You are on page 1of 246

Atmospheric Oxidation

and Antioxidants
VOLUME I
This page intentionally left blank
Atmospheric
Oxidation and
Antioxidants
VOLUME I

G. Scott, editor
Green Ridge
Newby Nr. Middlesbrough
Cleveland TS8 OAH
U.K.

ELSEVIER
AMSTERDAM - LONDON - NEW YORK - TOKYO 1993
ELSEVIER SCIENCE PUBLISHERS B.V.
Sara Burgerhartstraat 25
P.O. Box 211,1000 AE Amsterdam, The Netherlands

ISBN: 0-444-89615-5 (volume I)


0-444-89616-3 (volume II)
0-444-89617-1 (volume III)
0-444-89618-x (set)

© 1993 Elsevier Science Publishers B.V. All rights reserved.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any
means, electronic, mechanical, photocopying, recording or otherwise, without the prior written permission of the
publishers, Elsevier Science Publishers B.V., Copyright & Permissions Department, P.O. Box 521, 1000 AM
Amsterdam, The Netherlands.

Special regulations for readers in the U.S.A. This publication has been registered with the Copyright Clearance
Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from the CCC about conditions under
which photocopies of parts of this publication may be made in the U.S.A.. All other copyright questions,
including photocopying outside of the U.S.A., should be referred to the copyright owner, Elsevier Science
Publishers B.V., unless otherwise specified.

No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of
products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions
or ideas contained in the material herein.

This book is printed on acid-free paper.

Printed in The Netherlands.


V

CONTENTS

Preface vii
List of authors ix

CHAPTER 1. AUTOXIDATION AND ANTIOXIDANTS:


HISTORICAL PERSPECTIVE, by Gerald Scott 1
1. The Ageing of Rubber 1
2. Hydroperoxides and the Autoxidation Chain Reaction 2
3. Antioxidants 4
4. Synergism 23
5. The Effect of the Reaction Environment upon Oxidation and
Antioxidant Mechanisms 24
6. Biological Oxidation 30
7. The Present Position 36
References 38

CHAPTER 2. AUTOXIDATION, by S. Al-Malaika 45


1. Introduction 45
2. Mechanism and Kinetics of Autoxidation 46
3. Technological Effects of Oxidation 68
4. Other Techniques used for Oxidation Studies 78
References 78

CHAPTER 3. INITIATORS, PROOXIDANTS AND SENSITISERS,


by Gerald Scott 83
1. Reactions of Oxygen 83
2. Peroxides 102
References 116

CHAPTER 4. ANTIOXIDANTS: CHAIN-BREAKING


MECHANISMS, by Gerald Scott 121
1. The Chain-Breaking Donor Mechanism 121
2. Chain-Breaking Hydrogen (Electron) Acceptor Mechanism . . . 140
3. Catalytic Antioxidants 142
References 157
VI

CHAPTER 5. ANTIOXIDANTS — PREVENTIVE MECHANISMS,


by S. Al-Malaika 161
1. Catalytic Peroxidolytic Mechanisms and the Role of Sulphur-
Containing Compounds 164
2. Stoichiometric Peroxidolytic Mechanisms (PD-S) and the Role
of Phosphite Esters 208
3. Metal Ion Deactivation 211
4. Ultra-Violet Light Deactivation 218
References 221
Subject Index 225
vii

PREFACE

Oxidation by molecular oxygen is one of the most practically important of


all chemical processes. It is the basis of energy production in animals and,
at the same time, a major cause of irreversible deterioration and ultimate
death. Man uses oxygen positively in the production of energy by combus­
tion, and many important industrial processes in the petrochemical in­
dustry are based on the controlled oxidation of hydrocarbons. At the same
time, oxidation is the main cause of deterioration of foodstuffs and of many
industrial polymers.
It is clearly of great practical importance that the mechanisms of oxida­
tion and its prevention should be understood in order to utilise the reactions
of oxygen more effectively but, equally importantly, to control the adverse
effects of oxygen on man-made products and in biological systems. The three
volumes of this work are directed toward these objectives. Volume I reviews
current understanding of autoxidation, largely on the basis of the reactions
of oxygen with characterised chemicals. From this flows the modern mech­
anisms of antioxidant action and their application in stabilisation technol­
ogy. Volume II examines the oxidation chemistry of carbon-based materials
in more detail with emphasis on the technological phenomena that result
from the attack of oxygen and the practical procedures developed to prevent
them. Volume III addresses our present understanding of how oxidation is
involved both positively and negatively in life processes. This is a more
recent and rapidly developing aspect of oxidation chemistry and many of the
concepts still have to be proved by rigorous scientific investigation. Never­
theless, the mechanistic principles developed as a result of studies in vitro
over the years now provide the basis for understanding the complex oxida­
tion chemistry of life processes and its control by biological antioxidants.
The three volumes, although complementary to one another, form a single
whole and it is hoped that, by frequent cross-reference, the reader will be
enabled to utilise ideas and experience from other disciplines to enlighten
his own. The first edition of this work was published a quarter of a century
ago in a single volume. The increase in size of the second edition reflects the
growth of interest in the subject in the intervening period. Nevertheless, the
mechanisms outlined in the first edition still form the basis of our present
understanding of oxidation chemistry and there will therefore be frequent
reference to it in this edition.
GERALD SCOTT
This page intentionally left blank
IX

LIST OF AUTHORS

S. AL-MALAIKA (Volume I)
Department of Chemical Engineering and Applied Chemistry, Aston
University, Aston Triangle, Birmingham B4 7ET, U.K.

N.C. BILLINGHAM (Volume II)


School of Chemistry and Molecular Sciences, University of Sussex,
Brighton BN1 9QJ, U.K.

JOHN A. BLAIR (Volume III)


Pharmaceutical Sciences Institute, Aston University, Aston Triangle,
Birmingham B4 7ET, U.K.

E.B. BURLAKOVA (Volume III)


The Institute of Chemical Physics, Kosygin-str. 4, Moscow 117334,
Russia

G. CAMINO (Volume II)


Dipartimento di Chimica Inorganica, Università Degli Studi di Torino,
Via Pietro Giuria, 10125 Torino, Italy

D.J. CARLSSON (Volume II)


Division of Chemistry, National Research Council of Canada, Ottawa,
K1A 0R9, Canada

T. COLCLOUGH (Volume II)


Exxon Chemical Technology Centre, Abingdon, Oxon. 0X13 6BB, U.K.
NANCY E. DAVIDSON (Volume III)
Oncology Center, Johns Hopkins Medical Institutions, 615 N. Wolfe
Street, Baltimore, MD 21205, U.S.A.

H.H. DRAPER (Volume III)


Department of Nutritional Sciences, University of Guelph, Ontario,
Canada NIG 2W1
X

H. BRIAN DUNFORD (Volume III)


Department of Chemistry, University of Alberta, Edmonton, Alberta,
Canada T6G 2G2

GILL FARRAR (Volume III)


Pharmaceutical Sciences Institute, Aston University, Aston Triangle,
Birmingham B4 7ET, U.K.

JOHN M.C. GUTTERIDGE (Volume III)


National Institute for Biological Standards and Control, Blanche Lane,
South Mimms, Potters Bar, Herts. EN6 3QG, U.K.

KATHRYN Z. GUYTON (Volume III)


Department of Environmental Health Sciences, Johns Hopkins
Medical Institutions, 615 N. Wolfe Street, Baltimore, MD 21205, U.S.A.

BARRY HALLIWELL (Volume III)


Department of Biochemistry, King's College (KQC), Strand Campus,
London WC2R 2LS, U.K.

THOMAS W. KENSLER (Volume III)


Department of Environmental Health Sciences and Department of
Pharmacology and Molecular Sciences, Johns Hopkins Medical
Institutions, 615 N. Wolfe Street, Baltimore, MD 21205, U.S.A.

S.P. KOCHHAR (Volume II)


SPK Consultancy Services, 48 Chiltern Crescent, Earley, Reading RG6
IAN, U.K

R.P. LATTIMER (Volume II)


The B.F. Goodrich Research and Development Center, Brecksville, OH
44141, U.S.A.

R.W. LAYER (Volume II)


The B.F. Goodrich Research and Development Center, Brecksville, OH
44141, U.S.A.

DIANA METODIEWA (Volume III)


Department of Chemistry, University of Alberta, Edmonton, Alberta,
Canada T6G 2G2

ETSUO NIKI (Volume III)


Department of Reaction Chemistry, The University of Tokyo, Hongo,
Bunkyo-ku, Tokyo 113, Japan
XI

Z. OSAWA (Volume II)


Faculty of Engineering, Gunma University, Kiryu, Gunma 376, Japan

D.G. POBEDIMSKIJ (Volume III)


The Kazan Institute of Chemical Technology, Karl Marx-str. 68, Kazan
420015, U.S.S.R.

GREGORY A. REED (Volume III)


Department of Pharmacology, Toxicology, and Therapeutics,
University of Kansas Medical Center, Kansas City, KS 66103, U.S.A.

C.K. RHEE (Volume II)


The Uniroyal Goodrich Tire Company, Brecksville, OH 44141, U.S.A.

TADEUSZ SARNA (Volume III)


Department of Biophysics, Institute of Molecular Biology, Jagiellonian
University, A. Mickiewicza 3, 21-120 Krakow, Poland

GERALD SCOTT (Volumes I, II and III)


Department of Chemical Engineering and Applied Chemistry, Aston
University, Aston Triangle, Birmingham B4 7ET, U.K.

HAROLD M. SWARTZ (Volume III)


University of Illinois, College of Medicine at Urbana-Champaign, 506
South Mathews, Urbana, IL 61801, U.S.A.

PAUL J. THORNALLEY (Volume III)


Department of Chemistry and Biological Chemistry, University of
Essex, Wivenhoe Park, Colchester C04 3SQ, Essex, U.K.

PETER WARDMAN (Volume III)


Cancer Research Campaign, Gray Laboratory, P.O. Box 100, Mount
Vernon Hospital, Northwood, Middlesex HA6 2JR, U.K.
This page intentionally left blank
1

Chapter 1

AUTOXIDATION AND ANTIOXID ANTS:


HISTORICAL P E R S P E C T I V E

GERALD SCOTT

1. THE AGEING OF RUBBER

The phenomena associated with oxidation were recognised long before the
mechanism of autoxidation was developed. Many materials used by man
were known to undergo slow deterioration in the atmosphere. The earliest
investigations of oxidation were carried out on the first technologically
important polymer, natural rubber. Hoffman [1] has been credited with the
discovery that "perishing" of rubber involves the absorption of oxygen. In
retrospect, natural rubber and particularly vulcanised natural rubber, is the
most difficult medium that could be envisaged for the fundamental study of
chemical reactions. Not only is it chemically "impure" when first prepared,
but from the moment at which the latex leaves the tree, oxidation reactions
of both the hydrocarbon and non-hydrocarbon constituents are initiated.
Vulcanisation with sulphur increases the complexity of the system and it
ceases to behave like a simple hydrocarbon. More recent research has shown
that the sulphur cross-link behaves initially as an accelerator of oxidation
(pro-oxidant) but that oxidation products of the sulphides in the cross-link
retard the oxidation process [2]. The oxidation chemistry of sulphur com­
pounds will be discussed in later chapters but it should be noted that the
complex behaviour of vulcanised rubber almost certainly accounts for the
phenomenological approach subsequently adopted by scientists and technol­
ogists in studying its deterioration under practical conditions. The techno­
logical and even the scientific literature contain many anthropomorphic
terms such as "ageing", "fatigue", "perishing" and "poisoning" to describe the
loss of useful properties of rubbers caused by oxidation. These terms reflect
the attempts of early rubber technologists to understand technological
changes by biological analogy. They believed that rubber was in some way
"alive" as it left the rubber tree. It is interesting that biological ageing is
increasingly being seen [3] to involve the same chemical processes that
attracted the attention of early rubber scientists. Somewhat ironically, the
study of the oxidation of the polyunsaturated fatty esters (lipids), which is
2 GERALD SCOTT

believed to be the reason for biological ageing, was pioneered by technolo­


gists with practical interests in the deterioration of foodstuffs and in the
"drying" of polyunsaturated fatty esters used in paint technology [4].
A characteristic of both processes is the existence of an "induction period"
before changes in the properties of the materials can be detected. The reason
for this was recognised at the beginning of the present century when Genthe
[5] was able to show that the characteristic autoaccelerating oxygen absorp­
tion curve [6] was associated with the initial absence and subsequent build
up of hydroperoxides in linseed oil. The addition of partially oxidised turpen­
tine confirmed the conclusion that peroxides are the main initiators for
oxidation. The addition of other peroxides was also found to reduce or
eliminate the induction period. The discovery of peroxides as auto-initiators
for oxidation by molecular oxygen was later found to apply to other sub­
strates, including rubber [6]. It proved to be a key to the development of
antioxidation and antioxidant theory since it applies across the whole
spectrum of autoxidation phenomena. Since peroxides were first recognised
as being important as initiators of oxidation, they have never been far from
the centre of the autoxidation mechanisms.

2. HYDROPEROXIDES AND THE AUTOXIDATION CHAIN REACTION

The chemical structure of peroxides remained rather obscure for many


years. Since unsaturation was found to disappear when olefins were oxi­
dised, it was generally assumed that addition of oxygen to the double bond
was involved and Staudinger [8] gave the name "moloxide" to the four-mem-
bered ring structure, I, believed to be formed,

(a) | |
=
Iv2C CR.2 * R2C—CR2 (1)

An analogous structure (II) was assigned to cyclohexene peroxide by


Stephens [9] but the elegant studies by Criegee [10] and Farmer [11] and
their co-workers, showed that the chemistry of cyclohexene peroxide is
inconsistent with the structure II but can be accounted for by the hydro-
peroxide structure, III.

H Hk /OOH
r^S—O
k>° H
II III
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 3

The elucidation of the structure of cyclohexene and related peroxides


formed by the reaction of oxygen with hydrocarbons prepared the way for
the general theory of hydrocarbon autoxidation [12]. Bäckström had applied
Bodenstein's chain reaction theory to the autoxidation of aldehydes in the
presence of light [13]. Unlike earlier workers who discussed chain reactions
in terms of energetic intermediates, Bäckström recognised the free radical
nature of the peroxyl radical intermediate and formulated clearly for the
first time the now well-known radical chain process: Reactions (2) and (3).

RC=0 > RCOO (2)

O O
II II
RCOO+RCHO > RCOOH + R C = 0 (3)
In the hands of Bolland, Bateman and Gee and their co-workers [14-23],
at the British Rubber Producers Association, this theory provided a basis for
understanding the role of the hydroperoxide in the auto-initiation Reaction
(4),
A,ÄV
ROOH » RO+OH- (4)

and by analysis of the kinetics of olefin oxidation they were able to show that
the rate constant of Reaction (5) is normally very much faster than that of
Reaction (6).
OO
I
R'CH = CH-CH- + 0 2 > R'CH=CHCH-
(RO (ROO) (5)

ROO + R'CH = CHCH 2 - > ROOH + R'CH= CHCH- (6)

and hence the overall rate at normal oxygen pressures is to a large extent
dependent on the rate constant of Reaction (6).
The full implications of the Bolland, Bateman, Gee mechanism will be
discussed in Chapter 2 of this volume but some of them have great signifi­
cance for the subsequent development of antioxidant theory and will be
briefly noted here:
(a) Termination at atmospheric oxygen pressures is normally by bi-
molecular reaction of two alkylperoxyl radicals. Russell showed [24-26] that
when the peroxyl radical contains an a-hydrogen, this is predominantly by
disproportionation:
4 GERALDSCOTT

R
\ P~°
2RR'CHOO'->C O' > RCOR' + RR'CHOH + O, 07)
R
R HO
I
RCHR'
However, at low oxygen pressures or when the intermediate alkyl radical is
stabilised by resonance, then termination Reactions (8) and (9) involving
alkyl may also occur [21,23].

ROO+R- » ROOR (8)

2R- ► R-R (9)

(b) Although propagation occurs in saturated molecules and in many


olefins by Reaction (5), if there is no reactive méthylène group in an olefin,
as in the case of styrene, then Reaction (5) cannot occur and copolymeriza-
tion results to give a polyperoxide by Reaction (6) [27], as proposed earlier
by Milas [28].

PhCH=CH2 + ROO ► PhCH-CH2OOR

(10)

PhCHOO
I
ROO[CH2CHOO]n-CH2CH-00- *— CH2OOR
I
Ph Ph PhCH
n PhCH=CH2
+ n0 2

3. ANTIOXIDANTS

As we have seen, rubber technology was well advanced before the theory
of autoxidation was developed and the early discovery of protective agents
for rubber was entirely empirical. Thus a patent for the use of phenol and
p-cresol appeared in 1870 [29] and for hydroquinone and pyrogallol in 1901
[30]. An important discovery was that some chemicals (e.g. aniline) used in
the "curing" or "vulcanisation" of rubber could have a profound retarding
effect on subsequent ageing [31]. It was later shown that sulphur-containing
accelerators for vulcanisation (e.g. mercaptobenzthiozole and the thiuram
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 5

disulphides) [32] have similar antioxidant activity and some of the more
effective heat-resistant rubbers used in modern rubber technology are based
upon these discoveries [33].
The development of effective antioxidant systems for the modern motor­
car tyre was also an empirical process. The fact that many rubber manufac­
turers even today will not disclose the formulations used in tyres has made
the study of the mechanisms of rubber antioxidants particularly difficult.
This situation is further complicated by the fact that many commercial
antioxidants are complex mixtures of different chemical species. It was not
until a systematic kinetic study of the individual effects of well-charac­
terised additives in standard formulations was initiated by Shelton and his
co-workers [34] in the early 1950s that some light was thrown on the way
arylamines act as antioxidants.
The beginnings of antioxidant theory go back to the extensive investiga­
tions of Moureau and Dufraisse in the 1920s. Although the mechanisms they
proposed are now of only historical significance, since they were propounded
before the radical chain mechanism of autoxidation had been established,
nevertheless their findings provided a phenomenological basis upon which
the modern theories of antioxidant action are founded.
Moureau and Dufraisse believed that "antioxygens", as antioxidants were
then called, were essentially reducing agents. They proposed that the pri­
mary role of an antioxidant (B) was to react with peroxides (A[021) with
eventual regeneration of oxygen [35]:

A[0 2 ] + B > AO + BO

AO + BO > A + B + 02

Although these authors were unable to define peroxides in chemical terms


(A[0 2 ] indicates a loose association of the substrate A with oxygen), they
nevertheless arrived intuitively at what is now recognised to be one of the
most important mechanisms of antioxidant action: the catalytic conversion
of peroxides to inert products [36,37].
Moureau and Dufraisse were the first authors to report pro-oxidant/anti-
oxidant inversion. Thus thiophenol was initially a pro-oxidant for linseed oil
but it showed the phenomenon of auto-retardation with time [35]. This is the
characteristic behaviour of many peroxidolytic sulphur-containing anti­
oxidants in which effective ionic catalysts for peroxide decomposition are
formed after redox reactions with hydroperoxides [36].
The theory that antioxidants are essentially reducing agents was criti­
cised by Milas [28] who pointed out that the evidence accumulated by
Moureau and Dufraisse themselves did not accord with this view. Many
non-reducing antioxidants such as inorganic acids (e.g. sulphuric acid) were
6 GERALD SCOTT

known to be antioxidants, an observation which has since been explained on


the basis of their ability to destroy hydroperoxides by an ionic mechanism
[36,37]. Moreover, the fact that some antioxidants were actually oxidising
agents was not fully explained until the radical chain theory of autoxidation
had been fully developed.

3.1 Chain-Breaking Donor (CB-D) Antioxidants [38]

The crucial observation that free radical chain reactions could be in­
hibited by reducing agents was made by Bäckström in 1927 [23], He showed
that aliphatic alcohols were inhibitors for the oxidation of sodium sulphite
and benzaldehyde and that there was a correlation between the oxidisability
of alcohols and their inhibitory power. This theme was taken up by Lowry
and his co-workers [39] who showed that the antioxidant activity of phenols
in petrol as measured by the induction time to the onset of autoxidation was
related to Fieser's critical oxidation-reduction potential [40]. Although this
concept was later shown to be an oversimplification since it ignored steric
factors involved in the oxidation of phenols and in the reaction of the derived
phenoxyl radicals (see below), it did nevertheless prepare the way for the
systematic study of the kinetics of inhibited autoxidation
Bolland and ten Have [41] followed up their kinetic studies on the
oxidation of olefins to obtain a measure of the rate of hydrogen abstraction
by alkylperoxyl from a number of phenolic antioxidants.

ROO+AH > ROOD + A- (11)

Antioxidant activities as measured by the rate of this reaction were found


to be related to the critical oxidation-reduction potential of the phenols. The
pioneering studies of Bolland were followed by intensive activity by physical
chemists on the antioxidant mechanism of both hindered phenols and
arylamines [42]. These confirmed and extended Bolland's conclusions, but
placed certain boundary limits on the relationship between antioxidant
activity and hydrogen transfer activity as measured by critical redox poten­
tial. Two other reactions had to be taken into account. These were:
(a) the rate of reaction of oxygen with the reducing agent which results in
radical generation [34];

AH + 0 2 > A- + OOH in organic solution (12)

or

AH + 0 2 > A- + 0 2 + H + in aqueous solution (120


AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 7

(b) the reaction of the resulting "stable" aryloxyl with the substrate,
Reaction (11) and oxygen, Reaction (12) [43-45];

A +RH AH + R- (13)

A +Oo AOO (14)

All these reactions lead to chain initiation rather than inhibition. Reac­
tion (12) places a lower limit on the critical oxidation potential of the
hydrogen (or electron) donor. Reactions (13) and (14) emphasise the impor­
tance of the stability of the radical produced and Reactions (12) and (14)
indicate that the activity of hydrogen donor antioxidants is dependant on
oxygen pressure.
Kinetic studies of phenol and amine inhibition of autoxidation were
preceded by extensive studies by organic chemists of the products formed by
oxidation of these same species [46]. The early studies of Pummerer and his
co-workers [47,48] had shown the universality of oxidative dimerisation
reactions of phenols and Saunders and his co-workers [49-51] later carried
out similar investigations in the arylamine series.
In the late 1950s Ley and Müller and their co-workers prepared and
studied a series of "hindered" aryloxyls (V) which are the primary oxidation
products of hindered phenols (IV) [52-55].

OH
tBu tBu tBu
Oxidation

(15)

Many of these were found to be stable in the absence of oxygen but most
react rapidly and irreversibly with oxygen to give peroxidic products (e.g.
(VII) [56]).

tBu tBu
(16)

VII
00

tBu _ tBu

o 0 CHCH-
OR 6 OR
s-O-1

tBu - 0

5-Q-a
I
PQ

tBUyYtBU tBu
y+
PQ
3

m
tBuqtB -E.....

3
I ~ u ROO. tBuqtBu
----+

""
A
I

Me Me Me ~H2 ~ tBu

BHT lROO. o=9=CH o.


~
yU
tBu tBu
o o
o=<
~ t B U h tB
PQ

tBUQtBU

Me OOR 3
o

. transform at'Ion products ofBHT


Xl ative
Scheme 1. O'd
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 9

None of the phenols studied by Müller were important antioxidants and


Waters and his co-workers showed that the reason for this was that in all
cases the initial phenoxyl formed from the more effective antioxidants was
converted to either dimeric or oxidised products [57]. Scott and co-workers
[58] subsequently showed by electron spin resonance that 2,6-di-fer£-butyl-
4-methyl phenoxyl formed from the most widely used antioxidant, BHT
(butylated hydroxy toluene), had a half life of minutes in the presence of air
but that secondary radicals derived from dimerisation or oxidation products
were longer lived.
The salient features of the chemistry of BHT during autoxidation which
resulted from the investigation of the 1950s are summarised in Scheme 1
[57]. The detailed mechanisms of some of these reactions is still somewhat
obscure but it is now known that many of the oxidation products themselves
have antioxidant activity [59].
Many publications during the 1950s were concerned with an empirical
correlation of antioxidant activity and structure in the 2,4,6 trisubstituted
phenols. Most of these were addressed to the problem of oxidation during
storage and use of petrols and lubricating oils [60] but some were also
concerned with the deterioration of foodstuffs [61] and of rubbers [62]. Some
general rules emerged which were capable of interpretation in the light of
the more fundamental studies described above.
The salient conclusions were generalised as follows [60]:
1. Electron releasing groups increase the antioxidant activity of phenols
and electron attracting groups decrease it.
2. Groups which increase the délocalisation of the unpaired electron in the
aryloxy radical increase antioxidant activity.
3. Bulky alkyl groups in the ortho position increase activity whereas
branching on the para position in general decreases activity.
The first observation related to the electron donor activity (oxidation-
reduction potential) of the phenol toward the electrophilic alkylperoxyl
radical (Reaction (11)). In the transition state (Reaction (110) the oxygen of
the alkylperoxyl assumes a partial negative charge and the aryloxyl oxygen
a partial positive charge.
The transition state energy will be decreased by both release of electrons
from X and by délocalisation of the unpaired electron in the aromatic ring
[63]. Both effects have been experimentally substantiated. Boozer and Ham­
mond [45] were able to show, using the Hammett relationship, that the
activity of phenols with different substituents in the aromatic ring corre­
lated with electron releasing power, and in the case of polycyclic phenols, a
linear relationship between antioxidant efficiency and the resonance energy
of the aryloxyl radicals was demonstrated by Davies et al. [64].
The effect of ortho tert-aïkyl groups relates to the stability of the phenoxyl
radical produced by the primary hydrogen transfer but in this case not by
electronic stabilisation but by steric protection. In general the stability of
10 GERALD SCOTT

tBu tBu

X- O —H OOR X—f V-*6 —H —OOR

tBu tBu

(110

tBu

HOOR

tBu
aryloxyl radicals increases with increasing steric hindrance [58] and Bickel
and Kooyman demonstrated [65,66] that increasing steric hindrance in the
ortho position increased the probability of alkylperoxyl radical attack in the
intermediate radical (10a) over bimolecular dimerisation processes (10b)
[63].

R CH
R<oo '°° 2R
R,

(17)

CH,R R
\ . R R y^i

o=^ \ = c - c = / V=o
R, R,

3.2 Chain-Breaking Acceptor (CB-A) Antioxidants

In principle it should be possible to trap or otherwise deactivate alkyl


radicals during autoxidation. Alkyl radicals are electron donors whereas
alkylperoxyl radicals are electron acceptors [66].

R- -> R+ (18)
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 11

ROO *e > ROO" (19)

Whether the first process can occur will depend on the concentration of alkyl
relative to alkyl peroxyl in the system.
Bateman and Morris [19,23] in an important and fundamental kinetic
investigation of the termination step in inhibited autoxidation showed that
the following factors influence the [R]/[ROO] ratio and hence the relative
contribution of the three termination steps, 7-9.

2ROO > Inert pi oducts (7)

ROO + R- > ROOR (8)

2R- > R-R (9)

(a) The oxygen concentration


Figure 1 [19] shows that for three different hydrocarbons, the contribu­
tion of Reactions (8) and (9) to the termination step decreases with increas­
ing oxygen pressure.

(b) Hydrocarbon structure


It can also be seen from Fig. 1 that the pressure at which reactions of alkyl
radicals become dominant in termination differs with the structure of the
hydrocarbon.
2,6-Dimethyl heptane-2,5-diene (VIII) gives a very stable alkyl radical
(IX) on oxidation which is much less reactive toward oxygen than the
corresponding radicals from ethyl linoleate and phytene. Consequently
Reaction (8) plays a major part in the termination process of VIII even at
atmospheric pressure. The oxygen partial pressure at which the rate of
oxidation of a simple olefin such as hexadec-1-ene becomes independent of
oxygen pressure, is relatively low; VIII, on the other hand, shows oxygen
pressure dependence up to 1 atmosphere (see Fig. 2).

/
\ / ROO- \ •
C= CHCH 2 CH= C > C = CHCHCH= C
/ \ / \
CH3 CH3 CH3 CH 3 (20)

VIII CH3 ' CH3


\ /
CCH= CHCH= C
/ \
CH3 CH3
IX
12 GERALD SCOTT

100.

(a)

(b)

1 10 100
Oxygen pressure ( m m )

100i

(C)

10 100 1000
Oxygen pressure (mm)

Fig. 1. Termination characteristics of (a) ethyllinoleate, (b) phytene and (c) 2,6-dimethyl-
hepta-2,5-diene oxidations at various oxygen pressures (a and b at 45°C, c at 25°C).
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 13

200 400 600 Ö00


Oxygen pressure (mm)

Fig. 2. Effects of oxygen pressure on the oxidation rates of unsaturated hydrocarbons. A


= hexadecene-1 (45°C); B = ethyllinolenate (45°C); C = 2,6-dimethylhepta-2,5-diene
(25°C).

An important consequence of the increased probability of the cross-termi­


nation Reaction (8) in the oxidation of readily oxidisable hydrocarbons is
that the latter may act as oxidation retarders when added to less oxidisable
substrates. Thus, Russell [67] showed that triphenyl methane is a powerful
retarder for the oxidation of cumene due to Reaction (21a) which is able to
compete with the normal propagation reaction with oxygen (21b).

Ph 3 COOC(CH3) 2 Ph
Ph(CH3)2COO Termination

Ph3CH-522l^ p h 3 C . (21)

PI13COO > Propagation


PMCHahCH

A similar explanation has been recently advanced by Burton and Ingold


[68] to explain the retardation of methyl linoleate oxidation by ß-carotene.
The relatively stable radical (X) formed by addition of alkylperoxyl to the
conjugated system, like triphenyl methyl, is an effective trap for alkyl
peroxyl radicals at the relatively low oxygen pressures found in muscle
tissues. However, stable alkyl radicals are not true antioxidants since they
can also react with oxygen (21b). At best they are efficient retarders and at
higher oxygen pressures they actually enhance oxidation by direct attack of
molecular oxygen at the weak C-H bond, thus introducing new initiating
14 GERALDSCOTT

free radicals into the system. Holman showed [69] that ß-carotene and the
chemically related Vitamin A are preferentially oxidised in linoleic esters to
give hydroperoxides which initiate oxidation of the substrate [4]. This
emphasises the importance of oxygen pressure as a determinant of anti-
oxidant/pro-oxidant activity.

CH,
CTL O O R C H

CTL

(a) Oxidising agents CB-A as antioxidants


At an early stage in the development of antioxidant theory it was noted
that some oxidising agents are able to retard oxidation [28]. These are now
recognised as falling into the same class as polymerisation inhibitors. Thus
quinones effectively inhibit polymerisation [70] and Waters and co-workers
[71,72] showed that they function as alkyl radical traps. They are also
antioxidants under conditions where the partial pressure of oxygen is low
[59,73]. Chain-breaking electron acceptors are therefore in direct competi­
tion with oxygen which is itself a good electron acceptor;
OH

-> Dimers, etc.


(Termination)
R" (22)

RH
ROO ^ ^ ROOH + R
(Propagation)

The efficiency of Reaction (22a) depends on the [R]/[ROO] ratio and this
in turn depends on the oxygen pressure and on the oxidisability of the
substrate. This is why CB-A antioxidants have not, until recently, seemed
to be as important as the CB-D class. However, as will be seen later, there
are a number of practical situations both in polymer technology and in
biological systems where one or both of these conditions may be satisfied and
today this is one of the most interesting developments in the elucidation of
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 15

antioxidant mechanisms. Chain-breaking acceptor antioxidants will be dis­


cussed in more detail in later chapters in this volume but the main historical
developments will be discussed here.

(b) "Stable" free radicals as CB-A antioxidants


In the early 1960s Neiman and his co-workers reported [74,75] that stable
nitroxyl radicals are effective antioxidants in some polymeric substrates. Of
particular significance because of their high chemical stability were the
2,2,6,6-tetramethyl piperidinoxyls (XI).

R
O
CH 3
CH
CH, C^^3
I
O
XI XII

At about the same time, the diaryl nitroxyls (e.g. XII) were shown [76,77] to
be products of the oxidation of the corresponding arylamines by oxidation
both by alkylperoxyl radicals and by hydroperoxides, Reaction (23).

o*-o
H
ROO
\ / \ / (ROOH)

ROO

(23)
o
OK3 +
RO'
Nieman and Rozantzev suggested [78] that the nitroxyl radicals were able
to trap alkyl radicals in direct competition with oxygen.
In spite of these and subsequent investigations of nitroxyls by Ingold and
his co-workers in the mid 1960s [79], the practical significance of this species
was not fully appreciated at that time and in a standard text on polymer
stabilisation published in 1972 [80], it is stated that "the dialkyl nitroxide
radicals are of little practical value as stabilisers against thermal oxidation
compared to the more conventional chain-breaking antioxidants because the
former must compete with oxygen for R- radicals". It is now known that the
hindered piperidinoxyls are effective antioxidants in a wide variety of
16 GERALDSCOTT

polymers and that the mechanism involves the continual regeneration of the
nitroxyl from the derived hydroxylamines in the polymer. The reasons for
the negative initial conclusions as to their effectiveness can be again at­
tributed to the unfavourable conditions under which they were examined.

3.3 Catalytic Chain-Breaking Antioxidants

During the development of the modern synthetic polymer industry, a


number of unusual stabilisation systems had emerged whose activity could
not be explained by any of the currently accepted mechanisms of antioxidant
action. Indeed, the effectiveness of some of these appeared to be in direct
conflict with scientific thinking at that time. Thus, for example, it was
known that some of the most effective thermal antioxidants for polyamides
and polyesters contained copper salts [81,82]. Copper salts were regarded
by the rubber technologist as very active pro-oxidants ("poisons") and they
were rigorously excluded from rubber formulations [83]. Parallel scientific
studies provided evidence that all transition metals including copper were
able to catalyse the decomposition of hydroperoxides to free radicals [84],
Reactions (24) and (25),

ROOH + Cu 2+ > ROO + H + + Cu + (24)

ROOH + Cu + > R O + OH + Cu 2+ (25)

thus reducing the activation energy of hydroperoxide decomposition and


catalysing oxidation. However, studies in liquid hydrocarbons indicated
that this was an oversimplification. It was shown by Ohta and Tezuka [85]
that inp-xylene, cupric naphthenate is a pro-oxidant at low concentrations
and an antioxidant at high concentrations. A similar concentration inver­
sion was observed for copper stéarate in dodecane by Emanuel and his
co-workers [86]. This has since been shown to be a general phenomenon with
a variety of transition metal ions [87]
Two theories have been advanced for the antioxidant activity of transition
metal ions. The first is that the metal ion in its higher oxidation state
oxidises an alkyl radical to a carbonium ion;
R. + M(n+1) > R+ + M"+ (26)
There is a good deal of evidence that this can occur in the case of copper
since Kochi was able to isolate an intermediate copper alkyl which under­
went partial elimination to give an olefin [88-90];

R + CU2+ > R-Cu + > C = C + Cu+ + H+ (27)


AUTOXIDATI0N AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 17

Recent studies have shown [91] that Reaction (27) can occur during
photo-oxidation of copper pigmented paint media. The extensive conjugated
unsaturation that results is the cause of the brown discolouration which is
evident in old paintings and also occurs in polypropylene when processed at
high temperatures in low oxygen pressures [73].
The second mechanism involves the reduction of the alkylperoxyl by the
lower oxidation state of the metal followed by subsequent reduction of the
hydroperoxide to inert products.
ROO'+Mn+ > ROÖ + M (n+1)+ (28)
This reaction was suggested as an inhibition process in the case of copper
stéarate as early as 1946 by George and Robertson [92,93]. Denisov [104]
invoked oxidation by cupric ions to explain the complete inhibition of in­
itiated oxidation of cyclohexanol by concentrations of Cu + as low as lO^5
mol l"1. In this case the species oxidised the intermediate a-hydroxyperoxyl
radical, Reaction (29a), and the cuprous ion is regenerated in a parallel
reaction of the same species (Reaction (29b)).

^ C = o + 0 2 + H+ + Cu+

(29)

It was suggested by Scott in 1971 [94] that, based on this kind of evidence,
a genuinely catalytic antioxidant mechanism for hydrocarbons was in prin­
ciple possible by repeated cycling of the metal ion through Reactions (26)
and (28) and this could account for the high antioxidant activity of copper
ions in many technological systems. This concept was subsequently con­
firmed in polypropylene during processing [95] and was extended to a
number of other redox systems both inorganic [96] and organic [97-100] of
which the most important are iodine and organic iodides, stable nitroxyl
radicals and phenoxyl radicals (see Table 1). A detailed discussion of the
evidence for this antioxidant mechanism will be deferred until Chapter 4 in
this volume but the salient features of the catalytic redox mechanism which
has also been shown to operate in liquid hydrocarbons [102,103] is depicted
in Scheme 2.
18 GERALD SCOTT

ROOH ROO
Scheme 2. General catalytic (redox) mechanism for the stabilisation of polymers by
"stable" radicals.

TABLE 1

Redox systems which have been shown to have catalytic antioxidant activity in polymers
[59]

Oxidant (A-) Reductant (AH)

Cu2+ Cu +
I"
r Me
x>r

R—( N-O' R-/ N —OH


Me
TeMe J/
Me
OH

R R N-

tBu tBu tBu tBu

0=/ >=CH O- 0==/~~y=CH-/""V-OH


tBu tBu tBu tBu
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 19

An essential condition for the operation of the above scheme is that the
CB-A antioxidant must be able to compete with oxygen in reaction with the
substrate radical (R). This in turn implies that the [R]/[ROO] ratio must
be significant either due to limited availability of oxygen at the site of the
reaction or because the substrate is readily oxidisable (see Section 3.2).
It is interesting that these conditions are satisfied in many biological
systems which appear to be ideal for the operation of the catalytic chain-
breaking mechanism outlined above and indeed the evidence suggests that
regeneration of a-tocopherol from its phenoxyl radical is achieved a t the
expense of more oxidisable components of the system (e.g. ascorbic acid).
Although no direct evidence has been reported for a similar regeneration
from alkyl radical species this process must be favoured in biological sys­
tems due to the stability of the conjugated allylic radicals toward dioxygen
and by the low oxygen pressures in the lipid environment of many cells [68].

3.4 Peroxidolytic Antioxidants

It was demonstrated many years ago that mineral oils become much more
susceptible to oxidation after purification [105] and Denison and Condit
[106,107] were the first to demonstrate that this is due to small amounts of
sulphur and/or nitrogen compounds in crude mineral oil. They showed [106]
that the addition to desulphurised base oil of sulphides such as dicetyl mono
sulphide and cetyl phenyl sulphide increased the oxidation resistance of the
oil almost to that of the unpurified oil. A characteristic of the sulphur-con­
taining oils was the very low concentrations of hydroperoxides which they
contained compared to similarly oxidised purified oils. Denison and Condit
suggested the presence of a peroxide decomposer but were unable to identify
it. Monosulphides were shown to be converted to sulphoxides and sulphones
but they were not themselves antioxidants [107]. An observation which was
to assume great significance later was that water soluble sulphur acids were
also formed. Kennerly and Patterson [108] were the first to unambiguously
demonstrate the catalytic nature of the decomposition of hydroperoxides
with sulphur compounds but they were again unable to identify the nature
of the catalytic species. A major contribution was made by Oberright et al.
[109] who showed that the products formed from a-cumyl hydroperoxide
(CHP) were strongly dependant on the molar ratio of the hydroperoxide to
sulphur compound. At high ratios the products of the ionic catalytic decom­
position of CHP predominated whereas at stoichiometric ratios both ionic
and radical products were formed (see Scheme 3). This observation proved
to be of critical importance in the further elucidation of the mechanism of
sulphur-containing peroxide decomposers [110].
In the 1950s, chemists interested in polymer stabilisation began to inves­
tigate the oxidation chemistry of sulphur. It was known that sulphur, in the
form of a polysulphide cross-link profoundly modified the oxidation suscep-
CH3 CH, CH 3 CH,
+ -CH, I
CH3 — C— 0 + - 0 H C=0 CH,C—OH = CH2

CH 3 + H20
I
CH, —C—OOH HOMOLYTIC BREAKDOWN PRODUCTS
M+/M2 +

or hv

+
H+
CH3 CH3CCH3
I
CH.C—0 + O OH
H20
» I + (CH 3 ) 2 C=0 + H +

HETEROLYTIC BREAKDOWN PRODUCTS

Scheme 3. Ionic and radical mechanisms in the decomposition of a-cumyl hydroperoxide, CHP.
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 21

tibility of the vulcanised rubbers [111]. Whereas all the hydrocarbon rub­
bers oxidised in an autoaccelerating mode before vulcanisation, the chemi­
cal incorporation of sulphur as mono, di and polysulphides initially accel­
erated the rate of oxidation but subsequently led to autoretardation. This
behaviour depended profoundly on the nature of the vulcanising system and
in general the higher the ratio of accelerator to elemental sulphur in the
cross-linking system, the more oxidatively stable was the vulcanisate [112].
Bateman et al. [113] at the British Rubber Producers' Research Associa­
tion studied the oxidation of unsaturated monosulphides and disulphides
with chemical structures analogous to that of the sulphur cross-link. They
were able to show that the derived sulphoxides and thiosulphinates were
effective antioxidants in squalene, a model of the rubber molecule.
Hawkins and his co-workers carried out a parallel study of the effects of
a variety of sulphur compounds in polyethylene, particularly in the presence
of carbon black with which they were known to synergise [114,115]. These
authors also concluded the peroxide decomposition must be involved in the
antioxidant activity of this class of compounds and in an elegant study of
diphenyl disulphide, XIII, and its oxidation products, XIV, XV, Reaction (30).

^s-s^3 — Q-ï-s-Q
XIII XIV
O (30)

- Of-O o
XV
They demonstrated [116] that in the autoxidation of cumene, only XIV was
immediately effective as an antioxidant. Both XIII and XV had to undergo
prior reaction to give an effective antioxidant. As a result of these studies,
the principle of prior oxidation of an inert sulphur compound to give an
antioxidant was established in a number of quite different technologies but
there was at this time no agreement as to what the active agent was.
Kennerly et al. [108] and Hawkins and co-workers [116] believed it to be
thiyl radicals formed from the sulphur compounds. Bateman and his co-
workers believed it to be the sulphoxides or thio sulphinates themselves
[113]. They did however show [117] that the antioxidant activity of oxidised
sulphur compounds increased with their thermal instability, a reaction later
studied in some detail by Shelton and his co-workers [118]. This is il­
lustrated for di-fer^-butyl sulphoxide in Eqn (31).
22 GERALD SCOTT

o
II
(CH3)3CSC(CH3)3 > CH2 - C(CH3)2 + (CH3)3CSOH) (31)

A number of common mechanistic features linked these early studies of


the antioxidant mechanisms of sulphur compounds [36]. These were:
(1) The sulphur compounds themselves were normally inactive as anti­
oxidants and at high concentrations they were often pro-oxidants. Prior
oxidation was an essential requirement for high antioxidant activity.
(2) The kinetics of sulphur-inhibited autoxidation generally showed the
phenomenon of autoretar dation and some of the oxidation products showed
a high level of antioxidant activity.
(3) The parent sulphur compounds appeared to act as a reservoir for the
slow release of highly antioxidant-active oxidation products.
A clue to the chemistry involved in these phenomena was provided by the
work of Scott and co-workers [119] who showed that a variety of antioxidant-
active sulphur antioxidants are, under oxidative conditions, the source of
sulphur oxides which can function as catalysts for the ionic decomposition
of hydroperoxides. S 0 2 was identified as a product of the oxidation of the
metal dialkyl dithiocarbamates (XVI) which were well known antioxidants
in rubbers vulcanised by the so-called thiuram "sulphurless" system and
Hawkins and Sautter [116] similarly identified S 0 2 as a product of the
oxidation of diphenyl disulphide (XIII).
S S S S
R2NC^ ^ Z n ^ ^CNR2 (RO)2P^ ^ Z n ^ ^P(OR),
\ s / ^ ^ ^
XVI XVII

R 2 NCSSCNR 2 RSH || I yCSH

XVIII XIX XX
The sulphur compounds XVI-XX have all been shown to be readily
converted by hydroperoxides to low molecular weight sulphur acids and this
mechanism is common to the stabilisation of quite different technological
systems [36,37]. Thus XVI, XVIII and XX have for many years been impor­
tant antioxidants in vulcanised rubbers and recently XVI and related metal
complexes have become important stabilisers for plastics. The zinc dialkyl-
dithiophosphates (XVII) are long established antioxidants in lubricating
oils, but they and the analogous nickel complexes are thermal and photoan-
tioxidants in polyolefins. It is also now known that chain-breaking mecha-
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 23

nisms are also, involved in the antioxidant activity of many sulphur antiox-
idants [36,37,121-123] but there seems little reason to doubt that the
dominant process in hydrocarbon oils and polymers is the slow release of low
molecular weight sulphur acids which act as ionic catalyst for hydroperoxide
decomposition. A detailed discussion of the chemistry of these processes will
be deferred until Chapter 5.

4. SYNERGISM

The addition of additives to foodstuffs is severely limited by toxicological


considerations and this restricts the freedom of the food technologist to
design effective antioxidants for foodstuffs. There has therefore been a
tradition of exploiting the use of naturally occurring chemicals as antiox­
idants [61]. Among the more important of these are the E group of vitamins
(particularly a-tocopherol (XXI)), Vitamin C (ascorbic acid (XXII)), the gal-
late esters (XXIII) and nordihydroguaiaretic acid (XXIV). The phenols often
show co-operative effects with other natural products which may by them­
selves have little or no antioxidant activity (e.g. phospholipids, citric acid,
etc. [61].
It was no accident then that the co-operative behaviour of different
antioxidants to produce effects which were not possible with single anti­
oxidants, was first realised by Mattil and his co-workers in the stabilisation
of foodstuffs [124]. Synergism, as the phenomenon was called, has since
assumed dominating significance in other technologies and, as will be seen
in Section 1.6, is of key importance in antioxidant processes in vivo.
It was the discovery of synergism between conventional chain-breaking
antioxidants, sulphur compounds and UV stabilisers, that made the devel­
opment of polypropylene, initially as a plastic and later as a fibre-forming
polymer, a practical reality [125].
A stabilising "package" for readily oxidisable polymers such as the rub­
bers or polypropylene may contain several antioxidants acting by different
mechanisms. It is becoming decreasingly viable to develop multicomponent
stabiliser systems by empirical selection of antioxidants. An alternative
approach based on an understanding of the complementary role of the
individual additives is seen to be an economic necessity in polymer tech­
nology and explains the considerable attention that has been directed in
recent years to the study of antioxidant mechanisms [2]. The landmarks in
the development of antioxidant theory have been outlined in the previous
sections. It now remains to be seen how these provide a framework for the
development of practical stabilising systems in specific media and under
widely differing environmental conditions. This will be discussed in more
detail in Volume II.
24 GERALDSCOTT

er V
CH3
1
HO
R o
/C~^C—OH
II
CH3 CH3 W -° H
C
CH3 I
CHOH XXII
CH, CH 3 I
| 3 | 3 CH 2 OH
R = —(CH 2 ) 3 CH(CH 2 ) 5 CH(CH 2 ) 5 CH(CH 3 ) 2
XXI

COOH
XXIII

O H
HO-<f y-CH2CHÇHCH2-/ V

XXIV

5. THE EFFECT OF THE REACTION ENVIRONMENT UPON OXIDATION AND


ANTIOXIDANT MECHANISMS

So far in this chapter, the emphasis has been on the deterioration of


organic substrates under conditions of uncatalysed "thermal" oxidation. The
initiating step in this case is the thermolysis of indigenous peroxides or by
the adventitious formation of radicals by direct reaction of the substrate and
oxygen. This last reaction is generally believed to be the attack of ground
state dioxygen on the substrate (Reaction (32)).
RH + 0 2 > R- + OOH (32)
There is, however, relatively little evidence to prove this because very
minor impurities particularly in technological media may have a profound
effect on the oxidation process so that Reaction (32) cannot normally be
observed (see, however, Chapter 3, Section 2.1).
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 25

5.1 Metal Ion Contaminants

Adventitious contamination by metal ions is very difficult to avoid in most


man-made products. The accelerating effects of transition metal ions on
autoxidation was recognised in the early 1900s by Titoff [126] who also
proposed that an important role of antioxidants was to deactivate these
catalysts which were later shown to generate radicals by the redox reaction
discussed in an earlier section (e.g. Reactions (24) and (25)). Redox reactions
of this type and their inhibition are also of extreme importance in biological
systems and in foodstuffs due to the tendency of transition metal ions,
particularly iron, to escape from damaged haemoglobin or from its trans­
porting protein (see Section 1.6.2). Activation by transition metal ions is also
important in lubricating oil technology due to the corrosion of metal sur­
faces, in polymer technology where there is an increasing tendency to leave
trace catalyst residues in the polymer after manufacture for economic
reasons [127].
Early attempts to reduce the effect of pro-oxidant metal ions were largely
centred around efficient complexation of the metal ion to its maximum
co-ordination number so that hydroperoxides could not enter the metal
orbital and thus undergo electron transfer to give free radicals [128]. Al­
though this kind of process often works well in aqueous media and is the
basis of sequestering agents for metals, it has been found that in organic
media, the most effective metal deactivators are those which have at least
one additional antioxidant function. Thus, the metal dithiophosphates
XVIII used in lubricating oils are both metal deactivators and peroxidolytic
antioxidants and the most effective "copper deactivators" for polyolefins
contain a chain-breaking antioxidant function as well as a metal complexing
group [129]. This will be discussed in more detail in Volume II.

5.2Environmental Contaminants

It was recognised during the second world war that rubber tyres fitted to
vehicles in combat readiness were very susceptible to a component of the
atmosphere. This proved, after extensive investigation, to be ozone formed
by photochemical reaction of industrial pollutants, notably nitric oxide with
hydrocarbon products of the petrochemical industry in the upper atmos­
phere [130,131]. Ozone cracking only occurs in rubber subjected to stress
and the reason for this was apparent from the extensive mechanistic studies
of Criegee [132] and Bailey [133] who followed up the earlier work of
Staudinger [134]. These authors showed that the initially formed molo-
zonide (XXV) undergoes rearrangement through a zwitterion (XXVI) with
transient chain scission [135] (Reaction (33)).
The effect of stress therefore is to inhibit the recombination of the zwitte­
rion with the carbonyl compound so that paramount scission of the polymer
26 GERALD SCOTT

chains occurs in the surface of the rubber. Reaction (33) does not involve
autoxidation in the sense that it has been discussed so far in this chapter,
but it does give rise to peroxidic species, XXVII, XXVIII, which are initiators
for autoxidation. Antiozonants (e.g. XXIX) which were developed empiri­
cally to combat ozone cracking are also powerful chain-breaking donor
antioxidants.
O
/ \
O O
(a) (b)
RCH=CHR' RCH — CHR' RC + HOO"+R'CHO
XXV XXVI
(d)

R R (33)
I
-CHOOCHOO — RCH
A CHR'
XXVIII \ /
O-O XXVII

04^A NHR XXIX

Ozone does not attack saturated compounds as readily as it does olefins


but there is evidence [136] that it can initiate autoxidation by a direct
hydrogen abstraction mechanism (Reaction (34)) and its role as an initiator
for the normal free radical chain reaction cannot be ignored. It is probably
its role as an initiator of autoxidation of the lipids that leads to its toxic
effects in biological systems.

RH + O3 RO + OOH (34)

Another form of oxygen which has been implicated in autoxidation follow­


ing its discovery in the 1960s is excited state singlet oxygen [137]. The lower
energy QiDg) state is thought to be possibly involved in attack on polymers
since it can be readily formed from ground state O2 by energy transfer from
photo-excited triplet state species (particularly carbonyl groups)

% +3 0 2 % + ^ 2 (Ag) (35)

As will be seen later it is now believed that the attack of 0 2 is only


AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 27

important in unsaturated polymers (Reaction (36)) since it is physically


quenched extremely rapidly by many organic and inorganic compounds
(including water) and only the most reactive olefins can compete with this
process [138],

OOH
-CH 2 CH=CH ^-> -CH=CH-CH- (36)

Hydroperoxides produced in Reaction (36) are very powerful sensitisers


for autoxidation (see Section 1.2). Photo-excited forms of S 0 2 and oxides of
nitrogen are also capable of initiating photooxidation in polymers but these
processes are normally only of importance in heavily polluted atmospheres
[139].

5.3 Mechanooxidation

The accelerating effects of physical stress in the oxidation degradation of


polymers has been recognised phenomenologically since the discovery of
"mastication" by Hancock [140] over 150 years ago. The ability to convert
raw rubber from its resilient high molecular weight solid state to a plastic
product into which vulcanising ingredients and antioxidants could be
readily incorporated marks the point at which natural rubber became a
technological product and Hancock's discovery of chemical plasticisation of
raw rubber must rank with Goodyear's discovery of vulcanisation as a
landmark in the development of the polymer industry.
Although the practical importance of mastication was recognised in the
first half of the nineteenth century, its mechanism was not understood until
almost a century later when Busse and Cunningham [141] showed that the
mastication of rubber below 100°C in an atmosphere of air became less
efficient, not more efficient, as the temperature increased; that is, it did not
obey the normal Arrhenius relationship. This led Kauzmann and Eyring
[142] and later Watson and his co-workers [143], to conclude that the
determining factor must be mechanical scission of the polymer chain to give
macroalkyl radicals followed by their normal reaction with dioxygen (Reac­
tion (37a)). Watson et al. [114] went on to show that in the absence of oxygen
macroalkyls behaved very much like low molar mass alkyl radicals. Thus,
for example, they initiated vinyl polymerisation (Reaction (37b)) and were
trapped by quinones (Reaction (37c)).
Mechanooxidation has no analogy in normal solution chemistry. It cannot
occur unless localised stresses are present and are severe enough to break
carbon-carbon bonds. It is not, however, limited to polymers in the molten
state at high temperatures. Rubbers particularly when cross-linked are
susceptible to the phenomenon of "fatigue" when subjected to cyclical stress
28 GERALD SCOTT

CH 3 CH33
I I
C=CR y
C=CH
— CH,/ Ori^Crl') \
CH,
Shear I
C=CR ^C=CH^
— CH9/ CH7 + 'CH,
(a),
02/RH (c)
CH 3
I (37)
.C=CH^
CH, CH.OOH
(b) nCH 2 =CHR CH,

OH

CH 3 .
;C=CH
— CH9 \ CH 2 +CH 2 CH+ n

R
at ambient temperatures [145]. This was recognised by Russian workers,
notably Kuzminsky [146,147] and Slonimsky [148] to be accelerated oxida­
tion due to free radicals produced by mechanical scission of the cross-linked
network. Most antioxidants which protect rubber effectively against ther­
mal oxidation are almost ineffective in inhibiting stress cracking of rubber
which is the outward manifestation of fatigue. However, one class, the
4-alkylamino diphenylamines (XXIX), has been found by empirical selection
to be highly effective. The reason for their effectiveness was not known until
relatively recently when it was recognised [99] that diphenylamines of this
type are rapidly converted to the corresponding diphenyl nitroxyls (see
Reaction (23)) at an early stage during their mechano-antioxidant action.
The mechanism of their action is complex and will be discussed in more
detail in a later chapter but it is the best known example of the operation of
the catalytic CB-A/CB-D mechanism (see Section 1.3.5.) in rubber tech­
nology. Essentially the same antioxidant mechanism has also been demon­
strated to occur at high temperatures during the melt processing of
polypropylene and some of the most effective melt stabilisers for polyolefins
are stable aryloxyl and nitroxyl radicals of the type described in Section
1.3.5.
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 29

5.4 Photo-oxidation

The catalytic effect of light upon chemical reactions has been known for
many years and is frequently used in preparative organic chemistry to
catalyse free radical chain reactions. The recognition of the importance of
traces of peroxides in light catalysed radical reactions by Kharasch and
Mayo in 1933 [149] provided an explanation for the well-known free radical
addition of hydrogen bromide to olefins. Bateman and Gee [150] and, almost
simultaneously, Bamford and Dewar [151], demonstrated that photo-acti­
vated oxidation of olefins involved the same chain propagation steps as
autoxidation in the dark. However, Bateman and Gee also showed [150] that
if the olefinic substrate is pure then an induction period precedes the onset
of rapid autoxidation and this is associated with the development of hy­
droperoxides in the substrate. Hydroperoxides do not absorb light above 330
nm but Norrish and his co-workers [152,153] showed that photolysis of
hydroperoxides occurred readily at 313 nm. The photolysis products of
hydroperoxides were found to absorb UV light much more strongly than
hydroperoxides and this fact obscured the importance of hydroperoxides as
primary chromophores in subsequent studies [137].
During the 1960s and early 1970s, there was a very strong emphasis on
the importance of photo-excited species as sensitisers for photo-oxidation
[154]. These included carbonyl compounds, the primary photolysis products
of hydroperoxides and many theories were suggested to explain how car­
bonyl compounds might be involved in the initiation process. One of the most
ingenious was due to Winslow and Trozzolo [155] who proposed that singlet
oxygen might also be involved. They suggested that this species could be
O o
— CH 9 CCH ? CH 2 CH,CH;
(b)/zv II
■CH9CCH3 + C H 2 = C H C H 2

(o,
Norrish II
o
OOH I
I CH2CCH3 (38)
"CH^CHCH^CHoCHoCH-

o (d)

— CH 2 C—CH 3 + ' 0 2

Radicals HOOCH,CH=CH—
30 GERALDSCOTT

formed by quenching of triplet carbonyl by ground state oxygen (Reaction


(38c)) followed by reaction of 1 0 2 with vinyl groups formed in the polymer
by the Norrish II process (Reaction (38a)).
This suggestion had a profound influence on other workers in the field,
particularly photochemists who seemed in general to be unaware of the
earlier studies in rubber which had demonstrated that carbonyl compounds
are in any case generated by photolysis or thermolysis of hydroperoxides. In
an extensive review of photo-oxidation of polymers up to 1975, Ranby and
Rabec [137] devote only a few lines to the role of indigenous hydroperoxides
as initiators, whereas many pages are taken up with excited state photo-
sensitisers which are now recognised to be less important in polyolefins than
hydroperoxides [156].
One of the attractions of the photo-sensitisation theory of photo-oxidation
was the possible explanation it gave for the photo-stabilising effect of some
transition metal complexes (e.g. XXX and XXXI) in terms of energy transfer
from triplet carbonyl or singlet oxygen [154,157]. Subsequently it was
shown that the theory of "quenching" did not support such a conclusion [158]
and more detailed studies of the light stable antioxidants acting by the
peroxidolytic or chain-breaking mechanisms [138,159].

HOCH3

/~S-0N V \ R2N<S;Ni^>NR2
C H33 M
^ V/V
\ NL - V CH 3

/ \ \ = /
H3 OH

XXX XXXI

6. BIOLOGICAL OXIDATION

Triglycérides of unsaturated fatty acids are important components of cell


membranes. The fully saturated triglycérides tend to be higher melting than
the monoenic and the dienic and trienic esters are lower melting still. This
presumably explains nature's preference for the polyunsaturated esters in
higher mammals. In terms of oxidative stability, however, the reverse
arrangement would seem to be preferable since the introduction of double
bonds into a saturated chain increases the rate of oxidation many times.
Oleic esters (XXXIII) are much more oxidisable than saturated esters such
as stearic (XXXII) and it was shown by Holman and Elmer over forty years
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 31

ago [160] that linoleic esters oxidise about 40 times more rapidly than oleic
esters.

CH3CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2CH2COOR
XXXII

OH0OH2OH2OH2OH2OH2OH2O -"2^ H = 0 HO H2O -^2^ " 2 ^ "2^ y ^ 2 ^ " 2 ^ H2OOOR


XXXIII

CH3CH2CH2CH2CH2CH=CHCH2CH=CHCH2CH2CH2CH2CH2CH2CH2COOR
XXXIV

The phenomenon of rapid oxidative cross-linking of the triglyceryl linoleates


is the basis of drying oil technology and involves the facile removal of the
C n hydrogen atom to give a resonance stabilised radical and, after reaction
with oxygen, 1,3-dienic peroxides are formed [161] (Reaction (39)). The
latter undergo facile thermolysis (particularly with the presence of transi­
tion metal ions) to give both lower molecular weight scission products and
higher molecular weight cross-linked products. Oxidation is less favoured in
the case of the oleate and stéarate esters since the radicals produced are
much less stabilised. Moreover, the end products in the case of oleate and
stéarate esters are aldehydes and further oxidation products of lower mole­
cular weight; normally macromolecular products are not obtained. To the
organic chemist, it is very surprising that cells containing a high proportion
of polyunsaturated lipids can survive for long periods without oxidative
change in biological systems. This is particularly true of the erythrocytcs
which by virtue of their function have to retain their physical integrity in
the presence of a large excess of oxygen and in the close vicinity of haemoglo-

- C H = C H CH C H = C H -
13 12 11 10 9 Polymers

i
- C H C H = CHCH= C H - (
V RH
>
OOH
- C H C H = CHCH= C H -
13 12 11 10 9 (39)

I P°H
- C H = CHCH= CHCH- ^ a ^ _CH= CHCH= CHCH-

Polymers
32 GERALDSCOTT

bin with its potential for catalysing oxidation [162,163]. Over the past
twenty years, a considerable body of information has built up from biochemi­
cal investigations to suggest that this is achieved by means of a complex
synergistic system of antioxidant defences against oxidation by dioxygen.
Oxidation of the lipids is now also recognised to be closely linked with ageing
and age-related diseases [3,164,165]. Not only does free radical induced
damage in animals increase with age but the antioxidant levels also
decrease [164].
Harman [166] was the first to recognise the close connection between
ageing and free radicals in cellular metabolism. Much of the evidence is
indirect and in view of the complexity of biological processes, this is to be
expected. Harman also recognised [167] that the polyunsaturated fats were
potentially more damaging in the diet than the more saturated fats but a
less expected conclusion was that although this did lead to reduction in
lifespan in mice the reason for this was an increased incidence of cancer. A
large amount of subsequent research has confirmed the relationship be­
tween free radical generators in vivo and age-related diseases such as
cancer, Parkinson's disease, Alzheimer's disease and autoimmune disorders
[3,165]. This subject will be reviewed in some detail in Volume III of this
series by authors who have been involved in the recent advancement of the
subject. One of the most striking developments has been the recognition of
the role of biological antioxidants in controlling age-related diseases.
Moreover, the modes of action of biological antioxidants can be clarified by
the same mechanistic scheme which developed from a study of oxidation in
technological systems although the latter preceded the former by more than
a decade [168]. The major types of biological antioxidants and our present
understanding of the way in which they control degenerative diseases will
be briefly reviewed.

6.1 The E Group of Vitamins (The Inhibitols)

The antioxidant function of Vitamin E was first recognised by Olcott and


Mattill and their co-workers in the mid-1930s [124,169], although the
essential biological role of this group of phenols was known in the early
1920s [170]. Interestingly it was not the biological antioxidant activity of the
"inhibitols" that interested Olcott and Mattill but the fact that they syner-
gised effectively with other biological antioxidants in the preservation of fats
and oils (see Section 1.4). Olcott and Mattill found that fats of vegetable
origin were effectively stabilised by ascorbic acid whereas animal fats
responded much less readily [169]. This was shown to be due to the fact that
vegetable oils contained considerable concentrations of the inhibitols (toco-
pherols) which are widely distributed in plants. Animals do not have the
ability to synthesise the tocopherols and they consequently have to be
constantly replaced in the diet. Ascorbic acid, on the other hand, is widely
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 33

distributed in animal fats (e.g. dairy products) but is not present in veget­
able oils and fats. The addition of 5-10% of vegetable oils to animal fats
therefore markedly improves the keeping qualities of the latter [171].
The tocopherols, as the inhibitols were subsequently renamed, were thus
one of the earliest and most potent members of the chain-breaking group of
antioxidants whose mechanism was recognised to be due to its ability to
remove free radicals from the substrate whilst being itself irreversibly
destroyed (Reaction (11)).
The synergism between a-tocopherol and ascorbic acid and its esters was
suggested by Golumbic [172] to involve a regenerative cycle in which the
labile hydrogen in ascorbic acid reduces the aryloxyl radical formed in
Reaction (11) (see Scheme 4).

Asc-H Asc '

X
X
Toe Toc-H

ROOH ROO•
Scheme 4. Regenerative synergism between a-tocopherol and ascorbic acid. Toc-H =
a-tocopherol (XXI); Asc-H = ascorbic acid (XXII).

This regenerative process which was subsequently confirmed by electron


spin resonance [173] therefore requires relatively little tocopherol to convert
a large number of alkylperoxyl radicals to hydroperoxide at the expense of
ascorbic acid which is probably eventually converted to the corresponding
triketone by further oxidation. This synergistic action is a specific example
of the general phenomena of homosynergism since it involves two anti­
oxidants acting by the same mechanism [174] in an electron cascade.
It was noted by Golumbic and Mattill [175] in their pioneering studies on
the antioxidant activity of ascorbic acid that under certain conditions this
reducing agent is not an antioxidant and more recent work has shown that
it can also be a pro-oxidant [176]. It thus strongly resembles the unhindered
phenols with low oxidation-reduction potential which are poor antioxidants
in their own right but which synergise effectively with hindered phenols
[174].
The idea that the vitamins might be involved in the natural defence
mechanisms of living organisms did not gain popular currency until some
time after the work of Mattill and other technologists interested in the
preservation of natural products in vitro. Dann [177] was among the first
biochemists to champion the role of the tocopherols as widespread anti-
34 GERALD SCOTT

oxidants in the body and his views sparked a controversy concerning the role
of Vitamin E which continues among biochemists to the present day. In
1965, Horwitt [178] stated that "Nearly all biochemists would like to see
a-tocopherol directly implicated in specific enzyme reaction that cannot be
related to its antioxidant activity: to date, no such enzymic activity has been
unequivocally demonstrated. In all cases where such specific enzymic activ­
ity has been claimed, some other chemically unrelated antioxidant could
serve as a substitute for a-tocopherol". Since that time the antioxidant
function of a-tocopherol has been widened and deepened and there are now
probably few biochemists who would not admit the fundamental importance
of Vitamin E as a biological antioxidant in the lipids in vivo [179].
Of equal importance, however, has been the recognition of the same kind
of synergism between a-tocopherol and other important constituents of the
human diet, namely ascorbic acid, selenium and glutathione. Tappel [180]
was able to link the antioxidant chemistry of these compounds with the
synergism exhibited by these antioxidants in vitro. In particular he drew
attention to the analogy between the function of selenocystine diselenoxide
in vivo and organic disulphides in vitro. Tappel [181] also related Golumbic
and Mattill's regenerative mechanism for synergism between a-tocopherol
and ascorbic acid, to their behaviour in living organisms. The mechanism of
this process has recently been studied quantitatively by Ingold et al. [182]
in phospholipid bilayers, confirming the mechanism originally proposed by
Mattill and his co-workers in unsaturated fatty esters almost fifty years ago.

6.2 Preventive Antioxidants

The key role of peroxides in the oxidation of fats and oils in vitro was
recognised many years ago [5,183,184] but the significance of this process in
vivo was not appreciated until relatively recently. The reason for this
anomaly appears to result from the difficulty of detecting hydroperoxides in
biological systems.
The recognition by Fridovich [185] of the importance of the Superoxide
anion radical (Op in biological systems led the way to the investigation of
the damaging role of peroxides in cell metabolism. Hydrogen peroxide is now
believed to be the main source of initiating free radicals and not Superoxide
itself since this is a relatively stable and unreactive radical [186]. Thus,
Superoxide dismutase can only be considered to be an antioxidant in combi­
nation with a catalase or peroxidase which destroys hydrogen peroxide.

+
e . _ (H)
0-0 > 0-0 > H202 + 0 2 (40)
(a) (b)
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 35

The main source of free radicals in the presence of traces of transition


metal ions is thought to be through the well known Fenton reaction [187].

H 2 0 2 + Fe 2+ > HO + OFT + Fe 3 * (41)

H 2 0 2 + Fe 3 * > HOO'+H + + Fe 2 + (42)

The evidence for hydroxyl radical formation via redox reactions with Fe 2+
and other reducing agents has been extensively reviewed by Halliwell and
Gutteridge [186] and will be discussed in more detail in Volume III of this
series. However, it is important to note that the evidence for the role of
Superoxide as a source of free radicals came from the discovery of the
antioxidant activity of Superoxide dismutase, which catalyses Reaction
(40b) [188] and of catalase which has the ability to destroy hydrogen
peroxide without producing free radicals [189].
Indirect evidence for the importance of antioxidants which destroy hy-
droperoxides directly has also come from investigations of the protective role
of glutathione and of selenium in living organisms [178,180,181]. It is
important to note that the behaviour of glutathione peroxidase, which
contains both glutathione and selenium, has a formal similarity — at least
in so far as their effect in autoxidation is concerned — to the in vitro
peroxidases discussed in Section 1.3.6.
In the first edition of this book, antioxidants whose function it is to divert
potential radical generators into chemistry less likely to initiate autoxida­
tion, were categorised as "Preventive" [190]. Superoxide dismutase, catalase
and glutathione peroxidase all fall within this generally accepted classifica­
tion. The metal deactivators are also preventive antioxidants and it is
therefore not surprising that they also have a role in the arsenal of biological
antioxidants. This is particularly true in diseases caused by iron overload
[191] which leads to increased lipid peroxidation in the spleen [192]. Iron
overload is probably one cause of arthritis since iron is known to be present
in increased concentration in the sinovial fluid [193]. Excess iron is stored
in the body in the form of ferritin and haemosiderin and it is transported as
transferrin and lactoferrin [194]. In these complexes the iron appears to be
effectively deactivated so that it does not normally play much part in the
Fenton reaction. Removal of iron as haemoglobin is the simplest and most
efficient way of removing excess iron from the body. Synthetic complexing
agents are not normally effective deactivators since, as in vitro [195], they
often activate iron in redox reactions with hydroperoxides [186]. However,
the synthetic complexing agent desferrioxamine is used to treat iron over­
load in conditions where venesection is not possible and there are sugges­
tions that, as in the case of metal complexing in vitro, this metal deactivator
may also have other antioxidant functions [186].
36 GERALD SCOTT

7. THE PRESENT POSITION

In this chapter we have identified some of the important landmarks in the


theory of autoxidation from which has developed a comprehensive theory of
antioxidant action. The key intermediates in oxidation by molecular oxygen
are alkyl and alkylperoxyl radicals and the first isolable products of oxida­
tion, the hydroperoxides (see Scheme 5). From this follows the two main
mechanisms of antioxidant action: the chain-breaking and the preventive
[1961.
CB—A

CB—D
ROO
RO-+OH

CB—D

PEROXIDE -ROOH RH
DECOMPOSITION (PD)
A, h*/,M + /M 2 +
METAL
DEACTIVATION (MD)
UV ABSORPTION (UVA)
Scheme 5. Mechanism of autoxidation and the role of antioxidants.

Initiation is normally by reaction of hydroperoxides to give free radicals.


This process is accelerated by UV light and chemical inducing agents of
which metal ions and reducing agents are the most important and occur
widely in both technological and biological systems.
Chain-breaking hydrogen (or electron) donor (CB-D) antioxidants are
known to operate in all autoxidising systems, provided they are stable under
the conditions prevailing. Phenols acting by this mechanism are common to
foodstuffs, lubricating oils, polymers and biological systems. The chain-
breaking hydrogen or electron acceptor (CB-A) mechanism has been shown
to operate in oxygen deficient systems in polymers where alkyl radicals in
the substrate are the reducing species. The same mechanism may well be
involved in specific situations in vivo [168]. However, biological systems
contain a variety of other reducing agents which are capable of regenerating
cc-tocopherol from the primary aryloxyl radical so that the electron cascade
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 37

outlined typically for a-tocopherol and ascorbic acid in Scheme 4 will be in


competition with the regeneration of AH by reduction by substrate radicals
(see Scheme 2).
However, it has to be recognised that the chain-breaking mechanism of
antioxidant action deals with a situation once it has been initiated by the
introduction of radicals into a substrate. Moreover, it produces hydro-
peroxides which are the further source of initiating radicals. Considerable
attention has therefore been directed in recent years to the preventive
antioxidant mechanism since this is potentially a more powerful antioxidant
process. In practice, synergistic combinations of chain-breaking and preven­
tive mechanisms have been developed in technological media and are found
to be widely distributed in biological systems. The most powerful of these
are the peroxidolytic antioxidants which are either stoichiometric reducing
agents or catalysts for peroxide decomposition. Many sulphur compounds
are effective catalysts for peroxide decomposition in technological systems
and appear to act by an ionic process in which oxidation and reduction does
not occur [36,37]. The biological antioxidant, glutathione peroxidase, by
contrast, appears to function by a coupled single electron transfer process in
which selenium plays a crucial role. This process leads to the reduction of
hydroperoxides to alcohols and water [197] (see Scheme 6).

ROOH ROH

2 (E)G—S — SeH [(E)G — S — Se] 2 + H 2 0

2e + H +
(E) is the enzymic residue
Scheme 6. Catalytic electron transfer mechanism of glutathione peroxidase.

Reductive peroxidolysis is a known antioxidant mechanism in vitro. For


example, phosphite esters are widely used as peroxide decomposers in
polymers (Reaction (43)). However, their mechanism is frequently much
more complex than Reaction (43) indicates [198].

ROOH + (RO)3P > ROH + (RO) 3 P= O (43)


38 GERALD SCOTT

The destruction of hydrogen peroxide by catalase, although it involves no


overall change in the oxidation state of the metal, appears to involve single
electron transfer rather than ionic reactions [189]. It seems then, that there
is considerable scope for cross-fertilisation between in vitro and in vivo
peroxidolytic antioxidant mechanisms. The remaining preventive antiox-
idant mechanisms are closely related in technological and life processes.
Suitably designed metal complexing agents inhibit autoxidation by transi­
tion metal ions and UV and high energy initiated oxidation have a common
basis in the accepted free radical chain reaction. The latter differs from the
former in that it is not possible to protect the substrate from the initial
injection of free radicals by high energy irradiation [199] whereas photo-
antioxidants are capable of effectively retarding photooxidation [200]. How­
ever, the main inhibiting processes involved, irrespective of the process of
initiation, are the chain-breaking and peroxidolytic mechanisms which in
general show the phenomenon of heterosynergism.

REFERENCES

1 A.W. Hoffman, J. Chem. Soc, 13 (1861) 87.


2 N. Grassie and G. Scott, Polymer Degradation and Stabilisation, Cambridge Uni­
versity Press, 1985, p. 107 et seq.
3 D. Harman, in W. Pryor (Ed.), Free Radicals in Biology, Vol. V, Academic Press,
1982, Chap. 8, p. 255.
4 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 357 et seq.
5 A. Genthe, Z. Angew., Chem., 19 (1906) 2087.
6 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 4 et seq.
7 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, Chap. 2.
8 H. Staudinger, Ber. 58 (1925) 1075.
9 H.N. Stephens, J. Am. Chem. Soc, 50 (1928) 568.
10 R. Criegee, H. Pilz and H. Flygare, Ber., 72 (1939) 1799.
11 E.H. Farmer and A. Sundralingam, J. Chem. Soc, (1942) 121.
12 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 10 et seq.
13 H.L.J. Backstrom, J. Am. Chem. Soc, 49 (1927) 1460.
14 J.L. Bolland and G. Gee, Trans. Faraday Soc, 42 (1946) 236, 244.
15 J.L. Bolland, Trans. Faraday Soc, 44 (1948) 669.
16 J.L. Bolland, Q. Rep., 3 (1949) 1.
17 J.L. Bolland, Trans. Faraday Soc, 46 (1950) 358.
18 L. Bateman and G. Gee, Proc R. Soc, A195 (1948-9) 376, 391.
19 L. Bateman, Q. Rep., 8 (1954) 147.
20 L. Bateman, J. Chem. Phys., 22 (1954) 2090.
21 L. Bateman, G. Gee, A.L. Morris and W.F. Watson, Discuss. Faraday Soc, 10 (1951)
250.
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 39

22 L. Bateman, H. Hughes and A.L. Morris, Discuss. Faraday Soc, 14 (1953) 190.
23 L. Bateman and A.L. Morris, Trans. Faraday Soc, 48 (1952) 1149; 49 (1953) 1026.
24 G.A. Russell, Chem. Ind., (1956) 1483.
25 G.A. Russell, J. Am. Chem. Soc, 79 (1957) 3871.
26 G.A. Russell, J. Chem. Ed., 36 (1959) 111.
27 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 24 et seq.
28 N.A. Milos, Chem. Rev., 10 (1932) 295.
29 J. Murphy, U.S. Pat., 99, 935 (1870).
30 L.R. Moore, U.S. Pat., 680, 387 (1901).
31 N.A. Shepard, Ind. Eng. Chem., 25 (1933) 35.
32 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 428 et seq.
33 P.M. Lewis, in G. Scott (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science, London, 1984, Chap. 3.
34 J.R. Shelton and W.L. Cox, Ind. Eng. Chem., 43 (1951) 456; 45 (1953) 392; 46 (1954)
816.
35 C. Moureau and C. Dufraisse, Chem. Rev., 3 (1926-7) 258.
36 G. Scott, in G. Scott (Ed.), Developments in Polymer Stabilisation-6, Applied
Science Publishers, London, 1983, p. 29.
37 S. Al-Malaika, K.B. Chakraborty and G. Scott, in G. Scott (Ed.), Developments in
Polymer Stabilisation-6, Applied Science Publishers, London, 1983, p. 73.
38 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, Chap. 4.
39 CD. Lowry, G. Egloff, J.C. Morrell and CG. Dryer, Ind. Eng. Chem., 25 (1933) 804.
40 L.F. Fieser, J. Am. Chem. Soc, 52 (1930) 5204.
41 J.L. Bolland and P. ten Have, Discuss. Faraday Soc, 2 (1947) 252.
42 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 116-125.
43 A.F. Bickel and E.C Kooyman, J. Chem. Soc, (1956) 2215.
44 A.F. Bickel and E.C. Kooyman, J. Chem. Soc, (1957) 2217.
45 CE. Boozer, G.S. Hammond, CE. Hamilton and J.N. Sen, J. Am. Chem. Soc, 77
(1955) 3238.
46 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 125 et seq.
47 R. Pummerer, D. Melamed and H. Puttfarken, Ber., 55 (1922) 3116.
48 R. Pummerer and H.M. Ulrich, Ber., 58 (1925) 1808.
49 B.C. Saunders and P.J.G. Mann, J. Chem. Soc, (1940) 769.
50 B.C. Saunders and G.H.R. Watson, Biochem. J., 46 (1950) 629.
51 D.G.H. Daniels and B.C. Saunders, J. Chem. Soc, (1951) 2112.
52 K. Ley, E. Müller, R. Mayer and K. Scheffler, Ber., 91 (1958) 2670.
53 E. Müller, K. Ley and W. Schmidhuber, Ber., 89 (1956) 1738.
54 E. Müller, A. Ruber, K. Ley, R. Mayor and K. Scheffler, Ber., 92 (1959) 2278.
55 E. Muller, K. Schurr and K. Scheffler, Ann., 627 (1959) 132.
56 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 132 et seq.
57 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 138 et seq.
40 GERALDSCOTT

58 J.K. Becconsall, S. Clough and G. Scott, Trans. Faraday Soc., 56 (1960) 459.
59 G. Scott, in G. Scott (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science, London, 1984, p. 65.
60 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 120 et seq.
61 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 361 et seq.
62 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 445 et seq.
63 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 145 et seq.
64 D.S. Davies, H.L. Goldsmith, A.K. Gupta and G.R. Lester, J. Chem. Soc., (1956)
4926.
65 A.K. Bickel and E.C. Kooyman, J. Chem. Soc., (1953) 3211.
66 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 161 et seq.
67 G.A. Russell, J. Am. Chem. Soc, 78 (1956) 1047.
68 G.W. Burton and K.U. Ingold, Science, (11th May 1984) 569.
69 R.T. Holman, Arch. Biochem., 21 (1949) 51; 26 (1950) 85.
70 C. Walling, Free Radicals in Solution, Wiley, 1957, p. 166.
71 A.F. Bickel and W.A. Waters, J. Chem. Soc., (1950) 1764.
72 R.A. Jackson and W.A. Waters, J. Chem. Soc., (1960) 1653.
73 T.J. Henman, in G. Scott (Ed.), Developments in Polymer Stabilisation-1, Applied
Science Publishers, London, 1979, p. 39.
74 F.G. Rozantzev, L.A. Kalashrukova and M.B. Neeman, Izvest. Akad. Nauk SSSR,
Otdel. Khim. Nauk, (1962) 501.
75 M.B. Neiman (Ed.), Aging and Stabilization of Polymers, Consultants Bur., New
York, 1965, p. 33 et seq.
76 O.L. Harle and J.R. Thomas, J. Am. Chem Soc., 79 (1957) 2973; J.R. Thomas, J. Am.
Chem. Soc, 82 (1960) 5955.
77 R. Hoskins, J. Chem. Phys., 25 (1956) 788.
78 M.B. Neiman and E.G. Rozantzev, Report to Acad. Sei., USSR, Feb. 7,1964, p. 1095.
79 K.U. Ingold, in R.F. Gould (Ed.), Oxidation of Organic Compounds, Adv. in Chem.
Series, 75-1, ACS, Washington, D.C., 1968, p. 296.
80 J.R. Shelton, in W.L. Hawkins (Ed.), Polymer Stabilisation, Wiley Interscience, New
York, 1965, p. 330 et seq.
81 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 330 et seq.
82 M.B. Neiman (Ed.), Aging and Stabilization of Polymers, Consultants Bur., New
York, 1965, p. 249 et seq.
83 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 402 et seq.
84 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 43.
85 N. Ohta and T. Tezuka, Chem. Abs., 50 (1957) 278.
86 D.G. Knorre, L.G. Chuchrukena and N.M. Emanuel, Zh. Fiz. Khim., 33 (1959) 877.
87 H.S. Laver, in G. Scott (Ed.), Developments in Polymer Stabilisation-1, Applied
Science Publishers, London, 1979, p. 172 et seq.
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 41

88 J.K. Kochi, Science, 155 (1967) 415.


89 J.K. Kochi and R.V. Subramanian, J. Am. Chem. Soc, 87 (1965) 4855.
90 J.K. Kochi and H.E. Mains, J. Org. Chem., 30 (1965) 1862.
91 F. Rasti and G. Scott, Eur. Polym. J., 16 (1980) 1153.
92 P. George and A. Robertson, J. Inst. Pet., 32 (1946) 400.
93 P. George and A. Robertson, Trans. Faraday Soc, 42 (1946) 217.
94 G. Scott, Br. Polym. J., 3 (1971) 24.
95 R. Bagheri, K.B. Chakraborty and G. Scott, Polym. Deg. and Stab., 11 (1985) 1.
96 R. Bagheri, K.B. Chakraborty and G. Scott, Polym. Deg. and Stab., 9 (1984) 123.
97 R. Bagheri, K.B. Chakraborty and G. Scott, Chem. Ind., 865 (1980); Polym. Deg. and
Stab., 5 (1983) 45.
98 K.B. Chakraborty and G. Scott, Polymer, 21 (1980) 252; Polym. Deg. and Stab., 4
(1982) 1.
99 A.A. Katbab and G. Scott, Chem. Ind., (1980) 573; Eur. Polym. J., 17 (1981) 559.
100 L.P. Nethsinghe and G. Scott, Eur. Polym. J., 20 (1984) 213.
101 R. Bagheri, K.B. Chakraborty and G. Scott, J. Polym. Sei., Polym. Chem. Ed., 22
(1984) 1573.
102 T.A.B.M. Bolsman and D.M. Brouwer, Rec. Trav. Chim. Pays Bas, 97 (1978) 320;
T.A.B.M. Bolsman, A.P. Block and J.H.G. Frijns, Rec. Trav. Chim. Pays Bas, 97
(1978) 313.
103 H. Berger, T.A.B.M. Bolsman and D.M. Brouwer, in G. Scott (Ed.), Developments in
Polymer Stabilisation-6, Applied Science Publishers, London, 1983, Chap. 1.
104 E.T. Denisov, in G. Scott (Ed.), Developments in Polymer Stabilisation-3, Applied
Science Publishers, London, 1980, Chap. 1.
105 G.H. von Fuchs and H. Diamond, Ind. Eng. Chem., 34 (1942) 927.
106 G.H. Denison, Ind. Eng. Chem., 36 (1944) 477.
107 G.H. Denison and P.C. Condit, Ind. Eng. Chem., 37 (1945) 1102.
108 G.W. Kennedy and W.L. Patterson, Ind. Eng. Chem., 48 (1956) 1917.
109 E.A. Oberright, S.J. Leonardi and A.P. Kozacik, in Additives in Lubricants Sym­
posium, ACS Dev. Pet. Chem., Atlantic City, 1958, p. 115.
110 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 188 et seq.
111 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 417 et seq.
112 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 432 et seq.
113 D. Barnard, L. Bateman, M.E. Cain, T. Colclough and J.I. Cunneen, J. Chem. Soc.,
(1961) 5339.
114 W.L. Hawkins, R.H. Hanson, W. Matreyek and F.H. Winslow, J. App. Polym. Sei., 1
(1959)37.
115 W.L. Hawkins, V.L. Lanza, B.B. Loeffler, W. Matreyek and F.H. Winslow, J. App.
Polym. Sei., 1 (1959) 43.
116 W.L. Hawkins and H. Sautter, J. Polym. Sei., AI (1969) 3499.
117 L. Bateman, M.E. Cain, T. Colclough and J.I. Cunneen, J. Chem. Soc, (1962) 3570.
118 J.R. Shelton, in G. Scott (Ed.), Developments in Polymer Stabilisation-4, Applied
Science Publishers, London, 1981, Chap. 2.
119 J.D. Holdsworth, G. Scott and D. Williams, J. Chem. Soc., (1964) 4692.
120 M. Husbands and G. Scott, Eur. Polym. J., 15 (1979) 429.
42 GERALD SCOTT

121 A.J. Burn, Tetrahedron, 22 (1966) 2153.


122 J.A. Howard and J.H.B. Chernier, Can. J. Chem., 54 (1976) 382, 390.
123 S.K. Ivanov, in G. Scott (Ed.), Developments in Polymer Stabilisation-3, Applied
Science Publishers, London, 1980, p. 55.
124 H.S. Oleott and H.A. Mattill, J. Am. Chem. Soc, 58 (1936) 1627, 2204.
125 S. Al-Malaika and G. Scott, in N.S. Allen (Ed.), Degradation and Stabilisation of
Polyoleflns, Applied Science Publishers, London, Chaps. 6, 7.
126 A. Titoff, Z. Phys. Chem., 45 (1903) 641.
127 N.C. Billingham and P.D. Calvert, in N.S. Allen (Ed.), Degradation and Stabilisa­
tion of Polyolefins, Applied Science Publishers, London, 1983, p. 3 et seq.
128 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 172 et seq.
129 Z. Osawa, in G. Scott (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science Publishers, London, 1984, p. 193.
130 A.J. Haagen-Smit, CE. Bradley and M.M. Fox, Ind. Eng. Chem., 45 (1953) 2086.
131 A.J. Haagen-Smit and M.M. Fox, Ind. Eng. Chem., 48 (1956) 1484.
132 R. Criegee, Ber.. 88 (1955) 1878.
133 P.S. Bailey, Chem. Rev., 58 (1958) 925.
134 H. Staudinger, Ber., 58 (1925) 1088.
135 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 477.
136 S.D. Razumovskii and G.E. Zaikov, in G. Scott (Ed.), Developments in Polymer
Stabilisation-6, Applied Science Publishers, London, 1982, Chap. 6.
137 B. Rânby and J.F. Rabek, Photo-degradation, Photo-oxidation and Photo-stabilisa­
tion of Polymers, J. Wiley and Sons, 1975, p. 260 et seq.
138 G. Scott, in B. Rânby and J. Rabek (Eds.), Singlet Oxygen, Wiley, 1978, p. 230.
139 N. Grassie and G. Scott, Polymer Degradation and Stabilisation, Cambridge Uni­
versity Press, 1985, p. 190 et seq.
140 H.J. Stern, in CM. Blow and C Hepburn, Rubber Technology and Manufacture,
Second edition, Butterworth, 1982, p. 2.
141 W.F. Busse and E.N. Cunningham, Proc. Rubber Tech. Conf. 1938, p. 288.
142 W. Kauzmann and H.J. Eyring, J. Am. Chem. Soc, 62 (1940) 3113.
143 M. Pike and W.F. Watson, J. Polym. Sei., 9 (1952) 229; W.F. Watson, Trans. IRI, 29
(1953) 32.
144 C Ayrey, CG. Moore and W.F. Watson, J. Polym. Sei., 19 (1956) 1.
145 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 469 et seq.
146 A.S. Kuzminsky, A.G. Maizels and N.N. Lezhnev, Dokl., Akad. Nauk. SSSR, 771 (2)
(1950) 319.
147 A.S. Kuzminsky and M.G. Maizels, Chemistry and Physico-Chemistry of High-
Molecular Compounds, Nauka, Moscow, 1952.
148 G.L. Slominsky, V.A. Kargin, G.N. Biuko, E.V. Rezsova and M.L'yuis-Riera, Stareni
i Utomlenie, UN. I. T. O. Rezinshcikov Konf., 1953, p. 100.
149 M.S. Kharasch and F.R. Mayo, J. Am. Chem. Soc, 55 (1933) 2468.
150 L. Bateman and G. Gee, Proc. R. Soc, A195 (1948-9) 376, 391.
151 C.H. Bamford and M.J. S. Dewar, Proc. R. Soc, A198 (1949) 252.
152 J.T. Martin and R.G.W. Norrish, Proc. R. Soc, A220 (1953) 322.
153 R.G.W. Norrish and M.H. Searby, Proc. R. Soc, A237 (1956) 464.
AUTOXIDATION AND ANTIOXIDANTS: HISTORICAL PERSPECTIVE 43

154 A.M. Trozzolo, in W.L Hawkins (Ed.), Polymer Stabilisation, Wiley Interscience,
1972, p. 159 et seq.
155 A.M. Trozzolo and F.H. Winslow, Macromolecules, 1 (1968) 98.
156 D.J. Carlsson, A. Garton and D.M. Wiles, in G. Scott (Ed.), Developments in Polymer
Stabilisation-1, Applied Science Publishers, 1979, p. 219 et seq.
157 D. Bellus, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, Wiley, 1978, p. 61 et
seq.
158 H.J. Heller and H.R. Blattman, Pure App. Chem., 36 (1973) 141.
159 S. Al-Malaika and G. Scott, in N.S. Allen (Ed.), Degradation and Stabilisation of
Polyolefins, Applied Science Publishers, London, 1983, Chap. 7.
160 R.T. Holman and O. Elmer, J. Am. Oil Chem. Soc, 24 (1947) 127.
161 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 368 et seq.
162 A.L. Tappel, Arch. Biochem. Biophys., 44 (1953) 378.
163 D. Chiu, B. Lubin and S.B. Shohet, in W.A. Pryor (Ed.), Free Radicals in Biology,
Vol. V, 1982, Chap. 5.
164 R.S. Sohal and R.G. Allen, Adv. Free Radical Biol. Med., 2 (1986) 117.
165 R.J. Melhorn and G. Cole, Adv. Free Radical Biol. Med., 1 (1985) 165.
166 D. Harman, J. Gerontol, 11 (1956) 298.
167 D. Harman, J. Gerontol., 26 (1971) 451.
168 G. Scott, Chem. Br., (July 1985) 648.
169 H.S. Olcott and H.A. Mattill, Chem. Revs., 29 (1941) 257.
170 W.H. Selrelland and R.S. Harris (Eds.), The Vitamins: Chemistry, Physiology and
Pathology, Vol. 3, Academic Press, N.Y., 1954.
171 R.W. Reimenschneider and W.C. Ault, Food Ind., 16 (1954) 892.
172 C. Golumbic, Oil Soap, 19 (1942) 181; 20 (1943) 105.
173 J.E. Packer, T.F. Slater and R.L. Willson, Nature, 278 (1979) 737.
174 G. Scott, Atmospheric Oxodation and Antioxidants, First edition, Else vier, London
and New York, 1965, p. 204 et seq.
175 C. Golumbic and H.A. Mattill, J. Am. Chem. Soc, 63 (1941) 1279.
176 B.M. Watts and R. Wong, Arch. Biochem., 80 (1951) 110.
177 H. Dann, Pharmacol. Revs., 9 (1957) 1.
178 M.K. Horwitt, Fed. Proa, 24 (1965) 68.
179 See, for example, A.L. Lehninger, Biochemistry, Worth, 1977, p. 357.
180 A.L. Tappel, Fed. Proa, 24 (1965) 73.
181 A.L. Tappel, Geriatrics, 23 (1968) 97.
182 C.W. Burton, D.V. Foster, B. Perly, T.F. Slater, I.C.P. Smith and K.U. Ingold, Phil.
Trans. R. Soc, B311 (1985) 565.
183 W. Treibs, Ber., 75B (1942) 953; 76B (1943) 670; 80 (1947) 423.
184 F.D. Gunstone and T.P. Hilditch, J. Chem. Soc, (1946) 1022.
185 I. Fridovich, Science, 209 (1978) 875.
186 B. Halliwell and J.M.C. Gutteridge, Mol. App. Med., 8 (1985) 89.
187 C. Walling, in T.E. King, H.S. Masai and H. Morrison (Eds.), Proa 3rd. Int. Symp.,
Oxidases and Related Redox Systems, Pergamon Press, 1982 p. 85.
188 I. Fridovich, Adv. Enzymol., 41 (1974) 35.
189 P. Nicholls and G.R. Shonbaum, in P.D. Boyer (Ed.), The Enzymes, Vol. 8, Academic
Press, N.Y., 1963, p. 149.
190 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
44 GERALD SCOTT

and New York, 1965, Chap. 5.


191 A. Jacobs, Semin. Hematol., 14 (1977) 89.
192 A.D. Heys and T.L. Dormandy, Chem. Sei., 60 (1981) 295.
193 H.R. Schumacher, Arthritis Rheum., 25 (1982) 1460.
194 D.R. Blake, P.A. Bacon, F.J. Eastham and K. Brigham, Br. Med. J., 281 (1980) 715.
195 G. Scott, Atmospheric Oxidation and Antioxidants, First edition, Elsevier, London
and New York, 1965, p. 173 et seq.
196 N. Grassie and G. Scott, Polymer Degradation and Stabilisation, Cambridge Uni­
versity Press, 1985, p. 119 et seq.
197 L. Flohe, in W.A. Pryor (Ed.), Free Radicals in Biology, Vol. 5, Academic Press, 1982,
Chap. 7.
198 D.G. Pobedimskii, N.A. Mukmeneva and P.A. Kirpichnikov, in G. Scott (Ed.),
Developments in Polymer Stabilisation-2, Applied Science Publishers, London,
1980, Chap. 4.
199 N. Hayawara and T. Kagiya, in H.H.G. Jellinek (Ed.), Degradation and Stabilisation
of Polymers, Elsevier, 1983, Chap. 8.
200 G. Scott, in N.S. Allen and J. F. Rabek (Eds.), New Trends in Photochemistry of
Polymers, Elsevier Applied Science, London, 1985, Chap. 14.
45

Chapter 2

AUTOXIDATION

S. AL-MALAIKA

1. INTRODUCTION

All organic compounds are susceptible to attack by molecular oxygen. A


basic autoxidation scheme based on a free radical chain reaction theory
(Scheme 1) was developed in the early forties after extensive kinetic studies
of autoxidation of olefins which was carried out mainly by researchers at the
British Rubber Producers Association [1-3]. The study of autoxidation of
hydrocarbons and polymers was continued in the fifties and sixties by
leading researchers; their work has been fully covered in the literature
[4—15]. The oxidation of hydrocarbons and saturated polyolefins is auto-
catalytic, that is, the reaction starts slowly, possibly with a short induction
period, followed by gradual increase in rate (rate increases as hydroperoxide
concentration builds up) which will eventually subside giving rise to a
sigmoid curve as a function of time. The short induction period is removed
by addition of initiators (e.g. peroxides, azo compounds) and extended by
antioxidants and stabilisers.
Hydrocarbon polymers are normally processed (e.g. by extrusion and
injection moulding) at high temperatures to produce the final fabricated
article. The temperatures required for these processes are such that the
polymer is susceptible to thermal breakdown but this process is accelerated
by the presence of atmospheric oxygen or even the small concentrations of
residual oxygen (dissolved in the polymer or trapped in the polymer feed).
Various oxygen-containing groups are introduced in the oxidised substrate
which in the case of polymers will lead not only to changes in the molecular
structure and formation of low molecular weight products, but also to
deterioration in mechanical and electrical properties of the final polymer
article.
This chapter aims to cover the scientific aspects of autoxidation (mecha­
nism, kinetics and effect of molecular structure) in addition to the techno­
logical effects of oxidation and the main techniques used in studying effects
of autoxidation.
46 S. AL-MALAIKA

Initiation: production of R- or ROO- radicals (1)

ko
Propagation: R+02 ROOH+ R- (2)

ROO- + RH —2—> ROOH+ R- (3)

ROO+C=C * 3a > ROO-C-O (3a)


/ \

2R- kA
Termination: (4)

R. + ROO- -^_ Non-radical products (5)

2ROO- *6 > (6)

Scheme 1. Basic autoxidation mechanism.

2. MECHANISM AND KINETICS OF AUTOXIDATION

The free radical chain mechanism in autoxidation, like other radical


reactions, consists of the three basic steps of initiation, propagation and
termination as shown in Scheme 1 (R- and ROO- are alkyl and alkylperoxyl
radicals derived from the organic substrate RH). The initial classical kinetic
investigations of Bolland and Gee [1,16] on the photo-initiated oxidation of
cyclohexene in the presence of a free radical generator showed that the rate
of oxygen consumption corresponded closely to the rate of formation of
hydroperoxides and the rate law is shown in Eqn (7) (experimentally, at
oxygen pressure >50 torr, the functional dependence on oxygen concentra­
tion, f [0 2 ], is taken as equal to one). In the presence of sufficient oxygen
pressures, the relative steady state concentrations of radicals are [ROO-1 > >
[R-] and chain termination (whose rate is equal to the rate of chain initia­
tion, R) occurs entirely by Reaction (6). Based on the above, the general rate
law for autoxidation of large number of organic compounds is given by Eqn
(7a). The overall rate of oxidation of organic substrates, therefore, depends
on Ri and absolute values of ft3 and k$ (the ratio k^/ik^1' is a measure of the
substrate oxidisability) which is in agreement with the observed rate law,
Eqn (7).

d[Q2l
1/2.
Ä[RH]Äf /z
/I0 2 ] (7)
dt
AUTOXIDATION 47

l
dt dt dt ( Ä 6 )i/2

In the case of polymers, the rate of oxidation at ambient temperature in the


dark is usually very small and the free radical chain reaction is initiated
either thermally, photochemically, mechanically, or by high energy radia­
tion. Commercial hydrocarbon polymers normally contain chemical impuri­
ties such as small concentrations of oxygenated compounds, unsaturation,
peroxides, catalyst residues, e.g. metal ions, which are a consequence of the
polymerisation process and will have significant effect on the rate of initia­
tion and the overall rate of oxidation.

2.1 Initiation Reactions

The nature of the primary initiation reaction of molecular oxygen with


hydrocarbons, which is responsible for the production of the first free
radicals in the chain sequence, is complicated by interference from other
reactions such as hydroperoxide initiation or metal catalysis. A number of
mechanisms have, therefore, been proposed for the chain initiation process
(in absence of initiators, e.g. hydroperoxides), Reactions (8-10). Initiation by
direct reaction of most hydrocarbons with molecular oxygen in a bimolecular
process (Reaction (8)), is both thermodynamically and kinetically un­
favourable. An exception to this is the observation that radicals are formed
during the primary uncatalysed initiation step of indene [17], styrene [18]
through the direct attack of olefin by molecular oxygen in a bimolecular
process (Reaction (8)). Bromberg and Muszket [19] showed that the initia­
tion process for the self-initiated and the inhibited rate of oxidation of the
dihydrophenanthrene, I, and its substituted analogue, II, can be repre­
sented by Reaction (9) with apparently high activation energy.

RH + 0 2 > Free radicals (8)

or more specifically:

RH + 0 2 > R+HOO (9)

2RH + 0 2 > 2R+H202 (10)

This reaction is very slow at room temperature (based on differences in


bond dissociation energy for C-H, > 313.8 kJ mol - 1 , and that of H-OO, 190
kJ mol" 1 but is moderate at high temperatures (above 100°C). The rate of
Reaction (9) was measured at high temperatures for a number of substrates;
48 S. AL-MALAIKA

e.g. 4.0 x ÎO" 8 M s"1 for cyclohexanone at 130°C and 2.6 x 10" 9 M s" 1 for
n-decaneatl50°C[20].
Denisov [21] proposed that the termolecular initiation Reaction (10) is
thermodynamically more favourable. Kinetic studies which supported this
view were reported by Denisov [21] for tetralin and indene and by Carlsson
and Robb [22] for tetralin, cyclohexanol and cyclohexanone. It was suggested
[22,23] that this reaction may occur via the formation of a complex between
the substrate and oxygen followed by a rate determining step which involves
a reaction between the complex and a second substrate molecule (Reaction
(10a)) with an overall initiation rate, Rv (Reaction (10b)).

k 10a 7
*10a'
RH + Oo [RH...Oo] + RH » 2R+H202 (10a)
-10a

(k 10a' ^lOa
R{ = n [RH]^[02] (10b)
k -10a

As oxidation proceeds and hydroperoxide concentration builds up in the


substrate, the energetically more favourable hydroperoxide decomposition
becomes the prédominent initiating step, Reactions (11) and (12). The
primary product of autoxidation is, therefore, the hydroperoxide which
decomposes to free radicals; increasing build up of peroxides in the substrate
leads to an increased rate of initiation, and hence oxidation, resulting in the
characteristic initial autocatalytic rates of oxygen absorption as is shown in
Fig. 1. The rate of hydroperoxide decomposition is influenced by a number
of factors as shown below.

ROOH R O + OH (11)

2ROOH -» RO- + ROO- + H 2 0 (12)


AUTOXIDATION

<
Q.

Z
LU
o
>-
X
o

TIME ^

Fig. 1. Schematic representation of the relationship between hydroperoxide decomposi­


tion and rate of oxidation.

(a) Hydroperoxide concentration


The concentration of hydroperoxide affects the molecularity of the reac­
tion; in general, bimolecular decomposition (Reaction (12)) occur a t high
hydroperoxide concentrations (above 1.9 M). The dependence of the type of
decomposition (uni- or bi-molecular) on the hydroperoxide concentration
was established from kinetic studies using infra red spectroscopy [24-26]:
decreasing peroxide concentration is accompanied by shifts of the broad
absorption of O-H groups (e.g. a t 3350 cm"1) of the neat hydroperoxide
(presumably due to hydrogen-bonded dimer) to higher frequencies together
with narrowing of O-H absorption band (at 3500 cm"1) due to the disso­
ciated (non-bonded) groups (see Reaction (12a)). Bateman and coworkers
[24,25] found that a t low oxygen pressures (i.e. low hydroperoxide concen­
tration) the rate of oxidation (ß oxdn ) at the early stages for various olefinic
hydrocarbons, followed first order kinetics (Reaction (11a), Fig. 2d), while

20 60 100
Burette reading, mm Hg

Fig. 2. Autoxidations of 2,6-dimethylhepta-2,5-diene at 25°C and pressure of (a) 750 mm,


(b) 144 mm, (c) 55 mm, and (d) 30 mm [24]. The bar lines are conversion scales for equal
extent of oxidation.
50 S.AL-MALAIKA

with increasing oxygen pressure the ability to define the earlier reaction
stages becomes more blurred; bimolecxilar peroxide decomposition may
dominate under these conditions (see Reaction (12b), Fig. 2a).

2(RO + OH) <- 2ROOH ^z=± [ R O O HOORJ -> R O + H 2 0 + R O O (12a)


H

*oxdn~[ROOH] 1/2 (lia)


Ä 0 xdn-[ROOH] (12b)

^oxdn ~ (Ä'tROOH] + k"[ROOH] 2 ) 1/2 (13)

If both uni- and bi-molecular decompositions of hydroperoxide are in­


volved in the initiation step then the oxidation rate (Ä0xdn) m a Y be repre­
sented by Eqn (13) where k' and k" are first order and bimolecular rate
constants for Reactions (11a) and (lib), respectively. A plot of Rox^n2/
[ROOH]2 vs. [ROOH]" 1 should give a straight line, see Fig. 3 [24]. Factors
affecting the equilibrium between monomeric and dimeric hydroperoxide
structures are important in determining the decomposition path. For ex­
ample, unimolecular decomposition is favoured by raising the reaction
temperature (dissociation energy of RO-OH is about 175 k J mol"1) or by
having an R group with unfavourable steric or electronic properties in the
associated state. Table 1 compares the rate constants for the unimolecular
and bimolecular hydroperoxide decomposition for various olefins [4]. In the
case of polypropylene, it was shown [25] that thermal decomposition of
adjacent hydroperoxide groups of isotactic polypropylene proceeds bimole-
cularly and many faster than the decomposition of isolated hydroperoxide
groups which decompose unimolecularly.

100 200
[ROOH]*1, Liter mole1

Fig. 3. Autoxidation of 2-methyl-4-phenylbut-2-ene at 55°C and 730 mm [24].


AUTOXIDATION 51

TABLE 1

Comparison of unimolecular and bimolecular RO2H decomposition rate constants for


various olefins [4]

Olefin Hydroperoxide type % fc-106 bimol fc-107


- (55°C)(a) unimol,
Primary Secondary Tertiary moles -1 1 s"1 s"1

Allyl benzene 100 0.28 2.9(75°©


Cyclohexene 100 0.54 0.15 (55°C)
Ethyl linoleate 100 0.47 0.20 (55°C)
1-Methylcyclohexene 70 30 1.14 0.45 (45°C)
3.9(65°C)
Dicyclohex-2-enyl 25 75 2.48 0.7(45°C)
6.9(65°C)
Digeranyl 100 2.62 0.54 (45°C)

(a)
Both coefficients should be corrected for an efficiency factor, i.e., the number of
oxidation chains initiated relative to the number of radicals liberated in the initial
decomposition process. This can be done; the factor varies from 50 to 75% in the olefins
listed, which does not affect the general conclusion drawn.

(b) Induced decomposition by free radicals


In general the rate of decomposition of tertiary hydroperoxides (£-ROOH)
is strongly affected by induced decomposition by self-generated radicals, see
Reaction (14) (under conditions where Reaction (11) is negligible). This was
shown [28,29] to be mainly due to the formation of alkoxyl radicals by
non-terminating reaction of the tertiary peroxyl radicals, Reaction (15a) (the
corresponding di-tertiary peroxide was shown to be one of the products,
Reaction (15b)). Thus, for example, the chain length in the thermal decom­
position of ter£-butyl hydroperoxide was found to be 10 compared with 4 for
cumyl hydroperoxide [30].

*-RO + *-ROOH *-ROH + *-ROO (14)

2*-RO + 0 2 (15a)
2*-ROO
J-ROOR-* + 0 9 (15b)
52 S. AL-MALAIKA

TABLE 2

Effect of solvents on the thermal decomposition of tetralyl hydroperoxide at 170°C in the


presence of an antioxidant [317]

Solvent fc-104 (s"1) E& (kJ mole"1)

Iso octane 1.31


Butyl stéarate 1.47 86
Tetradecane 2.32.
Octadecane 2.54 83.3
White oil 4.84 88.5
1-Hexadecene 7.92 82.4
Dioctyl ether 14.5 84.2
2-Ethyl-l-hexene 12.6 67.8
2,4,4-Trimethyl-1-pentene 16.7
Polybutene 21.7 65.6
Polypropylene 25.0

(c) The effect of solvent


The solvent can have a remarkable effect on the decomposition rate of
hydroperoxides. For example, it was shown [31] that in the case of tetralyl
hydroperoxide, the decomposition was more rapid and proceeded with low
activation energies in unsaturated hydrocarbons (see Table 2). This is not
due to radical induced reactions (see Chapter 3, Section 2.2.4) since decom­
positions were carried out in the presence of an antioxidant.

(d)Acid catalysed decomposition


Decomposition of peroxides by acids is a non radical reaction giving
molecular products (e.g. phenol and acetone from cumene hydroperoxide)
and hence is normally responsible for inhibition (see Chapters 1 and 5).

(e) Metal catalysed decomposition


Metal ions are known to catalyse the decomposition of hydroperoxides by
redox Reactions (16a) and (16b) [34,35]. This catalytic action is only ob­
served when both oxidation states of the redox couple have comparable
stability. If the metal ion is only capable of effecting one of these reactions
(i.e. no pathway is available to regenerate the other oxidation state of the
metal ion), a stoichiometric decomposition of hydroperoxide will result. The
overall effect is a lowering of the activation energy of the bimolecular
peroxide decomposition reaction, Reaction (16c) (activation energy for ther-
AUTOXIDATION 53

mal decomposition of hydroperoxides is about 170-191 kJ mol"1, whereas


that of the metal catalysed decomposition is only 42.5-85 kJ mol"1.
The relative rates of these two reactions may be roughly correlated with
the redox potential of the particular M n /M / l + 1 couple. Redox potentials are
influenced by the nature of the ligand and the solvent and careful considera­
tion must be given when using redox potential values for aqueous solutions
[36] since values in organic solvents are not readily available.

M" + ROOH > M n+1 + RO. + OH" (16a)

M ^ + ROOH > Mn + ROO. + H+ (16b)

Overall,

M
2ROOH " / M n * > ROO. + RO. + H 2 0 (16c)

Induced decomposition of hydroperoxide by metal ions (M7M"+1 couples)


such as Co, Cu, Fe, and to a lesser extent, Mn, has been thoroughly
investigated [8,37-42]. It was shown that, in general, when the metal ion in
the complex is a strong oxidising agent (e.g. Pb IV , CeIV) Reaction (16b)
predominates, whilst if it is a strong reducing agent (e.g. Cr , Cu , Fe )
Reaction (16a) predominates [38-44]. The role of metal ions in redox cataly­
sis is two-fold: (i) the formation of free radicals which initiate autoxidation
and, (ii) the conversion of hydroperoxides into secondary oxidation products
such as alcohols, ketones, aldehydes and carboxylic acids. In the case of Cr
II and Cu I, for example, both were shown to reduce hydroperoxides to the
corresponding alcohols, but whilst chromium II undergoes a stoichiometric
reaction, see Reaction (17a) and (17b) (since no route is available for re­
generation of Cr II), copper I reacts with hydroperoxide in a catalytic process
due to the possibility of regeneration of Cu I by different routes, e.g. Cu I can
be regenerated via electron transfer oxidation of alkyl radicals which are
produced from the fragmentation of alkoxyl radical, see Reaction (18a) and
(18b) [38,39,45]. The structure of the ligand containing the metal has a
marked effect on the catalytic activity in the oxidising system [36].

ROOH + Cr11 > RO + Cr m (OH) (17a)

RO + Cr n > RO + C r m (17b)

ROOH + Cu1 > Cun(OH) + RO- > R. + >C= O (18a)


54 S. AL-MALAIKA

R- + Cuil Cux + H + + >C=C< (18b)

It has been shown [40] that there is always a competition between the
metal catalysed decomposition and radical-induced decomposition of hydro-
peroxides which is influenced by the ratio of hydroperoxide to the metal
complex. At high hydroperoxide ratios, alkoxyl radicals compete effectively
with the metal complex for the alkyl hydroperoxide. Furthermore, in the
presence of a solvent, especially if it can easily be attacked by alkoxyl
radicals, the relative amounts of hydroperoxide and the solvent is also
important. Contribution from metal catalysed reactions, therefore, predom­
inates at low hydroperoxide concentration and in reactive solvents. The
faster rate of metal catalysed decomposition which is observed for tertiary
hydroperoxides (compared with those of primary and secondaryanalogues)
was attributed to the induced decomposition by alkoxyl radicals.

Fig. 4. (a) Variation in maximum rate of oxidation of tetralin with cobalt concentration
[151. (b) Rate of oxidation of tetralin in chlorobenzene at 65°C as a function of cobalt
decanoate concentration. Numbers on curves are tetralin concentration in M [49].
AUTOXIDATION 55

The rate of metal-catalysed oxidation at low concentration of catalysts


(e.g. below KP 3 M in case of cobalt) is proportional to one half order with
respect to the metal concentration ([M] ^ ) , see Fig. 4a, and is nearly one half
order with respect to the rate of initiation [15,46,46a]. However, it was
shown that the catalytic effect of metal salts reaches a maximum value at
some critical concentration, above which the rate becomes independent of
the catalyst concentration, see Fig. 4a. This observation of limiting rates
[15,47] has been explained in terms of an equilibrium between rates of
formation and decomposition of hydroperoxides which is established at high
catalyst concentration. From a practical point of view, the limiting rate is
very useful, since it helps to predict the upper limit of oxidation rate and
distinguishes the mechanism of catalysis [47]. This dual role of metal salts
(as catalysts and inhibitors below and above the critical concentration
respectively), was demonstrated for cobalt-catalysed autoxidation of tetra-
lin [48,49]. The rate of oxidation continues to increase until the catalyst
reaches a critical concentration at which point a further small increase in
cobalt concentration inhibits the oxidation and is accompanied by a sharp
drop in the oxidation rate, see Fig. 4b.

2.2 Propagation Reactions

(a) Coupling reaction


Alkyl radicals which are formed in the initiation step react immediately
(with little or no activation energy [50]) with oxygen (a biradical) in a fast
radical-radical coupling reaction, Reaction (2) (Scheme 1); the rate of oxy­
gen uptake in this reaction follows that of a normal bimolecular reaction,
Eqn (19).

dtOg]
- ^ = /e2[R-][02] (19)

Very few absolute values of &2 have been determined in the liquid phase.
The reactivity of the polstyrenyl radical towards oxygen was found [50] to
be much faster (with k2 of 10 8 M"1 s"1) than with styrene even though the
latter is known to be extremely reactive towards free radicals. Ingold [51]
has shown that the rate constants for the reaction of most alkyl radicals with
oxygen to be in the order of 10 9 M"1 s _1 . There are, however, exceptional
cases where Reaction (2) (Scheme 1) does not occur readily and Reaction (3)
(Scheme 1) can compete with Reaction (2) at normal oxygen pressure. This
occurs mainly when the alkyl radical is highly resonance stabilised [52] as
in the case of pentaphenylcyclopentadienyl radical (III), 2,6-dimethylhepta-
2,5-diene radical (IV), or (V).
56 S. AL-MALAIKA

~fr >—=<; OX
Ph

Ph Ph
III IV V

(b) Hydrogen abstraction


Compared with the oxygen-radical coupling reactions, the abstraction of
a hydrogen atom by peroxyl radicals requires a higher energy of activation
(Reaction (3), Scheme 1). Substrates which contain an activated C-H bond,
e.g. alkene, aralkane, aralkene, ethers, ketones, Reaction (3), which involve
the breaking of a carbon-hydrogen bond, have a higher activation energy.
This is the most important reaction in autoxidation and is the rate deter­
mining step. Increasing the rate constant of this reaction, &3, increases the
rate of oxidation and the concentration of hydroperoxide, and the propagat­
ing chain becomes longer.
Although most rate controlling propagation reactions involve inter-
molecular hydrogen abstraction, peroxyl radicals can abstract a hydrogen
via an intramolecular reaction, e.g. Reaction (20), particularly in aliphatic
systems containing two reactive hydrogen atoms in the 1,3 or 1,4-position
[53]. It was also found [53,54,47] that intramolecular reaction was the major
propagation path in the oxidation of a number of branched alkanes while
normal alkanes underwent very little intramolecular propagation [55]. For
example, [53], 2,4-dimethylpentane and 2,4-dimethoxypentane gave very
high yields of dihydroperoxide through intramolecular attack of the initially
formed tertiary peroxyl radical on the remaining tertiary hydrogen atom
followed by the formation of a second peroxyl group, Reaction (20).

CH 3 ) 2 -C-CH2-C(CH 3 ) 2 > (CH 3 ) 2 -C-CH 2 -Ç(CH 3 ) 2 — ^


O-O. H OOH

CH 3 ) 2 C-CH 2 -C(CH 3 ) 2 > (CH 3 ) 2 C-CH 2 -C(CH 3 ) 2 (20)


OOH 00- OOH OOH

Internal hydrogen abstraction occurs predominantly on the 7-carbon of


the peroxyl radical via a 6-membered cyclic transition state, VI, and not on
the ß- or ô-carbons due to the less stable five- or seven- membered cyclic
transition state [53]. The rate law for the oxidation of 2,4-dimethylpentane
(HRH) obeys closely Reaction (21) which is identical to the general rate law
of autoxidation (see Eqn (1)) except for the factor of two which accounts for
the formation of dihydroperoxide.
AUTOXIDATION 57

R ß R
R-Co yC—R (VI)
I I
O.0,H

Ä0xdn=7^^/2[HRH] (21)

Based on results from simple hydrocarbon model compounds, it was


shown [56-60] that intramolecular propagation plays an important role in
some polymers which contain branched alkane structures like polypropy­
lene (PP) and polystyrene. For example, Chien and co-workers [57] ex­
amined polypropylene hydroperoxides (prepared by autoxidation of PP) and
found that more than 90% were intramolecularly hydrogen bonded. In
contrast, intramolecular propagation is unimportant in the oxidation of
2,4-pentanediol [61] because of hydrogen bonding between peroxyl radicals
and the hydroxy group. Intramolecular propagation is, therefore, not impor­
tant in the oxidation of polyvinylalcohol. Howard and Ingold [62] have
shown that in the case of ethers, e.g. diisopropyl ether and dibenzyl ether,
both mono- and di-hydroperoxides were formed but with higher yield of the
former, suggesting a preference for intermolecular hydrogen abstraction in
these ethers.

(c) Addition reaction


The unsubstituted double-bond position of 1-ene olefins is particularly
susceptible to attack by peroxyl radicals via an addition reaction (Reaction
(3a), Scheme 1). Mayo and co-workers [6,62b,64] and Howard and Ingold
[62a] studied the autoxidation of a large number of unsaturated hydrocar­
bons with terminal méthylène groups, especially conjugated olefins, e.g.
styrene. The peroxyl radical always adds to the double bond in such a way
as to produce the more stable ß-peroxy alkyl radical, e.g. Reaction (22a),
which subsequently combines with oxygen to regenerate a peroxyl radical,
Reaction (22b), hence di- or poly-peroxides are produced which contain two
or more substrate molecules.

ROO + CH 2 =CH > ROO-CH2-CH- (22a)


Ph Ph

ROO-CH2-CH- + 0 2 > R00-CH2-CH-0O (22b)


Ph Ph
58 S. AL-MALAIKA

Some olefins, e.g. indene, oxidise to form a copolymer with oxygen, i.e.
polyperoxide formation [17,65], in addition to the formation of hydroper-
oxide. In this case the peroxyl radical can either abstract a hydrogen from
the highly reactive méthylène group of indene, Reaction (23a), or add to the
double bond to form ß-peroxy alkyl radical which on further reactions with
oxygen and indene lead to the polyperoxide, Reaction (23b). At 30°C, the
reaction leads to 90% addition and 10% abstraction [17,65].

ROO- + [ j II > ROOH +


ir ^H H H oo
(R'OO-)

(23a)
R'OO- +

H H
0 2 , indene
(23b)

H. M O—O

H
00

2.3 Effect of Substrate Structure on the Rate of Propagation

Measuring the overall rate of oxidation at a known constant rate of


initiation makes it possible to determine the ratio of rate constants, k^/k^ ' ;
see Eqn (7) [1,7,66-68].
The ease of hydrogen abstraction from the substrate by ROO. radicals
affects the magnitude of k% (see Reaction (3)). Bolland [66] studied the
oxidation of a large number of olefins and examined the nature and the
position of substitution on the different carbons of the olefin structure
(represented by the allylic system, VIII) which affects the reactivity of the
a-methylenic group (hence Ä3) and deduced the following general rules.
c-c=c
ab c
VII
AUTOXIDATION 59

1. Replacement of one or two hydrogen atoms at position a or c by an alkyl


group increases A3 of the a-group by 3.3" (n = total number of R groups
introduced) while replacing the hydrogen at position b has no effect.
2. Replacement of H at position a by phenyl increases k% 23 times, while
Alk-CH=CH group causes an increase of A3 by 107.
3. The presence of an allylic system in a cyclic structure leads to an
a-methylenic activity which is 1.7 times greater that that of the acyclic
analogue.
The effect of structural changes in the substrate on the yield of hy-
droperoxide (note that Reaction (3) leads to hydroperoxide while (3a) leads
to polyperoxides), hence the effect on the competing Reactions (3) and (3a),
was further examined by Hargrave and Morris [67] who studied the yield of
hydroperoxides from various olefines under high kinetic chain length condi­
tions, i.e. oxidation rate is equal to the rate of hydroperoxide formation (see
Eqn (24)) and hydroperoxide concentration is high.

k2[R'] [0 2 ] = /e3[ROO] [RH] (24)

The effect of the position of the double bond on the reactivity and relative
rate of oxidation can be seen from the difference in hydroperoxide yield
which follows the trend of the a-methylenic activity of the olefin substrates
listed below [67]. Moreover, in the case of the 2-ene olefins, substitution of
a phenyl group on the 4-position was shown to give almost theoretical
hydroperoxide yield due to the increase in the lability of the a-methylenic
hydrogen atom;

a-CH2 activity in: Cyclohexene > Hept-3-ene > hept-2-ene > Oct-1-ene
Hydroperoxide yield (%): 95 > 84 > 79 > 69

Similar effects of substitution on the 3-position of 1-ene olefins were also


observed [67]. The decrease in hydroperoxide yield in the case of unsubsti-
tuted terminal double bonds (1-ene olefins) is mainly due to the competing
addition reaction, Reaction (3a), leading to di- or poly-peroxides (see later).
It was shown [7,68], however, that the early attempts [1,66] to correlate
oxidisability with structure of the substrate were only qualitative since the
rate constant ratio, k^/kßl/2, can be affected not only by the substrate
reactivity, but also by reactivity of the peroxyl radical and the absolute
magnitude of the termination rate constant. Howard and Ingold suggested
that meaningful correlations between structure and reactivity can be
achieved if (a) reactivity of the substrate to a particular peroxyl radical is
determined and (b) comparison of overall rates of oxidation is confined to
compounds which give peroxyl radicals which show the same reactivity
towards propagation and termination. These workers [7,68] were able to
60 S. AL-MALAIKA

overcome these problems by oxidising the substrate (RH) in the presence of


a high enough concentration of hydroperoxides, e.g. £-butyl hydroperoxide,
so that all the peroxyl radicals derived from the substrate undergo chain
transfer with the added hydroperoxide, Reaction (25a), and the termination
is by peroxyl radicals derived from the hydroperoxide (see Reactions (25b)-
(25d)). The overall rate of oxidation is given by Eqn (25e).

Chain Transfer: ROO + *-BuOOH ROOH + *-BuOO (25a)

?25b
Propagation: *-BuOO + RH > *-BuOOH + R- (25b)

R- + 0 2 > ROO (25c)

k25d
Termination: 2£-BuOO Non-radical products (25d)

-d[Q2l ^25b
RY2 [RH] (25e)
At 1/2
të*25d)

where absolute values of &t25d a n ( * ^25b were calculated. Using this method,
Howard and Ingold (Ref. [15] and references therein) obtained data for the
absolute reactivities of substrates with different structural features towards
certain peroxyl radicals (derived from the hydroperoxides added) and com­
pared it with reactivities towards their own peroxyl radicals. Both steric and
polar effects play an important role in the propagation reactions. In general,
peroxyl radicals are quite selective in hydrogen abstraction (abstracting
tertiary hydrogen in preference to secondary and primary), especially at
lower temperatures, and primary and secondary peroxyl radicals were
shown [69,70] to be several times as reactive as tertiary radicals for hydro­
gen abstraction (Table 3). The difference in reactivity was attributed mainly
to steric effects. However, polar effects can be quite important in the
oxidation of some substrates, e.g. cumene, 1,1-diphenylethane, and a-methyl-
benzyl alcohol, For example, linear relationship was observed between the
ratio of rate constants, k^k2^ (where A3 is the rate of propagation of
substrate with its own peroxyl radicals) and the a-substituent (Hammett)
constants, £ a m . The effect of steric hinderence on the reactivity of the
substrate in the hydrogen abstraction propagation step (Reaction (3),
Scheme 1) was illustrated [71] by the effect of two different ortho-alkyl
substituents on the rate of hydrogen abstraction from phenols: 2,6-di-tert-
butylphenol is less reactive than 2,6-dimethylphenol.
Polar effects play a dominating role in determining the magnitude of the
rate constant for hydrogen atom abstraction. Rüssels [72] demonstrated this
AUTOXIDATION 61

TABLE 3

Relative reactivities of some peroxyl radicals [69]

Hydrocarbon Ä3(a) w» &3/&25b

Toluene 0.24 0.034 7.0


1-Octene 1.0 0.27 3.7
Ethylbenzene 1.3 0.21 6.2
Diphenylmethane 2.0 0.50 4.0
Tetralin 6.4 1.65 4.0
9,10-Dihydroanthracene 175 70 2.5
sec-Amylbenzene 0.07 0.07 1.0
Cumene 0.18 0.18 1.0
2,3-Dimethyl-2-butene 1.7 1.7 1.0
Styrene 41 2.1 19
a-Methylstyrene 10 3.9 2.6

(a)
R02- + RH -^2_> ROOH + R-

^ C6H5(CH3)2C02- + RH-^b> C6H5(CH3)2COOH + R-

TABLE 4

Relative reactivities of aralkyl hydrocarbons toward ROO- radicals at 90°C (per hydrogen
atom), [72]

Tertiary Secondary Primary

p-Di-isopropylbeneze 2.7 Dibenzylether 21 Mesitylene 0.26


p-fer£-Butylcumene 2.3 Tetralin 9.4 p-fer£-Butyltoluene 0.26
p-Cymene 2.3 Indane 5.4 p-Xylene 0.23
Cnmene 1.9 Allylbenzene 5.0 p-Cymene 0.21
p-Bromocumene 1.47 Diphenylmethane 2.1 m-Xylene 0.15
Methyl-p-cumate 1.06 Ethylbenzene 1.1 Toluene 0.15
p-Nitrocumene 1.00 Methylphenylacetate 0.4 p-Chlorotoluene 0.12
p-Cyanocumene 0.94 Benzylbenzoate 0.36 p-Tolunitrile 0.05
Phenylcyclohexane 0.47 Benzylchloride 0.28 p-Nitrotoluene 0.05
Phenyldimethylsilane <0.004Phenylacetonitrile 0.11
62 S.AL-MALAIKA

by measuring ft3 for the oxidation of a number of aralkyl hydrocarbons, e.g.


a-substituted toluene, at 90°C, and at a constant initiator concentration, see
Table 4.
Russell found that powerful electron-withdrawing groups (e.g. cyano,
benzoyloxy, carbomethoxy) and weak electron-withdrawing groups (e.g.
chloro) have little effect on the reactivity of the remaining a-hydrogens,
hence resonance stabilisation of the formed radical cannot alone influence
the rate of abstraction reaction (Reaction (3), Scheme 1). It was concluded
that dipolar contribution (see VIII) to the transition state of the reaction
must offset the increase in reactivity which is expected from resonance
stabilisation of the radical formed.

ROOô H Rô+
VIII

Howard and Korchek [73] have measured absolute reactivities of some


a-substituted toluenes towards their own peroxyl radicals, AJ3, and towards
f-butyl peroxyl radical, A25b (Table 5) and found that the former depended
TABLE 5

Absolute reactivities of some a-substituted toluenes toward their own peroxyl radical, ks,
and toward the Mmtyl peroxy radical, &25b, at 30° [73]

Substrate ks per a-H &25b per a-H &3/&25b


(AT1 s"1) (M-1 s"1)

Benzaldehyde 33000 0.85 «40000


Benzyl ether 7.5(a) 0.3 25
Benzyl £-butyl ether 5.75 0.55 10.5
Benzyl alcohol 2.4 0.065 37
Benzyl ketone 0.82 0.045 18
Benzyl cyanide 1.56 0.01 156
Benzyl chloride 1.50 0.008 190
Benzyl benzoate 2.6 0.0085 306
Benzyl acetate 2.3 0.0075 307
Benzyl phenyl ether 0.75 0.1 7.5
£-Butyl phenyl acetate 0.8 0.03 27
Ethylbenzene 0.65 0.1 6.5
Benzyl bromide 0.6 0.006 100
Toluene 0.08 0.009 9

Estimated intermolecular rate constant.


AUTOXIDATION 63

CH3C(0)0/
c
" / V
ABr

PhCH20/C<°>0R
CH* ,OH / • PhCH2 C(O)
Ph#
Vt-BuO
#
• ""/ P h
C H , / °
>&Ph,CH3

fc"3>3 , , ,1
-0.2 0 0.2 0.4 0.6

Fig. 5. Relationship between log (&3/&25b) and am values of the a-substituents of substrates
shown in Table 5

on the nature of the attacking peroxyl radical and on the nature of the
substrate; the ratio k^/k25\) gave a measure of reactivity of a-substituted
peroxyl radical relative to the £-butyl peroxyl radical. A reasonably good
correlation was obtained between log (k^/k25\) and a m (Fig. 5); this suggests
that the reactivity of a peroxyl radical increases with increasing the electron
withdrawing capacity of the a-substituent.
Acylperoxyl radicals were found [47,74] to be much more reactive than
alkylperoxyl radicals (e.g. benzoyl peroxyl radical is several hundred times
more reactive than tetralyl peroxyl radical), hence the ease of oxidation of
aldehydes is due to the high reactivity of their acylperoxyl radicals. Differ­
ences in reactivities of aldehydes towards the same peroxyl radical is
affected by the inductive effect of the R group attached to the aldehyde
function; increasing the electron donating capacity increases the rate con­
stant, &3, for the hydrogen abstraction reaction [74,75].
Bolland [66] attempted to correlate E% with the exothermicities, AH3, for
the hydrogen abstraction reaction, Reaction (3), using the Polanyi relation,
Eqn (26a), and had calculated the overall activation energies, Ec, for the
oxidation of olefins which have different bond strengths by means of Eqn
(26b) (which follows from Eqn (7)). The assumptions made by Bolland were
that the activation energy for chain initiation, E-v is constant, and the
activation energy for chain termination, JS76, is zero. Korech and co-workers
[68] have argued that Bollands assumptions [66] for calculating propagation
activation energies, E%, (for Reaction (3), Scheme 1) are not valid and that
the value of a = 0.39 obtained by Bolland based on these calculations are
only approximate. Korchek et al [68], on the other hand, have correlated
absolute rate constants for hydrogen abstraction from RH by both tertiary
and secondary peroxyl radicals with the strengths of the C-H bond of the
64 S. AL-MALAIKA

substrate, D [R-H], and found a straight line relation between activation


energies and exothermicities, AH. Two empirical relations, at 30°C, were
derived for tertiary peroxyl radical (Eqn (27a)) and for secondary alkyl-
peroxyl radical (Eqn (27b)) and were found to hold well for hydrocarbons and
reasonably well for substrates with hetero atoms.

Es = C + oAH3 (26a)

E3 = EC-1/2E{ + 1/2 E6 (26b)

log k3 ,_ R00 = 15.4 - 0.2 D [RH] (27a)

tog h S . R 0 0 = 16.4 - 0.2 D RH] (27b)

The effect of structure on the rate constant for addition of peroxyl radical
to a double bond, A3a, depends mainly on the stability of the resulting
ß-peroxy alkyl radical with minor contribution from polar and steric effects.
It is clear from Table 6 [63,64] that structural changes in the olefin which
lead to an increase in the stability of this radical cause an increase in the
extent of addition reaction at the expense of hydrogen-abstraction. Thus
addition is favoured over abstraction of an allylic hydrogen when the double
bond is conjugated with, for example, an aromatic, vinyl, nitrile or carbonyl
group. Thus, for example in the case of the three isomeric alkenyl benzenes
IX, X, XI, the propagation reaction of the conjugated structures, IX and X,
occur mainly via an addition reaction (Reaction (3a), Scheme 1) giving good
yield of polyperoxides, in spite of the presence of allylic méthylène groups,
while XI gives exclusively hydroperoxide, via Reaction (3), during the early
stages of the reaction [2,6].

CH 3
I
C=CH2 CH=CHCH3 CH2CH=CH2

6 à (IX) (X)
à (XI)

Terminal olefins, e.g. di-isobutene, which lack structural features to


stabilise derived ß-peroxy alkyl radicals oxidise mainly to an epoxide and a
ketone; decomposition of ß-peroxy alkyl radical, Reaction (28), is much
faster than scavenging of oxygen [50,62,76]. Olefins which oxidise to give
mainly hydroperoxide or polyperoxide do give some epoxide when oxidised
at higher temperatures or low oxygen pressures, e.g. to obtain higher yield
AUTOXIDATION 66

of styrène oxide, oxidation of styrène is carried out at low pressure (1 torr)


or at high temperature (100°C) [77]. Simmons and Van Sickle [78] have
shown that the co-oxidation of olefin and aldehyde produces much higher
yield of epoxides than in autoxidation of olefin alone, since acyl peroxyl
radicals are more selective than alkylperoxyl radicals in favouring addition
rather than abstraction.

II II
ROO + C=C > ROO-C-C- > RO + -C-C- (28)

TABLE 6

Addition and abstraction mechanisms in autoxidation of alkenes

Hydrocarbon Temp. (°C) Mechanism


Abstraction (%) Addition (%)

Ethylene 110 0 100


Propylene 110 50 50
1-Butene 70 64 26
trans-2-Butene 70 33 62
Isobutene 80 14 81
3-Methyl-l-butene 70 90 6
Trimethylethylene 60 39 52
Tetramethylethylene 50 54 42
2-Methyl-l-pentene 70 35 45
1-Hexene 90 48 33
fertf-Butylethylene 90 2 100
Neopentylethylene 90 23 54
Cyclopentene 50 86 11.2
Cyclohexene 60 88 4.4
Cycloheptene 60 63 18.3
Cyclooctene 70 30 71
Cyclododecene 70 63
Methylenecyclohexane 70 58 39.5
Vinylcyclohexane 70 74 8.3
66 S. AL-MALAIKA

2.4 Termination and Effect of Oxygen Pressure

The rate of oxidation of most organic compounds is generally independent


of oxygen pressure; the termination step in autoxidation occurs almost
exclusively by recombination reaction of alkylperoxyl radicals, Reaction (6),
Scheme 1. This is mainly a consequence of differences in rates for Reactions
(2) and (3), where k2 > > k%. Under normal oxygen pressures and moderate
temperatures, therefore, Reaction (6) predominates, while Reactions (4) and
(5) can be ignored, and the rate of oxidation can be given by Eqn (7).
However, at very low oxygen pressures or when the alkyl radical is very
stable toward oxygen, the chain termination step (Reaction (4), Scheme 1)
predominates and autoxidation becomes a function of oxygen concentration
as shown by Eqn (29). However, in the case of intermediate oxygen pres­
sures where all three termination steps (Reactions (4)-(6), Scheme 1) can
contribute, the general form of the kinetic equation becomes more compli­
cated (see Eqn (30)). Bateman and Morris [79] calculated the relative
contributions of the three termination steps at different oxygen pressures
(0-760 torr) for different olefins and found it to depend very much on the
substrate structure.

= k2[R'][02] = k2kf1/2 i?i[0 2 ] (29)


d*

-d[0 2 ] = /^[RHHO^igj172
d
* " (k%k6[02]2 + /e2£3/e5[RH][02] + A§/e4[RH]2)1/2

The contribution of the three termination steps (Reactions (4)-(6),


Scheme 1) for typical olefins [79] was discussed earlier. It is clear from
Chapter 1, Fig. 1 that crossed termination becomes more important with
increasing substrate oxidisability. Similarly, Fig. 2 (Chapter 1) shows [79]
that whereas the unreactive olefin, hexadec-1-ene, oxidises at a rate which
is independent of oxygen pressure above 1 torr, increasing the reactivity of
the substrate, e.g. with the reactive olefin, 2,6-dimethylhepta-2,5-diene
(which gives a resonance stabilised free radical), shows that the oxidation
becomes pressure dependent to beyond 800 torr. Cross termination reac­
tions, therefore, occur mainly under conditions where hydrogen abstraction
(Reaction (3)) leads to a stabilised radical which has much lower reactivity
towards molecular oxygen (hence rate of Reaction (2) is reduced) with the
result that the alkenyl concentration approaches that of the alkylperoxyl.
Howard, Ingold and coworkers [7,80-83] have examined in detail the
mechanism involved in self-termination reactions of peroxyl radicals and
have measured the rate constant for this reaction (k$ for a large number of
AUTOXIDATION 67

hydrocarbons. They have shown [80,81] that values of k$ increases in the


order, tertiary peroxyl < secondary peroxyl < primary peroxyl (&6 for pri­
mary and secondary peroxyl radicals are about 10 higher than those for
tertiary radicals). Tertiary peroxyl radicals are believed [70,83] to exist in
equilibrium with tetroxide, Reaction (31a), which decomposes irreversibly
to give two caged alkoxyl radicals and oxygen, Reaction (31b). The majority
of the alkoxyl radicals escape from the cage (small amounts recombine to
give peroxide) and undergo hydrogen abstraction and ß-scission to give
primary or secondary radicals, Reactions (31c,d), which react with oxygen
followed by reaction either with the substrate resulting in further propaga­
tion, or with another tertiary peroxyl radical in a termination step, Reaction
(31e). The low rate constant for termination of tertiary peroxyl radicals is
due to the higher activation energy required for the decomposition of the
corresponding tetroxide intermediate to alkoxyl radical and oxygen [84].

2J-ROO. ^ É-ROOOOR-* (31a)

► t-ROOR-t

(31b)

► 2*-RO

f-RO + RH ► *-ROH + R- (31c)


ß-scission
if *-RO = Cumyloxyl; *-RO ► Acetophenone + CH3- (31d)
*-RO.
CH3 + O2 ► CH3OO ► Molecular products (31e)

In the case of primary and secondary peroxyl radicals, Russell [85]


suggested the involvment of a cyclic transition state in which one of the
a-hydrogen atoms is transferred to give ketone, alcohol and oxygen, Reac­
tion (32). This mechanism, excludes the intermediacy of alkoxyl radicals.
Evidence was based on experimental observation which indicates the forma­
tion of alcohol in high yields, the absence of peroxide, and the observation of
an isotope effect on A, later studies [39,56,57] have supported this mecha­
nism.
It was shown earlier that the work of Twigg and Mayo showed that the
oxidation of some olefins lead to a non-terminating alternative to Reactions
(5) and (6) (see Scheme 1) via addition of the peroxyl radical to the double
bond to give epoxide and alkoxyl radical, see Reaction (25). Mayo and
co-workers [50,88] studied further the relationship between epoxide forma­
tion and oxygen pressure during AIBN-catalysed oxidation of styrene and
68 S.AL-MALAIKA

2-Ç-00-^^^| I I
H
I H" I (32)
rl
> > C = 0 + 0 2 + > C—OH
I
H

methylstyrene at 50°C. In the case of styrene, the rate of oxidation is almost


indépendant of oxygen pressure at > 20 torr and termination occurs mainly
by Reaction (6). Decreasing the pressure leads to cross termination via
Reaction (5) while the complete absence of oxygen gives rise to the conven­
tional vinyl polymerisation chain terminating step, Reaction (4). In contrast,
the rate of oxidation of methylstyrene is oxygen dependent even at atmos­
pheric pressure and termination step 5 predominates even at pressure as
high as 25 torr (compared with 1 torr in the case of styrene). This is due to
the higher stability of the co-polymer free radical XII (which is obtained
under all conditions since methylstyrene does not homopolymerise) leading
to high rate of epoxide formation.

<fH;3 1 <fH;
V^ß
CHo-COO -CH2-C-
Ph i Ph

XII

3. TECHNOLOGICAL EFFECTS OF OXIDATION

Autoxidation leads to chemical and physical changes in both natural


substrates (e.g. lipids) and technological substrates (e. g. oils, fats, and high
polymers), the extent and nature of which varies depending on the structure
of the hydrocarbon involved. Unwanted degradation of lipids, oils, and
polymers, the intensification of environmental pollution, and the conversion
of some hydrocarbons to useful industrial intermediates, are all manifesta­
tions of oxidative processess. The information obtained from fundamental
studies of autoxidation can be used to explain the technologically important
processess which take place during exposure of such substrates to the
environment (e.g. effect of heat, light, stresses). In the case of polymers,
oxidative processess which occur at high temperatures and shear rates
AUTOXIDATION 69

involved in melt processing of the polymer are primarily responsible for the
deterioration in mechanical and electrical properties of polymer articles.
These phenomena will be discussed in more detail in Volume II.
Oxidation of liquid hydrocarbons, in the absence or presence of stabil­
isers, is often examined because of its relationship to polymer oxidation (e.g.
saturated hydrocarbon oxidation is related to that of polyolefins). Decom­
position of hydroperoxides (the main product of autoxidation) leads to the
formation of various oxidation products such as alcohols, carbonyl com­
pounds, e.g. aldehydes and ketones, acids and esters. Oxidation products
themselves may have a profound effect on the rate of oxidation, e.g. acids
formed could destroy the hydroperoxides, thereby reducing the oxidation
rate [89-91].
Examination of the nature of the products and kinetics of oxidation is,
therefore, quite important. Different types of testing methods are used in
measurements of oxidative effects in organic substrates which lead to better
understanding of the nature, kinetics and thermodynamics of the oxidation
process and products formed from it. Furthermore, the role of the structure
of the polymer repeat unit (which may be studied through an appropriate
choice of similar structure of a liquid hydrocarbon) on the oxidation process
and the changes in molar mass of polymers as function of extent of degrada­
tion can also be investigated. This section will outline the main principles of
some of these techniques which are used particularly for polymers, and some
of which are used for hydrocarbon model compounds and oils, in order to
provide kinetic, thermal and chemical information relavent to the oxida-
tive/degradative processess involved.
Quantitative descriptions of oxidation processess require absolute rate
coefficients for all important elementary steps and methods used for the
determination of these basic kinetic parameters which are fundamental to
the understanding of autoxidation mechanism in liquid hydrocarbons and
polymers have been extensively covered in monographs and reviews [7,14,
92,93] and will not be covered in detail here.

3.1 Oxygen Absorption

Oxygen absorption is one of the simplest and most widely used methods
for studying the mechanism and kinetics of oxidation of low molecular mass
pure compounds and high molecular mass polymers. The utilisation of the
oxygen absorption technique, especially if used in combination with other
testing methods, in polymers (e.g. rubbers, plastics and rubber modified
plastics) provide valuable information on kinetics of reactions of these
polymers with oxygen and of product formation. Furthermore, there is
generally a good correlation between the amount of oxygen absorbed in
these polymers and the deterioration of their useful mechanical properties
70 S. AL-MALAIKA

[94-97] and hence is a useful method to measure substrate oxidisabilities.


Methods used for oxygen absorption allow the measurement of oxygen
uptake as a change in volume at constant pressure and temperature. A
typical oxygen absorption curve is shown in Fig. 6: oxidation of a hydrocar­
bon substrate is characterised by an initial autoaccelerating period followed
by a sharp increase in the rate of oxygen uptake. A relationship between
hydroperoxide concentration (the primary products of autoxidation) and the
rate of oxidation is shown in Fig. 1. Oxygen uptake can either be measured
by manometric method [96] or, by a more sensitive system based on differ­
ential pressure transducers (± 0.5 p.s.i. range) as shown in Fig. 6 [98,99].
The transducer is connected to the sample vessel and a ballest vessel of the
same dimention. As the sample absorbs oxygen, a pressure difference
develops between the two vessels, which is converted by the pressure
transducer into a proportional output voltage and fed into a recorder.

SAMPLE
VESSEL

Fig. 6. Schematic representation of oxygen absorption apparatus.

3.2 Chemiluminescence

Chemiluminescence is the emission of low levels of light during the course


of some chemical reactions. The first reported observation of a weak emis­
sion of light from oxidation of polymer (e.g. PP) was in the early sixties in a
process called "oxyluminescence" [100,101]. Thermoluminescence relates to
the emission of light from a sample during heating. Thermoluminescence
analysis which is concerned with measuring the energy of emitted light as
a function of temperature has also been carried out on polymers (e.g. PE)
[102].
The advantage of chemiluminescence technique in studies of polymers
lies in its high sensitivity (extremely low levels of light, 10~10 to 10 lumens,
AUTOXIDATION 71

can be measured accurately) which makes it possible to determine various


kinetic parameters during the early stages of polymer oxidation [103-111].
However, the main drawback is that interpretation of the results depends
on the assumptions made of the mechanism of the luminescence processes
observed [111]. Although a number of mechanisms have been proposed, they
all agree in that the luminescence is due to emission of light during elec­
tronic transition from an excited state (mainly triplet state of a carbonyl
group formed during polymer oxidation) to the ground state in an exother­
mic reaction. The main mechanisms which have been proposed as the main
cause for chemiluminescence are: (i) termination of alkylperoxyl radicals
(primary or secondary) which lead to the formation of an excited ketone
(>C=0*) [101,104], see Reaction (33); (ii) decomposition of polymer hydro-
peroxide to ketone [112,113], see Reaction (34); and (iii) disproportionation
of two alkoxyl radicals to give an excited ketone [114].

2ROO. [ROOOOR] >C=0* + ROH + 3 0 1 (33)

ROOH >C=0* + ROH (34)

Both chemiluminescence and thermoluminescence can be measured using


the same apparatus shown schematically in Fig. 7.

M Fast Photon
PI Counter

Low Noise 1f
Photomultiplier
1 s\ 1
Filter \ — ► Amplifier H Pen
Recorder
ài
Polymer
Sample v
\. M
Vent+= Temperature 1
^ Programmer 1
and
L-^ Gas Controller 1
?—
* \-^ Inlet
/
Heating Block

Fig. 7. Schematic diagram of an apparatus used for measuring chemiluminescence and/or


thermoluminescence (for fuller description see Ref. [103]).
72 S.AL-MALAIKA

Pet
— 8 h
TJ

4 r-
DC
c
o
Oh
1^
TI T2 i\
T3 (0
0)
t£ 200 600
1 ! Temperature,
! i VAT Temperature (°C)

Fig. 8. (a) Typical thermogravimetric curve (A) and its derivative (B). (b) Thermo-
gravimetric curves for different polymers [126].

3.3 Thermal Methods of Analysis

(a) Thermogravimetry (TG)


Thermogravimetry is a measure of weight changes of substance as a
function of time at a constant temperature (isothermal TG) or as a function
of temperature at a linear rate of heating (dynamic TG); both methods have
been employed in studies of oxidation, kinetics, and mechanisms of thermal
degradation of polymers [115-121].
The variables measured in TG analysis are weight, time and tempera­
ture; precise measurements of these quantities are essential for obtaining
reliable TG data for use in evaluating kinetic parameters. Figure 8a shows
a typical thermogram where the weight loss is plotted against temperature.
Usually, an initial small weight loss (WQ-WJ is observed due to desorption
of residual solvent or water in the sample. This is followed by a second
weight loss step {Wy-W^} and possibly a third one (W2-W£y due to thermal
decomposition of the sample. Derivative thermogravimetry, DTG, is the
change in weight with respect to time, dw/dt, obtained by differentiation of
the TG signal (electronically or mathematically) and is useful for defining
precisely the quantities measured. The area under the TG curve is propor­
tional to the total change in the sample weight.
Figure 9 shows a schematic representation of a thermobalance in a
vertical configuration. The apparatus generally comprises a sample con­
tainer placed in a controlled environment under accurately regulated temp­
erature regime. The use of controlled atmospheres, e.g, nitrogen, oxygen, are
essential for thermal decomposition and oxidation studies. Changes in
sample weight are monitored by a transducer which produces an electrical
output (associated with the weight change) which is amplified and recorded
against temperature or time. Modern TG instruments are computer-control­
led and have substantial data storage, retrieval and analysis facilities.
Sample size, temperature and heating rates can all be varied within the
range, 0.5 to 500 mg, up to 1200°C and heating rate up to 100°C min" 1 .
AUTOXIDATION 73

-Spring

Transducer Armature

Demodulator

I Atmosphere
-W Control

Furnace _ Sample
Power Temp. ' Crucible
Supply Programmer
& Control

Temp
Sensor

Temperature
Measurement I

Fig. 9. Schematic representation of a thermal balance in a vertical position.

Many methods have been documented in the literature describing differ­


ent approaches to obtaining kinetic parameters (e.g. pre-exponential factor,
order of the reaction, activation energy) from dynamic TG data. The deter­
mination of kinetic parameters and the elucidation of mechanisms of the
complex polymer degradative processess from these measurements is con­
troversial [119,122-125]. The accuracy and reproducibility of TG data de­
pends on several factors including, physical characteristics of the sample
(size and shape), shape of sample crucible, the rate of heating (e.g. a high
rate may cause two stage decomposition to appear as one), arrangement of
the temperature-measuring elements (in some instruments the thermo­
couple is not in contact with the sample), the surrounding atmosphere and
the prior thermal treatment of the sample. Furthermore, great care, has to
be taken when interpreting results from TG analysis as many polymer
articles may lose their useful properties without necessarily losing weight.
However, qualitative comparison of characteristic TG curves for different
polymers or products is a useful approach which offer a visual indication of
the order of their stability, see Fig. 8b [126].

(b) Differential thermal analysis (DTA) and differential scanning


calorimetry (DSC)
DTA is a thermal method which measures the temperature difference
(AT) between the sample and a stable reference material against either time
or temperature as the two are subjected to change in temperature at a
controlled rate under identical thermal environments. Factors which affect
the accuracy of TG discussed in the previous section are also applicable to
74 S. AL-MALAIKA

DTA. Simultaneous measurements of DTA and TG allows simultaneous


measurements of both thermal events and weight changes which occur
during isothermal or dynamic measurements for precisely the same sample.
This method suffers from lower sensitivity and is not generally favoured
because of the inherent problems of non-uniformity of temperature environ­
ment caused by the physical displacement of the balance beam during
weight changes causing great baseline drift in DTA signal and reduced
sensitivity.
DSC records the energy required to reach zero temperature difference
between sample and reference against either time or temperature as the two
materials are subjected to identical temperature regimes under controlled
rates of heating and cooling.
DSC and DTA are, therefore, used to measure heat effects associated with
physical changes, e.g. glass transition temperature (Tg) in amorphous poly­
mers crystalline melting point (T m ), temperature of maximum decomposi­
tion rate (7^), and chemical changes which may be either exothermic or
endothermic, e.g. oxidation, decomposition, polymerisation. Direct measure­
ments of heats of fusion, crystallisation, vaporisation, specific heats, activa­
tion energies, and entropies of transition, as well as the effect of sample
preparation, treatments (e.g. annealing), molecular weight distribution and
composition on these measurements, can all be determined. Under isother­
mal conditions, the induction time to the onset of exothermic or endothermic
processes may also be monitored.
A very important difference between DSC and DTA is in the design of the
heating system. In case of DSC, the sample and reference are heated
separately, see Fig. 10, and both are maintained under identical tempera­
ture by electronically controlling the rate at which the heat is electrically
supplied to them. The peak area of a DSC curve is a true quantitative
measure of the electrical energy input which is required to maintain equal
temperatures in the sample and reference, and, unlike DTA, is independent
of the thermal constants of the instrument or changes in the thermal
behaviour of the sample. However, in the case of DTA, both sample and
reference materials are heated by the same source, see Fig. 10. Since, heat
capacities and thermal conductivities of the sample and reference material
are seldom identical a shift in the baseline of a DTA curve is normally
observed even in the absence of chemical or physical change: see Fig. 10,
hence peak areas of DTA curves may not be used for quantitative evaluation
of kinetic parameters [125,127,128].
In addition to the sample cells the instruments must have a heating-cool­
ing system which typically operates between -170° and 500°C (although
DTA machines cover much wider temperature range than DSC), a tempera­
ture controller and programmer, an atmosphere control system, a signal
amplifier and a recorder. The solid sample (powder, crystals, thin films) size
used in both techniques is small (ca. 20-50 mg for DTA and 0.5-10 mg for
AUTOXIDATION 75

■AT

i m 1 wa\
-4-^AAAAAAA/M-

t
^^ArfW*WW»M*
H—WNAAAAAAH-

-î-
Single Heat Source Single Heat Source
-HWWNA-k T
7HVWWM-
Individual Heaters
(i) Classical DTA (ii) "Boersma" DTA (iii) DSC

Oxidation or
Crystslization Cross-linking
TEXO

AT Heat Capacity

lEndo
First order Thermal Decomposition
Transition Melting or Volatilization
TIME OR TEMPERATURE

Fig. 10. (a) Schematic diagram of DTA and DSC. (b) Idealised DTA or DSC curves.

DSC). As with other thermal methods, the reliability of data obtained from
both DSC and DTA relies heavily on sampling techniques (size, shape,
packing), the sample thermal history, heating rate and thermocouple size
[129-131]. A very large number of reviews are available which discuss in
detail the various factors which affect DTA and DSC measurements and
their applications to studies of polymer oxidation and kinetics [125,129-
137].
These methods have also been used to predict polymer stability and to
assess the efficiency of different stabilisers [134,136,139] The speed and
simplicity in obtaining DTA and DSC curves and the availability of different
commercial instruments has led to the use of these techniques as a routine
laboratory procedure comparable to the measurements of infrared spectra;
it is used for quality control of polymer formulations in the wire and cable
industries [136,140]. However, the apparent simplicity of the technique
sometimes obscures the fact t h a t interpretation of DTA curves often
demands considerable skill and experience and results of extrapolation from
the high temperatures of measurements to the much lower service tempera-
76 S. ALrMALAIKA

tures must be treated with great caution. In the case of polymer stabilisa­
tion, for example, it was shown [134] that extrapolation from DTA data to
temperatures below the melting point of the polymer with the aim of
predicting polymer stability, or the efficiency of series of stabilisers, at room
temperature is not valid due to complications caused by the effect of crystal­
lisation and the possibility of the antioxidant of becoming less soluble (may
exist in a supersaturated state) at these lower temperatures.

3.4 Infrared Spectroscopy

Molecular vibrations and rotations of chemical groups, e.g. -OH, > C = 0 ,


>C=C<, - C H 2 - , show characteristic infrared spectra which can be used to
provide insight into chemical and molecular structure.
Analysis by infrared (IR) is a simple, fast, accurate, non-destructive
technique which requires a small amount of sample (gas, liquid or solid). It
is used to provide information on chemical structure (identification of differ­
ent chemical groups) and physical structure (e.g. crystallinity, configura­
tion, stereoregularity). The kinetics of liquid hydrocarbon model oxidations
as well as oxidative degradation of polymers have been investigated at high
temperatures and under different environments [5,95,141].
The rate of formation of oxidation products formed from polymer films
and hydrocarbon liquid models, especially carbonyl groups (e.g. carboxylic
acid, ketone, aldehyde, esters) and hydroxide species, have been calculated
using the base-line technique. Figure 11 shows changes in IR spectra of a
LDPE film during photo-oxidation. Thermal oxidation of polymer films, e.g.
PP, can become diffusion controlled above a certain thickness. Furthermore,
an internal absorption standard (an absorption peak which does not change
with thickness and oxidation) is normally used to correct for any variation
in thickness during oxidation.
The advent of rapid scanning Fourier-Transform infrared (FTIR) spectro­
meters (use of interferometer instead of monochromator) has promoted
further the application of infrared spectroscopy to studies of kinetics of
polymer oxidation, nature and proportions of products formed during ther­
mal and photodegradation and stabilisation of polymers. The main advan­
tages is that FTIR offers an increased energy through-put (about 100-200
times), very fast scanning (about 1 s per scan), increased accuracy of the
frequency scale and advanced computer-assisted data handling and pro­
cessing facilities. When separation of two components with the aim of
obtaining a 'pure structure' of one of the components is required, a difference
absorption spectra (i.e. subtracting out the contribution of one component
using the appropriate absorption bands) can be calculated by the comput­
erised spectrometer and hence is one of the simplest and most widely used
approaches.
MICRONS

100

4000 3500 3000 2500 2000 1800 1600 1400 1200 1000 800
WAVENUMBER (CM-1)

Fig. 11. Changes in infra red spectra of LDPE on exposure to UV light. Numbers in box are exposure time (hours) in a sunlamp/black-
lamp cabinet.
78 S. AL-MALAIKA

Recently, FTIR has been used, with other techniques, to identify and
quantify several key intermediates formed during 7-irradiated polyolefins
[143]. Very elegant work was carried out to follow some of the major
oxidation products, for example, peroxyl radicals and hydroperoxides were
reacted with gaseous NO at low temperatures (-78°C and -20°C, respec­
tively) for a long time (e.g. 2-3 h and >100 h, respectively) to give the
corresponding nitrates which could be identified by IR. In the same study,
the concentration of ketonic groups were quantitatively determined by IR
after reaction of all - O H species in the oxidised polymer with SF 4 . The
treatment with NO and SF 4 was shown to increase the sensitivity and
selectivity of this analysis.
The interfacing of FTIR with other instruments, e.g. GC, TG, pyrolysis-
GC, offers a better understanding of the oxidation process.

4. OTHER TECHNIQUES USED FOR OXIDATION STUDIES

Many other techniques have been used to monitor polymer oxidation and
degradation are beyond the scope of this chapter. These include, Fourier
Transform NMR for detection and quantification of degradation products,
thermal volatilisation analysis for volatile products of degradation, molecu­
lar weight determination by number of methods such as light scattering and
ultracentrifuge, viscometry, mass spectrometry and Chromatographie deter­
mination of products, or combination of two techniques to achieve better
analysis. These techniques have been covered in recent literature [144].

REFERENCES

1 J.L. Bolland, Q. Rev. (London), 3 (1949) 1.


2 L. Bateman, Q. Rev. (London), 8 (1954) 147.
3 J.L. Bolland and P. Ten Have, Trans. Faraday Soc, 43 (1947) 201.
4 D. Barnard, L. Bateman, J.I. Cuneen and J.F. Smith, in L. Bateman (Ed.), The
Chemistry and Physics of Rubber-like Substances, McClaren and Sons, London,
1963, Chapter 17.
5 G.A. Russell and R.C. Williamson, J. Am. Chem Soc, 86 (1964) 2357.
6 F.R. Mayo, A.A. Miller and G.A. Russell, J. Am. Chem. Soc, 80 (1958) 2501.
7 J.A. Howard, Adv. Free Radical Chem., 4 (1972) 49.
8 N.M. Emanuel, E.T. Denisov and Z.M. Maizus, in Liquid Phase Oxidation of
Hydrocarbons, English Translation by B.S. Hazzard, Plenum, New York, 1967.
9 F.R. Mayo, Proc Int. Oxid. Sympos., Adv. Chem. Ser., 75-77 (1968)
10 G. Scott, in Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965.
11 L. Reich and S. Stivala, in Autoxidation of Hydrocarbons and Polyolefins, Marcel
Dekker, New York, 1969.
12 H.H.G. Jellinek (Ed.), in Degradation and Stabilisation of Polymers, Elsevier,
Amsterdam, 1983.
AUTOXIDATION 79

13 C.H. Bamford and C.F.H. Tipper (Eds.), Comprehensive Chemical Kinetics, Vol. 14,
Degradation of Polymers, Elsevier, Amsterdam, 1975.
14 K.U. Ingold, in J.K. Kochi (Ed.), Free Radicals, Vol. 1, Wiley, New York, 1973.
15 J.A. Howard, in J.K. Kochi (Ed.), Free Radicals, Vol. 2, Wiley, New York, 1973,
Chapter 12.
16 J.L. Bolland and G. Gee, Trans. Faraday Soc, 42 (1946) 236.
17 G.A. Russell, J. Am. Chem. Soc, 78 (1956) 1041.
18 A.A. Miller and F.R. Mayo, J. Am. Chem. Soc, 86 (1964) 5709.
19 A. Bromberg and K.A. Muszket, J. Am. Chem. Soc, 91 (1969) 2860.
20 N.M. Emanuel, Proc 7th World Petrol. Congress, Vol. 5, Elsevier, Amsterdam,
1967, p. 3.
21 E.T. Denisov, Russ. J. Phys. Chem., 38 (1964) 1.
22 D.J. Carlsson and J.C. Robb, Trans. Faraday Soc, 62 (1966) 3403.
23 L. Dulog, Makromol. Chem., 77 (1964) 206.
24 L. Bateman, H. Hughes and A.L. Morris, Discuss. Faraday Soc, 14 (1953) 190.
25 L. Bateman and H. Hughes, J. Chem. Soc, (1952) 4594.
26 C. Walling and L. Heaton, J. Am. Chem. Soc, 87 (1965) 48.
27 N.V. Zolotova and E.T. Denisov, J. Polym. Sei., Part Al, 9 (1971) 3311.
28 R. Hiatt, J. Clipsham and T. Visser, Can. J. Chem., 42, (1964) 2754.
29 P.D. Bartlett and T.G. Traylor, J. Am. Chem. Soc, 85 (1963) 2407.
30 R. Hiatt, T. Mill, K.C. Irwin and J.K. Castleman, J. Org. Chem., 33 (1968) 1421.
31 J.R. Thomas, J. Phys. Chem., 63 (1959) 1027.
32 W.A. Pryor, Free Radicals, McGraw Hill, New York, 1966.
33 T. Koenig and H. Fischer, in J.K. Kochi (Ed), Free Radicals, Vol. 1, Wiley, New York,
1973.
34 F. Harber and J. Weiss, Proc. R. Soc. London, Ser. A, 147 (1934) 332.
35 A.J. Chalk and J.F. Smith, Trans Faraday Soc, 53 (1957) 1214.
36 R.A. Sheldon and J.K. Kochi, Oxid. Combust. Rev., 5 (1973) 150.
37 J.K. Kochi, Science, 155 (1967) 415; J.K. Kochi, J. Am. Chem. Soc, 85 (1963) 1958;
ibid, 84 (1962) 774; C.L. Jenkins and J.K. Kochi, J. Org. Chem., 36 (1971) 3095,
3103.
38 J.K. Kochi and H.E. Mains, J. Org. Chem., 30 (1965) 1862; J.K. Kochi and A. Bemis,
Tetrahedron, 24 (1968) 5099.
39 J.K. Kochi and P.M. Mocadlo, J. Org. Chem., 30 (1965) 1134.
40 R. Hiatt, K.C. Irvin, and C.W. Gould, J. Org. Chem., 33 (1968) 1430; R. Hiatt, T. Mill,
K.C. Irvin, and J. K. Castleman, ibid, 33 (1968) 1421; R. Hiatt, T. Mill and F.R.
Mayo, ibid, 33 (1968) 1416.
41 M.S. Kharasch, F.S. Arimoto and W. Nudenburg, J. Org. Chem., 16 (1951) 1556; 19
(1954) 1977; D.D. Coffman and E.L. Jenner, J. Am. Chem. Soc, 80 (1958) 2872.
42 M.S. Kharasch and A. Fono, J. Org. Chem., 24 (1959) 72; D.D. Coffman and H.N.
Cripps, J. Am. Chem. Soc, 80 (1950) 2877.
43 R.A. Sheldon and J.K. Kochi, J. Am. Chem. Soc, 90 (1968) 6688.
44 P.D. Barlett and P. Günther, J. Am. Chem. Soc, 88 (1966) 3288.
45 J.K. Kochi, Tetrahedron, 18 (1962) 483.
46 Y. Kamiya and K.U. Ingold, Can. J. Chem., 42 (1964) 1027.
46a C.E.H. Bawn and J.E. Jolley, Proc. R. Soc, A 237 (1956) 297.
47 Y. Kamiya and E. Niki, in H.H.G. Jellinek (Ed.), Aspects of Degradation and
Stabilisation of Polymers, Elsevier, Amsterdam, 1978, Chap. 3.
80 S. AL-MALAIKA

48 Y. Kamiya, Bull. Chem. Soc. Jap., 38 (1965) 2156.


49 Y. Kamiya and K.U. Ingold, Can. J. Chem., 42 (1964) 2424.
50 A.A. Miller and F.R. Mayo, J Am. Chem. Soc, 78 (1956) 1017.
51 K.U. Ingold, Ace. Chem. Res., 2 (1969) 1.
52 G.A. Russell, J. Chem. Educ, 36 (1959) 111.
53 F.F. Rust, J. Am. Chem. Soc, 79 (1957) 4000.
54 D.E. Van Sickle, J. Org. Chem. 37 (1972) 755.
55 D.E. Van Sickle, T. Mill, F.R. Mayo, H. Richardson and C. Gould, J. Org. Chem., 38
(1973) 4435.
56 L. Dulog, E. Radlman and W. Kern, Makromol. Chem., 80 (1964) 67.
57 J.C.W. Chien, E.J. Vandenberg and H. Jabloner, J. Polym. Sei., Part A-16, (1968)
381,393,.
58 C. Decker and F.R. Mayo, J. Polym. Sei., Polym. Chem. Ed., 11 (1973) 2847.
59 E. Niki and Y. Kamiya, Bull. Chem. Soc. Jap., 48 (1975) 3226.
60 L. Dulog and K. H. David, Makromol. Chem., 145 (1971) 67.
61 F.F. Rust and E.A. Youngman, J. Org. Chem., 27 (1962) 3778.
62a J.A. Howard and K.U. Ingold, Can. J. Chem., 48 (1970) 873.
62b F.R. Mayo, J. Am. Chem. Soc, 80 (1958) 2497.
63 D.E. Van Sickle, F.R. Mayo and R.M. Arluck, J. Am. Chem. Soc, 89 (1967) 4824,
4832.
64 D.E. Van Sickle, F.R. Mayo, R.M. Arluck, and M.G. Syz, J. Am. Chem. Soc, 89 (1967)
967.
65 J.A. Howard and K.U. Ingold, Can. J. Chem., 44 (1966) 1113.
66 J.L. Bolland, Trans. Faraday Soc, 46 (1950) 358.
67 K.R. Hargrave and A.L. Morris, Trans. Faraday Soc, 52 (1956) 89.
68 S. Korcek, J.H.B. Chenier, J.A. Howard and K.U. Ingold, Can. J. Chem., 50 (1972)
2285.
69 J.A. Howard, W.J. Schwalm and K.U. Ingold, Adv. Chem. Ser., 75 (1968) 6.
70 J.A. Howard, K.U. Ingold and M.S. Symonds, Can. J. Chem., 46 (1968) 1017.
71 J.A. Howard, K.U. Ingold, Can. J. Chem., 41 (1963) 2800.
72 G.A. Russell, J. Am. Chem. Soc, 78 (1956) 1047.
73 J.A. Howard and S. Korcek, Can. J. Chem., 48 (1970) 2165.
74 G.E. Zaikov, J.A. Howard and K.U. Ingold, Can. J. Chem., 47 (1969) 3017.
75 C. Walling and E.A. McElhill, J. Am. Chem. Soc, 73, (1951) 2927.
76 G.H. Twigg, Chem. Eng. Sei., Suppl. 3 (1954) 5.
77 F.R. Mayo, Ace Chem. Res., 1 (1968) 193.
78 K.E. Simmons and D.E. Van Sickle, J. Am. Chem. Soc, 95 (1973) 7759.
79 L. Bateman and A.L. Morris, Trans. Faraday Soc, 49 (1953) 1026.
80 J.A. Howard and K.U. Ingold, Can. J. Chem., 43 (1965) 2729.
81 J.A. Howard and K.U. Ingold, Can. J. Chem., 45 (1967) 783, 789.
82 J.A. Howard, K. Adamic, and K.U. Ingold, Can. J. Chem., 47 (1969) 3793, 3803.
83 J.K. Kochi and P.M. Mocadlo, J. Org. Chem., 30 (1965) 1134, and references therein.
84 J.R. Thomas and K.U. Ingold, Adv. Chem. Ser., 75 (1968) 258.
85 G.A. Russell, J. Am. Chem. Soc, 79 (1957) 3871.
86 D.F. Bowman, T. Gillan and K.U. Ingold, J. Am. Chem. Soc, 93 (1971) 6555.
87 R. Hiatt and L. Zigmund, Can. J. Chem., 48 (1970) 3967.
88 F.R. Mayo, A.A. Miller, J. Am. Chem. Soc, 80 (1958) 2480.
89 V.M. YurW, A.N. Pravednikov, and S.S. Medvedev, J. Polym. Sei., 55 (1961) 353.
ÀUTOXIDATION 81

90 JHoldsworth, G.Scott and D.Williams, J. Chem. Soc, 906 (1964) 4692.


91 S. Al-Malaika and G. Scott, Polymer, 23 (1982) 1711.
92 T. Mill and D.G. Hendry, in C.H. Bamford and C.F.H. Tipper (Eds.), Comprehensive
Chemical Kinetics, Vol. 16, Elsevier, New York, 1980, Chapter 1.
93 N.M. Emanuel, G.E. Zaikov and Z.K. Maizus, in Oxidation of Organic Compounds,
Transi. A.K. Henn and I.G. Evans, Pergamon Press, Oxford, 1984, Chapter 2.
94 J.R. Shelton and H. Winn, Ind. Eng. Chem., 39 (1947) 1133.
95 J.R. Shelton, F.J. Wherely and W.L. Cox, Ind. Eng. Chem,, 45 (1953) 2080.
96 W.L. Hawkins, R.H. Hansen, W. Matryek and F.H. Winslow, J. App. Polym. Sei.,
(1959) 137.
97 A. Ghaffer, A. Scott and G. Scott, Eur. Polym. J., 11 (1975) 271.
98 F.A.A. Ingham, G. Scott and J.E. Stuckey, Eur. Polym. J., 11 (1975) 783.
99 S. Al-Malaika, M. Coker, G. Scott and P. Smith, J App. Polym. Sei., in press
100 G.E. Ashby, J. Polym. Sei., 50 (1961) 99.
101 M.P. Schard and C.A. Russell, J. App. Polym. Sei., 8 (1964) 985, 997.
102 A. Charlesby and R.H. Partridge, Proc. Royal Soc. Series A, 271 (1963) 170,188; 283
(1965) 312.
103 G.A. George, Polym. Deg. Stab., 1 (1979) 217.
104 G.A. George, in N. Grassie (Ed.), Developments in Polymer Degradation-3, Applied
Science Publishers, London, 1981, Chapter 6.
105 G.D. Mendenhall, Angew. Chem. Int. Ed. 16 (1977) 225.
106 K. Naito and T.K. Kwei, J. Polym. Sei., Polym. Chem. Ed., 17 (1979) 2935.
107 R.A. Lloyd, Trans Faraday Soc, 61 (1965) 2182.
108 Lev Zlatkevich, J. Polym. Sei. Polym. Physics Ed., 25 (1987) 2207.
109 Lev Zlatkevich, Polym. Deg. Stab., 19 (1987) 51.
110 N.C. Billingham, E.S. O'Keefe and E.T.H. Then, Proc. ACS Polym. Mat. Sei. Eng.,
58 (1988) 431.
111 N.C. Billingham and G.A. George, J. Polym. Sei., B: Polym. Phys., (1990) 257.
112 L. Reich and S.S. Stivala, Makromol. Chem., 103 (1967) 74.
113 J.A. Howard, in J.K. Kochi (Ed.), Free Radicals, Vol. 2, Wiley, New York, 1973,
Chapter 5.
114 E.M.Y. Quinga and G.D. Mendenhall, J. Am. Chem. Soc, 105 (1983) 6250.
115 N. Grassie and W.W. Kerr, Trans. Faraday Soc, 55 (1959) 1050.
116 S.L. Madorsky, in Thermal Degradation of Organic Polymers, Interscience, New
York, 1964, p. 21.
117 R. Simha and L.A. Wall, J. Polym. Sei., 6 (1951) 39.
118 R. Simha and L.A. Wall, J. Phys. Chem., 56 (1952) 707.
119 J.H. Flynn, in H.H.G. Jellinek (Ed.), Aspects of Degradation and Stabilisation of
Polymers, Elsevier, Amsterdam, 1978, Chapter 12.
120 H.H.G. Jellinek, J. Polym. Sei. Part A-l, 4 (1966) 2705.
121 J.H. Flynn, Polym. Eng. Sei., 20 (1980) 675.
122 E.S. Freeman and B. Carroll, J. Phys. Chem., 62 (1958) 394.
123 H.E. Kissinger, Anal. Chem., 29 (1957) 1702.
124 H.L. Friedman, J. Polym. Sei. Part C, 6 (1965) 183.
125 S.S. Stivala, J. Kimura and S.M. Gabbay, in N. Allen (Ed.), Degradation and
Stabilisation of Polyolefins, Applied Science Publishers, London and New York,
1983, Chapter 3.
126 J.Chiu, Appl. Polym. Sympos., 2 (1966) 25.
82 S. AL-MALAIKA

127 M.J. Richardson, in C. Booth and C. Price (Eds.), Comprehensive Polymer Science,
Vol. 1. Polymer Characterisation, Pergamon Press, Oxford, 1989.
128 Supplement Volume of Encyclopedia of Polymer Science and Engineering, Ed.
Mark, Bikales, Overberger and Menges, Wiley, 1989.
129 W.W. Wendlandt, in Thermal Methods of Analysis, Interscience, New York, 1974,
Chapters 5-7.
130 F.M. Barrell and J.F. Johnson, in P.E. Slade Jr. and L.T. Jenkins, Techniques and
Methods of Polymer Evaluation, Vol. 2, Marcel Dekker, New York, 1970, Chapter 1.
131 J.L. McNaughton and C.T. Mortimer, Int. Rev. Sei.: Phys. Chem. Series II, 10,1975,
p.l.
132 P.D. Garn, in Thermoanalytical Methods of Investigation, Academic Press Inc., New
York, 1965.
133 L. Reich and S.S. Stivala, in Elements of Polymer Degradation, McGraw Hill, New
York, 1971, Chapter 2.
134 N.C. Billingham, D.C. Bott and A.S. Manke, in N. Grassie (Ed.), Developments in
Polymer Degradation-3, Applied Science Publishers, London, 1981, Chapter 3.
135 C.B. Murphy, Anal. Chem., 44 (1972) 513R.
136 J.B. Howard, Polym. Eng. Sei., 13 (1973) 429.
137 E.L. Charsley and J.G. Dunn, J. Therm. Anal., 17 (1980) 535.
138 S.S. Stivala and S.M. Gabbay, Polymer, 18 (1977) 807.
139 D.E. Van Sickle and D.M. Pond, ACS Adv. Chem. Ser., 169 (1978) 237.
140 D.I. Marshall, E.J. George, J.M. Turnipseed and J.L. Gleen, Polym. Eng. Sei., 13
(1973) 415.
141 S.S. Stivala, L. Reich and RG.Kellehmery, Macromol. Chem., 59 (1963) 28.
142 N. Grassie and G. Scott, in Polymer Degradation and Stabilisation, Cambridge
University Press, 1985.
143 D.J. Carlsson, R. Brousseau, C. Zhang and D.M. Wiles, Am. Chem Soc. Symp. Ser.,
364 (1988) 376.
144 J. Rabek, in Experimental Methods in Polymer Chemistry, Wiley, New York, 1980.
83

Chapter 3

INITIATORS, PROOXID ANTS AND S E N S I T I S E R S


GERALD SCOTT

1. REACTIONS OF OXYGEN

In the previous chapters, hydroperoxides were seen to be the primary


products of autoxidation. This fact gives them a key position in the process
of autoxidation since without the continuous generation of hydroperoxides
the kinetic chain reactions would rapidly self-terminate by the radical
coupling reaction of the chain propagating species, alkyl and alkylperoxyl.
Hydroperoxides (which include hydrogen peroxide itself) are therefore the
most important initiators of the radical chain reaction. If hydroperoxides are
not initially present in a substrate which is known to be capable of autoxida­
tion then other initiators become significant.
Dioxygen itself in its ground (triplet) state is a "diradical" and it is
therefore potentially capable of hydrogen abstracting from a hydrocarbon to
give a hydroperoxyl radical and alkyl;

RH + 0 2 > R+ÔOH (1)

However, this reaction occurs only with very labile C-H bonds and when it
does occur it leads to the formation of a relatively stable alkyl radical which
terminates rather than propagates the oxidation chain reaction.
Ground state oxygen can also in principle add to conjugated double bonds
(Reaction (2)) to give new radical species.

02+ —CH=CH—CH = CH—CH=CH —

* (2)
•O — O —CH — C H — C H —CH — C H — C H —

Thus both Reactions (1) and (2) can lead to the formation of hydroperox­
ides by further reaction of the alkyl radical with oxygen (see below). How-
84 GERALDSCOTT

ever both are energetically unfavourable and occur only with highly acti­
vated substrates [1]. Much more common, particularly in biological systems,
is the single electron reduction of dioxygen by an electron donor present in
the system (Reaction (3)). The Superoxide radical anion (I) so formed has
been implicated in many biological autoxidations.

02 + e > 0-Ô - H - OOH (3)


! (Pka4.8)

Although Superoxide is both a radical and an ion and should formally be


depicted as in I, in practice the radical is generally omitted (particularly in
discussion of biological chemistry) and Superoxide is depicted as O2. This
convention will be followed here.
Superoxide itself is not highly reactive toward organic substrates [2]. In
its ionised form it is considerably less reactive than alkylperoxyl due to the
délocalisation of the unpaired electron on two oxygens in I. However, the
protonated form (<pH 4.8) is more reactive and the latter readily dispropor-
tionates or is reduced to hydrogen peroxide [3,4]. The latter is itself very
sensitive to reducing agents, notably transition metal ions [3];

+e
OÔ > H0O9 * e > OH + ÔH (4)
2 + 2 2
+H

The same chemistry is also involved in the photosensitising effects of


some dyestuffs and has been shown to be responsible for the "tendering" of
dyed cellulose fibres. The formation of hydrogen peroxide was clearly de­
monstrated by Egerton [5-8] in a classic experiment in which hydrogen
peroxide formed by photo-sensitisation of dyestuffs and white pigments (e.g.
Ti0 2 ) affected not only the dyed threads but also undyed threads spatially
separated from them. Hydrogen peroxide photolysis was shown to be in­
volved, particularly in a moist environment [9].
Other forms of oxygen, notably ozone, 0 3 and singlet oxygen ^AgO^ have
been frequently invoked as pro-oxidants. There is a good deal of evidence for
the former since ozone is a relatively stable and persistent component of the
atmosphere and generally gives rise to peroxidic species on reaction with
organic substrates [10]. The practical importance of singlet oxygen is much
more debatable however, since, being an excited state of dioxygen, it is
readily quenched by inorganic agents such as water and even ground state
oxygen. It is not therefore a normal component of the atmosphere and
although it may be formed under specific conditions in sensitised photo-ox­
idation, any general involvement in the initiation of autoxidation is very
much a matter of debate [11].
INITIATORS, PROOXIDANTS AND SENSITISERS 85

G*2p
O O O O o
**2p
© © © o © © 0 0 © ©
* 2p
0 0 0 0 0 0 0 0 0 0
o 2p
0 © © 0 0
a* 2s
© © 0 0 0
a 2s
0 0 0 © 0
o* 1s
0 0 0 © 0
a 1s
0 0 © ©
Ground State 0
0 Singlet 0 Superoxide Peroxide Ion Singlet 0

<V«> <V*> <°;> <°r> <*?>


Fig. 1. Electronic structures of dioxygens. (Reproduced with kind permission from Phil.
Trans. R. Soc., B311 (1985) 659).

Finally, some atmospheric pollutants have pro-oxidant activity, particu­


larly in the presence of UV light [12]. Notable among these are the oxides of
nitrogen, oxides of sulphur and possibly polycyclic aromatic hydrocarbons,
particularly in the form of soot. Photosensitisers (in addition to hydro-
peroxides) may also be formed in organic substrates as a result of oxidation.
The commonest examples are carbonyl species (aldehydes and ketones)
which are excited to highly reactive species (generally triplet states) by UV
light.
The following sections are concerned with a more detailed discussion of
the chemistry of pro-oxidants and sensitisers.
The electronic relationship between the different forms of dioxygen dis­
cussed above are outlined in Fig. 1 [2]. This shows that only ground state
dioxygen ( XgO^ and Superoxide, O2, can participate directly in radical
reactions. The remainder are the source of radicals through further reaction
(homolysis or redox reactions of derived hydroperoxides). Ozone (II) simi­
larly does not formally contain an unpaired electron; Ha and lib are the
main contributory structures [13].

6 . . 6 M
. . o
11 \_ <—> _/ w /\
0 0 0 0 0 0

lia lib lie

1.1 Initiation by Ground State Dioxygen

Oxygen is a very powerful oxidising agent. Its standard oxidation-reduc-


86 GERALD SCOTT

tion potential is +0.816 and is in principle able to oxidise a variety of organic


molecules with lower oxidation-reduction potentials. In practice, oxygen is
surprisingly inactive toward organic compounds due to the fact that it can
normally only react in single electron transfer steps and this is disfavoured
unless a catalyst is present; notably the transition metal ions. Thus many
thiols and catechol derivatives are oxidised directly by oxygen in the pre­
sence of transition metal ions and the rate at which this process occurs
appears to be dependent on the concentration of the metal ions (e.g. Cu2+,
Fe , Mn ) in the system [14,15]. Leuco méthylène blue (III) is readily
oxidised to méthylène blue by single electron transfer to oxygen [16],

Me,N S NMe, Me,N


III
+ 0,-+H+

This reaction is a model for the generation of Superoxide from the reduced
form of the flavin ring in xanthine oxidase (IV) [16].

H O
Œ3
CH,
NH _o^ yY N Y^f
CH (6)
I I
R H R
IV
+ 0,-+H+

Some antioxidants with low oxidation reduction potential can liberate


hydroperoxyl in non-aqueous substrates, particularly at elevated tempera­
tures. Thus, some arylamine antioxidants show an optimum concentration
effect [17] and the kinetics suggest that, at high concentrations, direct
attack of oxygen may occur;
Oo
ArNHAr' ArNAr' + OOH (7)
INITIATORS, PROOXIDANTS AND SENSITISERS

1 2 3
Time (h)

Fig. 2. Oxidative degradation of polybutadiene rubber in solution at 50°C in the presence


of phenyl hydrazine (PH) and iron naphthenate (Fe3+), added at point indicated by the
arrow. (A) PH; (B) PH + Fe3+ (10 mole %); (C) PH + Fe3+ (10 mole %) + PBN; (D) PH in
N2; (E) PH + Fe + in N2. (Reproduced with permission of Rubber Chem. Technol 32
(1959) 231).

Redox reactions of this type have been used to initiate the degradation of
rubbers and it has been shown that in the case of hydrazobenzene and
phenyl hydrazine, hydrogen peroxide is a major reaction product [18],

°2
PhNHNHPh > PhNNHPh + OOH (8)
Fe 3 + I
P h N = NPh + HOOH

The reaction occurs relatively slowly at 50°C in the absence of metal ions
but in the presence of ferric naphthenate, it occurs extremely rapidly. This
is illustrated for phenylhydrazine in Fig. 2.
It seems probable that similar reactions are involved in the action of a
variety of mercaptans used as chemical "plasticisers" in rubbers to reduce
their molecular weight [19]. Typical examples are V-VIL

Cl
V VI VII
88 GERALD SCOTT

There is little doubt that thiyl radicals are produced since disulphides are
generally byproducts in these reactions. However, the same products could
also arise by reaction of thiols with hydroperoxides (see Section 2.2.4).

RSH ° 2 > RS +ÔOH (9)


i
RSSR

Oxygen coordinated transition metal compounds have long been thought


to be capable of activating dioxygen to a reactive species that can initiate
autoxidation. Oxygen carriers such as haematin and cobalt disalicylidene-
ethylenediamine are effective pro-oxidants, presumably due to the forma­
tion of a peroxyl type species [20-22].
Although ground state oxygen does not normally react with hydrocarbons,
there is evidence that it can react directly with conjugated olefins to give a
peroxyl radical and an alkyl radical [23]. It is not easy to demonstrate this
unequivocally when the primary products of autoxidation are hydroperox­
ides since initiation by very small amounts of hydroperoxides make it very
difficult to detect the much slower direct initiation by oxygen itself. How­
ever, in the case of styrene at 60°C, hydroperoxides cannot be formed since
it has no méthylène group, and the rate of oxidation is therefore independent
of the amount of oxygen absorbed, unlike hydroperoxide initiated oxidation
in which the rate of oxidation is proportional to (0 2 abs) 1 . This rate, which
is given by Eqn (10), has been assumed to be the rate of oxygen initiation by
Reaction (11) [24].

^=(1.4.10-V/2 (10)
at
CH 2 =CHPh + 0 2 > -OOCH2CHPh (11)

Very oxidisable compounds must be expected to be pro-oxidants for more


stable substrates due to the facile formation of hydroperoxides. There is a
good deal of indirect evidence for the pro-oxidant action of polyunsaturated
compounds in multicomponent oxidising systems. It was shown by Holman
[25] for example that Vitamin A (VIII) and ß-carotene (IX) are preferentially
destroyed in admixture with methyl linoleate with concomitant rapid de-
colourisation and the rapid formation of methyl linoleate hydroperoxide (see
Fig. 3). Similar effects have been observed in other substrates [26,27], but
the evidence suggests that this phenomenon occurs primarily in the initia­
tion stage of autoxidation since Bolland [28] found that in initiated oxidation
the rate of oxidation of a mixture of hydrocarbons was a linear function of
INITIATORS, PROOXIDANTS AND SENSITISERS

CH,
,CH,

CH
a/V CH
*CH
CH
£c-
I
.CH
CH2OH
CH3
CH,
VIII

CH3
^i^CH 3

CH CH .CH
S ^ C^CH^CH-
CH C^ CH
CH, I I
CH, CH, J2

IX

0.02 0.06 0.10


Moles oxygen/mole ester
Fig. 3. Coupled oxidation of Vitamin A acetate and methyl linoleate. (A) Vitamin A acetate
UV absorbance at 328 nm, (B) methyl linolineate hydroperoxide absorbance at 235 nm.
(Reproduced with permission from Arch. Biochem., 26 (1950) 85.

the molar concentration of the more oxidisable component. In the above


initiation reaction direct oxygen attack almost certainly occurs at conju­
gated double bonds [24]. There is little evidence for oxygen reaction at the
saturated carbon atom on hydrocarbons at ambient temperatures but such
a mechanism has been proposed for the oxidation of acetaldehyde [29],
90 GERALD SCOTT

CH3CHO + O2 > C H 3 C = 0 + OOH (12)

and oxidation of alkanes to olefins at high temperatures (> 300°C) almost


certainly does involve the direct attack of oxygen on the hydrocarbon [30];
e.g.

CH3-CH3 ° 2 > CH3CH2+ ÔOH > CH2=CH2 + H202


02

C H 2 = CH 2 + ÔOH (13)

Other reactions of oxygen with hydrocarbons have been discussed in Chap­


ter 2.

1.2Initiation by Singlet Oxygen

Singlet oxygen (xAg02) is formed by energy transfer to ground state


oxygen from a variety of excited triplet states rSens*) [31]

hv 3
Sens > ^en* > Sens* (14)

3
Sens* + 3 0 2 (2g) > 1
0 2 *( 1 Ag) + ^ e n s (15)

This process is therefore a potentially important initiating mechanism in


photooxidation and a great deal of attention has been directed toward
establishing the importance of photosensitised oxidation in polymers and in
biological substrates in recent years [32].
Unlike ground state oxygen, singlet oxygen is not a radical species (see
Fig. 1) and it does not therefore hydrogen abstract from reactive méthylène
groups or add to double bonds as described in Reaction (11). It does however,
interact with double bonds either to form an endoperoxide reaction (16), a
dioxetane, Reaction (17), or an allylic hydroperoxide by means of an "ene"
reaction (18) in which the double bond is shifted. Endoperoxide formation
is extremely rapid when the double bond is surrounded by electron releasing
groups (k * 10 1 mol~ s~ ) and the activation energy for both endoperoxide
formation and for the "ene" reaction are almost zero [33].

lQ2
-» I Y II (16)
INITIATORS, PROOXIDANTS AND SENSITISERS 91

(17)

H -O
\
H2C O
•o, \
" CH 2 CH 2 CH2 CH2
(18)
/

H2C^ ^OOH

—CH2 CH2—

Dioxetane formation, by contrast, has a higher activation energy and


appears to occur only when the alternative reactions are not possible [34,35],
although the ratio of dioxetane to hydroperoxide is strongly solvent depend­
ant [36]. It was shown by Ng and Guillet [37] that, during the oxidation of
cis-polyisoprene with 1 0 2 generated by microwave discharge, no reduction
in molecular weight occurred during the period of hydroperoxide formation
in the dark at ambient temperature. When, however, the polymer was
exposed to UV irradiation (365 nm) rapid degradation occurred at 30°C with
a quantum efficiency that increased with hydroperoxide content, indicating
the potential importance of photosensitisation as a cause of polymer de­
gradation. Similar results were obtained by Rabek et al. [38], who used
several photosensitisers (anthracene, Rose Bengal and méthylène blue) for
the photooxidation of cis-l,4-polybutadiene in benzene solution, but these
authors were not able to separate the effects of singlet oxygen attack and
subsequent photolysis of hydroperoxides. In solid state rubbers cross-link­
ing occurred, a phenomenon characteristic of the oxidation of polymers with
pendant vinyl groups due to oxygen co-polymerisation (see Section 2.1.).
Singlet oxygen generated by non-photochemical means does not react
with polymers that do not contain ethylenic unsaturation. No hydroperox­
ides were detected in polyethylene under the conditions described above
[39]. Nevertheless, singlet oxygen remains a potentially important source of
photo-initiation in polymers containing unsaturation [40], or in biological
systems where a variety of pigments are capable of producing 1 0 2 in organs
of the body exposed to sunlight. For example, damage has been reported to
the lens of the eye [41] and a variety of skin diseases are probably in part
caused by singlet oxygen attack, sensitised by pigments [42].
92 GERALD SCOTT

kANX^Acl
CH 2 CH 2 CH 2 N(CH 3 )

Some constituents of cosmetics may also cause skin damage in the same
way [43]. It is not always clear however, how much radical damage is caused
by singlet oxygen and how much by other reactive species resulting from the
same basic photo-excitation process. Thus in the case of the drug chlorpro-
mazine (X), the photo-excited species appears to be capable of either elimi­
nating a chlorine atom in the absence of oxygen [44] or of giving singlet
oxygen when excess oxygen is present in the system [45].
Many photochemical processes give rise to free radicals as secondary
products but caution has to be exercised in interpreting the primary source
of these species.
An example from inorganic pigment technology illustrates the complexity
of the behaviour even of simple chemicals. Titanium dioxide is widely used
as a white pigment in polymers. The anatase form of TiC>2 is an effective
photosensitiser for the photooxidative degradation of the polyolefins [46,47]
and many theories have been put forward as to the source of the initiating
radicals. These range from electron transfer from photo-excited Ti0 2 with
the direct formation of hydroxyl radicals [48],

hv
H20 > H+ + e(aq) + OH (19)
Ti0 2

possibly through absorption of water on the surface of the pigment [49,50]


to the primary formation of Superoxide, Reaction (16) [51,52],

hv
Ti0 2 + 0 2 > Ti0 2 + 0 2 (20)

with secondary formation of 102 by radical annihilation, Reaction (17) [53].

2H
2 02 * > H202 +1 0 2 (21)

Which of these is more important probably depends on the conditions at


the site of the reaction (e.g. UV wavelength, [0 2 ], [H 2 0] etc.) but this can
only be determined in model systems using reactions which are diagnostic
for the oxygen species involved. For example, hydroxyl radical formation can
be diagnosed by hydroxylation of aromatic compounds [54] or by spin
INITIATORS, PROOXIDANTS AND SENSITISERS 93

trapping [55,56]. Superoxide generally gives hydrogen peroxide which is the


source of hydroxyl radicals (Reaction (4)) and singlet oxygen undergoes
chemical reactions that are quite distinct from the reactions of other acti­
vated oxygen species.

Sens* Sens

3 l
02 02 — ROOH (Photodynamic
activity)

CAR* CAR

Scheme 1. Photoprotective action of the carotenoid pigments (CAR).

One of the most useful diagnostic characteristics of singlet oxygen is its


quenching by the carotenoid pigments, a process which is widely used in
nature to inhibit the damaging effects of light. Foote and Denny [57] were
the first to propose that the photoprotective effect of ß-carotene (IX) in vivo
involves quenching of 1 0 2 They found that concentrations of ß-carotene (IX)
as low as 1(T* M effectively inhibited sensitised photooxidation and calcu­
lated that one ß-carotene molecule can quench up to 1 0 0 0 1 0 2 molecules [58].
The mechanism proposed for the protective action of carotene in biological
systems is outlined in Scheme 1, the ß-carotene triplet decays with the
emission of light at 520 nm [59]. ß-Carotene, therefore, falls into the preven­
tive class of antioxidants (see Chapter 5). However, its usefulness, which it
shares with other conjugated olefins [58], appears to be limited to biological
systems [60]. The effectiveness of polyenes as photoantioxidants for the
photo-bleaching of chlorophyll-a is related to the number of conjugated
double bonds in the quencher. ß-Carotene is quite ineffective in protecting
polypropylene against photooxidation, confirming that singlet oxygen is not
an important photo-initiator for the oxidation of essentially saturated poly­
mers. Under these conditions carotene like other conjugated hydrocarbons
is rapidly destroyed [62] (see section 3.1.1.) and may even behave as a
sensitiser for oxidation. The behaviour of the carotenoid 1 0 2 quenchers is
clearly complex and again depends very strongly on the conditions being
investigated.
A second and more unambiguous distinction between the reactions of
ground state oxygen, derived alkylperoxyl radicals and singlet oxygen can
94 GERALDSCOTT

be demonstrated by analysis of the products of their reaction with unsatu-


rated compounds. For example, many conjugated dienes such as a-terpene
form endoperoxides with singlet oxygen (Reaction (16)), whereas they oxi­
dise on the a-methylene group with 3 0 2 to give hydroperoxides (Reaction
(22)).

xy 3 o 2
»
(22)
OOH

Endoperoxide formation occurs by a concerted, 4-centre mechanism and


because the products do not readily thermolyse to free radicals, this kind of
reaction can sometimes be used semi-quantitatively to measure the amount
of singlet oxygen being formed (see Table 1 ) [63]. In particular the reaction
of 1 0 2 with diphenyl isobenzofuran (DPBF) is so fast (Reaction (23)) that it
can be used as a quantitative measure for 1 0 2 without interference from
physical quenching processes [64]. At high DPBF concentrations, all the 1 0 2
is captured by the diene and the reaction is zero order. It can therefore be
used to measure the lifetime of 102 in various solvents.

Ph
2
02
X)
o
(23)
Ph
XII XIII
A comparison of singlet oxygen-quenching by common molecules is in­
structive, since it provides an explanation for the fact that 0 2 is not a very
important environmental sensitiser for the autoxidation of technological
molecules [65]. Table 2 shows that relatively abundant environmental
agents such as 3 0 2 , C 0 2 and H 2 0 can deactivate 1 0 2 relatively rapidly in
competition with olefins. Kearns and his co-workers [66] have shown that
the quenching ability of solvents varies over a range of three orders of
magnitude. The lifetime of 1 0 2 in water is the shortest that has been
measured (10 s) and at the other end of the stability scale are the
halogenated solvents (CF3C1 » 10" 3 s).
Also listed in Table 2 is the very high physical quenching rate constant
for ß-carotene which is approximately three orders of magnitude greater
than its chemical quenching rate constant and is even greater than the
INITIATORS, PROOXIDANTS AND SENSITISERS 95

TABLE 1

Reactivity of 1Ag O2 with olefins in methanol at room temperature [63]

Olefin k(\ mol 1 s *) Ä/ÄTME(

Me-C^Me 4.0-10 8 8.51


1.010 8
Me^.Me
Me^^Me
Me 4.7-10 7 1.00
4.010 7

/—v 1.010 7 0.21

Q T
Me
4.7-10 6

7.3-10 5
0.10

0.015

O 3.8-10 6 0.08

w
in MeOH/flBuOH (1:1);
^ TME = 2,3-dimethyl-2-butene.

TABLE 2

Rates of physical and chemical quenching of singlet oxygen ('Ag)

Molecule Rate constant Reference


kq(\ mol"1 s"1)

02 1.40-103 * 65
N2 0.0610 3 * 65
CO2 2.3040 3 * 65
H20 9.00-10 3 * 65
CH 3 CH=CHCH 3 (Cis) 25.00-10 3 + 65
C4H 9 CH=CH 2 6.70-10 3 + 65
CH3(CH2)7CH= CH(CH2)7COOCH3 (Cis) 74.0103 + 63
ß-Carotene 1.4-10 10 * 57
ß-Carotene 1.010 7 + 57
DPBF 7.0-10 8+ 68

* By physical quenching;
+ By chemical reaction.
96 GERALD SCOTT

chemical quenching rate constant for DPBF making it so useful as a diag­


nostic indicator for singlet oxygen (see Ref. [32], Chapters 5, 6, 8, 9, 20 and
33 for examples of the use of DPBF in the measurement of quenching rate
constants).

1.3 Initiation by Ozone

Unsaturated hydrocarbons react with ozone with rate constants varying


between 10~21 mol - 1 s^5 for polyphenyl hydrocarbons to over 10 5 1 mol" 1 s
for olefins to give oxygenated species including peroxides [69], see Table 3.
Ozone attack is an important phenomenon in rubber technology, not only
because of the rate at which it occurs but because under stressed conditions,
each molecule of ozone leads to the scission of the rubber molecule with the
rapid development of cracks which, if not averted, destroy the rubber [70].
The phenomenon of ozone cracking which is peculiar to rubbers under stress
will be the subject of detailed discussion in Volume II of this series and will
be considered here only in so far as the ozone reaction products lead to
initiation of conventional autoxidation reactions.

1.3.1 Attack of ozone at a saturated carbon atom


Typical alkane hydrocarbons react slowly with ozone to give peroxidic
products. The introduction of an aromatic ring increases reactivity but at
the C-H bond, not at the aromatic ring. Table 3 shows that polyphenylene
in which the double bonds are localised has similar reactivity to a saturated
carbon chain although localisation of double bonds as in polynaphthalene
substantially increases reactivity, almost to the level of the polydienes. This
is consistent with the view that in the saturated chain component, the
primary reaction is hydrogen abstraction whose rate is increased by electron
délocalisation in the carbon radical, and is confirmed by the fact that
polystyrene and low molecular weight analogues of polystyrene (e.g. iso-
propyl benzene XIV R=H) give rise to alcohols and hydroperoxides on
treatment with ozone at room temperature [71].

H OOH OH
I I I
RCH2CCH2R RCH2CCH2R RCH2CCH2R
A A
0
XIV
Ô Ô
Moreover, Razumovskii et al. carried out this reaction in the cavity of an
ESR spectrometer and showed that peroxyl radicals were formed [72]; the
INITIATORS, PROOXIDANTS AND SENSITISERS 97

TABLE 3

Rate constants (ki) for the reaction of ozone with different polymers and numbers of
chains broken for each reaction event (cp) in CCU at 20°C [69]

Polymer Structure unit Äi(l mol s ) qp

Polyphenyl

Polynaphthalene
m 510:"2

240"5

Polycarbonate CH3

KOH^O-H
CH,

Polyisobutylene CH3 1.2-10" 0.05


l
4-CH 2 -Ç—

Polyethylene CH3Jn 4.6-10"2 0.1


-r-CH2-CH2-i,
Ethylene-propylene 6-10" 0.06
copolymer
t CH2-CH2-CH-CH2
t
Polypropylene 810" 0.1
-lrCH2-CH-4;

Polystyrene
I ÇéH5
-CH,-C-
0.3 0.001

Polyvinylcyclohexane 0.8

[_ ~-~ CH 2 ~~ CH "jj;

Polyphenylacetylene [ W 1.410 3
I—CH = C hi
-t-CH2-CH=CH-CH2-l-n
Polybutadiene 610 4 0.006
r CH, 1
[—CH 2 -CH=C-CH 2 ^
Polyisoprene 4.4-10° 0.002

Cyclododecatriene 3.5-10°
98 GERALDSCOTT

radical concentration being directly correlated with ozone depletion from the
gas stream as it passed over the surface of the sample (see Fig. 4). Further­
more, hydroperoxide formation is directly proportional to the ozone concen­
tration in the gas stream and, at least during the early stages, to the time
of ozone treatment.

Fig. 4. Formation of peroxyl radicals in polystyrene (surface area 120 m ) reacted with
ozone at the concentrations indicated on the curves. (Reproduced from Developments in
Polymer Stabilisation-6, G. Scott (Ed.), Applied Science Publishers, 1983, p. 247).

The mechanism of the initiation process, which has been shown to occur
primarily on the surface of polystyrene [73] is shown in Scheme 2. Hydro-
peroxides are major products of the reaction and give rise to initiating
radicals for autoxidation. Ozone also induces their decomposition, probably
by Reaction (25);
OOH
I
—CCH, (25)

( + O H + 02)

1,3.2 Attack of Ozone at Localised Double Bonds


Table 3 shows that cis-polyisoprene is attacked by ozone at a rate almost
10 7 faster than its saturated analogue, ethylene-polypropylene copolymer.
Moreover, the products of the reaction are quite different and involve
INITIATORS, PROOXIDANTS AND SENSITISERS

CH 92CH
^"2

» Radicals

+ OH +R
Scheme 2. Ozonation of polystyrene.

O
/ \
O O
o, I I
RCH = CHR' ■> RCH — CHR' XV
(a)

(b)

1
feï + (c) / \
RCHOOH <-££-
H RCHOO+R'CHO - ^ RHC CHR'
2° » \ /
kW)
o—o
RCH[OOCH]n —OOCH— XVI
I I
R R
XVII
Scheme 3. Reaction of ozone with olefinic double bonds.

destruction of the double bond. The classical studies of Criegee and his
co-workers showed that the compounds normally called ozonides or iso-
ozonides (XVI) are secondary products of attack of ozone on double bonds
(see Scheme 3). The primary ozonides (molozonides, XV) are too unstable to
isolate. Criegee et al. also showed that polymeric peroxides with structures
analogous to XVII are also formed [74,75]. Neither ozonides nor polyperox-
100 GERALD SCOTT

ides are stable species and substantial amounts of free radical oxidation
products (carbonyl, hydroxyl, etc.) are always formed in rubbers during
ozonisation [76,77]. The rates of chain scission of rubbers have been shown
[78] to be related to the active oxygen content of the rubber. It is clear then
that quite apart from the very rapid rate of chain scission which results from
the separation of the species in step (b) of ozonide formation (Scheme 3),
hydroperoxidic species are formed during the ozonisation of unsaturated
polymers probably by Reaction (e) in Scheme 3 which can initiate normal
autoxidation processes.

1A Initiation by Other Atmospheric Pollutants

The main pollutants in the industrial atmosphere are sulphur dioxide


and oxides of nitrogen. Both have been shown to be effective initiators for
polymer degradation in the presence of light.

1.4.1 Sulphur dioxide


Sulphur dioxide is readily activated by light to the triplet state which can
hydrogen-abstract from hydrocarbon substrates or add to double bonds
(Scheme 4) to give radical species which initiate autoxidation [79,80]. How­
ever, both S 0 2 itself and derived sulphinic acids undergo facile redox
reactions with hydroperoxides to give highly reactive free radicals which can
also initiate autoxidation [81,82] and these processes rather than light
sensitised reactions are probably the main reason for the sensitivity of most
substrates to sulphur dioxide. Such reactions are autoretarding since the
end products of oxidation by hydroperoxides are the stable sulphuric and
sulphonic acids and these are effective antioxidants [82]. The mechanisms
of pro-oxidant and antioxidant reaction of S 0 2 will be discussed in Chapter
5 of this volume.

3
S 0 2 -^^ SO*2 - (a)
Ä HSO. + R

>c=c< (b)

RH
>c—c< > RS02H

cr o
Scheme 4. Radical reactions of excited SO2.
INITIATORS, PROOXIDANTS AND SENSITISERS 101

1.4.2 Oxides of nitrogen


Nitrogen dioxide is an odd electron molecule which is potentially capable
of both initiating and retarding radical chain reactions. Additionally, it
absorbs UV light and can dissociate into NO and O* which themselves have
initiating potential and are precursors of ozone by reaction with ground
state oxygen [83].
Nitrogen dioxide readily adds to olefinic unsaturation to give mixtures of
nitro compounds and nitrite esters (12), Scheme 5.

— C H = C H — + NO

N02 N02 NO. ONO

Scheme 5. Reactions of NO2 with olefins.

Although it is less reactive toward saturation hydrocarbons, it does react


slowly, particularly in the presence of light to give again mixtures of nitro
compounds and nitrite esters. This process is associated with pronounced
pro-oxidant effects [84] but the nitroalkanes are photo-antioxidants due to
their ability to reversibly eliminate nitrous acid which is a chain-breaking
donor antioxidant [84] (see Scheme 6). The mechanism of this process is
similar to that of other catalytic chain-breaking antioxidants and will be
discussed in Chapter 4.

RH + N0 2 > R + HONO
y YN0 2

ROO RN0 2 + RONO

ROOH + N0 2 «—^ HONO + > C = C <


Scheme 6. Reactions of NO2 with saturated hydrocarbons.
102 GERALD SCOTT

2. PEROXIDES

It will be clear from the foregoing sections that whatever may be the
primary reaction occurring between pro-oxidant species and the substrate,
peroxides are the universal and chemically identifiable products formed in
the presence of oxygen.
Homolysis of the weak peroxide bond in peroxides (including hydrogen
peroxide itself) gives rise to highly reactive radical species (notably hy-
droxyl, and alkoxyl radicals) which then initiate a conventional radical
chain oxidation process with ground state oxygen. This is why peroxides,
and particularly hydroperoxides hold such a key position in the mechanism
of autoxidation. Not only are the radicals hydrogen abstracting agents, but
alkoxyl radicals also readily undergo ß-scission to eliminate smaller mole­
cules, which in the case of macromolecular substrates can have devastating
effects on their properties.
Hydrogen peroxide and alkyl hydroperoxides, unlike the dialkyl perox­
ides, are also very susceptible to induced decomposition, particularly by
reducing agents and many examples of the induced decomposition of hydro­
peroxides have been shown to be important in indicating oxidative damage
in biological and technological substrates. These will be reviewed in detail
in subsequent volumes of this series and only the salient features of hy-
droperoxide chemistry will be discussed here.

2.1 Formation of Hydroperoxides

Peroxides may be formed in autoxidising systems by three main


processes. The first, which occurs predominantly with alkyl aromatic hy­
drocarbons and saturated hydrocarbons, involves the displacement of hy­
drogen at a tetravalent carbon atom by the well known initiated radical
chain mechanism, Reaction (26).

H 00- OOH
X 2 RH
PhC(CH 3 ) 2 > PhC(CH 3 ) 2 ° > PhC(CH)3)2 > PhC(CH 3 ) 2
(26)
(RH) (RO (ROO) (ROOH)

Olefins react in essentially the same radical chain reaction with ground
state oxygen but in the case of singlet oxygen they undergo "ene" addition of
oxygen with shift of the double bond in a non-radical process (see Section
1.2.). Conjugated olefins oxidise by the third mechanism, co-polymerisation
with oxygen to give polymeric peroxides:
INITIATORS, PROOXIDANTS AND SENSITISERS 103

X
CH2=CHPh * > XCH 2 CHPh °2 > XCH 2 CHOO
Ph 02/nCH 2 ==CHPh
(27)

XCH^HtOOCH^Hl^OO-
Ph Ph
Copolymer of styrene and oxygen is the only peroxide formed in the case of
styrene, and the polymer contains up to 30 styrene units per molecule [85].
If a substrate contains both a double bond and an allylic carbon then both
reactions may occur together. Thus in the case of indene, there is competi­
tion between hydrogen abstraction from the benzylic carbon atom and
addition to the conjugated double bond giving a copolymer of indene and
oxygen containing 5-10 indene units per molecule [86] (see Chapter 2,
Section 2.2.(c)). The kinetic chain length of this polymerisation process is
about 430 and the formation of one hydroperoxide unit in the oligomer is due
to chain transfer to indene monomer, thus initiating a new polymerisation
chain reaction. Other allylic olefins give similar low molecular weight
copolymers with oxygen, generally with less than three copolymer units per
molecule. (See Atmospheric Oxidation andAntioxidants, First Edition, p. 24
et seq. for fuller discussion.)
Conjugation in the olefin appears to be a prerequisite for the formation of
good yields of oxygen copolymers [87]. Thus in the case of the three isomeric
alkenyl benzenes, XVIII-XX, good yields of polymeric peroxides were ob­
tained from the first two, in spite of the presence of an allylic méthylène
group in both [21,88].

CH=CHCH3 CH2CH=CH2

XVIII XIX XX
By contrast, XX gives hydroperoxide exclusively during the early stages of
the reaction [88]. However, isomerisation of the double bond in XX occurs
during autoxidation. This is characteristic of all 1,4 dienes and is driven by
the greater contribution from 1,3 conjugation energy in the intermediate
radical [87];

-CHCH= CHCH= CH- o -CH= CHCHCH= CH- o -CH= CHCH= CHCH-


104 GERALDSCOTT

Thus, the hydroperoxides formed from linoleic esters are predominantly


(>90%) the conjugated isomers XXI and XXIII, (see Scheme 7), and co-poly­
merisation with oxygen to give the final cross-linked product does not occur
until substantial amounts of these isomers have been formed [89,90]. The
conjugated A 'n and A '* isomers of linoleic acid which have the corre­
sponding double bond arrangement to the two isomeric hydroperoxides XXI,
XXIII, do not form hydroperoxides until one mole of oxygen has been
absorbed and oxygen absorption is associated with reduction in conjugation
[91]. This accords well with the formation of polymeric peroxides.

14 13 12 11 10 9 8
CH3(CH2)3CH2CH=CHCH2CH=CHCH2(CH2)6COOR

OOH / | \ OOH
XXI -CH-CH=CHCH=CH- I -CH=CHCH=CHCH- XXIII
13 9
OOH
I
—CH=CHCHCH=CH—
11
XXII

Scheme 7. Isomeric hydroperoxides formed in the autoxidation of linoleate esters.

Although very high yields of hydroperoxides can be obtained from some


doubly activated méthylène hydrocarbons [87] (see Table 4), yields are
frequently much lower in the case of less reactive hydrocarbons. This
depends, to some extent, on the stability of the peroxides produced under the
conditions of their formation. Thus for example, since the oxidation of
saturated aliphatic hydrocarbons (even tert-àïkyl hydroperoxides) is so much
slower than that of the olefins and particularly the 1,4 dienes, decomposition
competes with formation. This will be discussed in more detail below.
Equally important, particularly in the co-polymerisation of oxygen with
vinyl compounds, is the oxygen pressure in the system. The yields of
styrene-oxygen 1:1 copolymer at normal oxygen pressures is 100% and the
rate of the termination step is independent of oxygen pressure indicating
that this occurs largely through alkylperoxyl [85]. However, as was seen in
the last chapter, as the oxygen pressure is decreased in the system, alkyl
radicals play an increasingly important role in termination and below 100
mm the formation of benzaldehyde, formaldehyde and styrene oxide in­
crease as a result of side reactions, (see Fig. 5). Mayo and his co-workers
studied this process in some detail [92,93] and showed that the oxygen-con­
taining products did not arise by decomposition of the polyperoxide and
concluded that they were formed by breakdown of the intermediate macro-
INITIATORS, PROOXIDANTS AND SENSITISERS 105

TABLE 4

Yields of hydroperoxides from the autoxidation of olefins [87]

Olefin Hydroperoxide Yield, %

Unconjugated Conjugated

PhCH2CH=CH2 PhCH=CHCH2OOH 100

CH3
PhCH2CH=C(CH3)2 1 100
PhCH=CHCOOH
1
CH3

CH3
(CH3)2C=CHCH2CH=C(CH3)2 - 1 100
(CH3)2C=CHCH=CHCOOH
1
CH3

CH3 CH3 CH3


1 1 1 80
RCHCH=CH2 RCCH=CH2 20 RC=CHCH2OOH
1
OOH

RCH2CH=CH2 RCHCH=CH2 50 RCH=CHCH2OOH 50


1
OOH

R = alkyl.

alkyl radical (see Scheme 8). The alternative processes increase with
decreasing oxygen pressure but the unzipping radical can only proceed so
long as there is an alternating sequence of monomer and peroxide along the
chain. The presence of two styrene units together stops the unzipping
reaction and at very low pressures depolymerisation decreases to zero.
With some aliphatic olefins in which a "stable" alkyl radical is formed by
attack of alkylperoxyl at the double bond, epoxide may be a major product
of the reaction. Thus in the case of ß-di-isobutene, Twigg found [94] that
epoxide was the main product with tert-butyl hydroperoxide and acrolein as
minor products. The mechanism proposed is shown in Scheme 9.
In general, the more reactive the intermediate carbon centred radical, the
more favoured is epoxide formation at the expense of oxygen attack. Thus
the yield of di-isobutene epoxide is independent of oxygen pressure to a
much higher pressure than is styrene [95].
106 GERALD SCOTT

0.06-fl
-<C e H 8 >»
Total C8HS

20 40 ' 6 0 80 720 3150


Oxygen pressure (mm)

Fig. 5. Effect of oxygen pressure on the rate and products of styrene oxidation at 50°C.
(Reproduced with permission from J. Am. Chem. Soc, 80 (1958) 2470).

ROO+nCH 2 =CHPh RO[OCH 2CHO]nOCH 2CH


D2/ Ph \ Ph

RO[OCH2CHO]OCH2CHOO " RO[OCH2CHO]OCH2CHO ■


1 I Ph
Ph Ph
etc + PhCH,—CH,
o
RO+CH.O + PhCHO

Scheme 8. Alternative reactions of macroradicals during the co-polymerisation of styrene


and oxygen.

Dialkyl peroxides are in general more stable than hydroperoxides since


they are much less susceptible to radical or solvent induced decomposition
(see below). However, there is evidence that electron-withdrawing substitu-
ents reduce the stability of dialkyl peroxides and diacyl peroxides are less
stable than dialkylperoxides. Thus copolymers of oxygen and polar mono-
INITIATORS, PROOXIDANTS AND SENSITISERS 107

ROO
I .
ROO+(CH3)3CCH2CH=C(CH3)2 > (CH3)3CCH2CH—aCH3)2

ROOH + (CH 3 ) 3 CCHCH=aCH 3 ) 2 (CH3)3CCH2CH—aCH3)2 + RO

0 2 /RH

OOH
I
(CH3)3CCHCH=CHC(CH3)2

O*
I
(CH3)3CCHCH=C(CH3)2 > (CH3)3C*+(CH3)2C=CHCHO

0 2 /RH

(CH3)3COOH

Scheme 9. Autoxidation of di-isobutene.

mers such as acronitrile or chloroprene decompose even at room tempera­


ture [87]. This tendency is offset by electron releasing groups such as methyl
so that methacrylonitrile and methylmethacrylate polyperoxides are rela­
tively stable compared with the corresponding copolymers of acrylonitrile
and methylacrylate [90]. This is primarily a direct effect on the 0 - 0 bond
strength [87].

2.2 Reactions of Hydroperoxides

Hydroperoxides are much more susceptible to the nature of the substrate


than are dialkylperoxides. They readily take part in electron and hydrogen
transfer reactions and the presence of oxidising and reducing agents either
already present in the substrate or formed by their own decomposition,
generally determine the stabilities of hydroperoxides rather than the homo-
lytic stability of the peroxide bond [87].

2.21 Self-induced decomposition


Most hydroperoxides decompose in a first order reaction at low concentra­
tions but by a second order process at high concentrations (see Chapter 2).
Thus Stannet and Mesrobian [97,98] found that 1,4-dimethylcyclohexane
108 GERALDSCOTT

hydroperoxide in 1,4-dimethylcyclohexane was first order at concentrations


below 0,03 molar with an activation energy of 91 kJ mol - 1 whereas at 1.9
molar concentration in the same solvent, second order kinetics were ob­
served and a lower activation energy (83.5 kJ mol -1 ). Bateman and co-work­
ers came to a similar conclusion from a study of the kinetics of cyclohexane
autoxidation [99] and they were able to demonstrate the existence of a
strong hydrogen bonded association between hydroperoxide molecules
which is dependent both on concentration and on the nature of the medium.
Thus stearic acid disrupted the hydrogen bonding and the decomposition
became first order again [100]. It was proposed that at high concentrations,
radical formation occurs through the hydrogen bonded dimer;

2 ROOH ^ [ROOH OOR] > ROÔ + H 2 0 + OR (28)


H

2.2.2Radical-induced decomposition
The alkoxyl radical produced in the above reaction is itself able to further
induce the decomposition of the hydroperoxide. Thus, tert-butyl hy­
droperoxide decomposes rapidly in chlorobenzene at 140°C to give only
tert-butyl alcohol and oxygen [101] by the radical chain reaction shown in
Scheme 10 (cycle A). In the presence of an oxidisable solvent (i.e. a solvent
with a labile hydrogen), much more alcohol is formed than can be accounted
for by the A cycle alone and Kharasch and co-workers [102,103] proposed a
reductive induced decomposition involving solvent radicals (R' •). The B cycle
(Scheme 10) is readily inhibited by oxygen [104].

ROOH ROH ROOH ROH

R B RO A ROO

ROH RH
Scheme 10. Radical induced decomposition oftert-butylhydroperoxide (ROOH).

A particularly powerful solvent effect on the rate of hydroperoxide decom­


position was observed by Stannet and Mesrobian [105], who found that
cumene hydroperoxide decomposed more rapidly in styrene than it did in
aromatic solvents at 113°C. Walling and Chang [106] could not find hydroxyl
groups in the product as required by the Kharasch reductive mechanism but
INITIATORS, PROOXID ANTS AND SENSITISERS 109

dialkyl peroxides were present and they concluded that in this case induc­
tion is by macroalkyl (P). This seems unlikely in view of the nucleophilic
character of alkyl radicals but there is little doubt that radical induction is
involved [87].

ROOH + P ROO + PH (29)

ROO + PhCH= CH 2 ROO[CH 2 CH-] n CH 2 CH (P) (30)


Ph Ph

Radical induced decomposition of hydroperoxides can be eliminated by


efficient hydrogen transfer (chain-breaking donor) antioxidants. Thomas
[107] found that increasing quantities of phenyl-a-naphthylamine (PAN)
caused a decrease in the rate of decomposition of a number of hydroperox­
ides to a minimum at a given temperature and in a given solvent. Above a
certain concentration, no further reduction in rate of decomposition was
observed. Furthermore, Oberright and co-workers [108] found that in the
case of cumene hydroperoxide, the ratio of the yields of cumyl alcohol to
acetophenone approaches infinity under these conditions, whereas it is 3 in
the absence of antioxidant; clear evidence of rapid removal of the inducing
cumyl radical, (see Scheme 11).

CH.—C=0

4- *CH3 Alkyl radicals


(or derived peroxyl
CH, CH3 radicals) induce
I CH,
CH3—C—OOH CH,—C—O decomposition
CH 3 — C — O H of cumene hydro-
RH ^ /W. peroxide
f J +R"
(ROOH) (RO) AHN
CH3
I
CH3—C—OH

No induced
+ A" decomposition
(ROH)

Scheme 11. Effect of an antioxidant on t h e induced decomposition of cumene hydro-


peroxide.
110 GERALD SCOTT

Under conditions of complete radical removal by PAN trapping, Thomas


measured the first order rate constants and activation energies for the
decomposition of several typical hydroperoxides [107]. These are listed in
Table 5.
TABLE 5

Decomposition of hydroperoxides at 150°C in white oil in the presence of phenyl-a-


naphthylamine (PAN) [107]

Hydroperoxide 104 k\ (mm *) AE (kJ mol *)

a-Tetralyl 81 84.2
Cumene 81 84.2
n-Octyl 56 80.4
2-(2,4,4'-trimethyl) pentyl 56 80.4

2.2.3 Photolysis of hydroperoxides


Hydroperoxides are among the most powerful initiators of photooxidation
[109-114]. The key chemical reactions involved were recognised by
Bateman and Gee [115] to involve photolysis of hydroperoxides.
hv
ROOH > RO'+OH (31)
Martin and Norrish [116] showed that alkyl hydroperoxides do not absorb
above 350 nm and are relatively unaffected by light > 350 nm. However, at
313 nm, tert-butyl hydroperoxide and cumene hydroperoxide both decom­
pose at significant rates [116,117] with the formation of more strongly
absorbing chromophores such as acetophenone. More recent studies have
suggested [118-120] that such photolysis products of hydroperoxide photo­
lysis are responsible for sensitising them to photolysis. Thus Li and Guillet
[120] found that macromolecular hydroperoxides in combination with
ketonic decomposition products give much higher quantum yields of macro-
molecular chain scission products than did ketone alone. More than 87% of
the light was absorbed by the ketone and these authors argued for efficient
energy transfer from ketone to hydroperoxide. Kilp et.al. [119] showed that
hydroperoxides are effective quenchers for triplet carbonyl and suggested
that alkoxyl radicals result directly from this process. However, there is
little question that whether photosensitised or not, hydroperoxides are
photolysed by UV light with a quantum efficiency approaching 1 [112,121]
and are the major source of photooxidation in substrates in which they are
present.
INITIATORS, PROOXIDANTS AND SENSITISERS 111

2.2.4 Redox reactions of hydroperoxides


The reactions of hydroperoxides are dominated by their ability to react
with reducing agents in redox reactions. In some cases, for example in their
catalytic decomposition by transition metal ions, both reduction and oxida­
tion of hydroperoxides are involved with continual production of radical
species;

ROOH + M+ > RO+OH"+M2+ (32)

ROOH + M 2+ > R O O + H + + M+ (33)

There is good evidence that the substrate may in some cases act as a
reducing agent with radical formation by Reaction (34);

ROOH + R'H > R O + H 2 0 + R' (34)

Under conditions where the radical-induced reactions discussed in Sec­


tion 2.2.2 are eliminated by means of hydrogen donor antioxidants, the rate
of hydroperoxide decomposition is still highly sensitive to the oxidisability
of the solvent. In Table 2, Chapter 2, the rate constants for the decomposi­
tion of tetralin hydroperoxide in a range of solvents [122] are given. There
is clearly a substantial solvent effect which cannot be due to induction by
alkylperoxyl radicals. It is not so clear however that it is not due to induction
by alkyl radicals, since electron (hydrogen) donor antioxidants are ineffi­
cient traps for alkyl radicals. There is no doubt, however, that the nature of
the substrate profoundly influences the rate of hydroperoxide initiation
during autoxidation. Thomas and Harle [122] were able to demonstrate a
linear relationship between the induction period to inhibited oxidation of a
variety of substrates and the rate of tetralin hydroperoxide decomposition
in the same substrates and under the same conditions, see Fig. 6. Under
these conditions, initiation cannot be due to alkyl radical-induced decom­
position of hydroperoxide, since the concentration of the latter must be
negligible in the presence of oxygen. It must be concluded then that the
redox Reaction (34) is the main contributor to the initiation reaction in the
presence of hydroperoxides.
More powerful reducing agents are even more effective generators of
initiating radicals from hydroperoxides and the technology of chemical
plasticisation of rubbers (mastication) is based on this fact [123]. This will
be discussed in more detail in Volume II (Chapter 3) of this series, but in
general any compound which releases an electron or hydrogen to hydro­
peroxide can act as an effective pro-oxidant. The best known examples are
thiols, sugars, amines, hydrazines, keto-enols and even phenols [123].
The role of sulphur compounds is particularly interesting because they
112 GERALDSCOTT

OEthyl palmitate
20.0 -Bis ( J2-ethylhexyl ) sebacate

10.0(- )Tetnadecane

OctadecaneO
5.
^ydrogenated polybutene
JC .White oil
Hexa (2-ethylhexyl)Ö
Ï 2.0| disiloxane

1.0
S3 Dioctyl ether Cl o Polybutene
. O Polypropylene
0.5

0.2
OTetralln
100 1000 10,000
Peroxide decomposition rate X 1 0 4 (min - 1 )

Fig. 6. Relationship between the induction periods of various substrates containing


tetralin hydroperoxide (containing an antioxidant) and the rate of decomposition of
tetralin hydroperoxide at 171°C in the same solvents. (Reproduced with permission from
J. Phys. Chem., 63 (1959) 1027).

have both pro-oxidant and antioxidant activity depending on the conditions,


and in particular the molar ratio of hydroperoxide to sulphur compound.
Thus, many sulphur antioxidants whose mechanism involves the formation
of sulphur acids which decompose hydroperoxides heterolytically show an
initial pro-oxidant effect (see Chapter 5). This phenomenon of pro-oxi-
dant/antioxidant inversion is also important in rubber technology, since it
explains why sulphur cross-linked rubbers initially oxidise more rapidly
than carbon-carbon cross-linked rubbers, but whereas the latter show
normal auto-accelerating characteristics, the former auto-retard to give
ultimately an oxidatively stable product.
The oxidation of ß-thiodipropionate esters has been studied in some detail
because of their importance as antioxidants in polymers [125]. Pro-oxidant
processes can be observed at each stage in the oxidation of these compounds
by hydroperoxides as outlined in Scheme 12. The hydroperoxide oxidation of
inorganic acids such as sulphurous acid or hydrogen chloride have been
shown to have a free radical component and even a mixture of tert-butyl
hydroperoxide and sulphuric acid has been reported to be an initiator for
methyl methacrylate [126] at molar ratios [r^SOJ/MBuOOH] > 1. Similarly,
hydrogen chloride in combination with cumene hydroperoxide is an effective
pro-oxidant for the oxidation of cumene at molar ratios [HC1]/ [CHP] > 1
[127,128], although at molar ratios < 1, it is an effective antioxidant.
The two competing processes lead to quite different products in the case
of cumene hydroperoxide. At [CHP]/[HC1] > 1, phenol and acetone are the
INITIATORS, PROOXIDANTS AND SENSITISERS 113

(R'OCOCH2CH2)2S -5222» [(R'OCOCH2CH2)2SOH + RO]

\ /
(R'OCOCH 2 CH 2 ) 2 S=0

ROOH
R'OCOCH=CH2 + R'OCOCH2CH2SOH ► R'OCOCH2CH2SO + RO+ H 2 0
ROOH

ROOH
R'OCOCH2CH2S02H -^—» R'OCOCH2CH2S 0 2 + R O + H 2 0

/
R'OCOCH2CH2S03H (R'OCOCH2CH2)
2^n2'2

+ S02 -522», HSOa + RO*

SO, + ROH

H 2 S0 4

Scheme 12. Prooxidant reactions of dialkylthiodipropionates and their oxidation products


[1251.

90 - Phenol

Acetone \ a - Cumyl alcohol 1

60 -

30

Acetophenone
a - Cumyl alcohol 7
tf
— Methanol
J 1 1 J 1 I I > =1 1
80:1 40:1 1:1 1:40 1:80
Molar Ratio ([CHP] : [HCl])

Fig. 7. Products formed from the decomposition of cumene hydroperoxide (CHP) at


0.910~ 2 Minthe presence of HCl at various molar ratios ([CHP]/[HC1]). Reproduced with
permission from Eur. Polym. J., 16 (1980) 175).
114 GERALD SCOTT

dominant products (see Fig. 7), whereas at molar ratios < 1, a-cumyl alcohol
was the major product with smaller amounts of acetophenone and methanol.
The chemistry of these competing processes was discussed in Chapter 1 of
this volume (see Scheme 3) and constitutes a useful diagnostic measure of
the importance of homolytic and heterolytic reactions to the mechanism of
hydroperoxide decomposition. In each case discussed above, and in others
which will be discussed in Chapter 5, there is a characteristic sharp change
in behaviour at [CHP]/[HX] = 1, where HX is HC1, H 2 S0 3 , RSOH, RS0 2 H,
etc. [81,82].
Oxidative decomposition of hydroperoxides is also known. For example
the oxidation of cyclohexene hydroperoxide by lead tetraacetate to give the
corresponding acetate and oxygen was used by Criegee and co-workers [129]
to elucidate the chemical structure of hydroperoxides formed by oxidation of
olefins.
OOH H^ /OAc
Pb(OAc)4 ^ I 1 , n U A A X , m ^5)
>2 I + Pb(OAc) 2 + 0 2

Other strongly oxidising metal ions (e.g. Ce +) also give quantitative yields
of the corresponding alcohols [130];

2 ROOH + 2Ce 4+ > 2 ROH + 0 2 + 2Ce 3+ (36)

However, for most transition metal ions (e.g. Co, Fe, Mn etc.), reduction
of hydroperoxides (Reaction (32)) is equally, if not more, important than
oxidation. Thus cobaltous acetate reduces tert-buty\ hydroperoxide to give
predominantly tert-butyl alcohol and oxygen. Dean and Skirrow [131] found
that 50% of the theoretical amount of oxygen was formed in the subsequent
reactions of alkylperoxyl.

2 ROO- > 2 RO- + 0 21


> Propagation
RO+ROOH > ROH + ROO- i (37)

2 ROO- > ROOR + 0 2 Termination

The products of the metal catalysed decomposition of hydroperoxides are


very similar to those of the uncatalysed process outlined in Scheme 10, cycle
A, and the effect of the transition metal ion is to reduce the activation energy
of the overall process (see Scheme 13). Transition metal ions also catalyse
redox reactions of hydroperoxides with reducing agents as discussed earlier
and many redox initiating systems for emulsion polymerisation are based
INITIATORS, PROOXIDANTS AND SENSITISERS 115

on combinations of hydroperoxides with reducing agents (polyamines,bisul-


phite, thiols, etc.), catalysed by transition metal ions [132].
The very pronounced pro-oxidant effect of many metal ions in autoxida-
tion is thus primarily due to their redox reactions with hydroperoxides (see
Chapter 2), and the kinetics of these processes is generally consistent with
the formation of a prior complex between metal ion and hyroperoxide [133].
Metal deactivators inhibit transition metal catalysed oxidation by competi­
tively complexing them (see Chapter 5).

RH

ROOH ROO KROOOOR] [ROOR + \02]

RO + ^ 0 2
2+ +
Co +H
ROOH
V

As in Scheme 11

RO+OHT ROOH

Scheme 13. Reaction of hydroperoxides with cobalt ions.

2.2.5Non-radical reactions of hydroperoxides


It has already been noted in Chapter 1 that the catalytic decomposition
of hydroperoxides by protonic species to non-radical products is an impor­
tant antioxidant process. The mechanism of peroxidolytic antioxidants will
be discussed in detail in Chapter 5 of this volume and their behaviour in
technological systems in Volume II. However, in biological systems, the
peroxidases and catalase appear to achieve the same effect by quite a
different mechanism. In spite of the extensive investigations carried out in
recent years on the peroxidolytic mechanism in vitro, none of these approach
the effectiveness of the biological peroxidases (see Chapter 1, Section 6.2.).
116 GERALDSCOTT

REFERENCES

1 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.


81 et seq.
2 B. Halliwell, J.M.C. Gutteridge and D. Blake, Phil. Trans. Roy. Soc., B311 (1985)
659.
3 B. Halliwell and J.M.C. Gutteridge, Mol. Asp. Med., 8 (1985) 89.
4 J.M. McCord and R.S. Roy, Can. J. Phys. Pharmacol., 60 (1982) 1346.
5 G.S. Egerton, J. Soc. Dyers Colour., 63 (1947) 161.
6 G.S. Egerton, Text. Res. J., 18 (1948) 659.
7 G.S. Egerton, J. Soc. Dyers. Colour., 64 (1948) 336.
8 G.S. Egerton, J. Text. Inst., 39 (1948) T293.
9 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
325.
10 S.D. Razumovskii and G.E. Zaikov, in G. Scott (Ed.), Developments in Polymer
Stabilisation-6, Applied Science Publishers, 1983, p. 239 et seq.
11 G. Scott, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, John Wiley and Sons,
1978, p. 230.
12 N. Grassie and G. Scott, Polymer Degradation and Stabilisation, Cambridge Uni­
versity Press, 1985, p. 190 et seq.
13 F.A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry, Second Edition,
Interscience, 1967, p. 366.
14 H.A.O. Hill, Phil. Trans. R. Soc., B294 (1981) 119.
15 B. Halliwell and C.H. Foyer, Biochem. J., 155 (1976) 697.
16 A.L. Lehninger, Biochemistry, Second Edition, Worth Pub. Inc., 1975, pp. 447, 486.
17 J.R. Shelton and W.L. Cox, Ind. Eng. Chem., 46 (1954) 816; Rubber Chem. Technol.,
27 (1954) 671.
18 E.I. Tinyakova, B.A. Dolgoplosk and V.N. Reikh, Izv. Akad. Nauk, SSSR, Obk.
Khum. Nauk, 9 (1957) 111; Rubber Chem. Technol., 32 (1959) 231.
19 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
391 et seq.
20 B. Anderson, Arkiv. Kemi., 33 (1950) 451.
21 L. Bateman, Q. Rev., 8 (1954) 147.
22 N. Uri, Nature, 177 (1956) 1177; Chem. Ind., 23 (1956) 515.
23 G.A. Rüssel, J. Am. Chem. Soc, 78 (1956) 1047.
24 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
82 et seq.
25 R.T. Holman, Arch. Biochem., 21 (1949) 51; 26 (1950) 85.
26 F. D. Gunstone and T. P. Hilditch, J. Chem. Soc., (1946) 1022.
27 T.P. Hilditch, Nature, 166 (1950) 588.
28 J.L. Bolland, Trans. Faraday Soc, 44 (1948) 669.
29 A. Combe, M. Niclause and M. Letort, Rev. Inst. Fr. Pétrole, 10 (1955) 786.
30 J.A. Gray, J. Chem. Soc, (1953) 741.
31 R.H. Young, K. Wehnby and R. L. Martin, J. Am. Chem. Soc, 93 (1971) 5775.
32 B. Rânby and J.F. Rabek, Singlet Oxygen, Reactions with Organic Compounds and
Polymers, John Wiley and Sons, 1978.
33 R.D. Ashford and E.A. Ogryzlo, J. Am. Chem. Soc, 92 (1970) 3293.
INITIATORS, PROOXIDANTS AND SENSITISERS 117

34 P.D. Bartlett and A.P. Schaap, J. Am. Chem. Soc., 92 (1970) 3223.
36 S. Mazur and C.S. Foote, J. Am. Chem. Soc. 92 (1970) 3225.
36 N.M. Hasty and D.R. Kearns, J. Am. Chem. Soc., 95 (1973) 3380; S. Mazur and C.S.
Foote, J. Am. Chem. Soc., 92 (1970) 3225.
37 H.C. Ng. and J.E. Guillet, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, John
Wiley & Sons, 1978, p. 278.
38 J.F. Rabek, Y.J. Shur and B. Rânby, in Singlet Oxygen, John Wiley & Sons, 1978,
p. 264.
39 A.K. Breck, C.L. Taylor, K.E. Rüssel and J.K.S. Wan, J. Polym. Sei., AI 12 (1974)
1505.
40 J.F. Rabek, Mechanisms of Photophysical Processes and Photochemical Reactions
in Polymers, John Wiley & Sons, 1987, p. 552 et seq.
41 J.S. Ziegler and J. D. Goosey, Photochem. Photobiol., 33 (1981) 869.
42 B. Frank, Angew. Chem. (Int. Ed.), 21 (1982) 343.
43 A.G. Motton, C F . Chignell and R.P. Mason, Photochem. Photobiol., 38 (1983) 671.
44 F.W. Grant and J. Greene, Toxic. Appl. Pharmacol., 23 (1972) 71.
45 A.K. Davies, S. Navartnam and G.O. Phillips, J. Chem. Soc. Perkin Trans. I, (1979)
22.
46 N.S. Allen and J.F. McKellar, Photochemistry of Dyes and Pigmented Polymers,
Applied Science Publishers, 1980, p. 247.
47 Degradation and Stabilisation of Polyolefins, N.S. Allen (Ed.), Applied Science
Publishers, 1983, p. 355 et seq.
48 H.G. Voelz, G. Kaempf and H.G. Filsky, Prog. Org. Coat., 3 (1974) 223.
49 H.G. Voelz, G. Kaempf and A. Klaern, Farbe + Lack, 82 (1976) 805.
50 A.H. Boonstra and C.A.H.A. Mustaers, J. Phys. Chem., 79 (1973) 1694.
51 W.F. Sullivan, Prog. Org. Coat., 1 (1972) 157.
52 G.S. Egerton and K.M. Shah, Text. Res. J., 38 (1968) 130.
53 S.P. Pappas and W. Kuhhirt, J. Paint Technol., 47 (1975) 42.
54 R. Richmond and B. Halliwell, J. Inorg. Biochem., 17 (1982) 95.
55 E. Finkelstein, G.M. Rosen, E.J. Raukman and J. Paxton, Mol. Pharmacol., 16
(1979) 676.
56 E.G. Janzen, D.E. Nutter, E.R. Davis, B.J. Blackburn, J.L. Poyer and P.B. McKay,
Can. J. Chem., 56 (1978) 2237.
57 C.S. Foote and R.W. Denny, J. Am. Chem. Soc, 90 (1968) 6233.
58 C.S. Foote, R.W. Denny, L. Weaver, Y. Chang and J. Peters, Am. N.Y. Acad. Sei.,
171 (1970) 139.
59 E.J. Land, A. Sykes and T.G. Truscott, Photochem. Photobiol., 17 (1973) 43.
60 M.M. Mathews-Roth, M.A. Pathak, T.B. Fitzpatrick, L.C. Harber and E.H. Kass,
New Eng. J. Med., 282 (1970) 1231.
61 D. Bellus, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, Wiley and Sons, 1976,
p. 61.
62 S.M. Anderson, N.I. Krinsky, M.J. Stone and D. C. Clagett, Photochem. Photobiol.,
20 (1974) 65.
63 K. Gollnik, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, Wiley & Sons, 1976,
p. I l l et seq.
64 J.A. Howard and G.D. Mendenhall, Can. J. Chem., 53 (1975) 2199.
65 G. Scott, in B. Rânby and J.F. Rabek (Eds.), Singlet Oxygen, Wiley & Sons, 1976, p.
230.
118 GERALDSCOTT

66 D.R. Kearns, Chem. Rev., 71 (1971) 395.


67 F.H. Doleiden, S.R. Fahrenholtz, A.A. Lamola and A.M. Trozzolo, Photochem.
Photobiol., 20 (1974) 519.
68 B. Stevens, S.R. Perez and J.A. Ors, J. Am. Chem. Soc, 96 (1974) 6846.
69 S.D. Razumovskii and G.E. Zaikov, in G. Scott (Ed.), Developments in Polymer
Stabilisation-6, Applied Science Publishers, 1983, p. 239 et seq.
70 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965,
Chap. 10.
71 V.Ya. Shlyapintokh, A.A. Kefeli, V.l. Goldenberg and S.D. Razumovskii, Dokl. Akad.
Nauk SSSR, 196 (1969) 1132.
72 T.V. Pokholok, R.M. Vikhlyaev, O.N. Karpukhin and S.D. Razumovskii, Vysokomol.
Sold., B l l (1969) 692.
73 S.D. Razumovskii and G.E. Zaikov, Neftekhimiya, 13 (1973) 101.
74 R. Criegee, Ber., 88 (1955) 1878.
75 P.S. Bailey, Chem. Rev., 58 (1958) 925.
76 R. Criegee, Angav. Chem. (Int. Edn.), 14 (1975) 745.
77 R.L. Kuczkowski, Ace. Chem. Res., 16 (1983) 42.
78 A.D. Delman, B.B. Sims and A.E. Ruff, J. Polym. Sei., 45 (1960) 415.
79 W. Funke and H. Haagen, in D.P. Garner and G.A. Stahl (Eds.), The Effects of
Hostile Environments on Coatings and Plastics, ACS Symp. Ser. 229,1983, p. 309.
80 H.H.G. Jellinek, in H.H.G. Jellinek (Ed.). Aspects of Degradation and StabiUsation
of Polymers, Elsevier, 1978.
81 M.J. Husbands and G. Scott, Eur. Polym. J., 15 (1979) 249; C. Armstrong, M.J.
Husbands and G. Scott, Eur. Polym. J., 15 (1979) 241.
82 B.B. Cooray and G. Scott, Eur. Polym. J., 17 (1981) 233.
83 R.M. Harrison and CD. Holman, Chem. Br., 18 (1982) 563.
84 S. Al-Malaika, T. Czeckaj, L.M.K. Tillekeratne and G. Scott, Polym Deg. Stab., in
press.
85 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
102 et seq.
86 G.A. Rüssel, J. Am. Chem. Soc, 78 (1956) 1035,1041.
87 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
23 et seq.
88 F.R. Mayo, A.A. Miller and G.A. Rüssel, J. Am. Chem. Soc, 80 (1958) 2500.
89 J.A. Cannon, K.T. Zilch, S.C. Birkett and H.J. Dutton, J. Am. Oil Chem. Soc, 29
(1952) 447.
90 J.K.T. Fuyger, J.A. Zilch, J.A. Cannon and H.J. Dutton, J. Am. Chem. Soc, 73 (1951)
2861.
91 R.R. Allen, A. Jackson and F.A. Kumerov, J. Am. Oil Chem. Soc, 26 (1949) 395.
92 F.R. Mayo and A.A. Miller, J. Am. Chem. Soc, 78 (1956) 1017.
93 F.R. Mayo and A.A. Miller, J. Am. Chem. Soc, 80 (1958) 2480.
94 G.H. Twigg, Chem. Sei. (Proc Conf. Oxid. Processes), Suppl. 3 (1954) 5.
95 F.R. Mayo, J. Am. Chem. Soc, 80 (1958) 2499.
96 S.F. Strause and E. Dyer, J. Am. Chem. Soc, 78 (1956) 136.
97 V. Stannett,and R.B. Mesrobian, Discuss. Faraday Soc, 14 (1953) 242.
98 V. Stannett, A.E. Woodward and R.B. Mesrobian, J. Phys. Chem., 61 (1957) 360.
99 L. Bateman, H. Hughes and A.L. Morris, Discuss. Farad. Soc, 14 (1953) 190.
INITIATORS, PROOXIDANTS AND SENSITISERS 119

100 L. Bateman and H. Hughes, J. Chem. Soc, (1952) 4595.


101 E.R. Bell, J.H. Raley, F.F. Rust, F.H. Seubald and W.E. Vaughan, Discuss. Faraday
Soc., 10 (1951) 242.
102 M.S. Kharasch, A. Fono and W. Nudenberg, J. Org. Chem., 15 (1959) 763.
103 M.S. Kharasch, A. Fono and W. Nudenberg, J. Org. Chem., 16 (1951) 113.
104 B.K. Morse, J. Am. Chem. Soc, 79 (1957) 3375.
105 V. Stannett and R. B. Mesrobian, J. Am. Chem. Soc, 72 (1950) 4125.
106 C. Walling and Y.N. Chang, J. Am. Chem. Soc, 76 (1954) 1578.
107 J.R. Thomas, J. Am. Chem. Soc, 77 (1955) 246.
108 E.A. Oberright, S. J. Leonardi and A. P. Kozacik, ACS Symp. Div. Pet. Chem., (1956)
115.
109 G. Scott, in N. Grassie (Ed.), Developments in Polymer Degradation, Applied
Science Publishers, 1977, p. 205.
110 G. Scott, in Stabilisation and Degradation of Polymers, Adv. in Chem. Ser., 169
(1978) 30.
111 B.B. Cooray and G. Scott, Polym. Deg. Stab., 3 (1980/1) 127.
112 D.J. Carlsson, A. Garten and D.M. Wiles, in G. Scott (Ed.), Developments in Polymer
Stabilisation-1, Applied Science Publishers, 1979, p. 219.
113 G. Scott, J. Photochem., 25 (1984) 83.
114 G. Scott, Br. Polym. J., 16 (1984) 271.
115 L. Bateman and G. Gee, Proc R. Soc, A195 (1948-9) 376, 391.
116 J.T. Martin and R.G.W. Norrish, Proc. R. Soc, A222 (1953) 322.
117 R.G.W. Norrish and M.H. Searby, Proc. R. Soc, A237 (1956) 30.
118 G. Geuskens, D. Baeyens-Volant, G. Delaunois, Q. Lu-Vinh, W. Peret and C. David,
Eur. Polym. J., 14 (1978) 291, 299.
119 T. Nismontski-Knittel and T. Kilp, J. Polym. Sei., Polym. Chem. Edn., 21 (1983)
3209.
120 S.K.L. Li and J.E. Guillet, Macromolecules, 17 (1984) 41.
121 D.J. Carlsson and D.M. Wiles, Macromolecules, 2 (1969) 597.
122 J.R. Thomas and O.L. Harle, J. Phys. Chem., 63 (1959) 1027.
123 G. Scott, Atmospheric Oxidation and Antioxidants, First Edition, Elsevier, 1965, p.
391 et seq.
124 C. Armstrong, F.A.A. Ingham, J.G. Pimblott, G. Scott and J.E. Stuckey, Proc. Int.
Rubber Conf., Brighton, May 1972, F2.1.
125 G. Scott, in Developments in Polymer Stabilisation-6, Applied Science Publishers,
1983, p. 29 et seq.
126 E. Rizzardo and D.H. Solomon, J. Macromol. Chem., A14 (1980) 33.
127 B.B. Cooray and G. Scott, Chem. Ind., (1979) 741.
128 B.B. Cooray and G. Scott, Eur. Polym. J., 16 (1980) 169.
129 R. Criegee, H. Pilz and H. Flygare, Ber., 72 (1939) 1799.
130 M.S. Kharasch, A.C. Poshkus, A. Fono and W. Nudenberg, J. Org. Chem., 16 (1951)
1458.
131 M.H. Dean and G. Skirrow, Trans. Faraday Soc, 54 (1958) 849.
132 R.G.R. Bacon, Q. Rev., 9 (1955) 287.
133 H.S. Laver in G. Scott, (Ed.), Developments in Polymer Stabilisation-1, Applied
Science Publishers, 1979, p. 167 et seq.
This page intentionally left blank
121

Chapter 4

ANTIOXIDANTS: CHAIN BREAKING MECHANISMS

GERALD SCOTT

1. THE CHAIN-BREAKING DONOR MECHANISM

In Chapter 1 it was seen that the development of antioxidants was a


practical response to the problem of the deterioration of organic based
materials, notably natural rubber and foodstuffs. In practice, then, empiri­
cal solutions were available to solve the problem of oxidative deterioration
before the underlying chemistry had been elucidated. It was not until the
mechanism of autoxidation emerged that antioxidants were recognised as
being inhibitors of a radical chain reaction involving alkyl and alkylperoxyl
radicals as the chain propagating species and hydroperoxides as the in­
digenous initiators of autoxidation. The development of this theory provided
the basis for the subsequent classification of antioxidants into "chain-break­
ing" and "preventive" [1], (see Chapter 1). The first is concerned with the
removal of the chain-carrying species (CB mechanism) and the second with
inhibiting the catalysed homolytic decomposition of hydroperoxides or with
their removal by reactions which do not give radicals (preventive mecha­
nism). The latter will be the subject of Chapter 5 and in the present Chapter
the removal of radicals by reduction (CB-D) and oxidation (CB-A) will be
discussed. Scheme 1 summarises these processes and it will be seen (Section
3) that under favourable conditions, both CB-D and CB-A mechanisms may
operate together resulting in the continuous removal of both alkylperoxyl
and alkyl radicals in a catalytic mechanism [2].
For the reasons discussed above, the most widely recognised and most
thoroughly studied mechanism of antioxidant action is the removal of
alkylperoxyl radicals from autoxidising systems by reduction; the CB-D
process, Reaction (1).

ROO+AH > ROOH + A' (1)

Early work on autoxidation and antioxidant action was carried out in


petroleum based liquid hydrocarbons or in low molar mass analogues of the
unsaturated rubbers where the complication of oxygen diffusion was
122 GERALDSCOTT

Oo RH
R' ROO* R+ROOH

A- AH

RA ROOH + A

I
_ = C+AH
\> Stable products
/ \
CHAIN-BREAKING CHAIN-BREAKING
ELECTRON (OR HYDRO­ ELECTRON (OR
GEN) ACCEPTOR HYDROGEN) DONOR

CB-A CB-D

Scheme 1. The chain-breaking mechanisms of antioxidant action.

deliberately eliminated in order to simplify the theoretical treatment of the


results. This has had important implications for the development of the
subject since it led to the neglect of one of the major parameters determining
the effectiveness of antioxidants in polymers, in lubricating oils and, more
recently, in biological systems, where the actual oxygen pressure may be
much lower than the partial pressure of oxygen in the atmosphere.
Early studies of antioxidant mechanisms were largely concerned with
physical chemical studies of the kinetics of inhibited oxidation on the one
hand and the investigation of the oxidation chemistry of phenols and amines
on the other. (See the first edition of Atmospheric Oxidation and Antiox-
idants for a discussion of the early investigation of antioxidant mecha­
nisms.) Both of these stratagems were necessary to the full understanding
of the relationship between antioxidant structure and activity and the dual
approach still persists in the current literature. In the present treatment,
the chemistry of antioxidant transformations under oxidising conditions
will first be reviewed in order to provide a basis for the understanding of
structure-activity relationships.

1.1 The Oxidative Transformation of Phenols and Aromatic Amines

In principle, a range of electron or hydrogen donors are capable of reduc­


ing alkylperoxyl radicals to the corresponding hydroperoxides (Reaction (1)).
In practice, only those that give radical products (A-) which do not readily
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 123

continue the kinetic chain by hydrogen abstraction from the substrate,


Reaction (2), or by reaction with oxygen of the atmosphere, are useful as
antioxidants;

RH AH + R- ° 2 > ROO

(2)

°9 ^^* AOO RH
> AOOH + R- ° 2 > ROO.

The very complex chemistry of phenol [3] and aromatic amine [41 oxida­
tion has been thoroughly surveyed by Pospisil and the reader is directed to
these excellent reviews to supplement the present discussion.
Both phenoxyl radicals and arylaminyl radicals formed in Reaction (1)
undergo further reaction by two main processes: (i) dimerisation and (ii)
reaction with oxygen derived radicals. This is summarised in simplified
form for the most widely used commercial hindered phenol, BHT in Scheme
2 [3,51.
The formation of all the quinonoid products shown in Scheme 2 involves
the removal of more than one hydrogen and the number (f) of alkylperoxyl
radicals deactivated by the CB-D mechanism is normally two or less. There
is no doubt that the initial step is the removal of the phenolic hydrogen by
alkylperoxyl since the first product always observed by electron spin reso­
nance during single electron oxidation of phenols is the phenoxyl radical
[6-8]. In the case of phenols without bulky alkyl groups in the 2 and 6
positions, the half-life of the phenoxyl is microseconds and it can only be
observed in ESR by a continuous flow procedure in which the radical is being
continuously formed. As the number of ortho fer£-alkyl groups is increased,
the half-life of the free phenoxyl increases. II has a half-life at room tempera­
ture of several minutes in cyclohexane solution [7]. The ESR spectrum of II
is a quartet of triplets due to a major interaction of the electron with the
p-methyl hydrogens (<Xp = 10.7 gauss) and a minor interaction with the ring
hydrogens (a m = 1.8 gauss). This indicates that 6.4% of the electron density
is delocalised in the methyl group by hyperconjugation. Other phenoxyl
radicals are much more stable than II. Thus X, which is the ultimate stable
radical product formed in the single electron oxidation of BHT, is one of the
classical "stable" radicals, galvinoxyl [9]. The electron is delocalised equally
in both aromatic rings giving a doublet of quintets (a p = 5.9 gauss, a m = 1.4
gauss).
Although galvinoxyl (X) and other phenoxyls containing a delocalising
group in the para position are stable in the absence of air they, like less
stable aryloxyls, react slowly with oxygen (Scheme 2) and more rapidly with
o
tBu^JL^tBu

y™
CH 3 OOR

-A:
O O

; XX
iBu tBu /„2/RI1 tBu

I,BHT CH 3 CH 3 OOH 7
CH, CH 3 O

0Jy—\U
OH tBu tBu
tBu

Y^ CH3 CH 3 \=<
CH; tBu tBu

tBu tBu

0=/ \=CH-/ V-O- X

tBu tBu

tBu tBu tBu tBu

HO-H^ V-CH 2 CH 2 —( \— OH -^-> O CHCH O

tBu tBu tBu tBu


VIII IX

Scheme 2. Oxidation of a hindered phenol by alkylperoxyl.


ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 125

tBu

tBu tBu tBu


R'H + R'

tBu tBu

a
R O [Co(III)(acac)2OH]
XII
Scheme 3. Transformation of hindered phenoxyl radicals in the presence of oxygen and
cobalt acetylacetonate (Co(acac)2).

alkyperoxyls to give peroxydienones (e.g. III-V) [3,5,10-18]. These are


relatively stable compounds at ambient temperatures in the dark, but at
elevated temperatures [19,20] and particularly in the presence of UV light
[21,22], Reaction (3), they give initiating radicals VI which, in the absence
of a hydrogen abstractable substrate, give stable rearrangement products
(XI).

tBu. .tBu tBu tBu x ^ A ^ tBu


hv
XT (3)
CH3 OOR CH, O ^COCHj
III VI XI
126 GERALD SCOTT

Reaction (3) provides an explanation for the fact that hindered phenols
are ineffective photoantioxidants in polymeric substrates [3,23] although
they synergise strongly with UV absorbers (UVA) which are able to protect
them from photolysis [23-25].
TkâC [8] in an elegant study of the oxidation of hindered phenols by ESR,
has recently provided convincing evidence for the importance of the reaction
of oxygen with phenoxyls (Scheme 2). When a variety of phenols are oxidised
by alkylperoxyl radicals coordinated to cobalt in the absence of oxygen, the
radical immediately observed in the ESR is always the primary aryloxyl.
However, in the presence of oxygen, a second and much more complex octet
signal due to interaction with the Co nucleus, is observed which Tkafi
interprets as being due to the formation of a cobalt coordinated cyclohex-
odienoneoxyl radical, XII, without aromatic character (see Scheme 3). The
ESR coupling constants for a number of phenoxyls and derived quinones are
listed in Table 1.

TABLE 1

Coupling constants in the ESR signals of free phenoxyl and complexed cyclohexodienon-
oxyl radicals [11]

R Signal iro XII (+)

a 3,5 a H
4 aco aHCß aj?2(6)

C(CH3)2 Octet-doublet 0.18 _ 1.05 _ 0.35


C(CH3)2Ph do 0.18 - 1.05 - 0.35
CH(CH3)Ph Octet-triplet 0.18 0.66 1.05 0.35 0.35
C 6 Hn do 0.18 0.46 1.05 0.25 0.25
H do 0.19 0.96 1.05 0.21 0.27
CH2CH3 Octet-quartet 0.18 0.90 1.05 0.35 0.35
CH2Ph do 0.16 0.87 1.05 0.35 0.35
CH3 Octet-quintet 0.18 1.07 1.03 0.43 0.34

(*) Subscripts refer to the positions in the ring shown in II', Scheme 3.
(+) Subscripts for H and C refer to the positions in the ring XII, Scheme 3.

Tkâc has identified both primary aryloxyl and secondary cobalt coordi­
nated radicals as oxidation products from a wide range of hindered and
partially hindered phenols including méthylène and sulphur bridged com­
pounds (XII-XVI) widely used as industrial antioxidants;
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 127

tBu tBu

K>-Q-|-(S-OH HO

R
R" ' R"
XIII XIV

XV
He draws some general conclusions from the rate of formation and the
stability of the phenoxyl radicals produced which are relevant to their
antioxidant activity [8]. The most important of these are:
(i) The activation energy of hydrogen abstraction by alkylperoxyl in­
creases and the stability of the phenoxyl decreases with decreasing steric
hindrance in the ortho positions.
(ii) Substitution of the a-hydrogens in a 4-alkyl group by alkyl substan­
tially increases the stability of the phenoxyl and hence CB-D activity.
(iii) Increasing bulk of the 4 substituent decreases the ability of the
phenoxyl to deactivate a second alkylperoxyl (to give III).
(iv) The incorporation of an alkylidine bridge in the 2 or 4 positions
decreases the tendency of the initially formed phenoxyl to disproportionate
with loss of the second phenolic group.
(v) The presence of a sulphur bridge increases the délocalisation of the
unpaired electron, thus permitting a lower degree of steric hindrance with­
out sacrificing CB-D activity.
Phenoxyl radicals, in spite of their stability, can also act as oxidising
agents in the presence of a reducing substrate. Thus thiols readily reduce
phenoxyls (e.g. II) to the parent phenols [26].
Similar reactions are known with unhindered phenols [27,28] and with
other readily oxidised compounds such as ascorbic acid. The latter process
is the basis for the well known synergism between a-tocopherol and ascorbic
acid [29] (see Chapter 1).
Irradiation of arylamines by UV light or high energy irradiation, gives
rise to aminyl radicals in the absence of oxygen but these are so unstable
that they can be observed only by the dynamic (continuous flow) procedure
in an inert medium [30-32]. Reaction of alkylperoxyls with arylamines gives
128 GERALDSCOTT

OH

u
tBu^/L/tBu
+RS
II + R'SH
;-
CH 3 i (4)
O

y
tBu^JL^tBu

Π3 SR'
the much more stable nitroxyl radicals [18,33-371 and the concentration
obtained is much higher than in the case of the phenoxyls from phenols.
Moreover, nitroxyls are also formed from diarylamines by oxidation with
hydroperoxides even at room temperature [38]. However, when used as
antioxidants, the concentration of alkylperoxyls is quite low and under these
conditions dark conjugated dimerisation products are also formed by linking
through the reaction 2 and 4 positions in the aromatic ring, Scheme 4 [4].
The products are complex mixtures of high molar mass arylamines, imides
and nitroxyls, all of which can behave as antioxidants under specific condi­
tions.
These transformations account for the generally higher antioxidant activ­
ity of the arylamines than the hindered phenols. The detailed mechanisms
of the action of ary lamine transformation products will be discussed in more
detail in the next section.
Like the hindered phenols, both arylaminyl radicals (e.g. XVII) [39-41]
and nitroxyl radicals (e.g. XX) [42-44] can react with a further alkylperoxyl
through the aromatic ring (see Scheme 5) and in this respect they differ from
aliphatic amines and their derived nitroxyls. Tka£ has shown the formation
of XXII as the primary product during the oxidation of pheny-ß-naphy-
lamine (XXI) with Co(III) *BuOO but when Co(III) *BuO is used as the
oxidising agent, the para coupled product, XXIII, is obtained [45].
The different behaviour of alkylperoxyl from that of alkoxyl (see Scheme
5), provides confirmation for the view that nitroxyl formation occurs by the
stepwise mechanism shown in Scheme 4.
Quinonoid products are frequently formed in the oxidation of dipheny-
lamines by alkylperoxyl radicals. These can only result from attack at the 4
position and Berger and co-workers [46] have proposed the reaction shown
in Scheme 6 to explain the formation of benzoquinone among the products
of oxidation of diphenylamine.
Nethsinghe and Scott [47] have shown that phenolic nitrones can also
generate nitroxyl radicals by hydrogen abstraction. Oxidation of XXVII
H

Q K ^ Q ^[0-*-0
t—i

I + ROOH
O
X
>
XVII

H H
ROO
I
OOR O
»
w

O
w
o
x
>
XVIII g
k ROOH
XX +OR S
ROO -

N N
0- <X> 0 XIX
Oligomers

Scheme 4. Reaction of diphenylamine with alkylperoxyl.

to
130 GERALDSCOTT

NH—t.-TS
\=/
7 co(m)tBuoo'
CodIDtBuO*

XXI XIV
CoUIDtBuO'

XXII XXIII
Scheine 5. Reactions of phenyl-ß-napthylamine with oxyl radicals.

ok} - okx
XVIIa
ROO
H

o»o° -ROO'
XXV
O-Ä-OT ^
N
O f O - °K>°* O -° * O
ROO
NO,

° XXVI
Scheme 6. Reaction of diphenylnitroxyl with alkylperoxyl.

showed no evidence of aryloxyl in the ESR, only a strong nitroxyl triplet


(XXVIIIb) being observed (see Scheme 7).
Although aromatic nitroxyls, like the phenoxyls discussed above, are
stable radicals they can be reduced to the corresponding hydroxylamines
relatively easily. In the absence of air they react readily with alkyl radicals
to give an alkylhydroxylamine which can undergo elimination to give the
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 131

Me O
ROO.
HO- /VCH-N- tBu
Me XXVII XXVIIIa

O
I
CH—N—tBu

Me XXVIIIb
Scheme 7. Nitroxyl formation by oxidation of a phenolic nitrone.

\
;N—o +R- = N—OR'] ^ = * ^ N — OR

I
^;N—OH + ^:c=c^
XXIX
Scheme 8. Reduction of nitroxyl radicals by alkyl.

hydroxylamine (XXIX) and an olefin (see Scheme 8). This compound is itself
a powerful scavenger for alkyl peroxyls and under certain conditions this
process can lead to catalytic antioxidant activity (see Section 3.1).

1.2 Kinetics of Inhibited Oxidation

Correlation of antioxidant structure and activity depends on having an


accurate and realistic method of measuring the ability of antioxidants to
inhibit autoxidation. It should be noted that in technological systems other
factors such as solubility in the medium and loss from the medium at high
temperatures or in the presence of extraction solvents, may also be more
important. A discussion of this aspect of antioxidant activity will be deferred
until Volume II and in this section the activity of antioxidants in pure
132 GERALD SCOTT

hydrocarbon substrates which are saturated at equilibrium with the oxygen


in the environment, will be considered.
A convenient and generally reproducible method of monitoring antiox­
idant activity is the measurement of induction period to oxygen absorption
in either an initiated or uninitiated autoxidation process [48-50]. This
procedure is widely used as an initial "screening" test to rate antioxidant
structures in order of effectiveness and some typical results will be dis­
cussed in the next section. It does not, however, give information in the rates
of the individual steps involved in the chain-breaking antioxidant mecha­
nism and a study of the kinetics of inhibited oxidation provides complemen­
tary information. A kinetic treatment of oxygen absorption originally
developed by Bolland and his co-workers [51-53], has proved to be helpful
in understanding how chain-breaking donor antioxidants compete with the
substrate for oxygen of the environment
The starting point of Bolland's analysis was the kinetic theory of auto­
xidation discussed in Chapter 2. Bolland studied the oxidation of ethyl
linoleate in the presence of 10"4 moles of antioxidant and 3*10~2 moles of
benzoylperoxide per mole of ester. The rate of oxidation in the absence of
antioxidant is given by Eqn (5) (see Chapter 2, Section 2).

-d[0 2 ] = r u = kske-1/2 r{ 1/2[RH] (5)

where rx = total rate of chain initiation.


Two possible chain terminating reactions involving antioxidant, AH,
were considered;

ROO + AH > ROOH + A- (6a)


krj

R. + AH > RH + A- (6b)
ks

followed by dimerisation of the stable radicals,

2 A- > Non-radical products (7)

There are two limiting equations for the rate of inhibited autoxidation, ra.
The first, Eqn (8), holds when k7 > > k8

k^f% = ra[AH] [RH] = K (8)

and the second, Eqn (9), holds when ks>> k7


ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 133

* 6 * 2 ~ V V l O ^ o O ■T-\• ^f" [RHl^AHF1 = K (9)


* solubility of 0 2 in ethyl linoleate at 100 torr.

Equation (8) was found to be in agreement with experiment and since k%


and k§, depend (at least on the basis of the underlying assumptions of the
above scheme) only on the structure of the olefin, then 1/K is a quantitative
measure of the chain-breaking efficiency of the antioxidant. Bolland and ten
Have [53] measured 1/K for a number of phenolic chain-breaking anti-
oxidants and these are listed in the first column of Table 2. Davies et al. [54]
using the same basic premises measured kjk^1 by means of Eqn (10).
Antioxidant ratings relative to hydroquinone (100) are listed in the second
column in Table 2.

- 3 - = krjkf [RH]"1 j - tl (10)

The two methods give reasonable agreement between themselves and with the
induction period method [55] (Table 2, column 3) for a number of phenols.

TABLE 2

Efficiency of simple phenols as antioxidants relative to hydroquinone

Antioxidant rate
1/K [53] fcvA*"1 [54] IP [55]

1.6 _ 1.2
2-Hydroxyphenanthrene - 8 -
ß-Naphthol 7.7 1.5 1.2
p-Methoxyphenol 17 - -
3-Hydroxyphenanthrene - 18 -
4-Hydroxyphenanthrene - 56 -
a-Naphthol 56 71 26
1-Hydroxyphenanthrene - 78 -
Pyragallol 300 - >53
Catechol 63 - >46
Hydroquinone 100 100 100
Toluhydroquinone 150 - -
Trimethylhydroquinone 570 - -
1,4-Naphthahydroquinone 4000 - -
134 GERALD SCOTT

A more extensive investigation by Bickel and Kooyman [56,57] who also


included some aromatic amines in their study, showed that some antiox-
idants obeyed a relationship of the form;

Initial oxidation rate atRHKAH]"1 + c[RH]

They found it necessary to invoke a transfer reaction between antioxidant


and substrate, Reaction (11), to account for the kinetics.

A+RH * 10 > AH + R- (11)

Assuming stationary state conditions they derived Eqn (12) and were able
to measure both antioxidant efficiency {krjk^1) and chain transfer activity
(^io^9~ ) f° r a number of hindered, partially hindered and unhindered
phenols.

ra = rj + kzkrfx [RHKAH]-1 + r{ k10kf1/2 V _1/2 [RH]"1 (12)

Their results, which are listed in Table 3, are in broad agreement with ESR
studies discussed in the last section which indicated decreasing stability in
this series and increasing tendency for the derived phenoxyl radicals to react
further with the substrate.
The kinetic approach to the measurement of antioxidant efficiency dis­
cussed above assumes that a stationary concentration of reaction interme­
diates exists. This is of course true when an initiator is added to the system
which can be considered to be in practice the only source of initiating
radicals. It is not true in the case of a chain-branching reaction in the
absence of initiator, particularly during the induction period to autoxidation
when the concentration of both radical species and hydroperoxide is continu­
ously changing. Shlyapnikov has shown [58] that the mathematical treat­
ment of inhibition in a chain-branching system provides a theoretical basis
for the well-known fact that below a certain critical antioxidant concentra­
tion there is a discontinuity in the concentration activity relationship below
which the antioxidant becomes essentially ineffective.
Shlyapnikov assumes that initiation occurs by Reactions (13) and (14)

k
RH + 0 2 ° > R- + products (13)

AH + 0 2 2-* R- +products (14)


ANTIOXIDANTS: CHAIN BREAKING MECHANISMS

TABLE 3

Antioxidant activities of 2,4,6 substituted phenols [56,57]

OH

"•£r*
Ri R2 R3 Ä7Ä53"1 k10k9-1/2

*Bu tBu *Bu 32 0


tBxx tBu Me 33.5 0
tBu Me *Bu - small
tBu Me Me 250 0
Me Me tBu 16-31 large
Me Me Me 260-350 0.0035
H H *Bu 80-100 0.0086
tBu H tBu 118 0

and derives Eqn (16) for the concentration of alkylperoxyl where ô is the
chain-branching probability.

LKUU-j - h^[AH] _ ô ( / e s [ R H ] + A,7[AH] UW

A stationary concentration of ROO exists only when the denominator in (16)


is > 0. Therefore the critical inhibitor concentration, [AH]cr, the concentra­
tion of AH at the boundary between the stationary and non-stationary
states, can be determined from the equation

ÔA3[RH]
[AH] cr = Mö gw , (17)
(l-ô)/e7

Shlyapnikov concludes that inhibitors that do not show critical antiox­


idant concentrations or where [AH]cr is high are inefficient. Thus [AH]^ for
phenyl-a- and phenyl-ß-naphylamine at 200°C are 5-20 times higher than
for the p-phenylene diamines AT-phenyl-AT-cyclohexylamine and N,N-dicy-
clohexyl-p-phenylene diamine at 200°C.
136 GERALD SCOTT

1.3 Structure-Activity Relationships in the Chain-Breaking Donor


Antioxidants

In the 1950s there was a spate of activity in several technological sub­


strates to determine the most effective antioxidant structures by induction
period measurement. Although the substrates were rarely characterised
and reproducible, individual authors generally used the same procedure in
the same medium. Some broad correlations were found to be possible
between quite different media such as mineral oils, fats and oils and
petroleum 159].
Results obtained using the induction period as a criterion of antioxidant
activity and comparing this with the kinetic parameter k^k^1 have been
reviewed in some detail for phenolic antioxidants earlier and only the salient
conclusions will be discussed here. Much of the information concerns the 2,4,6
substituted phenols and the following generalisations have been drawn:
(i) Electron releasing groups (methyl, methoxy, alkylamino etc.) in the 2
and 4 positions markedly increase antioxidant activity, whereas electron
attracting groups reduce antioxidant activity.
(ii) Groups in the 2 and 4 position which délocalise the unpaired electron
on the derived radical (notably aryl), increase antioxidant activity.
(iii) Branched chain alkyl groups in the 2 and 6 positions increase antiox­
idant activity.
Criteria (i) and (ii) also apply to aromatic amines but not (iii).
The first two generalisations receive a satisfactory explanation on the
basis of the stability of the transition state in the reaction of alkylperoxyl
with a phenol or amine group;

X^(?)Vo-H O — OR
V^Y 8'
R

Since the alkylperoxyl radical is on electron acceptor it has a partial


negative charge in the transition state, leaving the aromatic ring of the
electron donor electron deficient. Group X decreases or increases the energy
of the transition state by releasing or attracting electrons from the ring.
Similarly, substituents in R or X which délocalise the electron density
formally on the oxygen also decrease the transition stage energy.

(i) Inductive effects


Early studies by Fieser [60] showed that electron attracting groups in the
4 position increased the critical oxidation-reduction potentials of phenols,
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 137

whereas electron releasing groups decrease it (see Table 4). This is more or
less the order of their antioxidant activity and Bolland [53] later demon­
strated a substantially linear relationship between the log of antioxidant
activity and the normal oxidation-reduction potential of phenolic antiox-
idants. Penketh [61] in a comprehensive study of substituted phenols and
amines by polarographic oxidation,concluded that to be a good inhibitor an
antioxidant had to have an oxidation potential ((OP)o) below 0.7 volts.
However, there appears to be a lower limit to the useful oxidation potential
range since some amines and phenols with very low oxidation potentials are
ineffective as antioxidants. Boozer and Hammond [62] were also able to
correlate the electron releasing power of substituents with their effect on
antioxidant activity. Effective releasing groups such as alkoxy and alkyl
amino produced the most favourable effects and many modern commercial
phenols and arylamines contain these substituents.

TABLE 4

Effect of substituents on the critical oxidation potential of phenols; R-/^ \ - O H


relative to phenol (R=H) \—/

R Change (volts) R Change (volts)

N02 +0.344 CH2CH=CH2 -O.037


COOH +0.268 CH2COOH -O.03
CHO +0.212 Ph -O.042
S0 3 H +0.084 CH3 -0.052
Cl +0.005 OPh -0.059
CH2OH -O.037 OCH3 ^0.0235
CH2Ph -0.037 N(CH3)2 -0.371

The lower limit on the oxidation potential of phenols and amines appears
to be associated with direct attack of oxygen which, as has been seen
(Reaction (14)) can give rise to initiating radicals. In a detailed kinetic study
of the oxidation of rubbers containing p-phenylenediamines (XXX) Shelton
and his co-workers [63] concluded that direct attack of oxygen occurred in
the case of the iV,AT-di-sec-butyl-/?-phenylene diamine (XXX, R = R2 = sec
butyl) since it behaved as a pro-oxidant. N,iV-diphenyl-p-phenylene diamine
on the other hand, did not oxidise directly with oxygen [64]. Shelton has also
explained the concentration optimum that is frequently observed with this
type of arylamine amine as being due to direct oxygen attack at higher
concentrations [63].
138 GERALDSCOTT

RtNH-f .-TV. >-NHR 2 xxx

There has been considerable dispute between mechanists as to whether


or not chain-breaking donor antioxidant action involves the initial formation
of a charge transfer complex between the aromatic ring and the alkylperoxyl
group. This was sparked by the observation [65] that tetramethyl-p-pheny-
lene diamine (XXXI), which does not have an abstractable hydrogen on the
nitrogen, is quite an effective antioxidant;

CH, CH,

ROO (18)
OOR

CH CH,
XXXI XXXXII

The Würster ion-radical, XXXII, can certainly exist under autoxidising


conditions [66,67]. Boozer and Hammond [67] showed that tetramethyl
p-phenylene diamine can terminate two kinetic chains, probably involving
oxidation of the alkyl groups These authors also reported [62,68] that the
kinetics of oxidation of some phenols involved second order dependence of
the rate of inhibited oxidation on the concentration of alkylperoxyl. They
proposed the reversible formation of a charge transfer complex followed by
the attack of a second alkylperoxyl (Reaction (19)).

OH OH

+ ROO ^
ROO*

XXXIII XXXIV

The main argument proposed to refute this suggestion was that both
phenolic and amine antioxidants containing a labile hydrogen showed a
weak isotope effect and that quinonoid products (e.g. VII) were normally
formed along with the peroxydieone (XXXV) under autoxidation conditions
[69-71]. At the time there was no evidence for the formation of the interme­
diate aryloxyl which is the precursor of VII but the ESR studies discussed
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 139

earlier in this chapter can leave no doubt now that hydrogen abstraction
precedes the formation of both peroxydienone and quinonoid products.

(ii) Electron délocalisation


Alcohols and aliphatic amines are virtually without antioxidant action.
The reason for this is of course that the unpaired electron formed by
hydrogen abstraction is highly localised on oxygen or nitrogen so that they
are effective transfer agents. At the other extreme are the hydroxylamines
which give very stable nitroxyl radicals. Although the rate of hydrogen
abstraction from these species is very high [72,73] they are not useful as
antioxidants under normal conditions because they are readily oxidised by
atmospheric oxygen (see, however, Section 1.3).

(in) Steric effects


It is clear from ESR studies discussed in Section 1.1 above, that the
nature of the 2,4 and 6 alkyl group is of critical importance to the stability
and subsequent reaction of the initially formed aryloxyl. In the phenol,
XXXIII, increasing branching in Rx and R2 increases the stability of the
derived phenoxyl. However, this is gained at the expense of hydrogen
transfer activity since Table 4 shows that Ay/^"1 is greater for the partially
hindered phenols. The optimal alkyl group in the 4 position appears to be
methyl. BHT (XXXIII, R1R2 = £Bu, R3 = Me) is 7.5 times more intrinsically
effective as an antioxidant on a molar equivalent basis in decalin as
measured by the induction period than the second most widely used com­
mercial antioxidant in which R3 is -CH 2 CH 2 COOCH 18 H 35 [74]. TkaS's ESR
studies (Section 1.1) confirm that increasing branching in the 4 position
decreases the ability of the phenoxyl radical to trap a second alkylperoxyl.
Most of the studies that have been carried out on the antioxidant effi­
ciency of CB-D antioxidants have centred on the parent hydrogen donor.
There is now a good deal of evidence, particularly from the work of Pospisil
[34], that oxidative transformation products are formed from both phenols
and aromatic amines at an early stage in their useful life and that some of
these have similar chain-breaking activity to the parent hydrogen donor.
Thus, for example, Pospisil has isolated the high molar mass trimer,
XXXVII, from the oxidation products of the widely used rubber antioxidant
XXXVI (Reaction (20)).
Scott and co-workers have shown [74] that deliberate oxidation of poly­
propylene containing simple hindered phenols such as BHT can lead to
much more effective oligomeric structures (e.g. IX, and its precursors,
Scheme 2) which are better able to resist the effects of loss by volatilisation
in a technological environment. Similar condensation products of aryl-
amines (e.g. XVIII) have been isolated from technological media during
oxidation [4], It is clear then that it is invalid to consider only the antiox­
idant activity of the starting hydrogen donor, particularly in technological
140 GERALD SCOTT

OH OH

Yr CH n rS-
i
,tBu

ïy »V
CH3
1
CH3
XXXVI
ROO

A
OH O OH OH
BU (20)
rV™'-rV
1 H
,tBu tBu^

y
""YT -CH 2 -

Y CH3
H
eu
V V
1
CH,
1 '
1
CH3
1 CH,
CH

tBu Y " ^^ tBu


OH
XXXVII
media where the derived oxidation product may be much more effective than
the parent compound for physical reasons. This aspect will be considered in
more detail in Volume II.
Oxidation transformation products of antioxidants are also of consider­
able practical importance from other points of view. There is increasing
emphasis on the toxicological and aesthetic aspects of antioxidant be­
haviour. In general, hindered phenols and many aromatic amines are
colourless and relatively non-toxic when pure but articles containing them
frequently become discoloured during service due to oxidation products.
Although the parent compounds may have been declared non-toxic by the
appropriate licensing authorities, their transformation products have
generally not been examined. This aspect of antioxidant chemistry is ex­
pected to become the cause of increasing concern in the future as antiox­
idants move into more critical applications such as foodstuffs packaging,
body implants and replacement organs [75].

2. CHAIN-BREAKING HYDROGEN (ELECTRON) ACCEPTOR MECHANISM

2.1 The Involvement of Alkyl Radicals in Inhibition Processes

Although it is obvious from Scheme 1 that alkyl radicals can in principle


be scavenged during autoxidation, it is only relatively recently that the
practical importance of the CB-A mechanism has been recognised. In the
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 141

first edition of this book [76], the subject was treated as a logical possibility
based on the theoretical reaction of alkyl radicals with radical trapping
agents which have been recognised for many years to be effective inhibitors
of vinyl polymerisation due to their ability to trap macroalkyl radicals. It
was also recognised [2,77] that the effectiveness of a chain-breaking ac­
ceptor antioxidant must depend on its ability to compete with oxygen (see
Scheme 1).
It was this last factor which led to the relatively late discovery of the
acceptor mechanism. However, the early literature of antioxidant action
contains many references to the use of oxidising agents of which the most
commonly quoted examples were quinones and nitro compounds [78]. The
former which were known to trap alkyl radicals [79] were shown to be
effective inhibitors of mechano-oxidation in rubbers by Slonimskii and his
co-workers [80].
Much of the evidence for the chain-breaking acceptor mechanism has
come from studies in polymers where oxygen concentration is much lower
than in liquid hydrocarbons. Denisov has reported [81] that the Henry
coefficient, 7, for a liquid petroleum hydrocarbon is 1.3 - 1.8-10 mol 1"1
atm"1 whereas in the amorphous phase of polyethylene it is 3.440 mol 1_1
atm -1 . Under these conditions, diffusion of oxygen to the site of the reaction
becomes important and Denisov has shown that the rate of diffusion of
oxygen in high density (crystalline) polyethylene at room temperature is
about 100 times slower than in liquid hydrocarbons. Thus in polymers we
would expect the alkyl to alkylperoxyl ratio to be much higher than in liquid
hydrocarbons and this has indeed been shown to be the case. On the basis
of the above solubility and diffusion rate differences, Denisov has calculated
that [R-]/[ROO] is 240"^ in polypropylene at 371 K and 540" 6 in isopentane
at 373 K [81].
It was seen earlier (Chapter 2) that oxygen concentration has a pro­
nounced effect on the contribution from self termination processes (21), (22),
involving alkyl radicals due to variation in the [R-]/[ROO] ratio.

R. + ROO- > ROOR (21)

2R- > R-R (22)

This will be equally applicable to the participation of alkyl radicals in


inhibition processes.
Chain-breaking acceptors are very weak inhibitors in oxygen saturated
liquid hydrocarbon solutions. Thus 2,2,6,6-tetramethyl piperidinoxyl
(XXXVIII, R=H) and p-benzoquine have no significant antioxidant activity
in liquid hydrocarbon oxidation at 10 mol 1"1 but both are equally effective
antioxidants in polypropylene at 387 K [81]. They were shown to act by
142 GERALD SCOTT

trapping macroalkyl radicals. Anthracene and o-dinitrophenol were also


effective but their activity was an order of magnitude lower. Table 5 com­
pares the rate constants for inhibition by the CB-A mechanism in polypropy­
lene with rate constants for the competing reaction with oxygen. It is clear
that the CB-A antioxidants can compete with oxygen but only when the
oxygen concentration in the medium is low.

XXXVIII

TABLE 5

Rate constants for the reaction of macroalkyl radicals (P-) in polypropylene (7 = 3-10
mol l"1 atm"1) [81]

Reaction T(K) k
1 mol"1 s"1 (or s"1)

P- + 02 370 106
P- + XXXVIII 387 4.7-104
P- + p-benzoquinone 387 3.810 4
P- + anthracene 387 1.810 3
P- + o-dinitrophenol 387 2.4-103

3. CATALYTIC ANTIOXIDANTS

It was seen in Section 1.1 that most chain-breaking antioxidants act


sacrificially and can normally not destroy more than two chain propagating
radicals per molecule. That is, the stoichiometric coefficient of inhibition fis
generally between 1 and 2. However, from time to time inhibition processes
have been reported that have a higher value of f. Thus for example, as early
as 1963, Denisov reported [82] that/* for a-naphthylamine in the stabilisa­
tion of cyclohexanol at 140°C was 30. The explanation put forward was that
both the intermediate hydroxyalkyl radical (XXXIX) and the derived alkyl-
peroxyl (XL) are oxidised by the aminyl radical.
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 143

COH + PhoN- > PhoNH+ C = 0 (23)


2 2
/ /

OH
C + Ph 2 N- » Ph 2 NH + 0 2 + C = O (24)
00

Both reactions were highly exothermic (Reaction (23), + 224 k J mol - 1 ,


Reaction (24), + 150 k J mol -1 ) but in the presence of oxygen, Reaction (24)
is the main reaction occurring. This process of course regenerates the
original antioxidant, Ph 2 NH, and makes possible a cyclical process in which
the antioxidant acts as a catalyst for the removal of the peroxyl radical
(Scheme 9).

HCL .00

Ph 2 N' * + Ph,NH + 0 ,

H O ^ ^OO HO OOH

+ Ph2NH * f J + Ph2N-

Scheme 9. Catalytic antioxidant behaviour of diphenylamine in cyclohexanol.

Other reversible oxidising systems were found to behave in the same way
[83], Thus, benzoquinone/quinhydrone in isopropanol had an /Value of 20,
and for Cu + /Cu + in cylcohexanol, f was believed to be infinity since the
species is not destroyed by radicals. Aliphatic amines were shown by Deni-
sov to be inhibited in the same way [84,85], the CB-A process in this case
involving oxidation (Reaction (25)).

00-
/ -H*
-CH > -C=NR +02 (25)
NHR

Alcohols and amines represent a rather special case due to the presence
of the heteroatom which labilises the abstractable hydrogen. Even so, it is
144 GERALD SCOTT

surprising that the arylaminyl radical Ph2N- can survive in the presence of
alkylperoxyl radicals in view of the findings of other workers, notably Tka£
(see Section 1.1) that the arylaminyl radical is highly unstable and reacts
rapidly with alkylperoxyl to give nitroxyl which is a relatively stable species.
It seems possible then that nitroxyl radicals may be involved in Reactions
(23) and (24).

3.1 Catalytic Antioxidant Action ofNitroxyls

Denisov and his co-workers [85,87] extended the concept of catalytic


activity to hydrocarbon substrates when it was found that nitroxyl radicals
were continuously regenerated during the induction period to oxidation of
polypropylene at 114°C. Berger, Bolsman and Brouwer [88-90] followed up
this observation by a detailed investigation of the catalytic activity of a
range of arylamines and their derived nitroxyls in liquid hydrocarbons. They
obtained fhy means of Eqn (26)

/lAHo] = (Äi)t=t* • (i>0)t " ("A>t àtAv)l_t* (26)

in which [AH0] is the initial inhibitor concentration (y0)t and (i;A)t are the
rates of the uninhibited and inhibited oxidations respectively after the same
amount of oxygen has been consumed and (R)t=t* and (v)t=t* are the rates
of initiation and oxygen consumption immediately after the addition of the
inhibitor.

TABLE 6

Stoichiometric inhibition coefficients if) of substituted diarylamines and their oxidation


products in paraffin oil at 130°C.

Compound X=H X=0' X=OH

jT~\ * r ~ \ 36 26 35
C VN-f VOEt XXXIX
X
N02-<^ V ^ ^ f VN°2 XL
15
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 145

Table 6 taken from their review [901 shows that in the case of an
arylamine (XXXIX R=H) containing an electron releasing group, and which
is therefore a powerful CB-D antioxidant, the catalytic activity of the amine,
nitroxyl and hydroxylamine are all similar, whereas in the case of 4,4,-
dinitrodiphenylamine (XL) which is not an effective CB-D antioxidant, for
reasons discussed in Section 1.2, only the nitroxyl is a catalytic antioxidant.
Katbab and Scott [91,92] have studied a related arylamine (XLI, X=H) in
rubber subjected to mechano-oxidation.

,!, _NHipr
(~y -0 XLI

In this system the rate of initiation is high and the oxidation is diffusion
controlled and it was shown that the corresponding nitroxyl (XLIX=0) is
formed at the expense of amine after a short induction period during which
alkylperoxyl radicals are formed in the system. The nitroxyl concentration
rises to a maximum and then decreases with the formation of the corre­
sponding hydroxylamine XLI (X=OH) to give a stationary state (see Fig. 1)
which persists to the end of the induction period. As in solution, the nitroxyl
and related hydroxylamine behave similarly to the parent amines and in
some cases they are more effective (see Table 7) suggesting that some of the
parent amine may be wasted in side reactions which does not give rise to
nitroxyl (cf Scheme 4).
Radical concentration (g mol/g x 10 10 )

70 80 90
Fatiguing time (h)

Fig. 1. Kinetics of formation and decay of alkylperoxyl and nitroxyl radicals in rubber
during mechanooxidation. A: alkylperoxyl in the presence of HLI, X=H (1 g/100 g); B:
XLI,R= - O (g = 2.00168, N = 8.33). (Reproduced by kind permission from Chem. Ind.
(1980) 573.)
146 GERALD SCOTT

TABLE 7

Effectiveness of diarylamines and derived oxidation products as mechanoantioxidants in


natural rubber (cone111 g/100 g)

Compound Hours to failure*

X=H X=0 X=OH

f\i
X
f VNHiPr 274 337

MeOH; \-N-f VoMe 207 441 56+

t 0 c t / V N / \ t O c t 33 gg 107

* Control without additive, 20 h.


+ Additive insoluble, rapidly lost to the surface.

The mechanism proposed by Bolsman et ah, unlike Denisov's stabilisa­


tion of alcohols and amines (Scheme 9), involves the oxidation of alkyl
radicals rather than alkylperoxyl radicals. This is represented in gener­
alised form to take into account subsequent studies in polymers in Scheme
10. In this scheme the parent amine is not itself involved in the catalytic
CB-A/CB-D cycle but has to be oxidised to the nitroxyl before it can partici­
pate in the cycle. Scott and co-workers have shown that the formation of
unsaturation in the substrate is an inevitable consequence of this process
and in the example referred to in Fig. 1 and Table 7, the formation of
conjugated unsaturation has been shown in rubber [92] and in the structur­
ally related squalene [93].
Berger et al. have shown [90] that the "hindered" piperidines (XLIV) are
also catalytic antioxidants. The rate of hydrogen abstraction by alkylperoxyl
is much lower for hindered piperidines (3 M" 1 s _1 ) Scheme 10, Reaction (a)

ROO^
ROOH >L J< (27)

XLIV XLV
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 147

0 2 , ROO , ROOH, RO*

Scheme 10. Catalytic antioxidant mechanism (CB-A/CB-D) of nitroxyl radicals.

[94], than for diarylamines (10 3 -10 6 M _ 1 s~1R) and they do not themselves
act as CB-D antioxidants. The amines and nitroxyls have again similar
stoichiometric inhibition coefficients (see Table 8), confirming that nitroxyl
is the effective antioxidant.
Nitroxyls have also been shown to be involved in the stabilisation of
polyolefins both in light and in the polymer melt. A detailed discussion of
the technology will be deferred until Volume II but the catalytic cycle
outlined in Scheme 10 broadly describes the catalytic activity of nitroxyl
radicals under these quite different conditions of autoxi dation and much of
the kinetic data discussed in the following section has come from fundamen­
tal studies related to these technological phenomena.

3.2 Chemistry of Nitroxyls

Although the stable "nitroxides" have been known for many years, most
of our knowledge of their chemistry is due to Rozantzev and his co-workers
[95] who have shown that these species are stable enough to undergo a wide
range of chemical reactions leaving the nitroxyl group intact. Several com­
prehensive reviews are available on the formation and reaction of nitroxyl
148 GERALD SCOTT

TABLE 8

Stoichiometric inhibition coefficients (/) for hindered piperidines and derived nitroxyls

f
X=H X=0

MeJ
oi I
jMe XLVI 420 510

MeJ 1 Me XLVII
400 410
Me^N^Me
I
X

radicals [96,97] but only those aspects of their chemistry which are relevant
to their antioxidant function will be discussed here.
The aryl nitroxyls and the alkyl nitroxyls differ considerably in their
stability, particularly in autoxidising substrates. The stability of both spe­
cies depends on the délocalisation of the unpaired electron on oxygen and
nitrogen.

N-O < ► N-Ö


/ r
XLVIIIa XLVIIIb
However, in the aryl nitroxyls the electron can be further delocalised in
the aromatic ring, e.g. XLIX, leaving it susceptible to attack by alkylperoxyl
[96]. The occurrence of quinones among the oxidation products of nitroxyls
is almost certainly due to alkylperoxyl attack in the aromatic ring [90,98].
The fully a-substituted nitroxyl radicals of which the 2,2,6,6-tetramethyl
piperidinoxyls are among the most stable do not undergo alkylperoxyl
attack and Berger and co-workers have suggested [901 that the higher f
values of the hindered piperidinoxyls (cf. Tables 6 and 8) are due to this fact.
The stability of the fully substituted nitroxyls is not due simply (as is often
implied in the literature) to protection of the nitroxyl by the steric bulk of
the alkyl groups since the contribution of resonance structure XLVIIIa and
XLVIIIb provides the basic stability. The instability of the less hindered
nitroxyls is due to the presence of at least one hydrogen on the a-atom (e.g.
in XLIX) which makes possible the disproportionation reaction (28a).
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 149

Me Me ^ I ^Me
Me H N Me

OH
LI
(28)

XLIX
A good deal of information is now available from recent publications on
the rate constants for the individual reactions in Scheme 10 [101,102].
An area of uncertainty centres on the mechanism by which the nitroxyl is
regenerated. This in turn poses the question of whether the free hydroxy­
lamine (N-OH) or the alkyl hydroxylamine (NOR) is the main product of the
CB-A process since both of these can in principle be considered reservoirs for
the nitroxyl (N-O). A further factor which has to be considered is whether
Reactions b and/or d in Scheme 10 can compete with oxygen.
Carlsson, Gratton and Wiles [94] have discussed the last question and
have shown that in cyclohexane initiated by tert-butoxyl radicals at 25°C, kd
is 4-108 M _1 s"1 and compares with values obtained by other authors using
different alkyl radicals at the same temperature (4-840 8 M"1 s _1 ). kjkc is
about 0.05 and is consistent with results obtained in other liquid hydrocar­
bons (see Table 9).
Shlyapintokh and Ivanov [10] and Maslov and Zaikov [102] have reviewed
the extensive investigations carried out in recent years in the Soviet Union.
Table 9 summarises the values offtd//ecobtained in a variety of substrates
and with a number of derivatives of 2,2,6,6-tetramethyl piperidinoxyl
XXXVIII. Although there are some obvious anomalies which probably result
from the different methods used by different workers, it seems clear that the
critical ratio can be one or even greater. It should be remembered, however,
that Reaction c in Scheme 10 must occur if the nitroxyl is to be regenerated.
Under stationary static conditions, the rate of Reaction c has to be equal to
the sum of the rates of reactions which oxidise the hydroxylamines (Reac­
tions f j) or alkyl hydroxylamine Reaction h, to nitroxyl. The concentration
of >N-0 and >NOH + >N-OR will adjust accordingly.
There is no doubt that at temperatures above 130°C the catalytic reaction
occurs by the sequence b and fin Scheme 10. Berger et al. have shown that
above 80°C the rate of Reaction g is so fast that the sequence b, f cannot be
distinguished from d,e,f. Scott and co-workers [2,113-116] have found that
150 GERALDSCOTT

TABLE 9

Values of Äd/^c obtained with hindered piperidinoxyl XXXVIII, R = benzoyloxy

Source of alkyl radical R in XXXVIII T(K) kà/kc Reference

Cumene H 333 0.056 104


Methyl methacrylate H 323 0.32 105,106
n-Butylmethacrylate H 323 0.73 105,106
iso-Butylmethacrylate H 323 0.74 105,106
n-Nonylmethacrylate H 323 1.20 105,106
Methylacrylate H 323 2.20 105,106
n-Butylacrylate H 323 5.20 105,106
Styrene H 323 0.1 105,106
Cumene OH 333 0.034 107
Polypropylene OH 353 0.20 108
Polypropylene OH 358 2.50 107
Propylene OCOPh 387 0.25 109
Polybutadiene OCOPh 293 0.10 110
Cyclohexylmethyl ether H 348 0.32 104
Styrene -0 338 0.1 110
Ethylbenzene =0 333 0.038 111
n-Octane =0 273 1.73 112
n-Octane =0 373 0.13 112
n-Octane OH 373 0.161 112

at temperatures above 180°C in polyolefîns, regeneration of nitroxyl occurs


entirely through the sequence b and f as evidenced by the rapid formation
of unsaturation in the polymer [114,116] and the absence of polymer bound
alkylhydroxylamine [116]. Furthermore, the hydroxylamines are virtually
indistinguishable from the nitroxyls as antioxidants under these conditions
[2,114]. The rate constant for Reaction fis extremely high (540 5 1 mol -1 s"1
at 65°C) [73] and the alkyl hydroxylamines compare with the hindered
phenols as CB-D antioxidants. They also react, although at a lower rate,
with oxygen {k} 0 2 = 13-10" 1 mol"1 s"1 at 130°C) [46] and with hydroperox-
ides (fcj ROOH = 0.461 mol"1 s"1 at 130°C) [101]. This is why hydroperoxides
cannot normally be detected in systems stabilised with nitroxyl radicals.
Consequently the "hindered" hydroxylamines can replace the corresponding
nitroxyls in the catalytic cycle without loss of activity [86,90-92,110,113]
(see Tables 6 and 7).
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 151

The involvement of alkylhydroxylamines at lower temperatures is much


more complex and is still not fully understood. Berger et al. have shown [90]
that even below 80°C there is always a contribution from the direct reaction
of nitroxyl with alkyl to give olefin without the intervention of N-OR (see
Scheme 10, Reaction b). k\/k& remains constant at 0.21 between 60°C and
80°C. This is not surprising since nitroxyls are known to hydrogen abstract
even from saturated hydrocarbons [117] and the ßC-H bond in a free radical
is only about one third of the strength of a normal C-H bond [2]. On this
basis, unsaturated groups that stabilise the developing unsaturation in the
transition stage, e.g. LIV, might be expected to favour Reaction b and
increase the catalytic activity of nitroxyls. The results of Berger et al. [90]
at 130°C in hydrocarbon substrates (see Table 9) support this conclusion.
This shows quite clearly that the presence of a ß-phenyl group markedly
increases /^relative to the aliphatic hydrocarbons. Deuteration of the ß-carb-
on atom decreases f, as expected on the basis of Reaction (29).

/ Y - C H — CH — H + O—N > / \—CH —CH —H Ö—N^


~~ R R LIV
(29)

CH = CH+HO — Ne
I ^
R

The presence of a relatively small proportion of diphenylmethane which


gives a radical without a ß-hydrogen which therefore cannot lead to re­
generation of the hydroxylamine, gives a much lower value of f(see Table
10). Diphenylethane, on the other hand, which can regenerate >NOH,
actually gives a higher value of fin paraffin oil. Taken together, these facts
provide very convincing evidence for the catalytic mechanism involving
reactions b and f (Scheme 10) at high temperatures and are entirely con­
sistent with results obtained in polymers at processing temperatures (150°-
220°C). Here stationary conditions do not hold and complementary fluctua­
tion of nitroxyl and hydroxylamine has been observed [2,114-116]. This will
be discussed in detail in Volume II (Chapter 3).
Reaction e (Scheme 10) occurs slowly even at 20°C and a decrease in the
strength of the C - 0 bond by electron délocalisation on the alkyl group and
increased steric hindrance in LIII between the substituents a to nitrogen
and Rlf R2 and R3, remarkably increase ke. Typical values at 20°C, are [118]
R, - R2 = CH3, R3 = CN, ke = MO"7 s"\ Ri = R2 - CH3, R3 = Ph, ke = 610" 7
s~ ; Rx = CH3, R2 = R3 = Ph, ke = 340 s~ and isobutene was shown to be
formed from tri-ter£-butyl hydroxylamine above 100°C [90]. The rate con­
stant at 100°C is 0.440" 4 s"1 and the activation energy between 100° and
152 GERALD SCOTT

TABLE 10

Effect of hydrocarbon structure on/*at 130°C in the presence of XXXVIII

6
Hydrocarbon R in XXXVIII Cone11 10 mol"1 f

Paraffin oil PhCOO 630


n-Hexylbenzene PhCOO 3900
ß,ß-Dideutero-n-hexyl benzene PhCOO 1100
Paraffin oil =0 410
n-Hexylbenzene =0 720
Paraffin oil H 20 510
Paraffin oil + diphenylmethane* H 20 190
Paraffin oil + 1,1-diphenylethane* H 20 580

10 wt%.

130°C was found to be 96.2 kJ mol -1 . Berger et al. have suggested that
higher molar mass tert-aïkyl substituents such as are found in paraffins
(and polymers) also increase the value of ke [90].
Wiles et al. [118] and Kovtun and co-workers [104] have shown that
nitroxyl is slowly regenerated from alkylpiperidinoxyls, LIII, at 25°C in the
presence of oxygen and other radical acceptors.

R - ^ ^ N - 0 - C - R 3 LIII
jNtfe R2
Me

Reaction e must be the rate determining step in each case (see Table 11)
and cannot by itself be a termination process. It must be succeeded by one
of the alternative processes which remove either alkyl (e.g. b, Scheme 10) or
alkylperoxyl. Reaction h 0 is considerably slower than reaction h R 0 at 25°C
(see Table 11) which is generally believed to be the main process by which
nitroxyl is regenerated at low temperatures [101,102,119].
Two alternative mechanisms have been suggested for reaction h R 0 0 . The
first, due to Denisov and his co-workers [86], involves attack of alkylperoxyl
on hydrogens on the ß carbon atom in the alkylhydroxylamine (Reaction
(30)).
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 153

TABLE 11

Rate constants and characteristic times (t) for the oxidation of l-(2'-cyano-2'-propoxy)-4-
hydroxy-2,2,4,4-tetramethyl piperidine (LV, R = OH, Ri, R2=pentamethylene, R3 = H)
[101]

Temperature ÄhROO* ThROO* Ah02 Th02


°C 1 mol"1 s"-1 hr 10V1 hr

25 0.032 lxlO-3 2.3 1.2xl04


80 9.5 0.03-30 4000 7

RH R
| 1
ROO 1
>N-OC-C- > >N--0- + C= C- (30)
1 1 11 11
1 1
R2R3 R2 R3
However, Kovtun and his co-workers showed [104] that in the reaction of
(CH3)2(CN)COO with a variety of fully substituted piperidines, the rate of
regeneration of nitroxyl from the phenyl hydroxylamine LIV was similar
(ÄhROO* = 11 mol"1 s"1) to compounds of the type LIII even though LIV does
not have a hydrogen ß to oxygen. They suggested direct substitution of
alkylperoxyl at the N - 0 bond (Reaction (31a)).

Me

N—O-/ \ LIV

Me

Although the dialkylperoxide, ROOR, will be relatively stable below


100°C in the dark, Reaction (31a) cannot be considered to be a termination
process above 130°C (e.g. during polymer processing), or in light, since it
homolyses readily to give new free radicals. However, Klemchuk et al. have
shown the formation of disproportionation products, Reaction (31b), prob­
ably by a non-radical mechanism [103,124]. Qualitative evidence from a
number of sources [94,101,109] indicates that hydroperoxides and peracids
can react with alkylhydroxylamines (reaction h R00 HX and that this reaction
may contribute toward the regeneration of nitroxyl. However, this process
may be complicated by the decomposition of hydroperoxides to give alkoxyl
or the thermolysis of the alkylhydroxylamine itself under the conditions
154 GERALD SCOTT

examined. During photo-oxidation, photolysis of both dialkyl peroxides and


alkyl hydroperoxides have to be considered.

VOR 2^22* V O - + ROOR'


/ \ (a) /
\ (b)
R'OO X
N - 0 • + R'OH + R= O
/

In conclusion then, at temperatures above 100°C, the most likely


sequence of reactions leading to the catalytic CB-A/CB-D mechanism of
nitroxyl radicals is the sequence b,f in Scheme 10. The absence of hy­
droperoxides suggests that JROOH a n c * JRO m a y contribute to the regenera­
tion of nitroxyl. At lower temperatures in the dark, the main sequence is
probably d, h R Q.but the evidence does not rule out a contribution from b,f
and d, g, f. Regeneration of nitroxyl may also involve h R 0 0 H , particularly
when ROOH is a peracid.

1.3.3 Other "stable"radicals as catalytic CB-A/CB-D antioxidants


There is now a good deal of evidence that other stable radicals which are
capable of being reversibly reduced and reoxidised, can behave as catalytic
antioxidants in the same way as the nitroxyls at high temperatures. These
studies have mainly been in molten polymers at relatively high tempera­
tures under conditions of mechano-oxidation. This will be discussed in detail
in Volume II (Chapter 3), but in view of their similarity to the nitroxyls, the
salient features of the chemistry will be reviewed here.
It was seen earlier (Section 1.1) that some phenoxyl radicals are stable for
long periods of time. Galvinoxyl was found to be an effective processing
stabiliser for polyolefins by Henman [120] and subsequent work by Scott
and co-workers [22] showed that it is rapidly reduced to the corresponding
phenol LV in an oxygen deficient environment. There is no evidence in this
case for radical coupling; all the galvinoxyl (G,X) can be accounted for either
as the reduced or oxidised form (Reaction (32)).

-o-
tBu tBu tBu tBu
ΠHO
*° \3~ KZ/ ° ^
tBu tBu tBu
\3" tBu
G,X LV (32)
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 155

The catalytic antioxidant mechanism can be represented generally as in


Scheme 11 which illustrates the competition of the catalytic redox process
with the main autoxidation cycle. Other redox couples which also behave as
efficient mechanoantioxidants by the same mechanism are iodine and cupric
salts (see Scheme 11). A similar catalytic antioxidant effect of copper has
been observed in the photostabilisation of linseed oil, where a copper alkyl,
analogous to the alkylhydroxyamines may also be formed as an interme­
diate. The mechanism, which is outlined in Scheme 12, bears a resemblance
to that proposed by Denisov to account for the antioxidant activity of copper
in cyclohexanol although the oxidation product in that case is cyclohexanone
(see Section 1.3). In linseed oil the oxidation product is a conjugated polyene
and is the source of the brown discolouration that occurs in old paintings
pigmented with cupric acetate (verdigris).
It has long been known that copper ions are effective stabilisers for
polyamides at very low concentrations. It seems probable that the mecha­
nism involved is similar to that for copper in paint medium. Denisov and

o2
\
\\

;c=c: +H +

ROOH

/ \
\
ROOH RH
A- AH

NO NOH
G GH
I IH
Cu2+ Cu+
Scheme 11. General catalytic chain-breaking mechanism of antioxidant action.
156 GERALDSCOTT

—CHCH=CHCH,— —CH=CHCH=CH—

I
± —CHCH=CHCH,—

Cu++H+

ROOH

OOH OO-
I I
—CHCH=CHCH,— -CHCH=CHCH2
J
Scheme 12. Copper catalysed oxidative conjugation of unsaturated oils.

co-workers [123] have also reported the catalytic antioxidant activity of I 2 in


the oxidation of cumene. As in the case of polypropylene, this mechanism is
very oxygen pressure dependant. Denisov prefers the formation of alkyl
iodide and regeneration of iodine atoms by alkylperoxyl attack as proposed
for the regeneration of nitroxyl from alkylhydroxylamine,
ROO. + RI ROOR + I- (33)
However in polypropylene at higher temperatures (180°C), olefinic unsat-
uration is a major product and peroxides cannot be measured during the
reaction due to the facility of their reduction by HI. Scheme 11 is preferred
in this case.

REFERENCES
1 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, Chapters 4 and
5.
2 G. Scott, in G. Scott, (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science, London, 1985, Chapter 2.
3 J. Pospisil, in G. Scott, (Ed.), Developments in Polymer Stabilisation-1, Applied
Science, London, 1979, Chapter 1.
4 J. PospiSil, in G. Scott, (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science, London, 1985, Chapter 1.
5 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, p. 1323 et seq.
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 157

6 J.K. Beconsall, S. Clough and G. Scott, Proc. Chem. Soc. (1959) 308.
7 J.K. Becconsall, S. Clough and G. Scott, Trans. Faraday Soc, 56 (1960) 459.
8 A. Tkâ£, in G. Scott, (Ed.), Developments in Polymer Stabilisation-8, Elsevier
Applied Science, London, 1987, p. 101 et seq.
9 H.R. Forrester, J.M. Hay and R.H. Thompson, Organic Chemistry of Stable Free
Radicals, Academic Press, London, 1968.
10 K.J. Kochi, in K.J. Kochi (Ed.), Free Radicals, Vol. II, J. Wiley & Sons, New York,
1973, p. 665.
11 E.R. Altwicker, Chem Rev., 67 (1967) 475.
12 J. Pospisil, Pure Appl. Chem., 36 (1973) 207.
13 J. Kovarova-Lerchova and J. Pospisil, Eur. Polym. J., 14 (1978) 463
14 J. Kovarova-Lerchova, J. Pilar, G. Samay and J. Pospisil, Eur. Polym J., 14 (1978)
601.
15 L. Tamir, H. Pivcoca and J. Pospisil, Collect. Czech. Chem. Comm., 37 (1972) 1912.
16 J. Kovarova and J. Pospisil, Eur. Polym. J., 13 (1977) 975.
17 J. Lerchova and J. Pospisil, Chem. Ind., (1975) 516.
18 J. Lerchova and J. Pospisil, Angew. Makromol. Chem., 38 (1974) 191; 39 (1974) 107.
19 I. Buben and J. Pospisil, Collect. Czech. Chem. Comm., 40 (1975) 987.
20 I. Buben and J. Pospisil, Collect. Czech. Chem. Comm., 40 (1975) 977
21 H. Lind, T. Winkler and H. Loeliger, J. Polym Sei., Polym Symp., 57 (1976) 225.
22 J. Lerchova, L. Kotulak, J. Rotschova, J. Pilar and J. Pospisil, J. Polym. Sei., Polym.
Symp., 57 (1976) 229.
23 K.B. Chakraborty and G. Scott, Eur. Polym. J., 13 (1977) 1007.
24 K.B. Chakraborty and G. Scott, Polym. Deg. Stab., 1 (1979) 37.
25 G. Scott and M. F. Yusoff, Polym. Deg. Stab., 2 (1980) 309.
26 K. Ley, E. Müller and W. Schmidhuber, Angew. Chem., 70 (1958) 460.
27 E. Müller, K. Ley and G. Schlechte, Ber., 90 (1957) 2660.
28 E. Müller, A. Schick and K. Scheffler, Ber., 92 (1959) 474.
29 C. Golumbic, Oil Soap, 23 (1946) 184.
30 F.A. Neuberger and S. Bamberger, Chem. Ber., 107 (1974) 1788.
31 M. Shanshal, Z. Naturforsch., 28a (1973) 1892.
32 G.S. Shonov, Zh. Fiz. Khim., 46 (1927) 324.
33 R. Hoskins, J. Chem. Phys., 25 (1956) 788.
34 V. Cholvald and A. Tkâc, Collect. Czech. Chem. Comm., 46 (1981) 1071.
35 G.M. Coppinger and J.D. Swalen, J. Am. Chem. Soc., 83 (1961) 4900
36 A.A. Katbab and G. Scott, Chem. Ind. (1980) 573.
37 A.A. Katbab and G. Scott, Eur. Polym. J., 17 (1981) 559.
38 J. Lerchova and J. Pospisil, Angew. Makromol. Chem., 38 (1974) 191, p. 89 et seq.
39 D.F. Bowman, B.S. Middleton and K.U. Ingold, J. Org. Chem., 34 (1969) 3456.
40 R.F. Bridger, J. Am. Chem. Soc, 94 (1972) 3124.
41 K. Adamic, M. Dunn and K.U. Ingold, Can. J. Chem., 47 (1969) 287, 295.
42 A. Calder and A.R. Forrester, J. Chem. Soc. (C) (1969) 1459.
43 M.B. Neiman and E. G. Rozantzev, Izv. Akad. Nauk. SSSR, Ser. Khim., (1964) 1178.
44 T.A.B.M. Bolsman, A.P. Blok and J.H.G. Frijns, Rec. Trav. Chim. Pays Bas, 97 (12)
(1978) 310.
45 A. Tkâc, in G. Scott, (Ed.), Developmnets in Polymer Stabilisation-8, Elsevier
Applied Science, London, 1987, p. 95.
46 H. Berger, T.A.B.M. Bolsman and D.M. Brouwer, in G. Scott (Ed.), Developments in
158 GERALD SCOTT

Polymer Stabilisation-6, Applied Science Publishers, London, 1983, p. 5 et seq.


47 L.P. Nethsinghe and G. Scott, Rubber Chem. Technol., 57 (1984) 779.
48 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, p. 241 et seq.
49 B.W. Evans and G. Scott, Eur. Polym. J., 10 (1974) 453.
50 R.P.R. Ranaweera and G. Scott, Eur. Polym. J., 12 (1976) 591.
51 J.L. Bolland, Trans. Faraday Soc, 44 (1948) 669.
52 J.L. Bolland and P. ten Have, Trans. Faraday Soc, 43 (1947) 201.
53 J.L. Bolland and P. ten Have, Discuss. Faraday Soc, 2 (1947) 252.
54 D.S. Davis, H.L. Goldsmith, A.K. Gupta and G.R. Lester, J. Chem. Soc, (1956) 4926.
55 H.A. Mattill, J. Biol. Chem., 90 (1931) 141
56 A.R. Bickel and E.C. Kooyman, J. Chem. Soc. (1956) 2215.
57 A.F. Bickel and E.C. Kooyman, J. Chem. Soc. (1957) 2217.
58 Yu.A. Shlyapnikov, in G. Scott (Ed.), Developments in Polymer Stabilisation-5,
Applied Science Publishers, 1982, Chapter 1.
59 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, p. 122.
60 L.F. Fieser, J. Am. Chem. Soc, 52 (1930) 5204.
61 G. Penketh, J. App. Chem., 7 (1957) 512.
62 C.E. Boozer, G.S. Hammond, C.E. Hamilton and J.N. Sen, J. Am. Chem. Soc, 77
(1955) 3238.
63 J.R. Shelton and W. L. Cox, Ind. Eng. Chem., 46 (1954) 816; Rubber Chem. Technol.,
27 (1954) 671.
64 O. Lorenz and C. R. Parks, ACS Meeting, Louise ville, April 1961; Rubber Age,
(1961) 985.
65 C.J. Pedersen, Ind. Eng. Chem., 48 (1956) 1881.
66 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, p. 144 et seq.
67 C.E. Boozer and G.S. Hammond, J. Am. Chem. Soc, 76 (1954) 3861
68 C.E. Boozer, G.S. Hammond, C.E. Hamilton and J.N. Sen, J. Am. Chem. Soc, 77
(1955) 3233.
69 J.R. Shelton and E.T. McDonel, J. Polym. Sei., 32 (1958) 75.
70 K.U. Ingold and I.E. Puddington, Ind. Eng. Chem., 51 (1959) 1322.
71 K.U. Ingold, Chem. Rev., 61 (1961) 563.
72 S.M. Kavun and A.C. Buhachenko, Vysok. Soed., 9B (1967) 661.
73 J.T. Brownlie and K.U. Ingold, Can. J. Chem., 45 (1967) 2419, 2427.
74 G. Scott, in G. Scott (Ed.), Developments in Polymer Stabilisation-4, Applied Science
Publishers, London, 1981, p. 181.
75 D.F. Williams, Biocompatibility of Implant Materials, Vols. I and II, CRC Press,
Florida, 1982.
76 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, 1965, p. 161 et seq.
77 K.B. Chakraborty and G. Scott, J. Polym. Sei., Polym. Lett. Ed., 22 (1984) 553.
78 N.A. Milas, Chem. Rev., 10 (1932) 285.
79 A.F. Bickel and W.A. Waters, J. Chem. Soc, (1950) 1764.
80 E.Y. Reztsova, B.G. Lipkena and G.L. Slonimskii, Zh. Fiz. Chim., 33 (1984) 656.
81 E.T. Denisov, in G. Scott (Ed.), Developments in Polymer Stabilisation-5, Applied
Science Publishers, London, 1982, p. 23.
82 E.T. Denisov and V.V. Kharitonov, Izv. AN SSSR, Ser. Khim, (1963) 2222.
83 E.T. Denisov, Izv. AN SSSR, Ser. Khim, (1969) 327.
84 E.T. Denisov, N.L Mitskevich and V. E. Agsbekov, Liquid Phase Oxidation of
Oxygen-containing Compounds, Consultants Bureau, New York, 1977.
ANTIOXIDANTS: CHAIN BREAKING MECHANISMS 159

85 E.T. Denisov, in G. Scott, (Ed.), Developments in Polymer Stabilisation-3, Applied


Science Publishers, London, 1980, Chap. 1.
86 Yu.B. Shilov, R. M. Battalova and E.T. Denisov, Dokl. Akad. Nauk. SSSR, 207
(1972) 388.
87 Yu.B. Shilov and E.T. Denisov, Vysok. Soed., A16 (1974) 2313.
88 T.A.B.M. Bolsman, A.P. Blok and J.H.G. Frijns, Rec. Trav. Chim. Pays Bas, 98
(1978) 310.
89 T.A.B.M. Bolsman and D. M. Brouwer, Rec. Trav. Ghim. Pays Bas, 97 (1978) 320.
90 H. Berger, T.A.B.M. Bolsman and D.M. Brouwer, in G. Scott (Ed.), Developments in
Polymer Stabilisation-6, Applied Science Publishers, London, 1983, Chapter 1.
91 A.A. Katbab and G. Scott, Chem. Ind., (1980) 573.
92 A.A. Katbab and G. Scott, Eur. Polym. J., 17 (1981) 559.
93 L.P. Nethsinghe and G. Scott, Eur. Polym. J., 20 (1984) 213.
94 D.W. Grattan, D.J. Carlsson and D.M. Wiles, Polym. Deg. Stab., 1 (1979) 69.
95 M. Dagonneau, V.B. Ivanov, E.G. Rozantsev, V.D. Scholle and E.S. Kagan, Rev.
Macromol. Chem. Phys., C22 (1982-83) 169.
96 A.R. Forrester, J.M. Hay and R.H. Thomson, Organic Chemistry of Stable Free
Radicals, Academic Press, London, 1968, p. 180.
97 H.G. Aurich and W. Weiss, Topics in Current Chemistry, Vol 59, Springer-Verlag,
Berlin, 1975, p. 65.
98 M.B. Neiman and E.G. Rozantsev, Izv. Akad. Nauk SSSR., Ser. Khim., (1964) 1178.
99 G. Coppinger and J. Swallen, J. Am. Chem. Soc, 83 (1961) 4900.
100 E.G. Rozantsev and R.S. Burmistrova, Dokl. Akad. Nauk SSSR, 166 (1966) 129.
101 V.Ya. Shlyapintokh and V.B. Ivanov, in G. Scott (Ed.), Developments in Polymer
Stabilisation-5, Applied Science Publishers, London, 1982, Chapter 3.
102 S.A. Maslov and G.E. Zaikov, in G. Scott (Ed.), Developments in Polymer Stabilisa­
tion^, Elsevier Applied Science, London, 1987, Chapter 1,
103 P.P. Klemchuk, M.E. Gande and E. Cordola, Polym. Deg. Stab., 27 (1990) 65.
104 G.A. Kovtun, A.L. Aleksandrov and V.A. Golubev, Izv. Akad. Nauk. SSSR., Ser.
Khim., (1974) 2197.
105 E.M. Pless and A.L. Aleksandrov, Izv. Akad. Nauk. SSSR., Ser. Khim, (1977) 753.
106 A.L. Aleksandrov, E.M. Pless and V.F. Shuvalov., Izv. Akad. Nauk. SSSR., Ser.
Khim., (1979) 2446.
107 G. Baient, T. Kelen, F. Tudos and A. Renek, Polym. Bull., 1 (1979) 647.
108 I.A. Shlyapnikova, V.A. Rogenskii and V.B. Miller, Vysok. Soed., 21B (1979) 521.
109 Yu.B. Shilov and E.T. Denisov, Vysok. Soed., A16 (1974) 2313.
110 M.Yu. Shlyapintokh, S.G. Burkova, A.B. Shapiro, E.G. Rozsntsev and V.B. Ivanov,
Donaulandergesprsch über die naturliche und kunstliche, Alterung von Kunstsof­
fen, Dobruunik, 1978.
111 M.K. Khloplyankina, A.L. Buchachenko, M.B. Neiman and A.G. VasiPeva, Kinetika
i Kataliz, 6 (1965) 394.
112 M.F. Romantsev, N.V. Lysun, M.B. Sudnik and G.F. Pavelko, Khimiya Vysok.
Energii, 6 (1972) 100.
113 R. Bagheri, K.B. Chakraborty and G. Scott, Polym. Deg. Stab., 4 (1982) 1.
114 R. Bagheri, K.B. Chakraborty and G. Scott, J. Polym. Sei., Polym. Chem. Ed, 22
(1984) 1573.
115 J.B. Adeniyi, S. Al-Malaika and G. Scott, J. Appl. Polym. Sei., 32U (1986) 6063.
116 S. Al-Malaika, E.O. Omikorede and G. Scott, J. Appl. Polym. Sei., 33 (1987) 703.
160 GERALD SCOTT

117 A.I. Bogatyreva, A.G. Sklyarova and A.L. Buchachenko, Khim. Vylok. Energ., 5
(1971) 37.
118 D.W. Grattan, D.J. Carlsson, J.A. Howard and D.M. Wiles, Can. J. Chem., 57 (1979)
2197.
119 D.J. Carlsson, A. Garton and D.M. Wiles, in G. Scott, (Ed.), Developments in
Polymer Stabilisation-1, Applied Science Publishers, London, 1979, p. 246 et seq.
120 T.J. Henman, in G. Scott, (Ed.), Developments in Polymer Stabilisation-1, Applied
Science Publishers, London, 1979, Chapter 2.
121 R. Bagheri, K.B. Chakraborty and G. Scott, Chem. Ind., (1980) 865
122 R. Bagheri, K.B. Chakrabarty and G. Scott, Polym. Deg. Stab., 5 (1973) 81.
123 A.L. Alexandrov, T.I. Sapacheva and E.T. Denisov, Zh. Fiz. Chim., 44 (1970) 1122.
124 P.P. Kemchuk and M.E. Gande, Polym. Deg. Stab., 22 (1988) 241.
161

Chapter 5

ANTIOXID ANTS — PREVENTIVE MECHANISMS

S. AL-MALAIKA

It was seen in Chapters 1-3 that by far the most important class of radical
generators is hydroperoxides since they are the universal products of the
reaction of oxygen of the atmosphere (including singlet oxygen and ozone)
with hydrocarbon substrates. The critical role of hydroperoxides in auto-
xidation emphasises the importance of antioxidants which remove or deac­
tivate this radical generating species. This antioxidant mechanism, which
falls within the class of preventive antioxidants (see Ref. 1 for a fuller
discussion), has to be distinguished from agents which decompose hy­
droperoxides homolytically to free radicals, for example metal ions, or UV
light. Table 1 shows the effect of heat, light, metal ions, and ozone on
hydrocarbon substrates [1]. This confirms the importance of hydroperox­
ides; the different modes of decomposition probably account for the major
proportion of the free radicals introduced into hydrocarbon oils and poly­
mers during oxidation.
Preventive mechanisms, therefore, also include the deactivation of metal
ions and ozone, the absorption or screening of UV light as well as the
decomposition of hydroperoxides to non-radical products. The importance of
each of the different preventive mechanisms may vary depending both on
the chemical constitution of the substrate and the conditions to which it is
subjected (heat and metal ions are more important for lubricating oils and
fuels; heat, metal ions and ozone for rubbers; light and heat for high
molecular weight saturated hydrocarbon polymers; and light for polymers
containing UV absorbing groups). The first two mechanisms (deactivation
of metal ions and ozone and absorption of UV light) are palliative and are
therefore only of limited effectiveness, since they retard rather than inhibit
the generation of free radicals from hydroperoxides. The decomposition of
hydroperoxides by chemical reactions which do not lead to the formation of
free radicals (the peroxidolytic mechanism), is, therefore, by far the most
important preventive mechanism of antioxidant action.
Peroxidolytic antioxidants range from the simple stoichiometric peroxide
decomposers such as phosphite esters (I) through a whole family of sulphur-
containing compounds such as mono sulphides, e.g. (II) and metal complexes
TABLE 1

Reactions responsible for initiation of autoxidation

Reactant Radical product


A B C D
Heat Light Metal ions Ozone

M+ M2+

1. Saturated R- + H- R- + H- — R- (+ H+) RO- + OOH


hydrocarbons (RH)

2. Unsaturated RCH=CHCHR' + H- RCH=CHCHR + H- — RCHCHCH2R'


hydrocarbons and or RCH=CHCHO- *
conjugated olefins RCH-CHCH2R' + OOH-
(RCH=CHCH2R')

3. Chlorinated R- + Cl- R- + Cl- — — —


hydrocarbons (RC1)

4. Carbonyl compounds RCO + -R' R


and conjugated )C-0
carbonyl compounds
(RCORO or RCO- + -R'

5. Hydroperoxides RO- + -OH or RO- + -OH RO- + OH ROO- (+ H+)


((2) ROOH) ROO- + H 2 0 + OR

*The destructive influence of ozone on rubber is not due to this reaction although it probably occurs. This will be discussed in detail
in a later section.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 163

of dithioic acids, e.g. the metal dithiocarbamates (III) which act primarily as
catalytic peroxidolytic antioxidants. Most sulphur-containing compounds
act by other mechanisms as well such as chain-breaking or in some cases U V
screening activity, hence exhibiting "autosynergism". This will be discussed
in a later section in this chapter.

(RO)3P (ROCOCH 2 CH 2 ) 2 S [R2NCS2]„M

(I) (ID (HD

The study of peroxidolytic mechanism began independently in a number


of different technologies. It was first observed in hydrocarbon oils in the
early 1940s. It was shown that unrefined lubricating oils are much less
sensitive to oxidation than are the purified components which comprise
them and this was attributed to the presence of antioxidant-active naturally
occurring sulphur impurities in the oil. The early literature has been re­
viewed in the first edition of this work [1] and is summarised in Chapter 1
of this volume and will not be considered in detail here except to note that
the Group II dithiophosphate complexes (IV) (e.g. the zinc dialkyl dithio-
phosphates) constitute one of the most important classes of antioxidant/an-
ticorrosive agents for most modern lubricating oils. About the same time
dialkyl monosulphides were found to be effective as antioxidants in food­
stuffs [2] and this was followed by the acceptance of a number of esters of
thiodipropionic acid (II, R = alkyl) as additives for foodstuffs by the US
government in 1949 [3]. In the field of polymers the most important dis­
covery was also made in the early 1940s when it was demonstrated that
rubbers cured with thiuram disulphides (V) in the absence of sulphur but in
the presence of zinc oxide gave vulcanisâtes with a high level of resistance
to thermal oxidation [4]. This was shown [5,6] to be associated with the
formation of the corresponding zinc dialkyl dithiocarbamates (III, M = Zn)

[(RO)2PS2]„M (R 2 NCS 2 )

(IV) (V)

In the late 1950s, the dithiocarbamates (III), like the metal dithio-
phosphates (IV), were found to be effective antioxidants for lubricating oils
[7] and similar activity was shown for the dithiophosphates in rubbers [8].
With the later development of the polyolefins, the metal dithiocarbamates
(III) and metal dithiophosphates (IV) were found to be equally effective as
antioxidants in these substrates [9]. The simple monosulphides (II) were
found not to be very effective when used alone in polyolefins, although in
combination with chain-breaking antioxidants they became the basis of
most modern stabilising systems [10,11]. The use of these peroxidolytic
164 S.AL-MALAIKA

antioxidants in polymers will be discussed in Volume 2. In this chapter


mechanisms of peroxide decomposers will be discussed along with other
preventive mechanisms. Synergistic effects will be discussed in a later
section of this chapter.

1. CATALYTIC PEROXIDOLYTIC MECHANISMS AND THE ROLE OF


SULPHUR-CONTAINING COMPOUNDS

The pioneering studies of Denison [12] in the early 1940s and later of
Kennerley and Patterson [13] led to the recognition of a variety of peroxido-
lytic agents which function by a catalytic mechanism (PD-C). About the
same time Hock and Lang [14] suggested a catalytic decomposition of
hydroperoxides under the influence of electrophilic reagents, e.g. mineral
acids or Lewis acids. Some years later, fundamental studies by Kharasch
and his co-workers [15] demonstrated that the behaviour of aralkyl hy­
droperoxides (e.g. cumyl hydroperoxide, CHP, VI) in the presence of acidic
catalysts is characteristically different from its decomposition in the pre­
sence of transition metal ions. Analysis of the products formed in the
catalytic decomposition of CHP has proved to be an invaluable diagnostic
technique in the study of the mechanism of the peroxidolytic reactions (see
Chapter 1, Scheme 3). As will be seen in later sections, not only does it
distinguish between the two extreme types of catalytic activity, but when
the contribution of the two alternative processes varies, as frequently hap­
pens, depending on the experimental conditions, e.g. molar ratio of hy­
droperoxide to sulphur compound, it allows an estimate to be made of the
relative contribution of each.

1.1 Mechanisms of Simple Sulphides

The characteristic behaviour of sulphides in autoxidation was apparent


from early studies in lubricating oils which showed long-term oxidative
stability due to the presence as impurities of small amounts of naturally
occurring sulphur compounds [12]. In sulphur-vulcanised rubbers it was
found that increasing sulphur content increased the initial oxidation rate
and resulted in rapid loss in mechanical properties of the vulcanisate [16].
In PP, a very high level of synergism was observed between dialkyl sul­
phides, e.g. thiodipropionate esters, and chain-breaking antioxidants (e.g.
hindered phenols) [17], due to the elimination of an initial pro-oxidant
process involved in the oxidation of the sulphide. These and subsequent
investigations [11,18,19] showed that the sulphides and other sulphur-con­
taining antioxidants have three features in common:
1. The effective antioxidant is not the parent sulphur compound but an
oxidation product formed from it in the autoxidising medium.
2. In all cases, the effective antioxidant is a catalyst for the ionic decom-
ANTIOXIDANTS — PREVENTIVE MECHANISMS 165

position of hydroperoxides (PD-C)


3. The antioxidant stage is frequently preceded by a pro-oxidant stage
which varies in intensity depending on the structure of the sulphur com­
pound. Hence sulphide inhibited oxidations are generally autoretarding.

(a) The sulphur cross-link in rubber


Comprehensive studies by Bateman and his co-workers [20-23] (at the
British Rubber Producers Research Association, BRPRA) of the behaviour
of a wide range of model sulphides (e.g. unsaturated mono- and di-sulphides)
related structurally to the sulphur cross-link revealed the following.
1. Model sulphides like sulphur vulcanisâtes oxidise in an autoretarding
mode [20], suggesting the formation of an inhibitor during oxidation.
2. Chain branching on the carbon atom to sulphur increases the anti­
oxidant effectiveness of the sulphide [21].
3. Dialkyl disulphides are more effective than dialkyl monosulphides [21].
4. Derived sulphoxides and thiolsulphinates are more effective than the
parent sulphides and disulphides [21].
5. The effectiveness of sulphoxides and thiolsulphinates as antioxidants
appears to be related to their thermal instability [22].
6. All the active sulphur compounds are capable of destroying hydroperox­
ides [23].
7. Sulphoxides form stoichiometric hydrogen-bonded complexes with hy­
droperoxides and the antioxidant activity of the sulphides is associated with
the "deactivation" of hydroperoxides by this mechanism [22].
These experimental conclusions were subsequently confirmed when other
alkyl sulphides and thiodipropionate esters were studied in details by other
workers, notably by Hawkens [42] and Scott [25-27] and later by Shelton
[11,24], and this will be discussed in the following sections.

(b) Simple dialkyl sulphides and esters ofthiodipropionic acids


It was shown as early as 1945 [12] that simple sulphides are themselves
not antioxidants and their peroxidolytic activity depends primarily on their
reaction with hydroperoxides to form sulphoxides and sulphones (and thiol-
sulphinate in the case of disulphide). It was also shown [20,24,28] that the
sulphides are thermally stable but not the derived sulphoxides and thiol­
sulphinates. It was further demonstrated [29] that the antioxidant activity
of these oxidised sulphur compounds is related to their thermal instability
and the formation of more reactive species by decomposition and further
oxidation is involved. There is, however, an optimum stability for a given
system and oxidation conditions (e.g. ease of oxidation of the original sul­
phur compound to the mono-oxy derivative), since the instability required
for activity must be balanced by thermal stability to persist long enough to
perform its function as preventive antioxidant [24]. Marshall [29] investi­
gated the rates of decomposition of hydroperoxides in the presence of
166 S. AL-MALAIKA

dilauryl thiodipropionate, DLTDP (VII, R ■ C12H25) and observed an induc­


tion period (at 30-90°C) consistent with the oxidation of the sulphide to the
corresponding sulphoxide (VIII). Furthermore, the rate of hydroperoxide
decomposition by sulphoxide (VIII) was found to be 8.6 times faster than
with the original sulphide (DLTDP), and at least 20 moles of hydroperoxide
was decomposed per mole of sulphur compound. An important observation
was that in the latter case an induction period was also present suggesting
that the sulphoxide must decompose further to give the active catalytic
peroxide decomposer species.

ROOH
(ROCOCH2CH2)2S (ROCOCH2CH2)2SO (1)

VII, DLTDP VIII

It was clearly demonstrated in the early 1960s [28,30-32] that sulphox-


ides with one or more hydrogens on a carbon ß to the sulphinyl group
undergo thermal decomposition (eis -elimination) at moderate temperatures
to give olefins, which were isolated in most cases, and sulphenic acid
(RSOH), which could not be isolated in pure form. However, in a detailed
kinetic studies (using quantitative NMR) Shelton and Davis [33] followed
the reactive intermediate species formed from the decomposition of di-tert-
butyl sulphoxide (RSOR, R = -C(CH 3 ) 3 ) in different solvents at 80°C and
showed the formation of sulphenic acid intermediate (RSOH) which in­
creased to a maximum concentration and then decreased as it was converted
to the corresponding thiolsulphinate product (RSOSR) (Fig. 1). The same
workers [34] have calculated first order rate constants for the decomposition

Fig. 1. Concentrations by NMR vs. time of heatig 0.5 M di-tert-butyl sulphoxide in benzene
at 80°C. (a) Sulphoxide; (b) sulphenic acid; (c) thiosulphinate. (Reproduced with kind
permission from Int. J. Sulf. Chem., 8 (1973) 205.)
ANTIOXIDANTS— PREVENTIVE MECHANISMS 167

TABLE 2

First-order rate constants for the decomposition of sulphoxides at 100°Ca [34]

R1R2S -> 0
Ri R2 lO6*^1) Relative k

n-propyl n-propyl 0.06 0.02


isopropyl isopropyl 6.0 0.70
methyl tert-butyl 6.3 1.0
isopropyl tert-butyl 46 4.4
tert-butyl tert-butyl 1170 93
phenyl isopropyl 1.3 0.31
phenyl tert-butyl 205 33
n-heptyl CH2CH2C02CH2CH3C 390 280
CH2CH2CO2CH2CH3 CH2CH2C02CH2CH3C 850 300

a 0.05 M in toluene solvent; propionate systems 0.015 M in o-dichlorobenzene solvent.


b Corrected for the number of ß hydrogens.
c Extrapolated for data at lower temperatures assuming an activation enthalpy of 25
kcal mol" .

of several sulphoxides at 100°C (Table 2). The highest relative rate of


decomposition shown in Table 2 was for diethyl-ß,ß-sulphinyldipropionate
(VIII, R = C2H5) which was 300 times that of methyl-ter£-butyl sulphoxide.
The carbonyl of the ester group activates the hydrogen (which is ß to the
sulphinyl group) by delocalising the developing negative charge as the
hydrogen is transferred to the sulphinyl oxygen in the transition state.

0+
O O- --H- -O
I
R—O—C—CH2CH2—SNS- / C H — C—O—R (VIII)
N
CH 2
0+

Shelton and Davis have also shown that steric effects alone could not
explain the reason for the 100 times higher rate observed in Table 2 for
fert-butyl phenyl sulphoxide when compared to that of isopropyl phenyl
sulphoxide. The phenyl group stabilises the transition state by délocalisa­
tion of the charge on the developing sulphinyl anion as the carbon-sulphur
bond begins to break.
In a detailed study by Scott and co-workers [25-27] of the chemistry of
168 S. AL-MALAIKA

oxidation of thiodipropionate esters it was shown that in the absence of


hydroperoxides, sulphenic acid (IX) was formed from the sulphinyl di-
propionate ester (VIII) in a reversible process, Scheme la, which became
first order in the presence of an oxidising agent (e.g. CHP and the "stable"
radical galvinoxyl) which removed the sulphenic acid as it was formed [25].
In the presence of hydroperoxides the amount of sulphenic acid (IX) con­
sumed depends on the concentration and the ratio of hydroperoxide to the
acid. At low peroxide/sulphur compound ratios, sulphenic acid can promote
homolytic decomposition of hydroperoxide leading to alkoxyl radicals
(Scheme lb) which is the most probable cause for the observed pro-oxidant
effects at these low ratios. At high peroxide/sulphur compound ratio, how­
ever, sulphenic acid is readily oxidised to the corresponding sulphinic acid
(X), see Scheme lc. This acid was found [27] to be relatively stable at room
temperature but it readily disproportionated at 60°C to the corresponding
thiolsulphonate (XI) and sulphonic acid (XII), see Scheme Id. At high
temperatures, sulphinic acid decomposes to give sulphur dioxide and dialkyl
adipate (XIII) as major products, Scheme le. Homolysis was proposed to
account for these products and the minor by-products, alkylpropionate (XIV)
and alkyl acrylate (XV), the alkyl group used in this study was methyl. In
addition to the peroxidolytic action of the sulphur acids (sulphenic, sulphinic
and sulphonic acids) and sulphur dioxide they were also found [27,35] to
have chain-breaking activity (see Scheme If as an example). In all cases the
radical produced was incapable of continuing the chain reaction and a
cyclical regenerative mechanism has been proposed (involving both CB-A
and CB-D activity (see Chapter 4) for the antioxidant function of the
sulphoxide (see Scheme 2).
Comparison of the rates of decomposition of hydroperoxide (e.g. CHP) by
the three sulphur acids, the sulphinic, sulphonic acids (X and XII, respec­
tively, R = Me) and S0 2 , at approximately equivalent molar concentration
[27] showed that S 0 2 and the sulphinic acid, X, are considerably more active
than the sulphonic acid, XII. The first two tend to the same limiting rate
with time and satisfy the requirement that the necessary concentration of
the catalyst is produced during the induction period whilst the sulphonic
acid does not [27].
It was shown earlier that dialkyl sulphinyl dipropionate (VIII) shows an
induction period (75°C) before it becomes an effective catalyst for the ionic
decomposition of CHP. In the absence of hydroperoxides the products
formed from this sulphoxide were the corresponding thiolsulphinate, XVI (a
powerful peroxidolytic agent), and its stable disproportionation products
XVII and XVIII which are not peroxidolytic agents [25,37], Scheme 2a-d.
Decomposition of thiolsulphinates (e.g. XVI and di-feri-butyl thiolsulphi­
nate) was shown [27,24,36] to give thiolsulphoxylic acid (e.g. XIX, Scheme
2e) and at least part of the catalytic activity was attributed to this species
[24]. It was also suggested [37] that thiolsulphoxylic acid is readily oxidised
R02H
(ROCOCH2CH2)2S > (ROCOCH2CH2)2S = 0 i ROCOCH = CH 2 + ROCH 2 CH 2 SOH
O
hv VIII R'O X
IX
£
b(ROOHdefic)

ROCOCH 2 CH 2 - + ROCOCH 2 CH 2 SO
CB —D 3
ROCOCH = CH 2 + R'SOH ROCOCH 2 CH 2 SO + / R ' 0 2 H RO + H 2 0 + ROCOCH2CH3 2
Pro-oxidants <
w
H
O
ROCOCH2CH2SR
se
Regenerative mechanism RCOCH 2 CH 2 SSCH 2 CH 2 C0 2 R

/ O XI

Removal of antioxidant species

ROCOCH 2 CH 2 S^
Lo + R'0-
OH

ROCOCH 2 CH 2 S0 3 H
XII

ROCOCH = CH 2 + ROCOCH 2 CH3+(ROCOCH 2 CH2)2+S0 2 > S03(H2S04)


X* XIV XIII Antioxidant species

Scheme 1. Antioxidant mechanism of dialkyl thiodipropionates in the presence of a hydroperoxide.


170 S.AL-MALAIKA

(ROCOCH 2 CH 2 ) 2 S=0 > ROCOCH 2 CH 2 SOH + R O C O Π= C H 2


(VIII) 0 (IX) I
l
Il
ROCOCH 2 CH 2 —S—S—CH 2 CH 2 COOR + H 2 0
(XVI) \e

ROCOCH 2 CH 2 SSOH + ROCOCH=CH 2


b/
(XIX) If

ROCOCH 2 SOH + S 0 2

ROCOΠ2 CH 2 SO + ROCOCH 2 CH 2 S
X2 c
l
o
II
d
ROCOCH 2 CH 2 —S—S—CH 2 CH 2 COOR v?\

(XVII) I
ROCOCH 2 CH 2 —S—S—CH 2 CH 2 COOR
(XVIII)
Scheme 2. Mechanism of antioxidant action of dialkyl sulphinyl dipropionate in the initial
absence of hydroperoxides.

to further sulphur acids with the elimination of sulphur dioxide (Scheme 2f).
However, no evidence was found for the direct formation of S 0 2 from the
thiolsulphinate, XVI, at 75°C and the thiolsulphonate, XVII, was quite
stable under these conditions [27].

(c) The role of sulphur dioxide


The formation of sulphur oxides was postulated in the early 1960s inde­
pendently by Hawkins [19] and Scott [18] in many instances to account for
the peroxide decomposing activity of a variety of sulphur-containing antiox-
idants. Studies of the behaviour of S 0 2 in model systems [38] showed that
although sulphur dioxide is more effective than sulphuric acid as a catalyst
for peroxide decomposition, the former cannot survive in the strongly oxidis­
ing environment of a typical autoxidation and it was therefore suggested
that the effective catalystfe) are either sulphur trioxide or sulphuric acid.
The behaviour of S 0 2 was found [38] to be very similar to most of the sulphur
compounds discussed in the previous section. For example it causes a
ANTIOXIDANTS — PREVENTIVE MECHANISMS 171

pro-oxidant effect at low peroxide, e.g. CHP/S0 2 molar ratios associated


with predominantly homolytic products of the decomposition of CHP. The
pro-oxidant effect disappears at higher ratios giving rise to essentially
heterolytic products (phenol and acetone) of the decomposition of CHP. It
was further demonstrated [38] that an organic insoluble phase, attributed
to H 2 S0 4 , separates at low molar ratios of CHP/S0 2 while at high ratios a
soluble strong acid, suggested to be S0 3 , was formed by the alternative ionic
breakdown of the unstable peracid XX (Scheme 3c). Consistent with this
mechanism is the observation [39] that hydroxyl and sulphinyl radicals can
be identified by ESR in the reaction of S 0 2 with hydroperoxides (Scheme 3a
andb).
The conversion of S 0 2 to a catalyst for hydroperoxide decomposition was
demonstrated [38] by incremental addition of tert-butyl hydroperoxide,
TBH, to an excess of S 0 2 in organic solutions. It was shown that the reaction
which was initially stoichiometric became catalytic well before a stoichio-
metric equivalent of hydroperoxide had been added. This leads to the
important conclusion that Lewis acid catalysis can operate even in the
presence of an excess of a radical generating redox system.

OH
a I
ROOH + S 0 2 —^—► ROOS= O (XX)
b / \ c

OH \
RO + OS ROH + SOQ

i
Prooxidants Antioxidants

Scheme 3. Reactions of sulphur dioxide with hydroperoxide.

(d) Diaryl and diaralkyl sulphides


Diaryl and aralkyl sulphides, like the dialkyl sulphides, are not effective
peroxidolytic agents unless synergised with chain-breaking antioxidants
and in fact some aryldisulphides, e.g. XXI (see Scheme 4), were found [40,41]
to be effective pro-oxidants in rubbers and polyolefins. In common with the
dialkyl monosulphides, the diaryl disulphides rapidly autoretard to give a
long term antioxidant effect and the further oxidation products, rather than
the original disulphide, are the main catalysts for peroxide decomposition.
Hawkins and Sautter [42] studied the activity of the diaryl disulphide, XXI,
172 S. AL-MALAIKA

and its derived oxidation products in the autoxidation of cumene at 120°C,


Scheme 4. Whereas the disulphide XXI became effective as an antioxidant
only after an initial period of inactivity, the derived thiolsulphinate, XXII,
was effective immediately. The end disproportionation product of XXII, the
thiolsulphonate, XXIII, was less effective but became an antioxidant after a

<C^s-s^O>

S 0 2 / S 0 3 , etc.
Antioxidants

^-—®
XXI
XXXIII

'X2

^f--@
o

Scheme 4. Antioxidant activity of diaryl disulphide (XXI, R=H)


ANTIOXIDANTS— PREVENTIVE MECHANISMS 173

longer period of inactivity. It was shown [43,44] that thiolsulphonates


undergo thermolysis relatively slowly at temperatures below 130°C,
whereas thiolsulphinates are unstable at much lower temperatures and
readily undergo further reaction with hydroperoxides to give sulphur acids.
During the initial stages of autoxidation, Reaction b in Scheme 4 and
subsequent homolytic processes tend to predominate, but once hydroperox­
ides have built up in the system, Reaction c (Scheme 4) effectively inhibits
further oxidation.
Benzyl monosulphides were shown [41] to behave generally in a similar
manner to the diaryl disulphides. The derived benzylic sulphoxide, XXIV,
have been shown [45,46] to undergo facile reversible homolysis (Scheme 5a)
at high temperatures and this is responsible for the pro-oxidant reactions
observed during the initial stages of oxidation [41]. The effective ionic
catalystfs) for peroxide decomposition, most likely a low molecular weight
sulphur acid, was shown [47] to be produced by further oxidation of the
rearranged sulphenate ester, XXV, (Scheme 5e-g) irrespective of the nature
of the benzyl moiety. This will be explained in detail in Chapter 9 of Volume
II.

<0>-CH 2 - S—R

^y-CH 2 -S-R ; s ^ _)j-^n + • SK 5 ==± \\Jr-Utt2U5K


2
xxrv / XXV
X2 c
e ROOH

O o
II
<g)-CH2CH2^(g> RS—S R ^0/~CH2°SR
II
o f H2CXROOH)

(S0 3 ,etc) «-— RS02H /0V" CH 20H


Scheme 5. Antioxidant activity of benzyl monosuilphide.
174 S. AL-MALAIKA

1.2Mechanisms of Metal Complexes Containing Sulphur Ligands

Transition metal complexes containing at least one sulphur ligand have


been widely used as antioxidants (for thermal, UV stabilisation and control­
led photodegradation) in lubricating oils, rubbers and plastics and review of
the early literature is covered in the first edition of this book (Ref. [1] p 192
et seq.). Like other sulphur-containing compounds, thiolates function as
antioxidants by more than one mechanism. Although their peroxidolytic
activity was first discovered by Kennerley and Patterson [13] in the early
1950s their chain-breaking activity was not demonstrated until the mid
1960s [48-50]. Mechanistically, therefore, these metal complexes show a
generally similar behaviour to the sulphide antioxidants discussed in the
previous sections; they all oxidise during their antioxidant function to
sulphur acids. They differ from the latter, however, in that the nickel, copper
and cobalt complexes have the ability to function as UV stabilisers as well
as thermal antioxidants and this seems to be associated with the strong UV
absorbance of the metal ligand bond in the region 310-350 nm.

1.2.1 Metal mercaptothiazolates


The high thermal oxidative stability of rubbers cured by an "efficient
vulcanisation" system (i.e. high accelerator, low sulphur) was shown [51-53]
to be due to products (zinc mercaptobenzthiazolate, XXVII, and its hydrate)
formed from the accelerator (cyclohexylbenzthiazyl sulphenamide, XXVI)
during vulcanisation (Reaction 2). If these products, e.g. XXVII, ZnMBT, are
removed from the vulcanisate by extraction, the very high level of oxidative
stability disappears.

(gj^t-s-M^O ^ [(6rVs Zn (2)

(XXVI) (XXVII, ZnMBT)


The zinc complex, XXVII, ZnMBT, and the corresponding acid, MBT,
XXVII (Scheme 6), were both found [54] to be very powerful catalysts for the
decomposition of hydroperoxides and their behaviour as peroxidolytic anti­
oxidants was very similar to that of other sulphur compounds discussed
earlier. In the presence of molar excess of hydroperoxide (e.g. CHP) the
products formed, in both cases, were found to be predominantly those
expected on the basis of an ionic decomposition, but at lower ratios pro-oxi-
dant effects and autoretardation were observed, though to a lesser extent in
the case of MBT. Addition of a base (e.g. pyridine) inhibited the antioxidant
process and increased the pro-oxidant effect confirming that the effective
antioxidant is an acidic species.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 175

ROOH
©;>—
MBT, XXVIII
o
N N
H20 + RO+ [ 0 \ £—S' /
C—S +ROOH

PRO-OXIDANTS ROOH

m~ 2 MBTS, XXIX

ROOH

N
j O L P - H + S0 2 -
XXX, BT
i§0 £ — S — OH
II
o
SO,(H 7 S0 4 )

ANTIOXIDANTS

N
[ O J ^ )C-0H + S02 < OI > - | - Ö
b s
o o
ANTIOXIDANT
XXXI, BTSO +ROOH
Scheme 6. Transformations of mercaptobenzthiazole and its derivatives in the presence
of a hydroperoxide.
Figure 2 shows the products formed from MBT in the presence of different
concentrations of ter^-butylhydroperoxide, TBH [54]. At low molar ratios
([TBH1/[MBT] = 1) benzthiazolyl disulphide (MBTS, XXIX) was the primary
and major product formed (Scheme 6) but at a molar ratio [TBH]/[MBT] =
5, benzthiazole (BT, XXX) and benzthiazole sulphonic acid (BTSO, XXXI)
become major products. At [TBH]/[MBT] = 20 the latter was found to be the
only product. All the reaction products, except BT, were shown to have
peroxidolytic antioxidant activity. The sulphonic acid, BTSO, also exhibits
chain-breaking activity while the disulphide MBTS does not become a CB-D
antioxidant unless oxidised further. Furthermore, this acid (BTSO) does not
show the pro-oxidant effects associated with MBT and MBTS in the pre­
sence of hydroperoxide, and hence it is a much more powerful antioxidant.
176 S.AL-MALAIKA

The competing pro-oxidant and antioxidant processes for MBT are sum­
marised in Scheme 6. As in the case of thiodipropionate esters, the pro-oxi­
dant processes predominate during the early stages of the reaction or at low
[ROOH]/[S] ratios.
100»

10 20 30
[TBH]/[S] molar ratio

Fig. 2. Effect of tert-butyl hydroperoxide/MBT molar ratio, [TBHl/tS], on the yield of


products formed from MBT. (a) MBTS (MBT); (b) MBTS (MBTS); (c) BT (MBT); (d) BT
(MBTS); (e) BTSO (MBT and MBTS). Compound in parentheses is the starting material.
(Reproduced with kind permission from Eur. Polym. J., 15 (1979) 879).

The antioxidant performance of ZnMBT (XXVII) was shown [54] to be


somewhat different from that of MBT. Zinc sulphinate (XXXII, ZnBTS) is
formed in high yield instead of BTSO and benzthiazole (BT) is formed more
slowly (no zinc sulphonate was observed). Both ZnMBT and ZnBTS act as
reservoirs for the slow libration of S0 3 , see Scheme 7. Table 3 compares the
antioxidant effectiveness of ZnMBT with MBT and some of the derived
oxidation products, in paraffin oil at 140°C. Although BTSO is an effective
ANTIOXIDANTS — PREVENTIVE MECHANISMS 177

-N

®
> -s Zn

XXVII, ZnMBT
ROOH ROOH
(defic.) (excess)

H O - - Z n — S—C*

+
® (gnjt-i-o Zn

XXXIX, ZnBTS

RO + S —C
£@
N
+ S —C o ROOH
- so2/so3
XXIX, MBTS
Scheme 7. Transitions of ZnMBT in the presence of hydroperoxides.

antioxidant at temperatures below 100°C, it is much less effective t h a n the


zinc complex under these high temperature conditions.
The alicyclic analogues of MBT, the 4-alkyl-2-mercaptothiazolines
(XXXIII, RMT), differ considerably in technological performance; RMTs are
good UV stabilisers whilst MBT is not [55]. Moreover, the former stabilisers
were shown to have more effective UV stabilising role when the polymer was
oxidatively processed (in the presence of excess oxygen) than when it was
mildly processed (under limited amounts of oxygen). The primary transfor­
mation product formed during processing of RMT in PP was shown [56] to
have a UV absorption band around 312 nm (see Fig. 3c ); the corresponding

N
R
I *C—SH R-t; c-
RMT, XXXIII RTD, XXXW
178 S. AL-MALAIKA

TABLE 3

Induction periods (IP) for MBT, ZnMBT (VII and IV) and their derived oxidation products
in white paraffin oil, at 140°C [54]

Antioxidant Concentration * 104 (mol 1 *) IP (h)

MBT (VII) 13 1
ZnMBT (IV, M = Zn) 5 12
12 24
24 41
ZnMBT + TBH (1:5) 12 40
ZnMBT/Pyr (1:2 complex) 5 16
8 26
16 43
BTSO (X) 50 4
90 10
BTSO/Pyr(l:l) 90 1
ZnBTS (XII) 12 11

disulphide, RTD (R = Et), XXXIV, absorbs at 310 nm, while RMT do not
absorb in this region, see Fig. 3a. Examination of the kinetics of the reac­
tions of RMT (R = Ethyl, EMT) and RTD (R= Ethyl, ETD) with TBH in
solution revealed [56] that more than one compound absorbing at 310 nm
was formed during the reaction of the former with TBH (Fig. 4) and that the
first transformation product formed from EMT during its reaction with
hydroperoxides is the corresponding disulphide, ETD (cf. Fig. 3a and b).
Product analysis of reactions of EMT oxidised by TBH using GC-MS
technique showed that two pairs of products (XXXIVa,b and XXXV a,b,
Scheme 8, see also Fig. 4) are involved in the antioxidant action of RMT
compounds. The rapid formation of the new transformation products (the
isomers having m/e = 260, XXXIV a,b) is consistent with the cross termina­
tion of the intermediate radical species formed by loss of S 0 2 from the
oxidation products of the sulphides (see Scheme 8). A similar S 0 2 loss from
the analogous MBT was shown above (see Scheme 6) except that in this case
the parent benzthiazole (BT) is formed rather than the radical coupled
products. However, radical coupled products have been observed in the case
of alkyl sulphides (e.g. DLTDP) at higher temperatures [27]. The initial
photo pro-oxidant effect observed during the early stages of the photo-oxida­
tion of PP containing RMT [55] must therefore be due to the formation of
thiolsulphinates which are known to be sensitive to UV light and readily
ANTIOXIDANTS — PREVENTIVE MECHANISMS 179

3.0

<
CD

CO

<

250 300 350 250 300 350


WAVELENGTH (nm) WAVELENGTH (nm)

| 280 c

1
\ 246 / \

1.0
iW 312

200 275
WAVELENGTH (nm)

Fig. 3. Ultraviolet spectra of EMT, 2-10"4 M ( — ) and ETD, 3.5-10"6 M ( —) in hexane


(a). The UV spectrum of processed (OM, 5 min) polypropylene containing EMT (2-KT3
moVlOOg) is shown in (c). In (b) the UV spectra of EMT (reactant) and product (—) formed
during its reaction with TBH (2.5-at 10"3 M) at a molar ratio of 10 in dodecane at 50°C
are shown. (Reproduced with kind permission from Polym. Deg. Stab., 13 (1985) 261).

fragment to give radical species (see Scheme lg). This process must precede
the formation of sulphur acids which are the effective peroxide decomposers
formed at a later stage (see Scheme 8).

1.2.2 Metal dithiolates


Metal complexes of dithioic acids, e.g. dithiocarbamates, MDRC, (III),
dithiophosphates, MDRP (IV) and xanthates, MRX (XXXVI), are particu­
larly interesting examples of peroxidolytic antioxidants and were among the
earliest to be investigated from a mechanistic point of view becaxise of their
180 S. AL-MALAIKA

Et—CH—N Et—CH—N N- CH—Et


ROOH
>
I
CH C
I I
CH C—S —S — C CH2
V V
2 2

ROOH

Et — C H — N O N- CH—Et
I
CH2 C—S —S —C CH,
x
s7 Y
ROOH
(-so2)

Et — CH —N Et- CH — N Et — C H — N -
I II I II I I
CH2 C- +
\V CH2 C
x
CH2 C
^y . \y v
s s-
Et—CH—N N- CH — E t Et — C H — N

CH 9 C S C CH -> CH2 C—N CH — E t


^y \y I
** c CH2
s s y \./
Villa, m / e = 260 s s
ROOHC-SO,)
VlIIb,ROOH
m/e = 260
(-so 2 )

Et — C H — N N- CH—Et Et — C H — N

CH2 C—O —C CH2 CH, C—N- CH—Et


^y ^y Y C CH2
I
s s S \ /
O S
IXa, m/e = 244 IXb, m/e = 244
Scheme 8. Proposed mechanism for the oxidative transformation of 4-ethyl-2-mercapto
thiazoline.
ANTIOXIDANTS — PREVENTIVE MECHANISMS 181

R2N-C^ M

M = Zn

S S
• \ / ^ R2N—CC- + R2N—C( MOH + RO
R 2 N—Ct Zn ÎC—NR2 V \/
s s S S I
II R'H
o ROH + R'

ROOH

S S S
// // \
R ' + R 2 N —C R 2 N —C C—NR2
\ \ /
Sulphur acids SJJ § §
ROOH

RN = C = S + S 0 3 / H 2 S 0 4 Sulphur acids

Scheme 9. Proposed mechanisms of antioxidant action of MDRC.

importance in different technological media, e.g. in rubbers, lubricating oils


and plastics. Metal dithiolates have a broad spectrum of antioxidant activ­
ity: melt, thermal and UV stabilisation, and hydroperoxides are not nor­
mally detectable when they are used to stabilise, for example, polymers. The
metal ion plays a crucial part in their overall effectiveness. For example,
transition metal complexes containing nickel, cobalt and copper are more
stable towards UV light and are, therefore, better UV stabilisers than group
II metal complexes e.g. Zinc.
S S
/ x / \ / %
(RO)2C M C(OR)22 MRX, XXXVI
% / \ /
S S
182

50 100 150 200


TIME (min)

50 100 150 200


TIME (min)

Fig. 4. (a) Kinetics of changes (build up and decay) of species at 310 nm during the reaction
of TBH with EMT (2.5-10"4 M) at different molar ratios is dodecane at 25PC (numbers on
curves are molar ratios of PTBH]/[EMT]. Inset shows the time taken for the formation of
the second maxima at the different molar ratios tested, (b) Kinetics of changes (build up
and decay) of species formed at 310 nm during the reaction of TBH (1-10"2 M) with EMT
at a molar ratio (molar ratios [TBH]/[EMT]) = 40 in dodecane at different temperatures
(numbers on curves show temperatures in °C). Inset shows the time taken for the
formation of the first (O) and second (•) maxima at the different test temperatures [56].
(Reproduced with kind permission from Polym. Deg. Stab., 13 (1985) 261).

Extensive mechanistic studies over the past thirty years [57-64] have
confirmed early findings [13,14,18,48] that this class of compounds undergo
a complex series of oxidation reactions involving free radicals with sulphur
acids as the main catalysts for the final catalytic ionic decomposition of
hydroperoxides. The antioxidant stage is, therefore, preceded by a pro-
oxidant step which varies in intensity but when they are used in polymers
(usually at low concentrations), the pro-oxidant stage is not observable [57].
Overall, the mechanisms of action of metal dithiolates are similar. In
common with other sulphur-containing compounds, metal dithiolates func­
tion both as hydroperoxide decomposers and as radical scavengers [57-59,
ANTIOXIDANTS— PREVENTIVE MECHANISMS 183

65-67]. The relative contribution of these processes to the overall stabilising


effect, and the nature of the intermediates involved during their antioxidant
action, depend on: (i) the type of the metal complex, (ii) the ratio of the
peroxide to the complex, and (iii) the nature of the metal ion. Generally,
metal complexes of dithioic acids were shown [48,61,66] to exhibit multistep
hydroperoxide decomposition curves which represent both of the above two
processes; radical and ionic mechanisms. This is exemplified in the be­
haviour of nickel complexes of dithiocarbamic, dithiophosphoric, and xan-
thic acids, Fig. 5a [65]. In all cases, the rapid initial catalytic stage is
favoured by a low molar ratio of hydroperoxide to complex and gives rise to
homolytic products (Fig. 5b). The second catalytic stage which follows an
induction period, assumes greater significance at higher molar ratios of
hydroperoxide to complex, and gives rise to an ionic decomposition of hydro-
peroxides.
The fact that metal dithiolates are themselves not responsible for the
heterolytic decomposition of hydroperoxides but rather their transformation
products (formed by oxidation in the presence of hydroperoxides) is now well

a
» i.o
eu
X
u
r... ^
1 x
^_ ^^
^^^NiBX
o 0.5 \ ^ -
VNÎDBP
NiDBC^*-

1 K .i „1 _l ._!
20 60 100
Time (min.)

80
b 1
^ "Phenol 1
fr - 1
o

° 40

2
% f"^*«^. a-Methyl Styrene
>< i ^^^^ Acetophenone
\ oc- Cumyl alcohol 1
nn
20 60 100
[CHP] / [NiDBP]

Fig. 5. (a) Decomposition of CHP (MO -2 M) in chlorobenzene at 110°C in the presence of


nickel dithiolates (2-KT4 M). (b) Product yields after complete reaction of NiDBP with
CHP at 110°C in chlorobenzene at various molar ratios [CHP]/[DiDBP]. (Reproduced with
kind permission from Eur. Polym. J., 16 (1980) 503).
184 S.AL-MALAIKA

established [57,58,64,68,69]. They are, therefore, precursors of the effective


catalytic peroxide decomposers. Since the publication of the first edition of
this book, a thorough investigation in many laboratories into the nature of
the ionic catalyst(s) for peroxide decomposition formed from metal dithio-
lates has been pursued. The similarity in the overall mechanisms of anti-
oxidant action and the nature of transformation products formed from the
two more important dithiolate classes, namely, the dithiocarbamates and
dithiophosphates, will, therefore be discussed in some detail in the following
sections.

(a) Dithiocarbamates
The fact that the antioxidant activity of metal dithiocarbamate (MDRC)
is due to their oxidation by hydroperoxides to give sulphur acids which are
ionic catalysts for the decomposition of hydroperoxides has been exploited
in the development of highly effective thermal and UV stabilisers for poly-
olefins [57], (Their role as stabilisers in polymers will be discussed in
Volume II of this series). Different metal dialkyl dithiocarbamates (e.g. Ni,
Fe and Zn), however, behave differently in the polymer, which is in turn
different from that of the corresponding metal-free derivatives, the thiuram
disulphides. While nickel and zinc dialkyl dithiocarbamates are used as
thermal and photo antioxidants [69], the iron complex is used as pro-oxidant
under photooxidative conditions in spite of the fact that it is a very effective
melt stabiliser for polyolefines [109]. This difference in behaviour of the
different metal complexes was exploited in developing a very effective
time-controlled stabiliser systems for polyolefins based on combinations of
these different compounds [69-72]. Their mechanisms are discussed below.
Although the overall mechanisms of antioxidant action of the three metal
(M = Ni, Zn, Fe) dithiocarbamates, MDRC, are similar (involving both
homolytic and heterolytic processes), the nature of the initial transforma­
tion product(s) formed from these complexes is quite different, and this in
turn leads to different decomposition kinetics at different molar ratios of
hydroperoxide to metal complex. Comparative studies of the kinetics of
decomposition of cumene hydroperoxide (CHP) at high temperatures (e.g.
110°C) in the presence of the three metal complexes [61-63] revealed that
the both NiDRC and FeDRC behave similarly but quite differently from the
corresponding zinc complex. For example, while the décomposition curves of
NiDRC and FeDRC show the typical three-step behaviour, similar to that
exhibited by other dithiolate metal complexes [64], the ZnDRC exhibits a
two-stage behaviour (the initial rapid catalytic stage is completely missing)
under all molar ratios examined (see Fig. 6a). Furthermore, in the case of
NiDRC and FeDRC the length of the second stage induction period increases
with increasing [CHP]/MDRC] molar ratio (i.e., with decreasing metal com­
plex concentration) and the second catalytic stage becomes slower (Fig. 6b).
This behaviour is the exact opposite of what happens in the case of the zinc
ANTIOXIDANTS— PREVENTIVE MECHANISMS 186

complex (see Fig. 6b). These results suggested that the nature of the inter­
mediate products must be different in these cases.

Control

40 80 120 160
Time (min.)

(b) •o
NiDE(X ZnDEcJ
* 0.8
•c
1
j*r
0.4
-^F rO*\
^ ^ d
20 40 1
40 _ [CHP]/[Antioxidant] 1

r^ ^ • ^ ^ Z n D E C ^ ^ ^ *
20
F e D M C ^ ^ < ^ _ g
^-"""^l—•*-~*' ^^T NiDEC

10 30 50
Molar Ratio of CHP / Antioxidant

Fig. 6. (a) Comparison of CHP decomposition curves for reactions of MDRC, (M = Fe, Ni,
Zn) with CHP in chlorobenzene at 110°C and a molar ratio of [CHP]/[MDRC] = 30. (b)
Changes in induction period of CHP decomposition curves of MDRC, (M = Fe, Ni, Zn) and
the rate constant of the second catalytic stage in these curves (inset) as a function of
[CHP]/[MDRC] molar ratios. The concentration of CHP in all curves was MO"2 M [63],
(Reproduced with kind permission from J. Appl. Polym. Sei., 33 (1987) 1455).

It was demonstrated [63] that the nickel and iron dithiocarbamates form
the corresponding disulphide (XXXVII) as their initial oxidation product.
This is analogous to the formation of the corresponding disulphides from
many other related sulphur-containing compounds such as mercaptobenz-
thiazoles (XXVIII) and its metal complexes (XXIX), metal complexes of other
dithioic acids such as dithiophosphoric and xanthic acids and mercapto-
thiazolines (XXXIII) during similar high temperature reactions with hydro-
peroxides (see Sections 1.2.1 and 1.2.1(b)). The zinc dithiocarbamate, how-
186 S. ALrMALAIKA

ever, does not form the disulphide at any stage of its reaction with hy­
droperoxides. The main initial transformation product in this case was
found [62] to be the zinc thiopercarbamate, ZnDRSO (XXXVIII), see Scheme
9. The ease of formation of the initial transformation product, whether it is
the disulphide or the dithiopercarbamate, may depend on the oxidisability
of the central metal ion (NiDRC undergoes oxidation much more readily
than ZnDRC [74]).

S S
R 2 N — C/ VC—NR, R2N—C Zn
\
s—s
/ V
II
O.
XXXVII XXXVIII

It has been shown [57] that the main catalysts produced at the later
stages of reactions of metal dithiolates and other related sulphur-containing
compounds with hydroperoxides are sulphur acids, irrespective of the na­
ture of the initial transformation products. In the case of dithiocarbamates,
this was demonstrated [61,63] by examining the effect of adding a strong
base on the auto-oxidation of cumene initiated by CHP in the presence of the
three metal complexes (a technique developed earlier for other metal com­
plexes [54]. Figure 7 inset c, shows the effect of adding pyridine to the above
reaction in the presence of MDRC (M=Ni, Zn, Fe), Fig. 7a, and demonstrates
the complete removal of the antioxidant function of these complexes by
pyridine. Addition of a base was used further to demonstrate the oxidation
of ZnDRC to form the acidic species, responsible for the ionic decomposition
of hydroperoxides [54]. This was shown in the case of ZnDRC to occur during
the "induction period". Addition of excess CaC0 3 led to the neutralization of
all the acids, and hence no rapid second stage reaction occurred, while
addition of a less than stoichiometric amount of this base cannot effect
complete inhibition of the peroxide decomposition since it neutralizes only
part of the acids formed (see Fig. 7, inset d). It was indicated earlier (Section
1.1c) that the main acidic species formed in reactions of sulphur-containing
compounds with hydroperoxides is almost certainly S0 3 . Addition of a
molecular sieve to the above reactions (cumene/CHP/MDRC) to trap any
trace of the gas, if formed, showed [63] that the antioxidant activity is only
partially removed (Fig. 7b) suggesting that other acid catalysts beside S 0 3
must be involved in the antioxidant action of MDRC. H2SO4, formed by
reaction of SO3 with water from the dehydration of a-cumyl alcohol, was
suggested [54]. Examination of CHP reaction products formed during high
temperature reactions of CHP with metal dithiocarbamates (Ni, Zn, Fe) at
different molar ratios (Figs. 8 and 9) shows that at [CHP]/MDRC] molar
ANTIOXIDANTS — PREVENTIVE MECHANISMS 187

ratios below 20, the homolytic free radical process predominates as reflected
by the high concentration of acetophenone and a-cumyl alcohol, whereas at
molar ratios greater than 30, the products formed are mainly those expected
by a heterolytic process. The high concentration of a-methyl styrene formed
at all ratios, suggests that the acidic species responsible for the formation of
phenol also cause dehydration of a-cumyl alcohol to the corresponding olefin
(Reaction (3)). Furthermore, the presence of a constant but low proportion
of the homolytic decomposition products at higher ratios indicates that the
radical process makes a contribution in the overall mechanism at all ratios.

20
Time, h

, ^ ^ CUMENE c

Î1
d È ZnDEC /
i/FeDEC
~40
c
.0

\ \ o 0 1 2 m in"1 fro My Ni DEC


1 CO
0 0 2 7 Vnin" \
? 1.0 "?
«4
60 180 ^r 10
i 20
i 1II
Time, h
Time, min.
I
/
1
20 y'Contr ol b
c / ja FeDMC
/
** ?/l ZnDEC
o ^ ^ NiDEEC
(0 1 _SrJA
< 10

CM
t/yr
1 to -SM
O

10 20 30
Time, h

Fig. 7. (a) Effect of metal dithiocarbamates (510 M) on oxidation of cumene in the


presence of CHP (1-KT2 M) at 110°C. (b) Effect of molecular sieve (5A). (c) Effect of
pyridine, 5-10 M on the oxidation of cumene in the presence of dithiocarbamates (5-10-5
M) and CHP (0.1 M). (d) Effect of a base (CaC03> on thermal decomposition of CHP (M0~ 2
mol) at 110°C in chlorobenzene in the presence and absence of ZnDEC. Molar ratios of
CHP:ZnDEC:CaCO: (A) 30:0:0.25; (0) 30:1:2; (O) 30:1:0.25; (■) 30:1:0. (Reproduced with
kind permission from J. Appl. Polym. Sei., 33 (1987) 1485).
188 S.AL-MALAIKA

CH CH, CH7
(^yC-OOU -c—o
> (Q^C-O' > < g ^ ç _ o H — <Q>-Ç
C + H20
I
CH -i CH 3o CH -l CH
C H 3-i

+ OH

(3)

In contrast to the metal complexes, the metal-free derivatives, the thi-


uram disulphide (XXXVII) behaved quite differently. The main difference is
that while the metal complexes act mainly by a free radical scavenging
process at low hydroperoxide to sulphur compound molar ratios, in the case
of the disulphides the ionic reaction dominates the free radical process at all
molar ratios including stoichiometric. Figures 8 and 9 show that at stoi-
chiometric ratio, the phenol yield is 30% in the case of TETD, but less than
60 -a
^0I00I* • ■

? 40 A
>*
20
* ——» m 1
, T — ¥
*» . ■ -T 1 ■■
20 60 100
[CHP]/[NiDEC]

N-
60«
i
% 40

«wS-,
JH }f ^ ^ __

tf /
20
s*^/V
T • 1
v^
■ mS >-*.


20 60 100
[CHP]/FeDMC]

Fig. 8. Product yield after complete reaction of FeDMC with CHP at 110°C in chloroben-
zene at various molar ratios of [CHP]/[DMC]; M = Ni(a) and Fe(b). (Reproduced with kind
permission from J. Appl. Polym. Sei., 33 (1987) 1455).
ANTIOXIDANTS — PREVENTIVE MECHANISMS 189

0 10 20 30 40 50 60 70 80 90 100
(CHP)/(ZDEC)

20 40 60 80 100
[CHP] / [TETD]

Fig. 9. Product yield after complete reaction of ZnDEC (a) and TETD (b) with CHP (110~ 2
M) at 110°C in chlorobenzene at various molar ratios of [CHP1/ITETD or ZnDEC].
(Reproduced with kind permission from J. Appl. Polym. Sei., 30 (1985) 237; 33 (1987)
1455).

3% for the metal complexes). This behaviour is similar to that of disulphides


of other dithioic acids, e.g. thiophosphoryl disulphide (see next section).

(b) Dithiophosphates
Like the dithiocarbamates, the metal dithiophosphates act as antiox-
idants by both chain-breaking and peroxide-decomposing mechanisms. In
the former case, the metal complexes themselves play the major role while
in the latter case the transformation products are the more important
190 S. ALrMÂLAIKA

contributors to the mechanism [64]. However, in both cases the nature of the
oxidation products formed from reactions of the metal complexes with
hydroperoxides is very important to the overall mechanism. The antioxidant
role of metal-free derivatives of dithiophosphoric acid will be discussed first.

Metal-free derivatives of dithiophosphoric acid

Thiophosphoryl disulphides, DRDS (XXXIX) and some sulphur acids, e.g.


dithio- (DRDPA, XL), thiono- (DRTnPA, XLI) and thio- (DRTPA, XLII)
phosphoric acids have been reported [49,64,65,68,74-77] to be transform­
ation products from reactions of many metal dithiophosphate complexes
with hydroperoxides. In particular the disulphide has been identified by a
large number of workers as the initial transformation product formed from
the nickel, zinc and iron complexes during their reactions with hydro­
peroxides. There is, however, much less work on the positive identification
of the acids from these reactions although the intermediacy of sulphur acids
has often been suggested in discussions of mechanisms of action of dithio-
phosphates. Recent work on the mechanisms of action of dithiophosphates
[64,78-80] has thrown light on the involvement of these acids in the anti­
oxidant mechanisms of metal complexes of dithiophosphoric acid.

S S S S O
// \\ // // //
(RO)2P P(OR)2 (RO)2P (RO)2P (RO)2P
\ / \ \ \
S-S S-H O-H S-H

DRDS (XXXIX) DRDPA (XL) DRTnPA (XLI) DRTA (XLII)

Both thiophosphoryl disulphides and dithiophosphoric acids were shown


to act primarily as effective peroxide decomposers (through their oxidation
products) at all molar ratios of peroxide to dithiophosphate. There is also a
small contribution from the homolytic process at both stoichiometric and
catalytic ratios [65]. Figure 10 compares the inhibition effect of DRDS and
DRDPA on the CHP-initiated oxidation of decalin (at 130°C). In the case of
the disulphide, oxidation of decalin occurs immediately followed by a second
much slower oxidation stage, while in the case of the acid, the initial rapid
oxidation stage is completely absent even at low concentrations (for ex­
ample, cf. curves 5-10 in Fig. 10 a and b) [78, 64]. The initial pro-oxidant
stage in the case of the disulphide indicates its inability to inhibit hydrocar­
bon oxidation in the presence of excess hydroperoxides. However, the pre­
sence of the second autoretarding oxidation stage suggests clearly that,
under these conditions, the disulphide must be oxidised to more powerful
catalysts during the first step which are responsible for the autoretarded
ANTIOXIDANTS— PREVENTIVE MECHANISMS 191

60 Time, h

Q. 5x10~4

.Q

< 2x10"3

1 mr^m^^t^^^^^^ |

20 40 60 Time, h

Fig. 10. Effect of (a) thiophosphoryl disulphide DRDs, R = i-Bu; (b) dithiophosphoric acid,
DRDPA, R = hexyl; and (c) thionophosphoric acid, DRTnPA, R = i-Bu: on the oxidation of
decalin at 130°C in the presence of CHP (110 m). The molar concentrations of the
phosphorus compounds are given on the curves.

oxidation in the second stage. This is supported by the two stage decomposi­
tion of hydroperoxide (shown in the inset to Fig. 10a): an initial induction
period involving no (or very little) peroxide decomposition during which the
disulphide is oxidised to the more powerful products responsible for the
second rapid catalytic stage. Although dithiophosphoric acid (DRDPA) itself
is more effective peroxide decomposer, even at low concentrations, when
compared to the disulphides, the overall retarded inhibition of the acid
(shown in Fig. 10b) does indicate that the acid is also oxidised to more
effective catalysts.
Examination of kinetics and nature of transformation products of reac­
tions of DRDS and DRDPA with CHP at high temperatures (e.g. 110°C)
[78,79,64] revealed some important differences:
192 S.AL-MALAIKA

1. While the acid showed a rapid one stage CHP decomposition at a wide
range of acid concentrations (covering molar ratios of [CHP]/[DRDPA] of 2
to 100), the disulphide exhibits a two stage behaviour when compared under
the same experimental conditions (c.f. insets of Fig. 10 a and b).
2. P NMR analysis of products formed from the above reactions showed
[78,79] that the disulphide DRDS was the major transformation product
(90% yield) of the acid, together with some tetrasulphide (DRTetS) and
thiophosphoric acid (DRTPA), Table 4. In the case of the disulphide the
oxidation products are mainly the tetra- (DRTeS) and tri- (DRTS) sulphides
together with thio- (DRTPA) and thiono (DRTnPA) phosphoric acids.
Thiophosphoryl disulphides show an interesting behaviour when used at
very high concentrations (in the region of 10"1 M) with hydroperoxides at
near stoichiometric ratios. The initial slow peroxide decomposition step
observed at low concentrations and in the presence of excess peroxide (Fig.
10a inset) is replaced by a rapid decomposition stage, during which the
initial disulphide concentration stays almost unchanged, see Table 5. This
suggests that at these high concentrations, the disulphide itself is re­
sponsible for the initial (80-90%) peroxide decomposition since almost no
transformation products were obtained [78,79] during this period. The
oxidation products contribute mainly to the final decomposition of the
peroxide. Scheme 10 shows the probable reactions involved in the antiox-
idant action of thiophosphoryl disulphide. It is important to mention here
that, in case of polymer stabilisation, the antioxidants are normally used at
low concentrations (in the order of (10 -3 to 10 M) and under such condi­
tions oxidation products of thiophosphoryl disulphide, rather than the di­
sulphide itself, play the major role in the stabilisation mechanism.

TABLE 4

Products of oxidation of DRDPA (2-10""1 M) during its reaction with CHP in chlorobenzene
at 110°C. The [DRDPA]/[CHP] molar ratio was 1:2.5

Reaction Phosphorus yield (%) CHP


time (min) decomposed
DRDPA DRDS DRTetS (RO)2P(0)SR DRTPA CHP
(85.6)a (85.2)a (84.2)a (24.8)a (21)a (%)

0 100 0 0 0 0 10
2 0 90 5 0 5 90
12 0 80 15 0 5 95
30 0 73 18 4 5 98

P shift, ôppm.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 193

TABLE 5

Products of oxidation of DRDS (1-10"1 M) during its reaction with CHP in chlorobenzene
at 110°C. The [DRDPA]/[CHP] molar ratio was 1:5

Reaction Phosphorus yield (%) CHP


time decomposed
(min) DRDS DRTetS DRTPA (%)
(85.2)a (84.2)a (21)a

0 100 0 0 0
2 99 1 0 75
10 92 6 3 90
30 91 6 2 95

131
P shift, ô ppm.

TABLE 6

Products of oxidation of DRTnPA (2-10"1 M) during its reaction with CHP in chloroben­
zene at 110°C. The [DRTnPA]/[CHP] molar ratio was 1:2.5

Reaction Phosphorus; yield (%)


time
(min) DRTnPA DRTPA (RO)3P=0 (RO)3P=S
(63)a (21)a (-0.1)a (64.8)a

0 100 0 0 0
3 0 99 1 0
20 0 91 9 0
45 0 88 9 3

131
P shift, ô ppm.

Much less is known about the antioxidant activity (e.g. peroxide decompo­
sition and radical trapping) of thiophosphoric (DRTPA) and thionophosph-
oric (DRTnPA) acids. Both were shown to be produced from nickel and zinc
dithiophosphate during their reaction with hydroperoxides at different temp­
eratures [78-81]. Al-Malaika et al. have recently studied the effectiveness of
thionophosphoric acid as a peroxide decomposer and the nature of its transfor­
mation products under similar conditions to those used for other derivatives
of dithiophosphoric acid [64,78,79]. Figure 10c [78a] shows that the oxida-
194 S. AL-MALAIKA

S S
// //
(RO) 2 P+(RO) 2 P
\
S—S*
s sv
v " ^ •
(RO) 2 P// \P(OR)2 y2 R O O H (RO)
, x /'
L2 Pt •
\\ // (a) V // \
P P(OR)2
S—S \ /
S—S—S—S
(d) DRTetS(84-2)

// \
S S (RO) 2 P P(OR)2
# Ce) 0 \ /
(RO) 2 P o <—— (RO) 2 P S—S—S
\n \ DRTS (83 • 8)
S—OH S—OH
O O
DRDSO

(0

P °v
// [RCX)H] _ ^ x Jf (h) // \
S02 + (RO) 2 P ^ (RO)22 P
\ (K) \ — (RO)2P P(OR)2
OH \ /
SH S—S
DRTnPA (63) DRDOS (21)
DRTPA (21)
XLIV
S0 3 (H 2 S0 4 ) (0

S0 2 + (RO)2P
// (j) F
(RO)2P0
\ \n
OH S—OH
DRPA(-0-6), XUII

Scheme 10. Oxidation of thiophosphoryl disulphide in the presence of hydroperoxide.


Numbers in the scheme are 1 P NMR chemical shifts.

tion of decalin (at 130°C) in the presence of DRTnPA is autoretarded from


the beginning of the reaction (no induction period) when the concentration
of the acid is high, but shows a rapid oxidation at low concentrations (e.g.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 195

5-10"^ M). This suggests that thionophosphoric acid itself is not a very
effective antioxidant but when present at high enough concentration it
oxidises to a very powerful antioxidant which is responsible for the effective
inhibition. This accounts for the observation that the same small concentra­
tion of the acid (540" 4 M) autoretards the CHP-initiated oxidation of decalin
in the same way as when the concentration of the acid is high (Fig. 10c,
inset). Studies of 31 P NMR [78,79] have shown that thionophosphoric acid
(DRTnPA) is quantitatively transformed by hydroperoxides at high tempera­
tures to thiophosphoric acid (DRTPA) during the early stages of the reaction
(after 3 minutes, see Table 6); DRTPA is the real catalyst for the ionic
decomposition of hydroperoxide. Furthermore, thionophosphoric acid is so
readily oxidised in the presence of hydroperoxides that even at room temp­
erature it is quantitatively isomerised to thiophosphoric acid [79], see Scheme
10g. Table 6 suggests that thiophosphoric acid is a stable end product though
a small amount (<10%) is oxidised further to phosphoric acid (DRPA, XLIII),
see Scheme 10 i and j , which has been shown [82] to be an ineffective peroxide
decomposer. It is also possible that thiophosphoric acid may oxidise further
to the corresponding unstable disulphide, DRDOS, XLIV, which is expected
to have the same P chemical shift as the acid. The intermediate formation
of DRDOS during the oxidation of ZnDRP by hydroperoxides has been
suggested by some workers [74,83] to occur by a different route (see below).

Metal complexes ofdithiophosphoric acid

(i) Zinc dithiophosphates (ZnDRP)


The antioxidant mechanism of the zinc dithiophosphates has been
studied extensively in the last two decades mainly due to its importance as
an antioxidant, detergent and antiwear agent in lubricating oil. Studies of
the reactions of ZnDRP with hydroperoxides in both oxidisable and non-ox-
idisable model substrates have given an insight to the detailed mechanism
of its antioxidant action [13,58,64,67,78,83,84]. It was shown [67,83,84] that
at low temperatures (typically 70°C) the reaction of ZnDRP with CHP in
chlorobenzene (non-oxidisable substrate) exhibited three stages. A rapid
initial stage favoured by a low molar ratio of the peroxide to the complex
which gave rise to homolytic products (e.g. a-cumyl alcohol) was followed by
a period of slow reaction (induction period) leading to a third fast stage
during which all the remaining peroxides were destroyed. This final stage
was found to assume a greater significance at higher molar ratios of per­
oxide to complex and gives rise to an ionic decomposition of hydroperoxide.
The pro-oxidant effect observed during the oxidation of hydrocarbons in the
presence of ZnDRP and CHP [13,58,83] corresponds to the initial stage of
decomposition of hydroperoxide [67]. The consumption of the zinc complex
[67] during the first stage suggests that the catalyst for the third ionic stage
is directly formed by reaction of the zinc complex with the peroxide.
196 S.AL-MALAIKA

There has been some dispute about the identity of the ionic catalysts
produced from reactions of ZnDRP with hydroperoxides. Dialkyl thio-
phosphoryl disulphide, DRDS, for example, was isolated [67] as a major
product of the first stage reaction but its role as a catalyst has been
questioned [67,58,78,84,85]. A number of other sulphur-containing com­
pounds were identified in different laboratories as products of the oxidation
of ZnDRP by CHP. At low molar ratios of peroxide to complex, basic zinc
dithiophosphate (b-ZnDRP, XLIV) and the disulphide, DRDS were isolated
[74] (see Scheme 11).

{(RO)2PSS}2Zn + CHP -> {(RO) 2 PSS} 6 Zn 4 0 + {(RO)2PSS}2


ZnDRP b-ZnDRP DRDS

{(RO)2PS}2 + S
DRODS
{(RO)2PSS}2S
DRTS
H?0
{(RO)2PSS}2 + {(RO)2PS}2
DRDS DRDOS
O O
II
(RO) 2 P —OH (RO)2P — SH
DRPA DRTPA
Scheme 11. Oxidation of ZnDRP by CHP as suggested by Sanin et al. [74] path b-d, and
by Rossi et al. [83] path e.

Further oxidation of the disulphide was found [74] to produce dialkyl-


phosphoryl disulphide (DRODS) which on hydrolysis gave dialkylthiophos-
phoric acid (DRTPA) and dialkyl phosphoric acid (DRPA), see Scheme 12b
and c. Dialkylthiophosphoryl trisulphide (DRTS) was also identified [74] as
a product of the prolonged oxidation of ZnDRP by CHP (Scheme lid). The
formation of the phosphoryl disulphide (DRODS) was proposed to be formed
[83] via oxidation of the basic zinc complex which is formed in the initial
reaction of zinc dithiophosphate with hydroperoxide (Scheme l i e ) . Several
other phosphorous containing species have been claimed to be formed from
reactions of ZnDRP with hydroperoxides (at low temperatures) such as
esters of dithiophosphoric (XLVI) and phosphoric (XLVII) acids [86] as well
as zinc dialkylthiophosphate (ZnDRT, XLVIII) [87b].
ANTIOXIDANTS — PREVENTIVE MECHANISMS 197

(RO)2P-SR (RO)3 P - O {(RO)2PSO}2 Zn


II
S

CTRDTP, XLVI) (TRPP, XLVII) (ZnDRT, XLVIII)

Al-Malaika et al [64,78,79] have conducted extensive studies at high


temperatures (110-130°C) on the behaviour of ZnDRP, and some of its
transformation products, as antioxidants in the presence and absence of

{(RO)2PSS}2Zn
ZnDRP (98 -2)

{(RO)2PSS}6Zn40 {(RO)2PSS}2
b-ZNDRP (103) DRDS (85 • 2)

{(RO)2PS}2S
DRMS (79)
(RO)2POH
DRPA(-0-6)

{(RO)2PSS}2S
DRTS (83 • 8)

(RO)2PSR {(RO)2PSO}2
{(RO)2PSS}2S2 DRDOS
TIP (24 • 5)
DRTetS (84 • 2)
Scheme 12. Transformation products of ZnDRP during its reaction with CHP at high
temperature (110°C). Numbers in parentheses are 31 P chemical shifts.
198 S.AL-MALAIKA

CHP. At 110°C, the reaction of ZnDRP with CHP in a non-oxidisable


substrate (chlorobenzene) was found [78b] to proceed in two stages, see Fig.
11a (compared with three stages at lower temperatures, 70°C [67,83,84]).
The initial fast decomposition stage which corresponds to the first stage
reported at lower temperatures gives rise to homolytic products (e.g. a-
cumyl alcohol and acetophenone), Fig. 11a, inset. The disappearance of
a-cumyl alcohol and the build up of a-methylstyrene and phenol during the
slower second stage suggests an acid catalysed reaction, hence the ionic
mechanism (see Scheme 11a). The initial stage of the homolytic decomposi­
tion of hydroperoxide leading to the generation of free radicals was found
[78] to be responsible for the initial pro-oxidant stage observed during the
inhibited oxidation of decalin by ZnDRP in the presence of CHP at high
^ioo\

i I
a
l"S^i& .^»-""ï^fr- ' ■ ■-JE-
CM 100 200 I
' O Time , mln.

S
CL
1.0
L—- • -
__
-4
1x10
I

X -4
Ü
° 0.6
c - ^ ^ _ -3
o * - — - ^ ^ ^ 2 x 10

8 0.2 h
c
o
Ü

100 200
Time , min.

40 60
Time , h

Fig. 11. (a) Decomposition of CHP (110~ 2 M) by zinc dithiophosphate, ZnDRP, R = i-Bu,
in chlorobenzene at 110°C. Numbers on curves are concentrations of ZnDIBP in M. Inset
shows products formed. • , CHP; ■, a-methystyrene; ▲, a-cumyl alcohol; T, aceto­
phenone; X, phenol, (b) Effect of ZnDRP, R = i-Bu, on the oxidation of decalin at 130°C in
i - 2z ,
the presence of CHP (l-10~ M).
ANTIOXIDANTS— PREVENTIVE MECHANISMS 199

temperatures, 130°C (see Fig. lib). The subsequent effective autoretarding


inhibition of decalin is a consequence of the second stage ionic decomposition
of hydroperoxide (see Fig. 11a). The basic zinc complex (b-ZnDRP) shows
very similar behaviour when examined under the same experimental condi­
tions [78b] except that a much greater amount of CHP is needed to break
down the basic zinc complex (b-ZnDRP) than is required to destroy the same
molar concentration of the ZnDRP. This may be due to the larger number of
thiophosphoryl ligands present in the former case.
Examination of the nature of the transformation products formed from
the zinc dithiophosphates during their reaction with CHP at 110°C (using
31
P NMR), Table 7 [64] shows that the nature and concentration of the
products formed depend on the initial ratio of metal complex to peroxide.
Table 7 shows clearly that the major products from a stoichiometric reaction
of ZnDRP with CHP are the basic-zinc complex and the disulphide (b-
ZnDRP, 0=103, & DRDS, 0=85.2). At higher molar ratios, e.g. [CHP]/
[ZnDRP] > 5, the basic zinc complex was not detected. The original zinc
dithiophosphate is completely consumed during the early stages of the
reaction (less than 4 minutes); the small amount found at the end of a
stoichiometric reaction must arise from subsequent decomposition of the
basic zinc complex itself (cf. Tables 7 and 8). The major transformation
products observed at these high molar ratios are polysulphides, including
di- (DRDS), tri- (DRTS), tetra- (DRTeS) sulphides in addition to monosul-
phide (DRMS) and thiophosphoric acid (DRTPA). The unstable nature of
zinc dialkylthiophosphate (ZnDRT) is indicated from its initial build-up and
subsequent consumption when the molar ratio of peroxide to complex is > 5;
a similar observation was made by Sher [87].
It was shown [79] that oxidation of thiophosphoryl disulphide (DRDS)
leads mainly to tetrasulphides, thio- and thiono-phosphoric acids, in addi­
tion to trisulphide. Subsequent oxidation of thiophosphoric acid (DRTPA)
may lead to the corresponding phosphoryl disulphide (DRDOS), see Scheme
11, h. However, although in the NMR studies [78,79], no further attempt
was made to separate these two products (which are expected to give a
similar NMR chemical shift), it is highly unlikely, that DRDOS is formed
directly from DRDS, under these conditions, and that it is subsequently
hydrolysed to the acid, DRTPA, as has been suggested previously (see
Scheme l i b and c) [74,83].
Reactions of the basic-zinc complex with CHP at 110°C were shown [78b]
to give predominantly disulphides, and the zinc thiophosphate complex
(ZnDRT), Table 8. The formation of the latter from the basic-zinc complex
(Table 8) suggests that this species, which was found to build up during the
early stages of oxidation of ZnDRP by CHP (see Table 7), may be derived
from the basic-zinc complex. Although, at higher molar ratios of peroxide to
complex, the basic-zinc complex was not detected, because of its fast oxida­
tion reaction with the hydroperoxide (see Table 8), its intermediacy and
TABLE 7

Products of oxidation of ZnDRP by CHP in chlorobenzene at 110°C where TP is (RO)3P=S

[ZnDRP]/ Reacticm Phosphorus yield (%)


[CHP] time
molar ratio (min) b-ZnDRP ZnDRP DRDS DRTetS DRMS ZnDRT DRTS DRTPA DRPA TP TIP
(103)a (98.2)a (85.2) (84.2)a (79)2 (48.3)a (46.3)a (83.8)a (21)a (-0.6)a (64.2)a (24.5)a

1:1 0 <1 99 0 0 <1 0 <1 _ __ _


1:1 30 33 9 39 6 4 7 2 - - - - -

5:1 0 <1 99 0 0 <1 0 <1 - 0 - - -


4:1 4 0 0 48 16 8 18 3 - 7 - - -
4:1 10 0 0 45 17 9 16 0 - 12 - - -
4:1 60 0 0 44 27 12 0 0 - 16 - - -

10:1 0 <1 99 0 0 <1 - - 0 0 0 - —


10:1 5 0 0 36 15 13 - - 11 17 0 - -
10:1 60 0 0 34 18 10 - - 6 9 13 - -

50:1 0 <1 99 0 0 <1 - - 0 0 0 - 0


50:1 10 0 0 23 11 12 - - 14 31 0 0 0
50:1 180 0 0 14 15 12 - - 0 0 8 16 12

P shift, ô ppm.
ANTIOXIDANTS — PREVENTIVE MECHANISMS 201

TABLE 8

Products of oxidation of b-ZnDRP during its reaction with CHP in chlorobenzene at 110°C

[b-ZnDRP]/ Reaction Phosphorus yield (%)


[CHP] molar time
ratio (min) b-ZnDRPZnDRP DRS DRTetS DRMS ZnDRT
(104)a (98.2)a (85.2)a (84.2)a (79)a (48.3)a (46.3)a

1:1 0 90 6 0 0 0 3 0
1:1 30 62 9 11 2 3 11 2

1:3 0 90 6 0 0 0 3 0
1:3 2 10 0 50 13 10 13 3
1:3 30 10 0 49 11 10 17 1

a31
P shift, Ôppm.

transformation to the disulphide and ZnDRT is highly probable. It is sug­


gested, therefore, that in addition to ZnDRT, part of the disulphide which is
formed at all [CHP]/[ZnDRP] molar ratios must originate from the basic-
zinc complex; the remaining amounts are derived directly from zinc dithio-
phosphate itself. In addition to the contribution of the disulphide to the
formation of thiophosphoric acid, the zinc thiophosphate, ZnDRT, must also
make some contribution to the formation of this acid, as the disappearance
of the latter at the end of the reaction of CHP with ZnDRP (e.g. molar ratio
5) is paralleled by the appearance of the acid DRTPA, see Table 7. Thiophos­
phoric acid, must therefore, be formed from both the disulphide and ZnDRT.
This is in agreement with the finding of Rossi [83] that transformation of
ZnDRT leads to phosphoryl disulphide (DRDOS), and this was shown here
to be the oxidation product of the acid DRTPA. At higher molar ratios of
peroxide to zinc dithiophosphate (e.g. 50) the formation of thiophosphoric
acid predominates, see Table 7, and this leads to the main catalysts re­
sponsible for ionic decomposition of hydroperoxides. Similarly, the forma­
tion of the acid was found to dominate at higher molar ratios of peroxide to
thiophosphoryl disulphide. Thiophosphoric acid was also found to be one of
the major product formed during reactions of nickel dithiophosphate and
thiophosphoryl disulphide with hydroperoxides at high molar ratios.
In the presence of excess hydroperoxide, the thiophosphoric acid (DRTPA)
may oxidise further to the corresponding dialkyl phosphoryl disulphide
(DRODS) and both can subsequently give rise to dialkyl phosphoric acid
(DRPA, d - -0.9) and SO2/SO3. Phosphoric acid was shown [74] to be
unreactive towards hydroperoxides and is, therefore, a stable reaction pro­
duct. Oxidation of thiophosphoric acid, gives, in addition to DRPA, esters of
202 S. ALrMALAIKA

dithiophosphoric acids such as TP, XLIX and TIP, XLL. These esters were
also observed [86] by other workers as products of the reaction of ZnDRP
with CHP at low temperatures. DRODS has been suggested by others from
studies at lower temperatures although different routes to its formation
were proposed. Scheme 11 summarises the mechanism of inhibition by Zn
DRP based on the above finding.

//
s 0
II
(RO)2P (RO)2P
\ \
OR SR

TP,XLIX TIP,XLL

GO Nickel Dithiophosphate (NiDRP)


Al-Malaika and Scott have shown [65] that decomposition of hydroperox-
ides by nickel dithiophosphate at high temperatures (e.g. CHP at 110°C)
occurs in three stages (Fig. 5). An initial rapid catalytic stage which gives
rise to homolytic products (Fig. 12a) followed by an induction period, during
which, oxidation products are formed which are the precursors of the second
stage catalysts (see Fig. 12b). The induction period leads to the slower
catalytic stage which assumes greater significance at higher molar ratios of
hydroperoxide to complex.
Figure 12a shows that for [CHP]/NiDRP]=100, at 110°C a clear change­
over is obtained from an essentially homolytic to an essentially heterolytic
reaction which predominates the latter stages of the reaction. The catalytic
free radical pathway, on the other hand, is more important at low molar
ratios a-cumyl alcohol and acetophenone (the homolytic products of CHP
decomposition) are exclusively formed at molar ratios of [CHP]/NiDRP] <
10, see Fig. 5, inset. The relative contribution of the initial catalytic process
to the overall stabilising effect of NiDRP was found [57] to decrease with
increasing complex concentration. A linear relationship was found [57]
between the amount of hydroperoxide decomposed during the first catalytic
stage and the concentration of the nickel complex. By contrast, an inverse
relationship was found [65] between the length of the secondary induction
period and the concentration of the nickel complex at constant peroxide
concentration. The length of this apparent period of inactivity was directly
related to the time required to achieve a critical threshold concentration of
the active catalyst from the nickel complex. The opposite relationship was
reported by Burn [67] for reactions of ZnDRP with CHP (at lower tempera­
ture, 70°C), where the length of the induction period was found to increase
with increasing initial zinc complex concentration. This observation led
Burn to discount the formation of active catalyst from the metal complex
ANTIOXIDANTS— PREVENTIVE MECHANISMS 203

a-cumyl alcohol phenol


70 U -•

1.0
[NiDBP]= 1 x 10 4 M
[CHP]=1x i a 2 M
O c l a t 110°C

Q.
Z
Ü
a-Methyl styrene
0.5

Acetophenone

**■ a-cumyl alcohol


i U.
10 20 30 40 50 60
Reactiontime (min)

10 15 20 25 30
REACTION TIME (MIN)

Fig. 12. Kinetics of ionic and radical products formation of the reaction of CHP (1-KT2 M)
with NiDRP, R = n-butyl (MO"4 in chlorobenzene at 110°C (a). The CHP decomposition
curve for the same reaction is also shown, (b) Formation and decay of products formed
from NiDRP in the above reaction.
204 S. AL-MALAIKA

A A
(RO)2Px J4i JWR) 2
NiDBP
(a)
ROOH

S S
// % s s
(RO)2P P(OR)2 J£- (RO)2pf- +HO-Nif Vo^a + RO'
\
S—S
/ \ V

v(RO)2P* +HO—Ni—S + 2ROH


(RO)2P
(k) nROOHW)

NiS04 nH20

//
(R02)P0
VoH
II DBMS
O
BTSA (RO)2P—OR

.1 0,0,0-TBTP

(RO)2P—OH + S0 2
DBTnPA

(RO)2P—SH
DBTPA

Scheme 13. Antioxidant mechanism NiDBP.


ANTIOXIDANTS — PREVENTIVE MECHANISMS 205

under these conditions. It was further shown [57] that at all molar ratios of
hydroperoxide to nickel dithiophosphate, the nickel complex is completely
destroyed before the onset of the secondary catalytic stage (see Fig. 12b).
This confirms the earlier findings (see Section 1.2.2) that the metal com­
plexes are a precursors for the effective catalytic peroxide decomposers.
The nature of the ionic catalystfe) for peroxide decomposition formed from
reactions of NiDRP with hydroperoxides at different temperatures (25-
150°C) was investigated [80,81,88]. Thiophosphoryl disulphide (DRDS) was
found [65,81] to be a primary intermediate of the oxidation of nickel dithio­
phosphate at high molar ratios of peroxide to complex and at elevated
temperatures (e.g. 110°C). The formation of DRDS is associated with the
first rapid homolytic peroxide decomposition stage. The disulphide is
further oxidised to unstable intermediates which breakdown to give the
catalysts responsible for the effective hydroperoxide decomposing activity of
the nickel complex. Figure 12b shows that sulphonoic (DRDSO) and thiono-
phosphoric (DRTnPA) acids are formed slowly during the induction period,
the former loses S 0 2 to give the latter (see Scheme 130 which can be seen
to be a relatively stable product. Figure 12b also shows that although CHP
is decomposed at the beginning of the reaction, no phenol (the ionic decom­
position product) was found initially during the build-up of the disulphide-
derived oxidation products. Instead the formation of phenol was found to
coincide with the onset of decomposition of oxidation intermediates. This
observation clearly identifies the antioxidant activity with the sulphur acids
formed from NiDRP by hydroperoxide oxidation.
Examination of products formed from reactions of NiDRP with TBH at
room temperature using P NMR [79,80] shows that the nature and propor­
tion of the transformation products is highly dependant on the ratio of the
peroxide to complex. Monosulphide, disulphide, thiono- and thiophosphoric
acids are the major transformation products which are observed during
oxidation reactions of NiDRP with CHP (or TBH) at different molar ratios
(see Fig. 13). A small amount (about 10% of the total products) of other
phosphorous-containing species, mainly esters of dithiophosphoric (DTP)
and thiophosphoric (TP) acid (Scheme 13g and j) are also formed. Thiophos­
phoric acid (DRTPA) which becomes more evident at higher molar [TBH]/
NiDRP] ratios (Fig. 13) is probably formed via reaction of the thiophosphoryl
radical with the hydroperoxide (Scheme 13, reactions d-f and k). Under the
low temperature conditions used in these studies, the formation of thiono-
phosphoric acid from the disulphide (route 1-f) seems to be unlikely in the
light of the observation [89] that disulphides of dithioic acids are unable to
effect thermal decomposition of TBH at temperatures below 70°C. Further­
more, the concentration of the thio- and thiono- acids in the system was
found [80,79] to be a function of the amount of available hydroperoxide.
Scheme 13 summarises the reactions involved in the antioxidant function of
nickel dithiophosphate.
206 S.AL-MALAIKA

[ T B H ] / [ N Î D B P ] =1 [TBH]/[NiDBP] = 4

94 60

94 94

NiDBP

J1
85 79

[TBH]/[NiDBp]=2 79 [TBH]/[NIDBP]=10
94

21

100

JJUL
95 83 8 60 17

4 5 6 7 10
Molar ratio of [TBH]/[NiDBP]

Fig. 13. Major product yield after complete reaction of NiDRP, R = n-butyl with TBH in
cyclohexane at 25°C at different molar ratios [TBHl/[NiDBP]: (a) NiDBP (Ô = 94 ppm);
(b) DBMS (Ô = 79 ppm); (c) DBDS (Ô = 85 ppm); (d) DBTnPa (Ô = 63 ppm), DBTPA (Ô = 21
ppm) and (BuOtePSH (Ô = 60 ppm). Inset shows 31 P NMR spectra of products formed at
the end of reactions of NiDBP (0.3 M) and TBH at different molar ratios in cyclohexane
at 25°C. Numbers on signals are chemical shifts in ppm.
ANTIOXIDANTS — PREVENTIVE MECHANISMS 207

(Hi) Iron dithiophosphates (FeDRP)


Much less work has been done on reactions of FeDRP with hydroperox-
ides and its role as an antioxidant, when compared with other metal
dithiophosphates, e.g. Ni and Zn analogues. The apparent lack of interest in
FeDRP must be attributed to earlier reports [90,91] of the instability of iron
complexes of dithiophosphoric acids in the presence of traces of hydrogen
chloride when iron (III) chloride was used in their preparation. Al-Malaika,
Scott and Smith [78b] studied reactions of stable iron (III) dialkyl dithio-
phosphate (using a recent preparative method [92]) with hydroperoxides at
high temperatures and evaluated its antioxidant role. Iron dithiophosphate
was found to be a reasonably good inhibitor of oxidation of hydrocarbons (e.g.
decalin at 130°C) in the absence of peroxides but it is totally ineffective when
excess hydroperoxide is present. The behaviour of the iron complex when
used as an inhibitor for the non-catalysed oxidation of decalin at 130°C (Fig.
14a) was shown to be analogous to that of the corresponding zinc complex
(examined under the same conditions) in that the length of the induction
period increases with increasing complex concentration but, on the whole,
the iron complex is less effective. In studies of the effect of MRDP on AIBN
catalysed oxidation of cumene at 60°C, Burn [93,94] found that the iron
complex is a good trapping agent for peroxyl radicals.
The decomposition behaviour of FeDRP in the presence of CHP (at 110°C)
was found [78b] to be similar to that shown by other metal complexes (e.g.
Ni and Zn). It was shown to involve a two-stage decomposition process —
the relative importance of each stage depending on the initial [CHP]/
FeDRP] molar ratio. The first homolytic stage (Fig. 14b) gives acetophenone
as the major product; this stage predominates at high concentrations of the
iron complex (see inset of Fig. 14b). The predominance of the free radical
process at lower molar ratios (<5) must be responsible for the inability of
FeDRP to prevent the CHP-initiated oxidation of decalin (inset Fig. 14a).
Burn [94] also found that the effectiveness of FeDRP as an antioxidant for
cumene (at 60°C) was seriously affected by the presence of CHP. The first
stage of the FeDRP/ hydroperoxide reactions is accompanied by a colour
change: the initial dark green colour which changes to orange before forming
an orange/ brown inorganic precipitate (most likely Fe2Û3 or its hydrated
form) with a clear colourless supernatant liquid [78b]. This suggests that all
the iron is oxidised by the end of the first stage. The corresponding thio-
phosphoryl disulphide (DRDS) was identified (by P NMR) as the major
phosphorous-containing product present in the supernatant liquid. The
second stage of CHP decomposition is an ionic process with phenol as the
major product (see inset Fig. 14b); this stage predominates at higher [CHP]/
[FeDRP] molar ratios). The ionic decomposition associated with this stage
must be caused by sulphur acids which are the major oxidation products of
disulphides formed during the first stage of CHP-initiated decomposition of
the iron complex.
208 S. AL-MALAIKA

1
|| I
E
a - 20
E §
- 30 - i
c -3
• e- io
8 11
o I 2x10
*> " '
'o. 20 I
S- 20
o
0) *
w J O Time, min
■°
"3
< »*10
-3 .
5x10*/
CM 1 0
o J _^ ^^\
20 40 60
Time , h

0 10 4 20
[FeDIPP] , x 10 M

100

50

0.0
15 30 45
Time , min.

Fig. 14. (a) Effect of FeDRP, R = isobutyl, on the oxidation of decalin at 130°C in the
absence and presence (inset) of MO"2 M CHP. The molar concentrations of FeDi-BP are
given on the curves, (b) Kinetics of products formed during reaction of CHP (MO -2 M)
with FeDBP, R = isobutyl (20110"4 M) in chlorobenzene at 110°C. The CHP decomposi­
tion curve for the same reaction is also shown. Inset shows changes in concentrations of
phenol and acetophenone (ionic and radical products of CHP) formed at the end of
reactions of CHP with FeDi-BP (chlorobenzene, 110°C) at different molar ratios.

2. STOICHIOMETRIC PEROXIDOLYTIC MECHANISM (PD-S) AND THE ROLE OF


PHOSPHITE ESTERS

The main requirement of stoichiometric peroxide decomposers is that


they should substantially reduce hydroperoxides to alcohols without the
substantial formation of free radicals. The most widely used PD-S antiox-
idants are the phosphite esters (I) of which the trisnonyl-phenyl phosphite
(I,R = C 9 H 19 0-Ph-) was first used to stabilise SBR in the early 1940s. They
are now also used commercially in combination with hindered phenols to
stabilise polyolefins. It has been known since the late 1950s [95,96] that
phosphite esters act mainly by reducing hydroperoxides to alcohols and the
ANTIOXIDANTS — PREVENTIVE MECHANISMS 209

corresponding phosphate esters in a stoichiometric reaction (Reaction (4)).


In addition to their stoichiometric peroxidolytic activity (one molecule of
peroxide is decomposed by one phosphite molecule), some phosphite esters
act also as catalytic peroxidolytic agents in addition to having chain-break­
ing activity; the contribution of each of these modes to the overall mecha­
nism

(RO)3P + ROOH > (RO) 3 P=0 + ROH (4)

depends on the structure of the phosphorous compound, the oxidisability of


the substrate and the reaction conditions. Studies of reactions of a number
of different phosphites with hydroperoxides (e.g. CHP, TBH) in solution has
been conducted by several groups [97-100] who demonstrated that their
reactivity towards hydroperoxides is governed by both steric and polar
effects of groups bound to the phosphorous atoms. Reaction rates decrease
with increasing bulk and electron accepter ability of these groups, hence,
simple saturated alkyl and cycloalkyl phosphites are more reactive than
aryl and hindered aryl phosphites [100].
Five-membered catechol phosphites, such as LI have been shown to
decompose hydroperoxides catalytically in an ionoic pathway, e.g. with
CHP, phenol and acetone are the major products, [97-100,102,103]. The
catalysts for peroxide decomposition are formed in an initial fast stage
during reactions of the phosphites with hydroperoxides. Scott and co-work­
ers [97-99] and later Schwetlick and co-workers [100,102,103] studied the
nature of transformation products formed from reactions of LI with CHP.
Two different phosphates are formed, the first, LII, which is formed by

20 40
Time , min.

Fig. 15. Products of reaction of 2,6-di-fe/tf-butyl-4-methylphenyl-o-phenylene phosphite


with cumyl hydroperoxide in chlorobenzene at 50°C.
210 S.AL-MALAIKA

stoichiometric oxidation of LI by hydroperoxide, builds up to a maximum


before it decays to give the second phosphate, LIII (see Fig. 15 and Scheme
14). The cyclic phosphate, LII, which is a powerful Lewis acid [98], is
transformed to the open chain hydrogen phosphate, LIII, either by a fast
reaction with water (e.g. dehydration of a-cumyl alcohol), and/or by a slower
reaction with CHP to give an unstable peroxyphosphate, LIV, which is the
precursor to the real peroxidolytic catalyst, LIII, in addition to phenol and
acetone [103,102]. Further reaction of LIII with CHP gives also phenol and
acetone (Scheme 14).
In contrast to the catalytic peroxidolytic action of the phosphite LI, the
cyclic ethylene phosphites, e.g. LV, do not decompose peroxides catalytically

CH3 CH,

( g j ^ >~OPh + < g ) ^ cI _ 0 0 H <Ô^Ç-OH


CH, CH,
LI CHP
LII

CH,

OH / O - o - c n f O )
W
CHP^ [OJL / = 0 '
CH 3
O OPh
LIV

-PhOH

@oc
-Me2C=0

LII
OH OH

\ ^ r»
O VOPh
IP

LIII

o
Me.
OPh = O -
<§V OH +
Me
.C=0

Scheme 14. Antioxidant action of cyclic phosphites.


ANTIOXIDANTS — PREVENTIVE MECHANISMS 211

at low temperatures (below 50°C), although they do so at higher tempera­


tures (e.g. 75°C) [100] giving products similar to those obtained from LI,
except for the fact that the open-chained hydroxyethyl phosphate, LVI are
unstable under the reaction conditions and condenses to give polyphos-
phates (Reaction (5)). Other phosphite derivatives which have been shown
to decompose hydroperoxides catalytically include arylamidophosphites,
thiophosphites, phosphorous acid esters of thiobisphenols and some diaryl
hydrogen phosphites [100].

H Q/RO
r V_ O P h 522g f ° P ^ ° * °*? HO
o' V OPh " ^ C T ^OPh
LV

3. METAL ION DEACTIVATION

It was mentioned earlier in this chapter that deactivation of metal ions is


only of limited effectiveness and hence it will be covered only briefly. The
pro-oxidant effect of transition metal ions in catalysing the decomposition of
hydroperoxides to free radical products is now well established and has been
discussed in detail in Chapter 2 of this volume. It was seen in Chapter 2 that
it was recognised as early as 1954 that catalysis occurs by a redox mecha­
nism (most effective ions are those which can easily undergo a one electron
oxidation or reduction) via the formation of unstable metal-hydroperoxide
co-ordination complexes, followed by electron transfer to give free radical
products [104]. Therefore, it was demonstrated soon after, that addition of,
for example, hexadentate or octadentate ligands, which co-ordinate metal
ions to their maximum extent, renders metal ions ineffective as catalysts
[105]. Factors which contribute to the effectiveness of metal ions as redox
catalyst such as the stability and ease of co-ordination of the formed metal
complexes and the redox potential of the metal in the complex have been
reviewed and the early literature was covered in the first edition of this book
[1] and will be only discussed briefly.

3.1 Stability of Metal Complexes

Factors which effect metal complex stability are briefly outlined below.

(i) Effect of base strength


It is known that in the formation of a co-ordinate bond, the metal ion
functions as a Lewis acid while the ligand operates as a Lewis base [106].
212 S.AL-MALAIKA

The base strength of a series of related electron donor molecules is, there­
fore, related to their chelate stabilities with a given metal ion. Table 9
[1,107] shows that the strongest complexes are formed with strongly elec­
tron releasing groups (e.g. OMe) and the weakest with electron attracting
(e.g. NO2). Table 10 demonstrates [1,108] the importance of this principle in
metal deactivation where the increasing electron releasing power associated
with oxygen, sulphur and nitrogen is reflected in a paralleled capacity to
inhibit the pro-oxidant effects of copper.

TABLE 9

Effect of electron releasing groups on the bond strength of copper chelates [1,107]

00
HC=N N=CH

x x
X «$ * (volts)

N0 2 +0.03
S03Na -0.09
O -0.10
H -0.12
Me -0.15
OH -0.17
OMe -0.21

*The more negative is el/2 the greater is the stability of the metal complex.

(ii) Effect ofpolydenticity


The formation of a favourable arrangement of ligands around the metal
ion is quite important. The greater the number of chelate rings that can be
formed per ligand, the greater is the chelate stability (see Table 11 [1,106])
due to the favourable entropy change on the formation of polydentate
chelates (i.e. the probability of the metal escaping from the complex is
reduced the greater the polydenticity) [109,110] and the favourable elec­
tronic effect of one complexed grouping upon a neighbouring one in the same
molecule [111].
ANTIOXIDANTS — PREVENTIVE MECHANISMS 213

TABLE 10

Influence of electron releasing qualities of a ligand on metal deactivating efficiency


[1,108]

Compound X ED

O 0
HOCH2CH2XH S 5
NH 25

,COOH O 5
M S 15
NH 45

TABLE 11

Effect of polydenticity upon metal complex stability [1,106]

Ligand LogKuA*
Cu(II) Ni(II) Co(II)

NH2CH2COOH 8.6 6.2 5.2


NH(CH2COOH)2 10.6 8.2 7.0
N(CH2COOH)3 12.7 11.3 10.6

_ [MA*-m]
MA
where Mn+ + Am"= MA"\-m
"[M a+ ][A m T

(Hi) Effect of chelate ring size


The effect of increasing the size of the ring formed by co-ordination is
demonstrated in Table 12; as the ring involving two nitrogens becomes
bigger, the chelate formation constant (Table 12a [1,112]) and the efficiency
of metal deactivator (Table 12b [108]) decrease to lower value.

(iv) Effect of ligand bridge


The effect of tying together of two chelating groups with a bridge of the
appropriate chain length is demonstrated in Table 13 both by direct stability
measurements [1,107] and by the deactivating efficiency of the chelating
agent [108].
214 S. AL-MALAIK4

(v) Effect ofsteric hindrance


Alkyl substitution can be disadvantages when it hinders the approach oi
the chelating group to the metal ion which is reflected in a decreased meta]
deactivating efficiency (Table 14) [1,108].
It was demonstrated from all examples above that strong complex forma­
tion does lead to effective deactivation. However, although strong chelatior
is necessary, it is not in itself a sufficient criterion of an efficient meta]
deactivator. Table 15 [1,113] shows the effect of variety of complexing agents
on the common transition metal auto-oxidation catalysts. It is clear thai
whilst all the tetradentate chelating agents were effective with copper, thej
were not uniformly effective with the other transition metals whose maxi-
TABLE 12a

Calcium chelate stabilities involving alkylenediamine-tetraacetic acids [1,112]

HOOCCH2 CH2COOH
\ /
N(CH)2)n N
/ \
HOOCCH2 CH2COOH

n Log ÜTcaA Log ÜTca2A

2 10.5
3 7.1 <0.7
4 5.2 2.0
5 4.6 2.6

TABLE 12b

Effect of ring size on the copper deactivating efficiency of diamino alkanes [1,108]

NH2(CH2)nNH2
n ED (%)

2 70
3 65
4 35
6 15
ANTIOXIDANTS — PREVENTIVE MECHANISMS 215

TABLE 13
Effect of ligand bridge on copper chelate stability [1,107]

HC=N N=CH
I I
Ri R2

Äi Ä2 eS) (volts) ED(%)

Me Me +0.02
Et Et 5

-CH 2 -CH,- -0.75 100


-0.12

ö -0.76

TABLE 14
Effect of steric hindrance on copper deactivating efficiency [1,108]

OL
.OH

1
R

R ED(%)

H 95
Me 85
Et 65

mum coordination numbers are greater than four. The only chelating agent
which was universally effective was the octadentate chelating agent e (in
Table 15). The inability of the quadridentate chelating agents, such as a (in
Table 15), to deactivate metal ions of co-ordination number six seems to be
associated with the ability of hydroperoxides to enter the co-ordination shell
of the metal with consequent electron transfer reactions.
216 S.AL-MALAIKA

TABLE 15

Effects of chelating agents on the autoxidation of petroleum catalysed by metals [1,113]

Chelating agent (0.002%) ED metal oleate molar concn . = 1.6X10"8


Mn Fe Co Ni Cu

-103 -43 -833 100


\^CH=NCH2CH2N=CH^^

*Q - 0 -96 100 100

c ^v^CH=NOH

coo -84 100 96 -55 100

<cor -73 100 - -124 100

e
Q
\>^CH=NCH 2 .
4C
100 100 100 100 100

3.2 Redox Properties of Metal Complexes

The activating effect of some chelating agents upon the pro-oxidant


activity of metal ions implies that, it is not only necessary to inhibit the
co-ordination of hydroperoxides into the valency shell of the metal, but that
its redox properties are more favourable to the achievement of the optimum
ratio of the rates of the electron transfer processes (Reaction (6a and b)). It
is well known that the effect of electron donation from ligand to metal, the
charge residing on the complex and the electronic structure of the complexed
metal atom all affect the redox potential of metal chelates.
The very interesting phenomenon of catalyst-inhibitor inversion which
was reported by many groups of workers ([114] and references therein) was
discussed in earlier chapters. According to the above principle, if the concen­
tration of metal ions (e.g. Co+2, Cu+, Mn+2) is increased beyond a certain
value, catalysis of autoxidation abruptly ceases and catalysed oxidation
does not resume until a lengthy induction period has ended. Many metal
chelates, such as, Schiffs bases, oximes, and salicylato chelates, e.g. nickel
ANTIOXIDANTS— PREVENTIVE MECHANISMS 217

HO-OR
N
M / + OH+OR
/ : \ (6)
HO-OR
\:+/ b
M — M + H + OOR

complexes of butylsalicylaldimino, NiNBS, diisopropyl-salicylato, NiDIPS,


and oximes, NiOX and FeGIDOX, have been reported to exhibit catalyst-in­
hibitor inversion when present in autoxidation systems in both hydrocarbon
liquid models and in polymers, e.g. PE [115-117]. It was proposed [59] that
these chelates can remove the chain propagating radicals by either oxida­
tion or reduction with the transition metal ion, a mechanism which was
initially suggested by Black [118]. However, it was argued that the Black
mechanism cannot account for the reported antioxidant effectiveness of
ZnNBS [116] since the zinc ion has only one stable oxidation state. Uri [116]
provided an evidence that ZnNBS functions as a radical trap and shows
synergism with peroxide decomposers. It is possible, therefore that the
antioxidant activity of ZnNBS is due to reaction of alkylperoxyl radicals
with the ligands.

i-Pr
Ni
CH=N
n-Bu Me OH

NiNBS NiDIPS NiOX

Ranaweera and Scott [119] and Benbow and coworkers [120] have pro­
vided an alternative explanation of the dual behaviour of certain NBS, DIPS
and OX chelates as inhibitors of oxidation (e.g. see Scheme 15 for NiNBS).
The metal chelate removes alkylperoxyl radicals in two one electron transfer
reactions to form stabilised phenoxyl radicals [120]. The radical formed from
the ligands of NBS, DIPS and OX chelates would be strongly resonance
stabilised and therefore unlikely to propagate the kinetic chain. It was
suggested that initially the chelate functions as an antioxidant by destroy­
ing alkylperoxyl radicals by the two reactions shown in Scheme 15. At the
end of the induction period, and when the metal chelate has reacted, rapid
oxidation occurs due to metal-catalysed decomposition of hydroperoxides.
218 S. AL-MALAIKA

Benbow and coworkers [120] also explained Uri's finding [108] of the inhibi­
tion function of ZnNBS on the same basis of reactions of Scheme 15.
Ranaweera and Scott [121] have also reported that NiOX destroys hydro-
peroxides in a stoichiometric reaction in which the major product is the
alcohol. Similar product distribution were obtained in the presence of UV
light and heat and a non-radical mechanism was proposed, Scheme 16, to
account for the products of the chelate decomposition. Many of these che-
lates (e.g. Schiff bases and oxime chelates) have been shown to also act by
chain-breaking mechanism.

N=CH

ROO +

NiNBS

N = CH RN=CH

+ ROO

RN = CH

+
"O—(O/ ROO-+ M 2 +
Scheme 15. Antioxidant action of Ni NBS

4. ULTRA-VIOLET LIGHT DEACTIVATION

In spite of the absorption of the shorter wavelengths of the sun's rays by


the atmosphere which reduces the harmful near-UV light (below 290 nm) to
a very small amount, there remains enough energy in the residue to break
a variety of chemical bonds. In polymers, chromophoric groups are almost
always present either as adventitious impurities from their manufacture or
as oxidation products formed during their processing and fabrication, con­
sequently most polymers are subject to photodegradation (see Volume II,
Chapter 9).
The earliest class of UV absorbers to be developed commercially was
based on the principle of screening the polymer from the damaging effect of
ANTIOXIDANTS — PREVENTIVE MECHANISMS 219

H OR
\ /
O

CH, CH, Of, I I "CH 3


HO OH HO OH
NiOX

OH

®( C
II
,CH 3 + 0 = N —Ni

+ ROH
o

H
<

@C° ^ ©C ^CH,
H
I
( O X /OH + ROCH
3

O
I
R

Scheme 16. Antioxidant action of Ni OX.

harmful photons by eliminating or reducing them below the threshold level


needed to initiate photo-oxidation. Pigments or UV absorbers which reflect
or absorb the harmful UV light are widely available. The role of UV ab­
sorbers as stabilisers is discussed in details in other chapters of this book
and will only be briefly discussed here.
Colourless UV absorbers which have high extinction coefficient in the
region 300-350 nm, primarily by absorbing UV-light and re-emitting it as
vibrational (thermal) energy without being themselves destroyed and with­
out sensitising the oxidation of the substrate, eg. ortho-hydroxybenzophen-
220 S. AL-MALAIKA

ones, LVII (e.g. Cyasorb UV 531) and ortho-hydroxybenztriazoles, LVIII


(e.g. Tinuvin 327). The stabilizing effect of the -OH group, therefore, results
from a rapid tautomerization in the excited state (see Reaction (7)) as­
sociated with the release of harmless thermal energy [122].

0 OH HO Rt

R2

LVII LVIII
Cyasorb UV 531 Tinuvin 327, Rx = R2 - t-Bu

H H
/ \ / \
o o o o (7)

It was shown [123] that the concentration of hydroxybenzophenones, e.g.


UV 531, decreases while it exerts its photostabilising action in polymers.
These loses can not be due to instability of the hydroxy benzophenones, to
UV-light, because they are extremely stable to UV-irradiation [123,124].
The change in their concentration was found to be minimal in solution in the
absence of oxygen [125]. Furthermore, hydroxybenzophenones were shown
to be ineffective peroxide decomposers and thermal oxidative stabilisers at
elevated temperatures [126], but they show a mild antioxidant activity at
lower temperatures [124]. It was demonstrated that this class of compounds
function, at least in part, as sacrificial antioxidants removing chain-initiat­
ing radicals formed from hydroperoxides and derived carbonyl compounds
[124,126], the loss of 2-hydroxybenzophenones during photoxidation is,
therefore, largely due to their activity as free radical scavengers. Fuller
discussion on UV light deactivators is given in Volume II.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 221

REFERENCES

1 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965.


2 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965, p. 363
3 ANON, US Government Regulations Governing Meat Inspection (1949), US Dept.
of Agriculture, Washington, DC.
4 E.H. Farmer, J. Faraday Soc., 38 (1942) 356.
5 W.P. Fletcher and S.E. Fogg, Rubber J. Ind. Plast., 134 (1958)16; 135 (1958) 687.
6 J.R. Dunn and J. Scanlan, Trans IRI, 34 (1958) 228.
7 J.R. Thomas, O.L. Harle, W.L. Richardson, and L.O. Bowman, Additives in Lubri­
cants Symposium, ACS Div. Pet. Chem., 1956, p. 138.
8 J.R. Dunn and J. Scanlan, J. Applied Polym. Sei., 35 (1959) 267.
9 B. Baum and A.L. Perun, ACS Meeting, Div. Org. Coatings Plast., 21 (2) (1961) 79.
10 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965, p. 295.
11 J.R. Shelton, in W.L. Hawkins (Ed.), Polymer Stabilisation, Wiley, New York, 1972,
Chapter 2.
12 G.H. Denison, Ind. Eng. Chem., 36 (1944) 477; G.H. Denison and P.C. Condit, Ind.
Eng. Chem., 37 (1945) 1102.
13 G.W. Kennerley and W.L. Patterson, Ind. Eng. Chem., 48 (1956) 1917.
14 H. Hock and S. Lang, Ber., 257 (1944).
15 M.S. Kharasch and J.G. Bart, J. Org. Chem., 16 (1951) 128; M.S. Kharasch, A. Fono
and W. Nudenberg, J. Org. Chem., 15 (1950) 748.
16 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965, p. 433
et seq.
17 Hercules, British Patent 851, 670 (1960).
18 J.D. Holdsworth, G. Scott and D. Williams, J. Chem. Soc., (1964) 4692.
19 W.L. Hawkins and H. Sautter, J. Polym. Sei., AI, 7 (1963) 3499.
20 L. Bateman and J.I. Cunneen, J. Chem. Soc., (1955) 1596.
21 D. Barnard, L. Bateman, M.E. Cain, T. Colclough and J.I. Cunneen, J. Chem. Soc.,
(1961) 5339.
22 L. Bateman, M.E. Cain, T. Colclough and J.I. Cunneen, J. Chem. Soc, (1962) 3570.
23 L. Bateman, and K.R. Hargrave, Proc. R. Soc, A224 (1954) 389, 399; D. Barnard,
K.R. Hargrave and G.M. Higgins, J. Chem. Soc. (1956) 2845.
24 J.R. Shelton, in G. Scott (Ed.), Developments in Polymer Stabilisation-4, Applied
Science Publishers, London, 1981, p. 23.
25 C. Armstrong and G. Scott, J. Chem. Soc, (1971) 1747.
26 C. Armstrong, M.A. Plant and G. Scott, Eur. Polym. J., 11 (1975) 161.
27 C. Armstrong, M.J. Husbands and G. Scott, Eur. Polym. J., 15 (1979) 241.
28 N.P. Neureiter and D.E. Bown, Ind. Eng. Chem., Prod. Res. Div., 1 (1962) 236.
29 B.A. Marshall, in R.F. Gould (Ed.), Stabilisation of Polymers and Stabilisation
Processes, American Chemical Society, Washington, DC, Adv. Chem. Ser. 85,1968,
p. 140.
30 C.A. Kingsbury and D.J. Cram, J. Am. Chem. Soc, 82 (1960) 1810.
31 T. Colclough and J.I. Cunneen, Chem. Ind., (1960) 626.
32 C.H. Bepuy and R.W. King, Chem. Rev., 60 (1960) 431.
33 J.R. Shelton and K.E. Davis, Int. J. Sulf. Chem., 8 (1973) 205.
34 J.R. Shelton and K.E. Davis, Int. J. Sulf. Chem., 8 (1973) 197.
35 P. Koelewijn and H. Berger, Rec Trav. Chem., 91 (1972) 1273; 93 (1974) 63.
222 S. AL-MALAIKA

36 E. Block, J. Am. Chem. Soc., 94 (1972) 644.


37 G. Scott, in G. Scott (Ed.), Developments in Polymer Stabilisation-6, Applied Science
Publishers, London, 1983, p. 29.
38 M.J. Husbands and G. Scott, Eur. Polym. J., 15 (1979) 249.
39 B.D. Flockhart, K.J. Ivin, R.C. Pink and B.D. Sharman, Chem. Comm. (1971) 339.
40 G. Scott, Atmospheric Oxidation and Antioxidants, Elsevier, London, 1965, p. 396.
41 G. Scott, Eur. Polym. J., Suppl. (1969) 189.
42 W.L. Hawkins and H.J. Sautter, Polym. Sei., AI (1969) 3499.
43 J.C. Kice, F.M. Parham and R.M. Simons, J. Am. Chem. Soc, 82 (1960) 660.
44 C. Chatgilialoglu, B. C. Gilbert and M.D. Sexton, J. Chem. Soc, Perkin II (1980)
1141.
45 E.G. Miller, D.R. Rayner, H.T. Thomas and K. Mislow, J. Am. Chem. Soc, 90 (1968)
4861.
46 D.J. Abbott and C.J. Stirling, Chem. Comm., (1968) 165.
47 V.M. Farzaliev, W.S.E. Fernando and G. Scott, Eur. Polym. J., 14 (1978) 785.
48 T. Colclough and J.I. Cunneen, J. Chem. Soc, (1964) 4790.
49 S.K. Ivanov, and D. Shopov, Compt. Rend. Acad. Bulgar Sei., 18 (1965) 845.
50 S. Al-Malaika in R.B. Seymour and R.D. Deanin (Eds.), History of Polymeric
Composites, VNU Science Press, The Netherlands, 1987, p. 223.
51 J.E. Stuckley, in G. Scott (Ed.), Developments in Polymer Stabilisation-1 Applied
Science Publishers, London, 1979.
52 L.A. Brooks, Rubber Chem. Tech., 36 (1963) 887
53 F.A.A. Ingham, G. Scott and J.E. Stuckey, Eur. Polym. J., 11 (1975) 783.
54 M.J. Husbands and G. Scott, Eur. Polym. J., 15 (1979) 879.
55 S. Al-Malaika, K.B. Chakraborty, G. Scott and Z.B. Tao, Polym. Deg. Stab., 10 (1985)
55.
56 S. Al-Malaika, K.B. Chakraborty, G. Scott and Z.B. Tao, Polym. Deg. Stab., 13 (1985)
261.
57 S. Al-Malaika, K.B. Chakraborty and G. Scott, in G. Scott (Ed.), Developments in
Polymer Stabilisation-6, Applied Science Publishers, London, 1983.
58 S.K. Ivanov, in G. Scott (Ed.), Developments in Polymer Stabilisation-3, Applied
Science Publishers, London, 1980.
59 H.S. Laver, in G. Scott (Ed.), Developments in Polymer Stabilisation-1, Applied
Science Publishers, London, 1979.
60 D.J. Carlsson and D.M. Wiles, J. Macromol. Chem., C14 (1976) 65,155.
61 S. Al-Malaika, A. Marogi and G. Scott, J. Appl. Polym. Sei., 30 (1985) 789.
62 S. Al-Malaika, A. Marogi and G. Scott, Polym. Deg. Stab., 10 (1985) 237.
63 S. Al-Malaika, A. Marogi and G. Scott, J. Appl. Polym. Sei., 33 (1987) 1455.
64 S. Al-Malaika, in G. Scott (Ed.), Mechanisms of Polymer Degradation and Stabilisa­
tion^, Elsevier Applied Science, London, 1990, Chapter 3.
65 S. Al-Malaika and G. Scott, Eur. Polym. J., 16 (1980) 503.
66 J.A. Howard, Y. Ohkatsu, J.H.B. Chenier and K.U. Ingold, Can. J. Chem., 51 (1973)
1543; J.H.B. Chiener, J.A. Howard and J.C. Tait, Can. J. Chem., 55 (1977) 1644; 56,
157.
67 A.J. Burn R. Ceciel and V.U. Young, J. Inst. Pet., 57 (1971) 319.
68 J.A. Howard, in W.A. Pryor (Ed.), Frontiers of Free Radical Chemistry, Academic
Press, 1980, p. 237 et seq.
ANTIOXIDANTS— PREVENTIVE MECHANISMS 223

69 K.B. Chakraborty, G. Scott and W.R. Poyner, Plast. Rubber Process AppL, 3 (1983)
59.
70 S. Al-Malaika, A. Marogi and G. Scott, J. Appl. Polym. Sei., 31 (1986) 1685.
71 S. Al-Malaika, A. Marogi and G. Scott, Polym. Deg. Stab., 18 (1987) 89.
72 D. Gilead and G. Scott, in G. Scott (Ed.), Developments in Polymer Stabilisation-5,
Applied Science Publishers, London, 1982.
73 V.G. Vinogradova and A.N. Zverev, Izv. Akad. Nauk SSSR, Ser. Kim., 10 (1975)
2217.
74 P. Sanin, I. Blagovidov, A. Vipper, A. Kuliev, S. Krein, A.K. Ramaya, G. Shor, V.
Sher and Y. Zasalavsky, Proceedings of the Eighth World Petroleum Congress, 5
(1971) 91.
75 M. Johnson, S. Korcek and M. Zinbo, Society Automotive Engineers, Technical
Paper SP-558 (1983) 71.
76 A. Bridgewater, J.Dever and M. Sexton, J. Chem. Soc., Perkin Trans., 11 (1980)
1006.
77 O. Grishina and V. Bashinova, Neftekhimiya, 14 (1974) 142, (Chem. Abst.
80:145654c).
78aS. Al-Malaika, M. Coker, G. Scott and P.J. Smith, J. Appl. Polym. Sei., in press.
78bS. Al-Malaika, G. Scott and P.J. Smith, in preparation.
79 S. Al-Malaika, M. Coker, G. Scott and P.J. Smith, J. Appl. Polym. Sei, 44 (1992)
1297.
80 S. Al-Malaika, G. Scott and M. Coker, Polym. Deg. Stab., 22 (1988) 147.
81 S. Al-Malaika and G. Scott, Polym. Comm., 23 (1982) 1711.
82 S. Ivanov and I. Kateva, Neftekhimiya, 18 (1978) 417.
83 E. Rossie and L. Imperato, Chimica Industria, 53 (1971) 838.
84 Y. Ohkatsu, K. Kikkawa and T. Osa, Bull. Chem. Soc. Japan, 51 (1978) 3606.
85 I. Shkhiyants, N. Voevoda, N. Komissarova, S. Chernyav, L. Kaya, V. Sher and P.
Sanin, Neftekhimiya, 14 (1974) 312
86 P. Kozak and V. Rabel, Scientific Papers of the Prague Institute of Chemical
Technology, D39 (1978) 141.
87 V. Sher, E. Markova, G. Kuzminav and P. Sanin, Neftakhimiya, 13 (1973) 876.
88 S. Al-Malaika and G. Scott, Eur. Polym. J., 19 (1983) 235
89 S. Al-Malaika and G. Scott, Eur. Polym. J., 24 (1983) 25.
90 C.K. Jorgenson, J. Inorg. Nucl. Chem., 24 (1962)1571.
91 L. Lebedda and R. Palmer, Spectrochim. Acta (A) 29 (1973) 1371.
92 W. Hopkins and P. Mitchell, Reading University, Unpublished work.
93 A.J. Burn, Tetrahedron, 22 (1966) 2153.
94 A.J. Burn, International Oxidation Symposium, 1 (1967) 323.
95 C. Walling and M. Rabinowitz, J. Am. Chem. Soc, 81 (1959) 12.
96 D.B. Denhey, W.F. Goodyear and Goldstein, J. Am. Chem. Soc, 82 (1960) 1393.
97 K.J. Humphris and G. Scott, J. Chem. Soc, Perkin Trans., 2 (1973) 826.
98 K.J. Humphris and G. Scott, J. Chem. Soc, Perkin Trans., 2 (1973) 831.
99 K.J. Humphris and G. Scott, Pure Appl. Chem., 36 (1973) 163.
100 K. Schwetlick, in G. Scott (Ed.), Mechanisms in Polymer Degradation and Stabilisa­
tion, Elsevier Applied Science, London, 1990, Chapter 2.
101 J. Holcik, J.L. Koenig, and J.R. Shelton, Polym. Deg. Stab., 5 (1983) 373.
102 K. Schwetlick, C. Ruger and R. Noack, J. Prakt. Chem., 324 (1982) 697.
103 K.J. Humphris and G. Scott, J. Chem. Soc, Perkin Trans., 2 (1974) 617.
224 S. AL-MALAIKA

104 G.L. Banks, A.J. Chalk, J.E. Dawsen and J.F. Smith, Nature, 174 (1954) 274.
105 A.J. Chalk and J.F. Smith, Trans. Faraday Soc., 53 (1957) 1235.
106 S. Chaberek and A.E. Martell, in Organic Sequestering Agents, Wiley, New York,
1959.
107 M. Calvin and R.H. Bailes, J. Am. Chem. Soc., 68, (1946) 953.
108 R.W. Watson and T.B. Tom, Ind. Eng. Chem., 41 (1949) 919.
109 R.K. Her, J. Am. Chem. Soc., 69 (1947) 724.
110 G. Schwarzenbach, Helv. Chim. Acta, 35 (1952) 2344.
111 R.L. Smith, in The Sequestration of Metals, Chapman and Hall, 1959, p. 133.
112 G. Schwarzenbach and H. Ackermann, Helv. Chim. Acta, 31 (1948) 1029.
113 C.J. Pedersen, Ind. Eng. Chem., 41 (1949) 924.
114 Z. Osawa, in G. Scott (Ed.), Developments in Polymer Stabilisation-7, Elsevier
Applied Science, London, 1984, Chapter 4.
115 A.T. Betts and N. Uri, Macromol. Chem., 95 (1966) 22.
116 N. Uri, Chem. Ind. (1967) 2060.
117 M.U. Amin and G. Scott, Eur. Polym. J., 10 (1974) 1019
118 J.F. Black, J. Am. Chem. Soc, 100 (1978) 527.
119 R.P.R. Ranaweera and G. Scott, Chem. Ind., (1974) 774.
120 A.W. Benbow, CF. Cullis and H.S. Laver, Polymer, 19 (1978) 824
121 R.P.R. Ranaweera and G. Scott, Eur. Polym. J., 12 (1976) 591.
122 D.J. Carlsson, A. Garton and D.M. Wiles, in G. Scott (Ed.), Developments in Polymer
Stabilisation-1, Applied Science Publishers, London, 1979.
123 W. Klopffer, Polym. Sei., Sympos., 57 (1976) 205
124 P. Vink, in G. Scott (Ed.), Developments in Polymer Stabilisation-3, Applied Science
Publishers, London, 1980.
125 D.J. Carlsson, D.W. Grattan, T. Suprunchuk and D.M. Wiles, J. Appl. Polym. Sei.,
22 (1978) 2217.
126 K.B. Chakraborty and G. Scott, Eur. Polym. J., 15 (1979) 35.
SUBJECT INDEX 225

S U B J E C T INDEX

Absolute rate constants in hydrogen Alkyl radical traps, 111


abstraction, 63 — as antioxidants, 14
Acylperoxyl Alzheimers disease, 32
— reactivity in autoxidation, 63 Aminyl radicals
— addition to double bonds, 65 — stability, 127
Activation energy — dimerisation, 128
— in autoxidation, 64 Aniline
Aging — as antioxidant, 4
— biological, 1, 32 Antifatigue agents
Alcohols — in rubber, 145 et seq.
— as inhibitors for oxidation, 6 Antioxidant activity
Aliphatic amines — of arylamine transformation
— as inhibitors for oxidation, 143 products, 139
Alkenes — critical concentration, 134
— addition and abstraction — of phenol transformation products,
mechanisms of autoxidation, 65 139
Alkylaminodiphenylamines — and structure, 9,132 et seq.
— as antifatigue agents, 28 Antioxidants, 2, 4 et seq.
Alkyl chain branching — activities of phenols, 6
— in phenolic antioxidants, 139 — biological, 32
Alkylperoxyl, 3, 6 et seq., 56 — catalytic, 5,142 et seq.
— addition to double bonds, 64 — chain-breaking acceptor, 10,140
— in autoxidation, 36, 56 — chain-breaking donor, 6,121
— disproportionate, 3 — chain-breaking mechanisms, 121
— hydrogen abstraction from et seq.
phenols, 6 — classification, 35
— intramolecular hydrogen — colour due to, 140
abstraction, 56 — effect on products of cumene
— reduction by Cu+ hydroperoxide decomposition, 109
Alkyl radicals — hydrogen donor, 7
— in autoxidation, 36, 55, 66 — measurement of activity, 131
— involvement in inhibition, 141 — peroxide decomposing, 35,161 et
— reaction with oxygen, 55 seq.
— as retarders of oxidation, 13 — peroxidolytic, 161 et seq.
Alkyl hydroxylamines — phenolic, 4
— in the catalytic mechanism of — preventive, 35, 161 et seq.
antioxidant action, 47 et seq. — reaction with oxygen, 137
— reaction with oxygen, 152 — structure-activity relationship,
— reaction with peracids, 153 et seq. 136
— reaction with peroxyl, 153 Antioxygens, 5
226 SUBJECT INDEX

Antioxonants, 26 Carotenoid pigments


Arylamines — photoprotective effect, 93
— as antioxidants in rubbers, 5 Catalase, 34, 35
Aryloxyl radicals — mechanism of antioxidant action,
— irradiation to give prooxidants, 125 38
— dimerisation, 10 — as preventive antioxidant, 35
— oxidation by hydroperoxides, 128 — as Synergist with Superoxide
— reaction with oxygen, 7 dismutase, 34, 35
— reaction with peroxyl, 127 Catalytic antioxidants, 5,142 et seq.,
— reaction with the substrate, 7 161 et seq.,. 184 et seq.
— resonance energy, 9 — in hydrocarbon substrates, 144
— secondary oxidation products as Catechol phosphates
antioxidants, 129 — as antioxidants, 209
— steric hindrance in, 9,10 — reaction with hydroperoxides, 209
Ascorbic acid (Vitamin C), 32 et seq.
— as Synergist with a-tocopherol, 32 Catechol phosphates
Atmospheric pollutants — as antioxidants, 210
— as photosensitisers, 85,100 CB-A antioxidants, 10 et seq., 36,140
Autoxidation, 1 et seq. et seq.
— free radical chain reaction, 2, 3, 46 CB-D antioxidants, 6 et seq., 36
— of hydrocarbons, 56 — structure-activity relationship,
— inhibited, 6,132 et seq. 136 et seq.
— involving addition to double Chain-breaking antioxidants (CB), 36
bonds, 64 et seq.
— of in dene, 58 — acceptor (see CB-A antioxidants)
— kinetics of, 46 et seq. — donor (see CB-D antioxidants)
— mechanism of, 36, 46 — catalytic, 16 et seq., 36
— of vinyl alcohol, 57 Charge transfer complex
Autoinitiators, 2 — between phenolic antioxidants and
Autoretardation peroxyl radicals, 138
— by sulphur compounds, 22 Chelating ability
— in metal deactivation, 212 et seq.
Benzaldehyde — ring size and, 213
— autoxidation, 6 Chemical plasticisation of rubber, 27
p-Benzoquinone Chemiluminescence, 70
— as antioxidant, 141 — in study of autoxidation
— as catalytic antioxidant, 143 Chromophores
Benzoylperoxyl — as photoinitiators, 29
— reactivity in autoxidation, 63 Chlorpromazine, 92
BHT (see butylated hydroxy toluene) — as oxidation sensitiser, 92
Butylated hydroxy toluene (BHT), 8, 9 Cobalt cyclohexadienoneoxyl
— by oxidation of BHT in the
Carbonyl compounds presence of cobalt acetyl
— as photosensitisers, 85 acetonate, 126
ß-Carotene — ESR spectra of, 126
— as initiator of oxidation, 88 Cobalt ions
— as quencher of 1 02, 93 — as catalysts for peroxide
— as retarder of oxidation, 13, 93 decomposition, 114
SUBJECT INDEX 227

Copper ions, 16 et seq. 2,6-Dimethyl-hept-2,5-diene


— an antioxidants, 16,17 oxidation, 11, 66
— in antioxidant-prooxidant — termination step in autoxidation,
inversion, 16 11,12, 66
— as catalytic antioxidants, 143,155 2,6-Dimethyl phenol, 60
— deactivators for, 25, 212 et seq. Dioxetane formation
— in decomposition of — with singlet oxygen, 90, 91
hydroperoxides, 16 Dioxygen, 83
— in paint media, 17,155 — initiation of oxidation by, 85
Copper alkyls, 16 Diphenylamine
Critical oxidation-reduction potential, — oxidation to quinones, 128
6, 7,137, et seq. — oxidation to quinoneimines, 128 et
Cross-linking seq.
— alternative breakdown — as catalytic antioxidant, 143
mechanisms, 20,164 Diphenyl isobenzofuran (DPBF)
— decomposition by sulphur — in quantitative measurement of
compounds, 19 ^ 2 , 94
— decomposition products catalysed Diphenyl nitroxyl
by HC1,113 — reaction with alkylperoxyl, 130
— decomposition by Disalicylidene ethylene diamine, 88
dithiocarbamates, 187 2,6-Di-fer£-butyl phenol, 60
— decomposition by Disulphides
dithiophosphates, 195 et seq. — as antioxidants, 172
— of linoleate esters during — as prooxidants, 171
autoxidation, 31,104 Dithiocarbamates, 179 et seq.
Cyclohexene hydroperoxide, 2 — as antioxidants, 22,163
Cyclohexanol Dithiophosphates, 179 et seq., 189 et
— oxidation inhibitors for, 143 seq.
— as antioxidants, 22, 189 et seq.
Desferrioxamine — as antioxidants for lubricating
— as antioxidant, 35 oils, 163
— in treatment of iron overload, 35 Dithiophosphoric acids, 190
Dialkyl sulphides — as peroxidolytic antioxidants, 190
— as antioxidants, 165 — kinetics of C HP decompositions
— reaction with hydroperoxides, 165 by, 191
Diarylamines Dithiophosphoryl disulphides
— as mechanoantioxidants — kinetics of CHP decomposition by,
(antifatigue agents), 145 et seq. 190
Diaryl nitroxyls — prooxidant stage, 190
— as antifatigue agents, 145 et seq. — oxidation products of
— as antioxidants, 15 dithiophosphates, 196
Differential thermal analysis (DTA), Drying oils, 31
73
— in study of polymer stabilisation,
75 Electron attracting groups in phenols,
Dilauryl thiodipropionate, 166 et seq. 9
Dimerisation Electron donors
— of aryloxyls, 9 — as antioxidants, 9
228 SUBJECT INDEX

Electron releasing groups in phenols, — self-induced decomposition, 107 et


9 seq.
Electron withdrawing groups in — as sensitisers for autoxidation, 27
hydrogen abstraction, 62 — solvent induced decomposition, 51,
Endoperoxide formation 109
— with singlet oxygen, 90 — thermal decomposition rates, 110
— in semiquantitative single oxygen — unimolecular decomposition, 50
measurement, 94 Hydrogen abstraction
"Ene" reaction — polar effects in, 60
— with singlet oxygen, 90 Hydrogen chloride
Epoxides — redox reaction with
— by oxidation of di-isobutene, 107 hydroperoxides, 112
— as products of autoxidation, 64, Hydrogen peroxide
67,107 — as an initiator of oxidation in
— from styrene and a-methyl biological systems, 34
styrene, 68 — by reduction of Superoxide, 84
Ethyl linoleate — in tendering of cellulose fibres, 84
— as catalytic antioxidants, 146 Hydroxylamines
— termination in oxidation of, 11 — as CB-D antioxidants, 150
— in the catalytic antioxidant
Homolytic products mechanism, 147 et seq.
— in cumyl hydroperoxide 2-Hydroxybenzophenones
decomposition, 187 — as antioxidants, 220
— in presence of dithiolates, 188 — as UV absorbers, 220
— in presence of nickel 2-Hydroxybenzotriazoles
dithiophosphates, 202 — as UV absorbers, 220
Hydroperoxides, 2, 3, 5, 6,102 Hydroxyl radical, 35
— acid catalysed decomposition, 51
— as autocatalysts for autoxidation, Infrared spectroscopy
48,102 — in study of oxidation, 76
— bimolecular decomposition, 50 Induction period, 1, 132 et seq.
— decomposition, 161 Inductive effects
— formation by oxidation, 102 — in antioxidant activity, 136
— formation by ozone, 96, 98 Inhibition of oxidation
— macroalkyl radical induced — kinetics, 131
decomposition, 109 Inhibitors (see Vitamin E)
— metal ion induced decomposition, Initiation
51 et seq., I l l — of autoxidation, 46, 161,162
— oxidative decomposition of, 114 — by ground state oxygen, 85
— photolysis, 110 Iron overload, 35
— radical induced decomposition, 51, Iron dithiocarbamates
108,199 — as catalytic antioxidants, 184
— reaction of cobalt ions with, 115 — as catalysts for cumyl
— reduction by glutathione hydroperoxide decomposition, 188
peroxidase, 37 — kinetics of hydroperoxide
— redox reactions of, 111 et seq. decomposition by, 184
— reducing agents for, 111 — oxidation to thiuram disulphides,
— role in autoxidation, 48 et seq. 185
SUBJECT INDEX 229

— as thermal antioxidants, 184 — mechanism, 182


Iron dithiophosphates — as radical scavengers, 182
— mechanism of antioxidant action, Metal dithiolates
207 et seq. — as thermal and UV stabilisers, 181
— as peroxide decomposers, 207, 208 Metal ions, 25
— as radical traps, 207 — complexing agents for, 25, 211 et
— sulphur acids from, 207 seq.
Isoozonides, 99 — deactivators for, 25, 211 et seq.
Isotope effect — as inhibitors of oxidation, 55
— in reaction of phenolic — as initiators for autoxidation, 55,
antioxidants with peroxyl, 138 211 et seq.
— redox reactions with
Lactoferrin, 35 hydroperoxides, 16, 25
— stable iron complex, 35 Metal oxime complexes
Lipids — as antioxidants, 217
— in iron overload, 35 Methyl styrene
— oxidation, 1 — oxidation of, 68
— protection by tocopherol, 34
Linoleate esters
— oxidation 31,103 Natural rubber oxidation, 1
— formation of isomeric a-Naphthylamine
hydroperoxides, 103,104 — as catalytic antioxidant for
Linseed oil oxidation, 5 cyclohexanol oxidation, 142
Nickel dithiocarbamates
Macroalkyl radicals — as catalytic antioxidants, 184
— induced decomposition of — in catalytic decomposition of
hydroperoxides by, 109 cumyl hydroperoxide, 188
— during styrene oxidation — ionic reactions of, 183
— trapping agents for, 142 — kinetics of hydroperoxide
Mastication of rubber, 27 decomposition by, 184
Mechanoantioxidants — mechanism of antioxidant action,
— in rubber, 145,146 183
Mechanoxidation, 27 — oxidation to thiuram disulphides,
Mercaptobenzothiazolates 185
— mechanism of antioxidant action, — radical reactions of, 183
174 et seq. — as thermal antioxidants, 184
— oxidation by hydroperoxides, 175 Nickel dithiophosphates
— as peroxide decomposers, 174 — ionic reactions of, 183
Mercaptothiazolines — mechanisms of antioxidant action
— formation of SO2 from, 178 183, 202 et seq.
— kinetics of oxidation, 182 — oxidation products from, 202 et
— oxidative transformation products, seq.
177 et seq. — reactions with hydroperoxides,
— as UV stabilisers, 177 202 et seq.
Mercaptobenzothiazole Nickel oximes
— an antioxidant, 4 — mechanism of antioxidant action,
Metal deactivators 219
— as hydroperoxide decomposers, 182 Nickel thiolates
230 SUBJECT INDEX

— hydroperoxide decomposition by, — effect of substrate structure, 11


184 — technological effects of, 68
— ionic reactions of, 183 — of unsaturated fatty acids, 30,103
— mechanism of antioxidant action, et seq.
183 Oxidative cross-linking, 31
— as peroxide decomposers, 183 Oxides of nitrogen
— radical reactions of, 183 — as initiators for photooxidation,
Nickel xanthates 101
— ionic reactions of, 183 Oxidising agents
— mechanism of antioxidant action, — as CB-A antioxidants, 14
183 Oxygen
— radical reactions of, 183 — absorption, 2, 69
Nitroalkanes — absorption techniques, 69
— as photoantioxidants, 101 — addition to double bonds, 103
— photolysis of, 101 — carriers as prooxidants, 88
Nitroxyl radicals — concentration effect on antioxidant
— as alkyl radical traps, 15 activity, 137
— an antioxidants, 15 — concentration effect in
— aromatic, reduction to termination, 11 et seq.
hydroxylamines, 130 — diffusion in autoxidation, 121
— from aromatic amines, 128 et seq. — molecular as initiation of
— as catalytic antioxidants, 17,130, oxidation, 47
144 et seq. — pressure in autoxidation, 122
— chemistry of, 147 — reaction with antioxidants, 137
— disproportionate of, 149 — reaction with aryloxyl, 7
— as light stabilisers, 147 — reaction with reducing substrates,
— as mechanoantioxidants, 146 et 6
seq. Ozone, 84
— as melt stabilisers, 147 — as initiators of oxidation, 96
— reaction with alkyl radicals, 130 — as prooxidant, 84
— regeneration in hydrocarbon — reaction at localised double bonds,
oxidation, 149 98 et seq.
— structure of, 148 — reaction with olefins, 936 et seq.
Norrish II process — reaction with polystyrene, 99
— reaction at tetrahedral carbon, 96
Olefins Ozone cracking of rubber, 25
— oxidation of 2,3 Ozonides, 99
Oxidation, 1 et seq.
— of arylamines, 7, 122 et seq. Paint, 2
— ofBHT, 124 Parkinsons disease, 32
— biological, 30 PD-C, catalytic peroxide decomposing
— in cellular metabolism, 22 antioxidants, 164
— of di-isobutene, 107 PD-S, stoichiometric peroxide
— of linoleate esters, 30 decomposers, 208
— of oleate esters, 31 Peracids
— of phenols, 7, 122 et seq. — reaction with alkyl
— of polymers, 68 hydroxylamines, 154
— of styrene, 104 et seq. Perishing of rubber, 1
SUBJECT INDEX 231

Peroxidases, 35 Photolysis
Peroxides, 2, 5,102 et seq. — of hydroperoxides, 84
— from phenols, 7 Photooxidation, 29
— in oxidation, 34 — of cyclohexene, 46
— in reaction of ozone with Photosensitisers
polystyrene, 98 — for oxidation of polybutadiene, 91
— effect of structure on stability, 106 Polar effects
Peroxide decomposers, 36 — in hydrogen abstraction, 62
— catalytic, 36 "Poisoning", 1
— stoichiometric, 208, 218 Polymers
Peroxidolytic antioxidants, 19 et seq., — oxidation of, 47
36 et seq., 161 et seq. Polyperoxides, 3, 57
Peroxyl radicals — in ozonation of rubber, 99
— reaction with alkyl — effect of structure on stability, 106
hydroxylamines, 154 Polybutadiene
— reaction with double bonds, 57 — oxidation, 87
— secondary alkyl, 63, 67 — photosensitised oxidation, 91
— steric effects in reactions of, 60 Polyisoprene
— effects of structure on reactivity, — oxidation with singlet oxygen, 91
62 et seq. Polypropylene hydroperoxide
— relative reactivity in hydrogen — decomposition, 50
abstraction, 59 et seq. Preventive antioxidants, 35
— tertiary alkyl, 63, 67 Pyridine
Phenols — as acid scavenger for sulphur
— as antioxidants, 4, 6 acids, 186
— hindered, 7 et seq.
— oxidation by alkylperoxy, 6 Quinones
Phenolic antioxidants — as antioxidants, 141
— activity and structure, 9,133 — products of phenol oxidation, 123
— alkyl branching in, 139 et seq.
Phenolic nitrones
— nitroxyl formation by hydrogen
abstraction, 128 Radical chain reaction, 3
Phenoxyl radicals, 6,126 et seq. Radical induced decomposition of
— as catalytic antioxidants, 17 hydroperoxides, 108
— stability, 127 et seq. — inhibited by antioxidants, 109
Phenylhydrazine Redox reactions
— as prooxidant for rubber, 87 — in rubber oxidation, 87
Phenyl-ß-naphthylamine Rubber
— oxidation by phenoxyl and alkoxyl — heat resistant, 5
radicals, 128 — perishing of, 1
Phosphite esters, 208 et seq.
— as peoxidolytic antioxidants, 37, Singlet oxygen, 26, 84
163, 208 et seq. — in biological oxidation, 91
Phosphoryl disulphides — formation by energy transfer, 90,
— oxidation products of 92
dithiophosphates, 196 — as initiator of autoxidation, 26, 29,
Photoantioxidants, 38 84,90
232 SUBJECT INDEX

— quenching by carotenoid pigments, Sulphur trioxide as antioxidant, 186


93 Superoxide, 84 et seq.
— reactions with unsaturated — in biological systems, 34
compounds, 90 — formation by energy transfer from
— relative rates of quenching by Ti0 2 , 85, 92
environmental molecules, 95 — reduction to hydrogen peroxide,
— relative rates of quenching by 85,92
olefins, 95 Superoxide dismutase (SOD)
Sinovial fluid — as preventive antioxidant, 35
— oxidation in the presence of iron, — as synergist with catalase and
35 peroxidase, 34
Sodium sulphite Synergism, 23 et seq.
— oxidation, 6 — between tocopherols and ascorbic
Stability of phenoxyl radicals, 127 acid, 33, 34
Steric effects — between tocopherols and selenium
— in chelation by metal ions, 213 and glutathione, 34
— in phenoxyl radicals, 9,139 et seq.
Stilbenequinone Tendering of cellulosic fibres, 84
— oxidation products of BHT, 124 Termination in oxidation, 3,11,12,
Stoichiometric coefficients of 46,66
inhibition, 142 — effect of oxygen pressure, 11, 66,
Sulphenic acids 67
— as antioxidants, 168 — through alkylperoxyl, 67
— as peroxide decomposers, 168 Tetralylperoxyl
Sulphides — reactivity in autoxidation, 63
— an antioxidants, 19 Tetramethyl piperidinoxyl
Sulphoxides — as antioxidant, 15, 141 et seq.
— as antioxidants, 21,164,165 Tetramethyl-p-phenylenediamine, 138
— as peroxide decomposers, 165 Tetroxides, 67
— thermolysis, 170,173 Thermogravimetry, 72
Sulphur Thiophenol as antioxidant, 5
— compounds as antioxidants, 4,19, Thiophosphoric acids
37 — oxidation products of
— compounds as catalytic peroxide dithiophosphates, 190
decomposers, 164 et seq. Thiophosphoryl dilsulphides
— cross-links in rubber, 165 — as oxidation products of
Sulphur acids dithiophosphates, 190
— an antioxidants, 186 — as peroxide decomposers, 190,192
Sulphur antioxidants — as products of disulphide
— autoinhibition by, 21 oxidation, 196
Sulphur dioxide Thiosulphinates
— as an antioxidant, 170 — as antioxidants, 21,168
— as a peroxide decomposer, 22,168 — as peroxidase decomposers, 168
— as a photosensitiser, 27,100 Thiodipropionate esters, 163
— prooxidant effect, 171 — antioxidants for foodstuffs, 163
Sulphuric acid — oxidation by hydroperoxides, 112
— as an antioxidant, 5,168 — prooxidant activity, 110
— as a peroxide decomposer, 168 Titanium dioxide (anatase)
SUBJECT INDEX 233

— as sensitiser for photooxidation, 92 Vulcanisation of rubber, 1, 4


Tocopherols, 19, 32 et seq. Verdigris
— as biological antioxidant, 34 — as photoantioxidant in paint films,
— phenoxyl from, 19, 33 155
— regenerated from the phenoxyl, 33
— as Synergist with ascorbic acid, 32 Würster ion-radical, 138
et seq.
Transition metal complexes Xanthates
— as antioxidants, 174 — as antioxidants, 175 et seq.
— as UV stabilisers, 174
Transition metal ions Zwitterion formation
— as prooxidants, 115 — in ozone cracking, 25
Transferrin, 35 Zinc dithiocarbamates
Transition state — as catalytic antioxidants, 184 et
— in antioxidant action, 9 seq.
— in autoxidation, 62 — kinetics of peroxide decomposition
— in decomposition of sulphoxides, by, 184
167 — as thermal antioxidants, 184
Triphenylmethyl Zinc dithiocarbamates
— as retarder of autoxidation, 13 — as catalytic antioxidants, 184
Triplet carbonyl — kinetics of hydroperoxide
— in photoinitiation, 30, 85 decomposition by, 184
Triplet quenchers, 30 — as thermal antioxidants, 184
Zinc thiopercarbamates
UV light — as oxidation products of zinc
— deactivators for, 218 dithiocarbamates, 186
— in initiation of autoxidation, 36 Zinc dithiophosphates
UV absorbers, 218 — mechanism of antioxidant action,
195 et seq.
Vitamin A — oxidation products, 198
— as prooxidant for methyl linoleate, — reaction with hydroperoxides, 195
88 et seq.
Vitamin E (inhibitors, tocopherols) Zinc thiophosphates
— as biological antioxidant, 32 — oxidation products of zinc
Vitamin C (ascorbic acid), 32 dithiophosphates, 196

You might also like