71]

R k • 10 , s~ 5 !
E, kcal/mol log/t (s )
_1

50°C 65°C 85°C

CH 3 2.5 17.0 190.0 28.2 14.8

C H
2 5 3.5 26.0 330.0 29.6 15.6

«-C H 3 7 2.8 23.0 280.0 30.9 16.4

n-C H
4 9 3.5 20.0 270.0 29.5 15.5

A t early decomposition stages, the following reactions proceed:

CH N(N0 )3 2 -CH3NNO2 + N 0 2

CH N N 0 3 2 ^ CH N: + N 0
3 2

CH N: 3 + CH N(N0 ) 3 2 (CH ) NN(N0 ) 3 2 2

(CH ) NN(N0 ) 3 2 2 ^ 2CH 3 +N 2 + 2N0 . 2

If the degree o f transformation is less than 2%, N 2 is formed from amine


nitrogen. It is particularly remarkable that the nitramine radical C H N N 0 3 2

is inactive. It does not react with the molecules o f the parent substance and
is not subjected to the intramolecular oxidation

CH N'N0 3 2 - C H 3 O + N2O.

The fastest channel o f its consumption is the decomposition with the


formation o f methylnitrene. A s N 0 is accumulated, the following oxidation 2

reactions prevail:

CH3NNO2+NO2 -CH N(N0 )ONO 3 2

CH N(N0 )ONO 3 2 " N Q


- CH N(N0 )6 3 2

CH N(N0 )6 3 2 CH3NO+NO2

CH N: 3 +N0 2 ^ C H 3 N O + N O .

In contrast to the covalent compounds R N ( N 0 ) , the ionic salts 2 2

M N ( N 0 ) " are stable, at least, in the solid state [5.73].


+
2 2
90 Secondary Nitramines

REFERENCES

5.1. P. H . Stewart, J. B . Jeffries, J . - M . Zellweger, D . F . M c M i l l a n , D . M .


Golden, J. Phys. Chem., 93: 3557-3563 (1989).
5.2. Y . G . Lazaron, P. Papagiannakopouls, J. Phys. Chem., 94: 7114—
7119(1990).
5.3. L . Y u e , T . Zehna, X . Henning, Proc. XVII Int. Pyrotechnics
Seminar Combined with the II Beijing Int. Symp. on Pyrotechn. and
Explosives, Beijing, China, October 2 8 - 3 1 , 1: 466-472 (1991).
5.4. R. Shaw, F. E . Walker, J. Phys. Chem., 81: 2572-2576 (1977).
5.5. M . Farber, R . D . Srivastava, Chem. Phys. Letters, 64: 3 0 7 - 3 1 0
(1979).
5.6. M . Farber, R . D . Srivastava, Chem. Phys. Letters, 80: 345-349
(1981).
5.7. V . R. Stepanov, A . A . Fedotov, A . N . Pavlov, G . M . N a z i n , Kinetics
of Chemical Reactions, Chernogolovka, M o s c o w region: Joint
Institute o f Chemical Physiscs, Academy o f Sciences o f the U S S R ,
9 7 - 1 0 0 (1980).
5.8. J. J. Batten, Austr. J. Chem., 24: 945-954 (1971).
5.9. J . C . Hoffsommer, D . J. Glover, Comb, and Flame, 59: 303-310
(1985).
5.10. S. A . Shackelford, M . B . Coolidge, B . B . Goshgarian, B . A . L o v i n g ,
R. N . Rogers, J. L . Janney, M . H . Elinger, J. Phys. Chem., 89:
3118-3126 (1985).
5.11. M . Schroeder, XVI JANNAF Combustion Meeting, C P I A Pub.,
308(11): 17-34 (1979). XVII JANNAF Combustion Meeting, CPIA
Pub., 310(11): 395-413 (1981).
5.12. J. C . Oxley, M . Hiskey, D . N a u d , R. Szekeres, J. Phys. Chem., 96:
2505-2509 (1992).
5.13. J. C . Oxley, A . B . K o o h , R. Szekeres, W . Zheng, J. Phys. Chem.,
98: 7004-7008 (1994).
5.14. J. M . Flournoy, J . Chem. Phys.,36: 1106-1107 (1962).
5.15. Jr. R. Behrens,J. Phys. Chem., 95: 5838-5845 (1994).
5.16. T. L. Boggs, Progress in Astronautics and Aeronautics:
Fundamentals of Solid-Prop ell ant Combustion, K . Kuo, M .
Summerfeld Eds., N e w Y o r k , 90: 121-175 (1984).
5.17. C . F. Melius, J. de Physique. Colloque C4, Supplement aun 9, 48:
341-352 (1987).
5.18. Y u . M . Burov, G . M . N a z i n , Kinet. Katai, 23: 12-17 (1982).
5.19. R. S. Stepanov, V . N . Shan'ko, I. P. Medvetskaya, L . A . Kuznetsov,
In: Physical Chemistry, Krasnoyarsk, 2: 3-5 (1965).
5.20. F. I. Dubovitskii, B . L . Korsunskii, Usp. Khim., 1: 1828-1871
(1981).
Thermal Decomposition and Combustion 91

5.21. T. B . B r i l l , Comb, and Flame, 62: 213-214 (1985).


5.22. B . K . Laptenkov, V . P. Borisov, Y u . M . Grigor'ev, P. N . Stolyarov,
Chemical Physics of Combustion and Explosion. Kinetics of
Chemical Reactions, 107-110 (1989).
5.23. B . L . Korsunskii, F. I. Dubovitskii, Dokl. Akad. Nauk SSSR, 155:
4 0 2 ^ 0 5 (1964).
5.24. B . L . Korsunskii, F. I. Dubovitskii, G . V . Sitonina, Dokl. Akad.
Nauk SSSR, 174: 1126-1128 (1967).
5.25. B . L . Korsunskii, F. I. Dubovitskii, E . A . Shurygin, Izv. Akad. Nauk
SSSR, Ser. Khim., 7: 1452-1455 (1967).
5.26. G . V . Sitonina, B . L . Korsunskii, N . F. Pyatakov, V . G . Shvaiko, I.
Sh. Abdrakhmanov, F. I. Dubovitskii, Izv. Akad. Nauk SSSR, Ser.
A7z//w.,2:311-314 (1979).
5.27. B . A . Lur'e, V . N . Ivakhov, Chemical Physics of Condensed
Explosive Systems, M . : M k h T I , 104: 12-19 (1979).
5.28. H . Ronqru, Y . Zhengonan, L . Yanjun, Thermochim. Acta., 123:
135-151 (1988).
5.29. A . J. B . Robertson, Trans. Faraday. Soc, 45: 85-93 (1949).
5.30. B . S. Belyaeva, G . K . K l i m e n k o , L . T. Babaitseva, P. N . Stolyarov,
Chemical Physics of Combustion and Explosion. Kinetics of
Chemical Reactions, 4 7 - 5 0 (1977).
5.31. Y u . Y a . M a k s i m o v , V . N . A p a l k o v a , O. V . Broverman, A . I.
Solov'ev, Zh. Fiz. Khim., 59: 342-345 (1985).
5.32. Y u . Y a . M a k s i m o v , Theory of Explosives, M . : Vysshaya Shkola,
M K h T I , 53:73-84 (1967).
5.33. J. W i l d y , Symp. on Chemical Problems Connected with the Stability
of Explosives, Stookholm, 51-65 (1967).
5.34. R. N . Rodgers, E . D . Morris, J. Analyt. Chem., 38: 412-414 (1966).
5.35. F. C . Rauch, A . J. Fanelli, J. Phys. Chem., 73: 1604-1610 (1969).
5.36. P. G . H a l l , Trans. Faraday Soc, 67: 556-565 (1971).
5.37. R. N . G . Rodgers, G . W . Daub, Analyt. Chem., 45: 596-600 (1973).
5.38. K . Kishore, Analyt. Chem., 50: 1079-1083 (1978).
5.39. R. N . Rodgers, L . S. Smith, Thermochim. Acta., 1: 1-8 (1970).
5.40. Y . O y u m i , Prop. Explos. Pyrotechn., 13: 41-47 (1988).
5.41. K . K . Andreev, Thermal Decomposition and Combustion of
Explosives, M . : Nauka, 67-71 (1966).
5.42. Y u . M . Burov, G . M . N a z i n , Khim. Fiz., 3: 1126-1130 (1984).
5.43. S. B o n , Solid State Chemistry, V . M . Garner E d . , M . : Izd.
Inostrannoi Literatury, 335-353 (1961).
5.44. J. J. Batten, J. Chem., 23: 749-755 (1970).
5.45. J. J. Batten, J. Chem., 24: 2025-2029 (1971).
5.46. M . A . Mel'nikov, V . V . N i k i t i n , Fiz. Goreniya Vzryva, 8: 591-593
(1972).
5.47. G . K . Klimenko, Combustion and Explosion, M . : Nauka, 785-788
(1972).
92 Secondary Nitramines

5.48. R . N . Rodgers, Thermochim. Acta., 3: 4 3 7 ^ 4 7 (1972).


5.49. J. N . M a y c o c k , V . R. Pai Verneker, Explosivstoffe, 17: 5-9 (1969).
5.50. A . I. Medvedev, G . V . Sakovich, V . V . Konstantinov, Conf. on
Kinetics and Chemical Reaction Mechanisms in Solids: Abstracts of
papers, Novosibirsk: Institute o f Physical-Chemical Basis o f
Processing o f M i n e r a l R a w Materials, A c a d e m y o f Sciences o f the
U S S R , Siberian D i v i s i o n , 1: 163-165 (1977).
5.51. T. B . B r i l l , R. J. K a r p o w i c h , J. Phys. Chem., 86: 4260-4265 (1982).
5.52. Y u . M . Burov, G . B . Manelis, G . M . N a z i n , Khim. Fiz., 4: 956-962
(1985).
5.53. S. Zeman, M . D i m u n , Prop. Explos. Pyrotechn., 15: 217-221
(1990).
5.54. S. Zeman, M . D i m u n , S. Truchlik, Thermochim. Acta., 78: 181-184
(1974).
5.55. R. S. Stepanov, V . N . Shan'ko, I. P. Medvetskaya, V . M .
Gorodetskaya, Chemical Physics of Combustion and Explosion.
Kinetics of Chemical Reactions, Chernogolovka, M o s c o w region:
Joint Institute o f Chemical Physiscs, Academy o f Sciences o f the
U S S R , 56-58 (1977).
5.56. V. N. Shan'ko, R. S. Stepanov, In: Physical Chemistry,
Krasnoyarsk, 1: 190-197 (1974).
5.57. M . A . Stepanova, V . P. Tverdokhlebova, R. S. Stepanov, I. V .
Tselinskii, B . V . Gidaspov, Zh. Org. Khim., 13: 1364-1368 (1977).
5.58. B . L . Korsunskii, L . Y a . Kiseleva, V . I. Ramushev, F. I.
Dubovitskii, Izv. Akad. Nauk SSSR, Ser. Khim., 1778-1781 (1974).
5.59. A . J. B . Robertson, A . Yoffe, Nature, 161: 806-807 (1948).
5.60. M . A . Cook, M . T. Abegg, Industr. Eng. Chem., 48: 1090-1095
(1956).
5.61. F. I. Dubovitskii, Y u . I. Rubtsov, G . B . Manelis, Izv. Akad. Nauk
SSSR, Ser. Khim., 10: 1763-1766 (1960).
5.62. F. I. Dubovitskii, G . B . Manelis, L . P. Smirnov, Zh. Fiz. Khim., 35:
521-529(1961).
5.63. R. C . J. Farmer, J. Chem. Soc, 117: 1603-1606 (1920).
5.64. C . V . Hinshelwood, J. Chem. Soc, 119-120: 721-726 (1921).
5.65. M . A . Cook, The Science of High Explosives, N Y . : Reinhold;
London: Chapman and H a l l , Chapter 8, 440p. (1958).
5.66. R. N . Rodgers, L . S. Smith, Analyt. Chem., 39: 1024-1025 (1967).
5.67. G . J. Piermarini, S. B l o k , P. J. M i l l e r , J. Phys. Chem., 81: 3 8 7 2 -
3878 (1987).
5.68. E . R. Lee, R. H . Sandborn, H . D . Stromberg, Proc. VSymp. Int. on
Detonation, Washington: U S Gover. Print. Office, 331-333 (1972).
5.69. M . Graff, C . Gotzmer, Jr., W . E . M c Q u i t o n , J. Org. Chem., 32:
3827-3830 (1967).
5.70. M . Graff, C . Gotzmer, Jr., W . E . M c Q u i t o n , J. Chem. Eng. Data,
34: 513-514 (1969).
Thermal Decomposition and Combustion 93

5.71. B . L . Korsunskii, G . V . Sitonina, B . S. Federov, et a l , Izv. Akad.


Nauk, Ser. Khim., 4: 790-793 (1989).
5.72. V . V . Charskii, A . N . Pavlov, G . M . N a z i n , et al., Izv. Akad. Nauk,
Ser. Khim., 4: 794-797 (1989).
5.73. O . A . L u k ' y a n o v , V . P. Gorelik, V . A . Tartakovskii, Izv. Akad.
Nauk, Ser. Khim., 1: 94-97 (1994).
Chapter 6

ORGANIC AZIDES

Numerous studies were devoted to the study o f thermal transformation o f


organic azide compounds. These studies go back to the work by T. Curtius
[6.1]. Most papers on the subject were considered by G . L ' A b b e [6.2, 6.3].
A t present, a generalized description o f thermal decomposition o f azides is
possible, though particular details o f these reactions have not been revealed
until now.
The decomposition o f organic azides RN 3 in the overwhelming
majority o f cases proceeds by the RN—N 2 bond breaking with the
formation o f a singlet nitrene ( R N : ) and nitrogen molecule in the ground
state ( X Z ) .
! +
0 Sometimes, this process proceeds simultaneously with the
skeleton rearrangement o f the R N group. The other reactions, for instance
the N H elimination
3

CH3CH2N3 C H 2 = C H + NH3
2 ?

are not observed. The limiting stage o f the R N — N 2 bond breaking can be
classified as elimination, but not radical decomposition, because, in this
reaction, nitrogen molecule N 2 and one double valence-deficiency particle
nitrene are formed.
In the case o f R N 3 molecule and allyl anion,

(CH —
2 CH — CH )\2

at the initial state, the three-centered three-level ^--system, formed as a


result o f the overlapping o f three /?-orbitals, are populated by two pairs o f
electrons.

•4k

95
96 Organic Azides

One pair o f ^-electrons populates a low-energy level and provides a


binding interaction between all three atoms o f the - N fragment. Another
3

pair o f ^-electrons occupies a nonbinding weakly antibonding level.


The R N — N 2 bond breaking proceeds by a "heterolytic" type, i.e., a
new unshared electron pair o f the nitrogen molecule is formed not from one
electron o f the cr-bond and one /?-electron o f dipolarophilic - N group, but3

from the bonding pair o f cr-electrons localized between R N - and - N N


groups.

R N ( 5 / ) + N2OS0).

Exactly this process is allowed by the rules o f the spin and orbital symmetry
conservation and results in the formation o f nitrene in the singlet state S\.

6.1 Aliphatic Azides

A z i d e s o f aliphatic type are the most thermally stable compounds among


organic azides. The kinetic data on the decomposition o f these compounds
are presented in Table 6.1.1. The structure o f the alkyl group only weakly
influences the decomposition rate, and the main type o f influence is the
induction one, w h i c h , for example, can clearly be seen in the series o f
compounds 13 in Table 6.1.1. The H N — N 2 bond strength in hydrazoic acid
(46 kcal/mol [6.11]) is 5 kcal/mol larger than the Z ) ( C H N — N ) value. This
3 2

indicates that the positive induction effect o f the alkyl group stabilizes
nitrene, and, consequently, the transition state o f the reaction.
Singlet nitrenes, formed at the first decomposition stage, are subjected
to fast transformations. For instance, as was shown in 1933 by J . A .
Leermakers in [6.7], they are subjected to a very fast isomerization in
corresponding imines:

slowly rapidly
CH3N3 ^ CH N:
3 -J
HN==CH . 2
Thermal Decomposition and Combustion 97

Table 6.1.1. The kinetic characteristics of decomposition of aliphatic azides

Nos. Compound Medium T,°C E, \ogA Refs.


kcal/ (s" )1

mol

1.1 CH3N3 Gas 150-200 40.5 14.45 [6.4]


1.2 CH3N3 Gas 200-240 43.5 15.58 [6.5]
2.1 C H N
2 5 3 Gas 150-200 40.1 14.50 [6.6]
2.2 C H N
2 5 3
Gas 200-240 39.7 14.30 [6.7]
3. n-PrN 3 Gas 150-200 39.4 14.18 [6.6]
4. iso-?vN 3 Gas 150-200 38.5 13.86 [6.6]
5.1 cyc/oC H N 6 n 3 Gas 170-220 39.1 13.70 [6.8]
5.2 cyc/o-C H„N 6 3
Ethyl- 111-139 47.8 17.87 [6-9]
benzene
6.1 PhCH N 2 3 Gas 180-215 38.0 13.85 see*
6.2 PhCH N 2 3 Gas 160-200 39.0 14.34 [6.8]
7. N CH C=N—NCH
3 2 3
Gas 160-200 39.2 13.90 see*

N N

N CH C
3 2 N CH 3 Gas 150-210 37.3 13.41 see

N — N = N

I \
N3CH2C-C-CH2N3 Gas 170-230 36.8 13.90 see

N - O - N

10. CH COCH N 3 2 3 Gas 170-210 38.7 14.51 [6.8]


11. C H COOCH N
2 5 2 3 Gas 180-210 38.9 14.17 [6.8]
12. 4-N0 C H CH N 2 6 4 2 3 Gas 160-200 39.0 14.34 [6.8]
13. 4-RC H CPh CH N 6 4 2 2 3

13.1 R=H Nitro­ 169-190 31.7 11.8 [6.10]


benzene
13.2 R=H Hexa- 169-190 34.3 - [6.10]
decalin
13.3 R=H Dibutyl- 169-190 32.0 10.7 [6.10]
carbinol
13.4 R = C1 Dibutyl- 169-190 38.8 - [6.10]
carbinol
13.5 R=N0 2
Dibutyl- 34.3 - [6.10]
carbinol
13.6 R = CH 3
Dibutyl- 29.0 9.3 [6.10]
carbinol
98 Organic Azides

13.7 R = OCH 3 Dibutyl- 28.9 9.4 [6.10]


carbinol

13.8 R = N(CH ) 3 3 Dibutyl- 25.4 7.9 [6.10]


carbinol

* The authors' data

During the high-temperature pyrolysis ( 9 0 0 ° C ) o f methylazide, the


following components were obtained [6.12]: H C N , H N , and a solid 3

residual. The H N formation can be explained by the following reactions:


3

H N 3

CH N: + H - C H N
3 2 3 — C H N — C H - ^ HN3 + C H N = C H .
3 2 3 2

During decomposition o f benzylazide P h C H N 2 3 in inert solvents, the


following compounds were observed: 1M , H N , triphenylamine ( T P h A ) , 2 3

benzyl-benzylidenamine (BBA), dibenzylbenzamidine (DBA), 3,5-


diphenyl-4-benzyl-l,2,4-tetrazol ( D P h B T ) , and tetraphenylpyrazine (TPhP).
The formation o f the majority o f these products can be explained by the
reactions typical o f all azides: the reaction o f 1,3-dipolar cyclo-attachment
o f the parent benzylazide with the unsaturated compounds

C H CH N + C H C H = N H — - P h C H = N H — HN + PhCH= N - C H P h
6 5 2 3 6 5 3 2

: i (BBA)
PhH C 2 N 3

-N, H
\
/N-CH Ph 2

PhCH N +BBA—PhCH N
2 3 2 CHPh-^ CT I ^
Ph^ \N-CH Ph 2

N=N-N-CH^Ph

-PhCH N=C(Ph)-NHCH Ph (DBA)


2 2

-PhCH=N-N(CH Ph) 2 2-

The latter product reacts with D B A with the formation o f T P h A and


DPhBT.
In view o f the fact that aliphatic azides are stable compounds, which are
as stable as nitramines and nitroesters, they can successfully be employed as
explosives or components o f solid rocket propellants. A glycidylazide
polymer
H-(-OCH CH-)-OH 2

CH N 2 3
Thermal Decomposition and Combustion 99

has been considered as a binder for solid rocket propellants [6.13, 6.15].
Various modifications o f this polymer have a satisfactory stability (at 100°C
for 24 h the degree o f transformation is a fraction o f a percent). The
polymer is w e l l compatible with H M X and aluminum and exceeds
nitroesters and nitrocompounds in its compatibility [6.15].

6.2 Aromatic and Heterocyclic Azides


Aromatic azides are less stable in the reactions of N 2 detachment as
compared with aliphatic azides because o f stabilization o f nitrenes by a n-
system o f the ring. Comparison o f the difference o f bond energies in pairs
CH -N0 /PhCH -N0
3 2 2 2 (13 kcal/mol), C H - N F / P h C H - N F
3 2 2 2 (9 kcal/mol),
CH N-N /PhN-N
3 2 2 (6.6 kcal/mol), and Me N-N0 /Ph(CH )N-N0
2 2 3 2 (5
kcal/mol) gives an insight into the degree o f resonance stabilization by a
phenyl group o f a free electron on the C and N atoms bounded with the
ring. In the aromatic series, as well as in the aliphatic one, the rates only
slightly depend on the structure o f substitutes. The difference in the
decomposition rates between the most stable and the least stable A r N 3 at
150°C hardly attains a factor o f 5 [6.16-6.18]. Typical parameters o f the
Arrhenius equation for A r N 3 are as follows: E = 32-35 kcal/mol and A =
10 -10, 3 1 5
s" [6.16-6.18].
1

Solvents also exert only a weak influence on the decomposition rate.


In spite o f the ring stabilization by the ^-system, the primary
decomposition products and aromatic singlet nitrenes are highly reactive
compounds. The main pathway o f their transformation is the insertion into
the C — H and C — C bonds [6.17]. A good example is the reaction o f
intramolecular insertion o f nitrene in the C — H bond during decomposition
o f <9r//?o-azidobiphenyls, resulting in the formation o f carbazole with a high
yield

C^O €H2> - €HP>


N 3 J
N H
:

In this case, the decomposition rates o f azidobiphenyls are the same as in


the case o f phenylazides.
Aromatic azides with or//zo-substitutes containing a double bond
produce cyclic products during decomposition and have high decomposition
rates (Table 6.2.1).
The low values o f the pre-exponential factor are in agreement with the
synchronous reaction mechanism: the N detachment occurs simultaneously2

with the cyclic fragment closure. A c c o r d i n g to N . J . Dickson [6.23], "a


100 Organic Azides

motive force o f the process is the energy o f heterocycle d e r e a l i z a t i o n ,


w h i c h is partly formed in the transition state".

Table 6.2.1. The kinetic characteristics of decomposition of or//?o-substituted


arylazides
Nos. Reaction £, kcal/ log/1 Refs.
mol (s )
_1

12.8 [6.18-
6.20]

\- [6.18,
) 6.21]

11.9 [6.22]

However, it is probable that the nitrene mechanism enhanced by a


stabilizing interaction o f the vacant orbital o f nitrene with the unshared
electron pair o f the nitrogen atoms o f the azo-group or oxygen atoms o f the
nitro- or carboxyl group exists, at least, as a partly realizable channel.
6W/zo-N0 -substituted
2 arylazides have promising energetic
characteristics but cannot be used because o f the low stability.
Heterocyclic azides essentially differ from aromatic azides. The
reactions o f their decomposition are presented in Table 6.2.2.
Triazinylazides are more stable than A r N 3 and are close to aliphatic
azides because o f their stability. This heterocycle ranks below aromatic
systems in the resonant stabilization o f nitrenes.
The stability o f furazanylazides to N 2 loss is close to the stability o f
aromatic azides. On the contrary, isoxalylazides possess a higher
decomposition rate than A r N . Their reduced stability can be connected not
3
Thermal Decomposition and Combustion 101

with the decomposition o f the azide-group, but with the disintegration o f the
heterocycle by the synchronous mechanism.

Table 6.2.2. The kinetic parameters of thermal decomposition of compounds


containing N -group connected with the sp -carbon atom of the heterocycle
3
2

Compound State T,°C E, kcal/ \ogA Refs.


mol (s" )!

N C — N —
3 C — R 1

N =C(R )N 2

R = R = CI
1 2
Decalin 184.3 35.7 1
13.5 [6.24]
R = R = OCH
1 2
3 Decalin 174-196 36.7 13.77 [6.24]
R = R = NMe
1 2
2 Decalin 184.3 35.3 1
13.5 [6.24]
R =R =N
1 2
3 Decalin 150-190 39.5 15.40 see 2

R'=N , R = OCH CH N
3
2
2 2 3 Decalin 80-140 37.7 14.25 see 2

R (j=NOC(R )=CN
1 2
3

R = H, R = OMe
1 2
Decalin 60-90 25.5 12.21 [6.25]
R = H, R = CMe
1 2
3 Decalin 60-90 26.6 12.56 [6.25]
R = R = Me
1 2
Decalin 80-90 26.3 13.35 [6.25]
N (j:=NON=( :R
3 f

R = NH 2 Gas 130-180 34.4 14.10 see 2

R-COOH Gas 130-180 34.0 14.00 see 2

1
Calculated under the assumption that log4 = 13.5 (s
2
Unpublished data of the authors.

For certain heterocyclic azides, the rearrangement o f azide-tetrazole is


typical [6.26]. In the majority o f cases, this rearrangement proceeds rather
rapidly, and N 2 detachment occurs against the background of the
equilibrium isomerization

tetrazole ^==^ azide


-1

azide products.

The barrier o f reaction (1) is 20-25 kcal/mol [6.27, 6.28]. In this case, the
observable rate constant takes the form

*observ = , = * 2 * 0 + *) ,
102 Organic Azides

where K=kfk . A At K » 1, we w i l l have £ bserv


0
=
k\ at K « 1 we w i l l have
^observ ^ 2
=
k-l-< < C

A z i d e compounds o f other types (benzoyl-, v i n y l - , and formyl-azides),


whose decomposition is widely covered in the literature, do not possess the
explosive properties. Their decomposition occurs according to the nitrene
mechanism and, as a rule, is accompanied by a skeleton rearrangement o f
the R N fragment [6.2, 6.3, 6.29-6.31]. The decomposition rates o f all these
azides are nearly the same. The typical values o f activation energy and the
pre-exponential factor for these compounds are 26-28 kcal/mol and A =
10 12
-IO 1 4
s , correspondingly.
-1

REFERENCES

6.1. T. Curtius, G . Ehrhart, Berichte, Bd. 55: 1559-1568 (1922).


6.2. Azides and Nitrenes: Reactivity and Utility, E . F. V . Scriven E d . , N .
Y . : A c a d e m i c Press, 542 p. (1984).
6.3. G . L ' A b b e , Chem. Rev., 69: 345-363 (1969).
6.4. M . S. O ' D e l l , B . Darwent, Canad. J. Chem., 48: 1140-1147 (1970).
6.5. J. A . Leermakers, J. Amer. Chem. Soc, 55 : 3098-3105 (1933).
6.6. G . Geiseler, W . K o n i g , Zeitschrift fur Physikalishe Chemie Leipzig,
Bd. 227: 81-92 (1964).
6.7. J. A . Leermakers, J. Amer. Chem. Soc, 55 : 2719-2729 (1933).
6.8. R. S. Stepanov, L . A . Kruglyakova, E . S. Buka, Kinet. Katai, 27:
4 7 9 - 4 8 2 (1986).
6.9. P. Walter, W . A . Waters, J. Chem. Soc, 1632-1638 (1962).
6.10. W . H . Saunders, J. C . Ware, J. Amer. Chem. Soc, 80: 3328-3332
(1958).
6.11. I. S. Zaslonko, S. M . Kogarko, E . V . M o z z h y k h i n , Kinet. Katai, 13:
829-835 (1972).
6.12. F. O . Rice, G . J. Greleski, J. Phys. Chem., 61: 830-831 (1957).
6.13. N . Kubota, Prop. Explos. Pyrotechn., 13: 172-177 (1988).
6.14. L . Changqing, W . Ping, H . Y u e , Proc. XVII Int. Pyrotechnics
Seminar Combined with II Beijing Int. Symp. on Pyrotechnics and
Explosives, Beijing, 1: 451^159 (1991).
6.15. Z . X i a o y i , C h . Xuening, Proc. XVII Int. Pyrotechnics Seminar
Combined with II Beijing Int. Symp. on Pyrotechnics and Explosives,
Beijing, 1: 542-546 (1991).
6.16. K . E . Russel, J. Amer. Chem. Soc, 77: 3487-3488 (1955).
6.17. R. A . Smith, J. H . H a l l , J. Amer. Chem. Soc, 84: 480^185 (1962).
6.18. L . K . D y a l l , J. E . K e m p , J. Chem. Soc B., 976-979 (1968).
6.19. J. H . H a l l , F. E . Behr, R. L . Reed, J. Amer. Chem. Soc, 94: 4 9 5 2 -
4958 (1972).
6.20. L . K . D y a l l , Austral. J. Chem., 30: 2669-2678 (1977).
6.21. L . K . D y a l l , Austral. J. Chem., 28: 2147-2159 (1975).
Thermal Decomposition and Combustion 103

6.22. J. H . H a l l , F. W . Dolan, J. Org. Chem., 43: 4608-4612 (1978).


6.23. N . J. Dickson, L . K . D y a l l , Austral. J. Chem., 33: 9 1 - 9 9 (1980).
6.24. R. Kayama, S. Hasanuma, S. Sekiguchi, K . Matsui, Bull. Chem. Soc.
Japan, 47: 2825-2830 (1974).
6.25. G . L ' A b b e , F. Godts, Bull. Soc. Chim. Belg, 96: 229-236 (1987).
6.26. A . Konnecke, R. Dorre, E . Kleinpeter, E . Lippman, Tetrahedron., 35:
1957-1963 (1979).
6.27. B . Stanovnik, M . Tisler, Tetrahedron., 25: 3313-3320 (1969).
6.28. B . Stanovnik, M . Tisler, B . Stefanov, J. Org. Chem., 24: 3812-3813
(1971).
6.29. G . L ' A b b e , G . Mathys, J. Org. Chem., 39: 1778-1780 (1974).
6.30. G . Smolinsky, C . A . Pryde, J. Org. Chem., 33: 2411-2416 (1968).
6.31. D . S. Breslow, T. J. Prosser, A . F. Marcantono, C . A . Genge, J.
Amer. Chem. Soc, 89: 2384-2390 (1967).
Chapter 7

ORGANIC DIFLUOROAMINO-
COMPOUNDS

Polyfunctional difluoroamino-compounds, containing several N F - g r o u p s at 2

one carbon atom with the exception o f the terminal - C H ( N F ) 2 2 group, i.e.,
without the H atom in a-position with respect to the N F - g r o u p , have the 2

simplest kinetic regularities and reaction mechanism. These substances


decompose both in the gas phase and in the condensed states through the
C—NF 2 bond breaking.
The known kinetic data are presented in Table 7.1. Difluoroamines are
noticeably more stable than polynitro-compounds o f an analogous structure,
but have many similar features in their character o f decomposition. It is
seen from the comparison o f Tables 7.1 and 3.8 that the energy o f the C —
NF 2 bonds is somewhat higher than the D(C—N0 ) 2 value in polynitro-
analogues. In the methane derivatives, introduction o f each new N F - g r o u p 2

results in the decrease o f the D(C—N0 ) 2 value by 5-7 kcal/mol, i.e., nearly
the same as the decrease o f the D(C—N0 ) 2 value in polynitro-compounds
under the influence o f the N 0 group. The data available on monofunctional
2

difluoroamino-compounds also demonstrate that D(C—NF ) 2 > D(C—N0 ).2

For instance, in the case o f P h C H N F 2 2 and P h C H N 0 , their


2 2 D(C—N)
values are 51 [7.4] and 42 kcal/mol (see Table 3.8), correspondingly.
Apparently, for approximate estimation o f the D(C—NF ) 2 value in
difluoroamino-compounds, based on the data for nitrocompounds which are
easily accessible and have been studied in detail, one can accept that D(C—
N F ) = D ( C — N 0 ) + 4 kcal/mol.
2 2

In the case o f difluoroamines, the pre-exponential factor in the


expression for the rate constant has the increased value that is indicative o f
a semi-rigid type o f transition state.

105
106 Organic Difluoroamino-Compounds

Table 7.1. The kinetic characteristics of decomposition of /zewe-bis-difluoroamines


and related compounds

Compound State T,°C E, \ogA w2 Refs.


kcal/mol (s )
_1

220- 46.8 15.60 1.0 0.07 [7.1]


260
210- 47.9 16.42 2.0 [7.1]
230

Liquid 140- 46.9 16.15 2.1 [7.1]


165

F N 2 NF 2
Gas 210- 46.1 15.85 3.7 [7.1]
260

F N 2 NF 2

NF 2 Liquid 150— 47.1 16.00 [7.1]


180

F C(NF )
2 2 2 Gas 372- 53.7 15.75 io- 3
0.002 [7.2]
457
FC(NF ) 2 3 Gas 190- 48.4 16.45 1.2 - [7.2]
460
FC(NF ) 2 3 Gas 225- 45.9 15.73 3.5 0.02 [7.3]
250
C(NF ) 2 4 Gas 190- 40.5 16.40 5- 10 4
0.01 [7.2]
460

Note: w, is the relative reaction rate at 200°C; w is the ratio of the decomposition 2

rates of difluoroamines R C X _ (X = N F ) and nitrocompounds with a similar


n 4 n 2

structure (X = N 0 ) at 200°C. 2

The A values are the same for C ( N F ) 2 4 and C ( N 0 ) , as well as for


2 4

FC(NF ) 2 3 and F C ( N 0 ) . A positive influence o f the adjacent groups in


2 3

polynitro-compounds manifests itself in compounds with alkyl substitutes


(in the case o f M e C ( N 0 ) A = 1 0 2 2 2
1 7 1
s ). This effect is absent in the case
_1

o f difluoroamino-compounds, which is indicative o f a free rotation o f the


alkyl groups in /zewe-bis-difluoroamino-compounds. Hence, for instance,
the NF -derivatives o f propane and cyclohexane have the same /4/<r values.
2
Thermal Decomposition and Combustion 107

The secondary reactions during decomposition o f difluoroamines are


reduced to the transformation o f free radicals R C N F 2 2 with the F atom
elimination

R CNF 2 2 R C=
2 NF + F

or with the detachment o f R after rearrangement

R2CNF2 — R 2 C F N F — R C F = N F + R. :

Thus, the main products o f ( C H ) C ( N F ) decomposition are ( C H ) C = N F 3 2 2 2 3 2

and C H C F = N F [7.5].
3

During C ( N F ) decomposition, besides the following reactions


2 4

C(NF ) — - N F + C (NF ) 2 4 2 2 3

( F ) C = NF + F
N 2 2 FC(NF )= NF + N F 2 2

the secondary stage is probable [7.2]:

NF 2 + C (NF ) 2 3 ^ NF + (NF ) C=NF.


3 2 2

In addition, C F N F 3 2 and F C = N F are formed (apparently, on the vessel


2

walls). D u r i n g the decomposition o f F C ( N F ) , the molecules o f N , C F , 2 3 2 4

and N F are formed in equal amounts [7.3]. The formation o f these products
3

can also be explained by the reactions o f the N F 2 radical and various


rearrangements.
B y their absolute decomposition rate, polyfunctional difluoroamines
without the hydrogen a-atom belong to a number o f rather stable
compounds, and the heme-group - C ( N F ) - is one o f the most stable
2 2

fluoride-containing explosive groups and, in this regard, exceeds the nitro-,


nitrate-, azide-, and nitramine-groups.
A m o n g difluoroamines, the largest group o f substances are compounds
with single N F - g r o u p s (see Table 7.2). The gas phase decomposition o f all
2

these compounds, without exception, is drastically complicated by a


heterogeneous reaction on the surface. Hence, only a limited number o f
publications on the homogeneous gas phase reactions are available. They
are obtained under rather low pressures and for low S/V ratios, when the
surface influence does not manifest itself.
Monofunctional difluoroamines without the hydrogen oc-atom, i.e.,
substances o f C F N F 3 2 and ( C H ) C H F type, have to decompose according
3 3 2

to the radical mechanism with the primary C — N bond breaking and E =


D(C—N). However, these substances have not been studied. A s follows
108 Organic Difluoroamino-Compounds

from particular examples, compounds with the hydrogen ot-atom


decompose in vapors at high temperatures according to the radical
mechanism and at low temperatures ( T < 4 0 0 ° C ) through H F elimination.
Here, one can observe an analogy with the competition between the C — N
bond breaking and H N 0 2 elimination in mononitrocompounds.
In [5], decomposition o f l,2-bis(difluoroamino)propane was observed
at a pressure o f 10" Torr and temperatures 6 5 0 - 7 5 0 ° C . The activation 2

energy (55±3 kcal/mol) coincides with the Z)(C—N) value in this


compound (56 kcal/mol [7.17]). The main product is propylene (the yield is
95%). The reaction proceeds in two stages:

CH —^H-^H2 3 C H — C ' H — ^H +NF


3 2 2

NF 2 NF 2 NF 2

CH C HCH NF 3 2 2 CH CH=CH
3 2 +NF . 2

Table 7.2. The kinetic characteristics of decomposition of difluoroamines in the


liquid state [7.6]

Nos. Compound AK , cm7g


X T,°C E, kcal/mol log/*
(s- )1

1 C H CH NF
6 5 2 2 266 110-145 26.3 8.33 1.0
2 (C H ) CHNF
6 5 2 2 122 110-150 28.2 9.55 1.3
3 C H CH(NF )
6 5 2 2 315 110-150 26.0 8.30 1.3
4 C H CH(NF )CN
6 5 2 160 110-150 26.0 8.50 2.1
5 CH CH(NF )CH=CH-CH NF
3 2 2 2 303 110-145 24.0 6.88 0.7
6 C H CH(NF )CH NF
6 5 2 2 2 280 90-140 27.7 9.30 1.4
7 334 100-145 26.9 8.55 0.8
F2N<iH(CH ) CHNF 2 4 2

8 [F NCH CH(NF )CH ]


2 2 2 2 2 310 100-140 28.6 9.35 0.5
9 F NCH CH(NF )CH 00C-
2 2 2 2 268 100-150 24.0 6.80 0.7
C(CH )(NF) CH NF 3 2 2 2

10 F NCH (CHNF ) CH
2 2 2 3 3 332 107-157 25.4 8.00 1.6
11 F NCH C(CH )(NF )-
2 2 3 2 211 107-157 26.6 8.60 1.2
CH(NF )CH NF 2 2 2

12 (CH )CHC(CH )(NF )-


3 3 2 268 90-140 30.0 9.80 0.1
CH NF 2 2

13 [F NCH C(CH )(NF )-


2 2 3 2 248 110-160 28.6 8.73 0.1
C00CH ] 2 2

14 CH C00C(CH )(NF )-
3 3 2 180 110-160 30.9 9.60 0.05
CH C(CH )(C00CH )-CH NF
2 3 3 2 2

15 (CH ) C(NF )-C(NF )(CH )


3 2 2 2 3 238 110-160 30.0 9.38 0.1

* w is the relative rate at 100°C.


Thermal Decomposition and Combustion 109

The following products are formed in small amounts: C H C ( C N ) = N F (the 3

product o f elimination o f three H F molecules), F C H C N , H C N , and C H C N 2 3

(these products are formed on the walls).


For PhCH NF , 2 2 the exact parameters of HF elimination were
determined [7.6]: E = 42.5 kcal/mol and log A = 11.78. For this compound,
the radical decomposition should have the following kinetic characteristics:
E = 51.0 kcal/mol and log A = 14.5. The isokinetic temperature, when the
mechanism is changed, is 4 5 0 ° C .
The same activation energy (42.0 kcal/mol) is estimated for the thermal
reaction

CH NF —^ 3 2 CH —NF+HF2

from the data on the decomposition o f chemically activated C H N F 3 2 [7.8]


for A = 1 0 13 5
s~ . The isokinetic temperature for this compound is 7 0 0 ° C
l

[7.8], The radical decomposition should have the following parameters: E =


60.0 kcal/mol and ,4 = 1 0 1 7 1
s" . 1

Similarly, by the chemical activation method for the reaction

CH CH(NF2)CH NF2 —
3 2 HF+CH CCH NF 3 2 2

\ ^ NF

^ C H C H ( N F ) C H = N F + HF.
3 2

it was found that E = 38 kcal/mol [7.9] for the calculated value o f


^ = 10 s- .m 1

The data presented in [7.9] on Ph(CH )CNF C(NF )(CH )Ph 3 2 2 3

decomposition at 2 2 0 ° C do not agree with the scheme o f two competitive


reactions: the H F elimination reaction and radical decomposition. A l o n g
with the formation o f P h C N , for this relatively low temperature, a high
yield o f stilbene on the wall is observed. Similarly, a high yield o f
anthracene at 2 0 0 ° C was obtained in the reaction of 9,10-
bis(difluoroamino)-9,10-dihydroanthracene decomposition [7.10]. The N F 2 4

elimination is possible:

/ P h / h p h Ph

CH C—CCH 3 3 C H 3 — C = C — CH3 + N2F4.

NF 2 NF 2

Due to the low energy o f the C — N bond (50 kcal/mol), the endothermic
effect o f this reaction does not exceed 20 kcal/mol and this reaction cannot
be excluded for energetic reasons.
110 Organic Difluoroamino-Compounds

A l l difluoroamines with N F - g r o u p s , w h i c h are located


2 separately,
decompose considerably more rapidly in the liquid state than in vapors.
They demonstrate a close similarity among themselves by their
decomposition character and have a distinctive mechanism o f the initial
stage. The reaction product is always H F , w h i c h noticeably catalyzes the
decomposition. The kinetic reaction can be traced by the liquid-gas
chromatography method from the decrease o f the initial substance or
manometricaliy from the initial rate. The data o f manometric measurements
are rapidly distorted due to the polymerization o f gas phase products.
The data available on the liquid phase decomposition are presented in
Table 7.2. The following regularities are observed. The decomposition rate
o f benzyldifluoroamine P h C H N F 2 2 in the liquid state is higher than in
vapors by a factor o f 10 . In solutions, the decomposition rate o f P h C H N F
6
2 2

depends on the dielectric medium (Table 7.3) [7.6] and, in this case, at
132°C we have

log£ = - 4 . 6 4 4 - 3 . 3 3 / D ,

w h i c h is indicative o f a strong polar character o f the transition state. In spite


o f various structures, compounds 1-11 in Table 7.2 have the same
decomposition rates. Even P h C H N F , P h C H N F , and P h C H ( N F )
2 2 2 2 2 2 can
hardly be distinguished by the reaction rates. Such insensitivity to the kind
o f substitutes suggests a distant location o f these substitutes from the
reactive center, which is apparently located not on the C but on the N atom.
The l o w values o f the pre-exponential factor cannot be consistent with any
homolytic monomolecular mechanism and they are typical o f ionic
processes o f E l type. A l l these facts can be explained i f we assume that the
liquid phase decomposition o f difluoroamines proceeds through the ionic
dissociation o f the N — F bond according to the following scheme:

ki
R C H N F ^ = RCH NF F —
2 2 2
+ _
R C H = N F + HF
k 2

In this case, the observable values o f the rate constants and activation
energies are A : observ = kxk^k^ and £ bserv 0
=
£i + E - 3 E.
2

Compounds 12-15 from Table 7.2, having chain branching at a and (3-
carbon atoms, have a decomposition rate value which is one order o f
magnitude lower than these values for all other compounds. This fact can
also be explained within the framework o f the proposed mechanism by a
steric barrier for the ionic pair solvation caused by voluminous substitutes.
Decomposition o f the N F - g r o u p located near the tertiary carbon atom
2

proceeds at the same rate as that o f the R C C H N F 3 2 2 group sterically


protected on the side o f the (3-atom.
Thermal Decomposition and Combustion 111

Table 7.3. The kinetic characteristics of decomposition of n-benzyldifluoroamine


[7.6]

Compound r,°c E, kcal/mol log/1 (s" ) k • 10 , s" (132°C)


1 6 1

Gas 280-300 42.5 11.78 6.96- 10~ 6

Liquid 110-145 26.3 8.33 1.37


Solution in CC1 4 170-190 25.4 7.61 0.78
Solution in C H 6 6 170-190 25.6 7.90 1.21
Solution in CHC1 3 132 - - 5.20
Solution in CH CN 3 132 - - 18.00
Solution in CH N0 3 2 132 - - 22.00

Thus, in monofunctional structures, the presence o f the ot-hydrogen atom is


not a necessary condition o f a prompt decomposition in contrast to the
situation with /zewe-6/Xdifluoroamino)-compounds [7.1]. Apparently, the
cation component o f the ionic pair with the tertiary carbon atom is subjected
to rearrangement completed by the ejection o f H +

(CH ) CN F —
3 3
+
(CH ) C N(CH )F —
3 2
+
3 H C=C(CH )N(CH )F +H
2 3 3
+
.

However, in this case, the value o f k is smaller and E is higher than during 3 3

a direct detachment o f H +
from R C H N F .
2
+

In the case o f /zewe-bis(dirluoroamino)-compounds, the detachment o f


H F does not proceed i f there is no hydrogen atom in a-position. The cation
rearrangement o f the following type

CH (NF )C—N F — 3 2
+
(NF )C —NF(CH )
2
+
3

cannot proceed because the electronegative NF 2 group destabilizes a


positive charge near the a-carbon atom.
Shielded difluoroamines differ from unshielded ones by the
decomposition rate at low temperature ( 2 0 - 3 0 ° C ) by a factor o f 10 and can 2

be considered as rather stable compounds. At 20°C, the time of


decomposition o f these compounds by 0.01% is several tens o f years.

REFERENCES

7.1. A . V . Fokin, V . N . Grebennikov, G . M . N a z i n , Dokl. Akad Nauk


SSSR, 3 3 2 : 7 3 5 - 7 3 6 (1993).
7.2. J. M . Sullivan, A . E . A x w o r t h y , T. J. Hauser, J. Phys. Chem., 74:
2611-2620(1970).
7.3. E . W . Neuvar, Ind. Eng. Chem. Res., 28: 319-320 (1969).
112 Organic Difluoroamino-Compounds

7.4. V . I. Pepekin, Y u . A . Lebedev, G . G . Rozantsev, et al., Izv. Akad.


Nauk SSSR, Ser. Khim., 2: 452-453 (1969).
7.5. D . S. Ross, T. M i l l , M . G . H i l l , J. Amer. Chem. Soc., 94: 8776-8778
(1972).
7.6. V . N . Grebennikov, G . M . N a z i n , G . B . Manelis, Proc. 27th Int.
Conf. ICT, Karlsruhe, p. 106 (1996).
7.7. A . J. Dijkestra, J. A . Kerr, A . J. Trotman-Dikenson, J. Chem. Soc.
A, 105-110 (1967).
7.8. D . S. Ross, R. Shaw, J. Phys. Chem., 75: 1170-1172 (1971).
7.9. R. Shaw, Int. J. Chem. Kinet., 9: 689-691 (1977).
7.10. H . Cerfortain, J. Chem. Soc, 6602-6606 (1965).
Chapter 8

HETEROCYCLIC COMPOUNDS

A m o n g heterocyclic compounds, the derivatives o f furazan, furoxane, and


tetrazole are distinguished by their energetic characteristics. It is precisely
the combination o f these highly endothermic heterocycles with the ordinary
explosive groups ( N 0 , N , N N 0 , etc.) that, in principle, can provide the
2 3 2

creation o f new powerful explosives. The polymer compounds o f a


polyvinylmethyltetrazole type were also considered as prospective plastic
binders o f solid rocket propellants.

8.1 Thermal Decomposition of Furazanes and


Furoxanes

Chemical mechanism of decomposition

The main results concerning decomposition reactions o f furazanes ( F A ) and


furoxanes ( F O ) are presented in [1—7]. Mostly, the data were obtained in the
gas phase, and they demonstrate a strong dependence o f the decomposition
rate on the structure o f substitutes in the ring, especially in the furazane
series. A s follows from several examples, the passage from the gas phase to
the condensed phase is not accompanied by variation o f the decomposition
rate and mechanism. In the solid phase, as a rule (only the data for
benzotrifuroxane are available), the decomposition rate is considerably
lower than in the liquid state or in solutions.
In the majority o f cases, the decomposition o f furazanes and furoxanes
represents the reaction o f retro- 1,3-dipolar cyclo-attachment, which leads to
the formation o f appropriate nitriles and nitrile-oxides:
R —C=Nk
1
z -R 1 2
C N + R NCO
U 2

113
114 Heterocyclic Compounds

In the case o f furoxanes, nitrile-oxides are subjected to isomerization


into isocyanates, which are partly also subjected to polymerization and
oxidation, but are mainly stabilized as the final products. For the case o f
furazanes, the cyclization o f nitrile-oxides with nitriles into 1,2,4-
oxadiazoles is possible. Such a consequence o f transformations is typical o f
alkyl-derivatives o f furoxanes and furazanes, and was revealed in the first
studies. Thus, in [8.8], it was found that diphenylfurazane was sublimated
when partially decomposed. A s the result, benzonitrile, phenylisocyanate,
and 3,5-diphenyl-l,2,4-oxadiazole are formed. The reverse cyclization o f
nitriles and nitrile-oxides into disubstituted furazanes has never been
observed.

f
1.0-

Fig. 8.1.1. Dependence of the relative molar yield (<p) of the decomposition products
of dimethylfurazane on the degree of conversion (n) for T = 300°C, P = 200 mm 0

Hg, S/V = 3.5 cm" : (/) C H C N , (2) 3,5-dimethyl-l,2,4-oxadiazole (ODA), (3)


1
3

C H N C O , (4) O D A in the mixture with dimethylfurazane (DMF) and C H C N in the


3 3

ratio (1 : 3), (5) O D A at S/V = 30 cm" , (6) O D A in the mixture of D M F and PhCN
1

at a ratio (1 : 1.5).

Formation o f decomposition products in the course o f the reaction has been


studied most comprehensively for the case o f dimethylfurazane ( D M F )
decomposition [8.2]. A t earlier stages, the only products are C H C N and 3

C H C N O ( F i g . 8.1.1). The presence o f the latter product can be observed by


3

the method o f chemical traps, for instance, by the formation o f 3-methyl-5-


phenyl-l,2,4-oxadiazole i f the reaction proceeds in excess o f P h C N . In the
course o f the reaction carried out without additives, the concentration o f
C H N C O increases and 3,5-dimethyl-l,2,4-oxadiazole ( O D A ) appears. The
3

balance o f carbon atoms, taking into account the formation o f C H C N , 3

O D A , and C H N C O , at the end o f the reaction is fulfilled by no less than


3

95%. Such compounds as C O , C 0 , N , and polymerization products o f


2 2

C H C N O (on the surface) are formed in small amounts (up to 3%). Such
3

products as N O , N 0 , H C N , ( C N ) , and C H N 0
2 2 3 2 are detected in trace
amounts. O n addition o f N O and N 0 , the yields o f C H C N , C H N 0
2 3 3 2 and
Thermal Decomposition and Combustion 115

light gases are increased, but the reaction rate does not change. It has been
revealed that the following reaction proceeds on the surface:

C H C N O - ^ - CH3CN + O .
3

The following processes are also probable:

CH CNO+NO-^-- CH CN + N 0
3 3 2

C H C N O + N 0 ^ C H ( j > = N = 0 — - CH NO+CO +NO


3 2 3 3

ONO

CH NO + N 0 —
3 2 C H N 0 + NO
3 2

C H N O — 3 HCN + H 0 . 2

Methylisocyanate can also participate in similar reactions. Its oxidation


gives rise to the formation o f C 0 and C H N 0 . 2 3 2

During decomposition o f dimethylfuroxane (under the following


conditions: T= 2 8 0 ° C , S/V = 3.5 c m , P = 100-200 Torr), the main final - 1
0

product is C H N C O . The yield o f C H C N


3 3 is 7-8%, and the volume
fractions o f C O , C 0 , N O , N 0 , H 0 , ( C N ) , C H O N 0
2 2 2 2 3 2 are approximately
4.5%. A s the pressure PQ decreases up to 7 Torr, the yield o f C H C N and 3

volatile gases increases due to the increase o f C H C N O decomposition on 3

the surface. The addition o f N O results in the increase o f the C H N 0 3 2 yield.


Evidently, the secondary reactions are similar to the processes observed
during dimethylfurazane decomposition.
In some cases, products are formed in which the C — C bond is
preserved in a heterocycle that is indicative o f the stepwise decomposition
character. Thus, from l-methyl-2-nitrofurazane and similar furoxane, the
C H C O C N compound is formed in the amount o f 0.8 and 0.3 moles,
3

correspondingly:

H C^
3 N0 2 H C
3 N0 2
H 3 C
\ / ° / / N
~ 0

c-c c-c ^ c—c


11 w — ^ // w — 1 1/
N x N N 'N N N
0' o x
o
x
0 "i O O
O

C H C — C = N +NO + N 0 .
3 2

Kinetic data

The decomposition rates of furazane and furoxane were measured


manometrically (P 0 < 10 Torr, S/V < 0.6 c m ) and with the use o f the -1

ampoule method with the liquid-gas chromatographic analysis o f the parent


substance (P ~ 200 Torr, S/V= 3-4 cm" ).
0
1
Table 8.1.1. The kinetic characteristics of decomposition of furazanes (FA) and furoxanes (FO)
Substitutes R-—C=N
1
R --C=N.
1
Refs.
No. R- 1 >o
2
R—1 >o
2

Furazanes Furoxanes
-C=N C=N. 0

R 1
R 2
r,°c E, Iog04) "^relative T,°C E, log(/0 ^relative

kcal/mol (s-)
1
(200°C) kcal/mol (s-) 1
(200°C)
1 4-NH-benzo- - 54.9 13.60 4.6 - IO - - - [8.1]
-5
2

2 Benzo- 400-450 54.5 13.70 8.1 • 10~ -


5
39.4 12.0 2.10 [8.3]
3 4-N0-benzo-
2
- 54.4 14.30 3.5 • 10" -
4
37.7 11.80 1.90 [8.7]
4 5-N0-benzo-
2
- 54.2 14.50 6.6 • 10 -
-4
37.47 11.80 2.30 [8.7]
5 4,6-(N0)-benzo- 54.2 14.60 8.3 • 10" - 37.94 12.10 3.10 [8.7]
-
4
22
- -
6 NH NH 290-340 47.8 14.15 8.1 • 10~ - - [8.3]
2
2 2
-
7 NH 2 N0 2 290-340 47.6 14.17 0.1 - - - [8.3]
8 CH3 N0 2 260-320 46.7 14.30 0.3 200-240 40.0 40.20 63.0 [8.3]
-
9 Ph NCb - - - 180-220 38.5 14.80 900.0 [8.5]
10 H H 250-320 43.5 13.40 0.6 200-240 40.0 14.20 800.0 [8.4]
11 CH3 CH 3 260-320 46.0 14.54 1.0 220-300 45.0 14.75 1.0 [8.3]
12 CH3 CHs 240-300 46.1 14.93 2.2 - - [8.3]
13 C2H5 C3H7
2

240-290 45.8 14.91 2.8


- - [8.3]
14 Benzotri-FA and -FO 46.5 15.40 4.5
- -
45.3
-
15.30
-
0.70 [8.6]
- - - -
15 -C(NF)CHCHCH- 2 2- 2
- 290-320 46.2 14.58 0.65 [8.3]
16 -(CH ) - 250-300
2 4 45.1 15.20 9.8 270-300 46.7 14.77 0.11 [8.3]
Thermal Decomposition and Combustion 117

The data o f these two methods are very close [8.3]. The reactions obey the
equation o f the first order. The additives o f C H N C O , P h C N , N O , and
3

acetone change the composition o f the products but do not influence the
decomposition rate. The parameters o f monomolecular reactions of
furazane and furoxane decomposition are presented in Table 8.1.1.
The conservation law o f orbital symmetry allows the concerted
decomposition o f furazane and furoxane. However, as follows from the
kinetic data, this concerted decomposition is realized only in individual
cases.
Conservation o f orbital symmetry requires that in the ground state the
concerted processes o f cyclodecomposition o f furazane and furoxane, for
instance, o f a -[2 n + 2 ] type, have the identical allowed geometry o f the
n

combination o f molecular orbitals. In this case, under the influence o f


substitutes, the rates o f decomposition o f furazane and furoxane have to
vary in one direction.
A t the same time, as can be seen from Table 8.1.1, they vary in opposite
directions. Such experimental regularity is indicative o f a stepwise
mechanism.
The primary N — O bond breaking leads to the formation of
intermediate biradicals, whose structure is close to the structure o f the
activated reaction complex. In this case, the unpaired electrons o f N and O
atoms are in a - and (3-position with respect to the ^--system o f double
bonds; in the first case, they are destabilized and, in the second case, are
stabilized by this system.
A c c o r d i n g to the general quantum-mechanical rule, an antibonding
orbital loosens more effectively than a bonding orbital bonds [8.9]. Thus,
the unpaired electron o f the nitrogen atom plays a dominant role, but only in
the case o f furazane. A reduced electron density o f the N atom in furoxane
[8.10, 8.11] noticeably reduces the destabilizing role o f the a-electron, and
the opposite influence o f the /^-electron o f the oxygen atom becomes
dominant.
Substituents exert the same influence on the coupling in the ^--system o f
the transition state o f the opening o f heterocycles o f furazane and furoxane,
decreasing it in the series 1-16 (the coupling with substitutes decreases, the
disturbance o f a plane structure o f the cycle increases due to repulsion
between the substitutes as their volume increases or due to deformation
strain produced by the saturated carbon cycle). In accordance with this, the
stability has to decrease for furoxane and increase for furazane, which is
observed experimentally.
Against the background o f the expected variation o f the rate in the
derivatives o f furoxanes, a sharp peak is observed for unsubstituted
furoxane. In this case, obviously, the decomposition proceeds according to
the concerted mechanism. The main barrier for a concerted reaction is the
steric hindrance o f the transition state formation from the two linear
fragments, which are brought close together:
118 Heterocyclic Compounds

R'\
c-cw/ R
2
Ic-
// ^
N X N II;

oo
This barrier is eliminated when R = R = H . It is probable that in the case
1 2

o f plane substitutes, such as Ph and N 0 , a concerted splitting o f a


2

heterocycle also occurs, at least partially.


In the case o f a stepwise biradical mechanism, the scheme o f the
limiting decomposition stage takes the following form:

R —C=N
1
/ R-—C=NL R"—C—N* . 2

T > 0 ^ \ I > 0 7 • /O — products


R — C = N ^ ° ^ T R —C=N
2 2 U
' R —C=N"
2

.0

R4=N R - L N > ° - R - L N - °
>0; 2
0
2

,2_i_ ; 0
^ products.
R^— C = N .
x

Decomposition of furoxane takes place against the background of


considerably more rapid geometric isomerization. Hence, both isomers are
shown in the scheme as the parent ones.
The synthesis o f furoxane from nitrile-oxides, which is the reverse
process for a stepwise cyclodecomposition, is a special case o f the reactions
o f 1,3-dipole cycloattachment. A c c o r d i n g to the principle o f macroscopic
reversibility, the formation o f furoxane has to proceed through the biradical
transition state, which is the same transition state as for decomposition. The
conclusion obtained corresponds to the reaction mechanism proposed in
[8.14].

Table 8.1.2. The kinetic characteristics of decomposition of benzotrifuroxane [8.6]

State r,°c E, kcal/mol log/4 (s- ) 1


k- 10 , s" (70°C)
7 1

Solution in sy/w-trinitro- 190-230 41.5 13.8 2.1


benzene (5-20%)

Melt 200-250 41.5 14.2 5.2

Solid 170-190 68.0 26.0 0.28


Thermal Decomposition and Combustion 119

Individual observations carried out for such compounds as benzofuroxane,


tetramethylenefurazane, and methylnitrofurazane show that in the liquid
phase the rates differ only slightly from those in the gas phase. The
quantitative data were obtained for benzotrifuroxane ( B T F ) (Table 8.1.2).
The melting temperature o f this substance is 197°C. Decomposition in the
solid phase was studied within a narrow temperature range near the melting
temperature, and, thus, the E value is noticeably overestimated. However,
the reaction rates in the solid B T F and in the melt differ by a factor o f 20.
Probably, at lower temperatures, this difference is larger. B T F is a highly
reactive compound, probably due to the presence o f a nitroso-form, i.e., the
intermediate state during geometric isomerization.

8.2 Stability of Tetrazoles


A l l compounds o f the tetrazole series can be divided into three groups by
the stability o f heterocycle to splitting and, partially, by the mechanism o f
this process: tetrazole itself and its 5-monosubstituted, 1,5- and 2,5-di-
substituted tetrazoles.

1,5-Disubstituted tetrazoles

The primary degradation o f a tetrazole cycle for the 1,5-disubstituted


tetrazoles proceeds according to the following reaction:

R^C-N-R 1
l R — C=N-R
2
/
1
, ^ R — N = C=NR
2 1

- other products .

The first evidence o f this reaction was obtained in 1958 in the works on 1,5-
diaryltetrazoles decomposition [8.15, 8.16]. In addition to carbodiimides
A r N = C = N a r ' , the products o f their cyclization and disproportionation,
benzimidazoles were found during the decomposition o f N-naphthyl-
tetrazoles

which corresponds to the monomolecular reaction o f a nitrene atom


insertion in the C — H bond. Later, the direction o f decomposition according
to (8.2.1) was also confirmed in [8.17-8.22]. The kinetic parameters o f the
monomolecular decomposition o f 1,5-disubstituted tetrazoles are presented
120 Heterocyclic Compounds

in Table 8.2.1. The substitutes in the fifth position exert a weak influence on
the rate o f decomposition, in contrast to the substitutes in the first position,
w h i c h noticeably influence it [8.15, 8.24].

Table 8.2.1. The kinetic characteristics of thermal decomposition of 1,5-substituted


tetrazoles
R —C-N-R
1 2

R ,N products

R 1
R 2
State, r,°c £, kcal/mol log/1, Refs.
solvent (s- )
1

CH 3 CH 3 Gas 270-320 47.1 14.97 [8.23]


CH =CH
2 CH 3 Gas 270-320 47.5 15.05 [8.23]
Ph CH 3 Gas 240-290 47.7 15.56 [8.23]
CH 3 CH 0-CH
2 3 Gas 230-280 47.8 15.20 [8.23]
Ph NH 2 Nitro­ 160-210 39.8 14.34 [8.23]
benzene
CH 3 CH CH -
2 2 Nitro­ 190-210 44.5 15.31 [8.23]
COCH 3 benzene
Ph C-N-Ph Nitro­ 200-240 44.7 15.66 [8.23]
// \
benzene

There is evidence o f the existence o f the reaction o f l-(2-pyridyl)-5-


aryltetrazole pyrolysis

X
=N N x N N N
1
N = C-Ar

in boiling tetralin ( 2 0 6 ° C ) [8.22]. For any R and A r , the rate constants fall
within the range o f (1.4-3.6) • 10~ s 4 -1
which corresponds to £ a c t i v = 41
kcal/mol at/J = I O 1 3 5
s" .
1

Transformation o f tetrazoles in corresponding azides at stage 7 o f


reaction (8.2.1) is endothermic, and the value o f AH for this reaction is up 0

to 10 kcal/mol [8.25, 8.26]. The kinetic characteristics o f stage / o f channel


(8.2.1) can be obtained from the data o f Dimroth's rearrangement [8.26] o f
1-R-derivatives o f 5-aminotetrazole. This reaction proceeds in several
stages

H N-C-N-R HN-C-NHR
// \

/-
2
/ W
N x N

H NC=NR many
NH - C N H R
2

N-N rapid \
2
stages ,N
N ^2
Thermal Decomposition and Combustion 121

but only the opening o f the tetrazole cycle can be a limiting process. The
typical value o f the activation energy for this reaction is 36.5 kcal/mol for R
= alkyl and 33.7 kcal/mol for R = aryl [8.10]. The reverse process [stage -1
of reaction (8.2.1) and Dimroth's rearrangement] of cyclization of
azidoazomethanes into tetrazoles has been studied in detail [8.26, 8.27]. For
this process, the following parameters are typical: E_\ = 17-18 kcal/mol,
= - 7 e. u. Hence, decomposition o f azomethineazides R(N )C=NR
3

should proceed with the activation energy which is E = 28 kcal/mol, w h i c h


coincides with the E value o f decomposition o f vinylazides R ( N ) C = C R R
3
! 2

and carbonylazides R ( N ) C = 0 (see Chapter 5). Decomposition o f tetrazoles


3

occurs against the background o f a reversible isomerization similar to the


decomposition o f furoxanes which takes place against the background o f a
structural isomerization through the form o f dinitrosocompounds.

2,5-Substituted tetrazoles

A c c o r d i n g to R. Huisgen [8.28-8.31], the decomposition o f 2,5-substituted


tetrazoles proceeds according to the following reaction:

R — C = N R — C - N ^ N - R 1

/ \ i xX

N v ^ N - R 1
^ N R^-CEEN—NR^Ns (8.2.2)

The kinetic parameters are presented in Table 8.2.2.


2,5-Disubstituted tetrazoles are less stable than 1,5-substituted ones
mainly due to reduction o f the energetic barrier o f stage / (stabilization o f
the electron in the transition state by the 2 /r-system o f the N = C bond) and
increase o f the rate o f reaction 2, which is competitive with reaction - / .
In this stage, in the process o f nitrogen detachment, to form
nitrileimine, skeleton rearrangement is not required. It is required for
reaction 2 o f scheme (8.2.1) to proceed.

Tetrazole and 5-substituted tetrazoles

Different rates o f decomposition o f 1,5- and 2,5-disubstituted terazoles


suggest that prototropic tautomerism o f 5-substituted tetrazoles w i l l play an
essential role in their thermolysis. There is evidence that, in the solid state,
5-substituted tetrazoles exist in the l H - f o r m and exactly this form prevails
in solutions. O n the other hand, this form has a smaller decomposition rate
and, thus, the thermolysis reaction has to proceed through the isomerization
stage into the 2H-form.
122 Heterocyclic Compounds

Table 8.2.2. The kinetic characteristics of thermal decomposition of 2,5-di-


substituted tetrazoles

/ \ 9 products.
N x N - R 2 F

R R State T,°C £, kcal/ log/* Refs.


1 2

mol

CH 3 CH 3 Gas 190-260 42.2 15.01 [8.23]


CH 3 CH 3 Nitro­ 170-220 42.4 14.78 [8.23]
benzene
CH =CH3 CH 3 Gas 210-250 43.0 15.38 [8.23]
Ph CH 3 Gas 190-240 42.4 15.29 [8.23]
Ph CH 3 Nitro­ 180-230 42.0 14.85 [8.23]
benzene
CH 3 C H C O - C H Nitro­
2 3 190-230 39.9 14.97 [8.23]
benzene
CH 3 C H C O - C H Gas
2 3 160-210 42.2 14.82 [8-23]
Ph NH 2 Nitro­ 150-190 38.8 15.35 [8.23]
benzene
N0 2 C H C O - C H Gas
2 3 180-250 36.7 13.19 [8.23]
N0 2 Pic Liquid 75-95 34.1 13.76 [8.23]
0 N C = NON =C
2
C H
2 5 Gas 180-250 38.3 13.53 [8.23]

0 N C = NON =C
2
C H
3 7 Gas 180-250 38.3 13.53 [8.23]
1
1
0 NC=NON= C
2
(CH ) 2 2 Nitro­ 140-190 38.0 14.40 [8.23]
| -N0 2 benzene
Ph 4-C(0)- Liquid 207 k = 6.95 • i o - [8.33] 4

C H N0
6 4 2

Ph COPh Liquid 207 k = 7.42 • IO' [8.33] 4

Ph 4-C(0)- Liquid 207 k - 2 . 7 6 • 1 0 ' [8.33]


4

C H N0
6 4 2

4-N0 C H 2 6 4 COPh Liquid 207 k= 2.26 •IO"4


[8.33]
4-CH OC H 3 6 4 COPh Liquid 207 k= 1.06 • 10" 4
[8.33]
H COPh Liquid 110 k= 1.65 •10'
5
[8.34]
H cyclo- Liquid 110 k= 1.65 • IO" 5
[8.34]
C H,,CO
6

During the decomposition o f 5-substituted tetrazoles, products 3 4 - 3 7 are


formed, which are similar to the products o f decomposition o f 2,5-
disubstituted tetrazoles, i.e., mainly the derivatives o f triazole, tetrazine,
s/w-triazine, but not carbodiimides, which are typical o f the decomposition
o f 1,5-derivatives.
Thermal Decomposition and Combustion 123

The kinetic data presented in [8.36, 8.37] indicate that decomposition o f


tetrazoles, which are not substituted in the position o f the nitrogen atom,
and decomposition o f corresponding 2,5-substituted tetrazoles occur at
almost equal rates. Thus, the 2H-form plays a decisive role in
decomposition and the decomposition scheme has to include a prompt
prototropic tautomerism

N - % , N H » 2 -3-

The observable rate constant takes the form:

Kk k x 2

obs
™~(\ + K)(k_ k_ ) ]+ 2

A t k_\ >> k we w i l l have


2

_ Kk k
~o a:)jl, '
x 2

observ
+

and at k. x « k we w i l l have
2

Kk
k.
^observ " ' 1

(1 + / Q

The fact that K does not influence Observ g phase indicates that the
m t n e a s

2H-form o f 5-substituted tetrazoles prevails in vapors.

REFERENCES

8.1. V . V . Zverev, I. Sh. Saifullin, G . P. Sharnin, Izv. Akad Nauk, Ser.


Khim., 2: 313-317 (1978).
8.2. V . G . Prokudin, G . M . N a z i n , V . V . Dubikhin, Kinet. Katai, 22:
871-876 (1981).
8.3. V . G . Prokudin, G . M . N a z i n , G . B . Manelis, Dokl. Akad. Nauk
SSSR, 2 5 5 : 9 1 7 - 9 2 0 (1980).
8.4. V . G . Prokudin, G . M . N a z i n , Izv. Akad. Nauk, Ser. Khim., 1: 2 2 1 -
223 (1987).
8.5. V . G . Prokudin, G . M . N a z i n , I. V . Ovchinnikov, Izv. Akad. Nauk,
Ser. Khim., 2841-2844 (1987).
124 Heterocyclic Compounds

8.6. Y u . Y a . M a k s i m o v , E . N . Kogut, E . G . Sorokina, Chemical


Physics of Condensed Explosives, M . : M K H T I , 104: 19-22 (1979).
8.7. Y u . Y u . N i k i s h e v , I. Sh. Saifullin, O. R. K l u c h n i k o v , Kinet. Katai,
34:969-971 (1993).
8.8. K . A u w e r s , V . Meyer, Berichte, B d . 21: 784-817 (1888).
8.9. M . Orchin, G . Djaffe, Antibonding Orbitals, V . A . Cronghause
E d . , M . : M i r , 112p. (1969).
8.10. M . C a l l e d , L . Bonaccorti, D . Viterbo, Acta Crystallogr. B. 33:
3546-3548 (1977).
8.11. D . Britton, W . E . Noland, Acta Crystallogr. B. 28: 1116-1121
(1972).
8.12. F . B . M a l l o r y , A . Cammarata, J. Amer. Chem. Soc, 88: 6 1 - 6 4
(1966).
8.13. F. B . M a l l o r y , S. L . Manatt, C . S. W o o d , J. Amer. Chem. Soc, 87:
5433-5438 (1965).
8.14. R. A . Firestoun, J. Chem. Soc. A., 9: 1570-1575 (1970).
8.15. J. Vanghan, P. A . Smith, J. Org. Chem., 23: 1909-1912 (1958).
8.16. P. A . S. Smith, E . Leon, J. Amer. Chem. Soc, 80: 4647-4654
(1958) .
8.17. C . Wenirup, Tetrahedron., 27: 1281-1286 (1971).
8.18. P. G . Houghton, D . F . Pipe, C . W . Rees, J. Chem. Soc Chem.
Commun.,ll\-112 (1979).
8.19. N . A . K l u e v , Y u . V . Shurukhin, V . A . Konchits, at al., Chemistry
of Heterocyclic Compounds, 265-21 \ (1980).
8.20. R. J. Rao, K . K . Reddy, Ind. J. Chem. B, 22: 117-120 (1983).
8.21. Y u . V . Shurukhin, N . A . K l y u e v , I. I. Grindberg, V . A . Konchits,
Chemistry of Heterocyclic Compounds, 1422-1427 (1984).
8.22. V . V . Nedel'ko, V . P. Roshchupkin, G . G . Asatryan, High-
Molecular Compounds: Short Communications, 29: 2088-2094
(1987).
8.23. V . G . Prokudin, V . S. Poplavskii, V . A . Ostrovskii, Izv. Ros. Akad.
Nauk, Ser. Khim., 2209-2214 (1996).
8.24. G . I. K o l d o b s k i i , V . A . Ostrovskii, V . S. Poplavskii, Chemistry of
Heterocyclic Compounds, 1299-1326 (1981).
8.25. R . Huisgen, Angew. Chem. Intern. E d . : Engl., 19: 947-973 (1980).
8.26. R. A . Henry, W . G . Finnegan, E . Lieber, J. Amer. Chem. Soc, 77:
2264-2270(1955).
8.27. A . S. E n i n , G . I. K o l d o b s k i i , V . A . Ostrovskii, L . I. Bagal, Zh. Org.
Khim.,8: 1895-1901 (1972),
8.28. R. Huisgen, M . Seidel, J. Sauer, et al, J. Org. Chem., 24: 892-893
(1959) .
8.29. R. Huisgen, H . J. Sturm, M . Seidel, Chem. Ber., B d . 94: 1 5 5 5 -
1562 (1961).
8.30. R. Huisgen, H . J. Sturm, M . Seidel, Chem. Ber., B d . 94: 2 5 0 3 -
2509(1961).
Thermal Decomposition and Combustion 125

8.31. R. Huisgen, M . Seidel, C . W a l l b i l l i c h , H . Knupfer, Tetrahedron,


17: 3-29 (1962).
8.32. R. Huisgen, Angew. Chem., B d . 72: 359-372 (1960).
8.33. R. Huisgen, J. Sauer, M . Seidel, Liebigs Ann., B d . 654: 146-160
(1962).
8.34. H . Reimlinger, Chem. Industry, 94: 296-335 (1972).
8.35. C . Wentrup, Tetrahedron, 30: 1301-1311 (1964).
8.36. J. H . Markqraf, S. H . B r o w n , M . W . Kaplinsky, R. G . Peterson, J.
Org. Chem., 29: 2629-2632 (1964).
8.37. G . O . Reddy, V . R. M o h a n , B . K . M . M u r a l i , A . K . Chatterjee,
Thermochim. Acta, 43: 61-73 (1981).
Chapter 9

NITROESTERS

Nitroesters belong to one o f the most widely used classes o f organic


compounds employed in explosives and powders. For example,
nitroglycerin and cellulose nitrate are among them. A t the same time, a
relatively low thermal stability and compatibility with the other components
is a serious difficulty for their commercial use in various compositions,
especially if high-temperature action is required. The thermal
decomposition o f nitroesters is considered in detail in [9.1-9.5].
At present, it is revealed with high reliability that the thermal
decomposition o f nitroesters proceeds through O — N 0 2 bond breaking with
the formation o f nitrogen dioxide and corresponding alkoxyl-radical with
their following transformations into the final products. This reaction is
reversible for nitroesters. Small amounts o f N 0 2 reduce the decomposition
rate [9.2, 9.6, 9.7]. This mechanism is confirmed by fitting o f the
observable activation energy o f decomposition with the independently
determined energy o f the O — N 0 2 bond. For the majority o f nitrates, this
value varies within rather narrow limits o f 3 8 - 4 2 kcal/mol. Decomposition
through intramolecular or intermolecular H N 0 2 detachment, which was
proposed in some works, has not been confirmed. The energy o f the O —
N0 2 bond is noticeably lower than that o f the C — N 0 2 bond in similar
molecules. Thus, nitroesters are considerably less thermally stable in
comparison with the analogous nitrocompounds. Apparently, introduction
o f the second nitrate group into the position o f the same carbon atom would
result in such a reduction o f the energy o f the O — N 0 2 bond and molecule
stability that dinitrates cannot be separated in a pure form. Their existence is
only postulated in solutions, for instance, during R D X synthesis by
urotropine nitrolysis.
The simplest compound, methylnitrate, has the lowest value o f the
initial rate o f thermal decomposition in the gas phase. This rate increases as
any substitutes are introduced into this molecule, as well as at the
elongation o f the hydrocarbon chain. The adjacent nitrate group and the
other strong electronegative substitutes in /^-position exert the strongest
influence on the decomposition rate. Thus, the initial rate o f nitroglycerin

127
128 Nitroesters

decomposition is higher by a factor o f 10 than that for methylnitrate. O n the


whole, the rate variation as a function o f the nitrate molecule structure is in
a satisfactory agreement with the variation o f the inductive effects of
substitutes and their a constants. The decomposition rate o f nitroesters also
increases i f any electropositive substitutes are introduced, which is
determined by their attractive action on the N 0 - g r o u p . This results in a
2

more loosened structure o f the transition state, and the rate increases mainly
due to the entropy factor. The parameters for the rate constants o f
monomolecular decomposition and initial decomposition rates for the most
typical nitroesters in the gas and condensed phase are presented in Tables
9.1 and 9.2. In the condensed phase, the initial decomposition rate depends
on the free gas volume per 1 mole o f substance K , increasing with its 0
g

growth w h i c h is determined by the reversibility o f the initial stage and


variation o f the N 0 dissolubility.
2

Table 9.1. Kinetic parameters of the rate constants of monomolecular decompo­


sition of nitroesters in the gas phase

Nitroester log M s " ) 1


E, kcal/mol Refs.

Methylnitrate 15.7 40.5 [9.8]


Ethylnitrate 16.0 40.0 [9.8]
>7-Propylnitrate 16.5 40.0 [9.10]
/'so-Propylnitrate 16.5 40.5 [9.11]
ter/-Butylnitrate 15.9 40.2 [9.8]
Nitroglycol 14.3 35.7 [9.12]
Nitroglycerin 15.5 36.0 [9.7]

Table 9.2. Kinetic parameters of the initial decomposition rates of nitroesters in the
liquid phase

Nos. Nitroester log/:,, (s" ) 1


E, kcal/mol Refs.

1 Ethylnitrate 14.7 40.0 [9.9]


2 Ethyleneglycoldinitrate 14.5 38.8 [9.13]
3 1,2-Propyleneglycol-dinitrate 15.8 40.3 [9.13]
4 1,3-Propyleneglycol-dinitrate 14.9 39.1 [9.13]
5 1,3-Glycerin-dinitrate 16.8 42.4 [9.13]
6 Diethyleneglycol-dinitrate 16.5 42.0 [9.14]
7 Glycerintrinitrate 15.4 39.3 [9.13]
20.2 46.9 [9.7]
8 1,4-Butyleneglycol-dinitrate 15.1 39.0 [9.13]
9 2,3-Butyleneglycol-dinitrate 16.7 41.6 [9.13]
10 Dinitroxy-ethylnitramine (DINA) 16.5 41.5 [9.14]
11 Hexanitromannitol 15.9 38.0 [9.14]
12 Pentaerythritol tetranitrate (PETN) 15.8 40.0 [9.14]
(solution)
Thermal Decomposition and Combustion 129

In the gas phase, the decomposition rate is the maximal one. O n passing to
the liquid and then to the solid phase, it successively decreases, w h i c h is
natural for a homolytic decomposition o f the molecule into two radicals.
This decrease is determined by the necessity to overcome the intermolecular
forces on transition into the activated complex. Thus, for P E T N , the
transition from the liquid to the solid phase leads to the initial rate
decreasing by a factor o f 100, which makes the solid P E T N one o f the most
stable nitroesters [9.15].
The characteristic feature o f thermal decomposition o f nitroesters,
especially polynitroesters, is a very high level o f self-acceleration o f the
reaction. The value of this acceleration strongly depends on the
decomposition temperature and the magnitude o f V , decreasing with the 0
8

increase o f V 0
g
in contrast to the initial rate. A t T> 150°C and V 0
8
> 2 • 10 5

cm ,
3
decomposition proceeds almost without acceleration. If the
temperature and free volume decrease, this acceleration rapidly increases.
First o f a l l , the value o f self-acceleration is determined by the nitroester
structure. For an adjacent arrangement o f nitrate groups, as in the case o f
nitroglycerin and hexanitromannitol, the rate can increase by a factor o f 10 . 4

The process is separated into two stages: after the initial relatively slow
growth o f the rate, the period o f its fast acceleration starts and, during this
period, the rate increases proportionally to the square o f the product
pressure. The main product in the gas phase is N 0 . For a separate 2

arrangement o f nitrate groups under similar conditions, the rate increases


several times or by a factor o f 10. A pronounced acceleration is not
observed and N 0 2 is practically absent in the gas phase. The degree o f
decomposition, at which the change o f the stages proceeds, increases as V f (

goes up. If the self-acceleration is large, very often, the initial rate measured
turns out to be noticeably larger than its true value because, during the
heating o f the sample, a certain degree o f transformation has already been
attained and the observable rate includes the corresponding acceleration.
The main reason for self-acceleration is the development o f oxidative
and hydrolytic interaction o f the parent nitroester with the products o f its
decomposition: H N 0 , N 0 , and H 0 . Because o f the following equilibria,
3 2 2

which are attained in the gas phase:

3 N 0 + H 0 — 2HN0 +NO;
2 2 3 2N0 2 N 0 2 4 ) (9.1)

introduction o f any o f these components in the system results in an increase


o f the concentration o f all molecules. M a n y works have shown that
introduction o f these products gives rise to an increase o f the decomposition
rate o f nitroglycerin and other polynitrates with an adjacent arrangement o f
nitrate groups. In this case, the development o f acid hydrolysis o f the ether
bond accompanied by oxidation o f the alcohol formed plays the main role.
However, recently, it has been shown [9.16] that the rate constants o f
130 Nitroesters

nitroglycerin hydrolysis measured independently are too small to describe


the observable rates o f its hydrolysis within the autocatalytic stage. O n the
basis o f the detailed analysis o f the kinetics o f accumulation and
consumption o f the key intermediate compounds with the use o f the rate
constants o f the model reactions, it has been revealed that the main
interaction is the reaction o f nitroglycerin with N 0 , which leads to the 2 4

formation o f a nitrite group - O N O at the central carbon atom with the


following hydrolysis o f the O—NO bond. The rate constant o f this
interaction takes the form:

£ = 10 7 0
e x p ( - 1 7 0 0 0 / / v r ) , s" .
1

The rate o f hydrolysis o f nitrites is higher by several orders of


magnitude than that for nitrates. The reaction o f N 0 2 4 interaction with
nitroesters has the lower activation energy than N 0 2 detachment from the
molecule o f nitroester, and N 0 2 solubility and its dimerization decrease as
the temperature increases. Finally, the temperature dependence o f the self-
acceleration is very weak, and, for nitroglycerin and hexanitromannitol at
the stage o f fast acceleration, the rate practically does not depend on the
temperature [9.14]. Interaction o f nitrates with H N 0 3 and organic acids
gives a smaller, but noticeable contribution.
For hydrolysis o f nitroesters, the activation energy is also considerably
lower than that for N 0 2 detachment. Hydrolysis can develop from the very
beginning and its contribution grows as the temperature decreases because
the sample o f nitroester practically always contains a certain amount o f
impure water. Hence, below the temperature o f 8 0 - 1 0 0 ° C , even neutral
hydrolysis proceeds at a higher rate than thermal decomposition. This
results in the appearance o f a disturbance o f the straight line o f the
dependence o f log(w ) versus \/T. For nitroglycerin at 4 0 - 8 0 ° C and small
0

values o f V , the activation energy is approximately 17 kcal/mol [9.17],


0
g

and for cellulose nitrate it is 2 4 - 2 6 kcal/mol [9.18].


Recently, the influence o f high static pressures on the decomposition
mechanism o f nitroesters was studied [9.19] and it was revealed that within
the pressure range o f 0.4-0.8 G P a , the mechanism for the secondary nitrate
groups was changed: from O — N 0 2 bond breaking to H N 0 3 molecule
elimination.
B i n d i n g or elimination o f nitric acid, water, or nitric oxides from the
system should stop the acceleration o f the process for any mechanism o f
this reaction due to equilibria (9.1). This fact is used for stabilization o f
systems containing nitroesters. Diphenylamine is the most widely employed
stabilizer, capable o f binding both nitric acid and nitric oxides. Introduction
o f diphenylamine removes the self-acceleration o f the process up to the
moment o f its complete consumption. Then, the rate increases similarly to
nitroester decomposition without a stabilizer.
Thermal Decomposition and Combustion 131

Cellulose nitrate (Nc) is the most frequently used nitroester. This is the
basic part o f modern powders. It has a complex structure, a highly
developed surface, and represents a certain distribution o f three-, d i - , and
monocellulose nitrate units along the polymer chain. Thermal
decomposition o f cellulose nitrate is governed by all regularities o f
decomposition o f nitroesters. Some regularities are determined by a
polymer base o f the compound, alternative degree o f nitration, and by the
presence o f free hydroxyl groups. If the nitrogen content exceeds 4%,
cellulose nitrate decomposition proceeds with self-acceleration, which
rapidly increases as the degree o f nitration increases. In many studies, the
values o f the initial decomposition rate for pyroxylin (cellulose nitrate with
a nitrogen content o f 13.0-13.5%) were obtained, which are in satisfactory
agreement with each other. In the generalized form, the temperature
dependence o f the decomposition rate for V 0
8
= (3-6) • 10 c m takes the
3 3

form [9.20]:

w0 = 10 14 9 3
exp(-38000/y?r),s . M

In this case, as for other nitroesters, the initial decomposition rate


increases as the value o f V increases and the degree o f self-acceleration
8

decreases. It should be noted that to produce a sufficiently pure and


thermally stable cellulose nitrate, a rather time-consuming stabilization is
required, which is connected mainly with the necessity to eliminate acid
sulfoesters o f cellulose. Without this procedure, the decomposition rate can
be higher by several orders o f magnitude.
A s the degree o f cellulose nitration is changed within a wide range o f
nitrogen content, it was found [9.21] that the decomposition rate per one
nitrate group slightly increased as the content o f these groups went up. The
comparison o f the decomposition rates o f cellulose nitrate with various
degrees o f substitution should be carried out for equal V per one nitrate 0
8

group. Hence, for highly nitrated samples, the free volume has to be higher.
The decomposition rate o f the nitrate group located near the primary carbon
atom C is lower by a factor o f 14 than that located near the secondary C
6 2

and C carbon atoms. A c c o r d i n g to [9.17], these rates are determined by the


3

following ratio k ^ : k : k ^ - 1: 3.6 : 6 , but the necessity o f variation o f


c c c

the V0
8
value was not taken into account. Such variation o f the rate
constants is in perfect agreement with the inductive influence o f substitutes;
in this case, it is the adjacent nitrate group.
The initial decomposition rate o f the real cellulose nitrate samples
strongly depends on the presence o f water impurities and various oxidized
groups in them, which are formed during cellulose nitration and
stabilization o f the product obtained. This influence can be very significant,
especially at low temperatures. Thus, for a humidity o f 1%, below 7 0 ° C ,
hydrolysis occurs at a higher rate than N 0 detachment due to the presence
2
132 Nitroesters

o f free hydroxyl groups. Such an amount o f water cellulose nitrate can


easily be adsorbed from open air.
The decomposition rate o f solid cellulose nitrate is several times lower
than that for nitroesters with a similar structure. In solutions, the
decomposition rate o f cellulose nitrate increases.
Self-acceleration o f cellulose nitrate decomposition depends on a l l
factors that also influence other nitroesters. Self-acceleration increases
rapidly with the increase o f the degree o f nitration because the number o f
units with an adjacent arrangement o f nitrate groups increases. In the
samples o f cellulose nitrate with a nitrogen content up to 4%, self-
acceleration is almost zero. O n the whole, the increase o f the rate in
cellulose nitrate is noticeably smaller than, for instance, in nitroglycerin
under similar conditions. Even for nitrogen content higher than 13.5%, the
maximal rate exceeds the initial rate by no more than a factor o f 50 [9.18]
which, obviously, is determined by the relatively low solubility o f N 0 and 2

N 0
2 4 in cellulose nitrate.

REFERENCES

9.1. R . Boschan, R . T . M e r r o w , R . W . van Dolan, Chemical Reviews,


55: 4 8 5 - 5 1 0 ( 1 9 5 5 ) .
9.2. K . K . Andreev, Thermal Decomposition and Combustion of
Substances, M . : Nauka, 14-47 (1966).
9.3. T . Yoshida, N . Tsmagari, K . Namba, Journal of Industrial
Explosives Japan Society, 35: 325-338 (1972).
9.4. K . K . K l i m e n k o , Thermal Stability and Structure of Polynitro- and
Nitratecompounds, in Combustion and Explosive, M . : N a u k a ,
585-593 (1977).
9.5. R . A . Fifer, Fundamentals of Solid-Propellant Combustion, 90:
177-237(1984).
9.6. B . A . Lurie, B . S. Svetlov, Izv. Vyssh. Uchebn. Zaved., Khim.
Khim. Technol, 10(12): 1308-1314 (1967).
9.7. C h . E . Waring, G . Krastins, J. Phys. Chem., 74(5): 999-1006
(1970).
9.8. L . Batt, R . T . M i l n e , Int. J. Chem. Kinet., 8(1): 59-81 (1976); 9(4):
725-737(1977).
9.9. B . A . Lurie, B . S. Svetlov, Theory of Explosives, M . : Vysshaya
Shkola, 40-51 (1967).
9.10. G . D . Mendenhall, D . M . Golden, S. W . Benson, Int. J. Chem.
Kinet., 7(5): 725-737 (1974).
9.11. J. Jullien, J . M . Pechine, M . A . Sadek, Prop. Explos. Pyrothechn.,
8: 99 (1983).
9.12. K . K . Andreev, A . P. Glazkova, N . D . Maurina, Zh. Fiz. Khim., 32:
1726-1738 (1958).
Thermal Decomposition and Combustion 133

9.13. A . G . Afanas'ev, B . A . Lurie, B . S. Svetlov, Theory of Explosives,


M . : Vysshaya Shkola, 63-73 (1967).
9.14. B . A . Lurie, B . S. Svetlov, Theory of Explosives, M . : Vysshaya
Shkola, 51-63 (1967).
9.15. K . K . Andreev, B . I. K a i d y m o v , Zh. Fiz. Khim., 35(12): 2 6 7 6 -
2687 (1961).
9.16. B . A . Lurie, L . G . Afanas'ev, B . S. Svetlov, Chemical Physics of
Combustion and Explosion. Kinetics of Chemical Reactions,
Chernogolovka, 55-58 (1992).
9.17. B . S. Svetlov, V . V . Gorbunov, Theory of Explosives, M.:
Oborongiz, 190-197(1963).
9.18. H . N . Volltraner, F. Fontijn, Comb, and Flame, 41: 313-324
(1981).
9.19. D . L . N a u d , K . R. Brower, J. Org. Chem., 57(12): 3303-3309
(1992).
9.20. B . A . Lurie, B . S. Svetlov, Chemistry and Reactivity of Cellulose
and its Derivatives, Abstract o f Papers, Cholpon-Ata, 1273-1279
(1991).
9.21. E . A . Panina, V . A . Rafeev, T. V . Sorokina, Y u . I. Rubtsov, N . V .
Chukanov, Izv. Akad. Nauk SSSR, Ser. Khim., 6, p. 1273-1279
(1990).
Chapter 10

COMPOUNDS WITH MIXED


FUNCTIONS

In order to increase the melting point or energy content o f matter, quite


often compounds are synthesized, whose molecules contain several
different energetic groups, such as, for example, nitramine, d i - and
trinitromethyl, azide, etc. If there are no chemical, inductive, or any other
interactions among these groups, the stability o f such compounds with
mixed groups can be predicted, at least in the liquid state, from the known
parameters o f decomposition o f monofunctional compounds. The deviation
o f the actual stability from the predicted one can suggest the mutual
influence o f these groups.
The characteristic values o f the rate constants (at 150°C) and Arrhenius
parameters o f the decomposition for the main energetic fragments in the
compounds o f aliphatic series are presented in Table 10.1. Using these
characteristic parameters, one can estimate the expected rate and, in many
cases, with a large probability indicate the region o f the primary
decomposition o f the compounds with mixed groups. The combinations o f
pyramidal nitramine and nitrate groups, as well as the plane nitramine and
azide groups, are hardly distinguishable.
The compounds in which different functional groups are positioned
near one carbon atom are o f special interest. A n example o f such
compounds is a-(difluoroamino)-polynitro-compounds RCNF (N0 ) .
2 2 2 In
these compounds, the C — N 0 2 bond has the lowest strength and the N F 2

group, as the a-substitute, influences the D ( C — N 0 ) value with nearly the 2

same efficiency as a nitrogroup. Hence, the R C N F ( N 0 ) compounds have 2 2 2

the same stability as their analogs (see Table 10.2). The N F C ( N 0 ) , 2 2 3

FCNF CN0 ) ,
2 2 2 and CH CNF (N0 )
3 2 2 2 compounds, similar to
polynitroalkanes, are decomposed in the gas and liquid phases at the same
rate according to the radical mechanism. The first stage o f this mechanism
is N 0 2 detachment. These compounds do not exhibit any features o f the

135
136 Compounds with Mixed Functions

heterolytic mechanism, w h i c h is typical o f difluoroamines in the liquid


phase.

Table 10.1. Characteristic values of the kinetic parameters of decomposition of


monofunctional compounds

Group log/4 (s s" (150°C)


1

kcal/mol (150°C)

-CH C(N0 ) 2 2 3
43.0 17.2 1.0- 10" 5
1.0
-CH C(N0 ) CH - 2 2 2 2 47.5 17.1 3.7- 10 -8
0.004
-CH CF(N0 ) 2 2 2 47.5 17.0 2.9 • 10~ 8
0.003
-CH N ( N 0 ) C H - planar
2 2 2 41.0 14.5 2.0- IO -7
0.02
-CH N(N0 )CH - 2 2 2 38.0 14.5 7.4 • 10" 6
0.7
pyramidal
-CH N 2 3
39.0 14.4 1.8- 10 -6
0.18
-CH C(NF ) CH - 2 2 2 2 47.5 16.0 2.9- 10~ 9
0.0003
-CH NF 2 2 26.2 8.3 1.8- IO" 6
0.18
-R CNF * 3 2 30.0 9.0 3.2 • 10~ 7
0.03
-CH ON0 2 2 39.0 15.0 7.1 • 10" 6
0.7

* The NF -group is located at the third carbon atom or shielded by a voluminous


2

substitute R.

Table 10.2. Kinetic parameters of thermal decomposition of a- (difluoroamino)-


polynitrocompounds [10.1 ]

Compound State r,°c E, log/4 k, s-


1
D,2

kcal/mol (s- )
1
(100°C) kcal/mol

F NC(N0 )
2 2 3 Gas 110-175 38.5 16.0 3 • IO -7
0.3 44.9
F NC(N0 )
2 2 3 Liquid 80-90 38.7 16.4 5 • 10~ 7
1.3 -
FC(N0 ) - 2 2 Gas 170-220 43.0 16.2 1 • IO" 9
1.4 47.7
NF 2

CH C- 3 Gas 170-190 40.2 16.0 3 • 10~ 8


4.3 49.7
(N0 ) NF 2 2 2

CH C- 3 Liquid 120-140 40.2 16.0 3 • 10" 8


3.0 -
(N0 ) NF 2 2 2

[NF (N0 ) CCH 2 2 2 2 Liquid 130-180 39.5 15.4 2- IO" 8


1.2 -
-0] CH 2 2
3

w is the ratio of the rate constants of R C ( N 0 ) N F


l
2 2 2 decomposition and
corresponding polynitrocompounds R C ( N 0 ) . 2 3

2
D is the energy of dissociation of the C — N F bond [10.2]. 2

3
Unpublished data of the authors
T a b l e 10.3, The kinetic parameters of decomposition of compounds with mixed groups
No. Compound State r,°c E, logA w l
Refs.
kcal/mol (O (150 °C)
1 (N02) CCH2CH COOCH2-C(N02)2F
a 2
Melt 130-193 41.5 16.7 2 • IO -5
2 [10.5]
2 F(N0 ) CCH CH COO-CH C(N0 )3
2 2 2 2 2 2
Melt 130-193 41.5 16.5 1 • 10 -5
1 [10.5]
3 CH3N(N0 )CH2C(N0 )3 2 2
Melt 90-140 40.4 16.7 7 • IO -5
7 [10.6]
CH3N(N0 )CH2C(N02)3 2
Solution in TNT 130-180 40.3 16.8 7 • IO -5
7 [10.6]
4 [(N0 )3CCH2]2NN0
2 2
Melt 110-150 36.8 15.6 4 • 10" 4
38 [10,6]
[(N0 )3CCH2]2NN0
2 2 Solution in TNT 110-165 36.1 15.0 3 • IO" 4
33 [10.6]
5 [(NO^CCHzNNOzCH^ Solution in TNT 130-180 40.7 16.8 6 • 10 -5
6 [10.6]
6 FC(N0 ) CH N(NO )CH3 2 2 2 2
Solution in TNT 175-210 40.5 14.9 9 • 10" 7
4.5 [10.6]
7 [FC(N0 ) CH ] NN0 2 2 2 2 2
Solution in TNT 160-210 39.8 14.9 2 • 10"* 10.0 [10.6]
8 [FC(N0 ) CH2NNO CH2] 2 2 2 2
Solution in TNT 160-210 40.0 15.3 3 • 10 * -
15.0 [10.6]
9 NjCH fN(N0 )CH2]4N3 2 2
Solution in TNT 120-145 36.3 14.6 7 • 10" 5
40.0 2
10 (0 N) CCH N(N0 )CH Solution in TNT 170-210 38.0 14.5 I • 10^ 7.0 2
1 \
2 2 2 2 2

a.
CH N(N02)CH C(N02)2
2 2 O
2
o
11 FC(N0 ) (CH )2-C(NF )2CH3 2 2 2 2
Melt 150-200 46.7 16.6 3 • 10" 8
1.0 3
12 Melt 110-140 43.2 17.8 3.0 2 a*
(N0 ) C(CH >-C(NF ) CH3
2 3 2 2 2 3 • IO -5
c
13 (N0 ) C(CH2)200CCH2-CH(NF )CH NF2
2 3 2 2
Melt 110-140 43.2 17.9 4 • 10~ 5
4.0 2
14 / * ~ \ Solid 115-130 32.5 17.3 3.2 10 5
[10.7]

H
02N-N N-NO2

V
The places of a primary decomposition are bolded.
1
w is the ratio of the observable rate constant to the value predictedfromTable 10.1.
2
Data of the authors.
138 Compounds with Mixed Functions

In the case o f a-combination o f various substitutes, interaction among them


is o f stable character and, hence, the whole fragment acquires the properties
o f a monofunctional group with constant characteristics. Another example
o f this group are azidonitrocompounds R C ( N 0 ) 2 N 2 3 [10.3], which, by their
stability, are approximately equal to the R C ( N 0 ) 3 compounds [10.3, 10.4].
2

The data on the decomposition o f the actual compounds with mixed


groups are presented in Table 10.3. Based on the kinetic predictions using
Table 10.1, we indicated the parts o f molecules where the decomposition
starts. In the majority o f cases, the assignment o f the observable rate to a
certain structural group does not present difficulties. In 3, 5, 7, and 8
compounds presented in Table 10.3, the decrease o f stability o f the
nitromethyl and nitramines groups is observed due to their mutual inductive
influence. In compound 4, these interactions are strengthened, apparently,
due to conformational limitations and the joint influence o f the nitramine
group (the field effect), which is similar to those observed for
polynitrocompounds o f a pentane series [10.8]. The /?-nitramine group also
essentially influences the decomposition o f the azide group (compound 9 in
Table 10.3). The furazan ring in a-position (compound 14 in Table 10.3)
produces a very strong effect on the decomposition o f the nitramine group.
The discrepancy with the prediction is so considerable that it cannot be
explained within the framework o f the mechanism with N — N 0 2 bond
breaking. Probably, in this case, a totally different mechanism is realized,
w h i c h is connected with the intramolecular rearrangement.
Another case o f a very strong interaction o f the groups was observed
during the decomposition o f l-(azidomethyl)-3,5,7-trinitro-l,3,5,7-tetraaza-
cyclooctane 15 [10.9]:

\
N0 2

For a heating rate o f 2 ° C per min, the peak o f an intensive decomposition o f


this substance is observed at 140°C. Even aromatic azides do not have such
a low thermal stability. The decomposition mechanism o f compound 15 is
not clear. A heterolysis o f the C — N 3 bond is possible.
In the solid phase, the relative reactivity o f the group can be changed
considerably, and its prediction from the kinetic parameters becomes
impossible. In certain cases, information about the decomposition
mechanism o f compounds with mixed groups can be obtained with the help
o f the FTIR-registration o f the primary and secondary gas phase and solid
decomposition products during high-speed heating (up to 100°C per
second) o f the substance. Such unique investigations are described by T.
Thermal Decomposition and Combustion 139

B r i l l in [10.3, 10.4, 10.9-10.11]. In particular, by this method it was shown


that the decomposition of solid l,7-diazido-2,4,6-trinitro-2,4,6-
triazaheptane N CH —N(N0 )—CH —N(N0 )—CH —N(N0 )—CH N
3 2 2 2 2 2 2 2 3

occurs in complete agreement with the prediction, which takes into account
the inductive influence o f the /?-nitramine group mainly by splitting o f the
azido-group [10.11].

REFERENCES

10.1. V . N . Grebennikov, G . M . N a z i n , and G . B . Manelis, Izv. Akad


Nauk SSSR, Ser. Khim., 4, p. 649-651 (1995).
10.2. V . I. Pepekin, Khim Fizika, 13: 42-51 (1994).
10.3. J. T. Cronin and T. B . B r i l l , Appl. Spectrosc, 43: 649-653 (1989).
10.4. R. Subramanian and T. B . B r i l l , Prop. Explos. Pyrotechn., 15:
187-189 (1990).
10.5. V . G . Matveev, L . D . Nazina, G . M . N a z i n , D . A . Nesterenko, and
L . T. Eremenko, Izv. Rus. Akad Nauk, Ser. Khim., 12: 2455-2458
(1998).
10.6. B . L . Korsunskii, V . G . Matveev, L . D . Nazina, and G . M . N a z i n ,
Izv. Rus. Akad Nauk, Ser. Khim., 2: 259-264 (1998).
10.7. D . G . Patil and T. B . B r i l l , Thermochim. Acta, 235: 225-230
(1994).
10.8. V . N . Grebennikov, V . G . Prokudin, and G . M . N a z i n , Izv. Akad
Nauk SSSR, Ser. Khim., 12: 2822-2824 (1984).
10.9. T. B . B r i l l , R. J. K a r p o w i c z , T. M . Haller, and A . L . Rheingold, J.
Phys. Chem., 88: 4138-4143 (1984).
10.10. T. B . B r i l l , D . G . Patil, J. Duterque, and G . Lengelle, Combust, and
Flame, 95: 183-190 (1993).
10.11. Y . O y u m i and T. B . B r i l l , J. Phys. Chem., 91: 3657 -3661 (1987).
Chapter 11

GENERAL REGULARITIES OF
THERMAL DECOMPOSITION OF
ONIUM SALTS. NITRIC AND
PERCHLORIC ACIDS, DINITRAMIDE

11.1 General Regularities


The main distinctive feature o f onium salts, formed in the process o f a
proton transfer from the acid to the base, is the possibility o f a reverse
equilibrium proton transfer from the cation to anion with the formation o f
the initial acid and base molecules and their subsequent redistribution
between the condensed and gas phases. The general scheme o f such an
equilibrium and subsequent thermal transformations takes the form:

(BH A-) d^ B + + c c
HA C
BH —products +

\\K {B) g l^s(HA)


A" products
B G
HA g

B products

* V B *HA/ a A
B H Y HA"^ ° p r d u C t S
( ] 1 1
->
l

^ P B P H A B +HA-Products
d *wm BH +
+A ^ B H +A
P =K (B)a
B g B \

P A^ (HA)a A
H g H
p r
° d u C t S

PtotarPB + PHA

Here, B is base, H A is acid, K° , eq £ ( H A ) , and K (B) are the constants


g g

of corresponding equilibria in the condensed (c) and gas (g) phases, a , a A , B H

aBH+A a r e m e
activities o f the components in the condensed phase, P , HA P%
are the equilibrium pressures in the gas phase. Due to a lower degree o f the

141
142 General Regularities of Thermal Decomposition of Onium Salts

inverse dissociation o f a salt ( £ ° e q is small) and because the properties o f a


salt are almost unchanged, one can accept that a + A ~ = 1, i f the state o f a BH

pure salt is accepted as the standard state for the activity.


In principle, both ions o f the parent salt and acid and base molecules
formed from them can participate in the elementary stages o f thermal
decomposition. However, as a rule, the decomposition rate o f onium salts is
higher by several orders o f magnitude than this rate for the corresponding
salts o f metals or salts on the basis o f N H 3 and a stable acid (for instance,
H S 0 ) . The decomposition rates o f the ions forming the salt are too low.
2 4

The formation o f radicals by means o f electron transfer from an anion to


cation in onium salts has a high energetic barrier. A s a rule, such salts have
a proton but not electron conductivity. Hence, we can neglect the
contribution of radicals in thermal decomposition. Thus, thermal
decomposition o f onium salts occurs through the thermal transformations o f
acid and base molecules formed in equilibria (11.1.1).
Direct measurements o f the rates o f proton transfer in the liquid phase
[11.1] already at room temperature give a relaxation time o f 1 0 — 1 0 -5 -10
s.
Hence, equilibria are attained very rapidly and in all processes o f thermal
decomposition o f onium salts studied by the ordinary methods, the
concentrations o f the base and acid can be considered as equilibrium ones.
Although the amount o f such molecules at equilibrium is small, in the
majority o f cases the decomposition o f a salt proceeds through the thermal
transformation o f acid and base molecules because their decomposition
rates are considerably higher than those for the corresponding cations and
anions [11.2]. The decrease o f concentration o f such molecules due to
decomposition is rapidly compensated by an additional proton transfer on
the corresponding number o f anions. A t the limiting stage o f the reaction,
one or both products o f dissociation can participate. In the general form, the
kinetic equation for the reaction rate, which is written for 1 mole o f the
parent salt, takes the form:

_ P __._.M_A_. = £ C H A + /: B . (11.1.2)
M dt

Here, p and M are the density and molecular weight o f the parent salt, C A H

and C are the concentrations o f H A and B , k


B HA and k are the rate constants
B

of thermal transformations o f H A and B (these rate constants can differ


considerably from the corresponding values for pure compounds due to the
influence o f ions o f the parent salt), n and m are the reaction orders with
respect to H A and B , correspondingly. If the resulting system can be
considered as an ideal solution, we w i l l have the following relationships:
3HA =
NHA, B A =
NB, where N HA and N Q are the mole fractions o f the
components. In this case, N HA ~ ^ H A and N B ~ n , where n
B UA and n B are the
number o f moles o f the dissociation products per 1 mole o f the parent salt,
Thermal Decomposition and Combustion 143

because n , HA nB « n +
BH A and Zn in the system is close to unity. However,
x

for such a substitution, one should take into account that the N AH and N B are
dimensionless values and the dimension o f n HA and n B is mole. This allows
us to estimate the n HA and n B values based on equilibrium (11.1.1) and the
decomposition rate based on equation (11.1.2). For instance, for the salt
containing an excess o f the base / B and in the case o f absence o f the
components transition into the gas phase, the initial decomposition rate
takes the form:

/ \m~\
dn B H A ~
eq JL. (n + l )°
+

dt
HA B B

M M

K + /
B) »E = K -

If / >> n , one can neglect the contribution o f n


B B B

The dependence o f the decomposition rate o f an onium salt on the


variation o f its stoichiometry due to the introduction o f excess concentration
of acid or base is shown in F i g . 11.1.1.

Fig. 11.1.1. Variation of the decomposition rate of an onium salt as a function of the
stoichiometry of its composition: (0) the point of equivalence of an acid and base in
the salt, (/) the decomposition through an acid, (2) the decomposition through a
base, and (3) the decomposition of both products of dissociation.

Curve 1 corresponds to a salt decomposition in which this process proceeds


through an acid decomposition. The rate monotonically decreases at the
introduction o f an excess o f a base and, in principle, it can be reduced up to
a very small value, until the rate o f thermal transformations o f ions w i l l
become essential. Curve 2 demonstrates the decomposition rate o f a salt,
whose destruction proceeds through a base. In this case, the rate decreases
at the introduction o f an excess o f acid. Curve 3 corresponds to the case
144 General Regularities of Thermal Decomposition of Onium Salts

when both products o f dissociation are involved in the decomposition. In


this case, the decomposition rate has its minimal value at

O n deviation from this value, the decomposition rate increases.


The equilibrium vapor pressure o f the salt is also rapidly attained. This
pressure consists o f the vapor pressures o f the acid and the base. The ions
practically do not pass into the gas phase. In general, these pressures differ
considerably due to the differences in the K (HA) and K (B) values. The gas
g g

phase is enriched with a more volatile component and the condensed phase
is enriched with a less volatile component. A t evacuation o f the gas-phase
products o f dissociation, sublimation o f the onium salt w i l l occur with its
following condensation on the cold parts o f the system. A t the beginning o f
evacuation, a more volatile component w i l l be predominantly removed from
the system, and its content in the condensed phase and the equilibrium
pressure w i l l decrease. For the other component, these values w i l l grow
until the pressures o f both components are equal. After that, a pure salt w i l l
sublimate. Nevertheless, in the condensed phase, this salt w i l l be enriched
with the less volatile component. So, during evaporation, almost all
ammonium salts are enriched with an acid, because the acid decomposition
determines the decomposition rate o f ammonium salts. Their vacuum
evaporation results in a considerable increase o f the thermal decomposition
rate. Since at steady-state sublimation P HA = P, B taking into account
equilibria (11.1.1), we w i l l have:

^HA A
HA
£ (HA)
8

For a pure salt without sublimation we w i l l have the following relation:

In general, one has to distinguish between sublimation and thermal


decomposition. Decomposition should involve irreversible stages.
Sublimation cannot be considered as a kind o f thermal decomposition. After
vapor condensation, the parent salt forms again.
In principle, the reactions o f thermal decomposition o f acid and base
are also possible in the gas phase, but for the compounds under
consideration, under the temperatures that are usually utilized, the
contribution o f the gas phase reactions is small. The total rate is determined
Thermal Decomposition and Combustion 145

by the reactions in the condensed phase, though as the temperature


increases the contribution o f the gas-phase reactions also increases. The
kinetic scheme for the rate calculation is considered in detail in Chapter 12
for various conditions o f thermal decomposition using the example o f
ammonium nitrate. In this case, the kinetic regularities are rather more
complicated in comparison with the decomposition o f the nitrocompounds
considered above. The decomposition rate is determined not only by the
rate constants o f decomposition o f the acid and base, but also by their
equilibrium concentrations. Since the equilibrium state can be easily shifted
by introduction o f an excess acid or base, this offers new possibilities for
regulation o f the decomposition rate.
The quantitative calculation o f the rate can be carried out only i f the
value o f is known. For onium salts, the value o f A ^ e q for the equilibrium
o f the solid or liquid salt with the gas-phase products over the salt can be
calculated by the methods o f classical thermodynamics. Usually, these
pressures are small and gas can be considered to be the ideal that allows us
to express the equilibrium constant through the partial pressures and to
carry out the calculation based on the known thermodynamic relationships:

-RT\n(K*) = AG^ =(AH- TAS) , K!=P P . R HA

\ eq / ieact \ /react ' eq B HA

For the steady-state sublimation, we have P T O I A \ = P B + P A> H


=
PUA and
AH and AS o f the reaction are calculated from the values o f enthalpy o f
formation and entropies for the salt in the initial state and an acid and base
in the gas phase. When the vapor pressure over the onium salt can be
measured experimentally, its value is in good agreement with the calculated
value which confirms evaporation o f the salt through its dissociation. It is
rather more difficult to calculate the equilibrium constant K° eq in the liquid
or solid phase because it is difficult to determine the thermodynamic
properties o f the dissociation products dissolved in the parent salt or
adsorbed on the surface and defects o f its crystal lattice. These properties
differ considerably from the thermodynamic properties o f pure components
and it is difficult to calculate them. Today, the only way to determine A ^ e q

seems to be independent measurement o f the decomposition rate, the rate


constants k HA and k , and the calculation o f the equilibrium concentrations
B

and A*e based on relations (11.1.1) and (11.1.2). Examples o f such


q

calculations w i l l be presented on consideration o f the particular compounds.


However, for many onium salts, the A ^ q values are not known. So, from
calculation, one can obtain only the type o f the dependence o f the
decomposition rate on the acid or base excess.
Since thermal decomposition o f many onium salts (nitrates and
perchlorates) is determined by the decomposition rate o f the acids formed
during their dissociation, it is worthwhile to start the consideration o f these
salts from the study o f the kinetic regularities o f decomposition o f nitric and
146 General Regularities of Thermal Decomposition of Onium Salts

perchloric acids and, also, thermal decomposition o f dinitramide, w h i c h


represents a new group o f onium salts. Decomposition o f perchloric and
nitric acids is o f importance in itself. They can be used as the components
o f rocket propellants and are widely used in chemical and related industries.

11.2 Nitric Acid

Thermal decomposition o f nitric acid represents a reversible endothermic


reaction and almost quantitatively proceeds according to the following
equation:

H N 0 3 = N 0 + 0 . 5 H O + 0 . 2 5 O - 7 . 8 kcal/mol.
2 2 2 (11.2.1)

In a closed system, after the equilibrium pressures o f N 0 2 and 0 2 are


attained, the decomposition stops, although a major part o f H N 0 3 molecules
is preserved in the system. The attainable degree o f decomposition and the
heat o f decomposition Q depend on the experimental conditions, mainly on
T

the temperature and free volume V 0


G
o f the system. The kinetics o f thermal
decomposition o f nitric acid was repeatedly studied [11.3], mainly near its
100% concentration. The most detailed study o f the decomposition o f an
aqueous solution within the range o f concentrations o f 4 1 . 7 - 9 9 . 3 8 % H N 0 3

at temperatures 6 0 . 8 - 2 0 0 ° C was carried out in [11.4] mainly with the help


o f the heat release rate measurements. The reaction was studied inside
sealed and completely temperature-controlled glass ampoules with an
internal volume o f approximately 2 c m , with the use o f D A K - c a l o r i m e t e r s
3

[11.5, 11.6]. This technique has also been used in all other studies o f heat
release kinetics described below. This experimental technique made it
possible to hold all o f the volatile products in the reaction zone and to study
the aqueous solutions o f H N 0 3 at temperatures considerably greater than

HNO3
their boiling temperatures. Later, the kinetics o f decomposition o f a 100%
was studied in an open system at 0 - 6 0 ° C [11.7]. Decomposition
proceeds with considerable self-acceleration and all kinetic parameters are
complex functions o f the temperature, acid concentration, and free volume
o f the reaction vessel related to 1 mole o f H N 0 3 (Kq ). In the general form,
8

the dependence o f the decomposition rate on the degree o f conversion is


described by the following expression:

]_dQ
* '7(1-7), 7= J- (11.2.2)
dt a + brj
2

The rate constants o f a non-catalytic stage k =k fa and catalytic stage k


x 0 2

decrease as the concentration o f H N 0 3 decreases and k x decreases more


Thermal Decomposition and Combustion 147

rapidly. In the range o f concentrated solutions, the rate is mainly determined


by the first term. It was impossible to determine the value o f k 2 in this
region. The role o f k 2 increases as the solution is diluted. For the solutions
of 72% H N 0 3 and for lower H N 0 3 concentrations, the catalytic stage
provides the main contribution and the first term is essential only at the
beginning o f the process. If the solution is diluted, the rate rapidly
decreases, w h i c h requires an increase in the temperatures used. The
comparison o f the observed decomposition heat Q e with the calculated
decomposition heat of HN0 3 shows that the attainable degree of
decomposition ( c w ) changes within the limits o f 0.002-0.2. Based on
equilibrium (11.2.1) and taking into account the distribution o f N 0 2 and 0 2

between the gas and liquid phases, ionization o f P i N 0 3 molecules by water


and neglecting the passage o f H 0 and H N 0 2 3 molecules into the gas phase,
under an assumption that the reactive mixture represents an ideal solution,
one can derive the dependence o f the equilibrium degree o f decomposition
#max =
Qt / Qmax in the mixtures o f 1 mole o f H N 0 3 and n^Q moles o f
H 0 from the following expression:
2

[(i-r o) < o
H2 + 0-5« niax ] ^W
05 5 2S

0-r„No,-<0 W) 1 2 5
K (N0 )K (o )
g 2 g 2

Here, K N is the equilibrium constant o f H N 0 3 decomposition expressed


through the molar fractions o f the components, /H 0>/HNO 2 3
A R E T N E

ionization degrees o f the H 0 and H N 0 2 3 molecules in solution, A ^ f N O ^ ,


K (Oi) are the Henry constants o f the corresponding gases. Variation o f
g

<2m correspondingly changes the values o f k\ and k . Since the observable


ax 2

rate o f thermal decomposition characterizes the reactivity o f the whole


system, it is advisable to relate the experimental values o f k\ and k 2 to the
total amount o f acid in the system: k\ = k\ k =k
2 2 Om . ax

Table 11.2.1. The concentration and temperature dependences of the rate constant
of thermal decomposition of nitric acid

HNO3, wt % AT, °C k, s 2

41.70 179-200 5.2- 10 exp(-48600/RT)


16
1.9- 10 ,6
exp(-44000/RT)
49.84 165-197 1.4- 10 exp(-47000/RT)
17
1.7- 10
15
exp(-40700/RT)
57.80 163-197 6.2 • 10 expH4500/RT)
16
6.0- 10 ,3
exp(-36800/RT)
69.57 141-179 2.6- 10 exp(-38300/RT)
14
1.4 • 10 13
exp(-35200/RT)
76.30 140-167 7.7- 10 exp(-36000/RT)
13
4.7- 10 l2
exp(-33300/RT)
89.07 101-129 4.8- 10 exp(-26300/RT)
9
-
99.38 60-106 4.5 • 10 exp(-25600/RT)
10
-
100.0 0-60 9.0- 10 exp(-32000/RT)
12
-
148 General Regularities of Thermal Decomposition of Onium Salts

The dependences o f these constants versus \IT for the acids o f various
concentrations are presented in Figs. 11.2.1 and 11.2.2. The concentration
and temperature dependences o f k and k for V^Q = 1070 c m are presented
x 2
3

in Table 11.2.1.
Introduction o f N 0 into H N 0
2 3 (57.8% solution) in the amount o f 1.9%
at 166.5°C leads to the increase o f k by a factor o f 1.7 and introduction o f
x

N0 2 in the amount o f 2.7% results in the increase o f k\ by a factor o f 2.7. In


this case, self-acceleration o f the process is completely eliminated. One can
observe the decrease o f the activation energy and a very strong increase o f
the rate as the concentration o f H N 0 3 increases.
The mechanism o f H N 0 3 decomposition in the liquid phase was first
proposed in [11.3] and was later developed in detail in [11.4, 11.8]. The
observable kinetic regularities were explained based on the analysis o f the
real composition o f the solutions studied. In addition to the ordinary
hydration o f ions, the following set o f equilibrium reactions was considered:

2HN0 — N0 3 2
+
+ N0 "+ H 0 3 2 (11.2.3)

N0 2
+
+ N 0 3 N 2 0 5 (11.2.4)

HN0 3 + H 2 0 ^ H 0 +N0 " 3


+
3 (11.2.5)

HN0 4- n H 0 — H N 0 - n H 0
3 2 3 2

2HN0 +N 0 3 3 " ^ (HN0 ) .N0 " 3 2 3

Fig. 11.2.1 Fig. 11.2.2

Fig. 11.2.1. Dependence of log(/c,) on the temperature and concentration of H N 0 3

during decomposition of its solutions [ H N 0 ] , mol%: (7) 17.0, (2) 22.1, (3) 28.1, (4)
3

38.4, (5)48.0, (6) 70.0,(7) 97.9.

Fig. 11.2.2. Dependence of \og(k ) on the temperature and concentration of H N 0


2 3

during decomposition of its solutions [ H N 0 ] , mol%: (/) 17.0, (2) 22.1, (3) 28.1, (4)
3

38.4, (5) 48.0.


Thermal Decomposition and Combustion 149

The solutions o f nitrates o f alkali metals are stable over the temperature
range covered in the experimental measurements. The N 0 ~ ion is quite 3

stable and the reaction o f decomposition can occur only through the H N 0 3

and N 0
2 5 molecules. The rate o f monomolecular decomposition o f H N 0 3 in
the gas phase was studied in several works, for instance, in [11.9, 11.10].
The breaking o f the least strong bond H O — N 0 2 requires 47 kcal/mol and
decomposition proceeds according to the following scheme:

HN0 3 — OH+ N 0 2

O H + HNO3 = H 0 + N0
2 3

N0 3 + N 0 2 = N O + O2 + N 0 2

N O + NO3 = 2N0 2

N 0 3 + N 0 3 = 2N0 2 + 0 .
2

Unfortunately, the quantitative data for the pure gas-phase monomolecular


reactions are absent because the measurements are complicated by the
heterogeneous decomposition o f H N 0 3 on the surface o f the reaction vessel.
Thus, in [8], only an estimate o f the activation energy o f decomposition was
obtained within the range o f 4 0 - 5 0 kcal/mol. A t E = 47 kcal/mol and
ordinary pre-exponential factor 10 , this rate is lower than the actual rate o f
13

decomposition o f anhydrous H N 0 3 in the liquid phase by more than nine


orders o f magnitude. In [10], the shock tube study over the temperature
range 800-1200 K gave the following rate constant o f H N 0 3 decomposition
in argon:

k = 10 6 2
( £ / RT)* e x p ( - £ / RT), c m / ( m o l • s);
0 0
3
E = 47.3, kcal/mol
0

For the description o f the decomposition o f H N 0 and its solutions, a 3

mechanism is accepted that involves the equilibrium stages (11.2.3) and


(11.2.4) with the following monomolecular homolytic N — O bond breaking
in the N 0 molecule
2 5

N 0 2 5 — NO2+NO3 (11.2.6)

or an electron transfer between the two ions

N0 2
+
+ N0 " — 3 N0 + N02 3 (11.2.7)

If decomposition proceeds through reaction (11.2.6), taking into account


equilibria (11.2.3) and (11.2.4), one can obtain the following relationship:
150 General Regularities of Thermal Decomposition of Onium Salts

dC u

- 2k C
dt
b N I 0

/l=0 (11.2.8)
2k6 K3 K (C )4 m0}

( H,o)
C
a i

where k 6 is the rate constant o f N 0 2 5 decomposition, K 3 and K 4 are the


constants o f equilibria (11.2.4) and (11.2.5), ( C ^ Q ^ Q is the analytical

concentration o f the parent H N 0 , 3 ( C ^ Q )act a n c


* (Q-l o)act 2
a r e t n e

actual concentrations o f nondissociated H N 0 3 and H 0 molecules, w h i c h


2

were estimated from the ratio o f the vapor pressure o f H N 0 3 and H 0 over 2

the solutions studied and pure components and with the use o f the data
presented in [11].
If decomposition proceeds through reaction (11.2.8), a similar

dependence o f the rate on ( C ^ Q )act a n c


* (Q-J o)act
2
c a n
b e
obtained, but

without K and with k replaced by k . The dependence o f the initial rates o f


4 6 7

decomposition o f H N 0 3 solutions on the ratio o f (CHNO ) ^ H 3


2
2 0 l s

presented in F i g . 11.2.3. The first-order dependence is observed over the


major part o f the range, but in the region o f diluted solutions, a deviation
towards a decrease o f the decomposition rate is observed, and these
deviations decrease as the temperature increases.

-4

-d

Fig. 11.2.3. Dependence of \og(dC 2/dt)^ on log((CHN03) /C 2o) ^


mO * 0
2
H
o r v a r o u s

temperatures: (/) 107.9, (2) 159.8, (3) 193.5°C, A are the values obtained by
extrapolation of the temperature dependences.
Thermal Decomposition and Combustion 151

D u r i n g monomolecular decomposition o f N 0 , the deviation o f the rates 2 5

can be explained by the increase o f N 0 2 5 ionization with the growth o f


permittivity o f the medium at the dilution o f H N 0 3 with water. The
permittivity decreases as the temperature increases, which reduces N 0 2 5

ionization.
A t interaction o f N 0 2
+
and N 0 ~ ions, the deviations caused by this
3

should not exist because equilibrium (11.2.4) is shifted towards ions and the
total concentration o f ions is only slightly changed under the action o f an
additional ionization o f a relatively small amount o f N 0 . 2 5 Due to this,
decomposition through the N 0 2 5 molecule is more probable. Even more
convincing evidence that decomposition proceeds through an N 0 2 5

molecule and not through the interaction o f ions is a considerably lower rate
o f decomposition o f the solutions o f N 0 2 5 in H N 0 3 in comparison with the
solutions in organic solvents (see Table 11.2.3, [11.12]), which is caused by
a strong ionization o f N 0 2 5 molecules in the solution o f H N 0 . D u r i n g 3

decomposition through the interaction o f ions, the solution in H N 0 3 has to


decompose at higher rate.

Table 11.2.2. The temperature dependence of the rate constant of thermal


decomposition of N 0 2 5 in the gas phase and in the various solvents

Solvent Solvent
(gas) (gas)

Gas 4.5 • 10 exp(-24500/RT)


,3
CH C1- 2 4.1 • 10 exp(-24400/RT)
13

CH C1 2

CH N0 3 2 3.1 • 10 exp(-24500/RT)
I3
CHCI3 6.4- 10 exp(-24600/RT)
13

Br2 2.5 • 10 exp(-24000/RT)


13
CH CHC1 3 2 12.5 • 10 exp(-24900/RT)
13

CHC1 - 2 10.4 • 10 exp(-25000/RT)


,3
N 0 2 4 16.3 • 10 exp(-25000/RT)
l3

CC1

CCI4 HNO3
3

2.8 • 10 exp(-24100/RT)
13
25.0- 10 exp(-27800/RT)
I3

Table 11.2.3. The evaporation heats of H N 0 3 and H 0 from the solution of


2

HNO3—H 0 2

HNO3 0 20 30 40 50 60 70 80 90 100
wt%

17.0 15.8 14.9 13.7 12.4 11.3 11.0 9.7 9.4

kcal/mol
11.2 11.2 11.3 11.4 11.4 11.4 11.4 11.4 -

kcal/mol
152 General Regularities of Thermal Decomposition of Onium Salts

During decomposition of N 0 2 5 molecules, in view o f (11.2.8), the


experimentally measured activation energy o f decomposition takes the
form:

^obs = E + AH + AH + 2 [
6 4 3 (^ J -^hno NO sol 3 " [ {^H O) -K O
2 so] 2 >

where £ b
0 S is the observable activation energy o f the solution
decomposition, E is the activation energy o f N 0
6 2 5 decomposition, AH and 3

AH are the enthalpy variations in the corresponding processes, their sum (5


4

kcal/mol) is estimated based on the data [11],

HNO3
(^NO^sob^HpW %N0,> a n d
^H o 2
a r e t h e
evaporation heats o f
and H 0 from the solutions and individual liquids, which are
2

presented in Table 11.2.3. Taking into account the variation o f the


evaporation heats when the solutions are diluted and Cj-nv^ decreases from
100% to 41.7%, the value o f E obs should increase from 27 to 40 kcal/mol. I f
QiN0 3 ' higher than 65%, the experimental values o f E
s
ohs almost coincide

with the calculated ones. F o r lower values o f CJ-TNO » 3


t n e
experimental
values increase, which is determined by a stronger dependence o f the
degree o f ionization o f N 0 on the temperature in the weak solutions. O n
2 5

the whole, this calculation confirms the described mechanism o f H N 0 3

decomposition.
The values o f k in the gas phase and numerous organic solvents are
6

close to each other [11.12]. Their temperature dependences are presented in


Table 11.2.2. One can assume that these values are also conserved in H N 0 , 3

although the actual concentration o f N 0 molecules and the rate o f their 2 5

decomposition are lower than in other solvents due to a strong ionization o f

HNO3
N 0 in H N O 3 . W i t h the use o f the k value and the rates o f decomposition
2 5 6

of solutions measured in [11.13], one can calculate from equation


(11.2.8) that at 108°C in 99.38% H N 0 the N 0 concentration is equal to: 3 2 5

^N 0 2 5
= 1 0
~ ' 3 m o l / 1
'

and in 41.7% H N 0 3 this value is 1 0 ' 10


mol/1.
In order to obtain the quantitative relation between the initial
composition o f the solution and the rate o f its decomposition, one should
take into account the transition o f some o f the components into the gas
phase. Such transition shifts equilibria (11.2.3) and (11.2.4) in the liquid,
w h i c h results in the rate variation. This determines the dependence o f the
rate o f decomposition on V . The details o f such calculations are presented
0
g

in Chapter 12 for ammonium nitrate.


The self-acceleration o f decomposition is determined by the interaction
o f the molecule o f H N 0 3 with N 0 , which is the decomposition product. It
2
Thermal Decomposition and Combustion 153

was revealed that introduction o f N 0 2 increases the initial rate o f


decomposition, and this increase is close to those observed at N 0 2

accumulation due to decomposition. Introduction o f N 0 2 in the amount o f


approximately 140% from its accumulated value at the moment when the
maximum rate is attained gives the initial rate, which accounts for 125%
from the maximal rate o f decomposition o f an acid without additives. W i t h
regard to the first order o f the catalytic reaction and thermodynamic
analysis o f possible reaction pathways o f N 0 2 in the system, one can
suggest the two pathways o f decomposition acceleration:

HN0 +N03 2 = HNO2+ N 0 , 3 A H = 25 kcal/mol,


H N 0 + N 0 = H N 0 + N O + 0 , A H = 30 k c a l / m o l .
3 2 2 2

Further transformations into the final products proceed by the following


reactions:

2NO+0 2 = 2N0 2

N0 3 = NO + 0 2

HN0 + H N 0 2 3 ^ 2N0 + H 0. 2 2

Though the rate constant o f interaction o f N O with H N 0 3 [11.14] is rather


high, this reaction cannot determine the rate o f the autocatalytic reaction
because the content o f N O in the system is too low and it does not
accumulate in the course o f the reaction due to the interaction with oxygen,
whose concentration in the system is high because this is one o f the final
products.
The described mechanism o f H N 0 3 decomposition makes it possible to
explain all kinetic regularities o f the process determined by relation
(11.2.1). A complex form o f the kinetic law o f a non-catalytic stage (the
first term) is determined by the quadratic dependence on HN0 3

concentration in the numerator and, in the denominator, by the initial


concentration o f water in the solution and by its accumulation during
decomposition. The second, autocatalytic term is determined by the
accumulation o f N 0 . 2

11.3 Perchloric Acid


Thermal decomposition o f perchloric acid occurs with a very strong self-
acceleration and is frequently completed by explosion. Thermal
decomposition is sharply decelerated at an introduction o f water in the
system. A review o f the early works is presented in [11.15]. Under
conditions o f a closed and completely temperature-controlled system, the
154 General Regularities of Thermal Decomposition of Onium Salts

decomposition o f H C 1 0 4 proceeds mainly according to the following


equation:

2HC10 = 4 H 0 + C l + 3.50
2 2 2

A noticeable accumulation o f chlorine oxides does not proceed. In an open


system, considerable amounts o f chlorine are evolved in the form o f C 1 0 2

and CIO3, which is connected with their rapid elimination from the reaction
vessel and with the absence o f the secondary reactions. The kinetics o f
thermal decomposition o f aqueous solutions o f H C 1 0 4 was studied in
[11.16, 11.17] from the rate o f heat release by a technique similar to those

HCIO4 were
employed in studies o f nitric acid. Aqueous solutions with a concentration
of 65.9-100% studied over the temperature range o f 2 7 . 2 -
2 4 0 ° C . A t dilution, due to a very strong dependence o f the rate on acid
concentration, the range under consideration shifts towards the region o f
higher temperatures. For the same reason, after decomposition o f 1 0 - 1 5 %
o f the acid present in the solution, under isothermal conditions, the rate o f
decomposition reduced so strongly that the reaction was almost stopped. In
contrast to nitric acid, this inhibition is not connected with the reversibility
o f decomposition. In this case, elimination o f gaseous products from the
reaction zone does not result in the growth o f the rate. The reaction
proceeds at a very low rate and the complete termination of the
decomposition is not observed. In experiments using the microcalorimetry
technique, even for the anhydrous H C 1 0 4 the flashes have never been
observed, though in a series o f works it was argued that the heating or
storage o f perchloric acid in a closed volume inevitably results in explosion
[11.15].
A probable reason o f such explosions is thermal self-ignition which is
determined by insufficient heat dissipation during a strong self-acceleration
o f the decomposition. For all temperatures and samples utilized, the rate o f
heat release did not exceed 0.005 cal/s. Under conditions o f calorimeter heat
dissipation, the heating o f the substance was no more than 0.2°C, and the
reaction proceeded under almost isothermal conditions without thermal
acceleration.
B e l o w 120°C, one can observe the induction period on the curves o f the
rate o f heat release. After this period, the rate sharply increases and passes
through the maximum. The curves are described by the kinetic equation o f
the second order with the autocatalysis o f the first order. A n example o f
such heat release curves is shown in F i g . 11.3.1. Analysis o f the kinetic
regularities shows that above 120°C the induction period is shorter than the
time o f heating o f the reaction ampoule (about 10 min) and, thus, cannot be
revealed. One can argue that at higher temperatures at the beginning o f the
reaction, a strong self-acceleration also occurs. The actual initial rate is
considerably lower than the measured one and it is possible to measure only
Thermal Decomposition and Combustion 155

the rate o f the developing process, i.e., the values o f k . One can observe a 2

kink on the temperature dependence o f k (Fig. 11.3.2).


2

2.0 2.25 2.5 f/r-fO*

Fig. 11.3.1 Fig. 11.3.2

Fig. 11.3.1. Time dependence of the rate of heat release during decomposition of the
solutions of HC10 (55.4°C) for various [HC10 ], wt %: (/) 100, (2) 97.9, and (3)
4 4

95.0.

Fig. 11.3.2. Dependence of \ogk on \/T for the solutions of HC10 for various
2 4

[HC10 ], wt %: (7) 65.9, (2) 68.7, (3) 70.3, (4) 73.6, (5) 80.3, and (6) 84.8.
4

A t higher temperatures, the activation energy decreases sharply. The values


o f k\ were calculated from the dependence o f 77 on t (13.1.1) and the
anomalously high values o f the activation energy and pre-exponential factor
were obtained. This calculation is true i f the law utilized is preserved during
the induction period, but it is impossible to check it because the rate is
below the sensitivity o f the technique. The values o f k only formally can bex

considered as the measure o f the initial rate o f decomposition, which


satisfies the integral o f the kinetic equation. The initial rate can differ

HCIO4 is very low and the observable decomposition is connected


considerably from this value. Obviously, the initial rate even for 100%
with the
secondary processes. The values o f k\ and k 2 obtained are presented in
Table 11.3.1.
The acid with the concentration o f 50 mol % (84.8 wt %) corresponds
to the composition o f oxonium perchlorate H 0 C 1 0 " and represents a
3
+
4

solid ionic salt with r i t me


=
4 9 - 5 0 ° C . Over the temperature range studied,
this is a liquid melt. This melt can be considered as the simplest model o f an
onium salt.
156 General Regularities of Thermal Decomposition of Onium Salts

Table 11.3.1. The rate constants of thermal decomposition of HC10 4

HCIO4, wt % T,°C ki =f(T), s- 1


T°C h =f(T), s-
1

65.9 213-240 10 22,


exp(-61000/RT)
68.7 202-230 10 22 2
exp(-60000/RT) 230-240 10 7l
exp(-25500/RT)
70.3 202-220 10 I88
exp(-52000/RT) 220-240 10 67
exp(-24500/RT)
73.6 160-180 10 24 6
exp(-61000/RT) 180-240 10 57
exp(-22200/RT)
80.3 120-150 10 l89
exp(-45000/RT) 150-180 10 5,,
exp(-17700/RT)
84.8 100-125 10 122
exp(-29000/RT)

HCIO4, wt % T°C k =f(T), s"


2
1
T,°C

90.0 52-73 10 1()(,


exp(-20400/RT) 52-73 10
593
exp(-104000/RT)
95.0 48-71 10 1,,(,
exp(-20200/RT) 48-71 10
59(>
exp(-102000/RT)
97.9 40-64 10 lu
exp(-21800/RT) 40-64 10 573
exp(-98000/RT)
100.0 27-62 10 121
exp(-22700/RT) 27-62 10 448
exp(-78000/RT)

Due to the weak alkali properties o f water, the equilibrium o f dissociation

H 0 + CIO4" ^
3
+
H 0 + HC10
2 4

is shifted to the left-hand side considerably less than in the case o f other
perchlorates formed by stronger bases.
In order to obtain the dependence o f k on the concentration o f H C 1 0 ,
2 4

all o f the k values were extrapolated to 5 0 0 ° C . A t lower temperatures, the


2

comparison is difficult due to the kinks in the temperature dependence. The


degree o f ionization o f aqueous solutions o f H C 1 0 with a concentration 4

above 7 5 % was determined in [11.18]; for weaker solutions, it was


estimated from the vapor pressure o f H C 1 0 . The obtained dependence o f k
4 2

on the concentration o f H C 1 0 molecules is shown in F i g . 11.3.3. A s can be


4

/ \0 33
seen, k is proportional to I C
2 ) H C L 0

tg ^hco
i*
Fig. 11.3.3. Dependence of \og(k) on log(Cj_jcio ) for perchloric acid solutions.
4
Thermal Decomposition and Combustion 157

The analysis o f the whole complex o f the kinetic data shows that
thermal decomposition o f H C 1 0 and its aqueous solutions is a complex
4

branched chain reaction.


The ion C10 ~ is thermally stable at the temperatures considered and
4

cannot give a noticeable contribution to the observable rate. The rate o f


HCIO4 decomposition in the gas phase is lower by several orders o f
magnitude than the experimentally determined rates o f decomposition in
solutions.
Thermal decomposition o f H C 1 0 in the gas phase proceeds at
4

measurable rates at 3 0 0 - 4 0 0 ° C and represents a monomolecular reaction


[11.19]. The temperature dependence o f the rate constant takes the form:

£ = 10 13 7 5
exp(-45100//?r), s" . 1

The activation energy value corresponds to the decomposition o f the


HCIO4 molecule in two radicals O H + C 1 0 . A t high temperatures o f
3

decomposition, these radicals are very rapidly converted into the final
products 0 , C l , and H 0 . Introduction o f inert gases and variation o f the
2 2 2

initial pressure o f H C 1 0 give rise to only slight changes o f the rate o f


4

decomposition. The reason for the increase o f the rate o f H C 1 0 4

decomposition in the liquid phase, as for the case o f nitric acid, is the
possibility o f equilibrium formation o f perchloric anhydride having a low
thermal stability:

3HC10 ^ 4 CI2O7 + H 0 + CIO4"


3
+
(11.3.1)

CI2O7 = CIO3 + CIO4.

The correctness o f such a decomposition pathway is unambiguously


confirmed by the virtual coincidence o f the equilibrium constant (11.3.1)
determined from the initial rate o f H C 1 0 decomposition and the 4

independently determined rate constant o f C 1 0 decomposition (K joc2 7 C


=

1.94-10" ) with the K


4
cvalue calculated from the measurements o f
composition and vapor pressure over perchloric acid (Ac 20c 1 30-10 ) = 4

[11.15].
In [11.20], further possible pathways o f branching, continuation, and
chain termination (all in all, 87 reactions) were considered. The kinetic
analysis o f the process was carried out based on the estimates o f the
possible rates o f these reactions, taking account o f their energetic barrier. It
was revealed that the following processes mainly contribute to the
development o f decomposition:

1. Reactions o f chain branching:


158 General Regularities of Thermal Decomposition of Onium Salts

1. H C I O 4 + H C I O 3 = C1 0 + 2 6 H 0 2

C1 02 6 = 2C10 3

2. C 1 0
2 7 + CIO3 = C1 0 + C10 2 6 2 + 0 2

3. C 1 0
2 7 + CI = CIO3 + 2CIO2

C1 0 2 7 + C10 2 = C1 0 2 6 + CIO + 0 2

II. Reactions o f chain continuation:

1. H C I O 4 + C I O 3 = HCIO3 + C 1 0 2 + 0 2

2. C 1 0
2 6 = C 1 0 + CIO4
2

3. C 1 0 2 = CI + 0 2

4. CIO4 = C10 + 02 2

III. Reactions o f chain termination were considered in less detail. The


following reactions are probable:

1. C I O 2 + C I = Cl 2 + 0 2

2. CI + CI = Cl . 2

Formation and accumulation o f C 1 0 2 and CIO3 in the course o f HCIO4


decomposition in the liquid phase were repeatedly observed experimentally
[11.15]. A n attempt has been made to choose such theoretically possible
values o f the reaction rate constants in order to describe by this mechanism
the experimental kinetic curves o f decomposition o f aqueous solutions o f
HCIO4, first o f all, the value o f the induction period, the maximal value o f
the rate, and the time o f the attainment o f the rate maximum. The best fit
between the experimental and calculated curves is observed for branching
by the first reaction, but, in this case, in calculation, a weaker dependence o f
the induction period and the maximal rate on the acid content is realized. A
closer agreement is achieved for a combination o f the first and third
reactions when the first one dominates. The model o f a branched chain
process, obtained on the basis o f these reactions, explains qualitatively all o f
the observed experimental facts: a very high acceleration, a kinetic
termination o f the reaction after decomposition to a certain degree, and a
very strong dependence o f the rate on the initial acid and water
concentrations. A n additional confirmation o f the occurrence o f a branched
chain process at decomposition o f perchloric acid is the inhibition o f this
reaction by compounds containing the trichloromethyl group: chloroform,
tetrachloromethane, hexachloroethane, trichloroacetic acid, etc. [11.15].
These compounds interact with chlorine oxides and terminate chain
Thermal Decomposition and Combustion 159

branching. Their introduction in solutions o f H C 1 0 in the amounts o f 1-3%


4

before the beginning o f a rapid decomposition destroys accumulated C 1 0 2

and C 1 0 , increasing the induction period by a factor o f 10 or 100. After


3

consumption o f the inhibitor, decomposition develops and proceeds at rates


close to those observed in its absence. The employment o f other, better
known inhibitors o f chain processes is impossible in this system because
they are oxidized very rapidly by perchloric acid.
Quantitative calculations by this model are hampered by the absence o f
the values o f the rates o f elementary stages. Variations o f the activation
energy and the kinetic law within a wide range o f temperatures and acid
concentrations are obviously connected with variations o f the contribution
o f separate stages to the total process. These variations have not been
analyzed in detail. Unfortunately, today there is no theoretically proven
quantitative description o f the regularities o f perchloric acid decomposition,
w h i c h is similar to those presented above for nitric acid. Thus, it is
impossible to predict the variation of the parameters of thermal
decomposition o f H C 1 0 4 on variation o f the conditions o f its conducting,
and in particular, to estimate the rate o f acid decomposition in the melt or
crystal lattice o f onium perchlorate. A s we know, the decomposition o f
perchloric acid by the branched chain mechanism is the first example o f a
radical chain reaction in which acid-catalytic stages play an essential role.

11.4 Dinitramide
Recently, a series o f works on synthesis and study o f the properties o f
various salts o f dinitramide H N ( N 0 ) 2 was published, for example, [ 1 1 . 2 1 -
2

11.24]. In [11.21], the method o f dinitramide synthesis was described and it


was revealed that it represents a mobile liquid, which at room temperature,
after several minutes, is decomposed with self-heating and rich liberation o f
gas products. Sometimes, the decomposition was completed by an
explosion. A t the same time, the data available on the kinetics and
mechanism o f thermal decomposition o f the ammonium salt o f dinitramide
[11.25, 11.26] show that this salt is rather thermally stable even at elevated
temperatures, i.e., the stability o f the N ( N 0 ) ~ anion is considerably higher
2 2

than that o f dinitramide molecules. For correct and reliable estimation o f a


probable role o f dinitramide in the rate o f thermal decomposition o f its
onium salts and various solutions, it is necessary to have available
quantitative data on the rate constants o f thermal decomposition o f anions
and molecules o f dinitramide and also the rate constants o f the other
possible interactions o f dinitramide with the components o f the reacting
system. A quite detailed study o f these problems was carried out in [11.27,
11.28].
Dinitramide ( D N A ) was synthesized by the reaction o f K N ( N 0 ) 2 2 with
a dry H C l by the procedure described in [11.21]. A l l operations of
160 General Regularities of Thermal Decomposition of Onium Salts

dinitramide synthesis and the subsequent preparation for kinetic


experiments were carried out in a darkened room.
A l l attempts to study the kinetics o f the heat release o f a pure
dinitramide, even with the use o f small samples, were completed by an
explosion. T o conduct an isothermal experiment, it is necessary to reduce
sharply the rate o f heat release per unit volume that can be achieved for
diluted solutions. The choice o f organic aprotic solvent is hampered by the
high reactivity o f dinitramide. Thus, aqueous solutions were studied, in
w h i c h dinitramide molecules are strongly ionized, but the equilibrium can
be calculated quite accurately.
Within the temperature range o f 1 0 0 - 1 1 5 . 9 ° C , the decomposition o f
aqueous solutions o f H N ( N 0 ) o f 0.35-3.79 m o l % content (which
2 2

corresponds to the analytical concentration o f 0.19-2.70 mol/1) was studied.


In diluted solutions, the kinetic curves are described by the equation o f the
first order reaction. When the content o f H N ( N 0 ) exceeds 1.5 m o l % , the 2 2

self-acceleration o f the process starts. Since in aqueous solutions o f


H N ( N 0 ) the non-dissociated molecules and anions N ( N 0 ) " participate
2 2 2 2

in the decomposition, the rate curves o f decomposition o f a 3.35 mol %


solution o f K N ( N 0 ) in water at temperatures o f 9 6 . 5 - 1 5 9 . 8 ° C were
2 2

determined. These curves are described by the kinetic equation o f the first
order reaction. The kinetics o f the consumption o f N ( N 0 ) ~ anion and 2 2

variation o f the content o f H 0 ions as a function o f the degree o f


3
+

conversion was determined by the heat release in a 3.05 m o l % solution o f


H N ( N 0 ) at a temperature o f 106.2°C. The consumption o f anions
2 2

corresponds to the kinetics o f heat release and the content o f H C T ions 3

reduces only by 10%, i.e., the main product o f decomposition in the


solution is nitric acid. The IR-spectroscopic study demonstrated the
presence o f only N 0 in the gas phase, i.e., in general, the stoichiometric
2

equation o f decomposition takes the form:

HN(N0 )2 = 2 HN0 +N 0
3 2

T o study the rate o f thermal decomposition in media with higher acidity, the
decomposition o f solutions o f N H N ( N 0 ) and K N ( N 0 ) in aqueous
4 2 2 2 2

solutions o f H S 02 4 containing 0.18^10.07 mol % o f H S 0 was 2 4

investigated. Since an increase o f acidity o f a medium leads to a


considerable increase o f the rate o f thermal decomposition, in order to keep
the range o f rates o f thermal decomposition convenient for measurements,
the experimental temperatures were reduced. Therefore, a temperature
range o f - 9 . 0 to + 1 5 0 . 8 ° C was studied. The concentration o f the salt was
within the interval o f 0.1-0.3 mol/1.
In aqueous solutions o f dinitramide, ionization proceeds up to a
considerable degree o f conversion:

HN(N0 ) + H 0 — 2 2 2 H 0 + N(N0 ) ".


3
+
2 2
Thermal Decomposition and Combustion 161

The ions formed and the initial molecules are hydrated by water molecules.
Both the molecules o f H N ( N 0 ) 2 2 and anions N ( N 0 ) ~ contribute to the
2 2

observable rate o f decomposition. Only in very diluted solutions (0.03


m o l % and lower) are the molecules o f H N ( N 0 ) practically absent and the 2 2

rates o f decomposition o f the solutions o f K N ( N 0 ) , N H N ( N 0 ) , and 2 2 4 2 2

H N ( N 0 ) coincide.
2 2

Since in all solutions o f dinitramide studied the degree o f ionization o f


acid is high, the concentration o f anions is almost equal to the analytical
concentration o f the acid. The reaction order and the kinetic parameters o f
decomposition o f ions were obtained from the analysis o f the kinetic curves
o f decomposition o f the diluted solutions o f K N ( N 0 ) 2 2 and N H N ( N 0 ) .
4 2 2

The decomposition o f such solutions obeys the kinetic law o f the first order
reaction, the rate o f decomposition is proportional to the concentration o f
anions in solution, and a solvated ion decomposes in accordance with the
law o f a monomolecular reaction. The following temperature dependence
for the rate constant o f decomposition o f N ( N 0 ) ~ anions (k ) 2 2 an was
obtained:

k an =1.7-10 exp(-20.5-10 /r),


1 7 3
s . _ 1

The concentration o f non-dissociated HN(N0 ) 2 2 molecules in aqueous


solutions o f acids was estimated by the following relationship:

C =C /(\
k 0 + \0 °- ").
H pK
(11A1)

Here, H is the acidity function o f an acid solution. The value o f pK o f


0 a

D N A in the medium o f sulfuric acid was determined by spectrophotometric


analysis from the measurements o f the ratio o f the concentration o f ions and
neutral molecules in solutions o f H S 0 with a known value o f H [11.28]
2 4 0

and it was obtained thatpK = 4.85 ± 0.05. a

The values o f H for the solutions o f H N ( N 0 ) were not determined,


0 2 2

but within the range o f concentrations studied the acidity functions o f all
strong acids are close to each other and are determined mainly by the
activity o f H 0 ions.
3
+

The values o f H for solutions o f sulfuric acid with equal molar


0

concentrations were used. For all temperatures and solution concentrations


studied, the first order on the concentration o f H N ( N 0 ) molecules was 2 2

obtained with regard to the rate o f anion decomposition. The temperature


dependence o f the rate constant o f thermal decomposition o f dinitramide
(k ) takes the form:
ac

kac =3.3-10 exp(-13.7-10 /r),


1 4 3
s - 1
162 General Regularities of Thermal Decomposition of Onium Salts

If dinitramide salts are dissolved in sulfuric acid, their anions are protonated
with the formation o f dinitramide molecules:

HN(N0 ) ^ H 2 2
+
+N(N0 ) " 2 2 ^ = a - a /a H
+
a n a c (11.4.2)

Further protonation o f dinitramide molecules proceeds with the formation


o f their protonated form:

H N(N0 )2 2 2
+
^= H +
+ HN(N0 ) 2 2 K = a *- a / a
p H a c p (11.4.3)

In such solutions, the initial rate o f anion consumption is determined by the


sum o f the decomposition rates o f all three forms o f dinitramide in solution
and, under the assumption o f the first order o f the reaction o f H N ( N 0 ) 2 2 2
+

decomposition, takes the form:

-{dC/dt) =k C +kiC +k C . l=0 K m K p t (11.4.4)

where k is the rate constant o f decomposition o f dinitramide molecules in


k

the solutions o f sulfuric acid, k^ and C are the rate constant and the p

concentration o f the protonated form o f H N 0 / . 2 3

H a v i n g solved a set o f equations (11.4.2) and (11.4.3) under the


assumption that the initial concentration o f the salt is small and does not
change the activity o f hydroxonium ions, and using the following
expression for the acidity o f the medium h =a
0 H+ -y /^ +
B BH , one can
obtain the expression for the concentration o f anions, molecules, and the
protonated form o f dinitramide. In principle, at an essential difference in the
rate constants o f decomposition o f these forms, three regions o f the acidity
scale o f H S 0 solutions should exist, where the rate o f decomposition o f
2 4

the salt o f dinitramide dissolved in them is determined by the rate o f


decomposition o f only one o f the equilibrium forms N ( N 0 ) ~ , H N ( N 0 ) , 2 2 2 2

or H N ( N 0 ) . In the transition regions, the rate is equal to the sum o f the


2 2 2
+

rates o f decomposition o f two forms. The rate o f decomposition o f anions is


a determining one when H « pK and dinitramide exists almost only in
0 a

the form o f anions. In this case, the rate o f decomposition does not depend
on the acidity o f solution H and is described by the first term o f equation
0

(11.4.4). A s the acidity o f the medium increases, the concentration o f


H N ( N 0 ) molecules also increases and the concentration o f H N ( N 0 ) is
2 2 2 2 2
+

still insignificantly low. Starting from a particular value o f H , the rate o f 0

decomposition is determined by the rate o f decomposition o f H N ( N 0 ) 2 2

molecules. In this range, the concentration o f H N ( N 0 ) and the rate o f 2 2

decomposition linearly increase with the rise o f the acidity o f the system:

-(dC/dt),__ = 0 K -C -h /K ,
c 0 0 a (11.4.5)
Thermal Decomposition and Combustion 163

The concentration o f protonated dinitramide molecules increases


proportionally to the square o f acidity h . Thus, at a determining role o f 0

decomposition o f a protonated form o f dinitramide, the expression for the


initial rate takes the form:

-(dC/dtU = k -C -h /K -K ,
p 0
2
0 a p (11.4.6)

Within this range o f acidity, the second order on the acidity o f a medium
should be observed.
A t a further increase o f acidity, the determining role o f H N ( N 0 ) 2 2 2
+

decomposition is preserved, but the concentration o f N ( N 0 ) ~ anions w i l l 2 2

become low and equilibrium (11.4.2) w i l l not be essential, the concentration


^HN(N0 ) 2 2 * l°
s c s e t 0
and C H N ( , N 0 y w i l l be determined only by

equilibrium (11.4.3). In this case, the expression for the rate o f


decomposition takes the form:

-(dC/dt) =k .C .h /K ,
(=0 p 0 0 p (11.4.7)

i.e., again, the first order on the acidity o f solution w i l l be observed. If a


high degree o f protonation o f H N ( N 0 ) 2 2 molecules is attained, when it is
already possible to neglect the content o f these molecules, the concentration
o f the protonated form becomes constant and the rate should not depend on
acidity.

-1 -i

-2-

-7-

_8_] , , , , , , , r

- 8 - 6 - 4 - 2 0
H 0

Fig. 11.4.1. Dependence of \og(k) on H at various temperatures T°C: (1) 158.8, (2)
Q

136, (3) 121.4, (4) 83.3, (5) 30.6, (6) 20.7, (7) 3.7, and (8) 1.1.
164 General Regularities of Thermal Decomposition of Onium Salts

The dependence o f the initial rate o f decomposition o f solutions on H 0 is


shown in F i g . 11.4.1 for various experimental temperatures. The transition
from decomposition o f H N ( N 0 ) 2 to decomposition o f the protonated form
2

(lines 4) and the change o f the second order on acidity o f a medium to the
first one at the decomposition o f a protonated form (lines 6) are clearly
observed. The region in which H N ( N 0 ) 2 2 molecules are totally protonated
was not achieved in the experiments. In this region, the rates are too high.
F r o m the experiments at different temperatures and from relations
(11.4.5) and (11.4.6) with regard to the temperature dependence o f H , the 0

following relationships were obtained:

k[ IK = 5.7 • 10 exp(-l8.6 • 10 /T\


c a
15 3
s , -1
(11.4.8)

k /K -K
p p a = 6.6 • 10 exp(-16.9 • 10 / T),
10 3
s", 1
(11.4.9)

k /K =9A>lO"
p p exp(-14.3• 10 /T\
3
s .
_ 1
(11.4.10)

F r o m comparison o f equations (11.4.9) and (11.4.10) it was obtained that:

K a =1.4-10-exp(2.6-10 ), 3
at 2 0 ° C , K = 1 0
a
4 9 9
.

The value o f K determined by the spectral method is equal to K = 10 .


a a
4 8 5

These values coincide within the accuracy o f the experiment and


calculation. F r o m relationship (11.4.9) and with the use o f the expression
for K for the solution o f H S 0 , we w i l l have:
a 2 4

k'ac = 7 . 9 - 1 0 e x p ( - 1 6 . M 0 / 7 ) , s" ,
1 6 3 1
(11.4.11)

These values o f k ac are slightly lower and their temperature dependence is


somewhat higher than the values o f k ac obtained earlier for aqueous
solutions. The difference can be considered, along with the experimental
errors, with the difference in the degree o f solvation o f HN(N0 ) 2 2

molecules in aqueous and sulfuric solutions. Sulfuric acid competes with


molecules o f dinitramide for water molecules. In the solution o f K N ( N 0 ) , 2 2

at its dilution by water, the rate increases.


A t decomposition o f the protonated form o f H N ( N 0 ) , the values o f 2 2 2
+

the acidity o f the solutions, which would be close to or higher than the K v

value, were not achieved in the experiments, and we could not determine
this value. In order to obtain even very rough estimates o f the rate constant
o f decomposition o f a protonated form kp, the value o f K p was estimated
from the comparison o f the ionization constants o f dihydric acids o f the first
and second stages. In the majority o f cases, at 2 0 ° C , this difference is 1 0 - 4
Thermal Decomposition and Combustion 165

10 . It was accepted that the average value is as large as 10 . Then, we w i l l


5 4 5

have A: (20°C) = 3.3 • 10 and yfcp(20° C ) = 2.0. For decomposition o f a


p
9

protonated form, the cyclic transition state is most probable, for which the
average pre-exponential factor is 1 0 . Then, we w i l l have: 12

^ = 10 exp(-7.9-10 /r),
12 3
s , _ 1

Kp = l.l-exp(6.410 /T). 3

The dependence o f the total rate o f decomposition o f the solutions o f


dinitramide at 20 and 3 0 0 ° C on H , 0 calculated from the temperature
dependences o f the rates o f decomposition o f its various forms, is presented
in F i g . 11.4.2. A t 2 0 ° C and H > 1.27 (region I), the rate is determined by
0

the rate o f decomposition o f anions and does not depend on H ; within the 0

range o f - 1 . 9 < H 0 < 1.27 (region II), the determining factor is the
decomposition o f H N ( N 0 ) 2 molecules and the rate is proportional to the
2

acidity o f the medium. O n further increase o f acidity (decrease o f H ), 0 at


point b, the leading role passes to the decomposition o f a protonated form o f
dinitramide (region III), and, correspondingly, the second order on acidity
of a medium is obeyed. T o the left from point c (H 0 < -4.9), where C a n =
C ,
a c the acidity attains the value for which C a n >> C , C a c p and the
concentration o f H N ( N 0 )
2 2 2
+
is proportional to the first power o f acidity
(region I V ) .

-1 1 1 1 1 1 1 1 1 1 r
Fig. 11.4.2. Dependence of \ogk
- 8 - 6 - 4 -
exp on H Hat (1)
2 0
0
2
20 and (2) 300°C.
0

A s the temperature increases to 3 0 0 ° C , the character o f this dependence


changes considerably due to various temperature dependences o f the rate o f
decomposition and concentrations o f different forms o f dinitramide. Points
a and b are shifted towards a higher acidity because the activation energy o f
decomposition decreases i f the anion and then the molecule are protonated.
A t the same time, the location o f point c is determined by the value o f K , a

which decreases as the temperature increases, and point c is shifted towards


166 General Regularities of Thermal Decomposition of Onium Salts

lower temperatures. Thus, region III becomes narrower and at 2 2 0 ° C it


should completely disappear. T o the left from point c, as the acidity o f the
medium increases, the C A C value becomes close to C and the decomposition0

o f the H N ( N 0 ) 2 molecule still determines


2 the rate o f decomposition.
Hence, in the dependence o f l o g ® on H , a plateau should appear, and, at
0

point b, the leading role passes to the decomposition o f H N ( N 0 ) . Here, 2 2 2


+

the first order o f acidity is observed again.


A t 3 0 0 ° C , all rate constants are very high and isothermal calorimetric
measurements under such conditions are impossible. Nevertheless, these
data can be important, for instance, for the analysis o f the regularities o f
combustion and detonation o f dinitramide in acid media. .
Thermal decomposition o f the ammonium salt o f dinitramide in the
liquid phase [11.25] proceeds with considerable self-acceleration. The
increase o f the rate in the course o f decomposition occurs also in the
decomposition o f some solutions o f dinitramide. The passage from one
medium to another can considerably change the rate of thermal
decomposition, which should be taken into account in operations with
various dinitramide compounds. Thus, we studied the kinetic regularities o f
the possible interactions o f dinitramide with the products o f decomposition
and with the other components o f the system, which allows us to estimate
the rate variation in the course o f decomposition. The rate variations at a
solvent change permits one to choose between different possible pathways
o f thermal decomposition.
W e studied the influence o f nitric acid on the kinetic regularities o f
thermal decomposition. For this, we investigated the decomposition o f
solutions o f 0.1-0.3 wt % o f N H N ( N 0 ) 4 2 2 in 56-90 wt % aqueous solutions
o f nitric acid. In all experiments, the first order o f the reaction o f
decomposition is preserved during decomposition. A t low temperatures, the
rate o f the process in nitric acid with the same function o f acidity is
essentially higher than in sulfuric acid. A s the temperature increases, these
differences decrease, and, near 100°C, the rates o f decomposition in sulfuric
and nitric acids almost coincide.
Within the range o f acidity studied, the initial rate is proportional to the
square o f acidity h . This character o f the dependence cannot be explained
0

by the determining role o f a direct interaction o f dinitramide molecules and


nitric acid. Within the range o f acidity studied, the concentration o f H N 0 3

weakly increases, and for the dependence o f k /h exp 0 on C ^ Q ^ O N E C A N

obtain an improbably high order (near 20). The most probable explanation
is the equilibrium formation o f a mixed anhydride o f nitric acid and
dinitramide with its subsequent thermal decomposition w h i c h determines
the rate o f the whole process:

HN(N0 ) + H N 0
2 2 3 = N 0
4 6 + H 0 ,
2 K= ,0 1-4.12)
hno/ a
ac
Thermal Decomposition and Combustion 167

where a anh , a a c are the activities o f molecules o f anhydride N 0 4 6 and


nitroamide.

N 0 — N 0
4 6 2 5 + N 0 2

N 02 5 +H 0= 2 2HN0 . 3

W i t h regard to the dependence o f a on the acidity o f a medium and under ac

the condition that C , C a c a n h << C , we w i l l have:


0

-(dC I dt) f=0 = k • C - (k -K/K )


exp 0 anh a a ^ >\-CJ a
um Hi0 ,

where £ a n h is the rate constant o f anhydride decomposition in a solution, k Qxp

is the experimental rate constant, and K is the constant o f equilibria o f acid a

dissociation o f nitroamide. The following temperature dependence o f £ e x p

was calculated from the experimental data:

*exp = *anh ' K / =10 2 8


exp(-7.3 • 1 0 / T ) , s" .
3 1

In the course o f the reaction, the first order is explained by the fact that
C H N 0 ^ is practically constant during the decomposition, even slightly

increasing due to the formation o f nitric acid at thermal decomposition. The


dependence o f log(£ x ) ° #0 in sulfuric and nitric acids at temperatures o f
e P
n

- 1 ° C and 9 9 . 5 ° C is presented in F i g . 11.4.3.

\ 0
1 -2

-4

-6

-8

-10

-12

-14

-16

-10 -8 -6 -4 -2 0 2
H 0

Fig. 11.4.3. Dependence of log£ exp on H at (1) - 1 ° C and (2) 99.5°C in (a) sulfuric
0

and (b) nitric acids.


168 General Regularities of Thermal Decomposition of Onium Salts

A t - 1 ° C (lines 1) and H > 1, when H N 0


0 3 molecules are almost absent, the
rate is determined by the decomposition o f N ( N 0 ) ~ anions and is the same
2 2

in sulfuric and nitric acids.


Nevertheless, at higher acidity, the formation o f the m i x e d anhydride
and its decomposition in nitric acid give a higher rate and for all H < 1 the 0

rate o f decomposition in nitric acid is considerably higher than that in


sulfuric acid.
Due to the fact that the temperature dependence o f k exp is weaker than
that for the rate constants o f thermal decomposition o f dinitramide
molecules and its protonated molecules, with an increase in temperature the
difference in the rates o f decomposition in sulfuric and nitric acids
decreases, and at 9 9 . 5 ° C (lines 2) the interaction with H N 0 3 determines the
rate o f the whole process only at H < - 5 . 5 . A t a lower acidity, the rate is
0

determined by the decomposition o f H N ( N 0 ) 2 and H N ( N 0 ) 2


2 2 2
+
molecules,
and the observable rates coincide with the rates o f decomposition in sulfuric
acid.
Thus, interaction o f dinitramide with nitric acid considerably increases
the total rate o f decomposition at low temperatures and in media with a high
acidity, but at the thermal decomposition o f the ammonium salt o f
dinitramide at temperatures o f 100°C and higher, it does not play an
essential role.
W e studied the kinetics o f thermal decomposition o f the ammonium salt
o f dinitramide in anhydrous acetic acid (the content o f H 0 no more than 2

0.004 wt % was determined by the electric conductivity o f the acid) within


the temperature range 1 9 . 3 - 1 0 3 . 5 ° C . In the course o f the reaction and at
variation o f the concentration o f the salt in the parent solution, the kinetic
law o f the first order is observed. The rate o f thermal decomposition o f
acetous solutions o f N H N ( N 0 ) 4 2 2 is higher by several orders o f magnitude
than that for aqueous solutions. This rate depends on the water content, and
at the passage to a commercial glacial acetic acid containing 0.4 wt % o f
water, the rate o f decomposition decreases at 87.2°C by a factor o f 55. The
estimation o f the pK a value o f dinitramide in acetic acid was carried out
measuring the electric conductivity o f a solution with a concentration o f 1.1
• 10 -2
mol/1, and we obtained pK a = 8.1 ± 0.5 and, in the aqueous solution,
pK a = -4.85.
The characteristic features o f the kinetics o f thermal decomposition o f
dinitramide in acetic acid are o f importance for comparison o f homolytic
and heterolytic pathways o f decomposition. A low dielectric permittivity o f
acetic acid (6.2 in comparison with 81 for water) makes an ionic
dissociation o f dissolved compounds difficult. Thus, in acetic acid, all
heterolytic pathways o f decomposition should be strongly inhibited. A c e t i c
acid, possessing a low proton affinity, decreases the strength o f all acids
dissolved in it. For dinitramide, in comparison with aqueous solutions, the
pKk value in acetic acid increases by approximately 13 units. For other
strong acids, this variation is also equal to 13 ± 2 units. The relative strength
Thermal Decomposition and Combustion 169

o f acids, in particular, dinitramide and nitric acid, does not change at the
passage to acetic acid.
The extrapolation o f the data on the acidity function o f H S 0 2 4 solutions
in C H 3 C O O H [11.30] to a pure acetic acid gave the h value for C H C O O H 0 3

which is approximately 10. Comparison o f this h value and an K value for 0 a

dinitramide in acetic acid clearly showed that in acetous solutions all o f the
parent amount o f N(N0 )2~ 2 anions passes into the molecular form
HN(N0 ) 2 2 and further protonation o f these molecules is probable. The
observable rate o f decomposition in the solution o f acetic acid is higher by a
factor o f 10 than that in solutions o f H S 0 2 4 with an equal HN(N0 ) 2 2

content. There are no reasons to expect an essential increase o f the rate


constant o f decomposition o f H N ( N 0 ) molecules (A ) at the passage from 2 2 ac

an aqueous solution to acetic acid. The determining role o f decomposition


of H N ( N 0 )
2 2 2 molecules, whose rate o f decomposition is considerably
higher, is most probable. This is also confirmed by a sharp decrease o f the
rate at a slight increase o f the water content in acetic acid. In this case, only
content o f a protonated form can decrease considerably due to the reduction
o f the solution acidity.
In this case, the initial rate o f decomposition o f dinitramide in a
solution o f acetic acid can be written in the following form:

(dC/dt)^ =k -C =k -C -h /K ,
0 Qxp 0 p 0 0 p

where and K are the rate constant o f decomposition o f a protonated form


p

of H N ( N 0 )
2 2 2
+
and the equilibrium constant o f dissociation in a solution o f
acetic acid, correspondingly.

H N(N0 )
2 2 2
+
^ H + N(N0 ) " .
+
2 2

Based on the experimental data, the temperature dependence o f the rate


constant o f decomposition o f the solution o f N H N ( N 0 ) 4 2 2 in acetic acid
takes the form:
k Qxp =2.5-10 exp(-4.4-10 IT), 2 3
s" . 1

Under an assumption o f equality o f the rate constants of a


monomolecular decomposition o f the protonated form in sulfuric and acetic
acids, the following relation for K was obtained: p

Kp =10 - e x p ( - 3 . 5 - 1 0 / T ) .
10 6 3

The concentration o f the protonated form o f dinitramide decreases as the


temperature increases. Thus, the temperature dependence o f k exp is weaker
in acetic acid than in water.
170 General Regularities of Thermal Decomposition of Onium Salts

Based on the data available, we formulated some preliminary


conclusions about the possible mechanisms o f thermal decomposition o f
various forms o f dinitramide. Note that little is known about their structure
and bond energies. A n a l y z i n g the mechanism o f thermal decomposition o f
dinitramide molecule, one should take into account that it is similar to
nitronic acids, w well known in the chemistry o f nitrocompounds.
The reaction o f ic transformation o f a nitro-form into an acido-
form is typical of >r nitramide, this is a tautomeric equilibrium:

0 N-NH
2 „ [0 N-N-N0 ]" + H —
2 2
+
0 N-N=NOOH.
2

In media with a high dielectric permittivity, for instance, in aqueous and


acid solutions, the processes o f ionization and exchange by protons proceed
at higher rates. Thus, the tautomeric equilibrium is rapidly attained.
In [11.23], based on the analysis o f the IR-spectra, a somewhat different
structure o f the acid acido-form was proposed:

0=N^ /N-O-
N"
In the analysis o f the decomposition mechanisms, one should take into
account the possibility o f the reaction through either the aci- or nitro-form
o f the acid. The decomposition o f one o f these forms, due to a rapid
attainment o f the tautomeric equilibrium, w i l l cause the reduction o f the
content o f both forms.
The activation energy o f decomposition o f the N ( N 0 ) ~ anion in a 2 2

solution is approximately 41 kcal/mol, which is close to the corresponding


values for the secondary nitroamines [11.31], for which the determining
factor is the N — N 0 2 bond breaking. This mechanism is also most probable
for the N ( N 0 ) ~ anion and is also proposed as the main one at thermal
2 2

decomposition of dinitramide salts. Nevertheless, to confirm this


mechanism, it is necessary to determine the energy o f an N — N 0 2 bond
independently.
Within the temperature range o f 0 - 1 0 0 ° C , the rate o f decomposition o f
H N ( N 0 ) molecules is higher by a factor o f 10 —10 than that in the case o f
2 2
5 7

anions. Even the higher ratio between the rates o f decomposition o f nitric
and perchloric acids and their anions, described above, is observed. In these
acids, decomposition occurs through the equilibrium formation of
anhydrides C 1 0 and N 0
2 7 2 5 with the following decomposition o f relatively
weak bonds C 1 0 — O C 1 0 3 3 and N 0 — O N 0 . The second and even higher
2 2

orders on the acid are observed.


Obviously, equilibrium formation o f dinitramide anhydride by the
reaction:

2 H N ( N 0 ) ^ H N(N0 ) 2 2 2 2 2
+
+ N(N0 ) " ^
2 2 H 0 + N 0
2 6 7
Thermal Decomposition and Combustion 171

also occurs to a certain extent similarly to the other acids, but the increase
o f the rate o f decomposition o f anhydride does not compensate its low
concentration in comparison with HN(N0 ) 2 2 molecules. Here, the
contribution o f anhydride decomposition is not o f vital importance. The
decomposition o f dinitramide proceeds as a monomolecular reaction, very
likely due to the presence o f a relatively weak N — N 0 2 bond. The
molecular decomposition becomes more efficient than the formation o f
anhydride and its subsequent thermal decomposition. A t low temperatures,
in systems containing nitric acid, the formation o f a mixed anhydride by
reaction (11.4.12) proceeds until a considerably high degree o f conversion
is achieved due to a high concentration o f H N 0 3 molecules and because
they are protonated considerably more easily. Thus, the relative contribution
o f the pathway through the anhydride decomposition increases considerably
and becomes the determining one.
A high value o f the effective order o f the N — N bond in the anion due
to the contribution o f resonant structures, in which this bond is a double
one, can serve as an explanation o f the increase o f the activation energy o f
decomposition o f the N(N0 ) ~ 2 2 anion in comparison with H1M(N0 ) 2 2

molecules. It should also be supposed that the decomposition o f the


molecules proceeds through the decomposition o f their nitro-form, in which
the main contribution to the bond order gives the following resonance
structure:

+ + *°
N ^ - N - N

Protonation o f H " N ( N 0 ) 2 2 molecules should not influence the strength


o f the N — N bond. The orders o f the N — N bond in H N ( N 0 ) 2 2 2
+
and
HN(N0 ) 2 2 are close to each other. However, at the attachment o f the second
proton to the H N ( N 0 ) 2 2 molecule, the possibility o f the decomposition o f
the protonated form by the reaction o f dehydration appears. The detachment
o f water has to proceed more easily in the O-protonated acido-form o f the
acid through a six-member cycle:

o" y-H
0=TsL /N-CT

This structure is similar to the aci-form o f the acid obtained in [11.23].


The kinetic regularities and composition o f the final products observed
experimentally can correspond to the following reaction schemes:
For decomposition o f N ( N 0 ) " anions:
2 2
172 General Regularities of Thermal Decomposition of Onium Salts

N ( N 0 ) " — N 0 + *N(N0 )-
2 2 2 2

N0 2 + *N(N0 )-— N 0 + N 0 " 2 2 3

For decomposition o f H N ( N 0 ) 2 molecules: 2

HN(N0 ) — N 0 2 2 2 + HNN0 2

H N N 0 2 — N 2 0 + O H

N0 2 + OH — HNO3

For decomposition o f protonated molecules o f H N ( N 0 ) 2 : 2 2


+

H N(N0 )
2 2 2
+
— H 0 + N 0 2 3 3
+

N 3 0 3
+
— N0 2
+
+ N 0 2

N0 2
+
+ H 0 — 2 HNO3 + H 0 3
+

A t low temperatures, autocatalysis is connected with the accumulation o f


nitric acid, the equilibrium reaction o f formation o f a mixed anhydride o f
nitric acid and dinitramide N 0 , and its subsequent decomposition. In 4 6

diluted aqueous solutions o f dinitramide and in solutions o f dinitramide in


aqueous sulfuric and acetic acids, autocatalysis is not observed because the
concentration o f the mixed anhydride N 0 is low due to the decrease o f the 4 6

concentration o f H N 0 molecules in the equilibrium reactions o f ionization:


3

HNO3+ H 0 — N 0 " + H 0 2 3 3
+

HNO3 + 2 H S 0 — N 0 2 4 2
+
+ 2HS0 " + H 0 4 3
+

and due to the formation o f nitronium acetate:

HNO3 + CH3COOH ^ CH COON0 3 2 + H 0 .


2

Thus, the kinetic regularities o f thermal decomposition o f dinitramide


are in complete agreement with the general scheme o f decomposition o f
other oxygen-containing acids but the quantitative relations o f the rates are
different.

REFERENCES

11.1. R. B e l l , Proton in Chemistry, M : M i r , 137-158 (1977).


Thermal Decomposition and Combustion 173

11.2. G . B . Manelis and Y u . I. Rubtzov, Fiz. Goreniya Vzryva, 6(1): 3 -


11 (1970).
11.3. G . D . Robertson, D . M . Mason, and W . H . Corcolan, J. Phys.
Chem., 59: 683-690 (1955).
11.4. A . I. Kazakov, L . P. Andrienko, and Y u . I. Rubtzov, Zh. Fiz.
Khim., 53(4): 1054-1055 (1979).
11.5. O . S. Galuk, Y u . I. Rubtzov, G . F . Malinovskaya, and G . B .
M a n e l i s , , Zh. Fiz. Khim., 39(9): 2319-2322 (1965).
11.6. L . N . Gal'perin, Y u . R. Kolesov, L . B . Mashkinov, and Y u . Y a .
Gerner, Proceedings of VI- th All-Union Conference on
Calorimetry, T b i l i s i : Izd. Metznieraba, 539-543 (1973).
11.7. E . M . Horn, H . Keiser, and K . Schoeller, Chemical Monthly,
118(111): 1205-1218 (1987).
11.8. A . I. Kazakov, L . P. Andrienko, and Y u . I. Rubtzov, Chemical
Physics of Combustion and Explosion. The Kinetics of Chemical
Reactions, Chernogolovka, Joint Institute o f Chemical Physics,
71-74 (1977).
11.9. H . S. Jonston, L . Foerihg, G . Tao, and G . H . Messerly, J. Amer.
Chem. Soc, 73: 2319-2325 (1951).
11.10. H . Harrison, H . S. Jonston, and E . R. Hardwick, J. Amer. Chem.
fee, 84: 2478-2483 (1962).
11.11. V . I . Atroshchenko and S. I. Kargin, Manufacturing of Nitric Acid,
M . : Goskhimizdat, p. 523 (1962).
11.12. H . Eiring and F. Daniels, J. Amer. Chem. Soc, 52: 1472-1486
(1930).
11.13. A . A . Chicherov, Y u . M . Kargin, L . M . A m i r o v a , and M . N .
Agafonov, Zh. Prikl. Khim., 63(1): 191-193 (1990).
11.14. A . V . Baranov, V . L . Pogrebnaya, and G . V . M a l ' c h i k o v a ,
Proceedings of Krasnodar Polytechnic Institute, 29: 179-184
(1970).
11.15. V . Y a . Rosolovskii, Chemistry of an Anhydrous Chloric Acid, M . :
Nauka, p. 140(1966).
11.16. Y u . I. Rubtzov, G . B . Manelis, Z . I. Grigorovich, and V . Y a .
Rosolovskii, Zh. Fiz. Khim., 48(6): 1394-1398 (1974).
11.17. Y u . I. Rubtzov, G . B . Manelis, Z . I. Grigorovich, and V . Y a .
Rosolovskii, Zh. Fiz. Khim., 48(6): 1399-1402 (1974).
11.18. R. W . Duerst, J. Chem. Phys., 48(5): 2275-2284 (1968).
11.19. I. B . Levy, J. Phys. Chem., 66(6): 1092 (1962).
11.20. B . L . Psikha and A . I. Kazakov, Theoretical Study of the
Mechanism of Decomposition of Chloric Acid, Preprint o f Joint
Institute o f Chemical Physics, Chernogolovka, p. 30 (1980).
11.21. O . A . L u k ' y a n o v , V . P. Gorelik, and V . A . Tartakovskii, Izv. Rus.
Akad. Nauk, Ser Khim., 1: 94 (1994).
11.22. O . A . L u k ' y a n o v , Y u . V . K o p n o v a , T . A . K l i m o v a , and V . A .
Tartakovskii, Izv. Rus. Akad. Nauk, Ser Khim., 7: 1264 (1994).
174 General Regularities of Thermal Decomposition of Onium Salts

11.23. V . A . Shlyapochnikov, N . O. Cherskaya, O. A . L u k ' y a n o v , V . P.


Gorelik, and V . A . Tartakovskii, Izv. Rus. Akad. Nauk, Ser Khim.,
9: 1610 (1994).
11.24. V . A . Shlyapochnikov, N . O . Oleneva, N . O. Cherskaya, O . A .
L u k ' y a n o v , V . P. Gorelik, and V . A . Tartakovskii, Izv. Rus. Akad.
Nauk, Ser Khim., 8: 1508 (1995).
11.25. G . B . Manelis, Proceedings of 26th INT Annual Conference of ICT
"Pyrotechnics: Basic Principles, Technology, Application",
Karlsruhe, 15.1 (1995).
11.26. T. B . B r i l l , P. J. Brush, and D . G . Patil, Combust. Flame, 92: 178
(1993).
11.27. A . I. K a z a k o v , Y u . I. Rubtzov, L . P. Andrienko, and G . B .
Manelis, Izv. Rus. Akad. Nauk, Ser Khim., 2129 (1997).
11.28. A . I. K a z a k o v , Y u . I. Rubtzov, L . P. Andrienko, and G . B .
Manelis, Izv. Rus. Akad. Nauk, Ser Khim., 40 (1998).
11.29. M . I. V i n n i k , Usp. Khimii, 35: 1922 (1966).
11.30. C . H . Rochester, Acidity Functions, London-New York: Academic
Press 300 (1970).
11.31. B . L . K o r s u n s k i i , L . Y a . Kiseleva, V . I. Ramushev, and F. I.
D u b o v i t z k i i , Izv. Akad Nauk SSSR, Ser. Khim., 1778 (1974).
Chapter 12

AMMONIUM NITRATE

12.1 General Regularities


Numerous studies o f thermal transformations o f ammonium nitrarte ( A N ) ,
for instance, [12.1, 12.5], revealed a series o f fundamental features, which
made it possible to formulate general concepts o f the mechanism o f these
transformations and to explain the observed regularities o f thermal
decomposition: (1) the rate o f decomposition o f A N is considerably higher
than the rates o f decomposition o f K N 0 , ammonium salts on the basis o f
3

stable acids, and ammonia, which allows us to conclude that a direct


decomposition o f N H and N 0 ~ ions does not sufficiently influence the
4
+
3

decomposition o f ammonium nitrate; (2) additives o f nitric and some other


mineral acids considerably accelerate the decomposition o f A N , and at the
same time, bases inhibit this reaction; (3) evaporation o f A N proceeds with
the formation o f N H and H N 0 .
3 3

Based on the facts presented above and on the data o f Chapter 11, one
can conclude that the main processes during A N decomposition are the
proton transfer with the formation o f N H and H N 0 dissolved in the 3 3

condensed phase and the subsequent thermal decomposition o f nitric acid


(the kinetic regularities o f this reaction were considered in Chapter 11) and
oxidation o f N H4 ion by H N 0 and N 0 molecules, formed from H N 0
+
3 2 5 3

by equilibria (11.2.3) and (11.2.4). It was shown by direct measurements


[12.6] that in the presence o f the N H ion, the rate o f consumption o f
4
+

H N 0 molecules is higher than in pure nitric acid due to the thermal


3

decomposition. This additional rate is directly proportional to C N H + . The

kinetics o f this reaction can be described by a bimolecular interaction o f


NH 4 ion and N 0
+
and H N 0
2 5 molecules. In this case, the relative
3

contribution o f the reaction through H N 0 molecules increases as the 3

175
176 Ammonium Nitrate

temperature goes up and C H N Q 3 decreases. In a generalized form, the rate


o f thermal decomposition o f A N is the sum o f three parallel reactions:

"^ ANC 1
dt = 2k C o { N2 5 +k NHi N 0
2
C C
2 5 +^3 NHC
4
+ C
HN0 3 • (12.1.1)

The k value corresponds to k in Chapter 11.


x 6

Usually, a sample o f A N contains a certain amount o f water (0.1-0.3%)


and it is very difficult to eliminate this water without disturbance o f the
balance between N H and H N 0 . There are no studies o f absolutely dry
3 3

A N . Water leads to a decrease o f the concentration o f H N 0 molecules due 3

to ionization by equation (11.2.5) and considerably reduces the C o N 5

concentration due to the shift o f equilibrium (11.2.3). Reducing the absolute


reaction rates, water increases the relative contribution o f the reaction
through H N 0 . In [12.6], it has been shown that oxidation by the H1M0
3 3

molecule w i l l be the main process, when the ratio C ^ Q . , / C H 0 2


l s
t> ^ e ow

0.006 at 2 0 ° C and below 0.12 at 2 0 0 ° C . Usually, samples o f A N meet this


requirement and the reaction in the samples proceeds through oxidation o f
NH 4by the H N 0 molecule. The main contribution gives the third term o f
+
3

equation (12.1.1). Since the rate o f N 0 decomposition is lower than that 2 5

o f oxidation o f the N H ~ ion by these molecules, under these conditions, the


4

contribution o f the first term o f equation (12.1.1) w i l l be smaller than the


contribution from the second term. Based on the kinetic data available, the
following temperature dependence o f k was obtained: 3

k = 10
3
9 4
exp(-30000//?r), 1/mol-s.

In systems with a high excess of H N 0 , 3 especially at lower


temperatures, the transition to oxidation by an N 0 2 5 molecule and its
thermal decomposition can proceed. In this case, the first two terms o f
equation (12.1.1) w i l l become the determining ones.
The literature data available on the kinetics o f decomposition o f A N
melt unambiguously confirm the correctness o f equation (12.1.1). In [12.1,
12.4-12.7], the most detailed and explicit studies o f this process were
performed. Thus, the kinetics o f thermal decomposition o f A N melt with a
humidity o f 0.2% and p H = 5.7 for a 10% aqueous solution was studied
within the temperature range o f 1 9 0 - 2 3 0 ° C at V 0
g
= 80-100 c m [12.1]. For 3

the ideal equivalent salt, the theoretical value is p H = 4.8. The sample
contained a very small amount o f excess N H . 3

This sample and the given experimental conditions most accurately


reproduce the decomposition o f a pure A N . The experimental technique is
similar to those employed in studies o f nitric and perchloric acids. The
kinetics is described by the equation o f autocatalysis o f the first order. The
Thermal Decomposition and Combustion 177

temperature dependences o f the rate constants are as follows:

k =10
x
144
e x p ( - 4 7 2 0 0 / 7 ? r ) , s~ , ]

k = 10
2
7 3
exp(-24000//?r), s" . 1

Since the activation energy o f the autocatalytic stage is lower, the self-
acceleration rapidly increases with the temperature decrease. The cause o f
acceleration is accumulation o f H N 0 , which was confirmed by the analysis
3

o f its content in the course o f the process. The increase o f the rate is
proportional to C H N > accumulated at the given time. This dependence is
O l

shown in F i g . 12.1.1. A t the temperature o f 2 1 4 ° C and V = 80 c m , 0


g 3

approximately 0.11 mole o f H N 0 is formed per 1 mole o f decomposed


3

A N . L o w e r rate constants were obtained than those presented in [12.2,


12.3], which was determined by the low value o f V . In this case, the liquid 0
g

phase is only slightly enriched with H N 0 , and, under such conditions, the 3

equivalence o f the salt remains almost unchanged. In [12.2], decomposition


was studied in an open system, and self-acceleration was not observed.
Nitric acid was not accumulated in the melt, being removed together with
the gas products.

o o.o<f o.od
HNOj,
C m o l / l

Fig. 12.1.1. Dependence of the rate of heat release during A N decomposition on the
concentration of m t n e
ampoule for different temperatures: (/) 205.4, (2)
214.2°C.

12.2 Influence of Water, Acid, and Ammonia on


the Decomposition Rate
A detailed study o f the influence o f H 0 and excess H N 0 on the rate o f 2 3

decomposition o f A N in the liquid phase was carried out in [12.1]. With


regard to (12.1.1), in all mixtures studied, the rate is directly proportional to
178 Ammonium Nitrate

C H N 0 ^ and decreases with the introduction o f water, although the


proportionality factors are different within various concentration ranges o f
the solutions due to the variation o f ionization equilibria (11.2.3-11.2.5).
Since the work was performed to estimate the safety treatment o f such
mixtures at elevated temperatures, the rates o f heat release related to 1 k g o f
the mixture are presented. From these data, for the rate constants k\ and k , 2

the following dependences were derived:

(1) A t C H 2 o > 1 . 8 , mol/1, k^f + l/C^o+bCHHoJC^o

k =p + r/C ,
2 H20 / = 10 8 7
exp ( - 3 7 2 0 0 / ^ 7 ) , s" 1

6 = 10 7 4
exp ( - 2 6 2 0 0 / i v r ) , s" , / = 10
1 156
exp (-51000/RT\ s~ ]

r = 10 - exp ( - 3 0 2 0 0 / i v F ) , s" , p = 10
,0 15 1 5 9
exp ( - 2 3 2 0 0 / 7 ? ! ) , s" . 1

(2) A t C H O = 0 . 1 7 - 1 . 8 , mol/1, k = / +1 / C
2 x H i 0 +a C m o C

/ = 10 , 3 6
e x p ( - 4 7 8 0 0 / i ? F ) , s , a = 10 ° e x p ( - 3 4 0 0 0 / i ? r ) ,
- 1 11
s" , 1

/ = 10 5 0
e x p ( - 2 8 0 0 0 / i v r ) , s" . 1
(12.2.1)

A t Cm^Q < 1.8 mol/1, the values o f k have not been studied in detail. They 2

can be obtained only by a linear interpolation between the k value for 2

C H 2 o =
1.8 mol/1 by equation (12.2.1) and the k value for the parent salt 2

for which C p H = 0 . 1 7 mol/1. In general, in the presence o f excess H N 0 , 3

the acceleration occurs only in the case when the H N 0 formed can 3

noticeably increase the general acid content in the system. A t the


introduction o f more than 0.15% o f H N 0 , no acceleration is observed. 3

The influence o f excess N H on the rate o f decomposition o f A N was 3

also studied. W i t h regard to (11.1.1), introduction o f N H should strongly 3

reduce the value o f C ^ N Q and the rate o f thermal decomposition. In the

presence o f water, equilibrium (11.1.5) competes with equilibrium (11.1.1),


and the concentration o f water molecules in the melt is higher, though their
basicity is lower than the basicity o f N H . Moreover, the volatility o f N H 3 3

under the temperatures studied is very high, and its major part passes into
the gas phase, although there are no data on N H pressure over the A N 3

solutions. Finally, the amount o f N H in solution is small, and the shift o f 3

equilibrium (11.1.1) is insignificant. A t 2 4 0 ° C , introduction o f 0.012 mol/1


of N H into A N melt at F = 80 c m reduces the rate o f decomposition by
3 0
g 3

20% and introduction o f 0.029 mol/1 o f N H by 50%. 3

Large additions o f ammonia cannot be studied due to a high initial


pressure in the ampoule. A s can be seen from these data, an insignificant
excess o f N H , no more than 10~ mol/1 in the parent sample, cannot
3
4

essentially change the rate o f decomposition. Since in this case a lot o f N H 3

molecules are accumulated in the gas phase, the self-acceleration o f the


Thermal D e c o m p o s i t i o n and C o m b u s t i o n 179

process is not observed for a long time. A m m o n i a binds H N 0 3 molecules


formed during decomposition, and its consumption in the melt is
compensated by dissolving o f the new portions o f N H 3 from the gas phase.
Naturally, the influence o f N H 3 strongly depends on V . 0
g
A l l the other
stronger or less volatile bases produce the same influence. D u r i n g the
heating time o f the sample, the exchange reaction is already proceeding
with the displacement o f the equivalent amount o f N H 3 and with the
formation o f nitrate o f the introduced basis. The following influence is
determined by the amount o f N H . 3

12.3 The Heat and Macrokinetic Regularities of


Decomposition
The oxidation reaction o f the N H 4
+
ion by the H N 0 3 molecule, which
determines the rate o f decomposition o f the majority o f A N samples,
proceeds through their nitration accompanied by the rapid decomposition o f
a protonated nitroamide:

N H A HN0 3 = H 0 + NH N0
2 3 2
+
= H 0 + H 0 + N 0•
2 3
+
2

Together with the proton transfer from H 0 3


+
to N 0 ~ , this results in the
3

realization o f the main pathway o f A N decomposition:

NH N0 4 3
1 , q
= N O
2
g a s
+ 2H 0 2
1 , q
' g a s
.

The possible heats o f A N decomposition were analyzed in [12.7]. For the


main reaction, i f water is formed in the gas phase, the heat o f decomposition
is 11.7 kcal/mol, and, i f water is formed in liquid phase, this value is 32.8
kcal/mol. Actually, the heat w i l l differ considerably with variation in the
experimental conditions. In addition to the main reaction, the following
secondary reactions also proceed:

NH N03
4
, i q
= N + 0.5O + 2 H 0
2 2 2
I i q
+ 52.4 kcal/mol

NH N0 =
4 3 0.8N + 1.8H O2 2
liq
+ 0.4HNO 3
I i q
+ 57.5 kcal/mol m

A t water separation in the gas phase, the heat o f these reactions w i l l be


reduced. The heat o f evaporation o f 1 mole o f water is approximately 10.5
kcal/mol. If the temperature decreases and the content o f excess acid
increases, the nitrogen yield also increases. A t temperatures within the
range o f 1 3 0 - 1 5 0 ° C , more nitrogen is formed than N 0 , while at 2 3 0 - 2

2 8 0 ° C , no more than 2 % o f nitrogen is evolved. Finally, in the limiting


cases, the heat o f decomposition o f A N can range from 11.7-12.0 kcal/mol
up to 52-55 kcal/mol. For any specific experimental conditions, the heat
180 Ammonium Nitrate

can be calculated from the measured amounts o f n^^ and « H M O 3 formed

during decomposition o f 1 mole o f A N with regard to the water fraction in


the gas phase:

e „ = 144 « H N O j =52.4 ( n - 2 n N2 ) + 32.8 -0.5n HN )•

-10.5w fe as
H 0 , kcal/mol,

_^H 0(^+^AN-2F
2 H 2 Q )
(12.3.1)
RT{\-pl V IRT)
H,0
0 Hi0

In accordance with (11.1.1), the gas phase over A N consists o f N H 3 and


HN0 3 molecules. The vapor pressure over A N was measured in [12.8] in
the flow o f inert gas by the mass o f evaporated substance. Under such
conditions, the pressures o f P N H and / H N O 3
a r e
equal to each other in the
gas phase. For the equilibrium evaporation we w i l l have:

^evap = / N H / H N o , = 0 . 2 5 ^
>
t a l =exp (AS e v a p /R)exp (AH^/RT).

The experimental temperature dependences o f the vapor pressure for solid


and liquid A N , as well as the values o f AH and AS o f the salt evaporation,
derived from these dependences and calculated from the values o f the
standard enthalpies and entropies o f the components [12.9], are presented in
Table 12.3.1.
The coincidence o f the experimental and calculated values confirms the
actual attainment o f equilibria (11.1.1) for A N . Since the volatility o f N H 3

is considerably higher than that o f H N 0 , in a closed volume, the liquid 3

phase is enriched with nitric acid which results in the dependence o f the
reaction rate on V . g
Based on equations (11.1.1) and (12.1.1) and with
regard to the mass balance at the salt dissociation, the dependence o f the
initial rate o f decomposition on F 0
8
takes the form:

^G 0 ,(12.3.2)
K.
y

1+
dt V 2
RT RT
Y
AN

where K is the equilibrium constant (11.1.1) for A N , K is Henry's constant


p G

for N H dissolving in the melt o f A N ,


3 F A N is the volume o f one mole o f
A N . The value o f k x
2
should depend linearly on V . For the data available, 0
g

at 2 0 3 ° C , this dependence is presented in F i g . 12.3.1. From the slope o f the


straight line it was obtained that K = 120 atm, w h i c h is close to the critical g

pressure o f N H 3 (115 atm) at the critical temperature o f 132°C. W i t h i n the


range o f V 0
g
up to 100 c m , the values o f k\ only slightly depend on V .
3
0
g
Thermal Decomposition and Combustion 181

With regard to the temperature dependences o f k and k , from { 3 (12.3.2) one


can obtain:

K =\0
p
745
exp(-34400//?r).

Table 12.3.1. Thermodynamic parameters of evaporation for solid and liquid A N

Ammonium P =f(T) AH, kcal/mol AS, kcal/mol • K


nitrate
experi- calcu- experi- calcu-
ment lation ment lation

Solid log(P)(mm)= 42.7 42.1 72.1 73.7


10.708-4670/T
Liquid log(P)(mm)= 39.9 39.0 65.4 66.7
9.981-4360/T

Fig. 12.3.1. Dependence of {dQ/dtf on V for A N decomposition.


0
g

The degree o f self-acceleration o f decomposition (constant k ) is determined 2

by the accumulation o f H N 0 , and this value virtually does not depend on


3

VQ because, for considerable excess o f H N 0 , the equilibrium amount o f


3

N H is small and its redistribution between the gas and liquid phases is not
3

important.

12.4 Thermal Decomposition below the Melting


Temperature
O n transfer to the solid phase, the reaction rate o f thermal decomposition is
usually strongly reduced due to the counteraction o f the crystal lattice to the
proceeding o f the elementary act o f the reaction. A N melts at 169°C, and
182 Ammonium Nitrate

the rate o f its decomposition at lower temperatures is very low.


Nevertheless, sometimes it is necessary to have at least very rough
estimates o f this value. A s was shown in Chapter 2, at temperatures below
the melting temperature but close to it, the observable rate o f decomposition
is determined mainly by the reaction in the liquid eutectic inclusions and
has a complex and very strong temperature dependence, which is far from
the temperature dependence o f the rate for the solid phase. The data
presented in [12.10], where the authors tried to determine the kinetic
parameters o f A N decomposition for the solid phase, were derived exactly
within this range, and large values o f the activation energy presented in this
work cannot be considered as the true values o f the activation energy o f A N
decomposition. The data on the A N decomposition in the solid phase are
not available.
Since all A N samples possess a particular humidity and this water
cannot be eliminated by ordinary drying without the loss o f essential
amounts o f ammonia, and the eutectic A N — H 0 is crystallized only at the
2

temperature o f - 1 6 . 9 ° C , for all higher temperatures, a certain amount o f the


liquid solution o f A N in water is contained in the samples. In addition, all
excess H N 0 contained in the sample is also concentrated in the liquid
3

phase. Thus, the contribution o f the reaction in these liquid inclusions can
be an essential and even determining one. The initial rate o f the reaction in
the liquid inclusions was calculated from the water and acid content in the
sample, the solubility o f A N , and the kinetic parameters o f the reaction in
the solution [12.5]:

(dQ/dtl__ =g o (x +y )[f
0 0 + \8 / ( x + ^ )
0 0 (g + l)x V +0.29Z) y
0
l
0 x]
0
]

g = 10 32 5 > b K
^ ^xp{[-8350-4180y (x + ^ )
o )
0 0 0 (12.4.1)

where g is the solubility o f A N in the solution H 0 — H N 0 calculated from2 3

the data presented in [12.11], x , y are the mass fraction o f H 0 and H N 0


0 0 2 3

in the sample o f A N , / /, and b are the parameters o f equation (12.2.1), p is


the solution density.
The experiments on the estimation o f the rate o f decomposition o f the
sample previously studied and, also, two samples o f A N with excess o f
H N 0 , obtained from the parent one by exposure in H N 0 vapors for 5 and
3 3

15 min, were conducted at temperatures o f 1 2 3 - 1 4 0 ° C . The rate o f


decomposition was determined by the amount o f N and N 0 evolved inside 2 2

sealed ampoules under thermostatic control during 1-90 days. The gases
were analyzed by the mass-spectrometric technique. The results obtained
are presented in Table 12.4.1 in comparison with the calculation by relation
(12.4.1). In accordance with the analysis o f the decomposition products and
experimental conditions, with regard to relation (12.3.1), the heat o f this
process is 45 kcal/mol. For the parent A N without excessive acid, the
reaction in the liquid inclusions gives 3 0 - 3 5 % from the total rate o f
Thermal Decomposition and Combustion 183

decomposition and in acid samples the calculated rate is even higher than
the experimental one. The contribution o f the reaction in the solid phase can
be neglected. Apparently, relationship (12.4.1) can be used to estimate the
rate o f A N decomposition below the melting temperature for neutral
samples within a humidity o f no less than 0.2% and for any acid samples.
Direct measurements at temperatures are very complicated and time
consuming. N o w , there are no reliable data on the rate o f decomposition
inside the crystal lattice o f A N . In order to obtain such data, an extra dry
sample with the ideal balance o f H N 0 3 and N H is required. What is more,
3

the eutectic mixture o f A N with the dissociation products o f the salt is


always present in the sample.

Table 12.4.1. The rate of thermal decomposition of the samples of A N with x = 0

0.002 below the melting temperature

Sample y T,°C w 10 , s"10 1

experiment calculation

the parent 0 131 0.025 0.008


the parent 0 140 0.053 0.03
1 2.8- 10~
4
123 2.5 4.7
1 2.8- 10- 140 17.0 23.0
4

2 1.3- io- 3
123 12.0 21.0
2 1.3- io- 3
140 21.0 102.0

12,5 Influence of Additives on the Rate of Thermal


Decomposition
There are some chemical compounds whose introduction into A N can
strongly change the rate o f decomposition, and this should be taken into
account in numerical estimations o f the rate o f thermal decomposition. The
accelerating influence o f C l ~ ions is the most pronounced one [12.12-
12.14]. Nitric acid, used for A N synthesis, always contains a particular
amount o f C P ions (usually at the level o f 10" %). Almost all o f these ions
4

pass into A N , and such amounts already noticeably influence the rate o f
decomposition. Under comparable conditions, the quantitative results were
obtained in [12.14], Introduction o f C P ions increases both the initial rate o f
decomposition (k\) and the degree o f self-acceleration (k ). Particular values2

o f the rate constants at 2 0 0 ° C are presented in Table 12.5.1. The catalytic


influence o f C P ions increases directly proportionally to the H N 0 content 3

in the system. Thus, the degree o f acceleration increases considerably more


l ii
than the initial rate. The increase o f k and k is proportional to ( C , )
x 2 c

In order to reduce the accelerating influence o f C P ions, one should


184 Ammonium Nitrate

eliminate excessive F T N 0 from the system by binding it with the help o f


3

the appropriate compounds, for instance, the salts o f phosphoric acid, w h i c h


possess buffer properties. A t the introduction o f diammonium phosphate
( D A P h ) into A N , the binding o f H N 0 proceeds by the reaction:
3

HN0 3 + HPO4" 2
— N C V + H2PO4-.

Introduction o f 0 . 2 - 0 . 5 % o f D A P h into the melt o f A N for a sufficiently


long time completely eliminates self-acceleration o f the decomposition
without decreasing the initial reaction rate. After consumption o f all D A P h ,
the rate begins to grow and attains almost the same values as for samples
without additions o f D A P h . Introduction o f D A P h considerably increases
the time before rapid self-acceleration begins.

Table 12.5.1. The influence of CI ions on the rate constants of decomposition of


A N at the temperature of 200°C

[CP], % 0 0.0005 0.005 0.014 0.5

k • 10 , s'
{
6 1
0.19 0.36 0.42 0.63 0.93
k - 10 , s-
2
6 1
34 66 450 700 1060

Fig. 12.5.1. The influence of D A P h introduction on the rate of decomposition of A N


with an impurity of C P : (7) the parent A N , (2) A N + 0.05% CP, (3) A N + 0.05% C P
+ 0.3% DAPh, (4) A N + 0.05% C P + 0.35% DAPh.

The influence o f the addition o f D A P h on the rate o f decomposition o f A N


in the presence o f C P ions is shown in F i g . 12.5.1. Undoubtedly, the content
o f C P ions in the sample o f A N determines the rate o f its decomposition to
a great extent.
Practically all ions o f metals with variable valence exert a considerable
accelerating influence on the rate o f A N decomposition. Such influence
strongly depends on the phase state o f the additive and its solubility in the
Thermal Decomposition and Combustion 185

A N melt. For insoluble additives, the composition and the surface quality
are o f considerable importance.
The quantitative description o f the accelerating influence o f additives
and their comparison with each other can be carried out only for soluble
additives (in the form o f metal nitrates). For relatively low concentration o f
the additive (up to 0.1 mol/1), the increase o f the initial rate is proportional
to the concentration o f ions:

(drj/dt) = kC d d +£,,

where k\ is the rate constant o f decomposition o f the parent A N .


The values o f k for some ions at 180°C are presented in Table 12.5.2.
d

A t this temperature, we have k\ = 5 - 1 0 -9


s . A catalytic influence o f C r ,
_1 + 3

F e , and C u
+3 + 2
is considerable. For 10~ mol/1 (0.05-0.1%), they can
2

increase the rate o f decomposition by a factor o f 200-400. The influence o f


other ions is weaker, and the P b + 2
ions almost do not accelerate the
decomposition o f A N . If ions exist in insoluble form (in the form o f oxides),
their catalytic influence is considerably weaker. A s the initial rate increases,
the additives practically do not increase the self-acceleration o f the reaction.
A t a considerable increase o f the initial rate, the self-acceleration becomes
almost insignificant. O n l y at the introduction o f C r + 3
are the maximal rates
higher than the initial rate by a factor o f 2.5-3. Apparently, the majority o f
ions accelerate the pathways o f oxidation o f the N H 4
+
ion, which do not
lead to the formation o f nitric acid.

Table 12.5.2. The influence of metal ions on the rate constants of decomposition of
A N melt at a temperature of 180°C

Additive Fe +3
Co +3
Cr +3
Cu + 2
Zn + 2
Mn + 2

V io , 6
14 0.9 24 12 3.5 9
1/s • mol

In [12.15], variations o f the rate o f decomposition o f the components in


mixtures o f A N with explosives at a ratio o f 1 : 1 were studied. Various
aromatic nitrocompounds were used: nitrobenzene, o-nitrotoluene, o-
nitroaniline, trinitrotoluene, trinitroaniline, 1,3,5-triamino-2,4,6-trinitro-
benzene, and similar compounds, as well as some organic nitrates such as
P E T N and cellulose nitrate. The thermographic technique and analysis o f
the variation o f the component content in the mixture were employed. It
was revealed that A N increased the rate o f decomposition o f practically all
substances studied, and the influence o f the organic explosive strongly
depended on its structure and thermal stability. Nitroamines noticeably
decrease the rate o f A N decomposition in accordance with their main
186 Ammonium Nitrate

properties, and nitrocompounds usually increase the rate of AN


decomposition. This increase is proportional to the rate o f decomposition o f
the compound studied. The observed regularities are determined by the
catalytic influence o f nitric acid, formed from A N during its equilibrium
dissociation, and by interaction of A N with N0 , 2 formed during
decomposition o f nitrocompounds, and the influence o f nitroamines is
connected with the corresponding decrease o f the equilibrium concentration
ofHN0 . 3

Organic compounds occupy a unique position among the additives. A


systematic study o f such mixtures was performed. There are only some
works on lubricants and surface coats for A N . The results available are
difficult to generalize due to the various nature o f the substances
investigated, but one can note that in the majority o f cases, oxidation o f
organic additive by the melt o f A N proceeds at a sufficiently high rate and
at temperatures o f 170°C and even higher, when the A N melt exists. H N 0 3

molecules are very active oxidants and react very rapidly with almost all
molecules capable o f oxidation.
During decomposition o f the melts o f commercial A N samples,
increased heat release is frequently observed at the beginning o f the
process, which is determined by the occurrence o f a small amount o f an
organic impurity in the sample. After consumption o f this impurity, the
reaction is terminated.
This process is not o f considerable importance for the estimation o f the
safety o f A N production and storage. The total amount o f heat is small
because o f a small impurity content. The situation when the amount o f
organic additive is higher and the heat o f oxidation is essential is much
more dangerous.
In mixtures o f A N with hydrocarbons, a zero oxygen balance is attained
when there is approximately 5% fuel in the mixture. Such a mixture is a
typical explosive. For a mixture o f A N with a 7 - 8 % hydrocarbon o i l , the
initial rate o f heat release is higher than that in the A N melt by a factor o f
5 0 - 6 0 and the activation energy is E act = 13 kcal; the additional heat o f
decomposition is approximately 330 kcal per 1 kg o f the mixture.
Even silicon oils and liquids, which are particularly stable to oxidation,
do not always preserve these properties as applied to the A N melt. The rate
o f heat release can increase considerably. Probably, this is determined by
the presence o f the impurity in silicone, because the heat effect is small, 2 0 -
200 cal/kg.
The presence o f significant amounts o f organic substances in the A N
melt almost always leads to an increase o f the initial rate and an increase o f
the heat release o f the process due to oxidation o f this additive. The ultimate
increase o f the heat release can be up to 11000 kcal per 1 k g o f organic
additive in the mixture. The presence o f approximately 1% o f impurity can
considerably facilitate the conditions for the heat ignition o f such melts.
Thermal Decomposition and Combustion 187

REFERENCES

12.1. Y u . I. Rubtzov, A . I. Kazakov, S. Y u . M o r o z k i n , and L . P.


Andrienko, Zh. Prikl. Khim., 57(9): 1926-1929 (1984).
12.2. P. I. Robertson, J. Soc. Chem. Industry, 67: 221-225 (1948).
12.3. W . A . Rosser, S. H . Inami, H . Wise, and B . I. W o o d , J. Phys.
Chem., 67(9): 1753-1757 (1963).
12.4. Y u . I. Rubtzov, A . I. Kazakov, N . G . Vais, A . P. Alekseev, I. I.
Strizhevskii, and E . B . M o s h k o v i c h , Zh. Prikl. Khim., 61(1): 1 3 1 -
132 (1988).
12.5. Y u . I. Rubtzov, A . I. K a z a k o v , L . P. Andrienko, I. I. Strizhevskii,
and E . B . M o s h k o v i c h , Zh. Prikl. Khim., 60(1): 3 - 7 (1987).
12.6. A . I. Kazakov, Y u . I. Rubtzov, and L . P. Andrienko, Izv. Akad.
Nauk SSSR, Ser. Khim., 5: 972-977 (1980).
12.7. Y u . I. Rubtzov, A . I. K a z a k o v , and L . P. Andrienko, Chemical
Physics of Combustion and Explosion. Kinetics of Chemical
Reactions, Chernogolovka: Joint Institute o f Chemical Physics,
86-90 (1989).
12.8. G . D . Brandler, N . M . Junk, and G . W . Lawrence, J. Chem. and
Eng. Data, 7(2): 227-228 (1962).
12.9. Thermal Constants of Substances, Reference Book, V . P. Glushko
E d . , M . : Izd. V I N I T I , 1-10 (1963-1981).
12.10. E . B . M o s h k o v i c h , G . N . Podshivalova, I. Y u . Sidorina, and I. I.
Strizhevskii, Zh. Prikl. Khim., 55: 901-904 (1982).
12.11. P. I. Protzenko, O . I. Razumovskaya, and I. A . B r y k i n a , Reference
Book on Solubility Nitrate and Nitrite Salt Systems, Leningrad: Izd.
K h i m i y a , 272 p. (1971).
12.12. A . G . Keenan and B . Dimitriades, J. Chem. Phys., 37(8): 1 5 8 3 -
1587 (1962).
12.13. A . G . Keenan, K . Natz, and N . B . Franco, J. Amer. Chem. Soc,
91(12): 3168-3171 (1969).
12.14. Y u . I. Rubtzov, I. I. Strizhevskii, A . I. K a z a k o v , E . B .
M o s h k o v i c h , and L . P. Andrienko, Zh. Prikl. Khim., 62(11): 2 4 1 7 -
2422 (1989).
12.15. J. C . Oxley, J. L . Smith, and Wen Wong, J. Phys. Chem., 98(14):
3893-3900, 3901-3907 (1994).
Chapter 13

AMMONIUM PERCHLORATE

The number o f works on various aspects o f the kinetics and mechanism o f


the thermal decomposition o f ammonium perchlorate ( A P ) exceeds the
number o f studies o f the thermal decomposition o f all other onium salts
taken together. M a i n l y , there are two reasons for this: (1) A P is the main
oxidizer in solid rocket propellants and the kinetic regularities o f its
decomposition largely determine the stability o f propellants prepared on its
basis and regularities o f their combustion (these relations w i l l be considered
below); (2) in studies o f A P decomposition, many peculiarities o f the
development o f the reaction o f thermal decomposition in crystal lattice were
revealed, which are o f general theoretical interest and for which A P was
found to be a convenient model. There are several detailed reviews on the
thermal decomposition o f A P (for example, a review by H a l l and Pearson
[13.1], a review by Keenan and Siegmund [13.2], collected articles [13.3],
and a book by Solymosi [13.4]).
Certainly, we cannot discuss here the details o f the works available. W e
w i l l present the kinetic regularities o f decomposition and consider the
peculiarities o f the reaction in a crystal lattice. Similarly to ammonium
nitrate, the primary chemical stage o f thermal decomposition o f A P is its
equilibrium dissociation into ammonia and perchloric acid with the
following decomposition o f perchloric acid during its interaction with N H 4
+

ions. The temperature dependence o f the vapor pressure over A P was


determined in [13.5] and was found to be as follows: log(PMivi) =
-
66283.1IT + 10.56. It almost coincides with the dependence obtained by
thermodynamic calculation with regard to the salt dissociation in vapors
into N H and H C 1 0 . A proton character o f the electric conductivity o f A P
3 4

crystals was experimentally revealed in [13.6].

13.1 Kinetics of the Low-Temperature


Decomposition
The formal kinetic regularities o f A P decomposition were studied within a
very wide temperature range using a combination o f various techniques. A P

189
190 Ammonium Perchlorate

was preliminary purified by double recrystallization from distilled water,


was dried at 9 0 - 1 0 0 ° C , and after grinding, a fraction o f (0.5-1.0) • 10" m in
4

size was selected. Usually, based on kinetic regularities, the following two
regions o f A P decomposition are distinguished (although this division is
rather arbitrary): the low-temperature region (up to 3 0 0 - 3 2 0 ° C ) and the
high-temperature region. A t relatively low temperatures o f 1 9 5 - 2 9 0 ° C , the
rate o f decomposition was measured by mass loss and the rate o f heat
release in open and sealed ampoules [13.7-13.8]. T y p i c a l kinetic curves are
shown in F i g . 13.1.1. W i t h i n this temperature range, the reaction has two
characteristic features i f the decomposition is conducted in an open system:

Fig. 13.1.1. Dependence of the rate of heat release during A P decomposition on the
external conditions (229°C): (7) the open ampoule, (2) V = 1.7 • 10 , cm , (3) V =
0
g 4 3
0
g

2.6 10 , cm .
3 3

(1) H a v i n g an autocatalytic character, the reaction is sharply inhibited after


decomposition o f 3 0 - 3 5 % o f all A P , but complete termination is not
attained. The decomposition continues at a relatively low rate, although the
residual represents almost pure A P . After treatment o f this residual by water
vapors at room temperature, the decomposition proceeds again by 25-35%)
in the new experiment.
(2) A t a temperature o f 2 4 0 ° C , A P is subjected to a phase transition from
the or/Zzorhombic into the cubic lattice, which is accompanied by a
considerable decrease o f the rate o f thermal decomposition. In a cubic
lattice, the reaction develops more slowly. A t a temperature o f 1 9 5 - 2 4 0 ° C ,
the kinetic curves are described by the equation o f autocatalysis in the
following form:

df]/dt = K (]-ri) +K N(L-ri?-


x
L5
2
5
9

and at a temperature o f 2 4 0 - 2 8 0 ° C , by the equation o f the second order o f


autocatalysis o f the first order. The value o f k was determined from the
2
Thermal Decomposition and Combustion 191

rectification o f the rate curve in appropriate coordinates, but the values o f k\


are small and cannot be determined directly. They were calculated from the
integral form o f the kinetic equations, which are the following:

2
~(1 + TJ )JT^}
0 (l + 7 o ) 1 5 n
( V ^+ V ^ 0 ) ( 1
~ V ^ o ' )
for(195-240°C),
1 77 1 7 + 7o ,
k t=2 — + In , n A = k\ I kj
l + T/o 1 - 7 (1 + 7 ) 2 7
/od-7)
for ( 2 4 0 - 2 8 0 ° C ) . (13.1.1)

Here, the value 77 = 1 corresponds to the degree o f conversion o f A P which


is equal to a = 0.3-0.35.
c

The value o f k\ preserves the same temperature dependence in both


crystalline modifications, and k , at the point o f phase transition, decreases
2

sharply and, at temperatures above the temperature o f the phase transition,


has a weaker temperature dependence:

k = 10
x
8 0
exp(-30000/#r), s" 1
(195-280 C ) ,

yt = 1 0
2
102
exp(-30000/#r), s"' (195-240 C ) ,

k = 10
2
4 5
exp(-18100//?r), s -1
(240-280 C ) .

Within the temperature range o f 2 0 0 - 2 8 0 ° C , the heat o f thermal


decomposition o f A P in an open system is 40.9 ± 1.3 kcal/mol and is almost
constant at phase transition. Based on numerous literature data, the main
equation o f the reaction o f A P decomposition can be presented in the
following form:

4NH4CIO4 = 8H 0+2C1 + 3.50 +N + N 0


2 2 2 2 2

The heat o f decomposition presented above corresponds to this equation.


The ratio between N and N 0 changes slightly as the temperature varies.
2 2

The kinetic regularities are essentially changed i f the decomposition


proceeds in a sealed and completely thermostatically controlled ampoule.
A n example o f the curves o f the rate o f heat release is shown in F i g . 13.1.1.
The value o f k increases with the growth o f V , and, in general, the
x 0
8

temperature dependence is close to those presented above. The value o f k is 2

smaller by a factor o f 1.5-2. The values o f the maximal rate are also
smaller, but they are attained earlier due to the growth o f k\. A sharp
deceleration o f the decomposition after consumption o f 3 0 - 3 5 % o f A P is
weakly expressed. The decay o f the rate proceeds rapidly only until a
192 Ammonium Perchlorate

particular, measurable and rather high level. The smaller the V 0


8
value the
higher this level is. After this, the rate decreases considerably more slowly
until the complete decomposition o f all A P occurs.
If the value o f F 0
g
decreases, starting from its particular value, the peak
o f heat release appears at the second, slower stage, which is immediately
replaced by its sharp decay almost to zero (curve 3, F i g . 13.1.1). A t the
further decrease o f the value o f V , this peak shifts towards earlier reaction
0
G

stages, so that the ratio Q /V T 0


G
remains constant, where Q is the total
T

amount o f heat released up to the moment o f the appearance o f this peak.


Consequently, the appearance o f this peak is unambiguously connected with
the attainment o f a particular pressure o f the decomposition products in the
ampoule, w h i c h is constant under a given temperature. In [13.9], it was
revealed that decomposition o f A P at 1 3 0 - 1 9 0 ° C is stopped, i f the pressure
o f water vapors, which can form in the liquid phase on the surface o f
crystals, is attained in the vessel. Unfortunately, the quantitative data are
presented only for the temperature o f 150°C, but their joint consideration
with the experimental data described above, which were carried out at
higher temperatures, confirms the determining role o f condensation o f water
vapor in the reaction termination with the formation in the crystal o f the
liquid saturated solution o f A P in water.
As V 0
G
decreases, the equilibrium pressure is attained earlier, but the
first maximum o f the rate always precedes the peak o f heat release and the
reaction halt, connected with the formation o f the liquid phase. Before the
peak, one can observe the decrease o f the rate. A t a temperature o f 2 2 9 ° C
and V 0
8
> 2600 c m , the value o f the first maximum o f the rate is constant,
3

but with the decrease o f V 0


G
this value also decreases. A particular critical
pressure o f the decomposition products exists, and after this pressure is
attained, the rate o f decomposition decreases independently o f the degree o f
conversion achieved at a given time. The complete reaction halt due to the
formation o f a liquid solution is always preceded by the decrease o f the
reaction rate, probably due to the beginning o f condensation o f water vapor
on the defects o f the crystal lattice o f A P . The peak o f heat release by itself,
before the decomposition halt, is connected with the heat release o f water
condensation. In the measurements o f the rate o f gas release, such a peak is
not observed.
A b o v e the point o f phase transition, the kinetics o f heat release in a
closed system becomes more complicated, and the differences with an open
system become more essential. In pre-evacuated ampoules, the induction
period is poorly reproduced and all its values are considerably higher than
those in the open ampoule. If evacuation is carried out for several minutes
at a temperature o f 2 5 0 ° C and the ampoule is sealed only after that, the
reaction proceeds without the induction period. Consequently, at the phase
transition, the inhibitors o f the reaction are separated from the crystal. M o s t
probably, this is the adsorbed or occluded water. In an open system, water is
eliminated and does not influence the development o f the decomposition.
Thermal Decomposition and Combustion 193

The first maximum o f the rate is considerably lower, and no more than 4 - 5
cal/g are released before the maximum is attained, which corresponds to rj =
0.015-0.02 and its value does not depend on V . Q
g

Further development o f the reaction is similar to the kinetics before the


phase transition, but the peak o f heat release is expressed considerably more
strongly due to the greater accessibility o f the reaction products o f all
regions o f the crystal which experienced phase transition. A t a constant
temperature, the ratio Q /Vo also remains constant, but it increases with the
t
g

temperature increase. This makes it possible to estimate the equilibrium


pressure o f water vapors over the saturated aqueous solution o f A P ( P \ \ Q )
with the use o f the data on the water content in the decomposition products.
The values o f P^ Q obtained together with the vapor pressure over pure

water (P^ Q ) are presented in Table 13.1.1. For a temperature o f 150°C,

the calculation is carried out with regard to the data [13.9]. The temperature
dependence o f P^^Q takes the form:

PH o = 10 5 4
exp(-10300//?r), atm.

Table 13.1.1. The equilibrium water pressure over the saturated aqueous solution
of A P

r,°c "AP H 0
1 n
2

150 1.6 0.33 1.02


229 9.5 0.34 0.97
250 11.6 0.29 1.25
260 13.8 0.30 1.18
270 16.6 0.30 1.20

The value o f AH in the exponent is close to the evaporation heat o f water,


which confirms its determining role in the reaction halt. A s s u m i n g that the
solution is ideal and with regard to dissociation o f A P , the solubility o f A P
in water at corresponding temperatures was calculated. It was revealed that
less than 1 mole o f H 0 corresponded to 1 mole o f A P , which results in
2

strong competition in the hydration o f ions and hampers the formation o f


high hydrates o f perchloric acid.
W i t h temperature growth, after the formation o f the liquid phase, the
reaction rate becomes more pronounced. A t a temperature o f 2 7 0 ° C , this
rate is comparable with the value o f the first maximum o f the reaction rate.
This rate slowly increases in the course o f the reaction, which can be
explained by the decomposition o f a saturated aqueous solution o f A P . Its
194 Ammonium Perchlorate

mass increases due to the condensation o f the water produced and the
concentration is constant until complete dissolution o f A P . In the solid
phase, the reaction does not occur because o f a preferential dissolution o f
the lattice defects. F o r this stage, a value o f activation energy o f 38 kcal/mol
was estimated.
Unfortunately, we could not study this stage in detail because, under
higher temperatures, the strength o f a glass ampoule is insufficient to hold
the products in the liquid phase. The heat o f decomposition in a closed
system considerably changes as the value o f K varies due to the 0
8

redistribution o f water between the liquid and solid phases and due to a
pronounced inhibition o f the reaction after the liquid phase formation. In the
majority o f the experiments, complete decomposition was not achieved.

13.2 Kinetics of the High-Temperature


Decomposition

A t temperatures up to 3 9 0 ° C , the kinetics o f decomposition o f A P was


studied by weight analysis in open glass ampoules [13.10]. A b o v e 2 8 0 ° C ,
the shape o f the kinetic curves essentially changes. Self-acceleration o f the
reaction becomes smaller and then disappears; the rate is maximal at the
beginning o f the process, which is determined by higher activation energy
for k\ than for k . Simultaneously, at the first stage, the degree o f conversion
2

also decreases and at 3 2 0 ° C , it can be only 0.05, but the rate o f


decomposition during the second stage increases. A b o v e 3 2 0 ° C , the
decomposition can proceed up to 100% during an acceptable time. D u r i n g
the first stage, within a temperature range o f 2 8 0 - 3 2 0 ° C , the variation o f
the degree o f conversion ( a ) is described by the following dependence:
d

a, -0.05 , ,
7

= 10" exp(45700//?r).
1 7 3

0.35 -a d

A t the first stage, above 3 0 0 ° C , the reaction time becomes comparable with
the heating time o f the substance, and, in this case, it is impossible to study
the kinetics o f this stage by weight analysis. During the second stage, above
3 2 0 ° C , along the largest portion o f the curve, the rate is described by the
equation o f the reaction o f the zero order and only at the end o f the process
does it start to decrease. The temperature dependence o f the rate constant
(k ) takes the form:
3

k = 10
3
6 7 5
exp(-32500//?r), s" .
1
Thermal Decomposition and Combustion 195

Within the temperature range o f 3 9 5 - 4 6 4 ° C , the kinetics o f the first stage


was studied from the rate o f gas release during A P decomposition on a
nichrome plate with a mica substrate, which was rapidly heated by the
electric current at a pressure o f nitrogen o f 0.2 atm. Since the walls o f the
vessel are cool, these conditions are close to decomposition in an open
system. Water and the other products with a high boiling point are
condensed on the walls. Sublimation proceeds simultaneously with
decomposition, and the amount o f the sublimate was taken into account in
the calculation o f the rate o f decomposition. Under such conditions, two
stages o f decomposition are clearly distinguished. The first stage is
described by the kinetic equation o f the first order, and the temperature
dependence o f k takes the form:
x

A, = 1 0 e x p ( - 3 2 0 0 0 / / ? r ) ,
90
s" , 1

which practically coincides with the values o f k\ obtained at lower


temperatures. W e did not succeed in the study o f the kinetic parameters o f
the second stage by this technique.
A t higher temperatures, the decomposition o f A P was also studied by
the linear pyrolysis method [13.10]. U p to temperatures o f 5 0 0 - 5 5 0 ° C , the
stationary linear decomposition o f the substance is realized, which is
determined by the kinetic parameters o f the volume reaction within the
heated layer that leads to a complete gasification o f A P . This corresponds to
the second reaction stage. The values, which are sufficiently close to k at 3

lower temperatures, were obtained:

k = 10
3
6 8
exp(-30200//?r), s" .
1

A b o v e the temperature o f 5 5 0 ° C , pyrolysis passes into a different mode,


which is determined by the combustion o f gas-phase products near the
plate. In this case, it is impossible to determine the kinetic parameters. A t
high temperatures, the decomposition o f A P was studied in [13.11] by the
quantitative thermography method within a narrow layer between titanium
plates. The rate was recorded by a time-of-flight mass spectrometer from
the total intensity o f the peaks o f HC1 and C l . A l l characteristic features o f
2

the reaction were confirmed, and, in general, the values o f the rate constants
obtained with regard to the difference in the conditions o f decomposition
are in good agreement. Slightly higher values o f the rate constants presented
in [13.11, 13.12] can be connected with catalysis by the metal surface o f
heating elements.
Based on the data available, one can distinguish between two stages in
the thermal decomposition o f A P :
(1) The initial relatively rapid reaction with a considerable self-acceleration
at low temperatures. In this case, at low temperatures, the value o f a can be
d
196 Ammonium Perchlorate

as high as 0.30-0.35 and, at higher temperatures, it can be 0.05-0.1. T h e


parameters o f this stage characterize the constants k and k . x 2

(2) The second, slower stage proceeds with the rates, which are sufficiently
large to be measured only within the high temperature range. In this case,
the decomposition o f A P proceeds without self-acceleration. The rate is
determined by constant k . In particular, the inclusion o f this stage is the
3

characteristic feature o f high-temperature decomposition, although with


very low rates it also proceeds under lower temperatures. The main kinetic
regularities o f A P decomposition within the temperature range o f 2 0 0 -
6 0 0 ° C are presented in Table 13.2.1. The values o f k are presented only i n
2

[13.10], and they were cited above. Thus, at least until 6 0 0 ° C , the kinetic
parameters o f decomposition o f A P are preserved, although the contribution
o f separate stages changes as the temperature and experimental conditions
vary. This provides good opportunities for calculation o f the rates o f
combustion.

Table 13.2.1. The kinetic regularities of thermal decomposition of A P at 195-600°C

Method &T,°C The kinetic law k = faexp(-EZRT) Refs.


k, s" 1
E, kcal

The first stage

Gravimetric 195-280 drj/dt- fa(l-r/) + 10 30.0 [13.10]


2 80

farid-n)2

Gas 395-453 10 ° 9
32.0
separation difdt = fa(l-ri)
C l + HCl 240-600 10 82
27.0 [13.13]
drj/dt- fa(l-ii)
2

separation

The second stage (fa)

Gravimetric 260-380 drj/dt = fa IO 32.5 [13.10]


675

Linear 280-500 drj/dt = fa io - 6 8


30.2 [13.10]
pyrolysis
Thermography 340-435 drj/dt = faO-rjf IO 32.0 [13.11]
5 765

Product 240-600 drj/dt = fa(l-ij) 10 32.0 [13.13]


72

separation

13.3 Influence of Excessive Acids and Bases


In order to confirm that the first stage o f decomposition o f A P is
dissociation into N H 3 and H C 1 0 , the determination o f the influence o f
4

these products on the rate o f decomposition is o f crucial importance. Such


studies were conducted repeatedly [13.3, 13.13-13.16, 13.18]. It has been
revealed that introduction o f excessive H C 1 0 4 always results in a sharp
Thermal Decomposition and Combustion 197

decrease o f the induction period and the growth o f the initial rate. The rate
increase is more pronounced upon introduction o f the acid directly into the
crystal during its crystallization than in the case when the acid is introduced
into the volume o f the reactive vessel. The exposure o f crystals in N H 3

vapors or their precipitation from aqueous ammonia solutions increases the


induction period. In [13.16], the influence o f N H in a closed vessel was
3

studied in detail. It was revealed that excess o f N H 3 considerably increased


the induction period and virtually did not change the rate o f the process.
The influence manifests itself considerably more strongly after the phase
transition into the cubic lattice. In order to reduce the rate o f decomposition
below the sensitivity o f the technique employed, in an orthorhombic lattice
it is necessary to have a pressure o f approximately 0.26 atm, and in a cubic
lattice, it is enough to have N H pressure o f 6 • 10~ atm.
3
4

If in the presence o f excessive N H the decomposition proceeds in the


3

orthorhombic modification, the temperature growth, accompanied by the


phase transition into the cubic modification, results in the immediate halt o f
the reaction. A t the reverse cooling below 2 4 0 ° C , the reaction starts again
only after a prolonged induction period, which gives the impression o f
differences in the mechanism o f decomposition o f these phases o f A P .
These differences are explained by the fact that the decomposition occurs in
the depth o f the crystal on the developing reactive sites, on which the
composition o f products is attained very rapidly, including the amounts o f
H C 1 0 and N H . This composition corresponds to the stoichiometry o f the
4 3

decomposition on this site and is independent o f the composition o f the


environment.
The flow o f the gas-phase products from the reactive site towards the
surface o f the crystal prevents the diffusion o f the excess o f base or acid
towards this site. The uniform distribution o f the components through the
crystal volume is absent here, and the products o f decomposition remove
from the site an excessive N H or H C 1 0
3 4 introduced there earlier. Thus,
their influence exhibits itself only in the induction period, when the gas
release is very low. When the lattice is rearranged during the phase
transition, NH 3 penetrates into the depth and its influence becomes
considerably stronger. Gradually, the flow o f the products removes an
excessive N H , and the rate attains the ordinary values.
3

Thus, during the thermal decomposition o f A P , the assumption of


quasi-stationary and constant concentrations o f products o f decomposition
throughout the crystal bulk cannot be employed. The process is localized at
separate centers and the conditions near them differ considerably from the
conditions in the bulk o f the crystal. The inhibition o f decomposition at the
attainment o f the equilibrium pressure o f water vapor, which is capable o f
forming the liquid phase, is a more reliable fact, which confirms the
determining role o f the reactions o f perchloric acid at all stages o f
decomposition o f A P . Only the reactions with participation o f molecules o f
perchloric acid can be stopped so abruptly during water condensation due to
198 Ammonium Perchlorate

a strong ionization o f these molecules. In principle, such regularities can


also be obtained under the assumption o f the leading role o f decomposition
o f nitronium perchlorate N 0 C 1 0 2 4 in this reaction, postulated in [13.17].
The formation o f such molecules should proceed during the oxidation o f the
NH 4
+
ion to N 0 2
+
, but, actually, it is difficult to expect an essential
importance o f this pathway in the decomposition o f A P . M a i n l y , oxidation
of N H 4
+
occurs up to N and N 0 , and the formation o f N 0
2 2 2
+
as the main
product is hardly possible. A t the same time, the reactions with participation
o f perchloric acid can describe this process.
A s shown in Chapter 11, the decomposition o f perchloric acid is a
complex chain process. For the development o f this process, the
intermediate species such as C 1 0 , C 1 0 , C 1 0 , and CI are o f considerable
2 7 3 4

importance. The self-acceleration is observed only for temperatures below


120°C. A t higher temperatures, the quasi-stationary concentrations o f the
intermediate species are attained very rapidly. Their passage into the A P
crystal makes this process even more complicated. A l l intermediate
products and H C 1 0 4 molecules are able to oxidize the N H 4
+
ion. In the
course o f the process, the regeneration o f the acid molecule proceeds by the
proton transfer on the adjacent C 1 0 " ion. The numerous estimates o f the 4

influence o f the N H 4
+
ion on the rate o f thermal transformations of
perchloric acid [13.14, 13.15, and 13.19] demonstrate the increase o f the
rate by at least a factor o f 1 0 - 1 0 . A s it was demonstrated by direct
2 3

measurements, the oxidation-reduction processes with participation o f


either H C 1 0 4 molecules or the products o f its equilibrium transformations
and thermal decomposition, for instance, C10 2 molecules, are of
considerable importance. Nevertheless, reliable quantitative measurements
either o f the concentration o f an excessive acid at the reactive site o f a
decomposing A P or the rate constants o f its transformations are not
available. Thus, it is impossible to develop the calculation scheme o f the
rate o f decomposition o f A P ; one can speak only about the qualitative
tendencies. In particular, the observable growth o f the rate o f decomposition
of A P cannot be explained by the kinetic characteristic features o f thermal
decomposition o f perchloric acid or by its accumulation in the course o f
decomposition. A t temperatures above 2 0 0 ° C , in an open system, these
causes cannot provide a noticeable increase o f the rate.

13.4 Topochemical Peculiarities of Thermal


Decomposition
It is most probable that the acceleration o f decomposition o f A P is
connected with the topochemical peculiarities o f this reaction. The
regularities and topographic peculiarities o f the initiation and development
o f the reactive center in an A P crystal were studied in detail in [ 1 3 . 2 0 -
13.23]. One o f the most comprehensive studies was carried out together
Thermal Decomposition and Combustion 199

with Raevskii [13.20-13.22] using specially grown monocrystals of


rhombic modification and zero-defect filamentary crystals o f A P . The
decomposition was carried out under a microscope, on a special heating
table. This technique made it possible to study isothermal decomposition at
temperatures up to 4 0 0 ° C .
In the case o f the rhombic modification o f A P , during the induction
period, the crystal surface remained without any visible changes or was
weakly etched at the outlets o f dislocations due to sublimation. Under the
outlets o f dislocations at a depth more than 5 mm, separate spherical semi-
transparent formations appeared whose size was about 1 jum by the end o f
the induction period. After that, their growth stopped. Such formations were
called "nuclei". During the induction period, the nuclei are immovable and
form clusters in the crystal bulk, which are stretched along the (010) -axis o f
the crystal. Due to the formation o f a large number o f nuclei at the end o f
the induction period, the crystal dimmed. During this process, the mass loss
did not exceed 0.1%.
The reaction acceleration started when, at separate places in the crystal
with the maximal density o f nuclei, their motion started together with the
further increase o f their density. Gradually, an ellipsoidal reactive center
was created, which was stretched along the (010)-axis o f the crystal. The
new traveling centers appeared through its peripheries. Their displacement
was the superposition o f two motions: transitional motion from the reactive
center and chaotic motion o f the Brownian type. Consequently, the nuclei
moved away from the center o f decomposition, their size increased up to 2
|um, and the larger nuclei moved slowly. A s the number and size o f the
nuclei increased, they merged with the formation o f an immovable core o f
the center o f decomposition. O n the whole, the reactive center represented
itself as a cloud o f traveling and breeding nuclei, covering a porous
immovable bulk. A t constant temperature, in the crystal, the rate o f growth
o f the centers was constant and varied from one crystal to another. The
activation energy was as large as 30 kcal/mol for longitudinal growth and
32 kcal/mol for transverse growth. The width o f the chemical reaction zone
increased with temperature, and the activation energy o f this process was
approximately 10 kcal/mol. After penetration up to the crystal face, the
center grew in other directions without variations in the rate. The removal
o f the products proceeded through micro cracks and micro defects, but the
new nuclei did not appear along the channels removing the gas-phase
products. A micro crack did not allow the center to grow further, and it
developed only due to transverse growth, which resulted in a decrease o f
the reaction rate for a small number o f centers in the crystal. Separate
nuclei, especially at the initial stages o f acceleration, inside the non-
destructed crystal, did not stop growing near the size o f 2 urn, but grew
further up to the point o f j o i n i n g with the zone, where the outflow o f the
products was facilitated.
200 Ammonium Perchlorate

Therefore, the increase o f the rate o f decomposition is determined both


by the increase o f the number o f nuclei and by the growth in their size. The
rate started to decrease when the majority o f centers penetrated up to the
crystal faces and they fused together, resulting in the reduction o f the total
reaction zone. After several centers had fused, a nearly plane reaction zone
was created with the breeding nuclei, which penetrated through the crystal,
leaving behind a porous residue with a block size o f 2-5 jam, w h i c h
preserved the shape o f the parent crystal.
A complex mechanism o f the reaction center development was
explained based on the studies o f the behavior o f dislocations during A P
decomposition. The parent crystals contain a large number o f dislocations
which, being heated up to the temperature o f decomposition, lead to
polygonization with the formation o f a three-dimensional network of
dislocations. O n dislocations and other defects, both the salt dissociation
into N H 3 and H C 1 0 4 and the subsequent oxidation-reduction stages are
essentially facilitated in comparison with a perfect lattice due to the excess
o f the isobaric-isothermal potential on the defect. The points o f intersection
o f dislocations are especially active. The gas-phase products, evolved on the
stationary dislocations, lead to the formation o f nuclei. The halt o f nucleus
growth is determined by the achievement in it o f the pressure o f water
vapors, w h i c h is sufficient for the formation o f the liquid phase similar to
those observed in studies o f the kinetics in a closed volume. In this case, at
the places with maximal dislocation density, the gradients o f mechanical
stresses appear, which are sufficient for the breeding o f dislocations. This
leads to the formation o f the new nuclei and to the appearance o f the
development o f the reactive center. The models o f the structure o f a
developed reactive center and the scheme o f distribution o f mechanical
stresses are presented in F i g . 13.4.1. Here, (1) is the region o f the residual o f
A P decomposition, (2) is the reaction zone, where the accumulation o f the
products creates mechanical stresses sufficient for breeding o f dislocations
in region (3). For regions (2) and (3), it is impossible to determine a clear
boundary and this division is arbitrary. The distribution o f mechanical
stresses is presented for the reaction center, which is connected with the
atmosphere. If the center is located deep inside the crystal bulk, in regions
(1) and (2), the stresses are close to each other. In all cases, the stresses,
w h i c h are generated by the reaction products in region (2), play the main
role. Their unloading towards region (1) occurs due to the sample
destruction and, towards region (3), due to the generation of new
dislocations.

It was shown, by direct experiments on sublimation etching o f the


crystals o f A P during their thermal decomposition, that in the region o f the
center o f decomposition, the density o f dislocations is increased. The hole
o f etching corresponds to each nucleus near the surface. There are some
holes o f etching without nuclei. The nuclei on this dislocation have not been
formed yet. The displacement o f certain nuclei from their initial holes o f
Thermal Decomposition and Combustion 201

etching, which were left by dislocations after the beginning o f etching, was
observed. This confirms the motion o f dislocations and their breeding in the
region o f the center o f decomposition. The only reason for the appearance
o f mechanical stresses in the crystal o f A P is the accumulation o f the gas-
phase products in the nuclei. The pressure inside the nucleus was estimated
from the increase o f the gas volume in the nucleus during the dissolution o f
the crystal. The eutectic mixture N H C 1 0 - L i C 1 0
4 4 4 was used as a solvent.
A t a temperature o f 2 2 5 ° C , the total pressure o f approximately 20 atm was
obtained, which is in good agreement with the data on the equilibrium
pressure o f water vapors over the saturated solution o f A P (Table 13.1.1)
with regard to the water content in the products o f decomposition, which is
approximately 50%.

Fig. 13.4.1. A schematic model of the center of decomposition (a) and distribution
of the mechanical stresses in it (/?): (7) the region of a porous core after the
coalescence of the nuclei, (2) the reaction zone, (3) the region of multiplication of
dislocations.

In studies o f the mechanical properties o f A P crystals by stretching, it


was revealed that a noticeable flow o f the crystal starts at pressures near 10
k g / c m along the crystal (o 10) -axis and approximately 50 k g / c m along the
2 2

other axes, w h i c h is obvious evidence o f the leading role o f mechanical


stresses in the development o f the reactive center. In addition, it was shown
that the creation o f a high density o f defects by radiation treatment o f a part
o f the crystal with the help o f the X-rays 0.02 cm in diameter or by pricking
of the crystal by glass or tungsten needles resulted in the development o f the
reactive center at this place. In this case, the initial shape o f the center
coincided with the shape o f damage. O n the other side, at the thermal
decomposition o f the filamentary crystals o f A P , which are nearly free from
dislocations, the self-acceleration o f the reaction is also almost absent.
O n transition into the cubic lattice, the essential changes in the
character o f initiation and development o f decomposition also occurred.
202 Ammonium Perchlorate

These changes confirm the mechanism o f the decomposition kinetics


developed above. The cubic lattice o f A P , in contrast to the rhombic one, is
not optically active. Hence, the phase transition process and the quality o f
the crystal obtained could be easily controlled in a polarized light. It was
shown by sublimation etching that the regions, which had an elevated
dislocation density in the rhombic modification, preserved their defects or
increased their number after the phase transition. W i t h i n a temperature
range o f 2 4 0 - 2 5 0 ° C , the phase transition proceeded inertly and the
conversion zone propagated through the crystal non-uniformly, forming the
regions o f increased patchiness, and even preserving the regions with the
rhombic modification, included into the matrix o f the cubic lattice. A b o v e a
temperature o f 2 5 0 ° C , the crystal did not contain the rhombic phase. D u r i n g
the induction period, the process is similar to the decomposition in the
rhombic modification. Immovable nuclei approximately 1 um in size are
formed in the crystal body. Recall that the temperature dependence o f the
initial rate after the phase transition is also preserved unchanged.
The reaction acceleration was accompanied by the appearance o f
centers o f decomposition o f two types, which had a different development
mechanism. In those regions o f the crystal where the dislocation density is
higher, reactive centers developed, which merged with each other leaving a
porous residual, i.e., they developed similarly to the reaction in the rhombic
modification but had a spherical shape in accordance with the properties o f
the cubic lattice. A l l o f the peculiarities o f these centers were indicative o f
the dislocation nature o f their development. In those regions o f crystal
where the initial dislocation density is low, the separate nuclei, w h i c h
appeared during the induction period, continued to grow, first at a nearly
constant rate and then with considerable acceleration. The center growth
proceeded until the center started to connect with the atmosphere. A t this
moment, its development stopped. Inside well-grown crystals with a l o w
dislocation density, such centers attained a considerable size, w h i c h was
determined by the size and strength o f the parent crystal. A t a high pressure,
the gas-phase products were contained inside the center, and the interfaces
were saturated by a liquid solution o f A P in the products, mainly, in water.
A n y solid residual is not formed in these centers, and the parent A P is
completely decomposed. The accelerated growth o f these centers is
determined mainly by the increase o f the amount o f the liquid phase on their
surface. A t high temperatures, such centers are o f crucial importance for
decomposition. O n mechanical action on the crystals o f A P o f the cubic
modification, the formation o f dislocation centers increases considerably
and the initiation and growth o f the centers o f complete decomposition vary
only slightly.
W i t h i n a temperature range o f 2 4 0 - 2 5 0 ° C , due to inert and incomplete
phase transition, specific conditions were created for the reaction: increased
patchiness and mechanical stresses, created by the residual o f the rhombic
modification. This circumstance stimulated the initiation o f the dislocation
Thermal Decomposition and Combustion 203

centers both in the cubic and rhombic lattice. In the latter, they had an
ellipsoidal shape, as at temperatures below 2 4 0 ° C . Nevertheless, the growth
o f such centers was hampered due to the difficulties o f the motion o f
dislocations in the reaction zone in the case o f increased patchiness. There
were very few centers of complete decomposition. They attained
insignificant size, and their growth rapidly halted due to contact with the
atmosphere through the mosaic structure. The contribution o f these centers
to the total rate was very small. Finally, within this temperature range, the
observable rate is noticeably smaller than at lower temperatures. Formally,
this is manifested in a smaller value o f the rate constant k . 2

The approximate analysis o f the development o f the reactive center


through crystal deformation and dislocation breeding for the plane front
propagation was considered in [13.24] and has already been presented in
Chapter 2. The results obtained are in good agreement with the kinetic
regularities o f decomposition.
T o summarize the microscopic studies o f thermal decomposition o f A P ,
we should note that processes o f three different types, proceeding
simultaneously, were revealed:

(1) Processes on stationary dislocations, which lead to the origination o f


centers,

(2) Processes leading to the growth o f centers o f complete decomposition,


especially at elevated temperatures,

(3) Processes determining the development o f the centers, consisting o f fine


nuclei and forming a porous residual o f A P .

Processes o f the first two types can be considered to be traditional for


topochemical reactions. This is a characteristic pathway for initiation and
development o f the reactive centers in numerous reactions o f thermal
decomposition o f the solid phase. In this case, there is a pronounced
boundary between the parent substance and the reaction products.
Apparently, the formation o f a reactive center, which represents a cloud o f
nuclei traveling around the central core, is a specific feature o f A P . This is
caused by the absence o f solid products o f decomposition, which usually
represent the centers o f growth o f the new phase and thus facilitate the
formation o f a pronounced boundary o f the reactive core, and also by the
strength properties o f the crystal. A plastic flow, caused by the breeding o f
dislocations, arises under relatively light loads.
Thus, the gradients o f mechanical stresses, created by the evolving o f
the gas-phase products in the nucleus, turn out to be sufficient for the
motion and breeding o f dislocations. The dislocation network by itself
facilitates the diffusion o f the products from the nuclei, and, in doing so,
prevents their development into centers o f complete decomposition.
204 Ammonium Perchlorate

13.5 Influence of the Preliminary Irradiation and


Additives
A great number o f works on the influence o f various additives on thermal
decomposition o f A P have been published. Here, we w i l l briefly consider
the results, w h i c h are o f principal importance for the treatment o f the
mechanism o f this reaction or w h i c h can be employed to control the rates o f
thermal decomposition and combustion o f AP-based compositions. The
influence o f excessive acids and bases has already been discussed above. A
practical inaccessibility o f the reactive centers o f decomposition for the
influence o f the environmental gas medium and a high volatility o f
ammonia and perchloric acid makes it difficult to control the rate by
introduction o f such products.
The following two ways of changing o f the rate of thermal
decomposition o f A P are used most widely:
(1) Variation o f the properties o f the crystal lattice by its preliminary
irradiation, mechanical treatment, introduction of isomorphically co-
crystallized additives or additives which, at elevated temperatures, are able
to form liquid eutectic mixtures with A P and, in doing so, to dissolve the
defect places o f a crystal.
(2) Heterogeneous introduction o f additives, which are able to influencg the
rate o f oxidation-reduction stages o f thermal transformations o f perchloric
acid. M a i n l y , this is the compound containing the ions o f the metals o f
alternating valence.
The influence o f the variation o f crystal defects was studied with the
help o f a preliminary irradiation o f A P . Irradiation made it possible to
change the rate o f decomposition without changing chemical composition,
and to study in greater detail the influence o f the rate o f decomposition o f
A P on the rate o f decomposition o f the composites on its basis and on their
combustion rates. The irradiation was carried out by X - r a y s (the absorbed
radiation dose varied from 0.001 to 0.1 mrad) and was conducted on a linear
electron accelerator with energy o f 1.6 M e V (the absorbed radiation dose
varied from 0.075 to 8.3 mrad). The radiation dose was controlled by the
exposure time. After the radiation treatment, up to the beginning o f
decomposition, the samples were stored at room temperature for periods
from 1 h to 1-2 days. The rate o f decomposition was independent o f the
storage time within this range, although, at the beginning o f the experiment,
immediately after the radiation treatment, a higher and poorly reproducible
rate was obtained.
A preliminary radiation treatment causes, mainly, growth o f the initial
rate with a corresponding reduction o f the induction period. O n l y for high
doses does the degree o f self-acceleration start to increase.
W i t h i n the limits o f 0.001-3 mrad, the dependence o f the rate constant
o f decomposition k\ on the dose D takes the form:
Thermal Decomposition and Combustion 205

k - 1.7-10 s
x
6 M
+ 1.15-10- D 4 048
s"
1
(200 C ) ,

k - 11.4- 10" s" + 2 . 5 - 1 0 ~ D V


x
6 ! 4 0 4
(230 C ) .

For higher doses, the growth o f k\ is stopped. For doses up to 0.1 mrad, the
rate constant k is almost constant, and, within the range o f 0.1-8.0 mrad, it
2

increases by a factor o f 2 - 2 . 5 . These results confirm the mechanism o f


initiation and development o f thermal decomposition in a crystal lattice o f
A P developed above. Irradiation increases the number o f defects in a crystal
and, thus, facilitates the development o f nuclei. For higher doses, crystal
deformation is facilitated which, in turn, leads to the acceleration o f the
development o f reactive centers.
After preliminary radiation treatment, the rate o f thermal decomposition
can be increased not only due to the growth o f the number o f defects and a
partial decomposition during irradiation, but also due to the accumulation o f
free radicals and other decomposition products, which, taking into account
the chain radical character o f the decomposition o f perchloric acid and A P ,
also can contribute to the increase o f the initial rate. A s shown by the E P R
method, the paramagnetic molecules of C10 3 and N H 3
+
radicals are
accumulated inside the irradiated A P . In studies o f N H 3
+
recombination at
elevated temperatures, the phenomenon o f a go-ahead rapid free valence
transfer through the crystal lattice o f A P was observed.
In [13.25, 13.26], the influence o f isomorphic ions C 1 0 3
_
on the thermal
decomposition o f A P was studied. These ions are almost always present in
the crystals in the form o f technological impurities, and only after a fivefold
recrystallization does their content become lower than 5 • 10~ m o l % . Since 5

HC10 3 is a weak and less stable acid than H C 1 0 , the proton transfer into 4

C10 ~ proceeds more easily and the rate o f decomposition o f the molecules
3

formed is higher. During the decomposition o f H C 1 0 , chlorine oxides are 3

formed, through which the chain branching proceeds at HC10 4

decomposition. Thus, the presence o f C10 ~ ions leads to a reduction o f the


3

induction period and an increase o f the rate o f nucleus formation. In the


amount o f 2 • 10~ mol % at 2 3 0 ° C , these ions already increase the rate o f
2

formation o f the reactive centers by a factor o f 10, reducing the induction


period from 40 to 12 min. The rate o f growth o f the reactive centers and the
formal kinetic regularities o f the developed reaction almost do not change.
After a preliminary radiation treatment, C10 ~ ions are accumulated in the 3

crystal. The set o f data obtained made it possible to assume that in the
majority o f A P crystals studied the rate o f decomposition was determined
by the decomposition o f the impurity ammonium chlorate. The question
about the rate o f decomposition o f A P for the complete absence o f C 1 0 ~ 3

ions remains open. Such samples have not been studied. A t the same time, it
is obvious that the isomorphic impurity o f C 1 0 " can be considered as a 3

kind o f defect o f the crystal lattice, and the kinetic regularities presented
above are valid.
206 Ammonium Perchlorate

Introduction o f multiple charged cations into the crystal lattice o f A P ,


for instance B a + 2
[13.27], also changes the rate o f decomposition. But in
this case, the dependence o f the rate on the content o f doped cations is o f a
complex character; it passes through the m i n i m u m and m a x i m u m and is
determined by the action o f two factors: deformation o f the crystal lattice
and creation o f the cation vacancies during doping.
The kinetic regularities o f thermal decomposition of isomorphic
crystals N H C 1 0 — K C 1 0
4 4 4 were also studied. Since the ionic radii o f K +

and N H 4
+
are very close, such crystals could be obtained by joint
crystallization within the whole range o f the K content from 0 to 100%. A t +

all temperatures studied (up to 380°C), only A P decomposed, and


potassium perchlorate remained unchanged. The kinetic curves obtained are
similar to those for A P and are described by equations (13.1.1). U p to the
phase transition, the values o f k and k almost do not change. D u r i n g the
x 2

first stage, the degree o f conversion o f A P decreases, as well as in a pure


salt, attaining at the temperature o f 3 1 0 ° C the value a c = 0.06-0.08. In
composite crystals, the point o f phase transition shifts towards the region o f
higher temperatures as the content o f K increases. Thus, for a pure K C 1 0 , +
4

the temperature T pX = 2 9 8 ° C shifts, correspondingly, in the region o f


decrease o f the constant k , which is clearly seen in F i g . 13.5.1. For all
2

samples, with regard to the data scattering, the value o f k is independent o f x

the content o f K +
and o f the lattice type. Introduction o f K +
virtually does
not deform the crystal.
D u r i n g the second stage o f A P decomposition, where the rate is
determined by the development o f the centers o f complete decomposition in
the liquid phase, K +
ions exert a noticeable accelerating influence. In
crystals with 80%o o f K C 1 0 , the value o f k is higher by a factor o f 100 than
4 3

in a pure A P . Apparently, such influence is determined by the transfer o f K +

ions into the liquid phase, where they can increase the solubility o f A P , bind
the water in their hydrate shells and, thus, decrease the ionization o f
perchloric acid. A l s o , they can participate directly in a complex process o f
decomposition o f perchloric acid.
O n the whole, the transfer o f the reaction in the liquid phase makes it
possible to control the rate o f thermal decomposition more effectively. This
was demonstrated in studies o f the influence o f additives o f lithium and
guanidinium perchlorates, and lithium fluoride. O n addition o f 3 0 % o f
LiC10 4 into A P at the temperature o f 2 1 0 - 2 1 5 ° C , a liquid eutectic mixture
is formed, and, at a further temperature increase, a rapid and complete
dissolving o f the crystal proceeds. In the liquid phase, decomposition starts
and runs rapidly after a relatively long induction period. O n addition o f 2 -
3% o f L i C 1 0 , the liquid phase is formed only on the crystal surface, and its
4

major portion remains undissolved. Inside the crystal, the decomposition is


developed and inhibited after decomposition o f 3 0 % o f the parent substance
in absolutely the same way as without L i C 1 0 . After a long induction 4

period, the reaction in the liquid phase starts and proceeds till the end with
Thermal Decomposition and Combustion 207

the complete dissolving o f the residual o f the solid phase. A t temperatures


above 2 4 0 ° C , a liquid also wets the deep crystal layers. The phase transition
facilitates the penetration o f the liquid deep inside the crystal. Due to the
dissolving o f the sites, most enriched by defects, the reactive centers are
developed only at the sites not wetted by the liquid. Hence, the reaction in
the solid phase proceeds at a very small depth and is stopped. After the
induction period, the reaction in the liquid phase starts and proceeds up to
the end, dissolving the solid phase. The reaction is o f the zero order because
the consumption o f A P in the liquid phase is compensated by dissolving the
new portions o f solid residual. The form o f the kinetic curve is shown in
F i g . 13.5.2.

1.80 f.85 t.90 f.95 2.00 t/T-W 3


ffffis

Fig. 13.5.1 Fig. 13.5.2

Fig. 13.5.1. Dependence of l o g ^ ) and \og(k ) on \/T for isomorphic crystals of


2

N H 4 C I O 4 — K C I O 4 for different [KC10 ], %: (/) 0, (2) 4.5, (3) 11.8, (4) 28.2, (5)
4

55.8, (6) 80.0.

Fig. 13.5.2. A kinetic curve of thermal decomposition of A P in the solution of


L i C 1 0 (250°C).
4

When a reaction vessel made from stainless steel is used or upon


introduction o f stainless steel powder into the system, the reaction in the
liquid phase starts immediately after its formation without the induction
period. In a l l cases, finally, the residual L i C 1 0 4 remains, whose thermal
decomposition proceeds at a noticeable rate only above a temperature o f
400°C. The reaction develops in the same way in the presence o f
guanidinium perchlorate C N H C 1 0 , which in the pure form decomposes at
3 6 4

a noticeable rate above 3 0 0 ° C , melts at 2 4 8 ° C , and its eutectic mixture with


AP melts at 2 1 7 ° C [13.28]. The only difference is that guanidinium
perchlorate is partially involved in the process o f thermal decomposition,
being oxidized by the products o f decomposition o f A P . Its content in the
system decreases in the course o f the reaction. The amount o f the liquid
208 Ammonium Perchlorate

phase decreases and the reaction can be described by the kinetic equation o f
the first order. The temperature dependence o f the rate constant takes the
form:

k = \0 1 2 3
exp(-41000/RT),s~ . l

These results make it possible to treat the influence o f the additive o f L i F on


the kinetics o f thermal decomposition o f A P . The additive is commonly
used as a combustion inhibitor and stabilizer o f the compositions based on
A P . A t a temperature o f 2 1 5 ° C and also at higher temperatures, this
compound enters into the exchange interaction with A P :

11
NH4CIO4 + L i F ^ LiC10 + NH F
4 4

N H + HF
3

Similar to additives o f L i C 1 0 , its influence up to the phase transition is


4

insignificant; only the degree o f conversion increases slightly during the


first stage. However, above a temperature o f 2 4 0 ° C , the reaction is abruptly
inhibited, and the induction periods up to a rapid self-acceleration turn out
to be considerably higher than upon introduction o f only L i C 1 0 . The 4

reaction proceeds up to the complete decomposition o f A P . This influence


is connected with the formation o f the eutectic N H C 1 0 — L i C 1 0 , and the 4 4

presence o f F~ ions in it results in the additional stabilization o f thermal


decomposition. This stabilization can be explained by the interaction o f F "
ions with the radicals, determining the branching of the chain
decomposition o f H C 1 0 ; for instance, the reaction:
4

F"+ CI = CI" + F

occurs in the gas phase accompanied by an energy release o f 4 kcal/mol. F


atoms react with C 1 0 and C 1 0 radicals with the formation o f chloryl
2 3

fluoride C 1 0 F and perchloryl fluoride C 1 0 F , w h i c h are not able to


2 3

participate in chain branching radical reactions.


The inhibitory influence o f fluoride on the thermal decomposition o f
C 1 0 with the formation o f these compounds is described in [13.29]. This
2 7

is confirmed by the kinetic curves obtained at the decomposition o f the


solution o f A P in L i C 1 0 4 (50 : 50) with various amounts o f L i F added at
3 0 0 ° C , w h i c h are shown in F i g . 13.5.3. The induction period increases
proportionally to the content o f F~ ions. After the beginning o f a rapid self-
acceleration, the rate does not depend on the content o f these ions.
Apparently, a sharp self-acceleration o f the decomposition is determined by
the release o f H F into the gas phase. When the concentration o f F~ drops
Thermal Decomposition and Combustion 209

below the limiting value, the chain reaction of perchloric acid


decomposition develops to oxidize N H 4
+
ions. Thus, a very strong inhibition
o f A P decomposition upon introduction o f L i F above the point o f the phase
transition is connected with the dissolving o f the defect places o f the crystal
in the liquid melt N H C 1 0 — L i C 1 0 and with considerable inhibition o f the
4 4 4

reaction o f decomposition in the liquid phase in the presence o f F~ ions.


The influence o f compounds involving ions o f metals o f alternating
valences was studied with the help o f compounds o f iron, copper,
manganese, yttrium, lanthanum, nickel, chromium, cobalt, and cadmium,
and also compounds involving simultaneously two such elements, for
instance, copper chromite [13.17, 13.30-13.32]. The kinetic peculiarities o f
the influence o f the compounds o f these elements are considered in detail in
[13.4]. These ions are able to catalyze a great number o f thermal
transformations and oxidizing processes, including those with participation
o f perchloric acid. A l l o f these compounds essentially increase the rate o f
A P decomposition. They are frequently recommended and employed as
promoters o f combustion o f AP-based mixtures.
Additives o f metals and their alloys usually exert a similar influence on
the rate o f thermal decomposition because, as a rule, the metal surface is
covered by a layer o f oxide. The quantitative parameters o f this influence
strongly depend on the properties o f a catalyst and the procedure o f its
production, first on its dispersion, and also, on the method by w h i c h the
catalyst is introduced.

to

0.5

Fig. 13.5.3. The kinetic curves of decomposition of the mixture N H C 1 0 — L i C 1 04 4 4

(1 : 1) in the presence of LiF for various [LiF], %: (7) 0, (2) 0.25, (3) 0.5, (4) 1.25,
(5) 2.5,(60 5.0.

In particular, the isomorphic introduction o f K M n 0 4 into the crystal o f A P


gives a more pronounced catalytic effect than the introduction o f this salt as
a mechanical impurity. The initial rate increases by a factor o f 1.5-3, and
210 Ammonium Perchlorate

the degree o f self-acceleration increases by a factor o f 10-20 on retention o f


the kinetic law and activation energies o f both stages.
A l s o , a considerable advantage gives a preliminary coating o f a catalyst
on the surface o f A P . O n the whole, a higher catalytic effect provides the
use o f F e 0 and C u C r 0 . This effect usually increases with the temperature
2 3 4

increase due to the growth o f the degree o f dissociation o f A P . Organometal


compounds, in particular, derivatives o f ferrocene, are close to these
substances by their character o f action and the catalytic effect. Together
with the catalytic influence o f oxides, formed at the destruction o f these
compounds, the organic part o f the molecule is oxidized, which increases
the heat release in the condensed phase. The influence o f ferrocene
additives is considered in detail in the discussion o f thermal decomposition
o f A P mixtures with organic binders.

13.6 Thermal Decomposition of Composite AP-


Based Systems

In this section, we will consider in a generalized form thermal


decomposition o f the mixtures o f A P with different organic substances, first
o f a l l , with polymers, w h i c h represent the models o f AP-based rocket
propellants. Thermal decomposition o f such mixtures is o f considerable
importance both for the estimation o f thermal decomposition o f these
compositions and admissible shelf storage life and for understanding the
mechanism and regularities o f their combustion.
A m o n g compositions studied, the following three main groups can be
distinguished:
(1) Mixtures o f A P with organic compounds, whose rates o f thermal
decomposition are lower than that for A P . Thus, the process starts from
thermal decomposition o f A P . Mixtures o f A P with hydrocarbon polymers,
for instance, with polystyrene or polybutadiene [13.11-13.13] and so on,
can be attributed to this group. Nevertheless, in many cases, only the
differential thermal analysis ( D T A ) was used, and the quantitative kinetic
data are not available. In [13.11-13.13], it was shown that under relatively
low temperatures the kinetic regularities o f thermal decomposition in such
mixtures are very close to the regularities o f the decomposition o f a pure A P
under similar conditions. Products o f A P decomposition possess a rather
high oxidative activity and enter the secondary reactions o f interaction with
a polymer that results in an essential increase o f the heat release o f the
process and in the appearance o f C O and C 0 2 in the products of
decomposition. Thus, for the mixture o f A P with polystyrene at the ratio o f
4 : 1, the heat o f decomposition is 640 cal/g in contrast to 240 cal/g for a
pure A P under similar conditions. During thermal decomposition o f model
Thermal Decomposition and Combustion 211

mixtures o f A P with polyester oligomers, we revealed the decrease o f the


rate o f decomposition within a temperature range o f 2 4 0 - 2 5 0 ° C .
A s in the case o f a pure A P , when a preliminarily irradiated A P is used,
the rate o f decomposition increases, and the kinetic regularities of
decomposition and the rate constants for A P and its mixture with polyester
are also close to each other. This unambiguously confirms the determining
role o f A P in the decomposition o f such mixtures. Since the thermal
stability o f A P is high, such mixtures also possess a high thermal stability
and can be processed and employed within a rather wide temperature range.
Studies o f decomposition o f these mixtures have revealed that with a
temperature increase above 3 5 0 - 4 0 0 ° C the temperature dependence o f the
rate o f decomposition becomes considerably stronger, i.e., the activation
energy o f the process rapidly increases. If at low temperatures the value o f
£ a c t falls within the range 2 7 - 3 2 kcal/mol as for a pure A P , then at higher
temperatures the value o f E act rapidly increases up to 6 0 - 7 0 kcal/mol. For a
mixture o f A P with polybutadiene, the following dependence for & eff was
obtained, which describes the transition to higher values o f £ a c t :

jfc cff = io 7 5 8
^ ^ e x p ( - 2 7 0 0 0 / 7 ? r ) + 10 2 1 3 4
exp(-69600/7?r), s' ,
1

P =4.78-10 exp(-28800/7?r),
s
1 2
Pa,

where P 0 is the external pressure, P is the pressure o f the products o f A P


c

sublimation (at low temperatures, P « c P ). This transition is determined


0

by the growth o f the contribution o f the reaction o f polymer oxidation by


the gas-phase perchloric acid, formed during A P decomposition. During
decomposition o f a pure A P in the gas, the equilibrium pressures o f
ammonia and perchloric acid are attained, but the rate o f decomposition o f
HC10 4 in the gas phase is low. In the presence o f organic substances, with
temperature increase, the contribution o f the diffusion o f the gas-phase
HCIO4 into the layer o f this substance with its rapid oxidation increases.
The consumption o f perchloric acid is compensated by evaporation o f the
new portions o f A P . A t a sufficiently high pressure o f the products o f
dissociation, this pathway becomes the main route. In [13.12], it was shown
that the mixtures o f A P with polystyrene, polyacrylonitrile, and polyvinyl
chloride possess kinetic regularities, which are close to those presented
above. With temperature increase, the activation energy increases up to 6 0 -
63 kcal/mol. O n combustion o f the mixtures o f A P with stable organic
substances, this stage w i l l very likely determine the combustion regularities
to a large extent.
(2) Mixtures o f A P with substances whose rates o f decomposition are
higher than those for A P , but an essential interaction between the
components does not proceed. Mixtures o f A P with numerous nitro- and
212 Ammonium Perchlorate

fluoronitro-compounds, including polymer ones, are also among these. The


kinetic regularities o f decomposition o f such mixtures are determined by the
parameters o f thermal decomposition o f the organic component. For
nitrocompounds, these regularities were considered in the corresponding
chapters. The products o f decomposition o f nitrocompounds usually
demonstrate a relatively weak catalytic influence on the decomposition o f
A P . Thus, after completion o f the decomposition o f a nitrocompound with a
somewhat shortened induction period, the decomposition o f A P with self-
acceleration proceeds. So, during the thermal decomposition o f the model
mixture o f A P with phenol-formaldehyde polymers, obtained on the basis o f
ortho-, meta-, and /?<2ra-nitrophenol [13.34], it is revealed that introduction
o f nitrogroups into the polymer results in a sharp increase o f the rate o f
decomposition at the initial stage. In the D T A curves, two consecutive
peaks o f heat release were observed, which are connected with the
decomposition o f nitropolymer and A P . Unfortunately, there are no
quantitative kinetic data in that work.
(3) Mixtures o f A P with substances whose thermal stability considerably
changes in the presence o f A P . A s an example o f such mixtures, we w i l l
consider mixtures o f A P with alcohol nitrates and thiocol. Thermal
decomposition o f nitrates has been considered in Chapter 9. In this case, the
most typical peculiarity is a very strong increase o f the rate o f
decomposition in the presence o f water and acids or bases that is connected
with the development o f the hydrolysis reaction o f the O — N 0 bond. A t
2

the same time, A P can supply such a strong acid as perchloric acid which in
the presence o f a certain amount o f water in the mixture (in the case o f the
complete dehydration o f the parent compounds, water is accumulated as one
o f the products o f nitrate decomposition) rapidly leads to the development
o f hydrolytic processes.

Fig. 13.6.1. The kinetic curves of thermal decomposition of acrylic ester of glycerol
dinitrate (7) and its mixture with A P (2).

The kinetics o f thermal decomposition o f the mixtures o f A P with the


following nitroesters (30 wt %) was studied: nitroglycerin, 1,3-dinitrate-
Thermal Decomposition and Combustion 213

glycerin, acrylic and methacrylic esters o f dinitrate-glycerin, diacrylic and


dimethacrylic esters o f sorbite tetranitrate, and a methacrylic ester o f
nitroisobutyl-glycerin dinitrate. In all these mixtures, the initial rates o f
decomposition are close to the rate o f decomposition o f a corresponding
nitrate, but then they increase very sharply up to the thermal self-ignition.
Comparison o f the kinetic curves o f decomposition o f the acrylic ester o f
glycerin dinitrate and its mixture with A P is shown in F i g . 13.6.1. For all
other mixtures, qualitatively similar kinetic curves were obtained. The
increase o f the acidity o f the system in the presence o f A P can proceed in
two ways: (1) partial dissolving with the subsequent hydrolysis o f A P in the
products o f nitroester decomposition, (2) slight surface oxidation o f A P by
the products o f nitroester decomposition with formation o f an appropriate
amount o f perchloric acid. A rapid development o f hydrolytic processes is
typical, first o f all, for nitrates with the location o f O — N 0 2 groups at the
adjacent carbon atoms. In this case, the rate increases by a factor o f 10 and, 3

in the absence o f such groups, by a factor o f 10. It is important that due to


the accumulation o f acid and water in the course o f the process, the order o f
autocatalysis in such systems is higher than one and, usually, is close to
two. The activation energy o f thermal decomposition o f such systems,
estimated by various methods, falls within the limits o f 2 3 - 2 6 kcal/mol.
This is essentially lower than the activation energy o f decomposition o f
pure nitrates (38^42 kcal/mol), which also confirms an acid-hydrolytic
decomposition o f mixtures o f nitrates with A P .
Since the acceleration of decomposition is connected with the
accumulation o f acid in the system, its binding should lead to a decrease o f
the rate o f the catalytic reaction and, finally, to the elimination o f the
acceleration o f the process. Such an effect cannot be obtained by a
homogeneous introduction o f bases into the system, because the hydrolysis
o f nitroesters under the influence o f bases proceeds at rates comparable with
those for acid hydrolysis. Therefore, the bases should be introduced
heterogeneously, so that the neutralization o f acid would proceed due to its
diffusion towards the surface o f introduced particles. Indeed, introduction o f
10% o f M g O into the mixture A P - n i t r o g l y c e r i n almost eliminates self-
acceleration o f the reaction at 125 and 130°C, as demonstrated by the
kinetic curves in F i g . 13.6.2, but at 135°C, at a certain degree o f
transformation, an abrupt self-acceleration with thermal self-ignition o f the
mixture starts, i.e., a critical transition occurs from the state when
acceleration o f the process is almost absent, to a very rapid reaction with
self-acceleration. Obviously, this is connected with the competition o f the
autocatalytic reaction o f hydrolysis with catalyst diffusion towards the
inhibitor surface. A similar critical transition from the virtual absence o f
acceleration to a rapid self-acceleration proceeds also in other AP-nitroester
mixtures studied when solid particles o f M g O or N a C 0 are introduced.
2 3

In [13.35], the analysis o f the regularities o f the appearance o f critical


phenomena in such systems, without taking into account the consumption o f
214 Ammonium Perchlorate

the main substance, was carried out. The following two cases were
considered:
(1) The elimination o f a catalyst in the open system from the liquid phase
proceeds by its diffusion towards the interface and by evaporation into the
gas. The other components are assumed to be nonvolatile. In the case o f the
reaction o f the second order in the catalyst, the following critical condition
takes place in the system:

S /4r]
x
2
0 >1, where S =Nu l ] DJd) k {\-x\
2 (13.6.1)
JC = (1 + £> 2 d , ID X d K ).
2 g

Here, k\ and k are the rate constants o f a monomolecular stage, D and D


2 x 2

are the diffusion coefficients in the liquid and gas phases, rj = k /k , Nu is 0 x 2 x

the Nusselt number, d is the height o f a liquid layer, d is the distance from
x 2

the surface o f liquid at which the pressure o f a catalyst vapor is negligible,


K is Henry's constant in the temperature dependence o f the catalyst vapor
g

pressure. In this case, the maximal accessible stationary value o f the catalyst
concentration is C s t c r = TJ .0
V2

0 J 10 15

Fig. 13.6.2. The kinetic curves of thermal decomposition of the "AP-nitroglycerin"


mixture in the presence of 10% MgO for different temperatures: (/) 135, (2) 130, (3)
\25°C.

In the case o f sufficiently intensive autocatalysis (770= 1 0 and less), this - 3

concentration w i l l be l o w . W h e n the critical condition is violated, the


reaction w i l l pass into the nonstationary mode with rapid self-acceleration
and corresponding growth o f the catalyst concentration. This growth is
limited b y the value o f the external pressure in the system. The catalyst
vapor pressure cannot exceed the external pressure and at their equality, the
boiling o f the solution starts. Thus, the maximal possible concentration o f
Thermal Decomposition and Combustion 215

the catalyst in the liquid is C = K - P and, i f this value is smaller than m a x g txX

770 , the critical concentration o f the catalyst cannot be attained and the
112

transition to the nonstationary mode w i l l not proceed. Thus, for this system,
the second critical condition is the attainment o f the catalyst concentration
for which the boiling o f the solution starts at the given external pressure.
This condition o f the existence o f a stationary mode takes the form:

Ve^^O . 5
(13.6.2)

If this condition is violated at a critical point the catalyst concentration w i l l


jump from C .cr. to C
st m a x . T o consider the dependence o f the reaction
development on external conditions, let us assume that:

*,=4exp {-EJ RT), k = A Qxp 2 2 {-E I'RT), 2 D,=^ exp 3 (-E /RT)
3

D = A /P ,
2 5 xt K =A zx
g 4 V (E 1 RT).
4

Conditions (13.6.1) and (13.6.2) o f the existence o f the stationary mode w i l l


take the following form:

(B B\d x ^/P e
2
x t )exp[(E 1 + £ -2E 2 4)/RT]^

d\(\+{B d 2 , /P e x t )exp[(£ 3 - £ 4)/RT]f

(B JP )zx [{E -E -E )IRT>\


tx[ V x 2 4

B , A , B =A 2 2 5 IA 3 Ad, 4 2 B -{AJ
2 A )A
2
05
4' .

A s an example, the dependence o f rf I>cr on T at P = const is shown in F i g .


cr

13.6.3. The reaction is nonstationary within the region A B C . The


characteristic features o f the appearance o f critical phenomena in such
systems should be taken into account in the analysis o f the safety o f their
processing and storage. In laboratory studies with small values o f d\ < d ,
]cx

the reaction w i l l be stationary, but in transition to more large-scale charges,


within the region A B C , the reaction w i l l be abruptly self-accelerated up to
the heat self-ignition.
(2) U p o n introduction o f a heterogeneous inhibitor uniformly distributed
through the reaction mixture, the diffusion catalyst towards the inhibitor
particles is added, and the critical condition takes the form:

S 2 = 3Nu a D, /(l - a)(l - 2


4/3
a^)r k ,2
2

where a is the volume fraction o f the inhibitor, r is the size o f inhibitor


particles. Thus, introduction o f a heterogeneous inhibitor can abruptly raise
the total value o f 8 and transfer the reaction from the nonstationary to
216 Ammonium Perchlorate

stationary regime, which was observed in the experiments upon


introduction o f M g O and N a C 0 . The values o f N U J and N u have to be
2 3 2

determined either experimentally or by the solution o f the diffusion


differential equations under the corresponding boundary conditions. The
possibility o f the appearance o f critical phenomena at the thermal
decomposition o f composite AP-based systems is o f special importance for
estimation o f the thermal stability and safety o f the manufacturing process
o f these mixtures. In particular, in the case o f relatively low temperatures
and large-scale charges, the transition to the nonstationary regime is most
probable.
Thermal decomposition o f the model mixtures o f A P (70%) with
thiocol (30%) was studied in less detail. The kinetic regularities obtained
are close to those presented above for the mixtures o f A P with nitroesters.
A n example o f the curves o f the rate o f heat release is shown in F i g . 13.6.4.
Here, also, within a relatively low temperature range o f 1 2 0 - 1 4 5 ° C , after a
sufficiently long induction period, a sharp acceleration o f the process
proceeds. The curves are described by an equation o f the second order with
the autocatalysis o f the first order. The values o f the heats o f decomposition
and the rate constants k , obtained from the heat release and mass loss, are
2

presented in Table 13.6.1.

Fig. 13.6.3 Fig. 13.6.4

Fig. 13.6.3. The temperature dependence of the critical values of d\\ A B C is the
region of the existence of the critical phenomena.

Fig. 13.6.4. The rate curves of heat release in the AP-thiocol mixture for various
temperatures: (7) 145, (2) 140, (3) 135, (4) 130, (5) 125, (6) 120°C.

The calculated values o f k\ are very small and 770 k\/k - 10" . The
=
2
25

removal o f the gas-phase products o f their decomposition by continuous


pumping results in the reduction o f the induction period and growth o f the
m a x i m a l rate.
Thermal Decomposition and Combustion 217

Table 13.6.1. The heats and rate constants of thermal decomposition of the A P -
thiocol model mixture

T,°C a k . 10 , s-
2
3 1
r,°c a 10 , s"
3 1

cal/g cal/g
heat mass heat mass
release loss release loss

120 55 0.87 - 135 60 2.78 2.3


125 57 1.18 140 63 3.85 3.4
130 59 1.93 145 65 5.55 5.8

The kinetic analysis showed that the regularities o f the decomposition o f


this mixture might be described based on the four main processes:
(1) A slow thermal decomposition o f thiocol, which generates the catalysts
o f A P decomposition;
(2) A catalytic decomposition o f A P , whose decomposition products are
capable o f oxidizing thiocol rapidly at the temperatures studied;
(3) A n oxidative destruction o f thiocol. The products o f this reaction are not
able to catalyze the decomposition o f A P ;
(4) Oxidation o f the catalysts o f A P decomposition and their removal
through the external boundaries o f the system.
The reaction cannot be described on the basis o f a direct interaction o f
A P with thiocol. In this case, the rate should be maximal at the beginning o f
the process. The chemistry o f these reactions has not been studied. A
relatively small heat o f decomposition is determined by a rapid termination
o f oxidative processes due to termination o f the formation o f the products,
which catalyze the decomposition o f A P . The major portion o f A P remains
in the system and, at higher temperatures, its decomposition can be
observed. It was shown that the rate constant o f decomposition o f the
mixture is determined by the rate constants o f the processes enumerated
above with the corresponding notations: k = ky.k - k?/k . O n the whole, the
2 A

thermal stability o f the mixtures o f A P with thiocol is considerably lower


than that o f its mixtures with other binders, which do not contain the active
groups ( N 0 or O N 0 ) . The possibility o f stabilization o f the mixtures o f
2 2

A P with thiocol was not studied.


Among the composite AP-based systems, the A P mixtures with
organometal compounds are o f special interest. The derivatives o f ferrocene
are most important. Along with oxidation by the products of A P
decomposition, these compounds represent the catalysts o f its thermal
decomposition and, in a number o f cases, can be introduced into the mixture
as binder plasticizers. Since the volatility o f ferrocene is rather high, it is
more convenient to employ its derivatives with lower volatility. The
influence o f dimethylferrocene, diethylferrocene, and diisopropylferrocene
on thermal decomposition o f A P and its mixtures with butyl rubber and
218 Ammonium Perchlorate

plasticized dielectric o i l was studied. The substances considered are w e l l


dissolved in such a butyl rubber. The kinetics o f heat release, the dynamics
o f composition variation o f the condensed phase, and accumulation o f the
gas-phase products were studied in [13.36]. A temperature range o f 1 4 0 -
2 3 6 ° C was studied and it was revealed that within the low temperature
range, the two stages could be separated. The first stage is described by a
kinetic equation o f the first order, and the rate constant does not depend on
the type and amount o f organometal compound employed within the range
o f 3 - 5 % . The temperature dependence o f the rate constant takes the form:

£ = 2.5-10 exp(-24000/#r),
8
s" .
1

The thermal effect o f this stage is 3 5 ^ 5 cal per 1 g o f the mixture.


The heat release at the first stage is caused by oxidation o f the
organometal compound because it can be observed only in its presence and
it does not depend on the presence o f butyl rubber. The process proceeds
heterogeneously; the initial rate and heat o f the process are determined by
the specific surface o f A P .
A t the second stage, the reaction acquires a bulk character. The main
heat release is connected with oxidation o f butyl rubber and metalloorganic
compounds by the products o f A P decomposition and perchloric acid. In
this case, the rates o f heat release and thermal effect o f the reaction are
proportional to the content o f organometal compound in the mixture. For
the same amount o f additive, these values are directly proportional to the
iron content in the organometal compound used, i.e., they grow on passage
from diisopropylferrocene ( D P F ) to diethylferrocene ( D E F ) and further to
dimethylferrocene ( D M F ) . The reaction proceeds with self-acceleration. If
the additive amount is 3%, the dependence o f the maximal rate on the
temperature and reaction heat takes the form:

DMF w = 8.4-10 exp(-32100//?r),


9
s'\ 0 = 600 cal/g

DEF w = 8.2-10 exp(-32300//?r),


9
s" , 1
0 = 560 cal/g

DPF w = 7.9-10 exp(-32100//?r),


9
s" , 1
0 = 530 cal/g.

In the analysis o f the decomposition products, it has been revealed that the
gas phase consists o f C O and C 0 ; in the condensed phase, N H
2 ions are
4
+

mainly preserved, C 1 0 ~ ions are transformed into CI", and ferrocene iron is
4

transformed into iron oxide F e 0 . This means that the processes o f


2 3

oxidation o f organic molecules but not ammonia or perchloric acid are


catalyzed. The results obtained can be explained on the basis o f the model
involving the first stage, i.e., the surface reaction o f an organometal
compound with A P resulting in the oxidation o f the organometal compound
up to iron oxide. In the presence o f this oxide, the second stage develops,
Thermal Decomposition and Combustion 219

i.e., a selective oxidation o f the organic parts o f organometal compounds


and butyl rubber. Based on the kinetic data presented above, analysis o f the
regularities o f combustion o f these mixtures was carried out.

REFERENCES

13.1. A . R . H a l l and G . S. Pearson, Oxid. and Combust. Rev., 3: 1 2 9 -


239 (1968).
13.2. A . G . Keenan and R. F. Siegmund, Quart. Rev., 23(3): 4 3 0 - 4 5 2
(1969).
13.3. The Mechanism of Thermal Decomposition of Ammonium
Perchlorate, Collection of papers, Manelis G. B., Ed.,
Chernogolovka: Joint Institute o f Chemical Physics, 3-130 (1981).
13.4. F. Solymosi, Structure and Stability of Salts of Halogen Oxyacides
in the Solid Phase, L o n d o n : John W i l e y , 467 pp. (1977).
13.5. S. Inami, N . Rosser, and H . Wise, J. Phys. Chem., 67(5): 1 0 7 7 -
1083 (1963).
13.6. E . F. Khairetdinov, A . A . M e d v i n s k i i , and V . V . Boldyrev, Kinet.
Katai, 11(5): 1343-1345 (1980).
13.7. G . B . Manelis and Y u . I. Rubtsov, Zh. Fiz. Khim., 40(4): 770-774
(1966).
13.8. Y u . I. Rubtsov, L . P. Andrienko, and G . B . Manelis, Zh. Fiz.
Khim., 51(1): 70-73 (1977).
13.9. B . S. Svetlov, B . A . Koroban, and L . B . Rumyantseva, Theory of
Explosives, M . : Vysshaya Shkola, 119-125 (1967).
13.10. G . B . Manelis, A . V . Proshchin, Y u . I. Rubtsov, V . A . Strunin, and
A . S. Shteinberg, Fiz. Goreniya Vzryva, 3: 305-313 (1968).
13.11. E . P. Goncharov, A . G . Merzhanov, and A . S. Shteinberg, Gorenie
i Vzryv, Proceedings o f the Third Symposium on Combustion and
Explosion, Moscow, Institute o f Chemical Physics, 765-770
(1972).
13.12. A . G . Merzhanov, A . S. Shteinberg, and E . P. Goncharov, Fiz.
Goreniya Vzryva, 2: 185-190 (1973).
13.13. V . I. Orlov, O . P. Korobeinichev, and N . N . M a k s i m o v , Fiz.
Goreniya Vzryva, 6: 57-64 (1989).
13.14. V . A . Koroban, V . P. G u k , and B . S. Svetlov, Proceedings of
Moscow Chemical-Technological Institute, 112: 63-66 (1980).
13.15. V . A . Koroban, B . S. Svetlov, V . P. Guk, and T. I. Smirnova, The
Mechanism of Thermal Decomposition of Ammonium Perchlorate,
Chernogolovka: Joint Institute o f Chemical Physics, 5-29 (1981).
13.16. B . I. K a i d y m o v , Thermochimica Acta, 62(1): 87-99 (1983).
13.17. F. K . Galvey, P. I. Herley, and M . A . Mohamed, Thermochimica
Acta, 132: 205-215 (1988).
220 Ammonium Perchlorate

13.18. Y u . I. Rubtsov, and G . B . Manelis, Zh. Phys. Khim., Chemical


Kinetics and Catalysis, M . : Nauka, 98-101 (1979).
13.19. V . A . K o r o b a n , B . S. Svetlov, and V . M . Chugunkin, Gorenie i
Vzryv, Proceedings of the Third All-Union Symposium on
Combustion and Explosion, M . : Institute o f Chemical Physics,
741-744(1972).
13.20. A . V . Raevskii and G . B . Manelis, Dokl. Akad. Nauk SSSR, 151(4):
886-889 (1963).
13.21. A . V . Raevskii, G . B . Manelis, V . V . Boldyrev, and L . A .
Votinova, Dokl. Akad. Nauk SSSR, 160(5): 1136-1137 (1965).
13.22. A . V . Raevskii, The Mechanism of Thermal Decomposition of
Ammonium Perchlorate, Chernogolovka: Joint Institute of
Chemical Physics, 3 0 - 6 7 (1981).
13.23. V . V . Boldyrev, Y u . P. Savintsev, T. V . M u l i n a , and G . V .
Shchetinina, Kinet. Katai, 11(5): 1131-1139 (1970).
13.24. G . B . Manelis and E . V . Polianchik, Soviet Scientific Reviews,
Section B, Chemistry Reviews, 15(5): 61-101 (1991).
13.25. E . F. Khairetdinov, T. V . M u l i n a , and V . V . Boldyrev, The
Mechanism of Thermal Decomposition of Ammonium Perchlorate,
Chernogolovka: Joint Institute o f Chemical Physics, 101-123
(1981) .
13.26. E . F . Khairetdinov and V . V . Boldyrev, Thermochimica Acta,
41(1): 63-68 (1980).
13.27. T . G . D e v i , M . P. Kannan, and B . Hema, Thermochimica Acta,
285(2): 269-276 (1996).
13.28. Y u . I. Rubtsov, A . V . Raevskii, and G . B . Manelis, Zh. Fiz. Khem.,
44(1): 47-51 (1970).
13.29. V . Y a . Rosolovskii, Chemistry of Anhydrous Chloric Acid, M . :
Nauka, 140pp. (1966).
13.30. K . Raha, S. Ramamurthy, and D . G . Potil, J. Thermal Analysis,
35(4): 1205-1212 (1989).
13.31. D . B . L a i , G . Singh, and C h . Sh. Shukla, Fuel, 61(2): 129-132
(1982) .
13.32. A . A . Said, J. Thermal Analysis, 37(5): 959-967 (1991).
13.33. M . R. Prasad, K . Krishnan, I. David, K . N . N i n a n , and K .
Khrishnamurthy, High Temperature Chemical Process
Symposium, Bombay: 1: 330-336 (1982).
13.34. H . L . Girdhar, A . I. Arora, and G . N . M a l i k , Indian J. of Chemistry
/4, 21: 353-356 (1982).
13.35. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 37(1): 2 3 9 2 -
2396 (1964).
13.36. G . B . Manelis, V . A . Strunin, and A . P. Dyakov, Joint Meeting of
the Soviet and Italian Sections of the Combustion Institute, Pisa:
Tacchi Editore, 7(4) (1990).
Chapter 14

AMMONIUM DINITRAMIDE

The ammonium salt o f dinitramide ( A D N A ) N H N ( N 0 ) 2 4


+
2
_
is the key
compound o f a new class o f inorganic substances: the salts o f dinitramide
( D N A ) synthesized in 1970 [14.1]. A D N A is considered as a challenging
high-energy oxidizer for ecologically pure solid rocket propellants and,
thus, information about its thermal decomposition is o f considerable
interest. The applied problems o f the thermal stability o f A D N A are briefly
considered in [14.2]. In the same paper, a unique characteristic feature o f
A D N A was emphasized: a clearly expressed decomposition at temperatures
near 6 0 ° C . The term "anomalous decomposition" was introduced in [14.2]
to designate the unusual effects observed during the decomposition o f D N A
salts: the increase o f the decomposition rate on transition from the melt to
the solid state and the negative activation energy within a narrow
temperature range, which, as was found later, is adjacent to the melting
point o f the eutectic o f a salt with nitrate o f the corresponding cation. A t
present, many problems concerning either the stability o f A D N A as a
component o f rocket propellants (including the role o f impurities and the
problem of stabilization), or the mechanism of the slow thermal
decomposition o f the N 0 ~ anion have been solved to a great extent [ 1 4 . 3 -
3 4

14.6]. There are also indications o f A D N A decomposition in open systems


at high temperatures, approaching the combustion temperature [14.7, 14.8].

14.1 Decomposition in the Melt


The kinetics o f decomposition o f A D N A was studied most thoroughly in
[14.4] by the manometric method, whose correctness was checked with the
help o f a spectrophotometric determination o f the content o f the N 0 ~ 3 4

anion in the sample during decomposition. The spectral characteristics o f


ADNA in an aqueous solution are as follows: A max = 283.5 nm, the
extinction coefficient Z = 5600 ± 100.

221
222 Ammonium Dinitramide

A l l the kinetic curves are plotted in the AV-t coordinates, where AV is


the volume o f the gas phase products evolved from 1 g o f the substance at
time t reduced to the normal conditions ( 0 ° C , 760 Torr), for convenience o f
comparison o f the experimental results.
For calculation o f the rate constants, the kinetic curves were processed
according to the equation o f the first order or the autocatalysis o f the first
order. In those cases when the process could not be described by these
equations, or when only the initial portions o f the curves were registered,
the rate constants o f the primary decomposition stage were calculated at the
moments o f decomposition o f 0.1 or 1% o f the parent substance.
For a large free volume (m/V f < 10" ) and temperatures above 120°C,
4

the decomposition o f A D N A proceeds according to the law o f the first


order, and the rate o f the process is independent o f the presence o f an
ammonium nitrate impurity in A D N A (Fig. 14.1.1). One mole o f the
substance gives 4 moles o f the gas-phase products, consisting o f N 0 , N 2 2

(nearly 25 v o l % for each o f them), and H 0 (approximately 50%). Such 2

products as N O and N 0 were observed in small amounts. The rate constant


2

o f the first order is described by the following equation:

yt-10 , 4 4
exp(-35500/7?r), s" . 1
(14.1.1)

650

"9 ' I 1
I 1
l ' l • » • I ' l 1
I • I • I • I
0 20 40 60 80 100 120 140 160 180 200 220
t/min

Fig. 14.1.1. The kinetic curves of A D N A decomposition at m/V^- 4 • IO" and the 4

temperatures (1) 150 and (2) 170°C. The ammonium nitrate content in the samples
is (I) 0.3% and (II) 10%.
Thermal Decomposition and Combustion 223

A t temperatures o f 9 5 - 1 2 0 ° C and for m/Vf > 0.01, the reaction is o f a


clearly expressed autocatalytic character (Figs. 14.1.2 and 14.1.3).

t/min t/mln

Fig. 14.1.2 Fig. 14.1.3

Fig. 14.1.2. The initial stages of decomposition of pure A D N A samples (0.3% of


ammonium nitrate) at 100°C and various m/V values: (1) 0.003, (3) 0.024, (4) 0.08,
r

(5) 0.24, and (6) 1.0. Curve (2) corresponds to the following conditions: A D N A with
1 0 % o f N H N O at
4 3 m/V =0.\.
r

Fig. 14.1.3. Decomposition of A D N A in the melt at 104°C and m/V = 0.3: (1) the f

experimental curve, (2) the results of calculation by equation (14.1.1).

Table 14.1.1. Composition of the gas phase products of A D N A decomposition in


the melt

T,°C m/V { Product content, (%)

N 2 N 0
2 H 0
2 N0 2

104 0.3 2.5 43 53 traces traces


104 0.3 3.8 30 66 - -
104 0.3 4.8 28 68 - -
104 0.3 100 10 90 - -
170 io- 4
100 26 27 45 -

The gas-phase or volatile products are the catalysts. This follows from two
facts: first, the rate o f the autocatalytic stage depends on m/Vf (Fig. 14.1.2)
and, second, after evacuating the gas-phase products at the stage o f
acceleration, the rate always decreases to the initial level. It was shown by
direct experiments that H N 0 3 and N 0 2 could catalyze decomposition.
224 Ammonium Dinitramide

Under the conditions mentioned above, both o f these products, as w e l l as


water, are well dissolved in the liquid A D N A . M a i n l y , the gas phase
consists o f N 0 and N , and, in the course o f decomposition, the ratio
2 2

between these products essentially changes (Table 14.1.1). The character o f


the catalytic stage o f decomposition also varies. The overall kinetic curve o f
A D N A decomposition at 104°C and m/V { = 0.3 is presented in F i g . 14.1.3.
U p to the degree o f conversion o f approximately 10%, this curve is
formally described by the equation o f autocatalysis o f the first order:

drj/dt = k (1 - 77) + k rj(l - 77),


x 2
(14.1.2)

where rj = AV/AV^ is the degree o f conversion, k\ = 9.92 • 10" , s" and k = 7 1


2

5.60 • 10"*, s . The ratio o f the rate constants is k /k\ = 564. This means that
_1
2

already at the degree o f conversion o f 0.2%, the rates o f the initial and
catalytic stages w i l l be equal. This conclusion is in agreement with the data
o f F i g . 14.1.2: at 100°C, the linear portion o f the curves for various m/Vf
values is observed only up to the degree o f conversion o f 0.1-0.2%.
After decomposition o f 10% o f the substance, the character o f the
catalytic stage changes ( F i g . 14.1.3). The curve, calculated by equation
(14.2), differs sufficiently from the experimental one, w h i c h is described by
the following equation o f the first order after decomposition o f 15% o f the
substance:

drjldt = £ 3 ( 1 - 7 7 ) , (14.1.3)

where k 3 = 5.0 • 10" , s . After decomposition o f 15% o f the substance,


5 _1

water, nitric acid, and N 2 are not formed at all, and N 0 and N H N 0 2 4 3

become the only products o f decomposition. In spite o f variations o f the gas


phase composition, the stoichiometric reaction coefficient with respect to
the gas release remains constant and can be as large as 0.95-1.00 m o l o f the
products per 1 mole o f A D N A , w h i c h corresponds to AV W = 1.72-1.80. The
rate constants o f the initial stage, calculated at the time o f decomposition o f
0.1%) o f the substance, at temperatures o f 100, 105, and 115°C, nicely fit the
temperature dependence, described by equation (14.1.1) ( F i g . 14.1.4).
The occurrence of autocatalysis offers a possibility o f ADNA
stabilization in the melt by binding the catalysts by chemical reagents.
Experiments on A D N A decomposition at 9 8 - 1 2 0 ° C in the presence o f
various amines and ammonia, salts o f weak acids ( N H F , K P 0 ) , w h i c h are 4 3 4

able to reduce the acidity o f the medium, and, also, substances o f


nitrobenzene type, w h i c h are easily oxidized by H N 0 3 and N 0 , were
2

carried out. The amount o f additive was 1-3 m o l % . It turned out that all o f
these compounds reduce the rate o f the catalytic stage and increase the
induction period by a factor o f 2 - 3 . In addition, the strong amines o f the
urotropine type and ammonia influence the initial stage, reducing its rate by
Thermal Decomposition and Combustion 225

approximately a factor o f two. In the experiments with the deuterium


A D N A , the kinetic isotopic effect is observed for the rate constant o f the
initial stage: k /k
H D = 1.38.
The results obtained can be reasonably explained within the framework
o f the reaction mechanism presented in Scheme 1.

Scheme 1

H NN(N0 )
4 2 2 NH 4
+
+ NN0 2 + N0 2

H N N ( N 0 ) ^ N H + HN(N0 )
4 2 2 3 2 2

NNO2+NO2 ONNO2 + NO

20NN0 2 N 0 + N0 * + N0 "
2 2 3

N H + NNO" 4
+
2 NH + HNN0 3 2

NH 4
+
+ ONNO" 2 N H + HONN0 3 2

HN(N0 ) — 2 2 HNO3+ N 0 2

HNN0 2 OH +N 0 2

H O N N 0 — O H + 2NO 2

NH +O H ^ N H 3 2 + H 0 2

NH + N0 ^ 3 2 NH + HN0 2 2

12
NH2+NO—^N + 2 H 0 2

NH + N0 ^ 2 2 N 0 + H 0 2 2

2 H N 0 2 ^ NO + N O 2 + H 0 2

HN0 — 3 N 0 + 0 . 5 H O + 0.25O 2 2 2

16
2N0 + H 0 = ^ H N 0 + HNO3
-16
2 2 2

J 7

H NN(N0 ) 4 2 2 + HN0 2 N H N 0 + HN(N0 )


4 2 2 2

18
H NN(N0 ) + HN0 ^ = N H N 0 + HN(N0 )
19 -
4 2 2 3 4 3 2 2
1 8

NH N0 =N +2H 0
-19
4 2 2 2

2 N 0 ^ N 0
-20
2 2 4
2 4

N 0 — NO +N0 - +

-21
2 4 3

22

NO + H N N ( N 0 ) ^ y
+
4 2 2 NO N(N0 ) "+NH
+
2 2 4
+

N O N ( N 0 ) " i l N 0 + 2N0
+
2 2 2 2
226 Ammonium Dinitramide

Fig. 14.1.4. Arrhenius dependence of the rate constant of the initial non-catalytic
stage of A D N A decomposition: (1) the melt, m/V < 4 • K T , (2) the melt, m/V > 0.1,
f
4
r

(3) the solid phase, the water content in the sample is 0.1%, (4) the solid phase, the
water content in the sample is 0.5%.

The initial decomposition o f A D N A proceeds through two channels, w h i c h


represent the monomolecular decomposition o f the anion along the N - N
bond and dissociation o f the salt into a base and acid. The first reaction is
observed during the decomposition o f all D N A salts with metals and strong
bases [14.9, 14.10]. The second channel is opened for the reactions o f
onium salts with amines o f a medium strength [14.10].
Such facts as the kinetic isotopic effect, inhibition o f the initial stage by
ammonia and strong amines (which are able to displace ammonia from
A D N A ) , and, finally, sublimation o f A D N A in open vessels in vacuum are
indicative o f the decomposition o f A D N A , at least partially (no more than
50% at 100°C), through proton transfer.
In the case o f metal salts o f D N A , stages (3) and (4) are the only
possible pathways o f the loss o f the primary anion-radicals, formed after the
N0 2 detachment from the anion [14.9]. In the case o f onium salts, according
to stages (5) and (6), the proton transfer competes favorably with these
processes ( N 0 ~ and N 0 ~ are stronger nucleophilic radicals than N 0 ~ ) ,
2 2 2 3 3 4

and N 0 2 is accumulated in the system, to be, in turn, consumed on


ammonia oxidation. The ionic reactions proceed in the condensed phase, the
others reactions can proceed either in solution or in the gas phase.
For large values o f m/V , f the products are dissolved in liquid A D N A ,
and equilibrium (16) is attained. The decomposition is catalyzed by H N 0 2

and H N 0 . A s N H N 0
3 4 3 is accumulated, the efficiency o f the catalysis
reduces (due to the reversibility o f reaction (18)), and it proceeds as the
chain process (interchange o f stages (18) and (7)) at the stationary HN0 3
Thermal Decomposition and Combustion 227

concentration. The stabilizing influence of ammonium nitrate was


experimentally demonstrated (see F i g . 14.1.2). The influence o f N 0 2 can
be explained by the chain reaction (stages (20)-(23)), which is described by
the rate equation o f the first order.
Studies o f the liquid phase decomposition o f A D N A by the calorimetric
method [14.5] showed results which are in agreement with the data derived
by the manometric method. The following characteristic features were
revealed: the autocatalytic character o f decomposition, the catalysis by
means o f H N 0 , and the catalytic stage inhibition on addition o f ammonium
3

nitrate. T w o pathways o f the initial decomposition (stages (1) and (2) o f


Scheme 1) were revealed. The following expression was obtained for the
rate constant o f the initial stage:

k = 7.7 • 1 0 M
e x p ( - 3 4 6 0 0 / RT\s" ]
.

In [14.7], the decomposition of ADNA was studied by the


thermogravimetry ( T G ) method under isothermal and non-isothermal ( 0 . 0 1 -
10 7
K / s ) conditions with the IR-analysis o f the products. It was confirmed
that under conditions o f a good gas outlet, the reaction does not proceed
according to the first order. The compounds N 0 2 and N 0 were registered
2

as the primary products o f decomposition.


In [14.8], the yield o f N , N 0 , H 0 , and N H 2 2 2 3 during the decomposition
of the products of A D N A sublimation (i.e., the mixture o f N H 3 +
HN(N0 ) ) 2 2 under high-speed heating was measured by the mass-
spectrometric method. A kinetic scheme involving 152 reactions was
proposed, which described rather well the overall kinetics with respect to
the yield o f all products. The initial stage o f decomposition o f H N ( N 0 ) 2 2 is
the reaction:

HN(N0 ) —^ H N N 0
2 2 2 + N0 , 2

for which, £ a c t = 38 kcal/mol. Further stages are the decomposition o f


HNN0 2 into N 0 and O H and bimolecular reactions o f H N N 0
2 2 with O H ,
N O , and N 0 . 2

14.2 Decomposition in the Solid Phase


The rate o f decomposition o f A D N A in the solid phase was measured in
[14.4] at temperatures o f 4 0 - 8 0 ° C . The main results were obtained for the
sample containing 0.3% o f ammonium nitrate and 0.5% o f H 0 . The 2

reaction proceeds with the acceleration, which becomes noticeable after the
decomposition o f 0.05% o f the substance. The reason for the acceleration is
the autocatalytic decomposition o f A D N A in the liquid phase, the amount
228 Ammonium Dinitramide

of w h i c h increases as the products o f decomposition are accumulated. The


induction periods are very long (120 h at 80°C and 3 months at 6 0 ° C ) .
Nevertheless, in [14.6], the decomposition o f A D N A at 80°C was registered
up to the decomposition o f 10% o f the initial substance. A t these deep
stages, the reaction is subjected to inhibition (as well as in the melt).
Methyldiphenylurea was found to be the best stabilizer.
A t 8 0 ° C , the products o f decomposition o f A D N A are N (about 10%) 2

and N 0 . One mole o f the gas-phase products (180 N - c m ^ g " ) is formed


2
1

from one mole o f A D N A , as in the case o f large values o f m/V in the liquid f

phase. The rate constant, calculated from the rate o f decomposition on the
linear portion o f the curve takes the form:

k =lO
s
U M
exp(-33600/RT), s~\

The additives o f ammonia do not influence the initial rate o f decomposition


o f the solid samples. The kinetic isotopic effect observed during the liquid-
phase decomposition of N D N ( N 0 )
4 2 2 is absent in the solid phase.
Apparently, the decomposition channel through dissociation o f the salt into
a base and acid in the solid phase is not o f considerable importance.
A m m o n i u m nitrate is present as an impurity in all samples o f A D N A
and gives the eutectic mixture o f the following composition: ( 1 : 2 ) with the
melting temperature o f 6 0 ° C . A t 6 0 - 8 0 ° C , the decomposition o f A D N A in
the eutectic mixture proceeds at the same rate as in the melt, i.e., it is
described by equation (14.1). B e i n g chemically inert, ammonium nitrate
influences the rate o f decomposition by the mechanism o f submelting. Due
to the appearance o f the liquid phase, whose fraction increases as the
temperature increases (at 8 0 ° C , the melt already consists o f 1 part o f nitrate
and 6 parts o f A D N A ) , the rate and activation energy o f decomposition o f
ADNA within the range o f 6 0 - 8 0 ° C should depend on the content o f
ammonium nitrate.
The second impurity, which is also always present in A D N A , is water.
In small amounts (up to 0.5%), water does not accelerate decomposition.
Apparently, it exists in the adsorbed state and does not give the liquid
phase, in w h i c h A D N A can be dissolved. However, for a content o f 1%
H 0 , at 8 0 ° C , the rate increases by a factor o f six.
2

The accelerating influence o f water is connected with the dissolving o f


A D N A . Nevertheless, the other, more important and unusual phenomenon
is connected with water content in the ADNA sample: a so-called
anomalous decomposition, i.e., the increase o f the rate during the
decomposition o f dry (approximately 0.1% H 0 ) samples in vacuum (a 2

closed vessel, the initial pressure is 0.1 Torr) or in inert atmosphere ( N , dry 2

air) above the rate level in the liquid state.


O n the whole, at 40, 50, 70, and 80°C the increase o f the rate is not
large, but at 6 0 ° C , i.e., at the melting temperature o f the eutectic mixture
Thermal Decomposition and Combustion 229

with nitrate, a considerable peak o f the rate is observed ( F i g . 14.1.4). The


value o f this peak depends on the method o f drying and on the degree o f a
preliminary drying o f the sample. Thus, for instance, an A D N A sample with
an initial water content o f 0.65% was dried at 3 0 ° C in a vacuum o f 0.1 Torr
for 12, 24, and 52 h, and, after that, for an extra 6 h at 8 0 ° C . The water
content decreased, correspondingly, to 0.2, 0.12, 0.08, and 0.07%, and at
6 0 ° C the rate increased by a factor o f 1.4, 80, 900, and 8 • 10 . 3

Introduction o f 10 Torr o f H 0 vapors immediately terminates the


2

decomposition o f anomalous samples. A m m o n i a , methanol, acetone, N O ,


N 0,
2 and C H J act similarly but with smaller efficiency. A l l o f these
3

substances possess two common properties: the presence o f a dipole


moment and the ability to solvate charged particles.
Phenomenological description o f the anomalous decomposition in the
most detailed form is given in [14.11, 14.12], which studied metal salts o f
D N A . The characteristic features o f the anomalous decomposition are the
following: increase o f the decomposition rate in vacuum up to a level higher
than in the liquid phase, the peak o f the rate in the Arrhenius dependence at
the melting point o f a salt with nitrate o f a corresponding metal, and
instantaneous inhibition by water vapors. In contrast to metal salts, onium
salts o f D N A are not subjected to the anomalous decomposition [14.10]. In
this respect, A D N A occupies an intermediate position: the anomalous
decomposition is observed only at the melting point o f an eutectic mixture
and, at other temperatures, the rate growth is insignificant.
Similarly to [14.11, 14.12], one can relate the appearance o f the
anomalous decomposition o f D N A salts to the asymmetrical geometric and
electron structure o f the N 0 ~ anion in the crystalline state [14.13, 14.14].
3 4

In the anion, in which nitro-groups are not equivalent and the charge is
concentrated mainly near one o f them, the reactivity increases, and the new,
rapid channel o f decomposition is opened: N 0 ~ elimination to separate
3

N 0 . The hydrogen bonds between the anion and H-containing cations or a


2

direct solvation by water molecules reduce the anion polarization, smooth


its electron asymmetry, and impede the intramolecular rearrangement.
Further, one can suppose that during the formation o f the nucleus o f the
eutectic phase, the local rearrangement o f the crystal lattice proceeds, and
the stresses and new defects appear in the crystal. This can influence the
rate in the case when the reaction proceeds on the surface considerably
faster than in the bulk o f the crystal. Apparently, this takes place in the case
of the decomposition o f D N A salts without inhibition by water vapors. The
process o f eutectic formation w i l l be even more important for them, i f the
crystalline and occluded water can be eliminated. For A D N A , one should
also suppose that eutectic formation is accompanied by breaking o f
hydrogen bonds. The breaking o f H-bonds should result at least in
insignificant rearrangement o f the anion sublattice, and, thus, can be
observed by the method o f X-ray diffraction analysis. Thus, it becomes
230 Ammonium Dinitramide

possible to check the proposed mechanism o f the anomalous decomposition


with the help o f the structural studies o f A D N A at 6 0 ° C .

REFERENCES

14.1. O. A . L u k ' y a n o v , V . I. Gorelik, and V . A . Tartakovskii, Izv. Akad.


Nauk, Ser. Khim., 94-97 (1994).
14.2. Z . Pak, AIIA/SAE/ASME/ASEE 29 Joint Propulsion Conf, June
2 8 - 3 0 , Monterey, C A , FI1A-93-1755 (1993).
14.3. G . B . Manelis, AGARD Conf. Proceedings 599, "Environmental
Aspects of Rocket and Gun Propulsion", Aalesung (Norway), 15.1
(1994); Proc. 26th Int. Ann. Conf. oflCT., 1 5 - 1 - 1 5 - 1 6 , Karlsruhe,
Germany (1995).
14.4. A . N . Pavlov, G . M . N a z i n , and G . B . Manelis, Izv. Akad. Nauk,
Ser. Khim., (\99S).
14.5. A . I. K a z a k o v , Y u . I. Rubtsov, L . P. Andrienko, and G . B .
Manelis, Izv. Akad. Nauk, Ser. Khim., 395-401 (1998).
14.6. S. Lobbecke, H . Krause, and A . Pfeil, 28th Int. Annual Conf. of
ICT, June 2 4 - 2 7 , 1 1 2 - 1 - 1 1 2 - 8 , Karlsruhe, Germany (1997).
14.7. C h . A . Wight and S. V y a z o v k i n , MURI Workshop on Ammonium
Dinitramide (ADNA), T. B r i l l E d . , Reno, Nevada, January 15, 7 8 -
86 (1998).
14.8. J. Park and M . C. L i n , MURI Workshop on Ammonium
Dinitramine (ADNA), T. B r i l l E d . , Reno, Nevada, January 15, 6 5 -
77 (1998).
14.9. F. I. Dubovitskii, G . A . V o l k o v , V . N . Grebennikov, G . B .
Manelis, and G . M . N a z i n , Dokl. Rus. Akad. Nauk, 347: 763-765
(1996).
14.10. A . N . Pavlov and G . M . N a z i n , Izv. Rus. Akad. Nauk, Ser. Khim.,
1951-1953 (1997).
14.11. F. I. Dubovitskii, G . A . V o l k o v , V . N . Grebennikov, G . B .
Manelis, and G . M . N a z i n , Dokl. Rus. Akad. Nauk, 348: 205-206
(1996).
14.12. S. B . B a b k i n , A . N . Pavlov, and G . M . N a z i n , Izv. Rus. Akad.
Nauk, Ser. Khim., 1945-1950 (1997).
14.13. F. I. Dubovitskii, N . I. G o l o v i n a , A . N . Pavlov, and L . O .
A t o v m y a n , Dokl. Rus. Akad. Nauk, 355: 200-206 (1997).
14.14. R. G i l a r d i , J. Flippen-Anderson, C . George, and R. J. Butcher, J.
Am. Chem. Soc, 119: 9411-9416 (1997).
Chapter 15

THE SALTS OF HYDRAZINIUM,


HYDROXYLAMMONIUM, AND
NITRONIUM

Substitution o f the ammonium ion in the oxidizer for hydrazinium or


hydroxylammonium ions results in an increase o f the heat o f combustion
and energetic efficiency o f a composite, but in contrast to the N H 4
+
ion,
they are capable o f thermal transformations at a considerably higher rate by
themselves and, in a number o f cases, just an insufficiently high thermal
stability o f the salts o f hydrazinium and hydroxylammonium hampers their
commercial use. Since thermal transformations with participation o f N H 2 5
+

and N H 0 4
+
ions are rather complex and in the corresponding nitrates and
perchlorates these reactions are imposed on the processes considered above
with participation o f nitric and perchloric acids, it is appropriate to take into
consideration such salts, in which transformations of cations and
corresponding bases are the only possible thermal transformations and the
acids forming them are thermally stable. The salts o f hydrazinium are
melted at quite low temperatures and the reaction o f their decomposition is
usually studied in the liquid phase. Thermal decomposition o f all salts o f
hydrazinium proceeds through their preliminary dissociation into hydrazine
and acid. The rate o f decomposition is determined by the concentrations and
the rate constants o f transformation o f these products, and, in the majority
o f cases, thermal transformations o f hydrazine are the determining stages.

15.1 Hydrazinium Nitrate and Chloride


The salts o f hydrazinium examined can be divided into several groups by
the mechanisms o f their thermal decomposition. It is expedient to begin
with consideration o f hydrazinium nitrate (T meh = 7 0 . 5 ° C ) and chloride
(Tmeit =
9 1 . 5 ° C ) [15.1, 15.3]. The reaction o f thermal decomposition o f the
melts o f hydrazinium nitrate N H N 0 2 5 3 ( H N ) and chloride N H C 1 ( H C )
2 5

231
232 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

proceeds with self-acceleration and is described by an equation o f the


reaction o f the second order with the autocatalysis o f the first order.

F o r N H N 0 ; k = 10
2 5 3 x
1 2 2
exp(-38000/7?r), s" 1

£ 2 = 10 7 8
exp(-26500/ivr), s" 1

at K / = 9 - 1 0 , 3
cm . 3

The heat o f decomposition is 3 5 5 cal/g or 3 3 . 8 kcal/mol.

F o r N ^ C l ; ^ -10 , 3 2
exp(-40000/i?r), s _1

yt = 1 0
2
7 7
exp(-26000/7?r), s" 1

at V = (5-\0)-\0\
0
8
cm . 3

The heat o f decomposition is 527 cal/g or 36.1 kcal/mol.


For both salts, k\ very weakly and k strongly depend on V decreasing 2 Q
g

with its increase. The k values are almost equal for both salts. The overall 2

equations o f thermal decomposition take the form:

3N H N0 2 5 3 = 2.25NH4NO3+O.75N2O+N2 + N H 3

3N H C1 2 5 = 3NH4CI + N H 3 + N2.

Here, H N 0 partly participates in N H oxidation.


3 2 5
+

T o understand the mechanism o f thermal decomposition o f these salts,


it is important that hydrazine, formed during the equilibrium dissociation o f
the salt, forms the complex with a hydrazinium ion:

N H 4 + N H 5 ^ : N H4-N H5 .
2 2
+
2 2
+

The existence o f such complexes was confirmed in studies o f the fusibility


curves o f the N2H5NO3-N2H4 and N H C 1 - N H 4 systems and on the 2 5 2

measurement o f the equilibrium pressure o f hydrazine over the N H C 1 - 2 5

N H system [15.3]. Such complexes exist in the melts o f all salts o f


2 4

hydrazinium. For hydrazinium azide, the complex salt was separated in the
pure form. The complex N H C 1 • N H melts at 3 6 ° C and N H N 0 • N H 2 5 2 4 2 5 3 2 4

melts at 3 ° C , but the maxima in the fusibility curves are weakly expressed.
The complexes noticeably dissociate already near the melting point. The
rate constant o f the formation o f complex ions was obtained from the
difference o f the equilibrium pressure o f hydrazine over the pure hydrazine
and its solution:

= N H -N H
C
2 4 2 5
+ / C
N H C 2 4 N 2 H 5
+
=10- exp(10500//?r),
6 5
1/mol.
Thermal Decomposition and Combustion 233

A t 3 6 ° C , K = 8 1/mol, which corresponds to the dissociation o f the complex


c

by 30%, at 2 0 0 ° C , K = 2 • 10" 1/mol. The limiting stage o f thermalc


2

decomposition o f H N and H C is the decomposition o f these complex ions:

N H - N H — NH + N H
2 4 2 5
+
3 3 6
+

N H + N H — NH + N H
3 6
+
2 4 3 4 7
+

N H + N H — N H + N + 2NH
4 7
+
2 4 2 5
+
2 3

3 N H = 4NH + N
2 4 3 2

This is confirmed by direct measurement o f the rate o f thermal


decomposition in the mixtures o f H C and H N with hydrazine [15.4]. In such
a mixture, for both salts, the rate o f decomposition is the same:

- d C
K H 2
+
5
/ d t = kC
NH NH 2 A 2
+
5
= kK
C N H, C
2 ^ N ^ ; '

The temperature dependence o f the rate constant A eff takes the form:

k =k
eff K =\0 Qxp
c
l25
(-37000/T?r), 1/mol-s.

U s i n g this dependence and a similar dependence for k u one can calculate


the values o f Q s j ^ in the melts o f pure salts and K P
C
o f the dissociation o f
these salts into a base and acid:

N H N0 :
2 5 3 C N 2 H 4 = 1 0 " ° - e x p ( - 1 0 0 0 / r t r ) > rnolll\ K = 1 0 - e x p ( - 2 0 0 0 / / ? r )
3 c
p
3()

N H C1:
2 5 C N z H 4 =10" exp(-2000//?r), mol17; K = 10" e x p ( - 6 0 0 0 / / ? r ) .
()7 c
p
13

In this mechanism, the rate o f decomposition is determined only by the


concentration o f complex ions, which, in turn, depends only on the
concentration o f hydrazine in the melt, i.e., on the degree o f dissociation o f
the salt into hydrazine and acid. For such a mechanism, the lowest
decomposition rates are realized, and all supplementary interactions w i l l
result in an increase o f the decomposition rate. Apparently, hydrazinium
hexafluorophosphate N H P F , which was studied by differential thermal 2 5 6

analysis ( D T A ) , has the same mechanism and close decomposition rates


[15.5]. The decomposition proceeds with noticeable rates above 2 0 0 ° C , and
the main products are N H P F , N H , and N . There are no quantitative 4 6 3 2

kinetic data available. Sulphate ( N H ) S 0 , mono- and disubstititued 2 5 2 4

phosphates N H H P 0 and ( N H ) H P 0 , which were also studied by the


2 5 2 4 2 5 2 4

D T A method, with the peak o f decomposition at temperatures o f 2 0 0 -


2 5 0 ° C , can be also related to the same group [15.6]. A t the same time,
hydrazinium picrate N H C H ( N 0 ) 0 decomposes considerably faster, 2 5 6 2 2 3
234 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

and the peak o f decomposition under similar conditions is observed at 1 6 0 -


180°C. The reasons for this fact were not discussed in the work. Probably,
this is connected with the low thermal stability o f picric acid.

15.2 Hydrazinium Iodide


The most typical example o f interaction, leading to an increase o f the
decomposition rate, is the reduction or oxidation o f the hydrazinium cation
by the acid formed. The reaction o f reduction o f hydrazinium ion,
proceeding at thermal decomposition o f iodine hydrazinium, was studied in
detail in [15.7, 15.8]. The thermal stability o f hydrazinium iodine is
considerably lower than that for hydrazinium nitrate or chloride, and the
equal decomposition rates are attained at temperatures which are lower by
100°C. The growth o f the rate is determined by the interaction o f iodine
hydrogen, formed during the equilibrium dissociation o f the salt, with the
other molecule o f the salt by the following equation:

N H I+HI
2 5 = 2NH + I ; 3 2

_
^ N 2 H 5 I ldt = k C N 2 H 5 I C H I .

Iodine rapidly oxidizes hydrazine to nitrogen with the reduction o f the


parent HI molecules. The overall process is similar to the case o f
decomposition o f H C :

3N H I 2 5 = 3NH I + N H + N .
4 3 2

In this case, the dependence o f the decomposition rate on the acidity o f the
system is inverse in comparison with the decomposition o f N H C 1 , and the 2 5

rate increases at the introduction o f excessive acid into the system and
decreases at the introduction o f bases. Accumulation o f ammonia in the
ampoule, w h i c h is the final decomposition product, results in a rapid
reduction o f the rate, and the kinetic curves are described by the following
equation:

d7]/dt = k (\-rj) - ;
]
1 5
k =AC
x N 2 H I = 10
5
13 5 5
exp(-33000/7?r), s" .
1

The fractional order is determined by the redistribution o f ammonia


between the gas and liquid phases. Thermal decomposition o f the complex
salt o f a 2 N H I • N H
2 5 2 4 composition was also studied. The rate o f its
decomposition is considerably lower, and the temperature dependence o f
the initial rate takes the form:
Thermal Decomposition and Combustion 235

" ( < ^ N H I /d*)t=0 = °


2 5
1 1 6 3 5
e x p ( - 3 7 0 0 0 / RT), mol/1 • s.

T a k i n g into account that in the melt o f this salt C N 2 H 4


=
6.1 mol/1, C N H I 5

= 12.25 mol/1, in a pure melt o f N H I C N H 2 5 2 4


=
C H I , for N H I one can
2 5

obtain:
K P = 10 exp(-8000//?r).
30

Here, the value o f Kp and its temperature dependence are somewhat higher
than the dependences in hydrazinium chloride and nitrate.

15.3 Hydrazinium Azide


Decomposition o f hydrazinium azide N H N 2 5 3 (r meit = 7 5 ° C ) [15.9] occupies
an intermediate position between the two decomposition pathways
considered above. A c i d i t y o f hydrazoic acid is considerably lower than that
of the acids considered above. Thus, the degree o f reverse dissociation o f
the salt is noticeably higher, as well as the concentration o f the complex
ions N H 2 4 • N H . 2 5
+
Hydrazoic acid is able to reduce hydrazine and
ammonia, but the rate o f this reaction is lower than the rate o f the reduction
of hydrazine by the H I molecule. In the melt o f this salt, within the
temperature range o f 1 2 0 - 1 7 0 ° C , at the beginning o f the process, mainly,
the reduction o f the parent salt by the H N molecule occurs: 3

N H N +HN3
2 5 3 = 2NH +3N . 3 2

Nitrogen, formed in this reaction, is certainly not able to oxidize hydrazine


similarly to I in iodine hydrazinium. Thus, the reaction halts at this stage.
2

Accumulation o f ammonia rapidly reduces the equilibrium concentration o f


HN 3 and increases the concentration o f complex ions N H . The 4 9
+

contribution o f their decomposition to the total rate increases and the


contribution o f the reaction through H N decreases. The change o f the 3

mechanism is very rapid and, further, the reaction proceeds with slight
acceleration due to N H accumulation. Finally, the initial rate o f 3

decomposition o f N H N has an intermediate value between the


2 5 3

decomposition rates o f N H C 1 and N H I . The kinetics is described by the


2 5 2 5

equation o f autocatalysis o f the first order. The corresponding values o f the


rate constants take the form:

k = 10
}
8 0
exp(-24000/^r), s" ; k = 1 0
1
2
4 0
exp(-l6000/^7), s" .1

Irreversible consumption o f acid and accumulation o f ammonia result in the


growth o f the hydrazine concentration; thus, after consumption o f all N H 2 5
+
236 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

ions, a particular amount o f N H molecules remains in the system, w h i c h ,


2 4

at the temperatures studied, are rather stable in the lack o f ions and
represent the final products o f decomposition. The overall equation o f
decomposition takes the form:

N H N 2 5 3 = 0 . 2 5 N H + 1.33NH + 1.58N .
2 4 3 2

The heat o f decomposition is 910 cal/g or 68 kcal/mol. The growth o f the


heat o f decomposition in comparison with the other salts o f hydrazinium is
connected with i n v o l v i n g hydrazoic acid in the decomposition, and the l o w
effective activation energies o f decomposition are determined by a high
volatility o f the dissociation products. W i t h temperature increase their
content in the melt decreases. A t thermal decomposition o f hydrazinate o f
hydrazinium azide N H N 2 5 3 • N H 2 4 due to a sharp decrease o f the
equilibrium concentration o f H N , from the very beginning, the reaction 3

proceeds through the decomposition o f complex ions N H 4 9


+
and the reaction
rate is lower than in the case o f hydrazinium azide by a factor o f 17-20.

15.4 Hydrazinium Perchlorate and Diperchlorate


These salts occupy a special place among the compounds considered. The
mechanism o f thermal decomposition o f perchloric acid is very complex
and has been insufficiently studied. The presence o f N H molecules and 2 4

their ions, which are capable either o f oxidation to nitrogen or reduction to


ammonia, makes this process even more complicated. The rate o f their
decomposition strongly depends on the experimental conditions, especially
on the system volume K . It is rather difficult to compare the rates obtained
0
8

under different conditions. Thermal decomposition o f N H C 1 0 (r m e i t = 2 5 4

141°C) proceeds with a very weak self-acceleration [15.10] according to the


following equation:

8N H C10
2 5 4 = 7NH C10 +NH C1+ 4 N + 4 H 0 .
4 4 4 2 2

A t temperatures o f 1 3 0 - 1 6 0 ° C , the initial rate is close to the rate o f N H I 2 5

decomposition. The temperature dependence o f the rate is o f a complex


character, and the effective activation energy changes from 25 to 40
kcal/mol. It is difficult to expect that the molecules o f H C 1 0 4 are able to
reduce hydrazine to ammonia similar to H I and, on the other hand, the
consumption o f these molecules is insignificant and approximately 9 0 % o f
C10 ~
4 anions remain at the end o f the process. M o s t probably, H C 1 0 4

molecules initiate the reduction-oxidation decomposition o f hydrazine and


complex ions N H 4 9
+
to ammonia and nitrogen, playing the role o f an
oxidizer o f nitrogen-containing molecules to nitrogen. Thus, the bases are
not accumulated and self-acceleration is not pronounced. Such initiation is
necessary because the initial rate o f decomposition o f N H 4 9
+
ions is lower
Thermal Decomposition and Combustion 237

by several orders o f magnitude. Interaction o f hydrazine with HI0 3

proceeds at considerably higher rates. In attempts to synthesize N H I 0 , a 2 5 3

rapid decomposition was already observed at 0°C [15.5].


The kinetic regularities o f decomposition o f hydrazinium diperchlorate
N H ( C 1 0 ) 2 can serve as an indirect confirmation o f the interaction o f
2 6 4

HCIO4 with N H 2 4 or N H 4 9
+
. The equilibrium dissociation o f this salt
proceeds according to the following equation:

N H (C10 )
2 6 4 2 — N H C10 2 5 4 + HCIO4.

Thus, free N H 2 4 molecules and N H 4 9


+
ions in this salt are almost absent, and
the concentration o f H C 1 0 is higher by several orders o f magnitude than in
4

monoperchlorates considered earlier, and the concentration o f C N H + is

also high. The initial rate o f decomposition o f this salt also strongly depends
on the value o f K and under comparable conditions is lower than in the
0
g

case o f N H C 1 0 by a factor o f 10-12; at the same time, the degree o f self-


2 5 4

acceleration is higher. The reaction o f decomposition proceeds by the


following equation:

2N H (C10 )
2 6 4 2 = 2HCIO4+ 2 N + 5 H 0 + C l + 1.50 . 2 2 2 2

Obviously, during N H C 1 0 2 5 4 decomposition, the direct interaction o f N H 2 5


+

ions with the H C 1 0 4 molecule does not occur, otherwise, this interaction
w o u l d have resulted in a considerably higher rate o f decomposition o f
diperchlorate. O n l y the reactions o f N H 2 4 or their complexes can provide
higher rates of decomposition of N H C10 .
2 5 4 For N H (C10 ) ,
2 6 4 2

decomposition through the decay o f perchloric acid is quite probable. The


higher rate o f the process in comparison with ammonium perchlorate is well
explained by a higher concentration o f H C 1 0 4 molecules. The N H 2 5
+
and
N H
2 6
2 +
ions are oxidized either by perchloric acid or by the intermediate
products o f its decomposition. Finally, even ammonia is not preserved. This
confirms the leading role o f the decomposition o f perchloric acid.
Eventually, the aqueous solution o f perchloric acid remains, whose rate o f
decomposition is very low at the temperature studied. The process is
finished after a complete oxidation o f hydrazine and ammonia compounds.
The comparison o f the initial rates o f thermal decomposition o f all salts o f
hydrazinium considered is presented in Table 15.4.1. Eventually, during
thermal decomposition o f hydrazinium salts, the whole spectrum o f the
possible processes for onium salts is realized: from the reactions, involving
only transformations of hydrazine, through different intermolecular
interactions o f acid and hydrazine, to decomposition o f acid with the
subsequent oxidation o f hydrazine. In all cases, the preliminary equilibrium
dissociation o f the salt into acid and base is the first stage o f the process. It
should be emphasized that only the reactions o f hydrazine (through the
238 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

complex ions N H ) or only acid ( H C 1 0 ) give lower rates o f thermal


4 9
+
4

transformations than the interaction o f acid and base with each other.

Table 15.4.1. The initial rates of thermal decomposition of hydrazinium salts

Salt k =f(T), s-'


t

120°C 150°C 200°C

N 2 H N0 5 3 10 l22
exp(-38000/RT) ,0-8.9
io- 74
10""
N 2 H CI
5 10 l32
exp(-40000/RT) , -9 0
0 1(T 74
10" 52

N 2 H I
5 10 l36
exp(-33000/RT) 10^ 8
1(T 35
IO" 17

N 2 H N
5 3 10 80
exp(-24000/RT) 10'" io- 4 4
1(T 31

-
N H CI0
- IO"
2 5 4
, 0-3.55 2

- -
N 2 H (C10 )
6 4 2
,0-1.65 1<T 3

15.5 Hydroxylammonium Sulphate and Phosphate


For all hydroxylammonium salts studied, their thermal decomposition was
found to proceed through a preliminary dissociation into a hydroxylamine
and acid. The stability o f ions forming the salt is rather high. A distinctive
feature o f hydroxylammonium salts, in contrast to the other onium salts, is a
high rate o f decomposition o f the base. Decomposition o f hydroxylamine
and its solutions usually proceeds at a noticeable rate even at room
temperature. Similar to hydrazine, hydroxylamine can manifest itself as an
oxidizer and reducing agent. One can expect a variety o f reactions during
the decomposition o f its salts. A t the same time, it was revealed that the rate
of decomposition o f hydroxylamine in the solutions is very poorly
reproduced and depends on various factors, in particular, on the possibility
o f contact with the oxidizing atmosphere and on the content o f ions o f the
metals o f alternative valence even in very small amounts. B y the strength o f
the catalytic influence, the cations can be arranged in the following series
[15.12]: F e + 2
> Ni + 2
> Cu + 2
> Co + 2
> Ag + 1
> Z n . In all studies,
+ 2
the
observable rate o f decomposition o f hydroxylamine was determined by the
interaction with the other components present in the system or by the
catalytic influence o f metal ions. Attempts to attain the rate corresponding
to the thermal transformations o f hydroxylamine molecules has not met
with success.
It is advisable to begin with consideration o f the salts formed by stable
acids, in which only hydroxylamine transformations are possible. First o f
all, these are sulphate (NH 0) S0 4 2 4 and phosphate (NH 0) P0
4 3 4 of
hydroxylammonium. For these salts, the kinetics o f decomposition was
studied below their melting temperatures [15.13]. The kinetic regularities
are complicated by the dissolving o f the salt in the decomposition products
Thermal Decomposition and Combustion 239

in the course o f the process. The rate increases up to the moment o f the
complete dissolution o f the salt. After that, the rate is described by the
kinetic equation o f the second order and the temperature dependence o f the
rate constants takes the form:

( N H 0 ) S 0 , r =170-172°C:
4 2 4 melt k = 10 17 2
exp(-41600//?r), s~ ]

(NH 0) P0 , r
4 3 4 melt =160°C: k = IO exp(-30200//?r),
127
s" . 1

The rate o f thermal decomposition o f these salts is determined by


decomposition o f either NH OH molecules or their complexes with the 2

parent salt, and F e + 2


and, sometimes, C u + 2
ions, present in the form o f an
impurity in the parent salt, serve as the catalysts o f this process. The rate o f
decomposition o f ( N H 0 ) S 0 4 2 4 at 125°C depends on the concentrations
C F e + 2 and C C u + 2 (in m o l %) in the f o l l o w i n g way:

(drj/dt) t=0 = [1.44 • 1 (T (C 4


Fe+2 ) ' + 1.27-1 (T (C
0 87 4
Cu
+2
) ' ],
0 5 5
s" . 1

Usually, the content o f F e + 2


ions is considerably higher than that o f C u ; + 2

for instance, in the sample studied C F e + 2 = 3 • 10~ mol % , c 3


C u + 2
=
2 • 1(T 5

mol % . Thus, the influence o f F e + 2


is the determining one. A t 130°C, for
( N H 0 ) P 0 we w i l l have:
4 3 4

(dn/dt) (=0 =9-10" C 2


Fe+2 , s . - 1

In the parent salt, C F e + 2 = 2.8 • 1(T mol % and the influence o f C u


3 + 2
was
not analyzed. The attempt to study the rates o f decomposition under the
lower content o f impurities has not been successful. Repeated
recrystallization from distilled water does not reduce the content o f these
ions in the salt.
The determining role o f the reactions o f hydroxylamine during the
decomposition o f these salts was confirmed by the decrease o f the rate o f
decomposition on addition o f an excessive acid into the salt and by its
increase on introduction o f N H or S 0 ~ ions, w h i c h increased the 3 4
2

equilibrium concentration o f hydroxylamine. The decomposition according


to the equation o f the second order is indicative o f the bimolecular
decomposition o f hydroxylamine, most probably, by the following reaction:

2NH OH = NH NHOH + H 0
2 2 2

with the rapid decomposition o f hydroxylhydrazine to the final products.


The composition o f the products o f decomposition varies in the course o f
240 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

the process. A t the initial stage, for sulphate, the equation o f decomposition
takes the form:

(NH 0) S0 4 2 4 = 0.67NH HSO +0.33H OHSO + 1.67H O+0.67N .


4 4 3 4 2 2

The heat o f decomposition is near 390 cal/g or 64 kcal/mol. A t higher


degrees o f transformation, the yield o f N H 4
+
increases and N 0 appears, 2

with a corresponding decrease o f the amounts o f H 0 and N . The


2 2

decomposition products o f phosphate have not been studied, and the heat o f
decomposition o f 590 cal/g or 117 kcal/mol corresponds to the higher
content o f hydroxylamine in the salt.

15.6 Hydroxylammonium Chloride


The main kinetic regularities o f thermal decomposition o f chloride o f
hydroxylammonium are close to those for sulphate and phosphate
considered above, but the capability o f H C l to participate in reduction-
oxidation reactions makes this process more difficult. The kinetics was
studied within the temperature range o f 1 4 0 - 1 7 5 ° C at r = 152°C [15.14]. m d t

In the liquid phase, the reaction proceeds according to the order, w h i c h is


close to the first one, and the temperature dependence o f the rate constant
takes the form:

£ = 10 l34
exp(-34000/#r), s" .
1

The stoichiometry o f the reaction is expressed by the following equation:

N H O C l = 0.45NH Cl + 0.81H O + 0.18N O+0.55HCl+0.09N +0.01NO.


4 4 2 2 2

The heat o f decomposition is 312 cal/g or 21.7 kcal/mol. The content o f


Fe + 2
ions in the parent salt is near 1 • 10" m o l % , and their noticeable
3

influence starts from the value o f (5-10) • 10" m o l % . Introduction o f 3

excessive ammonia or hydrazine, whose basic properties are higher and


w h i c h almost quantitatively displace hydroxylamine in its salts, results in a
considerable increase o f the decomposition rate. Thus, introduction o f 33.5
mol % o f N H 2 4 increases the initial rate in the liquid phase by a factor o f 72.
Under such conditions, N H C 1 virtually does not decompose. A t the same
2 5

time, introduction o f excessive H C l results in a very weak decrease o f the


observable rate o f decomposition and, at the transition in the aqueous
solution, the rate even increases slightly with the increase o f C c i in this H

solution. The change o f the reaction order from the second to the first is also
indicative o f the appearance o f an additional decomposition channel, whose
contribution increases with the increase o f acidity. Obviously, this is caused
Thermal D e c o m p o s i t i o n and Combustion 241

by the interaction o f HC1 with the parent salt. M o r e efficient oxidizing


properties o f hydroxylamine in comparison with hydrazine make this
contribution considerable:

NH4OCI + HCI = N H + H 0 +C1 .


3 2 2

In [15.15], liberation o f chlorine during the decomposition o f N H O C l in an 4

open system is experimentally confirmed, and under conditions o f a closed


system, it oxidizes hydroxylamine, regenerating HC1:

2NH 0+ C l
3 2 = 2H 0+ N 2 2 + 2HC1.

Therefore, the rate o f decomposition o f N H O C l can be represented in the 4

form:

W = k\ C N H 2 Q H +k CNH OCI^HCI
2 4

but the values o f k\ and k were not obtained.


2

15.7 Hydroxylammonium Perchlorate and Its


Hydroxylamine Complex
Hydroxylammonium perchlorate ( H A P ) is an efficient oxidizer, but its
commercial use is considerably hampered by its high hygroscopicity. In this
case, together with hydroxylamine, perchloric acid is also capable o f its
own thermal transformations. Its reactions in other perchlorates have been
considered above. The kinetics o f decomposition o f hydroxylammonium
perchlorate was studied in the liquid phase at temperatures o f 1 1 3 - 1 5 0 ° C
[15.16]. A t the beginning o f the process, the decomposition rate is constant
or even slightly decreases, but at 77 = 0.01-0.02, it starts to grow and this
growth continues up to 77 = 0.7-0.8. In this region, the rate can be described
by the following equation:

drj I dt - k + k n. x 2

The initial decomposition rate increases with the increase o f the value o f
V0
6
and within the range K = (4-10) • 10 c m its temperature dependence
0
8 3 3

takes the form:


k = 1 0 - e x p ( - 2 7 5 0 0 / / ? r ) , s" .
x
8 8 1

The value o f k does not depend on K


2 0
g
and can be written in the following
form:
242 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

k = 10
2
160
exp(-37900//?7), s' . 1

The composition o f the decomposition products corresponds to the


following equation:

NH4OCIO4 = O.O4NH4CIO4+ 0 . 6 H C l O + 1 . 6 2 H 0 + 0.48N O+0.18C1 + 0.17N .


4 2 2 2 2

O n introduction o f N H , w h i c h results in the growth o f the hydroxylamine


3

concentration, the constant k increases, and this confirms the determining


x

role o f N H O H decomposition at this stage. The influence o f excessive


2

H C I O 4 is more complicated. For small additives o f H C 1 0 4 (up to 0.1%), the


initial rate decreases, but as the amount o f H C 1 0 increases, the rate starts to
4

grow: HCIO4 in the amount o f 3 mol % increases the rate o f decomposition


by a factor o f 4.
The influence o f both additives considerably increases as the value o f
F 0
g
decreases. The rate o f decomposition for a pure H A P in the course o f
decomposition changes similarly, which is connected with accumulation o f
perchloric acid. Obviously, the initial rate o f decomposition o f a pure salt is
determined by decomposition o f hydroxyl amine, and its lower value, in
comparison with sulphate and phosphate, is determined by the higher acid
properties o f H C 1 0 4 and the lower concentration o f N H O H in H A P . The 2

influence o f metal ions on the decomposition o f H A P was observed but has


not been studied in detail. In contrast to sulphate and phosphate, perchloric
acid and the products o f its decomposition oxidize hydroxylamine, and less
than 1 mole o f H C 1 0 is consumed per 1 mole o f N H O H . Thus, the content
4 2

o f HCIO4 increases and the content o f N H O H decreases in the course o f


2

the process. The rates o f decomposition o f these products are also changed.
Thus, the rate o f decomposition o f perchloric acid is rapidly equalized and
then exceeds the rate o f decomposition o f hydroxylamine. Exactly this
reaction is the determining one along the major portion o f the kinetic curve.
The acid is accumulated during the whole process, and this is one o f the
main final products, while the fraction of N H 3 and, correspondingly,
NH4CIO4 in the products is small. This confirms a high degree o f oxidation
o f the hydroxylamine. This mechanism also adequately explains the
absence o f the (1—77) term in the kinetic equation connected with the
consumption o f the parent product. The acid accumulated during nearly the
whole reaction course leads the reaction. Obviously, at the limiting stage o f
acid decomposition, hydroxylammonium ions are involved, because, during
the interaction with the molecules o f hydroxylamine, the decomposition rate
could not increase due to the constancy o f C N H O H ^ H C 1 0 > ov/'mg to the 2 4

equilibrium o f the salt dissociation. O n completion o f the decomposition,


the aqueous solution o f the composition o f H C 1 0 • 2 . 7 H 0 is formed. A t a
4 2

temperature near 150°C, even the dihydrate o f perchloric acid has a very
low rate o f decomposition. This rate increases under the influence o f
Thermal Decomposition and Combustion 243

hydroxyl ammonium ions. The reaction halts after their consumption. The
monohydroxyl amminate o f hydroxyl ammonium perchlorate N H O C 1 0 • 4 4

N H O H (T^eit = 9 8 ° C ) decomposed in the liquid phase at the initial rates,


2

which are higher by approximately a factor o f 1000 than in the case o f


hydroxylammonium perchlorates, which corresponds to a larger amount o f
N H O H molecules in this melt owing to the equilibrium:
2

NH OC10 - N H O H ^
4 4 2 NH OC10 + NH OH.
4 4 2

The kinetics o f decomposition is described by the equation o f autocatalysis


o f the first order. Within the range o f K = (15-20) • 10 , the rate constants
0
g 3

take the form:

k =10
x
, 8 7
exp(-39700/^r), s" 1

k =10
2
112
exp(-25700/#r), s" . 1

The stoichiometric equation o f decomposition o f N H O C 1 0 • N H O H takes 4 4 2

the form:

NH OC10 NH OH
4 4 2 = N H C 1 0 + 1 . 5 H 0 + 0.5N O.
4 4 2 2

The heat o f decomposition is near 400 cal/g or 66 kcal/mol. It is obvious


that all the time the reaction proceeds through thermal transformations o f
hydroxylamine, and C10 ~ ions are preserved. In the absence o f oxidation
4

by perchloric acid, a considerable amount o f N H is formed, which 3

displaces hydroxylamine from the salt, maintaining its high concentration in


the melt which, in turn, inhibits the dissociation o f H A P with the formation
o f a free perchloric acid. A t the introduction o f water into the melt, the rate
o f decomposition increases, apparently, owing to the displacement o f
N H O H molecules from the complex by water molecules.
2

15.8 Hydroxylammonium Nitrate

H y d r o x y l ammonium nitrate ( H A N ) N H O N 0 4 3 (T =mdt 4 3 ^ 4 ° C ) occupies a


special position, because its rate o f thermal decomposition is considerably
higher than that for other salts. The reaction is studied within the
temperature range of 85-121°C, and proceeds with a sharp self-
acceleration. It can be described by the equation o f autocatalysis o f the
second order up to n = 0.6:

dr//'dt = k (\ - 7 7 ) + k r/ (\ - 7 7 ) .
x 2
2
244 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

For higher degrees o f transformation, self-acceleration drastically increases,


and the m a x i m a l rate is higher than the initial one by a factor o f 2 0 0 - 5 0 0
[15.17, 15.19]. A t F 0
8
= 4- 10 c m , we w i l l have:
4 3

k = 10 exp(-16500/#r),
x
4 6
s~\

but the values o f k are not presented. The rate increases at introduction o f
2

either N H or H N 0 3 3 into the parent mixture, and the increase caused by


nitric acid is considerably higher. Introduction o f several m o l % o f H N 0 3

results in a violent decomposition o f the salt even at room temperature, and


introduction o f N H , increasing the initial rate, also increases the time up to
3

the occurrence o f sharp acceleration. The products o f decomposition have


not been quantitatively studied but, in contrast to the other salts o f
hydroxylammonium, considerable amounts o f N O and N 0 2 are separated.
Based on the experimental data, the following mechanism o f decomposition
was proposed, which incorporates two parallel reactions:

(1) Interaction o f the N H 0 4


+
ion with the N 0 2 5 molecule, which is formed
from H N 0 3 in equilibrium:

NH 0 +N 0 4
+
2 5 = HN0 +HNO+H N0
3 2 2
+
.

with the subsequent rapid thermal transformations o f these products.

(2) B i m o l e c u l a r decomposition o f hydroxylamine as in the case o f the other


salts o f hydroxylammonium.

This is the main pathway i f N H O H is in excess, because the equilibrium 2

concentration o f H N 0 and, especially N 0 , sharply reduces. O n l y during


3 2 5

the consumption o f hydroxylamine do the reactions through N 0 appear, 2 5

but they proceed at a considerably higher rate. The low thermal stability is
an essential difficulty for any commercial applications o f H A N .

Table 15.8.1. The initial rates of thermal decomposition of hydroxylammonium


salts

Salt kx=f(T). s- 1
kl9 s~ ]

80°C 120°C 150°C

io-
-6.0
(NH 0) P0 4 3 4 10 ,27
exp(-30200/RT) 10 41
IO" 29

(NH 0) S0 4 2 4 10 ,72
exp(-41600/RT) IO" 85
IO" 59
IO" 43

NH OCl
4 10 ,34
exp(-34000/RT) j -6
0 6 io- 45
io-
41

NH OC10
4 4 108 8
exp(-27500/RT) io- 82
IO" 65
io-
54

NH ON0
4 3 10 46
exp(-16400/RT) io- 55
IO" 45
io-
39
Thermal Decomposition and Combustion 245

The initial rates o f decomposition o f hydroxylammonium salts studied are


presented in Table 15.8.1. One can see from their comparison that the initial
rate o f decomposition o f nitrate is close to this value for hydroxyl
ammonium phosphate and at elevated temperatures is even lower but, in the
case o f nitrate, the rate growth is proportional to r/ 2
and, in the case o f
phosphate, it rapidly decreased proportionally to {\-rjf and, thus, for the
major part o f the process, the rate o f decomposition o f nitrate is
considerably higher than in the case o f the other salts. O n the whole, the
differences in the values o f the initial rates for hydroxylammonium salts are
smaller than those for hydrazinium salts.

15.9 Nitronium Perchlorate

Nitronium perchlorate (NP) N 0 C 1 0 2 4 was frequently considered as a


possible high-energy component o f rocket propellants, especially in hybrid
systems. This is a good reason to describe its thermal decomposition
together with the hydrazinium and hydroxylammonium perchlorates,
though, N P is not the onium salt and, in this case, decomposition through
the equilibrium proton transfer is impossible.
W e studied the kinetics o f thermal decomposition o f N P with the help
o f the rate o f heat release within a temperature range o f 7 5 - 1 1 9 ° C . Inside a
completely temperature-controlled ampoule, the reaction proceeds with heat
release, and the heat o f decomposition is near 50 cal/g or 7.3 kcal/mol. The
main reaction products are C l , N , 0 2 2 2 and N 0 . The enthalpy o f formation
2

o f N P is 9.3 ± 0.2 kcal/mol. During the decomposition o f the salt into the
elements, this amount o f heat should be released, and reduction o f the actual
heat release is connected with the formation o f the corresponding amount o f
N 0 . Chlorine oxides are unstable at the temperatures used and cannot be
2

preserved in the calorimetric ampoule, but in the presence o f cool walls in a


vessel or in an open system, they can be preserved. In [15.18], under such
conditions, decomposition with heat absorption was observed.
The kinetics o f heat release is rather complicated. A t the beginning o f
the process, a relatively low and almost constant rate is observed, and the
corresponding rate constant at the heat Q = 50 cal/g takes the form:

£ = 10
0
, 2 3
exp(-30200//?r), s" . 1

W i t h time, the rate starts to increases rapidly, and, then a decrease o f the
rate o f heat release is observed. Then, the growth again proceeds with the
attainment o f the previous level and, at last, the normal decay o f the rate
takes place, which is determined by the consumption o f N P . A b o v e the
temperature o f 9 0 ° C , the stage o f acceleration up to the beginning o f the
246 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

rate decay is described by the equation o f autocatalysis o f the first order.


The rate constants take the form:

k = 10
x
33 7
exp(-69000//?r), s - 1

k =10 exp(-21000//?r),
2
91
s" . 1

Comparison o f the values o f activation energies for k\ and k shows that 2

with temperature growth the self-acceleration o f the process rapidly


degenerates. The value o f k\ is only o f formal importance, determining the
time o f the beginning o f self-acceleration. The actual initial rate is
determined by the value o f k . B e l o w 9 0 ° C , the time up to the stepwise
0

acceleration o f the reaction increases by a factor o f 3. The temperature


dependence o f this time takes the form:

r = 10" 1 3 7
exp(26000/^r), s.

The kinetic curve below 9 0 ° C is even more complex and is not described by
the equation o f autocatalysis. The inclination o f the curve o f the rate o f heat
release above 9 0 ° C and its complex character under the lower temperatures
are apparently connected with the variation o f N 0 yields in the course o f 2

the process. The enthalpy o f formation o f N 0 in the gas phase is 8.1 2

kcal/mol, and at the formation o f 1 mole o f N 0 from 1 mole o f N 0 C 1 0 2 2 4

the heat o f reaction is near 1 kcal/mol.


Extrapolation o f the r value to 2 0 ° C gives a value o f 12-13 years.
Before this time, the decomposition w i l l proceed without acceleration.
Earlier, in [15.18], the decomposition was studied thermographically and
with the help o f pressure growth measurements. In vacuum, a considerable
part o f the salt was sublimated. There are no complete quantitative data
available, and the activation energy for self-acceleration under different
conditions is 15-25 kcal/mol; the initial rates have not been analyzed.
In the analysis o f the rate and mechanism o f decomposition o f N P , one
should take into account that this salt actually represents a mixed anhydride
o f nitric and perchloric acids and it should be compared with anhydrides o f
these acids and with N 0 and C 1 0 molecules. Their decomposition was
2 5 2 7

discussed in consideration o f thermal decomposition o f acids. Their


decomposition proceeds monomolecularly, and at 7 0 ° C , in the case o f
N 0 C 1 0 ko = 1.2 • IO" s" ; at the same time, for N 0 k = 8 • 1 0 s" , and
2 4
7 1
2 5 0
-3 1

for C 1 0 ko = 1.9 • 10~ . The rate o f decomposition o f N P is smaller by a


2 7
4

factor o f 10 —10 . The decomposition o f anhydrides proceeds through the


3 4

breaking o f relatively weak covalent bonds 0 N — O — N 0 and 0 C 1 — O — 2 2 3

C 1 0 . O w i n g to the symmetry, both bonds in the molecule are equal. The


3

main reason for higher thermal stability o f N 0 C 1 0 is the ionic structure o f


2 4
Thermal Decomposition and Combustion 247

its molecule. In the structure o f 0 N — O — C 1 0 2 3 the electrons o f the N — O


bond are almost completely shifted towards the oxygen atom, forming the
structure N 0 C 1 0 ~ , in which there are no weak bonds at all. A t the same
2
+
4

time, the following equilibrium should occur:

N0 C10 ~ - 0 N—O—C10 .
2
+
4 2 3

w h i c h is noticeably shifted to the left, towards the ionic structure. The


fraction o f covalent molecules is insignificant. The bond energies o f N —
O — C I should be close to similar bonds in N 0 2 5 and C 1 0 2 7 and, thus, the
decomposition has to proceed through the molecular form. The existence o f
such an equilibrium is confirmed by a noticeable volatility o f the salt: it can
be transferred into the gas phase only in the form o f molecules.
The kinetic scheme o f the first stage o f decomposition o f N P takes the
form:

+ Kc N0 + C10
2 4
N0 C10 ^: 0 N—O—CIO3
2
+
4
-
2

^ N0 +C10
3 3>

The initial rate o f decomposition per 1 mole o f N 0 C 1 0 2 4 takes the


following form:
W° = k K +k K l c 2 c ,

From the comparison o f the rate constants o f decomposition o f N 0 2 5 and


C 1 0 , one can conclude that the O — N bond is opened at a higher rate than
2 7

the O — C I bond, and k\ > k . The increase o f thermal stability o f N P is


2

determined by the value K c « 1. A similar phenomenon is observed on


dissolving N 0 2 5 in 100% H N 0 . 3

O w i n g to a high dielectric permittivity o f nitric acid, N 0 2 5 in solution is


mainly ionized into N 0 2
+
and N 0 " and the rate o f decomposition is
3

reduced by a factor o f 20, in comparison with the solutions in low-polarity


solvents. The mechanism o f the initial stage o f decomposition through the
oxygen detachment by the following reaction:

N0 C10 2 4 = N O C l O + 0.5O , 4 2

proposed in [15.18], is hardly possible with the observable rate constants


due to a high bond energy in the N 0 2
+
ion.
The radicals formed during decomposition of the salt rapidly
decompose and interact with each other by the reactions which have already
248 The Salts of Hydrazinium, Hydroxylammonium, and Nitronium

been considered in Chapter 11 for perchloric and nitric acids. In the


presence o f N 0 , the following additional reactions are possible:
2

N0 +C10
2 2 = N0 Cl+0 2 2

NO2+CIO = NOC1 + 0 2

2N0 C1 2 = N +Cl + 20
2 2 2

2NOC1 = N + Cl + 0 .
2 2 2

The causes o f self-acceleration o f the reaction have not been studied. The
most probable reason is the interaction o f N 0 2 or an active intermediate
product with the parent salt, for instance:

N0 2 +N0 C10
2 4 = N 0 2 4 + C10 4

N 0 2 4 = 2N0 . 2

N P rapidly interacts (frequently with self-ignition) at room temperature with


almost a l l organic substances, and the production o f composites on its basis
is virtually impossible.

REFERENCES

15.1. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 46(3): 6 2 7 - 6 3 0


(1972).
15.2. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 44(2): 3 9 6 ^ 0 0
(1970) .
15.3. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 16(3): 814-817
(1971) .
15.4. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 43(11): 2 9 7 2 -
2973 (1969).
15.5. K . S. M o h a m e d and D . K . Padma, J. Indian Institute of Science,
69(2): 125-129(1989).
15.6. K . C . Patil, I. P . Vittal, and C . C . Patel, Journal of Thermal
Analysis, 26(2): 191-198 (1983).
15.7. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 45(12): 3 0 4 2 -
3044(1971).
15.8. Y u . I. Rubtsov and G . B . Manelis, Zh. Fiz. Khim., 45(10): 2 5 6 9 -
2572(1971).
15.9. Y u . I. Rubtsov and G . B . Manelis, Izv. Akad. Nauk SSSR, Ser.
Khim., 2: 296-299 (1984).
15.10. V . A . K o r o b a n , T . I. Smirnova, and B . S. Svetlov, Chemical
Physics of Combustion and Explosion. Kinetics of Chemical
Reactions, Chernogolovka: 67-71 (1977).
Thermal Decomposition and Combustion 249

15.11. Y u . I. Rubtsov, Combustion and Explosion. Proceedings of the


Third All-Union Symposium on Combustion and Explosion, M.:
N a u k a , 4 9 - 5 2 (1972).
15.12. S. Lunak and I. Veprek-Siska, Collection of Czechoslovak
Chemical Communication, 39(2): 391-395 (1974).
15.13. V . A . Rafeev, Kinetics and Mechanism of Physical-Chemical
Reactions, Chernogolovka: Joint Institute o f C h e m i c a l Physics, 5:
52-53 (1982).
15.14. V . A . Rafeev, A . I. K a z a k o v , and Y u . I. Rubtsov, Zh. Fiz. Khim.,
53(3): 781-782 (1979), A v a i l a b l e from V I N I T I N 2759-78, (1978).
15.15. E . M . Marshak, Zh. Neorg. Khim., 13(4): 1212 (1968).
15.16. V . A . Rafeev, Y u . I. Rubtsov, E . S. Rumyantsev, L . P. A n d r i e n k o ,
G . I. L u b i m o v a , and G . B . Manelis, Izv. Akad. Nauk SSSR, Ser.
Khim.,1: 1463-1469 (1980).
15.17. V . A . Rafeev and Y u . I. Rubtsov, Izv. Rus. Akad. Nauk, Ser. Khim.,
11: 1897-1901 (1993).
15.18. J. N . M a y c o c k and V . R. Pai Verneker, J. Phys. Chem., 72(12):
4004-4009 (1968).
15.19. G . B . Manelis, AGARD Conference Proceedings, Neuilly-Sur
Seine: France, 29 A u g u s t - 2 September, 559 (1994).
Chapter 16

METAL PERCHLORATES AND


NITRATES, METAL SALTS OF
DINITRAMIDE

In contrast to onium salts considered above, the equilibrium formation o f


acids and bases is impossible for metal salts. In this case, all o f the
pathways of thermal decomposition are determined by thermal
transformations o f anions. O n l y at the decomposition o f crystal hydrates is
the hydrolysis o f a salt with separation o f oxide or hydroxide o f metal and
corresponding acid possible. This is observed for some heavy metal
perchlorates and nitrates, but for the majority o f salts such a pathway is not
realized. Crystal water is usually lost on heating long before the beginning
o f salt decomposition. Thus, the kinetic regularities o f decomposition o f
metal perchlorates and nitrates differed considerably from the regularities o f
decomposition o f onium salts and they are determined by other factors. The
detailed quantitative studies o f the kinetic regularities o f decomposition o f
these salts are almost absent. The studies were carried out mainly by the
D T A method without analysis o f the kinetic law and rate constants. Here,
we w i l l not consider the details o f these works. The detailed review o f metal
perchlorates is given in [16.1]. W e w i l l consider in a generalized form only
general regularities, which allow one to analyze the peculiarities o f thermal
decomposition and combustion o f mixtures containing such compounds.

16.1 Metal Perchlorates


Analysis o f the literature data available on decomposition o f perchlorates
shows that there are several possible pathways o f thermal transformations
o f the C 1 0 ~ anion and several factors determining the rate o f this process
4

and the composition o f the decomposition products. Consequently, the rate


o f decomposition can vary for different perchlorates within very wide
limits. First, let us consider the main chemical reactions during the
decomposition o f metal perchlorates. In this case, the following two ways
are possible:

MeC10 4 - MeCl + 2 0 2
(16.1.1)

2MeC10 4 = M e 0 + 3.50 + C l .
2 2 2
(16.1.2)
252 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

For each perchlorate, the composition o f the final products is determined by


the thermodynamic efficiency o f the formation o f chloride or oxide o f a
corresponding metal, but because the initial state is unique, the values o f the
enthalpies o f formation o f chloride and oxide related to 1 gram-atom o f the
metal are o f decisive importance, since the entropy variation during
decomposition, w h i c h is slightly higher for the second pathway, only
weakly changes for various metals. A comparison o f these values in the
standard state for various metals [16.2] is presented in Table 16.1.1.

Table 16.1.1. Comparison of the standard enthalpies of formation of metal oxides


and chlorides related to 1 g-atom of metal

Chloride - A / / " , Oxide -AH{\ Chloride -AH{\ Oxide -AH°,


kcal/ mol kcal/ mol kcal/ mol kcal/ mol

LiCl 97.6 Li 0 2 71.5 CoCl 2 74.1 Co 0 3 4 69.6


NaCl 98.3 Na 0 2 49.6 NiCl 2 72.7 NiO 57.3
KC1 104.3 K 0
2 43.3 A1C1 3
168.3 A1 0 2 3 200.2
RbCl 104.0 Rb 0 2 40.5 GaCl 3 125.4 Ga 0 2 3 130.2
CsCl 105.7 Cs 0 2 41.3 InCl 3 128.4 ln 0
2 3 110.7
BeCl 2 118.6 BeO 145.6 TICI3 74.4 T1 0 2 3 46.6
MgCl 2 154.1 MgO 143.8 MnCl 2 115.0 MnO 92.0
MgO 156.5 Mn0 2 123.5
MgCl 2

CaCl 2 190.2 CaO 151.8 CrCl 3 136.3 Cr 0 2 3 135.6


BaCl 2 201.7 BaO 131.0 Cr0 2 141.1
ZnCl 2 99.2 ZnO 83.8 TiCl 4 195.6 Ti0 2 225.6
CdCl 2 93.4 CdO 61.9 ZrCl 4
234.2 Zr0 2 263.0
CuCl 51.2 CuO 38.7 HfCl 4 236.7 HfD 2 267.1
AgCl 30.4 Ag 0 2 3.6 NbCl 5
190.6 Nb 0 2 5
226.8
FeCl 3
95.5 Fe 0 2 3
98.3 TaCl 5
205.1 Ta 0 2 5
244.7

O n the one hand, one can see that for the majority o f metals, the formation
o f chloride is more efficient, and thermal decomposition quantitatively
proceeds according to equation (16.1.1). Even trace amounts o f chlorine
cannot be detected in the products o f decomposition o f sodium, potassium,
rubidium, and cesium perchlorates. The most stable perchlorates belong to
this group. O n the other hand, for B e , A l , G a , N b , T a , T i , Z n , and H f oxide
formation is more efficient, and decomposition proceeds by equation
(16.1.2). The perchlorates o f this group are most unstable. L i t h i u m and
magnesium perchlorates occupy an intermediate position, and, in this case,
the difference in favor o f chloride is insufficient and is partially
compensated by a higher entropy increase in reaction (16.1.2). Here, along
with chloride, noticeable amounts o f oxide are formed. One interesting
peculiarity is observed during the decomposition o f magnesium perchlorate
[16.3] where, at the initial stage, only M g C l 2 is formed, but then M g C l 2
Thermal Decomposition and Combustion 253

interacts with the parent perchlorate with the formation o f the M g O M g C l 2

compound, which is more thermodynamically efficient than chloride or


oxide taken separately. A s a result, near 3 0 % o f chlorine is formed. B y their
thermal stability, these perchlorates occupy an intermediate position. For
manganese and chromium perchlorates, on retention o f a cation valence the
formation o f chloride is more efficient, and owing to further oxidation o f
cation during decomposition, the formation o f oxide becomes more
efficient.
Thermal decomposition o f metal perchlorates with the formation o f
chlorides proceeds mainly according to the following two mechanisms:
(1) . Detachment o f the oxygen atom from the C 1 0 ~ anion: 4

C10 " =
4 C10 " + O .
3

In this anion, the bond energy C I — O is relatively high (61-69 kcal/mol for
various salts) and quite high temperatures are necessary to proceed with the
decomposition at a noticeable rate. Such a pathway is probably essential
and may even be the main one during decomposition o f K , R b , C s , and
partly N a perchlorates. These perchlorates decompose at a sufficient rate
only at temperatures above 5 0 0 ° C , and their activation energies are rather
close to the C I — O bond energy. The chlorates formed possess a higher
decomposition rate, decomposing into chloride and oxygen, and, probably,
are partly subjected to disproportionation with the formation o f chloride o f
perchlorate.
In all cases, the rate o f decomposition o f perchlorate is determined by
the rate o f oxygen atom detachment from the C 1 0 ~ anion. 4

(2) . Interaction o f two C 1 0 " anions with separation


4 o f the oxygen
molecule:

2C10 " = 4 2C10 "+0 . 3 2

In this reaction, the enthalpy variation is 14-16 kcal/mol, but the interaction
o f two likely charged particles is connected with a high kinetic barrier. Such
interaction is considerably facilitated by polarization o f the anion and by the
decrease o f its effective charge that takes place for cations o f small radius
and, especially, for multiple charged cations. The polarizing ability is
determined by the ratio o f the charge to cation radius. Perchlorates o f L i ,
M g , C a , B a , C d , Z n , C u , and M n belong to this group. The activation
energies o f their decomposition fall within the range o f 4 0 - 6 0 kcal/mol, and
decomposition is usually observed within the temperature range o f 2 0 0 -
5 0 0 ° C . In [4], it was revealed that the very beginning o f the initial stages (77
< 0.04) o f decomposition o f alkali metal perchlorates proceeds with lower
activation energy. For this stage, it was postulated that decomposition
proceeds according to reaction (16.1.2).
254 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

For perchlorates, during the decomposition o f which the formation o f


oxides is more efficient, interaction o f two C 1 0 ~ anions proceeds through 4

the other pathway with the formation of C 1 0 2 7 molecules and the


subsequent thermal decomposition 2C10 ~ = C 1 0 4 2 7 + O " . Unfortunately,
2

there are no kinetic studies o f this reaction available, but for niobium [16.5]
and tantalum [6] perchlorates, separation o f C 1 0 2 7 was revealed by the
direct chemical analysis. Decomposition already proceeds at a considerable
rate at 7 0 - 1 2 0 ° C almost quantitatively according to the following equation:

Nb(Ta)(C10 ) 4 5 = Nb(Ta)0(C10 ) 4 3 + C1 0 . 2 7

The next stage o f decomposition requires a temperature o f 1 2 0 - 1 4 0 ° C , and


at such temperatures C 1 0 rapidly decomposes into chlorine and oxygen:
2 7

Nb(Ta)0(C10 ) 4 3 = Nb(Ta)0 C10 + C l + 3.50 .


2 4 2 2

Thermal decomposition o f the latter anion C 1 0 ~ proceeds only at 2 6 0 - 4

290°C:

2Nb(Ta)0 C10 2 4 = Nb (Ta )0 + C l + 3.50 .


2 2 5 2 2

For niobium perchlorate, all stages proceed at temperatures w h i c h are lower


than for tantalum perchlorate by 2 0 - 3 0 ° C in accordance with a smaller
radius o f cation. These reactions clearly demonstrate the increase of
perchlorate stability with the decrease o f the polarizing ability o f the cation
and the number o f anions coordinated on it. The parent niobium and
tantalum perchlorates possess a considerably higher decomposition rate
than, for instance, ammonium perchlorate.

Table 16.1.2. The melting and decomposition temperatures of metal perchlorates

Salt LiC10 4 NaC10 4 KCIO4 CsC10 4 Mg(C10 ) 4 2 Ba(C10 ) 4 2


Pb(C10 ) 4 2

r c
d 410 508 547 526 393 440 560
TC
m 248 468 580 577 600 470 280

Salt AgC10 4 Cd(C10 ) 4 2 Zn(CI0 ) 4 2 AI(CI0 )3 4


Nb(CI0 ) 4 5 Ta(C10 ) 4 5

r c
d 420 393 298 240 70 90
TC 476
m
- - - - -
There are some other factors, which influence the rate and kinetic
regularities o f decomposition o f particular perchlorates. For instance, an
increase in a cation's own polarizability results in a somewhat higher rate o f
decomposition o f C s C 1 0 in comparison with K C 1 0 . The melting
4 4
Thermal Decomposition and Combustion 255

temperatures o f perchlorates differ considerably. For particular salts, the


reaction starts in the solid phase and is accompanied by a progressive
melting o f substance that results in self-acceleration o f the reaction. For the
other salts, from the beginning the reaction proceeds in the liquid phase and,
thus, possesses a higher initial rate. Self-acceleration occurs in the liquid
phase, too, o w i n g to the catalytic influence o f C P ions. These circumstances
make it difficult to analysis the causes o f the difference in the rates o f
decomposition o f various perchlorates. The temperatures o f decomposition
Tj, corresponding to the temperatures at which the k value in equation 1-(1-
rj)1/3
= k- t attains the value o f 5 • 1(T , s~ , are presented in Table 16.1.2
3 !

based on the data o f [ 16.1 ].

16.2 Metal Nitrates


Thermal decomposition o f metal nitrates is frequently used in industry and
preparatory chemistry to obtain pure oxides. For the majority o f nitrates, the
thermal stability is rather high. The factors determining the rate o f thermal
decomposition o f a particular metal nitrate are almost the same as
considered above for the corresponding perchlorates. Primarily, this is the
polarizing force o f a cation, the type o f the cation bond with the nitrate
group and the phase state o f the nitrate [16.7, 16.8]. Almost all studies o f
decomposition o f metal nitrates were carried out by the D T A and D G A
methods, and the analysis o f the rates was conducted on the level o f "the
decomposition temperatures". Decomposition o f the majority o f nitrates
proceeds with the formation o f corresponding oxides. For the alkali metal
nitrates, the nitrites can be evolved as intermediate compounds. Nitrates o f
noble metals decompose with separation o f a free metal. The formation o f
metal nitrides is thermodynamically inefficient. Certain difficulties are
connected only with a possible variation at the decomposition o f the valence
state o f the metal due to its further oxidation by the decomposition products.
Nitrates o f alkali metals possess the highest stability. They represent the
pure ionic salts and decompose in the liquid phase. The decomposition rate
decreases when passing from L i to Cs in accordance with the reduction o f
the polarizing ability o f the cation. The absolute values o f the rates are
comparable with the rates o f decomposition o f corresponding perchlorates.
For the alkali-earth metal nitrates, the regularities are preserved and the
temperatures o f decomposition for C a , Sr, and B a nitrates are even slightly
higher than in the case o f alkali metals. In general, this is determined by a
higher melting temperature o f the salt. Decomposition starts immediately
after the melting o f the nitrate. Thus, at thermal decomposition of
Ca(N0 ) , 3 2 the activation energy is near 55 kcal/mol [16.9], and
decomposition proceeds through the pathway "nitrate-nitrite-oxide". A t the
same time, magnesium and, especially, beryllium nitrates have a higher
decomposition rate. A l o n g with a higher polarizing ability o f the cation, the
256 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

transition from ionic to covalent bond occurs, and beryllium nitrate has the
covalent structure. The decomposition and melting temperatures o f alkali
and alkali-earth metal nitrates [16.5] are presented in Table 16.2.1. F o r
nitrates of trivalent cations, as well as for perchlorates, higher
decomposition rates and successive decomposition are typical. A t the first
stage, oxynitrate is formed. For decomposition o f oxynitrate to oxide,
temperatures w h i c h are higher by 1 0 0 - 1 5 0 ° C are required:

Me(N0 ) 3 3 = MeON0 + N 3 2 + 2.50 . 2

Table 16.2.1. The melting and decomposition temperatures of metal nitrates

Salt LiN0 3 NaN0 3 KN0 3 RbN0 3 CsN0 3

T ,°C
d
474 529 533 549 584
T ,°C
m
253 306.5 334 321 409

Salt Be(N0 ) 3 2 Mg(N0 ) 3 2 Ca(N0 ) i 3 : Sr(N0 ) 3 2 Ba(N0 ) 3 2

T °C
d)
125 450 575 635 675
T ,°C
m
- 426 402 615 594

Table 16.2.2. The equilibrium pressures of P b ( N 0 ) decomposition products 3 2

T°C 223 230 250 274 297 357 448

P,mm Hg 6.2 6.9 11.8 32.6 78.4 514 1180

The successive mechanism o f decomposition is unambiguously determined


by the reduction o f the polarizing action o f the cation at the formation o f
oxynitrate.
D u r i n g decomposition o f hydrates o f metal nitrates, very frequently, at
temperatures near 100°C, along with water separation, decomposition o f
N0 ~ 3 proceeds, which makes it difficult or even impossible to obtain
anhydrous nitrates by drying their crystal hydrates. Thus, during U 0 ( N 0 ) 2 3 2

• 1 0 H O decomposition, U 0 , water and nitric oxides are formed [16.10].


2 3 8

A n a l y z i n g the decomposition o f nitrates, one should take into account


that for many nitrates their enthalpy o f formation is considerably lower than
that for the oxide formed. This difference has to be compensated by the
term TAS to make the decomposition possible. Under the given conditions
Thermal Decomposition and Combustion 257

of the process, this compensation is attained only at a particular value o f T,


and decomposition below this temperature is not observed. Since the
entropy o f gases depends on pressure, this temperature w i l l be also pressure
dependent. Under low pressures o f the decomposition products, the
decomposition w i l l proceed at considerably lower temperatures. For lead
nitrate, the equilibrium pressures are experimentally determined [16.7] and
presented in Table 16.2.2. This makes even more difficult the comparison
of the decomposition rates of various nitrates, especially if the
decomposition is based on the "decomposition temperatures". Thus, for
sodium and potassium nitrates, the estimating calculations show that the
observable decomposition temperatures are close to those at which the
equilibrium pressure o f the decomposition products attains the value o f 1
atm. Probably, the decomposition o f such nitrates is o f the equilibrium
character.

16.3 Metal Salts of Dinitramide. Decomposition of


Potassium Salt in the Liquid State
The main regularities o f decomposition o f metal salts o f dinitramide were
revealed by the example o f the potassium salt o f dinitramide (T \ me t =
127°C), which is the most available and convenient object to study the
reaction in the liquid and solid phases. Decomposition o f potassium
dinitramide within a temperature range o f 1 3 0 - 1 9 0 ° C has very simple
formal kinetic regularities [16.11]. The reaction is described by an equation
of the first order (Fig. 16.3.1), and the rate is independent o f the degree o f
vessel charging by the substance (m/V, g/cm ) within the limits o f m/V 3
-
0.001-0.1, and also o f the hundred-fold dilution o f potassium dinitramide
salts by the K N 0 / N a N 0
3 2 eutectic (5.5 : 4.5 wt fractions, T mdl = 140°C).
The composition o f the products is always in agreement with the following
stoichiometric equation:

K N ( N 0 ) — ^ N O + 0.5N O + 0.5KNO + 0.5KNO .


2 2 2 2 3

On addition o f the products, the rate does not change. The expression for
the rate constant takes the form:

k = 10 1 5 1 ± 0 3
exp[(-39500±0.6)/RT], s" .
1

In order to clarify the details o f the mechanism, the potassium dinitramide is


tagged by the 1 5
N isotope along the nitrogen o f the nitrogroup ( N ) or the H

central amine nitrogen ( N ) ; also, a , 5


N O , N 0 , N a N 0 , and K
l 5
2
l 5
2
, 5
N0 3 were
used. W i t h the help o f tags, the distributions o f N H
and N in N O , N 0 , and
a
2
258 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

solid residual were determined under various conditions o f decomposition,


as well as the mutual exchange between N O , N 0 , N 0 , N 0 ~ , N 0 ~ , and 2 2 2 3

N(N0 V. 2

A s for the exchange o f nitrogen atoms, the following facts were


revealed. The anion N ( N 0 ) ~ does not exchange with the products o f
2 2

decomposition; N O and N 0 do not exchange with N 0 " and N 0 ~ ; N 0


2 2 3 2

does not exchange with N 0 " , but very rapidly (apparently, by


3 the
mechanism o f electron transfer) exchanges with N 0 ~ . In the presence o f 2

N 0 , the liquid eutectic K N 0 / N a N 0


2 3 2 at 170°C is instantly hardened due to
N0 2
_
oxidation by nitrogen dioxide, but the exchange proceeds even faster,
because the tag from 1 5
N0 2 enters the solid product. There is also very rapid
exchange between N O and N 0 . 2

Fig. 16.3.1. Kinetic curves of decomposition of potassium salt of dinitramide at m/V


= 0.001. The numbers on the curves denote the temperature in degrees Celsius: (/)
the sample of aqueous ethanol, (2) the sample of water, (3) the sample of acetone,
(4) the sample of 1% solution in the salt eutectic.

The experiments on the decomposition o f a tagged potassium dinitramide


were carried out at m/V= 0.001 and T= 190°C. The analysis was conducted
after a complete and partial decomposition. First o f all, it was observed that
at the decomposition o f a 1% solution o f potassium dinitramide in the
eutectic N a N 0 / K N 0 ,
1 5
2 3 a 1 5
N isotopic tag transfers into the gaseous
products o f decomposition, i.e., N O and N 0 . This unambiguously indicates 2

that during the decomposition o f potassium dinitramide, N 0 2 is formed as


the intermediate product, through which extraction o f 1 5
N from the
condensed phase proceeds. Taking into account the data on isotopic tag
Thermal Decomposition and Combustion 259

exchange, it was revealed that during the decomposition N O is formed only


from nitrogroups and N 0 by combination o f N and N . A t the end o f the
2
a H

reaction, the fraction o f N O in N O is as large as 8% due to the exchange


a

reactions. The equivalent amount o f N is fixed in N 0 in the form o f N 0 . a


2
a
2

F r o m the experiments with the additives o f N O it was revealed that isotopic


composition o f N 0 always corresponds to the composition o f isotopes in
2

N O . Thus, at least in closed vessels, N 0 is formed with participation o f 2

NO.
Decomposition o f potassium dinitramide in the presence o f a threefold
excess o f N 0 was carried out to clear up some questions o f N oxidation. It
2
a

turned out that in this case, the rate o f decomposition did not change, but
the only products o f decomposition are N 0 and K N 0 . Moreover, N did 2 3
a

not enter N 0 ~ .
3

The data obtained make it possible to write down nearly unambiguously


the possible scheme o f decomposition, which takes the following form:

Scheme 1. Decomposition o f potassium dinitramide


in the melt.

0 NNN0 ~——
2 2 0 NNT + N 0
2 2

O ^ 1
+ N0 2 O ^ N O + N O (in a cell)1

202NN0 — — 2N0 " + 0 N N 1 3


2 2

0 NN + NO
2 N 0 + N0
2 2

N0 " + N 02 2 N0 " + NO3

In vacuum, instead o f (4) the following reaction can proceed:

0 N N + N0 " — - N0 ~ + N 0 .
2 2 3 2

In the presence o f N 0 , instead o f the reaction o f disproportionation (3), the


2

following reaction o f recombination proceeds:

0 NNCT+ N 0
2 2 — w
[0 NNON0 ]"—- 0 NN + N0 "
2 2 2 3

and, further, reaction (4) o f Scheme 1 proceeds.


The attempt to change at least one stage o f this scheme results in a
contradiction with the experimental observations. The first stage
(detachment o f N 0 ) is in good agreement with the kinetic parameters: the
2

increased pre-exponential factor and the value o f the activation energy,


w h i c h is typical o f the N — N bond breaking. The second stage is a rapid
process o f disproportionation, which is typical o f N 0 . The 2 reaction
proceeds so fast that the exchange N 0 / N 0 ~ has no time to proceed on the 2 2
260 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

cell boundaries. Reaction (3) is the best for separation o f the decomposing
anion into two forms, one o f w h i c h can be oxidized only into a nitrate-ion,
and the other is transformed into N 0 . Exactly this reaction provides the
2

constant composition o f the products during the decomposition. Biradical


0 N N possesses a large lifetime and has the time to react with N O or N 0 ~
2 2

before decomposition. Probably, at a higher temperature the following


decomposition w i l l be observed:

0 2 N N - ^ O+ N 0 2

and the composition o f products w i l l be changed towards the formation o f


N 0 and K N 0 . A n increase o f the N 0 yield is observed at 2 2 0 ° C .
2 3 2

The secondary reactions proceed rapidly and do not involve the parent
substance. Thus, equation (1), describing the rate o f decomposition,
corresponds to the first monomolecular stage o f decomposition o f the anion.

16.4 Decomposition of Potassium Salt in the Solid


Phase

The kinetic studies o f potassium dinitramide decomposition below the


melting temperature were carried out by the manometric and gravimetric
(the vacuum thermogravimetric scales with a sensitivity o f 0.1 ug) methods
with samples recrystallized from the aqueous alcohol with a residual solvent
content o f 0.2-0.25% [16.12].
The first difference from the liquid phase is that during decomposition
o f the solid salt only N 0 and K N 0
2 3 are always formed. D u r i n g vacuum
decomposition (a continuous evacuation up to 0.01 Torr), the reaction
proceeds with the acceleration and is characterized by a considerable
retention ( F i g . 16.3.2). The mass loss Am at a complete decomposition in
accordance with the following equation:

KN(N0 ) 2 2 KN0 + N 0 3 2

should be 3 2 . 3 % from the initial mass w , and, actually, at 7 0 - 1 0 0 ° C , the


0

reaction stops at Am/m 0 = 0.12, i.e., a retention effect up to 60%.


The kinetic curves cannot be described by a topochemical equation o f
the following type:
a = \-Qxp(-kt) n

for any o f the n values. The reason is an unusually pronounced retention


effect, caused by the action o f specific mechanisms o f the reaction halt.
Thermal Decomposition and Combustion 261

A s can be seen from F i g . 16.3.2, at 1 0 4 - 1 0 8 ° C the degree o f


conversion and the rate o f decomposition are considerably higher than at
other temperatures. It turns out that the temperature o f 108°C represents the
melting point o f the potassium dinitramide eutectic with K N 0 , which is
3

always present in potassium dinitramide as an impurity.


It is not the eutectic melting (in the melt, the rate decreases) but the
process o f its formation, generating the defects, that is o f importance. A t
115°C, the fast submelting o f the reaction centers proceeds, and the reaction
halts at the level o f 15% decomposition (Fig. 16.3.2).
Samples recrystallized from water decompose more rapidly than those
from alcohol or acetone in comparison with the parent sample from
recrystallized aqueous ethanol. Size reduction o f crystals leads to the
growth o f the initial rate, but, in this case, retention also increases.

Fig. 16.3.2. Kinetic curves of potassium dinitramide decomposition under the


pressure 0.01 Torr. The numbers on the curves denote the temeprature in degrees
Celsius: (/) a sample of aqueous alcohol, (2) a sample of water, (3) a sample of
ethanol, (4) a sample of acetone, (5) the moment of introduction of water vapors, (<5)
after the reaction interruption for the sample cooling up to 20°C.

The partially decomposed crystals are easily cracked when shaken and even
at simple cooling-heating o f the sample. In this case, a stepwise increase o f
the rate occurs and the degree o f transformation increases by 3 - 5 % . The
cooling-heating treatment can be repeated several times after the reaction
halt and so the degree o f transformation can be brought up to 80-85%).
A t the introduction o f water vapors (10 Torr) at any arbitrary stage o f
decomposition, the reaction immediately halts. A m m o n i a , ethanol and
262 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

acetone vapors and nitrous oxide also have an inhibiting influence, but they
act considerably more weakly than water.
In closed vessels, at an initial pressure o f 0.1 Torr and a degree o f
vessel charging o f m/V- 0.01, decomposition o f potassium dinitramide first
proceeds rapidly, as in vacuum, but as the vapors o f solvent pass into the
gas phase and N 0 is accumulated, the reaction is transferred into the
2

inhibited mode (Fig. 16.3.3).

Fig. 16.3.3. Kinetic curves of potassium dinitramide decomposition (0.3% H 0 ) in a


2

closed volume for m/V = 0.02 g/cm . The numbers on the curves denote the
3

temperature in degrees Celsius: (/) the moment of evacuation of the gas-phase


products, (2) a sample dried up to 0.05% H 0 , (3) the moment of the sample cooling
2

(up to 20°C) and heating.

The evacuation o f the products stimulates the decomposition. A rapid


cooling-heating treatment acts only on the rate o f decomposition o f the
sample, dried up to 0.05% o f H 0 . A t m/V > 0.5 or at the atmospheric
2

pressure in air, the reaction proceeds at a low rate from the beginning. The
rate o f such an inhibited decomposition was measured for the aqueous-
alcohol sample from the N 0 yield. For the rate constant the following
2

expression was obtained:

A = 10 156
exp(-41000/#r), s~ .
l

Provided that sufficient humidity is preserved, the solid potassium


dinitramide at 8 0 ° C is more stable than in the melt by a factor o f 10, and is
more stable than in vacuum by a factor o f 10 . 4
Thermal Decomposition and Combustion 263

The simplest explanation o f the observed facts can be given based on


the hypothesis o f variation o f the anion structure on transition from the melt
into the solid state. In the liquid phase, the anion has the most stable
symmetrical structure with the homogeneous charge distribution through
the anion. In the crystal state, the charges o f nitrogroups and the lengths o f
the N - N bonds in the anion are not equivalent [16.13, 16.14]. Such anions
have a low stability due to the tendency to decompose into N 0 and N 0 2 3
_
.
In ionic crystals, o w i n g to the volumetric restrictions, in the ideal lattice
the reaction, as a rule, can proceed only on the surface and on the defects.
For potassium dinitramide, the role o f the open surface increases, because
the decomposition rate sharply increases in the dry layers o f the substance.
The humidity dependence is explained by the fact that water sorption (or the
other polar dipole molecules) shields an anion from the polarizing influence
o f cations, smoothes the electron asymmetry o f the anion, and thus,
increases its stability. Polarization o f the anion, its structure and stability
can strongly depend on the microstructure o f a defect. It is quite possible
that because o f it the halt o f the reaction in vacuum is connected with the
consumption o f the defects and by the covering o f the surface by a
continuous layer o f the product.

16.5 Dinitramide Salts of Other Metals

Dinitramide salts with L i , N a , R b , C s , and B a


+ + + + 2 +
cations have been
studied in detail in [16.15]. In addition, there are some data on the salts with
T l , A g , and C d
+ + 2 +
[16.16]. Before the experiment, the products were twice
recrystallized from ethanol or aqueous isopropanol and were dried over the
anhydrone in a vacuum exiccator. In this case, the water content rapidly
decreases up to 0.2-0.4%, after which the drying process sharply
decelerates. A more strict drying up to 0.05% o f H 0 (the temperature is 2

6 0 - 7 0 ° C , in vacuum) is accompanied, as a rule, by a partial decomposition


o f the substance. L i t h i u m and barium salts on synthesis are obtained in the
form o f low-melting crystal hydrates. Their dehydration up to 0.5% o f H 0 2

is attained during their dehydration over P 0 2 5 at room temperature in


vacuum under pressure o f 0.01 Torr.
The decomposition rate was measured by the manometric and
gravimetric methods. Liquid-gas chromatography and mass-spectrometry
were used for the analysis o f the decomposition products.
For all metal salts, the kinetic regularities and the composition o f the
decomposition products were found to be the same as for potassium
dinitramide. In the liquid phase (the melt, the eutectic with metal nitrate,
and the solution in the eutectic N a N 0 / K N 0 2 3 or in potassium dinitramide),
the reactions proceed according to the equation o f the first order, and N O
and N 0 are formed as the gas phase products in the ratio o f 2 : 1. The rate
2
264 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

o f decomposition, i.e., the stability o f the N 0 ~ anion, linearly depends on


3 4

the polarizing force o f the cation [16.14].


Decomposition in the solid phase is accompanied by the production o f
N 0 only. In vacuum, the reactions proceed faster than in the melt or in the
2

closed vessel, and are characterized by a strong retention. The ratio o f the
rate constants o f decomposition in the solid and liquid phases at 100°C
changes from 10 ( L i D N A , B a D N A ) up to 10 ( N a D N A ) . Water vapors, as
4

in the case of potassium dinitramide, lead to immediate reaction


termination. R u b i d i u m and cesium salts o f D N A give eutectic mixtures o f
9/1 and 9.5/1 composition with the nitrates o f corresponding metals, and at
the melting points o f these eutectic mixtures (84.5 and 81.5°C), an increase
o f the decomposition rates is observed.
A l l kinetic characteristics for the salts studied are presented in Table
16.3.1 and in F i g . 16.3.4.

-4

-5

-6

-7

-8

10

2.6 2.8 3.0 r-MoVK

Fig. 16.3.4. Temperature dependence of the rate constants of decomposition of D N A


salts in the solid phase. The symbols on the lines correspond to cations: (/) the
reaction in vacuum, (2) the decomposition inhibited by water vapors.

A strong influence o f water vapors allows one to conclude that the


decomposition in the solid phase proceeds only in the dry layers o f a
substance and, thus, its rate w i l l depend on the degree o f preliminary drying
o f the substance and the conditions o f extraction o f the residuals o f a
solvent and gaseous products. L i t h i u m and barium salts o f D N A , having the
smallest ratios k$/k\, were obtained from crystal hydrates and, probably, the
residual moisture inhibits their decomposition in vacuum at the initial
stages. The same substances have the highest degree o f acceleration (which
Thermal Decomposition and Combustion 265

can be connected with water loss), and, at the acceleration stage, they are
equaled by the ratio k$/k\ with the other salts.

Table 16.3.1. The kinetic characteristics of decomposition of metal salts of


dinitramide in the solid and liquid states

Cation r„,°c Medium Ar,°c E, \og(A), s" J M O V k,s~ at 100°C


l

kcal/mol at 160°C

Li +
158 melt _ (36.0) a
(15.0) 67 b
8.1 • i o - 7

Li +
158 solid 62-120 26.0 10.0 5.7 • 10~ 6

(0.1 )
c

Li +
158 solid 60-98 40.6 17.3 3.2 • i o - 7

(760)
Na +
111 melt - (38.0) (15.0) 6.4 b
5.4 • i o - 7

Na +
111 solid 30-50 40.0 22.0 3.6 • i o - 2

(0.1)
Na +
111 solid 70-80 46.0 21.4 2.7 • 10" 6

(760)
K +
127 melt 130-190 (39.5) (15.1) 1.4 8.9 • i o - 9

K +
127 solid 80-100 26.6 10.3 5.1 • 10" 6

(0.1)
K +
127 solid 80-120 41.0 15.6 3.7 •IO" 9

(760)
Rb +
108 melt - (39.6) (15.0) 1.0 b
6.2 •IO" 9

Rb +
108 melt 130-170 (40.1) (15.3) 1.1 6.3 •IO" 9

Rb +
108 eutec­ 100 - - 6.1 • i o - 9

tic with
RbN0 3

Rb +
108 solid 50-100 20.0 6.6 7.6 • 10~ 6

(0.1)
Rb +
108 solid 70-102 38.0 12.8 3.4 IO" 10

(760)
Cs +
88 melt - (39.2) (15.0) 1.62 b
1.0 • 10 -8

Cs +
88 solid 60-85 13.0 3.0 2.4 • i o - 5

(0.1)
Cs +
88 solid 70-80 39.0 14.7 6.9 • IO" 9

(760)

Ba 2+
solution in 160-180 50.0 20.8 3.2 2.8 • IO" 9

potassium
dinitramide
Ba 2 +
solid 80-100 47.3 20.5 5.6 • 10" 8

(0.1)

a
Calculated from the k values at 160°C, presented in [16.14], under the assumption
that \o%(A), s~ has a value of 15 typical of alkali salts.
]

b
T h e data from [16.14].
c
The pressure of air over the sample (Torr).
266 Metal Perchlorates and Nitrates, Metal Salts of Dinitramide

The results obtained indicate that all metal salts o f D N A have a tendency to
anomalous decomposition, whose origin can be connected with the
asymmetrical electron structure o f the anion in the solid phase. In the
inhibited mode o f decomposition, the solid salts represent rather stable
compounds.

REFERENCES

16.1. F . Solymosi, Structure and Stability of Salts of Halogen Oxyacides


in the Solid Phase, L o n d o n : J. W i l e y , 467 pp. (1977).
16.2. Thermal Constants of Substances. Reference Book, V . P. G l u s h k o
E d . , M . : Izd. V I N I T I , 1-10 (1968-1981).
16.3. D . G . Lemesheva and N . V . Krivtsov, Zh. Neorg. Khim., 30(9):
2216-2221 (1985).
16.4. H . F . Cordes and S. R. Smith, J. Phys. Chem., 72(6): 2189-2193
(1968).
16.5. V . P . Babaeva and V . Y a . Rosolovskii, Zh. Neorg. Khim., 29(11):
2731-2735 (1984).
16.6. V . P. Babaeva and V . Y a . Rosolovskii, Zh. Neorg. Khim., 31(7):
1710-1715 (1986).
16.7. C . C . Addison and N . Logan, Advances in Inorganic Chemistry
and Radiochemistry, 6: 115-124 (1964).
16.8. L . A . Alekseenko, Proceedings of Tomsk University, 126: 2 0 - 2 6
(1954).
16.9. C . Ettarh and A . K . Galvey, Thermochimica Acta, 228(1-2): 2 0 3 -
219 (1996).
16.10. A . K . Rajagopalan, P . V . Ravindran, and T . P . Radhakrishnan, J.
Therm. Anal, 44: 88-96 (1995).
16.11. F . I. Dubovitskii, G . A . V o l k o v , V . N . Grebennikov, G . B .
Manelis, and G . M . N a z i n , Dokl. Rus. Akad. Nauk, 347: 763-765
(1996).
16.12. F . I. Dubovitskii, G . A . V o l k o v , V . N . Grebennikov, G . B .
Manelis, and G . M . N a z i n , Dokl. Rus. Akad. Nauk, 348: 2 0 5 - 2 0 6
(1996).
16.13. F . I. Dubovitskii, N . I. Golovina, A . N . Pavlov, and L . O .
A t o v m y a n , Dokl. Rus. Akad. Nauk, 355: 200-205 (1997).
16.14. R . G i l a r d i , J. Flippen-Anderson, C . George, and R . J. Butcher, J.
Am. Chem. Soc, 119: 9411-9416(1997).
16.15. S. B . B a b k i n , A . N . Pavlov, and G . M . Nazin, Izv. Rus. Akad.
Nauk, Ser. Khim., 1947-1950 (1997).
16.16. Z . Pak, AIIA/SAE/ASME/ASEE 29 Joint Propulsion Conf,
Monterey, C A , FIIA-93-1755, June 2 8 - 3 0 , (1993).
Chapter 17

BASIC ASPECTS OF THE COMBUSTION


MECHANISM

As follows from the previous chapters, the substances used for the
manufacture o f composite systems, including solid rocket propellants and
powders, are characterized not only by the pronounced variety o f their
chemical structure and reactivity but also by the particular difference o f
physical properties that, all together, predetermine a variety o f combustion
mechanisms. Each o f the components o f a complex system, such as an
oxidizer, a combustible binder, a metal, a catalytic or inhibiting additive,
and some others, are subjected during combustion to phase and chemical
transformations, as well as to a complex interaction with other components,
w h i c h start in the solid or liquid phase o f a surface layer, continue, and
come to an end in the gas phase.
A comprehensive description o f combustion, considering the interaction
o f all chemical and physical processes, already meets with difficulties at the
stage o f formulation o f the set o f input equations, including a complex
kinetics and heat- and mass transfer. These difficulties increase with the
solution and analysis o f a mathematical problem.
However, there are a number o f fundamental characteristic features o f
combustion, which make construction o f simplified combustion models
quite fruitful. In this approach, in each particular case, the factor which
plays the main role in the combustion process is selected and certain
secondary phenomena are excluded from consideration. This makes it
possible to complete the physical picture o f combustion and to describe it
mathematically by a comparatively simple set o f equations.
A pronounced exponential temperature dependence o f the rate o f a
chemical reaction is one o f the most important features o f combustion o f
gases and condensed matter. A s a result, the reaction is localized within a
narrow zone. However, because the values o f the rate constants can differ
considerably by many orders o f magnitude, the reaction rate in a complex
chain o f parallel and consecutive reactions, developing in various aggregate

267
268 The Basic Aspects of the Combustion Mechanism

states o f a substance, is determined by only one particular stage o f chemical


transformation.
For instance, during decomposition o f nitrocompounds, such a stage,
determining the rate, is the initial slow stage o f nitrogroup detachment,
w h i c h has a high activation energy. The subsequent reactions o f radicals
proceed relatively fast and virtually do not influence the total rate (see
Chapter 3). In this case, the introduction o f the concept o f a leading
combustion stage is completely adequate. This stage is responsible for the
formation o f the rate and for its dependence on the external conditions o f
combustion (the initial temperature and pressure) and on the internal
physical-chemical parameters o f the system. Thus, in a number o f cases, a
complex process o f chemical transformation in combustion can be
described by single-stage kinetics, o f course with a certain degree o f
approximation. However, under the variance o f the combustion conditions,
the changing o f a leading stage is possible, and such a change can proceed
within a comparatively narrow range o f the parameters and may have the
character o f a critical phenomenon. For instance, in a relatively simple case
o f only two stages o f combustion, and, correspondingly, two zones o f
chemical transformations, separated in space, various combustion modes
appear: the "governing" mode, when the leading stage is the reaction,
proceeding at temperatures close to the maximum one, and the heat flow
from this zone determines the reaction rate at the first stage; the "self-
ignition" or "separation" mode, when the leading stage is the reaction at
relatively low temperatures, and the heat flow from the subsequent zone is
negligible; and finally, so-called " m i x e d " combustion mode, when, due to
the closeness o f the reaction rates in both stages and the similarity o f their
variation, both zones jointly influence the combustion rate within a
sufficiently wide range o f variation o f the combustion conditions.
Thus, the most important problem in the study o f a combustion
mechanism is the selection of a leading stage o f combustion (its
localization, the temperature range, the dependence on the input parameters)
and the degree o f influence o f other processes. O n this basis, one can solve
the problems o f the efficient variation o f combustion characteristics having
practical importance (for instance, the ballistic coefficients), by the
influence o f the leading stage with the help o f catalysts or inhibitors o f
combustion, by variation o f the chemical nature, concentration, dispersion,
and some other physical-chemical constants o f the components.
O n combustion o f condensed substances, w h i c h is accompanied by gas
generation, the problem o f mathematical description o f the conditions o f the
combustion surface formation arises, and, naturally, o f the finding o f its
temperature, w h i c h is one o f the most important characteristics in a number
o f models.
A t transition from the condensed to the gas phase, such parameters as
the density, the thermal-physical constants, the flow velocities, the rate o f
chemical reaction, and some other characteristics drastically change,
Thermal Decomposition and Combustion 269

although, in some exotic cases, a smooth variation o f these characteristics is


observed (for instance, the mode of combustion with foaming).
Nevertheless, in the majority o f cases, a clearly defined surface is formed,
and, for its description, such processes as evaporation, sublimation, and
dispersion should be considered.
Naturally, the approach proposed for the modeling o f combustion is
limited. The models developed make it possible to determine the stationary
rate o f combustion and its dependence on the internal and external
parameters, to provide the basis for the solution o f a number o f problems o f
non-stationary combustion, and to propose methods o f control o f the rate
and some other characteristics o f combustion.
The choice o f the model and its validation are based on the study o f the
kinetics and mechanism o f chemical reactions and on the analysis o f the
whole complex of physical characteristics of compounds, and the
confirmation o f the validity o f models and their prediction capability are
based on comparison o f the experiments and calculations in w h i c h the
independently determined (if possible) kinetic and thermodynamic
characteristics are used.
The elementary models, even in a simple, mechanistic combination, can
provide at least a qualitative description of certain phenomena
accompanying the combustion o f complex systems, and also the basis for
the development of more comprehensive models, which will be
demonstrated below. For instance, these include the model o f combustion,
which takes into account combustion in the condensed and gas phases, and
the models o f combustion o f composite and layer systems.
It should be noted that all concepts mentioned above, as well as the
classical theory o f combustion, are based on the assumption o f complete
equilibrium on transitional, vibrational, and rotational degrees o f freedom.
A t the same time, the high rates o f chemical transformations, the short
characteristic times o f the processes, and the high temperature and
concentration gradients make possible, in principle, the deviation from such
an equilibrium. Analysis o f the kinetics o f non-equilibrium behavior in the
combustion o f particular high-energy compounds (hydrazine, ethylnitrate),
performed below, shows that the account o f non-equilibrium phenomena is
necessary in order to explain o f a number o f regularities. Studies in this
direction are under development and this case has to be considered as an
example o f a new approach to the solution o f up-to-date problems o f
combustion.
Chapter 18

COMBUSTION OF PURE SUBSTANCES:


REACTIONS IN THE CONDENSED
PHASE

18.1 The Model with Dispersion of the Solid


Substance
The process o f propagation o f a chemical reaction in the form o f a
combustion wave can develop completely and exclusively in the condensed
state and can be described by laws w h i c h are similar to those for the gas
phase combustion. In particular, the formulas for the combustion rate,
obtained by the Zel'dovich—Frank-Kamenetskii method, are the same for
both cases for the reaction o f zero order, when it is not necessary to
consider the diffusion equation [18.1]. The frontal polymerization, the
combustion o f thermite compositions at low pressures, and some other,
relatively sparse processes can serve as examples o f propagation o f the
reaction wave o f the substance in the purely condensed state.
However, in the majority o f cases, combustion o f the condensed
substances and their mixtures proceeds with a complete or partial transfer o f
the condensed phase into the gas phase at a particular stage of
transformation. If the parent substance is nonvolatile and is capable o f
exothermic solid phase decomposition, then the gas phase products formed
in the reaction mechanically destroy the solid phase, crush it and eject fine
particles o f the substance, forming a so-called dust-gas zone. This
phenomenon was called dispersion in combustion. For the first time, it was
observed and studied in combustion o f the initiating explosives and
powders [18.2], and it manifests itself most clearly at sub-atmospheric
pressures. A t low pressures (2-10 m m Hg), one can observe "flameless"
combustion: a tablet o f a pressed substance "vanishes" without visible
luminescence, forming a large amount o f smoke (70-80%), which
represents a dispersed parent substance. A s the pressure increases, a flame
appears, whose intensity increases, and the distance from the surface (a
"dark zone") decreases. The classical examples o f these phenomena,
described by A . F. Belyaev and P. F. Pokhil, are the combustion o f mercury
fulminate and pyroxylin powder.

271
272 Combustion of Pure Substances: Reactions in the Condensed Phase

The combustion process in the solid phase is described by the ordinary


equations o f heat- and mass transfer, in which the diffusion term is absent.
The formula for the combustion rate, which takes into account
incompleteness o f transformation in the solid substance, takes the following
form [18.3]:

r 2 = A R T Aexp
s
2
(-E/RT ) S

p Q E J Q

where T is the temperature o f the combustion surface, X is the coefficient


s

o f thermal conductivity, p is the density, Q is the thermal effect, A and E are


the kinetic parameters o f the reaction, R is the universal gas constant.

T = T + (Q/cX',
s 0 j =l!LzlLdTi.
0

o /(?)

Here, c is the heat capacity, r\ is the degree o f decomposition on the


x

combustion s u r f a c e , / ^ is the term, which takes into account the type o f


the kinetics.
For the reaction o f the zero order we w i l l have:

/0?) = 1; J ={tf)l2.
Q

For the reaction o f the first order we w i l l have:

f(Tj) = l-rj, y = 7 - ( l - ^ ) l n (l-7 ).


0 r r r

The degree o f dispersion is connected with the degree o f decomposition on


the combustion surface by a simple relationship:

A s an example, the results o f calculation o f the flameless combustion o f


mercury fulminate with the use o f the kinetic parameters o f the substance
decomposition in the solid phase in comparison with the experiment are
presented in Table 18.1.1.
The experimental observations showed that the amount o f smoke in
combustion considerably decreases with an increase in pressure. For
instance, in combustion o f pyroxylin powder it reduces from 7 0 % in
flameless combustion to 3 0 % at 100 m m H g and 0.6% at 100 atm. This
definitely indicates the functional connection existing between the
dispersion and pressure dependence o f the combustion rate.
Thermal Decomposition and Combustion 273

Table 18.1.1. The combustion characteristics of mercury fulminate

Experiment Calculation

>7d r, cm/s >7d r, cm/s


630-770 *0.7 0.4 0.76-0.66 0.1-0.58

In the simplest form this relation can be revealed, based on the following
reasons [18.4]. The force, acting on the solid substance from the gas flow, is
proportional to the pressure gradient in the pores o f the decomposing
substance, which, in turn, is connected with the rate o f filtration by D a r c y ' s
law:
V ozdPldx,
{

where V is the rate o f filtration, P is the pressure. Taking into account the
{

equation o f continuity o f the flow,

where p is the gas density and p is the density o f the condensed substance.
% c

Equating to each other the destructive force and the strength o f the
substance at the moment o f formation o f the combustion surface, we w i l l
obtain:
r =A fa,) P,

where the coefficient A(r)J incorporates in the generalized form the


dependences o f such characteristics as penetrability, strength, and others on
the volume o f pores, which is proportional to the degree o f decomposition

Considering this total dependence as a relatively weak one, and taking


into account the logarithmic character o f the relations between r/ and P, we r

w i l l obtain a nearly linear dependence o f r on P that in a number o f cases


corresponds to the experiments on combustion o f particular initiating
substances (for instance, potassium picrate, mercury fulminate, lead
dinitrophenolate) at low pressures [18.2].

18.2 The Model with Foaming of the Reacting


Substance
For the substances melting during combustion, the foaming o f the liquid
layer is typical, which represents the result o f decomposition reactions and
the formation o f the gas phase products. It was observed for the first time in
274 Combustion of Pure Substances: Reactions in the Condensed Phase

combustion o f dibasic propellants containing nitroglycerin [18.5]. The


foaming effect exhibits itself in a gradual increase o f the volume and
corresponding decrease o f the density due to the increase o f the number and
volume o f bubbles containing the gas phase decomposition products. In this
case, the decrease o f the heat release in the unit volume proceeds.
In its rectified form, the foaming model is valid for the nonvolatile
substance whose combustion front represents the foamed state, starting
from the softening temperature up to the highest temperatures. The theory
o f this process, developed by E . I. M a k s i m o v and A . G . Merzhanov [18.6],
gives the following formula for the combustion rate:

r l J X M T PAexp
n (-E/RTJ
p2
Q E Z

where T m is the m a x i m u m combustion temperature, M is the molecular


weight o f the gas phase products, Z is their weight fraction. F r o m this, it
follows that the coefficient n o f the pressure dependence o f the combustion
rate should be close to 0.5.
It was shown that there are some substances whose combustion is
quantitatively described by the theory o f volumetric dispersion (foaming).
Polyvinylnitrate and boranehydrazine are among them. Combustion o f the
latter substance was studied in [18.7].
Apparently, foaming on combustion o f boranehydrazine is caused by
the fact that the reaction o f decomposition with the formation o f boron
nitride proceeds in the liquid phase, and the suspension o f the solid B N in
the liquid substance possesses a high effective viscosity. The product o f
combustion represents a white, porous, fine substance with a density o f
approximately 0.01 g/cm . The reaction proceeds with the yield, w h i c h is
3

close to 100%, and follows the equation:

BH N H 3 2 4 -> B N + 0 . 5 N + 3 . 5 H .
2 2

B o r o n nitride is formed in the amount o f approximately 5 0 % from the


initial weight. This compound can be considered as an example o f an
efficient intramolecular oxidation o f boron and, o w i n g to the low molecular
weight o f the gas phase products and rather high combustion temperature ( «
1 5 0 0 ° K ) , is o f interest as a component for particular fuel composites. B o r o n
nitride as the product o f combustion also possesses particular valuable
properties.
W i t h i n a pressure range o f 2 0 - 1 0 0 atm, the pressure dependence o f the
combustion rate is described by the following equation:

r(cm/s) = 0 . 6 2 / )0 58
,(atm).
Thermal Decomposition and Combustion 275

Calculation o f the combustion rate o f boronhydrazine was carried out by the


formula presented above with the use o f the kinetic data on the liquid phase
decomposition o f the substance. The following dependence was obtained,
which is in a good agreement with the experiment:

r(cm/s) = 0 . 8 8 / , ( a t m ) .
)05

18.3 The Model with Evaporation or Sublimation

The majority o f individual high-energy compounds, including explosives


and the components o f rocket propellants, are capable o f evaporation or
sublimation at temperatures which can be attained in combustion.
Nitrocompounds, nitroethers, nitroamines, onium salts o f nitric and
perchloric acids are among them. The temperature o f transition into the gas
phase and the reactivity o f the substance are the criteria determining the
possibility o f favorable decomposition o f the substance on combustion in
the condensed or gas phases. If we suppose that the transition temperature,
w h i c h coincides with the surface combustion temperature, is close to the
boiling temperature o f a substance at a given pressure, which is true under
the assumption that the reaction o f gas formation in the liquid phase is
neglected, as was accepted by Y a . B . Zel'dovich [18.8], then the pressure
coefficient for the heavy substance, burning in the condensed phase, is
determined by the ratio between the activation energies o f decomposition
and evaporation.

r *Aexp(-E/RT ),
2
s P'« 5exp(-£ v /RT ) ,S

_P dr _ p E dT _ E s

U
~ r dP~ 2RT dP " 2E '
S
2
W

W e studied in detail the model with the leading stage o f combustion in the
condensed phase, in which, it was supposed, the after-burning reaction in
the gas phase proceeds in the "self-ignition" mode [18.4, 18.9]. The
temperature o f a combustion surface is determined by the heat release due
to the reaction o f decomposition and heat absorption due to evaporation
(sublimation).

r = r +(g /c)
s 0 r rj -{Q lc)
t w 77 .
V (18.3.1)

If the condensed substance, during reaction and evaporation, gives only


gases, then the condition o f formation o f the combustion surface takes the
form:
276 Combustion of Pure Substances: Reactions in the Condensed Phase

>7+>7v=1.
r ( - - )
18 3 2

The sum o f the partial pressures o f the products o f decomposition and


evaporation should be equal to the external pressure, i f we neglect the
increase o f dynamic pressure near the combustion surface:

P = P +P =[l T w + f(l-ri )/ri ]Bexp(-E /RT ).


y w w s (18.3.3)

The expression for the rate was obtained by the usual method o f exponent
expansion in a series for the reaction o f zero order:

r 2 = 2X R T Aexp(-E/RT ) s
2
s (18 3 4)
p E (1-77 ) [ a - ( 0 0 ) 7 ]' V r + v V

where T and rj are determined from (18.3.1), (18.3.2), and (18.3.3). The
s v

following designations are accepted: Qr is the thermal effect o f


decomposition, Q v is the heat o f evaporation (sublimation), n is the r

fraction o f decomposition, r/ is the fraction o f evaporation, / i s the ratio o f w

the number o f moles at decomposition and evaporation o f the substance, B


and Ey are the parameters characterizing evaporation.
In a simpler case, when Qy«Q , x which is equivalent to the assumption
about the compensation o f the heat consumption on evaporation by the heat
supply from the gas phase, we w i l l obtain the following expression for the
combustion rate, and the pressure and temperature coefficients:

r 2 = 2 A Q R T>A^-E,RT ) S

cp 2
E(T -T )S 0
2

E f-1
3 1 O D T 2 (^ni ~~ ^ s ) 0 r ^v)
omr 2 A 71 /
n= s J

3 l n P
l+ ^ t f - " ^ ) ( 1 - ^ riv)

a = d\nrdT _2RT s s
2
RT$ T -T s 0 f
8T S dT 0 , , E f - l
v f r r T V 1

l + ^ ( ^ - ^ s ) ( l - V ^ )
^ s /

where T = T + Q/C. m Q

In the limiting cases, we w i l l have:


Thermal Decomposition and Combustion 277

At p_>0, 77v >?->£/2£ , v

Atp_>oo, ^ ->0,
v H->0, a-^EIlRTl

W e calculated the characteristics o f combustion, in which the influence o f


the thermal effect in the condensed phase is reflected, which is determined
not only by the exothermic decomposition o f a substance, but also by the
heat transfer from the gas phase zone or by a preliminary heating o f the
substance (AQ > 0), by the heat consumption on evaporation, dispersion,
and heat dissipation through the lateral surface (AQ < 0).
A s the pressure increases, the degree o f evaporation decreases, the
temperature o f the surface and the combustion rate increase, and the
pressure coefficient decreases. If AQ > 0, the degree o f evaporation
increases, at AQ < 0, it decreases, but in this case, the values o f T , r, and n s

change insignificantly. Thus, evaporation acts by the feedback mechanism,


exerting control and a stabilizing action on the characteristics.
The pressure dependence o f the temperature coefficient has a non-
trivial appearance (Fig. 18.3.1) [18.10].

O
-6

Fig. 18.3.1. The pressure dependence of the temperature coefficient of the


combustion rate for various values of the thermal effect in the condensed phase,
which increases from curve 1 to curve 3.

Within the range o f low pressures, it decreases, and at high pressures it


starts to grow, so, consequently, the m i n i m u m is realized. This follows from
the transformation o f the formula for a; i f one w i l l assume that within the
range o f moderate pressures the following relation is true:

(£ //?r
v m
2
xr -r )»i
m s ; / = i.

Then. E 1 1
2E T -T
w m s T -T
s 0
278 Combustion of Pure Substances: Reactions in the Condensed Phase

A t low pressures, the variation o f cr is determined mainly by the second


term (the decrease o f a), at high pressures, by the first term (the increase o f
cr). In both cases, the variation o f a is connected with the variation o f rj v

and, consequently, T . This peculiarity is also reflected in the


s different
character o f the dependence o f a on T : at low pressures, the value o f a
0

decreases, and at high pressures, it increases.


The value o f the temperature coefficient determines essentially the
thermal stability o f combustion. The physical reason o f the appearance o f
thermal instability is the growth o f the excess o f enthalpy at the combustion
front o f the condensed substances o w i n g to the difference in the spatial
distribution o f the temperature and concentration, in contrast to a gas, for
which, as a rule, a similarity o f these variables is fulfilled. In [18.11], the
criteria o f combustion stability o f the condensed substances were obtained
in the general form by B . V . N o v o z h i l o v , based on consideration o f the
dependence o f perturbation o f the combustion rate on the temperature and
the temperature gradient at the combustion surface. A t T$ = const,
combustion is always stable i f

k= a(T -T )<\.
s 0

In the case when T s * const, at k > 1, stability is provided by the additional


relationship:

_ (k-\f . 3T
5 = r = —-, P = const.
S
— < 1,
r(k + \) dT 0

In the model with evaporation, this expression takes the following form (for
the particular case Q = 0 a n d / = 1) [18.12]:
v

[(E/2RT )
2
(r -r )-i] 0

S=-
S s

\ + (E/2RT ) s
2
(T -T )
S Q + (2E /RT ) V S
2
(T M -T )
S

F r o m this, it follows that the smaller E and the higher E , Q, and T values, v Q

the higher the degree o f combustion stability. Since n oc E/2E , one can V

conclude that thermal instability is more typical o f substances with a


pronounced pressure dependence o f the combustion rate. The following
equation:
S(T ) = l s 9

w h i c h determines the boundaries o f stability, is a cubic equation with


respect to T , and, consequently, has three roots, which for particular values
s

o f the parameters can be real ones and fall into the region o f possible T s

values. The dependence o f 8 on T , w h i c h is unambiguously connected with


s
Thermal Decomposition and Combustion 279

P, is shown in F i g . 18.3.2. Curve 1 corresponds to the case o f stable


combustion at all pressures (8 < 1, everywhere). Curve 3, crossing the
straight line 8 = 1 twice, characterizes combustion which is stable at low
and high pressures and unstable within the intermediate range. Curves 2 and
4 correspond to the case o f stable combustion at low pressures and unstable
combustion at high pressures.
A n important factor, which can influence the appearance o f a stability
criterion, as well as the other combustion characteristics, is the temperature
dependence o f the activation energy o f evaporation, which decreases with
the T growth.
A s follows from the aforesaid, the most important aspect o f a
combustion mechanism in the condensed zone is the question o f the value
o f the combustion surface temperature and a reasonable limitation o f the
reaction conversion degree in this zone. In the model presented, these
values are naturally determined by the joint influence o f the processes o f
decomposition and evaporation. The latter process produces a stabilizing
influence on the combustion characteristics on variation o f the different
parameters.

1.5

0.5
0
o,i o,8

Fig. 18.3.2. The dependence of the stability criterion on the temperature of the
combustion surface, which is functionally connected with the pressure for various
values of the activation energy of evaporation and decomposition.

Here, it is pertinent to make a remark about the choice o f the T values and
s

the thermal effects in the condensed phase, based on the arbitrary,


"reasonable" assumptions or in order to fit the results o f calculation to the
experiment, w h i c h can be found in some works. A s a result, the combustion
rate, which depends exponentially on the T value, also becomes a strong
s

function o f the undetermined parameter and the model o f combustion


acquires an incomplete and semi-empirical character.
A s regards the influence o f the gas-phase reaction on the characteristics
o f combustion, in the model considered it was taken into account by the
assignment o f particular values o f the heat transfer from the gas phase. A
more correct approach requires a joint consideration o f the processes in both
phases, which w i l l be done in the next section.
280 Combustion of Pure Substances: Reactions in the Condensed Phase

Let us consider the results o f the study o f combustion o f particular


substances with high-energy content.

Ammonium perchlorate (AP), NH4CIO4

Combustion o f this oxidizer is characterized by a particular characteristic


feature. The generalized pressure dependence o f the combustion rate
constructed based on the data o f many authors is presented by T. L . Boggs
in [18.13]. The lower limit o f deflagration is 20 atm. Within the pressure
range up to approximately 140 atm, the results o f measurements o f the rate
by various researchers mainly coincide. A t the same time, above this
pressure, the combustion rate o f A P becomes very sensitive to the grade o f
purity o f the substance, to the particle sizes, to the density, to the k i n d o f
sheath, and to other factors. Therefore, a considerable scattering o f the data
is observed: from a weak pressure dependence o f the combustion rate (a
plateau) to a strong decay with subsequent sharp increasing.
A number o f experimental facts point to the important role o f the
thermal decomposition o f A P in combustion: the presence o f bubbles in the
surface layer o f extinguished samples, caused by gas formation, the heat
release in the condensed zone, measured with the help o f thermocouples,
the peculiarities o f the influence on the combustion rate o f ionizing
radiation, alkali metal salts, and the compounds o f metals with variable
valence. The temperature gradients measured on the combustion surface
showed that heat inflow from the gas zone could not provide the heating o f
the condensed phase up to the temperatures observed [18.14] and
compensate the high endothermic effect o f the substance sublimation
(approximately 500 cal/g). Let us analyze the characteristic features o f
combustion o f A P within the framework o f the concept o f the determining
role o f thermal decomposition and sublimation, taking into account the
contribution o f the gas-phase reaction by the assignment o f a particular
value o f the heat inflow in the heat balance on the combustion surface [18.9,
18.15].
A s shown in Chapter 13, the initial (rapid) stage o f A P decomposition
proceeds according to the unique mechanism within the wide temperature
range o f 2 0 0 - 4 5 0 ° C studied with an activation energy o f approximately 30
kcal/mol. These kinetic parameters w i l l be used in our calculations.
According to the experiments, A P is capable of self-sustaining
combustion at atmospheric pressure i f an additional heat is supplied, whose
value is approximately 100 cal/g in the form o f irradiation energy, a
preliminary heating, or a small additive o f a fuel. The calculation showed
[18.9] that under these conditions, a good agreement o f the combustion
characteristics o f A P and the experimental results is observed (Table
18.3.1).
Thermal Decomposition and Combustion 281

Table 18.3.1. The characteristics of A P combustion at 1 atm

r, cm/s 7v

cal/g

Calc. 0.02 750 0.33 140 -160 100


Exp. 0.02-0.03 770 102-107

The characteristics o f A P combustion at elevated pressures calculated under


conditions o f compensation o f the heat losses on sublimation by the heat
inflow from the gas zone [18.15] are presented in Table 18.3.2.

Table 18.3.2. The characteristics of A P combustion

Characteristics Pressure, atm Calculation Experiment

r, cm/s 40 0.12 0.5


r, cm/s 100 0.19 1.0
7s, K 18 890 893
7s, K 35 920 923
*7v 20 0.4
>7v 100 0.32
n 20-60 0.47 0.77
n 70-140 0.44 *0.5
cr, grad -1
50-100 2.8- 10" 3
2.0- 10~ 3

8 100-300 1.5 unstable


combustion

The following conclusions can be formulated from these results:


1. The calculated rate differs from the experimentally measured one by
approximately a factor o f 5, which can be considered a quite acceptable
result, having in mind a particular error o f the kinetic data obtained for the
other conditions in comparison with the high temperature-combustion front.
Nevertheless, this fact can be explained i f one assumes the possibility o f
liquid phase formation in the surface layer on A P combustion, which was
observed in experiments by a number o f researchers. The appearance o f the
liquid phase is caused either by A P melting under these conditions or by the
formation o f the eutectic " A P - w a t e r " with a low melting temperature. The
possibility o f eutectic formation was proved by calculation o f the phase
diagram o f this composition at high pressures. It is known that the rate o f
decomposition o f the substance after melting increases usually by more than
a factor o f 10. In this case, a strong catalysis o f A P decomposition by the
dissolved perchloric acid and by the other acidic products is observed, as
well as its influence on the reaction rate, and, consequently, on the
282 Combustion of Pure Substances: Reactions in the Condensed Phase

combustion rate due to variation o f dissolubility o f these products in the


liquid phase at the pressure change [18.9].
Thus, in order to bring these results into correspondence with the
experimental data, it is necessary to increase the rate o f decomposition o f
A P in the condensed phase by a factor o f approximately 25.

2. The temperatures o f the combustion surface o f A P in theory and in


experiment are close to each other, which points to the character of
sublimation under these conditions, which is close to the equilibrium one.

3. The pressure coefficient and the temperature coefficient at a pressure o f


approximately 100 atm are in agreement with experiments, but at lower
pressures, the experimentally measured value o f n is higher than the
calculated one.

4. Finally, the value o f the stability criterion, which is higher than 1, is


indicative o f the thermal instability o f combustion which is also observed in
the experiments: the observation of luminescence pulsation at the
combustion front, and a high sensitivity to heat dissipation (the diameter o f
the sample and the type o f sheath).
Thus, the regularities o f A P combustion, as a first approximation, are
satisfactorily described within the framework o f the combustion model with
substance decomposition in the condensed zone and with the regulating role
o f sublimation.
Apparently, the role o f dispersion on combustion o f A P at a pressure
above 20 atm is insignificant, because this factor mainly appears at
subatmospheric pressures.
A review o f solid monopropellant combustion (deflagration) models
applied to A P , which include both decomposition and evaporation, is given
in [18.17].

Hydroxylammonium perchlorate (HAP), NH3OHCIO4

This salt is the analog o f A P and represents a potential oxidizer in


composite propellants. The elucidation o f the question as to what extent the
aforementioned peculiarities o f combustion are realized for this substance
and w h i c h peculiarities o f combustion can manifest themselves o w i n g to its
individual physical-chemical parameters is o f considerable interest [18.18].
The kinetics and mechanism o f thermal decomposition o f H A P in the
liquid phase (the melting temperature is 8 9 ° C ) has been well studied
(Chapter 15). Decomposition o f the substance, as well as A P , starts from the
equilibrium dissociation o f the salt into hydroxylamine and perchloric acid
and is considerably accelerated under the action o f excessive concentrations
o f these products. The influence o f the base is stronger. This is the main
Thermal Decomposition and Combustion 283

difference between H A P and A P .

Fig. 18.3.3. The pressure dependence of the combustion rate of hydroxylammonium


perchlorate for various contents of perchloric acid (%) in the substance: (1) 0.6, (2)
0.4, (3)0.0, and (4)0.19.

The pressure dependence o f the rate o f H A P combustion for various H C 1 0 4

concentrations over the stoichiometry is presented in F i g . 18.3.3. If the


concentration o f H C 1 0 4 is as high as 0.1%, H A P is incapable o f stationary
combustion. W i t h the increase o f the H C 1 0 concentration above this value
4

or with its decreasing to zero, as well as for excessive concentrations o f


N H O H (in this case, the complex with H A P is formed), the combustion
2

rate increases and the limits o f stable combustion are broadened (the regions
where combustion is absent are shown by dotted lines). The construction o f
the dependence o f the combustion rate on the concentration o f these
products at P - const leads to the curve with a minimum. In this case, the
influence o f the base is considerably stronger. The dependence o f the
combustion rate on the concentration o f perchloric acid and hydroxylamine
is in good agreement with the results o f the kinetic studies (Chapter 15) and
is indicative o f a particular role o f the decomposition reaction in the liquid
phase on H A P combustion.
The characteristic feature o f onium salts, which follows directly from
the mechanism considered (see F i g . 18.3.2) and most clearly manifests itself
in the case o f H A P , is the existence o f the range o f combustion instability
(pulsations, extinction) within the intermediate range o f pressure and,
consequently, the two combustion limits in addition to the third one, which
is located at l o w pressures and caused by heat losses. For H A P , the region
o f combustion instability is within the pressures o f 30-100 atm and for A P ,
it is within 100-300 atm. The addition o f A P in the amount o f 10%) and
larger makes H A P capable o f stable combustion within the pressure range
o f 5-120 atm. The complex o f H A P with hydroxylamine burns, starting
from 0.1 atm, and has rates o f combustion which are higher by
284 Combustion of Pure Substances: Reactions in the Condensed Phase

approximately a factor o f 3. In all cases, the pressure dependences o f the


temperature coefficient are expressed by the curve with a m i n i m u m , w h i c h
in the case o f H A P and its complex is located at approximately 40 atm; for
mixtures with A P this m i n i m u m is shifted towards the higher pressures.
Calculations o f the combustion characteristics o f H A P , carried out
based on the kinetic parameters by formula (18.3.5), showed a good
agreement with the experimental data (Table 18.3.3). The rates o f
combustion were measured with the help o f the samples, w h i c h were
pressed into the tubes made from organic glass with an inner diameter o f
0.7 cm up to the relative density o f approximately 1. The burning was
conducted inside a combustion bomb o f constant pressure in a nitrogen
atmosphere. The rates o f combustion were measured by the pressure
increase with the help o f a pressure gauge. The temperature o f the
combustion front was measured by the thermocouple method.

Table 18.3.3. The characteristics of H A P combustion at a pressure of 20 atm

Charac­ r, cm/s >7v n cr- 10 , grad"


3 1
8
teristic

Calcu­ 0.21 830 0.50 0.48 3.2 1.35


lation
Experi­ 0.28 780±60 0.50 3.5 close to
ment unstable
mode

Thus, in the case o f H A P , both the peculiarities caused by the kinetics o f


decomposition in the liquid phase and the characteristic phenomena
following from the combustion theory manifest themselves most clearly as
the m i n i m u m in the curves o f o(P) and discontinuity o f the r(P)
dependence.
The latter phenomenon, caused by combustion instability within a
particular region o f pressures and connected with the competition o f
decomposition and evaporation, was most conspicuous on combustion o f
organic perchlorates. Analysis o f the literature data [18.12, 18.16] showed
that a particular regularity is observed in the variation o f the size o f the
region o f instability and, correspondingly, the combustion limits as a
function o f the pressure coefficient n, characterizing the degree of
instability. A s can be seen from F i g . 18.3.4a, the region o f instability
broadens as the n value increases and the second pressure limit increases
more considerably than the first one, which is in qualitative agreement with
theory ( F i g . 18.3.4b). It is interesting to note that trimethylammonium
perchlorate, w h i c h is incapable o f dissociation and which does not belong to
the class o f onium salts considered, drastically differs from them both by
the level o f the combustion rate and by the absence o f extinction.
Thermal Decomposition and Combustion 285

300

200

100

0
n a

Fig. 18.3.4. The dependence of the limits of a stable combustion on the pressure
coefficient n. The lower branch is the first limit, the upper branch is the second
limit, (a) is the experimental dependence for the organic perchlorates and (b) the
calculation.

Ammonium nitrate (AN), NH4NO3

The basic features o f thermal decomposition and combustion o f this


substance have much in common with A P . The primary stage of
decomposition is also a proton transfer with the formation o f an acid and
base. The following reactions result in the formation o f nitrogen monoxide,
water, and other products. Simultaneously, evaporation o f H A proceeds
accompanied by dissociation into ammonia and nitric acid. However, the
heat o f thermal decomposition o f H A is lower and the thermal stability is
higher than in the case o f A P . Similar to A P , decomposition o f A N is
catalyzed by acids and metal compounds. In this connection, in contrast to
A P , the burning o f A N is not so efficient. A c c o r d i n g to the data o f
G l a z k o v a [18.19], the burning o f A N with additives o f halogenides o f
alkali, alkali-earth, and heavy metals starts from a pressure o f 10-50 atm
and A N with additives o f chromates and bichromates is capable of
deflagration at atmospheric pressure. A particular similarity with A P is also
observed for the r(P) dependence: at rather high pressures (approximately
400 atm) an area o f weak pressure dependence o f the combustion rate
appears, which represents the plateau or a decrease o f the rate (an unstable
mode), and, then, in some cases rate growth is observed once again.
The question arises as to whether decomposition o f A N in the
condensed zone could provide the level o f observable rates o f combustion
or whether the gas-phase reaction is responsible for it. Calculations o f the
combustion characteristics [18.20] were conducted in the form o f the
simplest evaluative variants: for the non-catalyzed thermal decomposition
o f A N in the liquid phase (see Chapter 12 and [18.21]) and decomposition
of N 0 in the gas phase [18.22]. It turned out that the gas-phase reactions
2

lead to a combustion rate which is lower than the experimentally measured


one by two orders o f magnitude, and thermal decomposition gives values o f
286 Combustion of Pure Substances: Reactions in the Condensed Phase

combustion characteristics which are in satisfactory agreement with


experiment (Table 18.3.4).

Table 18.3.4. The characteristics of A N combustion under the pressure 70 atm

Charac­ r, cm/s 7v A£ , cal/g


r
n 8
teristic

Calcula­ 0.34 790 0.51 147 1.04 3.4


tion
Experi­ 0.3-0.6 «600 0.8-0.9 unstable
ment combustion

The results presented suggest that the mechanism o f H A combustion is


close to those for A P and H A P , and the quantitative discrepancy is
connected with the difference in the heats o f decomposition and in the
kinetic parameters.

Hydrazinium nitrate (HN), N2H5NO3

Thermal decomposition and combustion o f hydrazinium nitrate are


distinguished by a number o f characteristic features. Within the pressure
range o f 3 0 - 1 2 0 atm, the burning rate o f H N is governed by the following
law [18.23]:

Kcm/s) - 0.033-P 08 2
,(atm).

The additives o f potassium,'sodium, lithium nitrates and potassium chloride


considerably increase the burning rate, but influence insignificantly the
lower limit o f stable burning. Within the range o f medium pressures, the
promotion effect increases in the series: K N 0 3 -> N a N 0 3 -> L i N 0 . The
3

concentration dependence o f the relative increase o f the burning rate o f


potassium nitrate at various pressures is shown in F i g . 18.3.5.
The dependence is expressed by the curve with a m a x i m u m , and the
promotion effect decreases as the pressure increases. In this case, the
m a x i m u m is shifted towards the range o f lower concentrations.
Calculations carried out based on the kinetics o f decomposition o f H N
in the liquid phase and the experimental data (Table 18.3.5) are in good
agreement and demonstrate an insignificant contribution o f the gas-phase
reaction to the heat balance on the combustion surface.
Thermal Decomposition and Combustion 287

J i » »
10 20
a,%

Fig. 18.3.5. The dependence of a relative increase of the combustion rate of the
mixtures of H N with potassium nitrate on the additive concentration under various
pressures: (1) 40, (2) 80, and (3) 130 atm.

Table 18.3.5. The characteristics of H N combustion under the pressure 50 atm

Charac­ r, cm/s 7s, K 7v A Q


* %
n G- 10 ,
3

teristic c(T -T ) s 0 grad -1

Calcula­ 0.70 882 0.68 0.90 2.5


tion
Experi­ 0.82 860 5 0.82 3.1
ment

A strong promoting influence o f metal salts on H N combustion, which is


also typical o f other similar oxidizers [18.19, 18.24], is a very surprising
fact because the assumption about the mechanism o f this effect caused by
the catalysis is not confirmed by the experiments on thermal decomposition.
For explanation o f the promoting effect the following hypothesis was
proposed: during combustion, the accumulation o f nonvolatile salt additive
proceeds on the combustion surface which should result in a decrease o f the
pressure o f vapors o f the main volatile component (onium salt), in an
increase o f the temperature o f combustion surface, and, correspondingly, in
an increase o f the burning rate.
Analysis o f the content o f potassium ions in the surface layer o f
approximately 200 jum thick after extinction o f the burning samples showed
that the mean concentration o f potassium in it is higher than in the parent
sample by approximately a factor o f 3. Taking into account that the size o f
the reaction layer is smaller by one order o f magnitude, the actual
concentration o f potassium in it should be even higher. The thermocouple
measurements revealed the increase o f the temperature o f the combustion
surface o f H N with 0.5% o f K N 0 3 by 85°C in comparison with a pure
substance.
288 Combustion of Pure Substances: Reactions in the Condensed Phase

Let us assume that the pressure o f saturated vapor o f H N with the


additive is expressed by the following equation, in which the effect o f
saturation o f the additive on the surface is taken into account:

P*=a- N
b N
) P

where TV is the concentration o f the additive in the parent substance, is


the concentration o f the additive on the surface, P v is the vapor pressure
without the additive, b is the coefficient.
W i t h the help o f this expression, and the equation o f heat balance, in
w h i c h the correction for the heating o f the additive was introduced, the
dependences were calculated, which made it possible to explain the
experimental data.
1. The existence o f the m a x i m u m on the curve o f the dependence o f the rate
increase on the concentration o f the additive.
2. The decrease o f the promoting effect, the shift o f the maximum towards
the lower concentrations, and decrease o f the power o f the dependence r(P)
with the pressure increase.
3. The increase o f the effect with the decrease o f the molecular weight o f
the additive.
Thus, the character o f the influence o f a chemically inert additive is
caused by two factors acting in opposite directions: by the decrease o f the
vapor pressure and, consequently, the degree o f evaporation o f substance,
and by the decrease o f the thermal effect o f the reaction in the condensed
phase (per one gram o f the mixture) owing to dilution.

Hydrazinium chloride (HQ, N2H5CI

This compound, w h i c h is not an oxidizer, but is a model substance, clearly


demonstrates the general and specific features o f combustion o f onium salts
[18.16, 18.25]. The ultimate pressure, from which H C deflagration is
possible, depends on the sheath diameter in which the powder o f substance
was pressed: at 0.7 cm, it can run as high as 50 atm, and, at 1 cm, it is 30
atm. The pressure dependence o f the rate is extremely weak. The increase
o f the initial temperature gives rise to the growth o f the burning rate,
weakly influences the lower pressure limit, and slightly increases the power
o f the r(P) dependence. The thermocouple measurements o f the combustion
front temperatures showed that almost all heat is released in the condensed
phase: after the temperature T is attained, a further increase o f only 4 0 -
s

7 0 ° C is possible. It is interesting to note that in the gas phase periodic


oscillations o f the temperatures are observed, which can attain the value o f
approximately 100°C. After the substance was burned out, a white porous,
Thermal Decomposition and Combustion 289

sometimes fused product remains in the sheath. This is ammonium chloride


( A C ) . The weight o f the residual increases with the pressure increase and
comes to 5 2 % at 60 atm and 6 1 % at 110 atm.
A t 80 atm, the overall reaction, determined from the analysis and
balance ratios, can be expressed by the following equation:

N H Cl->0.55NH4Cl
2 5 solid +0.45NH4Cl eva p+0.22NH +0.38N +0.17H
3 2 2

Thus, the scheme o f the combustion processes is as follows:

N H C 1 -> N H C 1 + N H , N , H
2 5 4 3 2 2

i t i t
N H 2 4 + HC1 N H + H C 1 . 3

In contrast to A P , the leading stage o f the reaction is connected not with the
anion, but with the cation part ( N H ) . 2 5

In accordance with this scheme, the balance heat and mass relationships
can be modified in the following way:

P = P +P*
V + P =(\ + £ l
^ L
) P + P*

7 i = 7 i - ^ + ^ ( l - 7 v ) - ^ v - ^ 7 ; ( l - ^ v ) .
c c c c

where Py and Py are the pressures o f saturated vapors o f hydrazinium


chloride and ammonium chloride, r/y and r/y are the fractions o f
evaporated H C and A C , and a = M /M . AC HC

The combustion characteristics, calculated based on the kinetic and


thermodynamic data (Chapter 15), revealed a satisfactory agreement with
the experimental data on the r(P) dependence. A s regards the pressure
coefficient n, in this case the discrepancy is significant (Table 18.3.6).

Table 18.3.6. The characteristics of HC combustion under pressure of 100 atm and
temperature of 293 K

Charac­ T, K
m 7v 7v* n cr • 10 3

teristic cm/s grad -1

Experi­ 0.14 680 760 0.41 «0 4.5


ment
Calcula­ 0.26 800 810 0.057 0.365 0.93 2.7
tion
290 Combustion of Pure Substances: Reactions in the Condensed Phase

The main reason is connected with the unstable character o f combustion (a


small reaction heat, the heat losses, the temperature variations at the front),
w h i c h is confirmed by calculations: the stability criterion, which within the
stable range should be less than 1, at T = 283 K has a value o f 2.5. A s
0

follows from the aforesaid, for the majority o f onium salts in the mode o f
instability, the plateau or the decay o f the burning rate appears in the r(P)
dependence. It is also important that the increase o f T enhancing the 0i

stability o f H C (S= 0.8), increases the value o f v u p to approximately 0.2.

Ammonium dinitramide (ADN), NH [N(N02)2j 4

This compound can be quite competitive with the ordinary oxidizers o f the
A P type either by its energy characteristics or by ecological properties
[18.26]. The data on combustion o f this substance were obtained by us in
1976 and could be published only recently [18.27].
Samples o f this substance, pressed into a tube made from PMMA,
burned at pressures up to 100 atm in a bomb o f constant pressure, equipped
with a pressure gauge and photorecorder for combustion rate measurement,
and at P < 1 atm in the glass device. In the latter case, the rate was
determined from the measurement o f the burning time o f a sample with a
fixed size.

P.atm * 3C,cm

Fig. 18.3.6 Fig. 18.3.7

Fig. 18.3.6. The pressure dependence of the rate of A D N combustion for various
sample diameters: (1)5 mm and (2) 7 mm.

Fig. 18.3.7. The temperature distribution through the combustion front of A D N : (1)
0.66 atm and (2) 20 atm.

The pressure dependence o f the combustion rate for the diameters o f the
sample o f 5 and 7 mm is shown in F i g . 18.3.6. It is characterized by the
Thermal Decomposition and Combustion 291

pressure coefficient value o f approximately 0.7 at P < -20 atm, and above
this pressure, the combustion rate becomes a pronounced function o f the
sample diameter.
A t P < 20 atm, the temperatures o f the combustion wave, measured by
tungsten-rhenium thermocouples o f 5 urn in diameter ( F i g . 18.3.7), are
relatively low, and, at larger pressures, a high-temperature zone appears.
Chemical analysis o f combustion products at sub-atmospheric pressures
showed that the amounts o f N O , N 0 , N 0 , and H N 0 2 2 3 remain constant,
N H 4 N O 3 , N H 4 N O 2 , and NH4N3O4 decrease with the pressure increase, and
the amounts o f N 2 and H 0 , determined from the balance, increase. A t a
2

pressure o f 0.66 atm, the reaction o f combustion o f A D N can be represented


by the following equation according to the data available:

N H N 0 - > 0.412NH NO + 0.0004NH NO + 0.007NH N O +


4 3 4 4 3 4 2 4 3 4

0 . 3 3 8 N O + 0 . 8 6 8 N O + 0 . 3 3 1 N 0 + 1.178H 0 + 0 . 6 9 8 N .
2 2 2 2

Introduction o f ammonium nitrate into A D N , which is the combustion


product, results in a decrease o f the combustion rate, which is already
noticeable at small concentrations o f the additive. The inhibition effect
decreases at high pressures. The additive o f C u 0 (2%) results 2 in a
insignificant increase o f the combustion rate at 1 atm; however, this effect
decreases with the pressure increase and, at 70 atm, the effect disappears.
The combustion rate with the K C r 0 2 2 7 additive is lower than for a pure
substance (by approximately 2 0 % at 40 atm). The mixture o f A D N and
aluminum (20%) burns more slowly than A D N at 1 atm by approximately a
factor o f 2, but slightly faster at 5-70 atm. It is interesting to note that in
spite o f the increase o f the r value (by approximately 4 0 % at 30 atm), the
flameless combustion proceeds, but in the case o f a pure substance, a
luminous zone is observed.
Small additives o f organic fuels (hydrocarbon polymers or plasticizers)
decrease the rate o f A D N combustion, especially at low pressures. For
instance, the decay o f r is as large as approximately 3 times at 1 atm on
addition o f 1-5% o f fuel.
The results obtained are indicative o f a considerable role o f thermal
decomposition o f A D N in the condensed (liquid) state. One o f the pathways
o f the decomposition o f A D N decomposition, as well as other onium salts,
is the dissociation o f the substance into the acid and base, which is followed
by the oxidation-reduction reaction in the condensed or gas phase.

N H N ( N 0 ) <-> N H + H N ( N 0 ) -> products.


4 2 2 3 2 2

A l o n g with this, another pathway o f A D N decomposition is possible (see


Chapter 14), through the decomposition o f the anion, which dominates
under particular conditions, for instance, at accumulation o f water during
292 Combustion of Pure Substances: Reactions in the Condensed Phase

the subsequent reactions:

N ( N 0 ) ~ -> N N ( V + N 0 .
2 2 2

The heat released during A D N decomposition in the condensed zone o f


combustion is consumed on the heating o f the substance and evaporation o f
the decomposition product, i.e., ammonium nitrate. However, as estimates
showed, the heat is insufficient for a complete evaporation o f A N .
Thus, this product abandons the combustion surface both in the form o f
vapor (NH3+HNO3) and in the form o f fine particles. Note that A N virtually
does not decompose in the condensed zone o f A D N combustion, because its
rate constant is smaller than those for A D N by four orders o f magnitude.
For quantitative estimation o f the combustion characteristics o f A D N at
low pressures, when the role o f heat inflow from the gas phase is
insignificant (approximately 3%, according to the experimental data), let us
consider the following model.
The leading process is thermal decomposition o f A D N in the liquid
phase with the rate constant o f the initial stage and the heat o f incomplete
decomposition. In this case, the temperature o f the combustion surface is
determined by the pressure o f the saturated vapor o f A N . The mass and heat
balances on the combustion surface are determined by the following
equations:

P = (l + — )Bexp(-E /RT ) w s ,
7v

c c

where a and z are the ratios o f the molecular weights o f A N and A D N and
their gasification products, correspondingly.
The following magnitudes o f the parameters were accepted:
F o r decomposition o f A D N : A = 1 0 1 4 4
, s" , E = 35.50 kcal/mol, Q = 280
1
r

cal/g (see Chapter 14). For decomposition o f A N : B = 1 0 , atm; £ 7 1


v
=
19
950 cal/mol, A = 10~ , cal/cm-s-K, p = 1.8g/cm , a = 0.695, z = 0.5.
3 3

The results o f calculation and the experimental data, which are in good
agreement, are presented in Table 18.3.7.
A strong dependence o f the combustion rate on the sample diameter
(Fig. 18.3.6) is indicative o f the unstable character o f combustion, because
the heat losses considerably influence the combustion process exactly in
this mode as compared with the stable one [18.28]. A t P = 2 0 - 4 0 atm, the
calculation o f the criterion 8 (see 18.3) gives the value 8= 1.4, which falls
into the region o f instability. In this respect, A D N is similar to A P and the
other onium salts.
Thermal Decomposition and Combustion 293

Table 18.3.7. The characteristics of A D N combustion

P, atm r, cm/s n

calc. exp. calc. exp. calc. exp.

0.20 0.27 600 560 0.84 0.7


20 1.5 1.82 710 700 0.82 0.7

The proposed mechanism o f A D N combustion is also confirmed by the


character o f the influence o f additives. In particular, C u 0 , w h i c h is a2

typical catalyst o f decomposition, promotes the combustion o f A D N . A t the


same time, A l and A N inhibit its combustion, because they absorb the heat
for heating and endothermic evaporation. However, a pronounced inhibiting
effect o f small concentrations o f the additives o f organic substances cannot
be explained only by the thermal factor. A s shown by the kinetic studies,
carried out by us with the help o f a differential calorimeter, the same
additives inhibited the decomposition o f A D N , and the rate o f oxidation o f
various polymer fuels by nitric oxides demonstrated the inverse dependence
on the combustion rate o f A D N mixtures with these fuels. This paradoxical
effect can be interpreted in the following way. A t the earlier stages o f
combustion, the products o f decomposition o f A D N (the acids and nitric
oxides) are partially absorbed by the fuel and react with it. The
concentration o f these products, which catalyze the A D N decomposition,
decrease in the liquid phase o f the oxidizer. A s a result, both the rate o f
decomposition o f A D N and the combustion rate o f the mixture decrease. A t
the low pressures o f burning, the influence o f a competitive process, i.e.,
oxidation o f the fuel in the liquid or gas phases by the products o f A D N
decomposition, is weaker due to their low reactivity.
Note that a series o f our results on A D N combustion, in particular the
conclusion about the leading role o f the liquid-phase decomposition in the
mechanism o f combustion, is in agreement with the results o f the work
recently published by A . E . Fogelzang, et al. [18.29].
In conclusion, let us consider the case o f onium salt combustion, w h i c h
illustrates the w e l l - k n o w n statement that there are no rules without
exceptions. The calculation o f the combustion rate o f hydrazinium azide
showed that its value is smaller by several times than the experimentally
measured one, which is caused by a high volatility o f the substance. One
can assume that, in this case, the leading stage o f the process is located in
the gas phase. However, this statement requires additional studies, including
the study o f the kinetics o f the gas-phase reactions.
A s a result o f the analysis o f combustion o f onium salts o f various
acids, the following conclusions can be formulated:
1. In the mechanism o f combustion o f these substances, the reaction o f
exothermic decomposition in the condensed phase and evaporation
294 Combustion of Pure Substances: Reactions in the Condensed Phase

(sublimation), which are the natural limiting factors and regulators o f the
degree o f conversion and the surface temperature, play an important role.
2. The main characteristics o f combustion, i.e., the rate, the degree o f its
dependence on the pressure and the initial temperature, and the combustion
stability, can be described quantitatively or qualitatively within the
framework o f this mechanism.
3. Dispersion (with the destruction o f the solid phase or with the foaming o f
the liquid phase) plays a secondary role in this mechanism, which under
particular conditions, is expressed in the additional increase o f the degree o f
pressure dependence o f the combustion rate.
4. Apparently, the gas-phase reactions in the mechanism o f combustion o f
onium salts play a minor role. Their influence is discussed in detail in the
following chapters.

REFERENCES

18.1. B . V . N o v o z h i l o v , Dokl. Akad. Nauk SSSR, 141(1): 151-153


(1961).
18.2. K . K . Andreev, Thermal Decomposition and Combustion of
Explosives, M . : N a u k a , 346 p. (1966).
18.3. V . A . Strunin, Zh. Fiz. Khim., 39(2): 433-435 (1965).
18.4. V . A . Strunin, G . B . Manelis, A . N . Ponomarev, and V . L . Tal'rose,
Fiz. Goreniya Vzryva, 4(4): 584-590 (1968).
18.5. B . L . Crawford, C . Huggett, and J. J. M c B r a d y , J. Phys. Coll.
Chem., 54: 854-862 (1950).
18.6. E . I. M a k s i m o v and A . G . Merzhanov, Dokl. Akad. Nauk SSSR,
157(2): 412-415 (1964).
18.7. G . B . Manelis, G . N . Nechiporenko, A . V . Raevskii, et a l . ,
Combustion of Boron-Based Solid Propellants and Solid Fuels,
Pennsylvania State University: C R C Press, 348 p. (1993).
18.8. Y a . B . Z e l ' d o v i c h , Zh. Eksper. Tekhnich. Fiz., 12(11/12): 4 9 8 - 5 2 4
(1942).
18.9. G . B . Manelis and V . A . Strunin, Combustion and Flame, 17(1):
69-77 (1971).
18.10. V . A . Strunin and G . B . Manelis, Fiz. Goreniya Vzryva, 11(5):
797-799(1975).
18.11. B . V . N o v o z h i l o v , Nonsteady Burning and Combustion Stability of
Solid Propellants, Progress Astron. Aeron., 143: 601-641 (1992).
18.12. V . A . Strunin and G . B . Manelis, Fiz. Goreniya Vzryva, 7(4): 4 9 8 -
501 (1971).
18.13. T . L . Boggs, AlAA Journal, 8(5): 867-873 (1970).
18.14. J. Wenograd and R. Shinnar, AlAA Journal, 6(5): 964-966 (1968).
18.15. G . B . Manelis and V . A . Strunin, Proc. 11th Int. Symp. on Space
Technology and Science, Tokyo, 97-104 (1975).
Thermal Decomposition and Combustion 295

18.16. G . B . Manelis, V . A . Strunin, and Y u . I. Rubtsov, Archiwum


Termodynamiki i Spalania, 6: 51-58 (1975).
18.17. C . Price, T. Boggs, and R. Derr, AJAA Paper, 219: 1-12 (1978).
18.18. V . A . B o k i i , V . A . Strunin, G . B . Manelis, et al, Fizika Goreniya
Vzryva, 15(4): 55-59 (1979).
18.19. A . P. G l a z k o v a , Catalysis of Combustion of Explosives, M.:
Nauka, 263 p. (1976).
18.20. V . A . Strunin, A . P. D ' y a k o v , L . B . Petukhova, and G . B . Manelis,
Proc. Zel'dovich Memorial in Combustion, Detonation, Shock
Waves, Int. Conf. on Combustion, M . : 2: 218-221 (1994).
18.21. M . W . Beckstead, 26th JANAF Combustion Meeting, Pasadena: 4
C P I A N 5 2 9 : 255-268 (1989).
18.22. E . S. Fishburne and R. Edse, J. Chem. Phys., 41(5): 1297-1304
(1964).
18.23. A . F. Zhevlakov, V . A . Strunin, and G . B . Manelis, Fiz. Goreniya
Vzryva, 12(2): 185-191 (1976).
18.24. T. L . Boggs and D . E . Zurn, Combust. Sci. Technol, 4: 227-232
(1972).
18.25. A . F. Zhevlakov, V . A . Strunin, and G . B . Manelis, Fiz. Goreniya
Vzryva, 10(2): 185-191 (1974).
18.26. G . B . Manelis, Environmental Aspects of Rocket and Gun
Propulsion, A G A R D C P 559, Norway, 1-13 (1994).
18.27. V . A . Strunin, A . P. D ' y a k o v , and G . B . Manelis, Combustion and
Flame, 1 1 7 : 4 2 9 ^ 3 4 (1999).
18.28. K . G . Shkadinskii and B . I. K h a i k i n , Gorenie i Vzryv, 3th All-
Union Symp. on Combustion and Explosion, M ' . : Nauka, 104-109
(1972).
18.29. A . E . Fogelzang, et al., Proc. 28th Int. Annual Conf. ICT,
Karsruhe, (1997).
Chapter 19

COMBUSTION OF PURE SUBSTANCES:


REACTIONS IN THE GAS PHASE

Combustion o f highly volatile substances can be described based on the


Belyaev-ZeFdovich mechanism. According to this mechanism, the
temperature o f the combustion surface is set to be equal to the boiling
temperature o f the substance and the chemical reaction proceeds within the
narrow gas-phase zone at temperatures which are close to the m a x i m u m .
The combustion rate is described by the well-known approximate formula:

r 2 = 2N\Ap?(RT>/E) A xp(-E/RT )
N+]
e m

Q (T -T f
m 0

The pressure and temperature coefficients take the following form:

n = N/2; a = E/2RT*.

However, there are only a few substances for w h i c h combustion follows


such simple regularities. The main reason is connected with a complex
kinetics o f chemical reactions in the gas phase. For this reason, the
formation o f the heat release zone can be caused by the influence o f several
processes and can be located in the region which is located rather far from
the maximal temperature zone.
The parallel and concurrent reactions are among these typical
processes. Destruction o f a substance and its oxidation, the catalytic and
non-catalytic reactions, interactions o f a fuel with various oxidizers or
interaction o f an oxidizer with various fuels and some others are examples
o f such processes. Let us consider two parallel reactions, proceeding with
various thermal effects [19.1]:

297
298 Combustion of Pure Substances: Reactions in the Gas Phase

A->C,
A-»D,

The expression for the combustion rate can be written in the following
form:

2A
](QiW +Q W )dT
c p\T T)
} 2 2 9
2 2
h

where Wis the reaction rate and T is the combustion temperature.


b

It is assumed that the activation energies o f the reactions do not differ


significantly from each other, so that the areas o f heat release are close to
each other in space. The temperature at the combustion front, which is not a
constant value, can be obtained in the following simple way. Integrating the
equation o f mass transfer over the whole space for each reaction, we w i l l
obtain that the degree o f decomposition is proportional to the reaction rate.

//, = - )w dx = -W,(T ); = - \W dx s -W {T )
rz r r• r
x h % 2 2 b

At 'o

F r o m this, one can obtain:

0 , W (T ) + Q W (T )\_
} h 2 2 b

W,{T ) W (T ) c b + 2 b

Since the reaction rate depends on the temperature exponentially, this


expression is a transcendental equation, from which the combustion
temperature can be derived, which is a function o f the kinetic and
thermodynamic constants and the pressure.
Let us analyze the following simplest case. It is assumed that the first
reaction is not pressure dependent and the second one is proportional to it.
Then, we w i l l have:

Q2 k g, kE T
n= •
2 2 2 0

2(0 k +Q
i 1 k) 2 F(k,+k Yc 2 2(2, *,+S 2 k )RT
2 b
2
T (T -T )
b b a

The temperature coefficient takes the following form:


Thermal Decomposition and Combustion 299

g, *, E Q ]+ 2 k 2 E 2 T0

r (r -r )_'
G -
2(2,
— + —

T - T F k )RT
2 b
2
b b 0
i
b -'o r

where

F = 1 | * i k (Q -Q )
2 { 2 (E -E ) x 2

(k +k ) c
} 2
2
R T b
2

The variation o f the combustion characteristics in a dimensionless form is


presented in Figs. 19.1 and 19.2 for the following parameters: s = 10, s x 2 -
10, q = 0.2-4, 0 = 0.3, and (p = 1, where s = Ec/RQ
O 0 U 0 = cT /Q
O 0 u ft =

Fig. 19.1 Fig. 19.2

Fig. 19.1. The calculated pressure dependence of the combustion temperature for the
parallel reactions at various values of the thermal and kinetic parameters: (1) q =
Q,/Q = 2, £ , > E , (2) q = Q /Q = 2, £ , = E , (3)q=\,E >
2 2 x 2 E , (4)q=\,E = E, 2 ] 2 l 2

(5) q = 0.5, £, - E , (6) q = 0.5, £ , > £ , (7) q = 0.2, £ , > E , (8) g = 4, £ , < £ .
2 2 2 2

Fig. 19.2. The calculated pressure dependence of the combustion rate for the parallel
reactions. The designations are the same as in Fig. 19.1.

The combustion temperature decreases with the pressure growth at q < 1,


increases at q > 1, and remains constant at q = 1. The combustion rate at q =
1 monotonically increases and the pressure coefficient increases from 0 to
0.5, which corresponds to the transition o f the combustion mode from the
reaction o f the zero order in pressure to the reaction o f the first order. In the
case o f q > 1, the combustion rate increases and the n value passes through
a m a x i m u m , which exceeds the n value o f each o f the reactions. The smaller
the E]-E 2 value, the larger this overshoot. In the case o f q < 1, the
300 Combustion of Pure Substances: Reactions in the Gas Phase

dependences under consideration can have a m i n i m u m . The larger E\—E 2

value, the deeper this m i n i m u m for the n value. This is caused by the
competition o f two factors such as the thermal (Q, T ) and the kinetic (W oc
b

P) ones, w h i c h can act in one or in opposite directions.


For q = 1 and E = E , the temperature coefficient is constant, for q = 1
x 2

and E\ > E , it decreases with pressure growth, for q > 1, it decreases, for q
2

< 1 and E] = E , it increases, and for q < 1 and E\ > £ , it.has a m a x i m u m .


2 2

If E\ is not equal to E and for the particular q values, multiple values o f


2

the combustion rate can correspond to the unstable combustion modes


within a particular pressure range ( F i g . 19.2, a dotted line) that was first
shown in [19.2].
Thus, the analysis o f even such a simple scheme o f parallel reactions
results in a complex and non-trivial appearance o f the dependences o f
combustion characteristics on pressure and, first o f a l l , in the arising o f
extrema, w h i c h are connected with the variability o f the combustion
temperature.
The consecutive reactions, as well as the parallel reactions, are typical
o f combustion o f gas and condensed systems. Such chemical processes as
nitrogen reduction ( N 0 2 - » N O - » N ) , consecutive transformations
2 of
carbon ( C -> C O -> C 0 ) , chlorine ( C l
2
+ 7
-> C l + 4
-> Cl° -> CI" ) and others
1

are connected with them.


The simplest case is the following reaction scheme:

A n a l y s i s o f the limiting combustion modes for the essentially different


kinetic parameters was carried out in [19.3]. It was shown that the
combustion process could be determined by only one o f the following
stages:
1. The leading stage is the first stage at the temperature o f this reaction.
The rate o f the second stage is very low and its zone is significantly
separated from the first one.
2. The leading stage is also the first stage at the combustion temperature,
w h i c h is determined by the heat release in both stages. The rate o f the
second reaction is very high and the zone corresponding to it is located
near the first one.
3. The leading stage is the second stage at the temperature, w h i c h is
determined by both stages. The first reaction proceeds very rapidly, but
with an insufficient thermal effect, and, thus, represents a dependent
process.
In all these cases, the characteristics are unambiguously determined by
the parameters o f only one stage. However, the study o f so-called m i x e d
modes, when interaction o f the stages manifests itself both on the kinetic
level (closeness o f the kinetic constants) and on the thermal level
Thermal Decomposition and Combustion 301

(contribution o f each stage to the effective combustion temperature), is o f


theoretical and applied interest. The solution o f the problem in a general
formulation is connected with particular mathematical difficulties. Let us
analyze one particular case in the approximate formulation. W e w i l l assume
that the rates o f both stages are comparable, but the heat release o f the
second stage is considerably higher than that during the first stage, which,
o w i n g to it, cannot provide a self-contained stationary development o f
combustion.
Let us express the combustion rate in the following form:

" " \IW {PJ ) x h + \IW {PJ )'2 h

The contribution o f the second stage to the heat balance depends on the
ratio o f the rates in the following way:

b 0
c \ + (WJW ) 2 c '

This gives:

At jp, - > o , r - > 0 .

A t
^->oo, (or^ ->0),
2 r =r + ^ '
b 0

c
and r —» 0 (in accordance with the assumption formulated).

At W 2 ^00,7; = T +Q /c +0 { Q /c2

and r oc .

Let W = A^expi-E^RT);
x W = AP *
2 2
N
exp(-£ /RT). 2

Then, the pressure coefficient takes the form:

clRT h
2
\ + y- 1
RT l
b
z
+ y

Et-EtQtJi+r- ) 1

y(\ + y- )]
1+
r

where y W IW 2 V

A s follows from these formulas, i f the rate o f the second stage depends
on pressure stronger than the first one, depending on the ratio o f kinetic
constants, the following types o f the curves are realized: a continuous
growth o f n or a dependence with a maximum and further decrease o f n. If
302 Combustion of Pure Substances: Reactions in the Gas Phase

N\ > N 2i then, together with a monotonic decreasing or increasing o f n, the


dependence with a m i n i m u m is possible.
Thus, the characteristic feature o f complex reactions on combustion is
the variation o f the effective combustion temperature on variation not only
o f the thermal, but also the kinetic parameters o f the reaction, which results
in the extreme pressure dependences o f r, n and cr in the cases o f parallel
and consecutive reactions studied.
The classical example o f complex reactions is a monomolecular
decomposition in the gas phase. In this case, under combustion conditions,
the kinetic nonequilibrium o f the process is observed:

A + M—^—> A* + M
A* + M—^—> A + M
A*—products,

where A denotes the molecule o f the parent substance, M is the neutral


molecule or A molecule, A is the activated molecule.
This characteristic feature o f a monomolecular reaction was taken into
account in the theory o f combustion o f gases and volatile substances and
when the results o f calculations were compared with experiments.
Let us consider the influence of a complex mechanism of
monomolecular reaction on the combustion characteristics [19.4, 19.5]. The
equations, w h i c h describe the combustion process, take the following form:

-m — + W + W + — W = 0,
x 2 3

c dx dx c c c
2
p D ^ - - m — -W x +^ =0,2

dx 1
dx
2
p D ^ - - m — +W x -W -W =0,
2 3

dx 1
dx

where Wx = a-a-^-k^xpi-EfRT), W 2 = a-b'p -k exp(-E /RT),


2
or 2 W 3 =
a-b-p-koyexpi-Ei/RT), a is the concentration o f A , b is the concentration o f
A , m is the mass velocity, a is the fraction o f A in the parent system.
In accordance with the theory o f monomolecular reactions and kinetic
studies, we w i l l consider that the first stage (activation) has a rather high
activation energy, whereas deactivation only slightly depends on the
temperature (usually, k 02 oc (7) ) and decomposition into the
N
products
proceeds with a zero activation energy. W e w i l l also suppose that the main
heat release proceeds at the third stage during the formation o f the final
products and within a narrow zone. This makes it possible to apply the
analytical method for the solution o f this problem: the method o f conjugated
Thermal D e c o m p o s i t i o n and C o m b u s t i o n 303

asymptotic expansions with subdivision o f the temperature range into the


regions with various asymptotic behavior, such as the internal region (the
reaction zone) and the external region (the heating zone). Representing the
functional characteristics connected with the temperature, concentrations,
and their spatial gradients in the form o f asymptotic expansions in terms o f
small parameters and restricting our consideration to the zero and first order
approximations, and taking into account the conditions o f the conjugation o f
zones, after integration, we w i l l obtain the following relationships:

2 a X (RT>) (PIRT fk m x k,
c (T -T ) {
m 0
2
E )k (P/RTJ
x i+ k 2

The same result can be obtained i f one assumes the quasi-stationary value
of the intermediate concentration and performs integration by the
Zel'dovich-Frank-Kamenetskii method under condition E\ > 0, E = E = 0. 2 3

The dependence o f the combustion rate on the pressure and the initial
temperature w i l l take the following form:

n =- 1+ - ; G -
2[ \+k 2 k~\PIRTJ_

From this, we w i l l obtain:


A t />->0,>?-> 1,
A t P -> OC,A7 -> 0.5.
The pressure for which n = 0.75 and, consequently, which is connected with
the most pronounced variation o f n, is equal to:

P*=k, k' R]
T.
m

Let us consider from the standpoint o f these results the experimental


data on thermal decomposition and combustion o f some compounds. The
decomposition o f hydrazine was studied behind the shock waves within the
wide temperature and pressure ranges under conditions close to combustion.
The calculation o f the combustion characteristics carried out based on these
kinetic parameters showed their good correlation with the experiment both
for the gas and liquid hydrazine [19.6]. Thus, under the pressure o f 50 m m
H g , the calculated value o f the combustion rate o f hydrazine vapors is 117
cm/s and the experimentally measured value is 116 cm/s. The pressure P (n
= 0.75) is approximately 20 atm (determined from the calculations) and
approximately 10 atm (determined from the experiments) (Fig. 19.3). The
temperature coefficient o f hydrazine is 2.9 • 10~ degrees" 3 1
(determined
from the calculation) and 3.0 • 10" 3
degrees -1
(determined from the
304 Combustion of Pure Substances: Reactions in the Gas Phase

experiment). Thus, the mechanism o f the monomolecular reaction o f


hydrazine decomposition makes it possible to explain the main regularities
o f combustion and, in this connection, it is not necessary to invoke the
hypothesis about the chain mechanism o f decomposition, which is
discussed in several works.

-1 0 1 2 3

Fig. 19.3. The calculated pressure dependence of the coefficient n\ (1) hydrazine and
(2) ethyl nitrate. The points are experimental data.

Let us consider the combustion o f ethyl nitrate. The combustion o f this


substance was studied in [19.7], and the kinetics o f decomposition was
investigated by the shock tube method in [19.8]. Comparison o f the
calculations and experiments gave the following results: at 35 m m H g , the
combustion rate o f vapors is 22.7 cm/s (determined from the calculation),
15 cm/s (determined from the experiment), P = 0.33 atm (determined from
the calculation) and P* = 0.2 atm (determined from the experiment).
Based on these results, one can assume that the non-equilibrium
character o f the kinetics o f monomolecular reaction also manifests itself on
combustion o f the other gas-phase and volatile compounds.

REFERENCES

19.1. V . A . Strunin and G . B . Manelis, Fiz. Goreniya Vzryva, 19(2): 8 9 -


94 (1983).
19.2. B . I. K h a i k i n and S. I. Khudyaev, Dokl. Akad Nauk SSSR, 245(1):
155-158 (1979).
19.3. B . I. K h a i k i n , A . K . Filonenko, and S. I. Khudyaev, Fiz. Goreniya
Vzryva, 4: 591-599 (1968).
19.4. G . B . Manelis, V . A . Strunin, and E . V . Lebedeva, Khim. Fiz.,
5(9): 1269-1276(1986).
19.5. G . B . Manelis and V . A . Strunin, 35th Congr. Int. Astronaut.
Federation, Lausanne, IAF-84-301 (1984).
19.6. A . P . Alekseev and G . B . Manelis, Fiz. Goreniya Vzryva, 16(4):
54-60 (1980).
Thermal Decomposition and Combustion 305

19.7. J. A . H i c k s , 8th Symp. (Int.) on Combustion, The Combustion Inst.,


487-496(1962).
19.8. I. S. Zaslonko, T. I. Kochergina, Y u . K . Mukoseev, et al., 2(8):
1060-1064 (1983).
Chapter 20

COMBUSTION OF PURE SUBSTANCES:


REACTIONS IN THE CONDENSED AND
GAS PHASES

The experimental facts indicate that, apparently, for numerous condensed


substances the combustion processes proceed according to a more
complicated mechanism than in the elementary combustion models
considered above. It incorporates various chemical reactions of
decomposition and oxidation in the condensed and gas-phase states, the
phase transition, and interaction o f combustion stages. The most important
question, w h i c h directly concerns the practice o f the control o f the
combustion characteristics, is the relative role o f the process in the
condensed and gas zones and their joint influence on the combustion rate in
the mode, w h i c h we w i l l call a mixed mode.
Let us consider the following scheme, which includes two parallel
reaction pathways [20.1, 20.2]:
1. The thermal decomposition in the condensed phase with heat release;
2. The substance sublimation and the gas-phase reaction of vapor
decomposition.
In general, this scheme reflects the combustion mechanism of
ammonium perchlorate.
This scheme can be described by the following set o f equations:
In the condensed phase (x < 0):

X ^--cm
c — +Q c W -Q c v W =0,
v
c
dx 2
dx c c v v

dx c

In the gas phase (x > 0):

307
308 Combustion of Pure Substances

8
dx2
dx ^ 8

H
dx
'» 12
dx 8

where m is the mass combustion rate, 77 and y are the degrees o f


decomposition in the condensed (index c) and gas (index g) phases, index
V denotes sublimation. It is assumed that the products o f decomposition in
the condensed zone are saturated by the vapors o f sublimate and, thus, the
rate o f sublimation W can be expressed through the decomposition rate in
w

the condensed phase W and the degree o f sublimation TJ :


c V

O n the combustion surface, the following relationship is true, which relates


the values o f T$, t]y, and the external pressure P:

P = P + P =(1+/- -^)BQxp(-E /RT ).


r w v s
(20.1)

Let us consider three possible combustion modes ( F i g . 20.1):


1. The rate o f the gas-phase reaction is low and it proceeds in the induction
mode, w h i c h is also called the self-ignition mode [20.3]. The determining
process is the decomposition reaction in the condensed phase.

Fig. 20.1. The scheme of the temperature distribution at the combustion front in the
(1) induction, (2) mixed, and (3) conjunction modes.

The known procedure o f separation o f the combustion front into two


zones, the zone o f heating, in which one can neglect the chemical source,
and the reaction zone, in which one can neglect the convective terms,
results in the following expression for the combustion rate (for the zero
order reaction):
Thermal Decomposition and Combustion 309

; ; ; 2 ^ 2^r /E )p ^ex (-y/?r )


s
2
c c P s

(1 - t],)[2c(T - T ) - Q (1 - 77 ) + Q tl }' s 0 c V v v

In the gas-phase zone, the reaction proceeds in the induction mode and,
consequently, in the vicinity o f T , one can neglect the first (conductive) s

term in the corresponding differential equation:

dT_
cm-
dx

and, then, we w i l l obtain another expression for the combustion rate through
the gas-phase characteristics, which gives an additional condition for
obtaining the T value. s

2 _ A Q (M /R'T )A Bexp[-(E
s g w s g g + E )/RT ]V S

YYl — •>
c[c(T -T )-Q (]-nv) s 0 c + Qrf ] v

where R* is the gas constant, expressed in atm-cm /grad-mol. 3

2. In the mixed mode, the reactions in both phases influence the combustion
process in nearly the same way. These reaction zones are separated by the
space interval o f the order o f the heating zone width in the gas phase. The
possibility o f the existence o f such a mode within a wide range o f variation
o f the external conditions is caused by the fact that the rates o f both
reactions increase with the pressure growth: in the condensed phase, due to
the suppression o f sublimation and growth o f the surface temperature, and,
in the gas phase, o w i n g to the ordinary mechanism o f the gas-phase
reactions. For the combustion rate, we have the following expression:

m2_ \
2
Q (M /R'T )(RT /E )B
g v b b
2
g A cxp(-E /RTJexpj-E^/RT )
& v b

3. In the conjunction mode, in which the gas-phase zone is adjacent to the


combustion surface, we have the following expression for the combustion
rate, which also takes into account the gas-phase reaction at the temperature
o f the combustion surface:

m 2 = IX&iMylRXXRTZlEJBA^ txp(-E /RT ) v s

[ c ( r - r ) - e ( l - 7 ) + C} ^ ]
s 0 c v v v
2

x[(exp(-V/?r )-exp(-£ //?r )]. b g s

It is convenient to calculate the combustion characteristics in the following


way. Let us accept 7^ as an independent variable. Equating the rates o f
310 Combustion of Pure Substances

combustion in both phases, one can find the value o f 77V, and then the values
o f m and P can be calculated. The pressure dependence o f the combustion
rate, calculated for various values o f the parameter cp, which is proportional
to the ratio o f the decomposition rates in the gas and condensed phases, are
presented in F i g . 20.2:

ej(p/pj

Fig. 20.2. The pressure dependence of the combustion rate in various combustion
modes: (s) the induction mode, (m) the mixed mode, (c) the conjunction mode; the
"cond" and "g" indices denote combustion in the condensed and gas phases. The
numbers designate the following: (1) (p = 0.1, (2) cp = 1, and (3) <p= 10.

A t cp = 0.1, within the range o f low pressures, the induction mode is


realized which, at P - » 0, asymptotically merges with the combustion mode
in the condensed phase. A t medium pressures, the mixed mode is settled,
and at high pressures—the conjunction mode, which at P -> 00, again passes
into the combustion mode in the condensed phase. The transition from one
mode to another is marked by circles at the places o f intersection o f the
curves. A t <p = 1, a qualitative picture o f the mode changing is preserved:
the region o f the mixed mode expands and the occurrence o f the induction
mode moves towards very low pressures. A t <p = 10, in the conjunction
mode, the m a x i m u m appears because the growth o f the combustion rate,
caused by the high rate o f the gas-phase reaction, gives way to a decay due
to the decrease o f the concentration o f sublimated vapor and transition to
the combustion mode in the condensed phase. For the same reason, the
extrema are formed both in the pressure dependences o f the temperature and
in the pressure coefficients.
Thermal Decomposition and Combustion 311

The combustion model presented was also studied by us by the methods


o f computer modeling. Comparison o f the results o f computer modeling
with the results o f the approximate solution showed a satisfactory
agreement o f the combustion rates within an error o f approximately 20%.
Thus, depending on the ratio o f the kinetic rate constants o f the
reactions in the condensed and gas phases, different structures o f the
combustion front and various combustion modes are possible: the induction,
mixed, and conjunction modes, in w h i c h the combustion characteristics
have various functional dependences on the determining parameters. In this
connection, let us consider critically a widely used concept o f "flame
standoff. A number o f models with a complex combustion mechanism
were developed in the literature based on this concept. Possessing a number
o f advantages, such as a relative simplicity and clearness, these concepts
demonstrate certain shortcomings, which restrict their application:
1. Actually, the dependence o f the width o f the standoff zone (the "dark
zone") x m on the combustion parameters is obtained based on the dimension
reasoning and, in this case, the relation between the kinetic and
thermodynamic constants o f the gas-phase reaction and the combustion rate
is too simplified, and, thus, a special analysis is required.
2. The dependence presented can be used within particular limits, when the
value o f x m corresponds to the width o f the heating zone by the order o f
magnitude and cannot be applied for the induction and conjunction
mechanisms.
3. The validity o f application o f the "flame s t a n d o f f concept for the
quantitative estimates of combustion characteristics needs further
examination by the solution o f the problem in a rigorous formulation and
with the help o f numerical calculations.
Let us analyze the experimental data on A P and nitroamines within the
framework o f the given combustion model with a complex mechanism,
including the reactions in the condensed and gas phases, and, also, in the
pores o f the substance. The calculations o f combustion characteristics o f A P
are based on the kinetic rate constants o f decomposition in the solid phase
and decomposition o f vapors o f perchloric acid with an activation energy o f
43 kcal/mol.
A typical picture o f the distribution o f the thermal characteristics at the
combustion front (the heat flow and the rate o f heat release), obtained from
the numerical calculations, is presented in a dimensionless form in F i g .
20.3. There are one maximum o f heat release from the reaction in the solid
phase and two peaks caused by the gas phase reaction: the first is inside the
pores o f the substance, and the second is at a certain distance from the
combustion surface completely in the gas-phase zone.
The rates o f combustion, calculated for the three modes mentioned
above, were found comparable with each other: for the purely condensed
phase combustion r = 0.05 cm/s (at P = 40 atm), for the gas-phase
combustion r = 0.025 cm/s, and for the mixed mode (the gas-phase and
312 Combustion of Pure Substances

condensed combustion simultaneously) r = 0.07 cm/s. These combustion


rates are lower than the experimentally measured rate (0.5 cm/s) by
approximately one order o f magnitude. The possible causes o f this
discrepancy were discussed in Section 18.3.

9 = rr-TjE/RT*

Fig. 20.3. Distribution of thermal characteristics of combustion at the combustion


front of A P : (1) the heat flux, (2) the rate of heat release in the solid phase, and (3)
the rate of heat release in the gas phase.

The multistage character o f transformation o f the components on A P


combustion is confirmed by the data o f mass-spectrometric analysis o f the
combustion products, carried out for a composite AP-based system at sub-
atmospheric pressures by O. P. Korobeinichev, et al. [20.4]. The m a x i m u m
concentration o f N H , C 1 0 , and 0 is attained inside the pores at a distance
3 2 2

o f approximately 0.05 m m but not at the combustion surface, the


concentrations o f the final products H C l and N O also intensively increase
inside the surface layer, and the concentration o f N 0 , as an intermediate
2

product, attains the maximum at a certain distance from the surface.


The pressure dependence o f the combustion rate o f A P within a wide
pressure range has a complex and unusual character. A c c o r d i n g to [18.13],
at pressures o f 2 0 - 5 0 atm, the value o f the coefficient n is approximately
0.8. In this case, the surface layer o f extinguished samples has the form o f a
"foam" (a solidified melt with bubbles).
A t pressures o f 7 0 - 1 4 0 atm, the dependence r(P) becomes weaker (n <x
0.5). Simultaneously, the character o f the combustion surface, which
consists o f cavities and convexities, is changed. Within the pressure range
o f 150-500 atm, the combustion is unstable (luminescence pulsation).
Finally, at high pressures, a pronounced growth o f the combustion rate is
observed, which is governed by a dependence close to those observed at
low pressures.
W e can give the following interpretation o f these facts. Apparently,
within the pressure range o f 2 0 - 1 5 0 atm, a variation o f the relative role o f
Thermal Decomposition and Combustion 313

the gas-phase reaction in the combustion mechanism takes place. A s


follows from the theoretical analysis (Section 19.3), the m a x i m u m in the
curve r(P) is possible for the consecutive and parallel reactions i f the
transition o f the combustion mode takes place from the reaction w h i c h has a
weaker pressure dependence (for A P , n = E /2E c y = 0.5), to the reaction with
a higher pressure dependence (probably, a bimolecular gas-phase reaction
with n = 1). In this case, the temperature o f the combustion surface can also
change with the pressure variation due to the changing o f the relative
contributions o f both processes. This results in the fact that the m a x i m u m
value o f the pressure coefficient can exceed the n values, calculated for the
single-stage modes. This conclusion is confirmed by the analysis o f the
literature data on combustion o f individual crystals o f A P and heated
pressed samples, which showed that in both cases the maximum in the n(P)
curve was obtained at a pressure o f 4 0 - 5 0 atm, which attained the value o f
approximately 1.2, then, at a pressure o f 80-100 atm, the decrease o f n up
to approximately 0.6 was observed.
A s for the combustion mechanism under higher pressures, it can be
connected with thermal instability (the plateau, the rate decrease) and with
the development o f the secondary exothermic reactions, caused by the
transformation o f nitrogen and chlorine into the final products (the rate
growth).
Apparently, such a complicated mixed combustion mechanism, which
includes the physical consequence o f the substance transformation (the
condensed and gas zones) and chemical consecutive and parallel reactions,
is typical o f many other onium salts—analogs o f A P . The other
characteristic feature o f their combustion mechanism is the unstable region,
w h i c h manifests itself to a certain extent within various pressure ranges
almost for all compounds.
Let us consider from the same point o f view the combustion o f
nitramines (RJDX and H M X ) , which are the most important components o f
propellants. There has been a great number o f publications on the
combustion mechanism o f these compounds and theoretical development o f
combustion models. A s a rule, in such works, the chemical processes in the
condensed and gas phases are o f equal importance. In [20.5], M .
Benreuven, et al. carried out the mathematical modeling of R D X
combustion based on the exothermic decomposition o f the substance in the
liquid layer, evaporation, the gas-phase decomposition o f R D X , and on the
reaction o f interaction o f C H 0 and N 0 . Then, in [20.6], the two-region
2 2

combustion model for combustion o f H M X with a large number of


chemical reactions was developed. The main result o f these works is the
conclusion about a relatively small fraction o f decomposition o f nitramine
in the liquid state (approximately 10%). A t the same time, in [20.7], the
process o f evaporation o f these compounds on combustion was not
considered at all. In [20.8], the model o f "foam reaction" in the liquid layer
with a large number o f chemical reactions in both phases was applied to
314 Combustion of Pure Substances

describe the combustion mechanism o f R D X . The conclusion was made


about the valuable fraction o f decomposition o f substance in foam, which
was as large as approximately 50%. In [20.9, 20.10], detailed studies o f the
chemical and thermal structure o f the combustion zone o f nitramines were
carried out. The fast thermolysis technique and spectroscopy method have
been successfully employed in [20.11] to establish the relationship o f
chemical reactions in the surface zones to combustion characteristics o f
nitramines. Some results obtained were used by theoreticians for validation
and confirmation o f their models.
Thus, we can conclude that for cyclic nitramines, the combustion mode,
when the contributions from the processes proceeding in the condensed and
gas-phase zones into the total value o f the combustion rate are equal, is also
typical.

REFERENCES

20.1. V . A . Strunin, A . N . Firsov, K . G . Shkadinskii, and G . B . Manelis,


Fiz. Goreniya Vzryva, 13(1): 3 - 9 (1977).
20.2. V . A . Strunin, A . N . Firsov, K . G . Shkadinskii-, and G . B . Manelis,
Fiz. Goreniya Vzryva, 22(1): 4 0 - 4 7 (1986).
20.3. A . G . Merzhanov and A . K . Filonenko, Dokl Akad. Nauk SSSR,
152(1): 143-146(1963).
20.4. O . P. Korobeinichev and A . G . Tereshenko, Acta Astronautica, 6:
385-390 (1979).
20.5. M . Benreuven, L . H . Caveny, R. J. Vichnevetsky, and M .
Summerfield, 16th Symp. (InU) on Combustion: Combustion Inst.,
P A , 1223-1233 (1976).
20.6. N . S. Cohen, G . A . L o , and J. C . C r o w l y , AlAA Journal, 23(2):
2 7 6 - 2 8 2 (1985).
20.7. A . B i z o t and M . W . Beckstead, Flame Structure, Novosibirsk:
Nauka, 1 : 2 3 0 - 2 3 5 (1991).
20.8. Y . - C . L i a u and V . Y a n g , J. Propul. Power, 11(4): 729-739 (1995).
20.9. O . P . Korobeinichev, L . V . K u i b i d a , and V . Z h . Madirbaev, Fiz.
Goreniya Vzryva, 20(3): 4 3 - 4 6 (1984).
20.10. A . Zenin, J. Propul. Power, 11(4): 752-758 (1995).
20.11. T . B . B r i l l , J. Propul. Power, 11(4): 740-751 (1995).
Chapter 21

COMBUSTION OF CONDENSED
COMPOSITE SYSTEMS

21.1 Quasihomogeneous Composites


If the particles o f an oxidizer are o f a rather small submicron size, the
condensed zone can be considered as a quasihomogeneous one with respect
to the temperature distribution through the transverse direction. In this
approximation, let us consider the combustion model o f a composite
propellant, consisting o f ammonium perchlorate and polymer binder [21.1].
The main processes, proceeding in the condensed zone, are the following:
1. Thermal decomposition and sublimation o f the oxidizer.
2. Destruction o f the fuel.
3. Oxidation o f the fuel by the products o f decomposition o f the oxidizer.
In this model, the influence o f the gas-phase reactions can be taken into
account by the assignment o f a certain value o f the additional heat in the
thermal balance on the combustion surface.
Equations, describing the combustion process, are the following:

X^L- m^L Q^-Q W ^Q,W,=Q


ax ax
C + 2 2

dx dx dx

The boundary conditions are as follows:

x= -oo: rj = cp = j = 0; T=T 0

x = 0: Tj j ;
= r s <p = <p ; j = j \s s T=T S

The first integral o f the first equation gives the following heat balance in the
condensed phase:

315
316 Combustion of Condensed Composite Systems

c(T -T )
s 0 = Q j - Q <p + QJ
iT s 2 s S + AQ • g

Using the Zel'dovich approximation, one can obtain the following


expression for the combustion rate:

2*](QW-Q lV +Q lV )dr
2 2 3 3

The values o f r/ , q>s, and j s s are the degrees o f transformation o f the


components in the reactions o f decomposition o f an oxidizer, destruction
and oxidation o f a fuel. They have to be derived from the additional
conditions.
W e w i l l consider that the leading process is exothermic decomposition
o f the oxidizer, whose fraction in a propellant is, as a rule, high (80-90%).
Then, <p$ and y can be found from the ratio o f the corresponding kinetic
s

equations and one can obtain the approximate expressions, in which these
characteristics depend on the ratio o f the kinetic constants on the
combustion surface:

The degree o f transformation o f an oxidizer is restricted by its sublimation


and dispersion.

7s = l - 7 v - % ; P = (\ + l
-^)B xp(-E /RT )
Q w s .
>7v

Let us consider the simplest cases o f the zero order reaction:

W =p A
] c l e x p ( - £ , / RT ); S W = pA
2 c 2 e x p ( - £ / RT );
2 S

W =p A exp(-E /RT )
3 g 3 3 s .

For a partial oxidation o f the fuel, within the region o f the condensed zone,
one can take into account the diffusion o f the oxidizer (in the dissolved and
gas-phase states) with the help o f the following well-known formulas:

W =SpjD A Qxp(-E /RT ),


3 c 3 2 s (for the melting fuel)
Thermal Decomposition and Combustion 317

, (for the solid fuel) >

where k D oc £) , S is the specific surface o f the particles, D


g c and D are the
g

diffusion coefficients in the condensed and gas phases.


The qualitative analysis o f these expressions allows one to make the
following conclusions:
1. The increase o f the decomposition rate o f the oxidizer causes effects o f

two kinds: the increase o f the combustion rate, because m oc JW ] , and the
decrease o f the combustion rate as a result o f the decrease o f the fraction o f
oxidized fuel, becausey sW /W In the limiting case W << W when the
x
3 h 2 h

oxidizer decomposes so rapidly that the fuel has no time to be oxidized in


the condensed zone and the heat inflow from the gas phase cannot
compensate the heat losses on the heating and destruction o f the fuel, the
combustion rate can decrease at the introduction o f the fuel in such a
propellant system (Fig. 21.1.1).

8r

0
i 2 3

Fig. 21.1.1. Dependence of the combustion rate on the decomposition rate of the
oxidizer: (1) the composite system and (2) the oxidizer.

2. The increase o f the gasification rate o f the fuel (destruction, evaporation)


leads to the decrease o f the combustion rate because these processes
proceed with heat absorption and, in addition, as a result o f competition,
decreases the oxidation level o f the fuel ( T decreases).
s

3. The increase o f the oxidation rate o f the fuel in the condensed zone leads
to the increase o f the combustion rate, mainly due to the temperature Ts

increase. There are several known ways o f increasing W : the increase o f


3

the rate constant (a catalysis), the specific surface, and the concentration o f
oxidizer (the pressure). Since the rate o f oxidative pyrolysis o f the fuel
318 Combustion of Condensed Composite Systems

increases, as a rule, similarly to the rate o f destruction, it is appropriate to


use a sufficiently thermostable fuel, but containing active groups, which can
be rapidly oxidized and with a high thermal effect. It should be noted that
the efficiency o f the factors promoting oxidation is considerably higher in
the condensed phase, where it is necessary to oxidize a very small amount
o f the fuel in order to increase the value o f Ts, in comparison with the gas
phase, from which the heat inflow is inversely proportional to the
combustion rate.
The influence o f the kinetic parameters o f the reactions considered on
the pressure and temperature coefficients is presented in Figs. 21.1.2 and
21.1.3. The value o f n for a pure oxidizer decreases with the pressure
increase due to sublimation suppression. In composite systems, it increases
i f the fuel oxidation proceeds with a higher activation energy than the
decomposition o f the oxidizer, and it decreases in the opposite situation. In
the case o f fuel destruction, the value o f n decreases, but the degree o f its
decrease also depends on the ratio o f £ and E\.
2

Fig. 21.1.2 Fig. 21.1.3

Fig. 21.1.2. The pressure dependence of the coefficient n: (1) the oxidizer, (2)-(5)
the composite propellants ((2) £ > £ , , (3) £ < £,, (4) £ > £,, (5) £ < £ , ).
3 3 2 2

Fig. 21.1.3. The pressure dependence of the temperature coefficient of the


combustion rate. The designations are the same as in Fig. 20.2.

For the temperature coefficient, the presence o f the m i n i m u m in the a(P)


curve is typical. The depth and width o f the " w e l l " or its disappearance are
also determined by the corresponding variations o f the kinetic parameters.
Let us consider the particular experimental observations within the
framework o f the quasi-homogeneous model. In cooperation with D ' y a k o v ,
we studied model composite uncured propellants, containing ammonium
perchlorate (< 50 um), plasticized butyl rubber ( B R ) , and various ferrocene
catalysts [21.2].
Thermal Decomposition and Combustion 319

Study o f the kinetics o f thermal decomposition o f these composites


determined from the heat release on a differential automatic calorimeter
showed the complex character o f the process. There are two stages in w h i c h
ammonium perchlorate and ferrocene catalyst interact, i.e., the ferrocinium
salt is formed on the surface o f A P particles and the bulk catalytic oxidation
o f the fuels (the organic part o f the catalyst and plasticized rubber) occurs.
The results obtained on the kinetic rate constants and thermal effects o f the
overall process o f decomposition o f the composite system were used to
calculate the combustion characteristics (Table 21.1.1).
It turned out that the first rapid stage o f decomposition has enough time
to be entirely completed at the initial regions o f the combustion front,
probably already in the heating zone, which influence is equivalent to the
increase o f the initial temperature by an appropriate value AT 0 =Q\/c. The
second stage proceeds in the main reaction zone, but the degree of
transformation in this stage is restricted by A P sublimation.

Table 21.1.1. The characteristics of combustion of the composite system AP(85%)


+ BR(15%) + 3% dimethylferrocene (above 100%) at the pressure 40 atm

r, cm/s n 7i,K c(T -T )


s 0 77r

calc. exper. calc. exper. calc. calc. calc.

1.1 1.9 0.46 0.40 1000 210 0.28

A n illustration o f a good correlation between theory and experiment is the


dependence o f the relative increase o f the combustion rate on the
concentration o f the alkylferrocene catalysts ( F i g . 21.1.4). In this case, a
specific character o f a complex kinetics, connected with the two-stage mode
o f the process and the chemical structure o f a catalyst, manifests itself in
full measure on the combustion characteristics.
The results obtained were used to reveal the factors which provide a
high efficiency o f ferrocene catalysts. It turned out that the content o f iron
in a compound and the structures o f an organic part were among these
factors. It was observed that the efficiency factor divided by the iron
content in the catalyst increases in the following series o f substitutes in the
ferrocene ring: alkyl, carbonyl, hydroxyl, nitrile, aldehyde group, fulvene
group, alkyne, and alkene, thus characterizing the activity o f the functional
groups in the condensed zone o f combustion.
The correlation between the thermal effects o f the first and second
stages o f thermal decomposition and the pressure coefficient was also
observed. Thus, the coefficient n decreases with the increase o f the ratio
Q\/Qi and the dependence o f n on the concentration o f the catalysts has the
m i n i m u m in the experiment and calculation.
320 Combustion of Condensed Composite Systems

Fig. 21.1.4. Dependence of the increase of the combustion rate on the catalyst
concentration at 40 atm (the experimental data (left) and the calculated data (right):
(1) dimethylferrocene, (2) diethylferrocene, (3) diisopropylferrocene.

The dependence o f n(P) for these composite systems is o f a rather complex


character (Fig. 21.1.5). The formation o f the maximum on the curves and its
comparison with the dependence n(P) for A P deflagration (although, in the
case o f a composite, it is shifted towards lower pressures) indicates that the
leading process is the transformation o f A P . The complex character o f the
n{P) dependence correlates qualitatively with the theoretical curves
presented in F i g . 21.1.2. However, certain peculiarities o f the behavior o f n
within the range o f high pressures (the m i n i m u m and a subsequent growth)
require the use o f other combustion processes, which w i l l be discussed
below.

Fig. 21.1.5. Pressure dependence of the coefficient n for the composite systems
based on A P and butyl rubber: (1) without a catalyst and (2) with the ferrocene
catalyst.
Thermal Decomposition and Combustion 321

Note also that introduction o f the ferrocene catalysts gives rise to a decrease
o f n at pressures P < 40 atm. The decrease is especially considerable near a
pressure o f 1 atm.
A n obvious relation between the regularities o f thermal decomposition
and combustion was revealed for many other inorganic and metalloorganic
catalysts and inhibitors. In particular, there was a correlation between the
inhibiting action o f lithium fluoride, especially in combination with lithium
perchlorate, on the combustion rate and decomposition rate o f composite
systems.
In the case o f potassium permanganate, which actively promotes
decomposition o f A P in the solid phase, a certain ambiguity o f its action
was revealed. A t isomorphic introduction o f K M n 0 4 into the A P crystals,
the combustion rate decreases. A t the same time, i f it is deposited on the
surface o f A P particles, the combustion rate increases, which was explained
by us as the competition o f two factors: the increase o f dispersion (in the
first case) and the promotion o f decomposition o f A P from the surface (in
the second case).
The pressure dependence o f the temperature coefficient for the model
composite propellant, which has a m i n i m u m , also correlates qualitatively
with the calculated curves (Fig. 21.1.3). In this case, introduction o f H M X
as a cooling agent (the combustion rate decreases) into this composition
naturally gives rise to the increase o f a in accordance with theory.
Thus, the model o f combustion o f a quasihomogeneous fuel, based on •
the processes in the condensed zone o f burning, to a certain extent can
explain qualitatively, and sometimes quantitatively, the observed
regularities and peculiarities of combustion, and give certain
recommendations on purposeful variation o f the combustion characteristics.

21.2 Layered Systems


For better understanding o f the combustion mechanism of composite
propellants, in research practice, the so-called layered combustion models
are widely used: a "sandwich", i.e., the system, which consists o f layers o f a
fuel and oxidizer, ordered in the longitudinal direction with respect to the
combustion front, and a "chemical arc", i.e., the system, which consists o f
plates or cylinders o f a fuel and oxidizer, which are pressed to each other by
their end-faces. In the gap between these-surfaces combustion occurs.

il
Chemical arc'

This system was first studied experimentally by W . H . Andersen, et al., in


[21.3] and by A . S . Shteinberg in [21.4]. In the latter work, it was proposed
to call it "a chemical arc".
322 Combustion of Condensed Composite Systems

The theoretical analysis o f this system was carried out by us in [21.5,


21.6] for the case o f a heterogeneous reaction o f oxidation o f the solid fuel
by the gaseous product o f oxidizer decomposition.
The heat and mass transfer in the system are o f a complex character,
w h i c h can be revealed by analysis o f the dependence o f the Nusselt number
on the Reynolds number. A s shown by S. S. Rybanin in [21.6], within the
range o f the Reynolds numbers smaller than 1, the dependence o f N u =
F(Re) is weak and N u oc 1. Thus, we w i l l study the specific character o f the
phenomenon, with respect to various modes o f the process in a simplified
formulation, assuming N u = const, and also representing the gradients in the
following form:

~ = ( \-<*I)IXQ\
A
^ - = (T -T )/X 2 X 0 ,
ox ox

where x is the distance between the components.


0

Comparison o f the results o f calculation in an approximate formulation


with the results o f calculation in which the gas dynamics was taken into
account showed that the difference between them was mainly o f a
quantitative but not qualitative character.
A s it follows from the similarity o f the fields o f concentrations and
temperatures, the relation between the concentration o f a gas-oxidizer and
the temperature is expressed by the following relationship:

a =C x +C T, 2

where C\ and C are proportionality factors.


2

Using the boundary conditions, we w i l l obtain the temperature


dependence o f the concentration on the fuel surface in the following form:

n _YQ2 + rQi-(v / +
r) c
(T -T )
2 0

where Q and Q are the heats o f oxidizer gasification and fuel oxidation,
x 2

correspondingly, ^/and y are the stoichiometric coefficients.


The combustion rate o f the fuel and oxidizer take the following form:

m = A e, (-E fRT )^^ ^-^


2 P 2 P 2 2
+
^
+
- T
^ ,
Qi -Q\
Q -c(.T -T )
2 2 0

c(T -T )-Q
} 0 }

The distance between the surfaces is expressed by the following formula:


Thermal Decomposition and Combustion 323

0
pA expi-E /KT )[
2 2 2 y Q +y 2 Qi~(V+r)c (T -T )][Q -c(T -T )]
2 0 2 2 0

The value o f T\ can be found from the following transcendental equations:

1. In the case o f a sublimable oxidizer:

v Q ^ r g , - ( ^ ) c ( 7 ; - r 0 ) = ( B / / ? ) e x p ( _ £ v I R T ^

2. In the case o f a decomposing oxidizer:

W Qi +r Q\ - + r M ^ - ^)] g - ^ 2 2 - ) =

4 exp(-£j/fl^)
yOy4 e x p ( - £
2 2 /RT ) 2

Finally, the relation between the gap width and the pressing force F
between the components can be found from the following relationship:

F= UP-P )27rydy0 =- ^ ,
o 2
Po x

where y 0 is the radius o f the sample.


Let us analyze the results obtained. For the case o f a sublimable
oxidizer and heterogeneously oxidable fuel the calculated characteristics are
presented in F i g . 21.2.1 for the following values o f the parameters:

cE /RQ 2 2 =2.75; cE IRQ


y 2 = 0.6; cE IRQ { 2 \n(BI p) = 0.2;

e /&=-(U;
v cr /e =0.05;0 2 ^ = 1; r = 3.

A t small values o f x , the surface temperatures o f both components are


0

low and close to each other. The rates are low and the ratio o f the rates is far
from the stoichiometry. The oxidizer plays the role o f a "cooler", which
removes the heat from the fuel for heating and gasification. Under such
conditions, the process develops in a slow kinetic mode.
After the critical condition is achieved, the process passes through an
intermediate unstable region in the diffusion mode, which features a sharp
temperature increase o f the fuel surface, approaching the maximal value,
and a decrease o f the oxidizer concentration.
324 Combustion of Condensed Composite Systems

0*2

0.^

x /x:
a

Fig. 21.2.1. The calculated dependence of the combustion characteristics on the


distance between the oxidizer and fuel in the "chemical arc" system for the
heterogeneous oxidation: (1), (3) the surface temperature and the combustion rate of
the fuel, (2) the surface temperature of the oxidizer, and (4) the ratio of the
combustion rates of the components.

Later, the rate o f the fuel regression decreases, and, at large values o f x 0> the
process is governed by a purely diffusion law:

First, the ratio o f the rates increases and, then, tends to a constant value.
The parameters o f sublimation o f the oxidizer exert an influence in the
following way. The surface temperature increase o f the oxidizer and an
adequate increase o f the external pressure result in an increase o f the
regression rate o f the fuel, and the increase o f the heat o f sublimation gives
rise to a decrease o f the rate. In both cases, the combustion diffusion mode
is observed to be achieved faster.
In the case o f a decomposing oxidizer, the characteristics o f the process
are determined by the ratio between the kinetic rate constants o f both
processes w h i c h restricts the possibility o f the existence o f stationary
modes. The surface temperature o f the oxidizer does not remain constant as
in the previous case, but can immediately decrease or pass through a
m a x i m u m . If the heat o f decomposition o f the oxidizer is positive and it is
capable o f deflagration ( A P ) , then the ratio o f the regression rates m /m\ in
2

the diffusion region decreases with the increase o f x , because in this case
0

m\ —> const and m - » 0.


2

The ratio o f the rates can also significantly differ from the
stoichiometric one, especially at x -> 0 and x - » oo. Thus, a considerable
0 0

part o f the gaseous oxidizer comes out from the interlayer into the
environment and remains unreacted, which is a specific feature o f this
system.
Thermal Decomposition and Combustion 325

In [21.7], we experimentally studied the layered system o f a "chemical


arc" type using a device in which the cylindrical columns o f the fuel and
oxidizer were pressed by their end-faces with the help o f a rod and a spring.
The pressing force applied to the samples was determined by the
compression o f the spring. T o fix the combustion zone, the column o f the
oxidizer was set against a thin tungsten needle. The displacement o f rods
and the electric light bulbs connected with them was recorded, in the course
of the burning o f the samples, on the film o f a photorecorder and, thus, the
combustion rates were measured. The experiments were conducted in a
combustion bomb o f constant pressure in a nitrogen atmosphere at pressures
o f 2 0 - 1 0 0 atm. The diameter o f the samples was 9 m m , and their height
was 10-15 mm.
Dependences o f the mass velocity o f combustion o f a pressed A P
(density, 1.88 g / c m ; particle diameter, 160-220 jam) and the columns o f
3

polymethyl methacrylate ( P M M A ) , polyethylene ( P E ) , and coal on the


thickness o f the gap between them, calculated by the formula for the case o f
a sublimable oxidizer, are presented in Fig. 21.2.2.

Fig. 21.2.2. The experimental dependence of the combustion rate of the components
on the distance between them: (1) and (V) A P + P M M A , (2) and (2') A P +
polyethylene, (3) and (3') A P + coal.

The curves have a maximum. The systems with coal have the highest
burning rates, and the lowest burning rates are observed for the systems
with P M M A . In the case o f polyethylene and coal, the deflagration rate o f
A P is higher than the combustion rate o f a pure oxidizer, and in the case o f
P M M A the deflagration rate is lower. The ratio o f the mass combustion rate
as a function o f the distance is also expressed by the curve with a
maximum. A n excess o f gasified fuel is observed in comparison with the
326 Combustion of Condensed Composite Systems

amount required according to the reaction, especially in the region o f small


values o f x . 0

The system o f a "chemical arc" type allows one to introduce the


combustion catalysts separately in each o f the components and, because o f
this, to reveal the mechanism o f their influence, in particular, the
localization o f an additive influence on the combustion process. W e used
[21.8] the following organometal compounds: diethylferrocene ( D E F ) and
epoxyethylferrocene ( E E F ) , which were introduced either into the fuel
components (butyl rubber, polyester, and polyethylene) or into the oxidizer
( A P ) . The catalyst was added mechanically or by dissolving in the fuel
before cure.
Dependence o f the rate o f regression o f the A P columns and cured
polyester ( P E S T ) , i.e., a copolymer o f tetrahydrofurane and propylene oxide
in the ratio 85/15, is shown in F i g . 21.2.3. The curves have a m a x i m u m ,
w h i c h is attained at the gap width o f 10-12 jam. Introduction o f E E F in the.
amount o f 0.6% into P E S T gives rise to an increase o f the combustion rates
o f both components in comparison with the system without a catalyst.

0 0.001 0.002
X , cm

Fig. 21.2.3. The experimental dependence of the combustion rate of the A P and
PEST system on the distance between the components at a pressure of 40 atm: (1)
PEST, (2) PEST + 0.6% EEF, (3) A P , (4) A P , and (5) A P without the fuel.

O n l y in the case o f small x are the combustion rates o f A P the same in both
cases and close to the combustion rate o f A P without fuel participation,
w h i c h is designated by a dotted line. The increase o f the combustion rate on
the introduction o f catalysts is larger by approximately 30%) at the
m a x i m u m for both components, and for larger distances it is 8 5 % for A P
and 70%) for P E S T + E E F systems. The ratio o f the mass velocities changes
considerably with the gap increase. A t small values o f x (approximately 8
urn), the value z = m/m ox ranges from 0.7 for the uncatalyzed pair to 0.8
for the catalyzed one, i.e., it considerably exceeds the stoichiometric ratio,
calculated for the complete burning, which is 0.13. A s the value o f x
Thermal Decomposition and Combustion 327

increases, the value o f z decreases and approaches the stoichiometric value:


a t x > 12 jim, it is as high as 0.35.
Comparison o f the experimental results with theory demonstrates a
qualitative correlation. Actually, a sharp increase o f the combustion rate up
to the ultimate value and the subsequent decay means the transition from
the kinetic mode to the diffusion one. In the case o f P M M A , whose
gasification proceeds with the heat consumption, the fuel plays the role o f a
peculiar cooler, absorbing the heat from A P for heating and gasification. A s
a result, the combustion rate o f A P by itself becomes lower. In the case o f
the system with coal, the situation is opposite. Here, the fuel is gasified as a
result o f a heterogeneous exothermic oxidation reaction and, thus, serves as
a heater with respect to P A , whose combustion rate becomes higher than the
rate o f deflagration o f this substance.
The mechanism o f the influence o f the organometal catalyst introduced
into the fuel can be represented in the following way. The active iron, w h i c h
is formed during decomposition o f E E F , initiates the reaction sites and
promotes the chain-radical reactions during destruction o f P E S T . A s a
result, its burning rate increases. A t the same time, iron particles, which
enter the gas-phase zone, promote the chemical processes occurring in this
zone. O w i n g to this, the heat flux towards the surface o f the oxidizer and,
correspondingly, the rate o f its burning increase.

A "sandwich type " system

For this layered system, we w i l l consider the case when the oxidizer is
capable o f exothermic decomposition and sublimation, and the fuel is
partially heterogeneously oxidized by the products o f decomposition o f the
oxidizer, and is partially destroyed and gasified. This problem was studied
in cooperation with A . N . Firsov and K . Y . Shkadinskii in [21.9, 21.10]. In
an effort to simplify the problem, all processes are considered as occurring
on the surface o f the components. The temperature and concentration o f a
gaseous oxidizer, as well as the kinetic characteristics are assumed to be
averaged within the boundaries o f each layer. Interaction between the layers
is accounted for by the introduction o f the heat and mass transfer
coefficients in the transversal direction both in the condensed and gas-phase
zones. Corresponding equations describing the processes take the following
form:
The gas phase (y > 0):
328 Combustion of Condensed Composite Systems

n da 2
l da ccp x

dy dy cph
r^d a da. ap
(a,-a ) = 0
1 g

" dy dy (\-(p)L
2
s 2

The condensed phase (y < 0):

dT dT 2
cc
A —~ - cm — ^ (7j -T ) = 0 ,
x x c

cn
— 2
^ 2

(7;-r ) = o
dy dy (\-<p)L
2
c 2

The boundary conditions (y = 0):

On the surface of the oxidizer:

dy s
dy

a m - pD
]S
Ity

a =P /P
]S w =B [exp(-E /RT )]/P, w ls

m - A e x p ( - £ /RT );
x T [S m =m- m v r

O n the surface o f the fuel:

8
* + 2oX OX-2d d: W W

da
a m-pD—± + jm
2

0x =0,
dy
2S

w
o x = Pg 2s4)x a e x
P(-^ox 7
); m
i = ' w

The conditions before the combustion front (y = - oo):

r , = r 2 = r 0 .

The conditions behind the combustion front (y = + oo):


Thermal Decomposition and Combustion 329

dy dy dy dy

The gas state equation:

P = RTp/M .

The designations are as follows: a c = NucAc/£, <2 = Nu /Lg/L, m is the g g

mass velocity, L is the thickness o f a "sandwich", and cp is the fraction o f


the oxidizer.
In the stationary mode o f combustion, the equality conditions o f the
mass burning rates o f the components are obeyed:

m - m + m;
] T y m =m 2 ox + m; d m -m -
x 1 m.

In this case, the decomposition rate o f the oxidizer (m ) and the oxidation x

rate o f fuel (m ) ox play the determining role. The rates o f sublimation (my)
and destruction (m ) are derived from the balance relationship.
d

The essence o f the solution o f the analytical problem reduces to the


following procedures. B y subtraction o f differential equations o f the heat
and mass transfer in the layers o f the components and subsequent
integration using the boundary conditions, one can derive the differences o f
the mean temperatures and concentrations on the combustion surface o f the
oxidizer and fuel layers.

2 [ g , (1 - ?7y ) - gy?7v + Q (1 - 77QX ) - S o X ^ O X ] / C


T - T
A

'IS 'IS" , ; / — ~i
1+ 4 N u
^ + 1+ 4 N U
^
<p{\ - (p)c L m 2 2 2
\ (p{\ - (p)c L m 2 2 2

2
(7v+77ox)
4Nu X 2

1 +J1 +
v <p(\-<p)c L m 2 2 2

After summation and integration o f the same equations, which were


preliminarily multiplied by <p and (1 - cp), correspondingly, one can obtain:

cp T ( 1 -<p)T = T
IS + 2S 0+ q , M - r i - Q M + t o x - a ( l - ^ x ) = ^
c c
9 a ls + (1 - <p)a = cp T] - (1 - q>)jij = a„.
2S v ox
330 Combustion of Condensed Composite Systems

Finally, we have a complex transcendental set o f equations. The following


algorithm turned out to be efficient.
The temperature 7*1 is accepted as the initial parameter for calculations.
s

Then, the value o f a s is derived. F r o m the balance relationships one can


{

express m and, excluding this parameter, obtain the quadratic equation with
respect to a s- 2 Finally, these two equations for the temperature and
concentration differences can be used to find the dependences F](L, T ) 2S
=
0
and F (L, T ) = 0. F r o m these equations, one can determine numerically the
2 2S

unknowns o f the problem. If necessary, one can find the spatial profiles o f
the values T a n d a in each o f the layers.
A l l parameters o f the given model can be subdivided by their functional
features. The internal chemical parameters characterize the kinetics, the
thermal-physical parameters characterize the heat and mass transfer. The
propellant properties (layer size, composition) should be referred to the
third group, and the external conditions (pressure and initial temperature)
have to be assigned to the latter group.
The values, corresponding to the oxidizer of AP type and
heterogeneously reacting fuel, were chosen as the parameters o f the first
and second groups. The calculated dependence o f the combustion rate on
the total "sandwich" size is presented in F i g . 21.2.4 for various ratios o f the
components. A t very small values o f L, the combustion mode corresponds
to the quasihomogeneous character o f composition: the temperatures and
concentrations in the layers are almost the same. W i t h an increase o f L, an
increase o f m is observed, which is more intensive at cp = 0.7 and 0.9.

Fig. 21.2.4. The calculated dependence of the combustion rate of a "sandwich" on


its size for various ratios cp of the oxidizer and the fuel: (1) cp = 0.9, (2) cp = 0.7, and
( 3 ) ^ = 0.4.

Let us call this portion o f the curve the first anomalous branch, because as is
accepted in all models o f combustion o f composite propellants, the
combustion rate can only decrease with an increase o f the characteristic size
o f the component. The anomaly is caused by the fact that in this range the
Thermal Decomposition and Combustion 331

imbalance o f T and a
s s in the layers increases. The temperature on the fuel
surface increases owing to the high exothermic effect o f heterogeneous
oxidation accepted in calculations, and the concentration o f the gas-oxidizer
decreases. However, on the whole, the oxidation rate o f the fuel, w h i c h
under given conditions is the leading combustion process, increases.
A t a further increase o f the L value, the dependence m(L) acquires the
ordinary typical character (the rate decreases) connected with the increase
o f the role o f the diffusion factor and with the decrease o f a s-
2

A t a certain size o f the layer, the values o f m, which are still decreasing,
become smaller than the values corresponding to the combustion o f a pure
oxidizer without fuel participation. The temperature o f the fuel surface is
lower than that o f the oxidizer surface, the heat flux is directed from the
oxidizer towards the fuel, and the oxidation level o f the fuel becomes low,
i.e., the fuel plays the role o f a "cooler" with respect to the oxidizer.
Finally, within the range o f a sufficiently high size, which is equal by
order o f magnitude to the thickness o f a heated layer, the second anomalous
branch appears (the heat exchange branch), which is connected with the
decrease o f the influence o f heat outflow from the oxidizer towards the fuel
up to zero and with the transition o f the process in the autonomous mode o f
combustion o f the oxidizer without the fuel participating (at L -> oo).
It should be noted that the combustion mode at the contact place o f fuel
and oxidizer, which can be realized in practice, remains beyond the
framework o f the averaged model considered. It can be analyzed within the
framework o f the model o f non-one-dimensional combustion, which w i l l be
considered below.

Fig. 21.2.5. The calculated pressure dependence of the combustion rate of a


"sandwich" for various values of its size, which increases from curve (1) to curve
(5).

The typical modes, which follow from the dependence m(L), manifest
themselves also in the pressure dependences o f the combustion rate ( F i g .
21.2.5): the kinetic mode at very small characteristic sizes (a high degree o f
dependence), the diffusion-kinetic region at medium sizes (reduction o f the
dependence, the appearance o f a plateau or a negative dependence), the heat
332 Combustion of Condensed Composite Systems

exchange mode at an increase o f the size (the appearance o f the region o f a


rapid growth o f m on P), and, finally, the return to the kinetic mode at large
sizes (an autonomous mode o f combustion o f a pure oxidizer).
Note that for systems with inert fuel, which is weakly oxidized at the
combustion front, the most typical mode is the transition from combustion
with a completely heated fuel to the autonomous mode o f combustion with
non-heated fuel at an increase o f pressure, which is accompanied by an
increase o f the degree o f the dependence m(P). A complex character o f the
dependence o f m on pressure and size gives rise to various forms o f the
pressure dependences o f the pressure and temperature coefficients: growth,
decay, and the appearance o f extrema.
Let us compare the results with the experimental data, but first, let us
make the following remarks. The first is concerned with the fact that the
majority o f the experiments have been carried out for disordered composite
systems with fine particles o f oxidizer and only a small number o f
experiments have been conducted for ordered systems with a relatively
large layer size. Nevertheless, there is evidence [21.11] that the combustion
rates o f ordered and disordered systems (the mixture of potassium
perchlorate with P M M A ) , by their absolute value and the character o f
dependence m(d), differ relatively weakly from each other.
The other remark is concerned with the model o f a limited, particular
character (a propellant with a thermostable fuel, which is capable o f
heterogeneous oxidation), whereas in the majority o f cases the composite
propellants are systems with gasifying polymer fuels. However, the
common factors, which are typical o f both systems, are taken into account
in the model: the inhomogeneity, the heat and mass exchange between the
components, and the pressure dependence o f the reaction rate o f the fuel
oxidation. Thus, keeping in mind the special character o f the model, we
considered a wider range o f literature data.
Let us enumerate some o f them, which correlate qualitatively with the
results obtained.
1. The m a x i m u m on the curve o f the dependence m(d) in the range o f small
sizes (the first anomalous branch o f a kinetic nature) was observed for non-
stoichiometric mixtures o f A P with polysterene [21.11].
2. The m i n i m u m in the range o f rather coarse particles (the second
anomalous branch o f the heat exchange nature) was repeatedly observed in
the experiments: for a layered system consisting o f A P and naphthalene
[21.12], a composite system with carbon, for which the m i n i m u m on the
curve shifts towards fine particles as pressure increases [21.13], and a
composite system with additives of aluminum, w h i c h reduced the
combustion rate in the following way: the smaller the particle sizes, the
stronger the influence on the combustion rate [21.14].
3. The types o f dependence m(P), which follow from the model considered,
were observed on combustion o f different condensed systems: mixtures o f
AP and potassium perchlorate with organic fuel (polysterene,
Thermal Decomposition and Combustion 333

polypropylene, polyester rubber, and some others) and with inorganic


(tungsten, aluminum) ones [21.11].
4. A complex form o f the dependence o f the temperature coefficient on the
pressure and the particle sizes (a monotonic variation or an extremum) was
observed for AP-based mixtures, potassium perchlorate with various fuels
and black powder [21.11].
Thus, the model using the averaging o f the characteristics in the layers,
in spite o f certain approximations and a restricted character, can explain a
series o f experimentally observed dependences, obtained on combustion o f
composite systems. However, it does not give a correct idea o f the
combustion front structure, which can be obtained only within the
framework o f the non-one-dimensional formulation o f the problem [21.10,
21.15].
As a zero approximation, we w i l l consider the initial combustion
surface to be plane and we w i l l find the distribution o f the temperatures and
concentrations over the combustion surface, and then on the basis o f
particular chemical reactions, we w i l l calculate the local rates o f the
reactions at each point and the shape o f the surface itself
Although such a simplified approach is o f a qualitative character,
nevertheless it can give certain information about the level o f the maximal
reaction rates, about the position o f the leading point, w h i c h leads the
combustion process, and about the comparability o f the results o f a given
approach and the results o f the model with averaging.
Thus, the equations describing the process are the following:
The gas phase (y > 0):

r
dT 2
d T^ 2
dT_
• cm 0,
dx dy dy
r ax
2 + 2

' o a da
2
3a_
^ + T
dx 2
dy 2 'dy'

The condensed phase (y < 0):

Q2 \
dT
2 T
dT_
-crn„ :0.
dx 2 +
*~
dy 22
dy

In order to simplify the analysis, on formulation o f chemical processes on


the surface, we w i l l restrict ourselves to the reaction o f exothermic
decomposition for the oxidizer, excluding its sublimation, and to the
heterogeneous reaction o f oxidation o f the fuel, excluding its destruction.
The boundary conditions (y = 0):
334 Combustion of Condensed Composite Systems

O n the surface o f the oxidizer (0 < x < (pi):

dT
dy

= am
>i mx
a /g
m = A e x p ( - £ / RT ).
r r r S
m

O n the surface o f the fuel (<pl<x< I):

m
ox = sP A>x
a
%
e x
P(-£ x 0
/RT
s)-

dT_ ar
=4
dy '*dy

8a
= -J m

dy
m

The conditions o f symmetry:

jc = 0 a n d j c = l : ^L = ^L = o.
dx dx

The conditions before and behind the combustion front:

8T dT da .
y = -oo: — = 0; y = +oo: — = — = 0.
dy dy dy

The following notations were accepted: m is the decomposition rate, m x ox is


the oxidation rate, / = L/2.
The solution takes the following form:

4(# + y ) s'm(k7T(p)cos(k7rx/l)

(
2knX %
V

1 + J1 +
\ c m
nJj

T = T 2(g o x -g )
r
sin( kncp) zos(knx II)
I -
cn f
2knX ^ c
+J1 +
\ ™Jcm
j

where, „=a<p-j(\-<p),
a T„=T 0 + &- +
i l < p ) Q o
\
Thermal Decomposition and Combustion 335

m-m -A x x e x p ( - £ /RT )
r S 9 (o < x < <p\) and

m =m =ox a^A^exp(-£ ox /'RT ),(<pl


s <x<l).

The surface profile is determined with the help o f the relationship


m - m m a x - c o s / ? , where J3 is the angle between the direction o f the
combustion front propagation and the normal at a given point o f the surface.
Since dy I dx - t a n ( / ? ) , one can obtain:

where the point with the maximal velocity is accepted as the zero point and
integration is carried out over both sides from this point.
The calculation procedure is as follows. The value o f i = m l max was
accepted as the initial independent parameter, and then the values o f a , T , s s

and m were calculated as a function o f £ = x/l (from 0 to I). The maximal


value o f m was found, w h i c h was considered as the rate o f propagation o f
the combustion front (m ). max Then, the layer size L = 2i/m max and the surface
profile £ = y/l = F(Q were determined.
The profiles o f the concentrations a , the temperatures 0
s S
=
cT*/Q , r

the linear rates o f decomposition r , the rates o f the fuel oxidation r x and
d 0

the surface shape ^ are shown in F i g . 21.2.6.

Fig. 21.2.6. Distribution of the characteristics on the combustion surface of a non-


one-dimensional "sandwich" model for various sizes: (1)1/ and (2) L , (L /L) = 5). 2 2

W i t h a small-sized "sandwich", one can-observe a weak variation o f these


characteristics, which corresponds to quasi-homogeneity. The maximal rate
o f the process, which, in this case, is equal to the oxidation rate ( r m a x ->
336 Combustion of Condensed Composite Systems

r x ) , is determined at the contact boundary o f the components. The surface


0

shape is formed by the two branches, one o f which (for the oxidizer) is
nearly straight and has a larger slope that corresponds to a lower and almost
constant burning rate.
A t larger size L, the variations o f the characteristics become stronger.
The maximal rate, as before, is realized at the contact o f the layers, but the
decomposition rate becomes higher, and the leading stage is changed ( r m a x

->r ).d

It should be noted that at certain ratios the maximal rate o f the process
(''max -> ^*ox) can be located not at the contact but at a certain distance from
it, w h i c h is determined by the additional condition:

dr
ox _ 1 das
dx a dx
s RTS
2
dx

A t a further increase o f the size o f a "sandwich", owing to a constantly


increasing diffusion difficulty o f the gas-oxidizer supply to the fuel surface,
its concentration and, correspondingly, the rate o f fuel oxidation decrease to
zero at a certain distance from the contact boundary. Thus, at large L, the
correction on incomplete combustion was introduced into the values o f
and which made it possible to a certain extent to compensate the
incorrectness o f the model.

5
o

Fig. 21.2.7. The calculated dependence of the maximal (at the contact boundary) and
mean rates of burning of the oxidizer and fuel on the "sandwich" size: "max" is the
maximal rate, "mean" is the mean rate, ox is oxidation, and d is decomposition.

The dependence o f the m a x i m a l rates o f burning o f the oxidizer and fuel on


the layer size and their comparison with the mean rates, which follow from
the model with averaging, are presented in F i g . 21.2.7. Qualitatively, these
dependences are similar: the predominant rate decreases and the leader
changes at a certain value o f L. The differences are as follows: in the t w o -
dimensional model the value o f m passes through the m a x i m u m , and, at L
d
Thermal Decomposition and Combustion 337

- » oo, both rates tend to a constant value; in the "averaged" model the
oxidation rate decreases to zero at a finite value o f L, and the decomposition
rate, passing through the m i n i m u m , at L-+ oo, tends to the combustion rate
o f a pure oxidizer.
In the case when oxidation o f the fuel proceeds with a small thermal
effect (the fuel is inert and requires heat consumption for heating), the
situation is qualitatively changed: the maximal rate r max is realized at the
center o f the oxidizer layer (x = 0), where the leading point is located. The
surface shape is different: there is a cavity at the center o f the oxidizer layer
with a sharp protuberance in the fuel layer.
The main goal o f the theoretical and experimental studies o f the layered
systems as elementary models is to determine the combustion mechanism o f
composite propellants, which, in the majority o f cases, represent a
disordered system o f particles o f an arbitrary shape introduced into the
matrix, i.e., the polymer binder or viscous fuel. Let us imagine the
composite system as a set o f ordered blocks o f one o f the components with
a characteristic size L x in the medium o f the other component with a
characteristic size o f interlayer L , and consider various interactions from
2

the point o f view o f the results presented.

a & c d e
R<R L R>R T r / r , r >r, r*r t

m H PPP
im m p pin

Fig. 21.2.8. The scheme of interaction of the components in the ordered composite
system.

In case (a) ( F i g . 21.2.8), the interlayer (a quasihomogeneous fuel with a


very fine oxidizer) burns at a higher rate than the particle ( A P , nitroamine).
In this case, interaction between these components is purely thermal. Then,
the combustion rate w i l l be determined by the rate o f propagation o f the
combustion front through a continuous matrix having regard to the heat
exchange, which depends on the characteristic size and concentration o f
particles. The surface profile represents a system of alternating
protuberances (component 1) and cavities (component 2) o f a symmetric
shape with the leading point at the center o f the interlayer.
338 Combustion of Condensed Composite Systems

In case (b), the combustion rate o f a particle (for instance, a quick-


burning explosive) is higher than the burning rate o f the matrix (a
quasihomogeneous propellant or fuel) and the leading point is located at the
center o f the particle, where a well is formed and a protuberance occurs in
the matrix. The mean combustion rate can be determined approximately by
the following formula [21.11]:

L + L
} 2

r
"(I /r ) + (L /r )'
1 1 2 2

where r x and r 2 are the rates o f the front propagation through both
components, w h i c h depend on the parameters o f heat exchange. It is also
necessary to take into account the non-stationary phenomena appearing at
transition o f the front from a quick-burning component to a slow-burning
one, and the ignition lag connected with this.
Examples (c)-(e) are the cases o f a chemical interaction o f the
components, at which the fuel in the matrix is oxidized by the products o f
decomposition o f an oxidizer. The combustion front propagates along the
contact boundary ((c)-(d)). The cavity has an asymmetric configuration:
both components are protruded, and the wall is steeper and higher for the
particle (r < r ) or the matrix (r > r ). In case (e), the combustion mode is
x 2 x 2

realized when the leading point is located at a certain distance from the
boundary. The burning surface o f the fuel at r < r has a complex form x 2

with a " h i l l " at the center, two pits, and two leading points.
Such a structure o f the combustion surface is in qualitative agreement
with the observations o f the surface o f the extinguished samples o f A P -
based propellant, known in experimental practice. In some cases, at l o w
pressures, particles o f the oxidizer are projected over the other components
and, at high pressures, cavities are formed on the surface. They are also in
agreement with the results o f studies o f the combustion surface structure o f
"sandwich"-type systems consisting o f A P , H M X , and polymers [21.16,
21.17].
Thus, the main conclusion, which follows from the study o f multiphase
layered systems, w h i c h to a certain extent can be considered as the model o f
composite propellants, can be formulated in the following way: the
combination o f chemical (reaction kinetics) and physical (heat and mass
transfer) interactions o f the components in the condensed and gas zones is
the cause o f the existence o f various combustion modes, such as the kinetic,
diffusion-kinetic, heat exchange, and autonomous ones, which, first o f a l l ,
depend on the reactivity o f the components and characteristic sizes.
Finally, the attainment o f this or that combustion mode determines the
level o f the combustion rate, its dependence on the pressure and the initial
temperature, and the sensitivity to catalysts, to high-energetic and
technological additives.
Thermal Decomposition and Combustion 339

21.3 Composite Propellants


Up-to-date knowledge o f the combustion mechanism o f composite solid
propellants, which follows from theoretical and experimental studies (see,
for instance, reviews [21.18, 21.19]), allows one to represent sufficiently
correctly a qualitative picture o f the development o f the combustion process
through the whole front. In the heating zone, the phase transitions o f
separate components proceed; also, thermal decomposition o f the most
reactive compounds is observed, which makes a certain positive or negative
contribution to the initial temperature. Then, the component—the leader
manifests itself (as a rule, this is an oxidizer which is capable of
deflagration), which determines the combustion process. Simultaneously,
the other components are transformed (the fuel binder, the energy-
containing, technological, and catalytic additives) and their interaction with
the main component or the products o f its decomposition occurs. A t the
beginning o f the reaction zone, the process occurs in the kinetic mode
o w i n g to a good contact between the components and comparatively low
temperatures. A s the reaction layer is burnt out, during its loosening and the
formation o f pores and bubbles, the mass and heat exchange factors come
into force. A s a result, the diffusion limitations o f chemical reactions and
thermal inhomogeneity o f the condensed medium appear. Finally, the
combustion surface is formed, i.e., the concept, which is widely used in
various theories o f combustion o f the condensed systems, but which, in
many cases, is not physically clearly justified. In a number o f cases,
researchers are restricted by the assignment o f a certain temperature value
to it, based on the thermal effects o f chemical reactions, frequently
determined arbitrarily or chosen for their correspondence with the
experimental data.
The only factors which are able to give a physical meaning and
quantitative validation o f the combustion surface are the processes of
evaporation (sublimation) and dispersion. If on the basis o f the first factor,
the balance relationships o f the mass and heat on the surface can be
obtained, which unambiguously determine the temperature T , then, in the s

case o f dispersion, the theoretical analysis is very complicated and hardly


can give rise to the quantitative relationships, because it is necessary to have
information about not only the kinetics and thermodynamics o f the process
in the condensed zone, but also the gas dynamics, the strength
characteristics o f the medium, which is destroyed by the gas flow.
In addition, the combustion surface characteristics are influenced by the
gas-phase flame (heat input, the mass flow o f active agents, and erosion). In
the gas phase, the combustion process is also complicated. For the
composite system, the following main flame types are considered in the
literature: the kinetic flame o f an oxidizer, the kinetic-diffusion flame o f
interaction between a gasified fuel and oxidizer, mainly in the vicinity o f
340 Combustion of Condensed Composite Systems

the boundary o f the components (in a so-called "tip o f a flame"), and,


finally, the distant diffusion flame with a finite maximal combustion
temperature, w h i c h only weakly influence the combustion rate. The degree
o f temperature and concentration inhomogeneity o f the gas medium is
determined, mainly, by the characteristic size o f the particles o f the parent
system, but can also be connected with the inhomogeneity and non-one-
dimensionality o f the combustion surface. A s a result, flow turbulence is
developed, which influences the mass exchange along with the diffusion.
At present, computer modeling is widely used for mathematical
solution o f the combustion problems o f complicated composite systems.
Nevertheless, this does not exclude the use o f approximate analytical
methods, especially in those cases when it is necessary to obtain a picture o f
the process development within a wide range o f the values o f the input
parameters and to determine the possible characteristic combustion modes.
In this chapter, we are not going to give a detailed and complete review
o f the experimental and theoretical studies on combustion o f the solid
propellants used in practice. Let us present only some arguments, which can
expand the data available about the combustion mechanism.
The model binary composite system based on A P and H M X studied by
us can serve as an illustration o f interactions o f various types: the chemical
and physical interactions between components, when each o f them is
capable o f deflagration. The dependence o f the combustion rate o f pressed
samples on the concentration ratio o f these components is presented in F i g .
21.3.1. In the case o f coarse particles o f both substances (160-330 p m for
A P particles and 1000-1200 p m for H M X particles), the combustion rate
does not depend on the ratio o f components within the range o f H M X
concentrations o f 4 0 - 1 0 0 % , i.e., under these conditions, particles o f H M X
are able to form a continuous matrix, through which the combustion front
can propagate in spite o f the presence o f the other component, as in the case
o f a pure substance (autonomous combustion). W h e n the content o f H M X is
less than 4 0 % , a decrease o f the combustion rate o f a binary system is
observed, i.e., in this case, the other component ( A P ) exerts an inhibiting
influence on the combustion process due to the thermal (cooling) action and
discontinuity o f the matrix. The picture does not change i f rather fine A P
particles (7-10 urn) are used in the system and i f the size o f H M X particles
remains the- same. Note that similar regularities are observed at other
pressures (within the range o f 2 0 - 1 0 0 atm). Only i f the fine fractions o f
both components (7-10 p m for A P particles and < 140 pm for H M X .
particles) are used is the dependence drastically changed: within the
concentration range o f H M X 20-80%), a considerable increase o f the
combustion rate occurs in comparison with the coarse components and the
appearance o f a m a x i m u m is observed. This is indicative o f a chemical
interaction between substances, when H M X and the products o f its
decomposition play, mainly, the role o f a fuel with respect to more reactive
products o f decomposition o f A P (chlorine oxides).
Thermal Decomposition and Combustion 341

Fig. 21.3.1. Dependence of the combustion rate of A P and H M X mixtures on their


ratio: (1) the mixture of the coarse and (2) fine particles of the components.

The fact that an essential contribution in this chemical interaction can


introduce the processes in the condensed zone is confirmed by the
experiments on heat release during decomposition o f binary systems,
studied with the help o f a differential automatic calorimeter: the maximal
rate o f heat release in the case o f coarse particles o f the components
(approximately 100 cal/g-h) occupies the average position in comparison
with pure substances, although it is shifted in time. In the case o f fine
particles, the maximal rate increases up to approximately 250 cal/g-h.
These peculiarities o f combustion also manifest themselves for ternary
composite systems containing a fuel. The maximal combustion rate is
observed for an H M X concentration o f approximately 2 5 % . Under these
conditions, this agent reacts actively within the leading combustion zone,
and, when its content is 5 0 - 6 0 % , the combustion rate becomes smaller than
for the parent composition. This is indicative o f a predominant cooling
action o f the component introduced.
The chemical and heat exchange interaction o f the components also
reflects on the r{P) dependence o f the composite systems, which can be
represented by the following scheme ( F i g . 21.3.2). Curve (1) refers to the
base mixture o f an oxidizer and a fuel. When the active component is
introduced, the rate increases and it is described by curve (2). In certain
cases, a chemical reaction o f the additive at the leading zone o f combustion
can be stopped by a pressure increase. This can proceed, for instance, due to
the variation o f the ratio between the characteristic particle size and the
width o f the reaction zone. Consequently, the combustion rate value can
come back to the value for the base composition. In this case, the transition
region (curve 2 ) w i l l be distinguished by a reduced pressure coefficient (a
"plateau"). O n the contrary, when an inert component-is introduced (curve
3), the transition region, in which a cooling influence o f the additive
342 Combustion of Condensed Composite Systems

gradually stops (curve 3'), w i l l have an increased value o f n. Introduction o f


a catalyst in the double-base propellant [21.20] and a comparison o f
combustion o f a pure H M X and its mixture with polybutadiene [21.21] can
serve as an illustration o f the r(P) dependence for the first and the second
cases, correspondingly.

Fig. 21.3.2. The scheme of possible dependences of r(P)\ (1) the base mixture of the
oxidizer and fuel, (2) the mixture with an active additive, and (3) the mixture with
an inert additive.

Useful information about the combustion mechanism o f composite


propellants can be obtained from detailed analysis o f the pressure
dependence o f the combustion rate with the help o f the smoothing spline
method and processing o f numerous experimental results, collected by us in
a database [21.22,21.23,21.24].
The essence o f this method is the description o f the r(P) dependence
within a finite interval by a cubic spline, w h i c h obeys certain conditions and
provides a prescribed closeness o f an approximate reconstructed function to
the experimentally measured data, as well as the absence o f sharp variations
o f this function between measurements.
H a v i n g a reconstructed function r(P), one can easily calculate the value
o f the pressure coefficient n within any interval, which can be as narrow as
necessary:

; ; =
l n
(^2^l)
ln(P //>,)'
2

and construct the dependence n(P).


Thermal Decomposition and Combustion 343.

* 0.6

P, atm
P

Fig. 21.3.3 Fig. 21.3.4

Fig. 21.3.3. The pressure dependence of the coefficient n for the composite
propellants according to [21.10]: (1) A P + P M M A , (2) A P + polystyrene, and (3)
A P + bitumen.

Fig. 21.3.4. The scheme of transformation of the dependence n(P).

In the majority o f cases, for composite AP-based systems, the dependence


n(P) is observed, w h i c h is presented in F i g . 21.3.3 for the composition o f
A P with polystyrene. A t a pressure o f approximately 20 atm, a maximum is
observed, and at 60 atm, a m i n i m u m appears. A s shown by analysis o f the
experimental data, under the influence o f such factors as the size o f the
particles o f the oxidizer, the ratio between the fuel and the oxidizer, the
nature o f the fuel, the additives o f catalysts, and others, the transformation
o f curves can proceed, which is schematically shown in F i g . 21.3.4. It
manifests itself in the following form: a shift o f the curve along the x-axis.
Consequently, one o f the extrema, which is observed within the pressure
range studied, disappears, and the resulting curve has only one m a x i m u m or
m i n i m u m , or one can observe a gradual smoothing o f the "pit" (the
minimum) and its transformation into the plateau with an increase o f n.
One can consider a logical assumption that the maximum on the curves
o f the dependence n{P) is caused by the mechanism o f combustion o f A P ,
by its complex chemical transformation in the condensed and gas phases,
which, under these conditions, controls the combustion process. Apparently,
the nature o f the m i n i m u m with a following growth o f n (let us call it a
"tail") and its transformation are connected with the interaction processes
between the oxidizer and the fuel in a so-called "hydrocarbon flame" o f a
diffusion-kinetic nature.
The following is indicative o f this fact:
1. The appearance o f the " t a i l " and its intensification are caused by factors
which intensify the interaction o f the oxidizer and the fuel: a high
dispersion o f the oxidizer, a relative thermal stability and a low volatility o f
344 Combustion of Condensed Composite Systems

the fuel, w h i c h facilitate a partial but very efficient (for the increase o f the
combustion rate) oxidation at earlier stages o f combustion, for instance,
bitumen ( F i g . 21.3.3).
2. The weakening o f the " t a i l " and its disappearance, i.e., the transition to a
perchlorate dependence, are equivalent to a reduction o f a chemical
interaction o f the components and to an increase o f a relative role o f the
oxidizer. This is a typical situation for systems with a coarse A P , with an
easily volatile fuel, for instance, P M M A ( F i g . 21.3.3, curve 1), as w e l l as
with catalysts promoting decomposition of AP (copper chromate,
organometal catalysts based on metals o f variable valence).
Consequently, one can conclude that the degree with w h i c h the " t a i l "
manifests itself correlates with the influence o f the chemical interaction o f
the fuel and oxidizer, w h i c h depends, first o f all, on the dispersion and
reactivity o f the components.
The following conclusions can be formulated based on the combustion
models considered and their comparison with experiments [21.25, 21.26]:
1. The combustion mechanism o f composite solid propellants cannot be
satisfactorily formulated and understood without a preliminary explanation
o f the mechanism o f combustion o f the oxidizers, w h i c h represent the most
essential fraction o f propellants, and, in the majority o f cases, play an
important and leading role in the combustion process.
The value o f the combustion rate o f inorganic oxidizers (onium salts o f
different acids) is determined, mainly, by the rate constant o f thermal
decomposition in the condensed phase at the temperature o f the combustion
surface. The latter depends on sublimation (evaporation) o f the substance,
which exerts a control and • stabilizing influence on the combustion
characteristics. The gas phase reactions influence the combustion process in
a m i x e d mode, introducing an additional contribution to the heat balance on
the combustion surface.
W i t h i n the framework o f the mechanism considered, one can explain
complex and specific regularities o f combustion: the pressure dependence
o f the combustion rate (the maximum o f the n(P) curve), the dependence o f
the combustion rate on the initial temperature (the m i n i m u m on the o(P)
curve), the existence o f the region o f unstable combustion or the occurrence
o f several limits o f stable combustion.
2. For composite propellants, the kinetic factor plays an important role in
the mechanism o f combustion: the ratio o f the rates o f decomposition o f the
oxidizer and fuel oxidation at earlier stages o f combustion. In the case o f a
rather thermally stable oxidizer (for instance, A P ) , a low volatile fuel has
time for a partial oxidation in the condensed zone and can introduce a
corresponding contribution to the heat balance, providing stability o f
combustion and an increase o f the combustion rate in comparison with a
pure oxidizer. In the case o f an easily decomposing oxidizer ( A D N ) and a
rather thermally stable fuel, the ratio o f the rates o f transformation o f both
components can be such that oxidation o f the fuel in the condensed zone
Thermal Decomposition and Combustion 345

w i l l not proceed, and due to the heat consumption for its heating and
gasification and insufficient heat flux from the gas-phase zone, the
combustion rate o f the composite mixture w i l l decrease in comparison with
a pure oxidizer. The same circumstance is the reason for an "anomalous"
character o f the dependence o f the combustion rate on the particle size o f
such an oxidizer: the increase o f r with d going up. F r o m this, one can see
an efficient way to increase the combustion rate: to use rather thermally
stable fuels and other components, containing active functional groups
rapidly oxidized with the high thermal effect.
3. A t all stages o f combustion o f the composite propellant, the heat and
mass exchange o f the components influences both the level o f the
combustion rate and the character o f its dependence on the external
conditions o f combustion and internal parameters o f the fuel (the size o f
particles o f the components, the reactivity). A s a result o f a joint action o f
these phenomena and the reaction kinetics, combustion can proceed in the
kinetic, diffusion-kinetic, heat exchange, and autonomous modes. These
modes can be influenced not only by varying the component dispersion (for
instance, a super fine ammonium perchlorate, a super fine aluminum), but
by increasing the degree o f homogenization o f the fuel by combining in one
component or one phase the properties o f an oxidizer and a fuel (for
instance, methylammonium perchlorate, the solution o f lithium perchlorate
in acrylamide), or an oxidizer, a fuel and a catalyst (perchlorate o f
ferrocinium derivatives), as well as using film coatings o f particles o f the
oxidizer, which possess catalytic, inhibiting, or high energy content
properties.

REFERENCES

21.1. V . A . Strunin and G . B . Manelis, F i z . Goreniya V z r y v a , 15(5): 2 4 -


33 (1979).
21.2. G . B . Manelis, V . A . Strunin, and A . P. D ' y a k o v , Joint Meeting of
the Soviet and Italian Sections of the Combustion Institute, Pisa:
Tacchi Editore, 7.4(1990).
21.3. W . H . Andersen, K . W . B i l l s , E . Mishuck, G . M o e , and R. D .
Schultz, Combustion and Flame, 3(3): 301-318 (1959).
21.4. A . S. Shteinberg and S. S. Rybanin, Heat and Mass Exchange,
M M F , M i n s k : 3: 132-134 (1988).
21.5. V . A . Strunin and G . B . Manelis, Proceedings of the 11th All-
Union Conf on Evaporation, Combustion, and Gas Dynamics of
Dispersed Systems, Odessa: 39 (1972).
21.6. S. S. Rybanin and V . A . Strunin, Proceedings of the V-th All-
Union Conf on Heat and Mass Exchange, M i n s k : 2: 3 0 - 3 9
346 Combustion of Condensed Composite Systems

21.7. A . P . D ' y a k o v , V . A . Strunin, and G . B . Manelis, Proceedings of


the IV-th Ail-Union Symposium on Combustion and Explosion, M . :
244-248(1977).
21.8. V . A . Strunin, A . P. D ' y a k o v , L . I. N i k o l a y e v a , and G . B . M a n e l i s ,
Fiz. Goreniya Vzryva, 35(3): 57-62 (1999).
21.9. V . A . Strunin, A . N . Firsov, K . G . Shkadinskii, and G . B . Manelis,
Fiz. Goreniya Vzryva, 25(5): 25-32 (1989).
21.10. V . A . Strunin, A . N . Firsov, K . G . Shkadinskii, and G . B . M a n e l i s ,
Flame Structure, Novosibirsk: Nauka, 1: 258-261 (1991).
21.11. N . N . Bakhman and A . F . Belyaev, Combustion of Heterogeneous
Condensed Systems, M . : Nauka, 226 (1967).
21.12. B . S. Ermolaev, A . I. Korotkov, and Y u . I. Fr.olov, Fiz. Goreniya
Vzryva, 6(3): 277-285 (1970).
21.13. V . A . Strunin, A . P. D ' y a k o v , and G . B . Manelis, Fiz. Goreniya
Vzryva, 17(1): 19-23 (1981).
21.14. S. S. N o v i k o v , V . Y u . Potulov, and S. V . C h u i k o , Combustion of
the Condensed Systems, Chernogolovka: 56-58 (1977).
21.15. V . A . Strunin, A . N . Firsov, K . G . Shkadinskii, and G . B . Manelis,
Fiz. Goreniya Vzryva, 26(5): 36-42 (1990).
21.16. E . W . Price, J. K . Sambamurthi, R . K . Sigman, and R . R . Panyam,
Combustion and Flame, 63(3): 381—413 (1986).
21.17. O . P. Korobeinichev, A . G . Tereschenko, V . H . Shvartsberg, et al,
Flame Structure, Novosibirsk: Nauka, 1: 262-267 (1991).
21.18. N . S. Cohen, C . F. Price, and L . D . Strand, AlAA Paper, 921 \ 1-11
(1977).
21.19. K . N . R . Ramohalli, Fundamentals of Solid-Prop ell ant Combus­
tion, ed. K . K u o and M . Summerfield, 90(8): 409-477 (1984).
21.20. N . Kubota, T . J. Ohlemiller, L . H . Caveny, and M . Summerfield,
15th Symp. (Int.) on Combustion, The Combustion Institute,
Pittsburgh, P A , 529-537 (1974).
21.21. N . S. Cohen and C . F . Price, Journal of Spacecraft, 12(10): 6 0 8 -
612 (1975).
21.22. L . B . Petukhova, T . V . Peregudova, V . A . Strunin, and G . B .
Manelis, Fiz. Goreniya Vzryva, 25(3): 36-39 (1989).
21.23. V . A . Strunin, L . B . Petukhova, and G . B . Manelis, Fiz. Goreniya
Vzryva, 29(2): 68-72 (1993).
21.24. V . A . Strunin, L . B . Petukhova, and G . B . Manelis, 25th Symp.
(Int.) on Combustion, Irvine: U n i v . C a l i f , Poster Session, 5.04.428
(1994).
21.25. G . B . Manelis, G . M . N a z i n , Y u . I. Rubtsov, and V . A . Strunin,
Thermal Decomposition and Combustion of Explosives and
Powders, M . : Nauka, 223 (1996).
21.26. V . A . Strunin and G . B . Manelis, Journal of Propulsion and
Power, 11(4): 666-676 (1995).
Author Index

Andersen, W . H . , 321 Robertson, R., 1


Apin, A . Ya., ix Roginskii, S. Z . , 1
Rybanin, S. S., 322
Belyaev, A . F., ix, 271,297
Benreuven, M , 313 Sapozhnikov, A . V . , 1
Boggs, T. L . 280 Semenov, N . N . , ix, 1
Schroeder, M . 77
Curtius, T. 95 Shaw, R. 74
Shkadinskii, K . G . , 327
Dickson, N . J . 99 Shteinberg, A . S., 321
Dubovitskii, F. I., 1 Siegmund, R. F. 189
D ' y a k o v , A . P., 318 Solymosi, F. 189

Farmer, R. C , 1 Ubbelohde, A . R. 21
Firsov, A . N . , 327
Fogelzang, A . E . , 293 Walker, F. 74
Frank-Kamenetskii, A . D . , 15,
271,303 Zel'dovich, Y a . B . , ix, 271,
275,297,303,316
Garner, W . E . , 1
G l a z k o v a , A . P., 285

H a l l , A . R. 189
Hinshelwood, C . N . , 1
Huisgen, R. 121

Keenan, A . G . 189
Khariton, Y u . B . , ix
Korobeinichev, 0 . P., 312

L'Abbe, G. 95
Leermakers, J. A . 96

MacCall, A . 38
M a k s i m o v , E . I., 274
Merzhanov, A . G . , 274

Novozhilov, B . V . , 278

Pearson, G . S. 189
P o k h i l , P. F., ix, 271
Raevskii, A . V . , 30, 199

347
Subject Index

Acid-base catalysis, 79 combustion, 221


Acidation, 53 decomposition in the liquid
A c r y l a m i d e , 345 phase, 221, 227
A l c o h o l nitrates, 212 decomposition in the melt,
Aliphatic azides, 96, 97, 98, 99, 2 2 1 , 2 2 2 , 223
101 decomposition in the solid
Aliphatic nitrocompounds, 13, phase, 227
35,36, 43,53 A m m o n i u m nitrate, 145, 152,
A l l y l anion, 95 175, 189, 2 2 2 , 2 2 3 , 2 2 7 , 2 2 8 ,
A l k y l group, 39, 48, 88, 96, 106 243,285,291,292
A l k y l substitutes, 40, 51, 79, 106 A m m o n i u m perchlorate, 26, 30,
A l k y l a t i o n , 40, 41 31, 1 8 9 , 2 1 8 , 2 1 9 , 237, 2 4 1 , 2 4 3 ,
A l u m i n u m , 99, 2 9 1 , 3 3 2 , 3 3 3 , 254,280,282,283,284,307,
345 3 1 5 , 3 1 8 , 3 1 9 , 345
A m m o n i u m bichromate, 26 Anharmonicity coefficient, 56
A m m o n i u m dinitramide A n i l i n e , 13
ammonium salt o f dinitramide Anthracene, 13, 109
( A D N A ) N H N ( N 0 ) - , 159,
4
+
2 2
Aromatic azides, 99, 101, 138
166, 168,221 Aromatic nitrocompounds, 9, 11,
anomalous decomposition,
2 2 1 , 2 2 8 , 2 2 9 , 230
1 3 , 6 1 , 6 8 , 69, 70, 185 213, 214, 216, 222, 224, 232,
Arrhenius dependence, 21, 69, 235,243,246
226,229 Autoprotolysis, 53, 54
Arrhenius equation, 2, 21, 99 Azides, 9, 10, 11,95-102, 120,
Arrhenius equation parameters, 121, 138
21 aliphatic azides, 96-99, 101
Arrhenius law, 2 1 , 3 1 aromatic azides, 99, 101, 138
Arylazides, 100, 101 heterocyclic azides, 99, 101
Autocatalysis, 13, 14, 53, 68, 70, Azidoazomethanes, 121
79, 80, 154, 172, 177, 189, 190, Azidobiphenyls, 97
Benzylazide P h C H N , 982 3

l-(azidomethyl)-3,5,7-trinitro- Benzyl-benzylidenamine ( B B A ) ,
1,3,5,7-tetraazacyclooctane, 138 98
Azomethineazides, 121 Benzyldifluoroamine, 110
Binary composite system, 340
Ballistic properties, 3 Binder, 3 , 9 9 , 210, 217, 2 6 7 , 3 1 5 ,
Benzimidazoles, 119 339
Benzofuroxan, 65 Biphenyl, 63
Benzonitrile, 63, 114 6/.s(difluoroarnino)-9, 10-
Benzotrifuroxane ( B T F ) , 113, dihydroanthracene, 109
118, 119

349
350 Subject Index

Bond breaking, 9, 11, 12, 23, 26, CH CH(NF )CH=CH-CH NF ,


3 2 2 2

35, 48-53, 55, 58, 61, 62, 73, 75, 108


80, 95, 96, 105, 107, 108, 117, CH CN,41,43, 109, 111, 114,
3

127, 130, 138, 149, 170,259 115


Boranehydrazine, 274 CH CNF (N0 ) ,136
3 2 2 2

Boron nitride, 274 CH C(N0 ) ,47, 48


3 2 2

BrC(N0 ) , 48 2 2
CH C(N0 ) CH , 44
3 2 2 3

BrC(N0 ) , 44, 56 2 3
CH -C(N0 ) CH , 44, 48
3 2 2 3

Br-C(N0 ) , 44, 49, 56 2 3


CH -C(N0 ) , 44, 48, 55
3 2 3

Bromonitrobenzene, CH C(N0 ) , 44, 55


3 2 3

m-bromonitrobenzene, 62 CH COCH N , 97
3 2 3

/7-bromonitrobenzene, 62 CH COOCH , 43
3 3

1,4-butyleneglycol-dinitrate, 128 CH COOH, 43, 169


3

2,3-butyleneglycol-dinitrate, 128 CH COOC(CH )(NF )CH C-


3 3 2 2

Butylnitrate, (CH )(COOCH )-CH NF ,109


3 3 2 2

tert-butylnitrate, 128 CH N(N0 )CH C(N0 ) ,137


3 2 2 2 3

Butyltrinitrotoluene, CH NO, 43, 54, 76


3

o-tetra-butyltrinitrotoluene, CH N0 , 11, 43, 47, 50, 51, 53,


3 2

66 54, 108, 111, 114, 115


(CH )(N0 ) C-C(N0 ) (CH3),
3 2 2 2 2

Calculations, 49
quantum-chemical (CH )(N0 ) C-C(N0 ) ,49
3 2 2 2 3

calculations 54, 67, 73 CH N ,973 3

CC1 , 56, 57, 111, 151


4
CH OH, 39, 43, 53
3

C(CH ON0 ) (PETN), 4, 22,


2 2 4
CH ONO, 43, 78
3

24, 128, 129, 185 CH ON0 ,43,78, 115


3 2

CH(N0 ) ,48, 52 2 2
CH SC(N0 ) CH ,45
3 2 2 3

CH =CHN0 ,42, 52
2 2
CH SC(N0 ) ,45
3 2 3

CH N0 ,48
2 2 CH S0 C(N0 ) CH ,45
3 2 2 2 3

CH CBr(N0 ) ,44
3 2 2 (CH ) CBr(N0 ), 40
3 2 2

(CH )(C H )CC1(N0 ), 40


3 2 5 2 (CH ) CC1(N0 ), 40
3 2 2

CH CC1 N0 ,44
3 2 2 (CH ) CHCBr(N0 )CH ,40
3 2 2 3

CH CF(N0 ) ,44,51,52
3 2 2 (CH ) CHCC1(N0 )CH ,40
3 2 2 3

CH CHFN0 ,52
3 2 (CH ) C(NF ) ,106, 107
3 2 2 2

(CH )CHF(N0 ), 40
3 2 (CH ) C(NF )-C(NF )(CH ), 108
3 2 2 2 3

CH CH=CHN0 ,42
3 2 (CH ) CN0 ,48
3 2 2

(CH )CHC(CH )(NF )CH NF ,


3 3 2 2 2 (CH ) N radical, 76
3 2

108 (CH ) (N0 )C-C(N0 )(CH ) ,49


3 2 2 2 3 2

CH CHN0 ,48
3 2 [CH C(N0 ) CH ] NN0 ,24
3 2 2 2 2 2

CH -CH(N0 ) ,48
3 2 2 (CH ) CN0 523 3 2>

CH CH(N0 ) ,44
3 2 2 [C(N0 ) CH ] NN0 ,24 2 3 2 2 2

CH CH(N0 ) ,44
3 2 2 C(NF ) ,106, 107 2 4

CH -CH(N0 )CH ,48


3 2 3 (CN)(N0 ) CCH CH(OAc) ,55 2 2 2 2

CH -CH(N0 )(CH ) ,48


3 2 3 2 (CN)(N0 ) CCH OCH ,55 2 2 2 3

CH -CH(N0 )C H , 48
3 2 2 5 (CN)(N0 ) CCH CH 0ac, 55 2 2 2 2

CH -CH (N0 ), 50
3 2 2 C(N0 ) ,47, 48, 55 2 3

C(N0 ) C(N0 ) ,24, 44, 50, 56


2 3 2 3
Thermal Decomposition and Combustion 351

C ( N 0 ) , 4 4 , 4 5 , 5 5 , 5 6 , 106
2 4 Cations,
C H CC1 N0 ,44
2 5 2 2 Ag ,238,263
+

( C H ) C H B r ( N 0 ) , 40
2 5 2 B a , 263,265
2 +

C H C H C 1 ( N 0 ) , 40
2 5 2 Cd ,2632 +

C H CHN0 ,48
2 5 2 Cs ,263,265
+

C H CH=CHN0 ,42
2 5 2 Li ,263,265
+

C H C H ( N 0 ) , 4 4 , 52
2 5 2 2 N a , 263-265
+

C H - C H ( N 0 ) , 49
2 5 2 2 Rb ,263,265
+

C H C(N0 )=CH ,42


2 5 2 2 Tl ,263
+

C H C(N0 ) ,44,48
2 5 2 2 Cation rearrangement, 108
C H C(N0 ) ,52
2 5 2 3 Cation vacancies, 205
C H - C ( N 0 ) , 49
2 5 2 3 Cellulose nitrate, 127, 130-132,
C H COOCH N ,97
2 5 2 3 185
C H N ,97
2 5 3 Chain reaction, 4, 11,30, 70, 77,
C H N0 ,43
2 5 2 157, 159, 2 0 9 , 2 2 7
C H C(N0 ) ,44
3 7 2 3 Chemical arc, 321,324-326
C H CH(N0 ) ,44
3 7 2 2 Chemical kinetics, 1, 5, 7, 9, 220
C H -CH(N0 ) ,49
3 7 2 2 Chemical physics, 1, 32, 72, 91,
C H -CH N0 ,49
3 7 2 2 92, 123, 173, 187, 219, 220, 248,
C H N0 ,
3 7 2 249
/7-C H N0 ,43 3 7 2 Chlorine oxides, 154, 158,205,
W0-C3H7NO2, 44 245,341
C H -C(N0 ) ,49
4 9 2 3
Chloroform, 158
C H -CH(N0 ) ,49
4 9 2 2
Chloronitrobenzene,
C H CH=CHN0 ,42
6 5 2 w-chloronitrobenzene, 62
C H CH NF ,108
6 5 2 2 o-chloronitrobenzene, 62
C H CH(NF ) ,108
6 5 2 2 /^-chloronitrobenzene, 62
C H C H ( N F ) C N , 108
6 5 2 Chloropicrin, 42
C H C H ( N F ) C H N F , 108
6 5 2 2 2 C h l o r y l fluoride C 1 0 F , 208
2

( C H ) C H N F , 108
6 5 2 2 Coefficient,
C1C(N0 ) ,48 2 2 coefficient o f thermal
C1C(N0 ) ,44, 45,56 2 3 conductivity, 272
C 1 C ( N 0 ) C ( N 0 ) C 1 , 56 2 2 2 2 pressure coefficient, 275, 277,
C 1 0 , 154, 158, 159, 198,208,
2 282, 284, 285, 289, 291, 299,
312 301,310,313,319, 341,342
CIO3,154, 157, 158, 170, 198, temperature coefficient, 276-
205, 208, 246, 247 278,282,283,297, 298,300,
C1 C(N0 ) ,41
2 2 2 303,318, 321,332,333
C 1 0 , 1 5 7 , 170, 198, 208, 246,
2 7 Combustion, 3, 5, 6, 7, 10, 33,
247,254 72, 90-921, 132, 133, 166, 173,
C1 CN0 ,443 2 187, 189, 195, 196, 203, 204,
CuCrO ,210 4 208, 2 0 9 - 2 1 1 , 2 1 9 - 2 2 1 , 2 3 1 , 2 4 8 ,
Calorimetric method, 10, 227 249, 251, 267-322, 324-341, 343-
Carbazole, 99 346
Carbodiimides A r N = C = N a r \ combustion instability, 283,
119, 122 284
Carbonylazides, 121
352 Subject Index

combustion o f explosives, 5, Degree o f transformation, 4, 13-


6, 92, 294, 295, 346 16, 18-20, 2 7 , 3 1 , 4 1 , 8 9 , 99,
combustion mechanism, 6, 129,213,261,316,319
267, 268, 279, 307, 311, 313, Detonation, 3, 53, 79, 166
314, 321, 336, 338, 339, 341, Detonation ability, 53
343 l,5-diacetyl-3,7-dinitro-l,3,5,7-
combustion rate, 6, 204, 268, tetraaza-cyclo-octane, 85
271, 272, 274-280, 282-285, Diaminohexanitrodiphenyl
287, 290-294, 297-301, 303, ( D A H N D ) , 66
304, 307-314, 316-322, 324- Diaminotrinitrobenzene
327, 330-333, 336-341, 343, ( D A T N B ) , 66
345 Diammonium phosphate ( D A P h ) ,
combustion temperature, 6, 184
221,274, 275,298-300,302, l,7-diazido-2,4,6-trinitro-2,4,6-
339 triazaheptane, 139
stationary combustion, 269, Dibenzofuran, 64
283 Dibenzylbenzamidine ( D B A ) , 98
Complete dissolution, 17-21, Diethyleneglycol-dinitrate, 128
194,239 Diethylferrocene (DEF), 217,
Compounds, 2 1 8 , 3 2 0 , 326
compounds with mixed Diethylnitramine, 79
functions, 135 Diffusion, 5, 9, 25, 28, 30, 197,
heterocyclic compounds, 113, 203, 211, 213-216, 271, 272,
124 316, 317, 323, 324, 327, 331,
polyvinylmethyltetrazole type 332, 336, 338-340, 343, 345
compounds, 113 Difluoroamines, 9, 10-12, 102-
Concerted decomposition, 117 105, 107, 109, 136
Concerted reactions, 12 Difluoroamino-compounds, 12,
Conductivity, 105, 106
electron conductivity, 142 1,2-difluorotetranitroethane, 46
proton conductivity, 189 1.2- dihydro-syw-triazine, 79
Consecutive reactions, 267, 300, Diisopropylferrocene ( D P F ) ,
302 2 1 7 , 2 1 8 , 320
Critical diameter o f detonation, Dimerization, 130
53 Dimethylenetriamine, 53
Crystal lattice, 1, 2, 16, 22-25, Dimethylferrocene, 217, 218,
69, 70, 80, 145, 159, 181, 183, 319, 320
189, 192, 204-206, 229 Dimethylfurazane ( D M F ) , 114,
C y c l e opening, 12, 64 115
C y c l i c nitramines, 73, 76, 77, Dimethylnitramine ( D M N A ) , 73
314 Dimethylnitrosoamine, 80
Cyclohexane, 56, 106 Dinitramide H N ( N 0 ) , 141,
2 2

cyc/o-C HnN ,97


6 3 146, 159-166, 1 6 8 - 1 7 2 , 2 2 1 , 2 5 7 -
Deflagration, 280, 282, 285, 288, 263,265,271,290
320, 324, 325, 327, 339, 340 1.3- dinitrate-glycerin, 212
Dehydration, 29, 171, 212, 263 Dinitroanthranyl, 68
Dinitrobenzene,
Thermal Decomposition and Combustion 353

w-dinitrobenzene, 62 Enthalpies o f formation, 39, 48,


odinitrobenzene, 65 49, 54, 252
p-dinitrobenzene, 61 Entropies o f activation, 89
l,4-dinitro-l,4- Entropy factor, 128
diazacyclohexane, 24 Epoxyethylferrocene ( E E F ) , 326
Dinitromethylnitramine, 89 Equivalent nitrogroups, 45
1.4- Dinitro-tetrahydro-imidazo- Ethyleneglycoldinitrate, 128
[4,5-d]-imidazole2,3(lH,3H>- Ethylnitrate, 128,269
dionine, 85 Eutectic mixture, 183, 201, 204,
Dinitroxy-ethylnitramine 206, 207, 227-229, 264
( D I N A ) , 128 Eutectic mixture N H C 1 0 - 4 4

Diphenylamine, 13, 130 L i C 1 0 , 201


4

3.5- diphenyl-4-benzyl-l,2,4- Eutectics, 16


tetrazol ( D P h B T ) , 98 Evaporation, 5, 6, 144, 145, 151,
Diphenylfurazane, 114 152, 175, 179-181, 193, 211,
Dipole moment, 38, 39, 229 214, 269, 275-279, 282, 284,
Dipropylnitramine, 81 285, 288, 292-293, 313, 314,
di-isopropylnitramine, 81 3 1 7 , 3 3 9 , 344, 345
Direct resonance coupling, 62 Exothermicity, 79
Dislocations, 2, 25, 26, 30, 31, Explosion, 5, 9, 15, 16, 72, 79,
199, 200-203 88, 91, 92, 133, 153, 154, 159,
Dispersion in combustion, 271 160, 173, 187, 219, 220, 248,
Double bond, 50, 99, 117 249, 295, 345
Explosives, 1-6, 9, 10, 12, 13, 19,
Effect, 57, 70, 72, 73, 90-93, 98, 102,
deuterium isotopic effect, 50, 113, 123, 127, 132, 133, 185,
51,75 219, 2 7 1 , 2 7 5 , 2 9 4 , 2 9 5 , 3 4 6
isotopic effect, 50, 51, 53, Explosive composition, 3
6 5 , 7 5 , 7 8 , 2 2 5 , 2 2 6 , 228
field effect, 48, 138 F B r C ( N 0 ) , 44 2 2

kinetic isotopic effect, 75, 78, F C 1 C ( N 0 ) , 44 2 2

225, 226, 228 FCNF (N0 ) ,136


2 2 2

mesomeric effect, 40, 47 F C ( N F ) , 1 0 6 , 107


2 3

secondary P-deuterium F - C ( N 0 ) F C ( N 0 ) F , 49
2 2 2

isotopic effect, 75 F(CN0 ) ,482 2

steric effect, 47 F-C(N0 ) CH ,492 2 3

Electron impact, 74 F - C ( N 0 ) F , 48
2 2

Electron transfer, 26, 40, 142, F C ( N 0 ) C ( N 0 ) F , 44


2 2 2 2

149,258 F - C ( N 0 ) C ( N 0 ) F , 49
2 2 2

Electronegativity, 38 F - C ( N 0 ) C ( N 0 ) , 44, 49
2 2 2 3

Electrophilic character, 62 F C ( N 0 ) C ( N 0 ) , 44
2 2 2 3

Electropositive substitutes, 128 [FC(N0 )CH ] NN0 ,24


2 2 2 2

Elimination, 2, 3, 10, 11, 12, 23, F C ( N 0 ) , 44, 106


2 3

35, 36, 38-40, 46, 51, 74, 75, 95, F C ( N 0 ) C ( N 0 ) , 48


2 2 2 2

107-109, 130, 1 5 4 , 2 1 3 , 2 1 4 , 2 2 9 F C ( N 0 ) N F , 136


2 2 2

Energetic characteristics, 47, F C ( N 0 ) C H N ( N 0 ) C H , 137


2 2 2 2 3

101, 113 [ F C ( N 0 ) C H ] N N 0 , 137


2 2 2 2 2
354 Subject Index

[FC(N0 ) CH NN0 CH ] ,137 2 2 2 2 2 2 Furoxan, 9, 10, 10


FC(N0 ) (CH ) -C(NF ) CH , 2 2 2 2 2 2 3

137 Glycerin dinitrate, 213


Fe O ,210,218,252
2 3 G lycerintrinitrate, 12 8
F I C ( N 0 ) , 44 2 2 Glycidylazide polymer, 98
[ F ( N 0 ) C C H ] , 44 2 2 2 2 Group,
[ F ( N 0 ) C C H ] C B r N 0 , 44
2 2 2 2 2 azide-group, 101
[ F ( N 0 ) C C H ] C H N 0 , 44
2 2 2 2 2 explosive group, 107, 113
[ F ( N 0 ) C C H ] C ( N 0 ) , 45
2 2 2 2 2 2 free hydroxy 1 groups, 131,
[ F ( N 0 ) C C H ] C F N 0 , 44
2 2 2 2 2 132
[ F ( N 0 ) C C H ] C C 1 N 0 , 44
2 2 2 2 2 heme-group - C ( N F ) - , 107 2 2

F(N0 ) CCH CH COO- 2 2 2 2 nitramine-group, 79, 138, 139


CH C(N0 ) 1372 2 3 /^-nitramine group, 138, 139
F ( N 0 ) C - C ( N 0 ) F , 44, 49
2 2 2 2 nitrogroup, 39, 41, 45, 47-49,
F ( N 0 ) C - C ( N 0 ) , 4 4 , 49
2 2 2 3 74, 80, 1 3 6 , 2 5 7 , 2 5 9 , 263,
[F(N0 ) C—CH 0] CH ,56 2 2 2 2 2 268
[F(N0 ) CCH CH COOCH C(N 2 2 2 2 2 Guanidinium perchlorate
0 ) F , 56
2 2 CN H ClO ,207
3 6 4

F C(NF ) ,106
2 2 2

F C N 0 , 48
2 2 H - C H ( N 0 ) C H , 49 2 3

F C ( N 0 ) , 44, 50
2 2 2 H - C H ( N 0 ) C H , 49 2 2 5

F NCH CH(NF )CH OOCC-


2 2 2 2 H - C H ( N 0 ) C H , 49 2 3 7

(CH )(NF) CH NF ,108


3 2 2 2 H - C H ( N 0 ) , 49 2 2

F NCH (CHNF ) CH ,108


2 2 2 3 3 H - C H N 0 , 49 2 2

F NCH C(CH )(NF )CH(NF )C


2 2 3 2 2 H C 1 0 , 154-159, 189, 196-198,
4

H N F , 108
2 2 200, 205, 208, 211, 236, 237,
[F NCH C(CH )(NF )-
2 2 3 2 242, 283
COOCH -] ,108 2 2 H C N , 4 3 , 7 8 , 98, 109, 114
[F NCH CH(NF )CH ] ,108
2 2 2 2 2 H - C ( N 0 ) ( C H ) , 49 2 3 2

F NC(N0 ) ,136
2 2 3 H - C ( N 0 ) C H , 49 2 2 3

[ F N C ( N 0 ) C H ] N N 0 , 137
2 2 2 2 2 2 H - C ( N 0 ) C H , 49 2 2 2 5

Ferrocene, 2 1 0 , 2 1 7 , 218, 318- H - C ( N 0 ) C H , 49 2 2 3 7

321 H - C ( N 0 ) , 44, 49, 50 2 3

Final products, 3, 6, 43, 114, 127, H C O O H , 78


153, 157, 171, 236, 239, 242, HCOOCH ,78 3

252,302,312,313 H C O O N H C H O , 79
First order reaction, 14, 17, 19, H F C ( N 0 ) , 44, 50 2 2

20, 54, 160, 161 H O C H N H C O H , 78 2

Foaming, 269, 273, 274, 294 H O O C C H C O O H , 24 2

Formation o f aci-forms, 53 H N ( N 0 ) , 162-166, 168, 169,


2 2 2
+

Formula o f Frank-Kamenetskii, 171, 172


15 H N 0 \ 162
2 3 4

Free valence, 25, 67, 77, 78, 205 H 0 i o n s , 160, 161


3
+

Freon-114V, 56 Heat transfer, 5, 6, 7, 277, 279


Freon-113, 56 /zeme-dinitroalkanes, 35, 43
Furazan, 9 Heptane, 56
Furazanylazides, 101 Heterocyclic azides, 99, 101
Thermal Decomposition and Combustion 355

Heterogeneous reaction, 11, 36, hydrocarbon chain, 127


80, 1 0 7 , 3 2 2 , 3 3 3 hydrocarbon polymers, 210, 291
Hexachlorbenzene, 67, 68 hydrogen donors, 53
Hexanitrobenzophenone hydrolysis
( H N B P h ) , 66, 69 acid hydrolysis, 129,213
1,1,3,5,5,7-hexanitro-3,7- nitroglycerin hydrolysis, 130
diazacyclooctane, 24 hydroxonium ions, 162
Hexanitrodibenzyl ( H N D B ) , 66 hydroxylammonium,
Hexanitrodiphenyl ( H N D P h ) , 66 hydroxylammonium chloride
Hexanitrodiphenylsulphide N H O c l , 240
4

( H N D P h S u l f i d e ) , 22, 66 hydroxylammonium nitrate


Hexanitroethane, 13, 45, 55 N H O N 0 , 243
4 3

Hexanitromannitol, 128-130 hydroxylammonium


Hexanitrostilbene ( H N S t ) , 66 perchlorate N H O H C 1 0 ,3 4

Hexogen ( R D X ) , 4, 15, 24, 73- 241,243,245,282,283


75, 77-80, 82-84, 88, 127, 313, hydroxylammonium
314 phosphate ( N H 0 ) P 0 , 238
4 3 4

High-energetic compounds, 1-3, hydroxylammonium sulphate


9, 269, 275 ( N H 0 ) S 0 , 238
4 2 4

High-vacuum pyrolysis, 10, 11,


74 I C ( N 0 ) , 44, 55, 56
2 3

H M X , 22, 24, 73-75, 77, 80, 84, Imines, 96


88, 9 9 , 3 1 3 , 3 2 1 , 3 3 8 , 340, 341 Impurities, 16, 1 9 , 2 1 , 4 3 , 131,
/ ? - H M X , 80, 88 205,221,239
H N 0 2 elimination, 12, 35, 38, Induction period, 5, 154, 155,
4 0 , 5 1 , 5 3 , 7 4 , 7 5 , 108 158, 159, 192, 197, 199, 202,
Humidity, 131, 176, 182, 183, 204-208, 212, 215, 216, 224,
262, 263 227,228
Hydrazine, 2 3 1 - 2 4 1 , 2 6 9 , 3 0 3 , Inhibitors, 4, 7, 10, 1 1 , 6 1 , 7 9 ,
304 159, 192,268, 321
Hydrazinium, Interface, 2 6 - 2 9 , 2 0 2 , 2 1 4
hydrazinium azide N H N , 2 5 3
Intermediate biradical, 117
2 3 5 , 2 3 6 , 293 Intermediate products, 10, 25, 29,
hydrazinium chloride N H C 1 2 5
4 3 , 7 8 , 8 0 , 198,237
( H C ) , 235 Internal rotation, 38, 41, 50
hydrazinium diperchlorate Intramolecular oxidation, 41, 89,
N H ( C 1 0 ) , 237
2 6 4 2
274
hydrazinium Intramolecular rearrangement,
hexafluorophosphate 138,229
N H P F , 233
2 5 6
Iodonitrobenzene,
hydrazinium iodide 234 /?-iodonitrobenzene 62
hydrazinium nitrate N H N 0 2 5 3 Ionic lattice, 25
( H N ) , 231, 234, 286 Ionization, 12, 53, 62, 147, 151,
hydrazinium perchlorate, 236 152, 156, 160, 161, 164, 170,
hydrazinium picrate 172, 176, 178, 198,206
N H C H ( N 0 ) 0 , 233
2 5 6 2 2 3 Isocyanates, 39, 114
Hydrazoic acid, 96, 235, 236
356 Subject Index

Isomerization, 12, 23, 41, 53-55, water detachment


64-66, 96, 101, 111, 114, 118, mechanism, 54
119, 121, 122 zone mechanism, 29, 30
isomerization o f nitroalkanes, M e l t i n g temperature, 2, 2 1 , 22,
54 24, 25, 80, 88, 119, 181-183,
Isoxalylazides, 101 228, 238, 254, 255, 256, 260,
281,282
Kinetic parameters, 41, 45, 65, Mercury fulminate combustion,
68, 100, 119, 121, 128, 135, 136 271,272
138, 146, 161, 182, 195, 196, Metal nitrates, 185, 255, 256
259, 272, 280, 284, 286, 299, Metal perchlorates, 2 5 1 , 2 5 3 ,
300,302,303,318 254
K M n 0 , 209, 321
4 l,5-methane-3,7-dinitro-l,3,5,7-
K N ( N 0 ) , 159, 160, 161, 164
2 2 tetraaza-cyclo-octane, 85
Method,
Laser heating, 52 ampoule method, 115
Lattice inhibitory effect, 24, 25 calorimetric method, 10, 227
Law, chemical activation method,
Darcy's law, 273 109
Layered systems, 321, 337, 338 classical thermodynamics
L i C 1 0 , 2 0 1 , 2 0 6 - 2 0 9 , 254
4 methods, 145
L i F , 208-210 differential scanning
L i m i t i n g stage, 6, 65, 95, 142, calorimetry method, 79, 88
233,242 differential thermal analysis
L i t h i u m fluoride, 206, 321 method, 10, 79, 8 8 , 2 1 0 , 233
L i t h i u m perchlorate, 321, 345 electron paramagnetic
Long-chain molecules, 23 resonance ( E P R ) method, 26
F T I R method, 88
Mass combustion rate, 308, 325 gas-liquid chromatography
Mechanism, method, 88
Belyaev-ZePdovich high-vacuum pyrolysis
mechanism, 297 method, 10, 11,74
bimolecular mechanism, 54. ignition method, 88
88 isothermal calorimetry
mechanism o f cycle opening, method, 88
64 linear pyrolysis method, 195
molecular mechanism, 11, 35, liquid chromatography
3 6 , 5 1 , 6 4 , 65 method, 88
monomolecular mechanism, manometric method, 10, 88,
1 1 , 4 3 , 6 1 , 8 8 , 110 221,227
radical mechanism, 35, 40, M I N D O / 3 method, 47, 52, 54
4 2 , 6 1 , 6 6 , 107, 108, 136 quantitative thermography
redox (oxidation-reduction) method, 195
mechanism, 198 termogravimetric analysis
stepwise biradical method, 88
mechanism, 118 thermochemical method, 49
surface mechanism, 29 thermocouple method, 284
Thermal Decomposition and Combustion 357

thermogravimetry method, N H N ( N 0 ) , 160, 161, 166, 168,


4 2 2

10, 227 169, 291


time-of-flight mass N H O C 1 0 , 243, 244
4 4

spectrometer method, 195 N ( N 0 ) " anion, 171 2 2

Zel'dovich—Frank- 4-N0 C H CH N ,97 2 6 4 2 3

Kamenetskii method, 271, N H , 232, 234-237, 240, 274,


2 4

303 289
M e t h y l a m m o n i u m perchlorate, N H , 231,232, 235,237
2 5
+

284,345 N H ,237 .
2 6
2 +

Methylazide, 98 N H C 1 0 , 2 3 6 , 2 3 7 , 238
2 5 4

Methyldiphenylurea, 228 N H (C10 ) ,237,238


2 6 4 2

Methylisocyanate, 115 N 0 , 4 6 , 130, 132, 1 5 1 , 2 2 5 , 2 4 8


2 4

Methylnitrate, 127, 128 N 0 , 149-152, 170, 175, 176,


2 5

3-methylnitro-butane 36 244,246,247
Methylnitrofurazane, 118 N H \ 235-237
4 9

1- methyl-2-nitrofurazane, 115 N 0 , 1 6 7 , 172'


4 6

2- methylnitro-propane 36 N H C 1 , 231, 232, 234-240, 288,


2 5

2-methyl-3-nitropropen 41 289
M o d e l o f combustion, N H I , 234-238
2 5

"chemical arc", 321, 324-326 N H N , 2 3 5 , 2 3 6 , 238


2 5 3

model with averaging, 333, ( N H 0 ) S 0 , 238,239,244


4 2 4

336 [ N F ( N 0 ) C C H 0 ] C H , 137
2 2 2 2 2 2

quasihomogeneous fuel, 321, 2 , 4 - ( N 0 ) C H S C ( N 0 ) , 45 2 2 6 3 2 3

337 2,4-(N0 ) C H SC(N0 ) CH , 2 2 6 3 2 2 3

"sandwich" type, 321, 327, 45


329-331,335-336, 338 ( N 0 ) C C H ( O A c ) C H , 55
2 3 3

two-dimensional model, 336 ( N 0 ) C C H C H , 55


2 3 2 3

M o d i f i e d Polyanyi-Wigner (N0 ) CCH —CH —


2 3 2 2

equation, 27 O C H C ( N 0 ) , 55 2 2 3

Molecular beam, 52, 73 (N0 ) CCH —CH —


2 3 2 2

Molecular crystals, 22 C O O C H C ( N 0 ) , 55 2 2 3

M o l e c u l e stability, 1, 127 (N0 ) C(CH )-C(NF ) CH ,137


2 3 2 2 2 3

Monomolecular decay, 2 (N0 ) C(CH ) OOCCH -


2 3 2 2 2

Mononitroalkanes, 10, 12, 35, CH(NF )CH NF ,137 2 2 2

36, 62 (N0 ) CCH CH COOCH -


2 3 2 2 2

Mononitrocompounds, 42, 108 C ( N 0 ) F , 137 2 2

M u l t i p l e charged cations, 206, ( N 0 ) C C H O C H C H , 55


2 3 2 2 3

253 ( N 0 ) C C ( N 0 ) , 48
2 3 2 2

( N 0 ) C C ( N 0 ) , 44
2 3 2 3

ND N(N0 ) ,228
4 2 2
( N 0 ) C ( N 0 ) C - C H , 49
2 3 2 2 3

N F C ( N 0 ) , 136
2 2 3
( N 0 ) C - C ( N 0 ) , 44, 49, 55
2 3 2 3

N H , 79, 95, 142, 175, 176, 178-


3
N ( C H N N 0 ) C H N , 24
3 2 2 4 2 3

181, 183, 189, 196, 197, 200, N C H [ N ( N 0 ) C H ] N , 137


3 2 2 2 4 3

205, 225, 227, 235, 239, 242- ( N H 0 ) P 0 , 238,239,244


4 3 4

244,289, 291,292,312 2,4,6-(N0 ) C H N(N0 )CH ,24 2 3 6 2 2 3

N H 0 C 1 , 241-244
4
[(N0 ) CCH ] BrN0 ,44 2 3 2 2 2

[(N0 ) CCH ] CHN0 ,44 2 3 2 2 2


358 Subject Index

[ ( N 0 ) C C H ( C H ) N H ] C = 0 , 55
2 3 3 2 Nitroethane, 12, 36, 38, 39, 53
[(N0 ) CCH ] CC1N0 ,44
2 3 2 2 2 Nitrohexane,
[ ( N 0 ) C C H ] C F N 0 , 44
2 3 2 2 2 1- nitrohexane, 36
[(N0 ) CCH ] NN0 ,I37
2 3 2 2 2 2- nitrohexane, 36, 37
[ ( N 0 ) C C H N H ] C = 0 , 55
2 3 2 2 Nitrogen dioxide, 127, 258
[(N0 ) CCH N(N0 )CH ] ,24,
2 3 2 2 2 2 Nitroglycerin, 127-130, 132,
137 212-214, 273
Naphthalene, 332 Nitroglycol, 128
Nitramines, 9, 10, 11, 13, 73-77, Nitromethane, 12, 42, 43, 50-55,
79, 80, 88, 89, 98, 1 3 8 , 3 1 3 , 3 1 4 57
Nitrates, 127, 130, 145, 149, 185, 2-nitro-3-methylbutane, 36, 37
212, 213, 231, 251, 255-257, 2- nitro-2-methylbutane, 36, 37
264,286 3- nitro-3-methylpentane, 37
monocellulose nitrate, 131 2-nitro-2-methylpropane, 36
N i t r i c acid, 9, 13, 130, 146, 147, 2-nitro-3-methylpropane, 37
154, 157, 159, 160, 166-169, N-nitro-N-fluoroamines, 88
171, 172, 175, 177, 180, 183, Nitronic acids, 12, 53, 54, 57, 64,
185, 186, 224, 244, 247, 248, 6 5 , 6 6 , 170
285 Nitronium perchlorate, N 0 C 1 0
2 4

N i t r i c oxides, 13, 88, 130, 256, 198,245


293 Nitropentane,
Nitrileimine, 122 1-nitropentane, 36
Nitrile-oxides, 113, 114, 118 Nitrophenol,
Nitriles, 113, 114 weta-nitrophenol, 212
1- nitroadamantane, 46 or/Zzo-nitrophenol, 212
Nitroalkane, partf-nitrophenol, 212
halogenated nitroalkanes, 53 N-nitropipyridine, 81
Nitroalkene, 4 1 , 42 Nitropropane,
Nitroaniline, 1- nitropropane, 36
w-nitroaniline, 61 2- nitropropane, 36
p-nitroaniline, 61 N-nitropyrolidine, 81
Nitrobenzene, 12, 13, 52, 61-65, Nitrosocompounds, 78, 80
185,224 Nitrostyrene, 41
Nitrobutane, Nitrotoluene,
1- nitrobutane, 36 w-nitrotoluene, 61
2- nitrobutane, 37 o-nitrotoluene, 65, 66
Nitrocompounds, p-nitrotoluene, 61, 62, 63
aliphatic nitrocompounds, 13, N 0 group, 4 0 , 5 0 , 5 1 , 105
2

35,36, 43,53 Nonequilibrium imperfections,


aromatic nitrocompounds, 9, 25
11, 1 3 , 6 1 , 6 8 - 7 0 , 185 Nusselt number, 214, 322
2- nitro-3,3-dimethylbutane, 36
2-nitro-2,3-dimethylbutane, 37 0 NCH -CH N9 ,49
2 2 2 2

Nitrodiphenyl, 2,4,6,2 ,4,6,4 ,6 -octanitro-w-


o-nitrodiphenyl, 62 terphenyl ( O N T ) , 66
Nitroesters, 9, 12, 13, 98, 99,
127-132,212,213,216
Thermal Decomposition and Combustion 359

Octogen ( H M X ) , 22, 24, 73-75, Phenol-formaldehyde polymers,


77, 80, 84, 88, 99, 313, 314, 321, 212
338, 340-342 Phenylazides, 99
Olefins, 37, 3 9 , 4 1 , 4 5 Phenylisocyanate, 114
O n i u m salts, 9, 141, 142, 144, Picric acid, 67, 78, 234
145, 146, 159, 175, 189, 226, /ra-picrylaminotriazine
229, 237, 238, 251, 275, 283, ( T P A M T ) , 66,71
284,288,290-294,313,344 Plastic binders, 113
Organic azides, 9, 95, 96 Polar coupling, 62
Organometal compounds, 210, Polarizability, 38, 254
217,219,326 Polyacrylonitrile, 211
Polybutadiene, 210, 211, 342
ort/zoderivatives, 61, 64 Polyester rubber, 333
ort/zoazidobiphenyls, 99 Polyethylene ( P E ) , 325, 326
Oscillations, Polymerization, 110, 114, 271
pendular oscillations, 50 Polymers ( C H N 0 C H ) , 7 9
2 2 2 n

Oxadiazole, 114 Polymethyl methacrylate


1,2,4-oxadiazoles, 114 ( P M M A ) , 325
3,5-diphenyl-l,2,4- Polynitroalkanes, 10-13, 135
oxadiazole ( O D A ) , 114 Polynitrocompounds, 43, 47, 55,
3-methyl-5-phenyl- 70, 136, 138
1,2,4-oxadiazole, 114 Polypropylene, 333
Oxidation-reduction processes, Polystyrene, 210, 2 1 1 , 3 4 3
200 Polyvinyl chloride, 211
O x o n i u m perchlorate H 0 C 1 0 " , 3
+
4
Positive mesomeric effect, 40
155 Potassium dinitramide, 257-263
Oxynitrate, 256 Powder, 3, 5, 6, 2 0 7 , 2 7 1 , 2 7 2 ,
288,333
Pentaerythritol tetranitrate Precipitation, 26, 27, 29, 30, 197
( P E T N ) , 4, 22, 24, 128, 129, 185 Pressures,
Perchlorates, 145, 156, 206, 231, high pressures, 55, 88, 277-
241,243,245,251-256,284,285 279, 281, 285, 291, 310, 312,
Perchloric acids, 141, 146, 170, 320,338
176, 2 3 1 , 2 4 6 , 275 high static pressures, 130
Perchloric anhydride, 157 low pressures, 11, 54, 65, 76,
Perchloryl fluoride C 1 0 F , 208 3
107, 257, 271, 273, 277-279,
Permittivity, 151, 168, 170, 247 283,291-293,310,312,338
P h C H N , 9 7 , 98
2 3
PrN ,
3

P h C H N 0 , 44, 47, 51, 105


2 2
A7-PrN , 97
3

P h S C ( N 0 ) , 45
2 3
/<>o-PrN ,973

P h S C ( N 0 ) C H , 45
2 2 3 Promotors, 10
P h S 0 C ( N 0 ) C H , 45
2 2 2 3
1,2-bis(difluoroamino)propane,
Phase transition, 6, 7, 16, 22, 108
190-193, 197, 2 0 2 , 2 0 6 , 208, Propellants, 1, 3-6, 98, 99, 1 13,
209,307,339 146, 189, 210, 221, 245, 267,
Phenol, 63 273, 275, 282, 313, 318, 321,
330, 332, 336, 338-340, 342-344
360 Subject Index

Propylene oxide, 326 Rearrangement,


1.2- propyleneglycol-dinitrate, nitro-nitrite rearrangement,
128 51-53,62, 73
1.3- propyleneglycol-dinitrate, skeleton rearrangement, 95,
128 102, 122
ft-propylnitrate, 128 Recrystallization, 22, 190, 239
Protonated nitroamide, 179 Relaxation time, 142
Prototropic tautomerism, 122, Resonance stabilization, 99
123 Resonance structure, 171
Pyridine, 53 Reversible isomerization, 121
Pyroxylin,131,271,272 Reynolds numbers, 322
pyroxylin combustion, 272 Rotation,
free rotation, 50, 106
Quasi-heterolytic processes, 38 free three-dimensional
Quasihomogeneous composites, rotation, 52
315 internal rotation, 38, 41, 50
Quasihomogeneous fuel, 321, R R K M theory, 5, 50
337
Secondary nitramines, 9, 13, 73,
Radiation treatment, 25, 201, 75, 79, 80, 88
204, 205 Self-ignition, 5, 1 5 4 , 2 1 3 , 2 1 5 ,
Radical, 248, 275, 308
1-adamantane radical, 46 Semi-ionic pairs, 39
nitroalkyl radical, 48, 49 Singlet nitrene ( R N : ) , 95
nitroxyl radicals, 64, 67, 68 Sodium azide, 26, 27, 28
R D X , 4, 15, 24, 73-75, 77-80, Solid rocket propellant, 4, 5, 6,
82, 8 3 , 8 4 , 1 2 7 , 3 1 3 , 3 1 4 98, 99, 113, 189, 2 2 1 , 2 6 7
Reactions, Solution,
autocatalytic reactions, 5 in acetone, 81-83, 88
branched chain reaction, 157 in benzene, 81, 83
chain termination reaction, in benzophenone, 83
157, 158 in dibutylphthalate, 85-86
heterogeneous reactions, 11, in dicyclohexylphthalate, 83
36, 80 in w-dinitrobenzene, 82
parallel reactions, 6, 176, in ethanol, 82
244, 297, 299, 3 0 0 , 3 1 3 in metadinitrobenzene, 84,
polymer oxidation reaction, 87, 88
211 in nitrobenzene, 82
reaction o f disproportionation in /so-octane, 80, 81
11,259 in polyphenyl ether, 83
recombination reaction, 10 in trinitrobenzene, 81, 82
retro-1,3-dipolar cyclo- in trinitrotoluene, 83
attachment reaction, 113 Solvation, 55, 110, 164, 229
secondary chain-radical Solvents, 12, 67, 7 8 , 9 8 , 9 9 , 151,
reactions, 61 152,247
tautomeric transformation Specific reaction rate, 14, 15
reaction, 170 Spectrophotometric analysis, 161
Thermal Decomposition and Combustion 361

Spectrophotometric effective combustion


determination, 221 temperature, 301, 302
Stability criterion, 279, 282, 290 melting temperature, 2, 21,
Stabilization, 4, 9, 40, 48, 53, 22, 24, 25, 80, 88, 119, 181,
69, 99, 100, 121, 130-132,208, 182, 183, 228, 238, 255, 256,
217, 2 2 1 , 2 2 4 260,281,282
Stabilizer, 130, 2 0 8 , 2 2 8 Terphenyl, 64
Stage, Tetrahydrofurane, 326
catalytic stage, 68, 79, 146, tetranitro-2,3,5,6-dibenzo-
147, 224, 227 1,3a,4,6a-tetraazapentalen
non-catalytic stage, 146, 153, (Takot-1), 67
226 tertanitro-2,3,4,5-dibenzo-
Steady-state sublimation, 144, 1,3a,6,6a-tetraazapentalen
145 (Takot-2), 67
Steric barrier, 37, 110 Tetranitroethylene, 45
Steric hindrances, 47 1,4,6,9-tetranitro-1,4,6,9-
Stilbene, 109 tetraazadecalin, 24
Stoichiometry, 143, 197, 240, 1,3,5,7-tetranitro-l,3,5,7-
283,323 tetraazacyclooctane ( H M X ) , 22,
Structure, 24, 73, 74, 75, 77, 80, 84, 88, 99,
plane structure, 51, 117 313,321,338,340, 341,342,
Sublimation, 6, 22, 24, 80, 144, Tetrahedral form, 51
145, 195, 199-200, 202, 211, Tetralin, 120
226, 227, 269, 275, 276, 280- Tetramethylenefurazane 118
282, 294, 307-309, 315, 316, 2,4,2,4 -tetranitrodiphenylamine
318, 319, 324, 327, 329, 333, ( T N D P h A ) , 66
339, 344 Tetraphenylpyrazine (TPhP), 98
Substitute, Tetrazine, 122.
electronegative substitute, 48, Tetrazole, 9, 10, 101, 113, 119,
50, 88, 127 120, 121, 122, 123
meta-substitutes, 61 1,5-disubstituted tetrazoles,
ortho-substitutes, 64, 99 119
para-substitutes, 61 1,5-diaryltetrazoles, 119
/-sulfone ( H N D S u l f o n ) , 66 N-naphthyltetrazoles, 119
Sulfuric acid, 161, 162, 164, 166, 1 -(2-pyridyl)-5-aryltetrazole,
168 120
Supercritical conditions, 54 T e t r y l , 4 , 1 3 , 7 8 , 7 9 , 86
Surface-to-volume ratio, 10 Tetryl picric acid, 13
Thermal effect, 5, 38, 218, 272,
Tautomeric equilibrium, 170 276, 277, 279, 288, 297, 300,
Temperature, 318,319,337,339,345
boiling temperature, 146, Thermal explosion, 5
275,297 Thermal instability, 278, 282,
combustion temperature, 6, 313
2 2 1 , 2 7 4 , 2 7 5 , 2 9 8 - 3 0 2 , 340 Thermal stability, 127, 132, 138,
critical temperature, 5, 180 157, 185, 210, 212, 215, 217,
362 Subject Index

221, 231, 234, 244, 246, 247, w-trinitrophenol, 62


253,255,278,285,343 2,4,6-trinitrophenyl ( T N P h ) , 66
Thermographic technique, 185 Trinitrophloroglucinol ( T N P h G ) ,
Thermogravimetry, 10, 227 66
T h i o c o l , 212, 216, 217 Trinitroresorcin, 68
Toluene, 13, 63 2,4,6-trinitrotoluene ( T N T ) , 4,
Topochemical acceleration, 70 56, 66, 68, 69
Topochemical peculiarities, 198 Trinitrotoluol, 65, 67
trans-1,4,5,8-tetranitro-1,4,5,8- Trinitrotriaminobenzene, 70
tetraazadecalin, 24 l,3,5-trinitro-2,4,6-
Transition state, triazidobenzene 23
polar transition state, 12, 38, l,3,5-trinitro-l,3,5-
39 triazacyclohexane ( R D X ) , 4, 15,
semi-rigid transition state, 52 24, 73, 74, 75, 77—80, 82, 83,
Triaminotrinitrobenzene 84, 88, 1 2 7 , 3 1 3 , 3 1 4
( T A T N B ) , 66 Tripicryl-s-triazine ( T P T ) , 67
1.3.5- triamino-2,4,6- Triphenylamine ( T P h A ) , 98
trinitrobenzene 185 Tungsten, 2 0 1 , 2 9 1 , 3 2 5 , 3 3 3
Triazine, 77
s/w-triazine, 122 Ultimate stress, 31
Triazinylazides, 100 Unpaired electrons, 117
Triazole, 122 Unsaturated nitrocompounds, 41
Trinitroaniline, 66, 67 Unshared electron pair, 96, 100
2.4.6- trinitroaniline ( T N A ) , 66 Urotropine, 224
Trinitroanisole, 68 Urotropine nitrolysis, 127
2,4,6-trinitroanisole ( T N A n ) , 66
s/w-trinitrobenzene ( T N B ) , 66 Validation, 2 6 9 , 3 1 4 , 3 3 9
Trinitrocarbazole ( T N C a r b ) , 66 van der Waals radius, 48, 74
Trinitrocresol, 66, 68 Vibrational pre-melting, 21
2,4,6-trinitrocresol ( T N C r ) , 66 Vibrations,
2,4,6-trinitroethylbenzene asymmetrical deformation
( T N E B ) , 66 vibration, 50
2,4,6-trinitromesethylene pendular vibrations, 50
( T N M e s ) , 66 torsional vibrations, 50
Trimethylammonium valence vibration, 56
perchlorate, 284 Vinylazides, 121
Trinitromethyl group, 55, 57
1,1,1-trinitroalkanes, 35 Water,
1,1,1-trinitroethane, 13, 43, 51 redistribution o f water, 194
Trinitromesethylene, 66
1,3,7,9-tetranitrophenothiazine- yield
5,5-dioxide ( T N P h T h D ) , 66 nitrile yield, 45
Trinitrophenylenediamine
( T N P h D A ) , 66

You might also like