You are on page 1of 55

Encyclopedia of Polymer Sceince and Technology Copyright c 2005 John Wiley & Sons, Inc. All rights reserved.

RADIATION CHEMISTRY OF POLYMERS


Introduction
The fundamental property of polymeric materials which sets them apart from other materials is their very large molecular weight. The arrangement of a large number of structurally similar chemical units into linear or branched chains provides polymers with their unique viscoelastic properties. This large molecular weight necessarily leads to both transient and permanent molecular entanglements, which determine the response of the polymer to applied loads. It is essentially this property which is altered, or deliberately manipulated, when a polymer is subjected to ionizing radiation. High energy radiation leads to a succession of chemical reactions which ultimately lead to either an increase or a decrease in the molecular weight of a polymer. As will be discussed below, either result may be desirable. The initial processes which occur when a high energy photon interacts with an organic polymer are reasonably well established and do not depend essentially on the chemical structure of the material. However, these primary processes lead to a cascade of further reactions, the nature of which depends sensitively on the nature of the polymeric material; for example, the ability to dissipate excess energy, the relative bond energies within various structures, and the presence of oxygen, radical scavengers, and other moderating factors. This article approaches the eld of radiation chemistry of polymers from a mechanistic viewpoint, and does not describe in detail the growing list of applications of high energy radiation in polymer science. The basic principles of radiation chemistry are described, and subsequently the interaction of radiation with specic classes of polymers. The mechanisms of reaction are described by reference to specic materials. Furthermore, the means by which these mechanisms have been determined, ie, the methods of analysis, are described in detail. Finally, an attempt is made to draw overall conclusions about the radiation sensitivity of the constituent units of synthetic polymers. Interaction of High Energy Radiation with Organic Matter. The passage of a high energy photon or particle through matter leads to a complex cascade of processes which results in the dissipation of the initial energy, ultimately as thermal energy and chemical reactions (15). If the source of radiation consists of high energy photons, the energy is deposited in a highly nonuniform manner via a number of processes. In organic polymers the most important mechanism of energy loss is Compton scattering, in which a photon interacts with an electron resulting in ejection of the electron and deection of the photon with reduced
1

RADIATION CHEMISTRY OF POLYMERS

energy. The probability of the photon undergoing Compton scattering and the resultant energy of the ejected electron and scattered photon depend on the incident energy of the photon and the electron density of the material through which it is traversing. Of less importance for high energy photons is the photoelectron effect, in which the entire energy of the photon is transferred to a single electron, which is ejected with energy equal to the incident photon energy minus the binding energy of the electron. Also of less importance for most commercial sources of ionization radiation is the mechanism of pair production. This also involves complete absorption of the incident photon energy, and results in the formation of an electronpositron pair. The combination of the two particles results in emission of two 0.51-MeV -rays, which may undergo further interaction with the material as described above. Pair production is only of importance when the incident photons have energies greater than 1.02 MeV. If on the other hand the source of ionization is high energy electrons, the initial steps toward loss of the energy to the material involves inelastic collisions resulting in ionization and excitation of the material and loss of energy of the incident particle. On ejection of the secondary electrons, the resultant processes occurring are identical to those occurring on irradiation with high energy photons. These can be summarized as follows. Initial interaction with photons: RR + + e Interaction of primary electrons with other structures: R + e R+ + 2 e (2) (1)

As the energy of the electrons is reduced, there is increased probability of recombination of the cations and secondary electrons to form excited states: R + + e R (3)

The excited-state molecules may return to the ground state through radiationless decay, or undergo homolytic dissociation reactions to form free radicals, which are believed to be the main protagonists in subsequent radiochemical reactions. Heterolytic bond cleavage may in addition result in the formation of charged species. Linear Energy Transfer and Recombination Reactions. The vast majority of studies on the radiation stability of polymeric systems have been concerned with either incident photons or electrons. This clearly reects the greater availability of sources or electron accelerators. However, there has been a smaller but increasing interest in the effects of heavier charged particles on polymers, either because of interest in ion implantation or concern for the effects of heavy ions on materials in, for example, the space environment, or indeed processing of polymers using ion beams. Of most interest is the effect of the increased linear energy transfer (LET) of heavier charged particles on the mechanism of the

RADIATION CHEMISTRY OF POLYMERS

radiation chemistry. The results presented so far are conicting. For example, While Parkinson and co-workers (6) and Boyett and Becker (7) did report differences between neutron and irradiation, others have indicated identical changes in properties (8). The effect of proton irradiation was found to be identical to electron irradiation for aliphatic polymers; however, some differences were observed for aromatic materials (9,10). However, Hill and Hopewell (11) found no difference between the mechanical properties of polyimides irradiated with rays or photons. A survey of several studies of the effects of ion beam irradiation on various polymers (1217) again does not present a clear picture. This is clearly an area for further investigation. Energy Transfer and Dissipation. From early times it was recognized that the presence of small concentrations of certain structures in a material can lead to a pronounced decrease in the radiation sensitivity of the material as a whole. The earliest observations of this type were of the evolution of gases from mixtures of small aromatic and aliphatic molecules. For example, Manion and Burton (18) reported in 1952 that the yields of reactions in irradiated mixtures of cyclohexane and benzene were reduced signicantly compared to the yield expected from the linear addition of the yields of the pure individual molecules. This led to the concept of protection effect, in which the introduction of aromatic groups results in signicant protection of the surrounding aliphatic structures. The mechanism of protection afforded by the aromatic molecules has been a subject of considerable discussion. Since these early studies of mixtures of aromatic and aliphatic liquids (18,19), the phenomenon of radiation protection has been observed in copolymers of styrene with other monomers (2025), in a range of alkylbenzenes (26), and in miscible polymer blends (27). The description of the mechanism of radiation protection has often in the past relied upon the assumption that the amount of energy deposited in each molecule or molecular segment within a polymer is proportional to its electron fraction. As a consequence the decrease in radiation sensitivity of one component must be due to either transfer of energy or charge prior to dissociation, or scavenging of intermediates in systems where signicant mobility is present. It has been reported, however, that the assumption that the cross section is proportional to electron density is not at all times obeyed (28). Therefore the assignment of protection effects wholly to transfer processes must be made with caution, although the relative stabililizing effect of aromatic structures is not in question. Yields and Units. The unit of radiation dose used in this text is the Gray (Gy); 1 Gy corresponds to an energy absorption of 1 J/kg of material. Typical applied doses in the eld of radiation modication of polymers are in the range of tens of kGy; however, in some cases irradiation up to 10 MGy has been required to induce sufciently large yields of new molecular structures. The yield of radiochemical events is expressed as the G value [written, for example, G(X) for the G value of cross-linking]. The G-value is dened as the number of products formed for every 100 eV, or 16 aJ, of energy absorbed by the material. Thus, 1 kGy of radiation will produce G 1.036 10 7 moles of product per gram of irradiated polymer (4). The SI unit for radiation yield is mol/J, which is equivalent to 10 Gy (29).

RADIATION CHEMISTRY OF POLYMERS

Summary of Radiation-Induced Reactions in Polymers


The preceding section has described briey the primary reactions occurring on interaction of high energy radiation with organic molecules in general. Following the succession of reactions detailed above, a number of subsequent combination and bond-breaking reactions lead to permanent changes in the properties of the material. Much experimental evidence exists to support the contention that most of these reactions are initiated by free radicals. The observation, by electron spin resonance (ESR) spectroscopy, of relatively large concentrations of radicals in polymers irradiated at low temperatures is one example of this. In the vast majority of experimental studies of the radiation chemistry of polymers, we observe usually the nal products of the passage of the high energy photons. Attempts to trap radical and charged species at low temperatures usually succeed in revealing part of the radiochemical history of the material. An indication of this can be obtained by consideration of the relative yields of products on radiolysis of simple alkane molecules. The respective bond energies of C H and C C bonds in linear alkanes are 4.3 and 3.4 eV (415 and 328 kJ/mol) (1). On consideration of these numbers alone, one would expect that the predominant products of radiolysis of linear alkanes would be due to rupture of C C bonds. However, the main product of the radiolysis of linear alkanes, and indeed of polyethylene, is hydrogen gas. Furthermore, there is some debate as to whether permanent cleavage of the main-chain C C bond actually occurs at all on irradiation of polyethylene. The origin of these results is of course the cage effect. The larger radical fragments formed on cleavage of C C bonds are unlikely to escape from the reaction cage, and recombination reactions tend to occur. The smaller, perhaps thermally excited, hydrogen radicals formed by cleavage of C H bonds have a much greater probability of escaping and undergoing recombination and other reactions. This simple and important observation also helps us understand the differences in the radiation chemistry of hydrocarbon polymers and peruorocarbon polymers, as discussed below. In addition the unusual temperature dependence of the radiochemical yields of a number of materials can be explained by the cage effect. For example, the radiochemical G-values for formation of stable free radicals in -irradiated uorinated ethylenepropylene copolymers (FEP) are 0.22 at 77 K, and 2.0 at 298 K (30), since at low temperatures cage recombination of the primary radicals is very effective, while at higher temperatures sufcient thermal energy is available to remove the radical fragments from the reactive cage. Additional evidence of the very large number of nonpermanent bond cleavages initiated by ionizing radiation is the observation of large changes in the stereochemistry of initially highly stereoregular polymers, such as isotactic poly(methyl methacrylate) (iPMMA) and related polymers (3134), as well as isotactic polypropylene (iPP) (3538). The loss of stereoregularity arises from a racemization reaction occurring as a result of temporary scission of either the main chain or the side chain. The latter mechanism is favored for syndiotactic PMMA. The yield of temporary chain scissions is approximately 10 times greater than the yield of permanent chain scissions determined by measurement of gel fraction or molecular weight of the soluble polymers.

RADIATION CHEMISTRY OF POLYMERS

Cross-Linking and Chain Scission. The most important reactions occurring during radiolysis of polymers are those that lead to permanent changes in their molecular weight. The reactions leading to either increases or decreases in molecular weight are referred to as cross-linking and chain scission, respectively. End-linking, ie, reaction between a polymer main-chain species and a polymer chain end, is regarded as a cross-linking reaction of lower functionality. In general cross-linking and scission processes can occur simultaneously in any irradiated material; however, it is often observed that one tends to dominate over the other, and thus polymers can be broadly placed into the categories cross-linking or degrading. Over the past 50 years a solid understanding of the relationship between polymer structure and the relative yields of cross-linking and chain scission has been acquired. Thus, for linear polyolens, the proportion of chain scission reactions increases as the number of tertiary and especially quaternary carbons along the backbone is increased. Polyethylene undergoes primarily cross-linking reactions, polypropylene undergoes approximately equal amounts of scission and cross-linking, which ultimately leads to an increase in molecular weight, and nally polyisobutylene undergoes chain-scission reactions exclusively. However, it must be clearly stated here that respective and absolute yields of the two types of reactions depend also on a number of extrinsic factors, such as the presence of a crystalline phase, the temperature, the presence of chain-end unsaturation, and perhaps most profoundly the presence of oxygen gas. In Table 1 below, we attempt to classify the various polymer types as either cross-linking or degrading on exposure to ionizing radiation in the absence of oxygen. Specic examples are discussed later in this article. Evolution of Gases. When polymers are exposed to high energy radiation, the reactions induced will lead to the formation of low molecular weight gaseous molecules. A study of the structure and yields of the gaseous products

Table 1. Cross-Linking vs Chain Scission for the Polymer Classes on Exposure to Ionizing Radiation in Vacuum Cross-linking dominates Polyethylene Polypropylene Polydienes Polystyrene Polyacrylates Polyacrylamide Fully uorinated elastomers Poly(vinyl chloride) Polyamides Polyesters Polysiloxanes Poly(vinyl alcohol) Poly(N-vinylpyrollidinone) Main-chain aromatic polymers Chain scission dominates Polyisobutylene

Poly(-methyl styrene) Polymethacrylates Polymethacrylamide Fully uorinated thermoplastics Poly(vinylidene chloride) Polychlorotriuoroethylene Cellulosics

RADIATION CHEMISTRY OF POLYMERS

can provide great insight into the mechanisms of the reactions occurring on radiolysis. In general, gaseous products can arise either from primary cleavage of the main-chain bonds or from cleavage at the side chains. Polymers that fall into the class of degrading polymers can, on irradiation at elevated temperatures (more specically at temperatures approaching their ceiling temperature), yield high concentrations of monomers among the gaseous products. The ceiling temperature is the temperature at which the reversible addition of monomers to a radical chain end is balanced by the tendency of the monomers to depolymerize. Thus, it was noted as early as 1953 (39) that poly(methyl methacrylate) foams on irradiation due to the formation of gaseous products, including methyl methacrylate monomer. Subsequently, it was reported that the yield of methyl methacrylate monomer increased substantially to a limiting yield around 180 C (40,41). Related to this is the pronounced tendency of these materials to undergo depolymerization after exposure to high energy radiation. The most important commercial exploitation of this phenomenon has been the development of positive radiation resists for use in the semiconductor industry. A number of methacrylates and butene sulfones have been developed for these applications (4244). Another important class of materials which undergo largescale degradation of the main chain are the aliphatic polysulfones. For example, in 1981 Bowmer and ODonnell (45) examined the yields of a number of aliphatic polysulfones as a function of temperature and discussed these results in terms of the change in the equilibrium between polymerization and depolymerization as the ceiling temperature is approached. This aspect of radiation chemistry has had far-reaching consequences for our modern society. The second major mechanism leading to the formation of volatile products on irradiation of polymers is cleavage of the side-chain groups. In the case of linear polymers, such as polyethylene, scission of the C H bond leads to the formation of thermally activated hydrogen atoms. As discussed above, these small entities are able to escape the reactive cage and undergo subsequent reaction. These include addition reactions to double bonds, and also abstraction of hydrogen atoms and recombination with hydrogen atoms to form molecular hydrogen gas. Thus, the main gaseous product formed on radiolysis of hydrocarbons is hydrogen gas. The yield of hydrogen gas in polymers of the cross-linking type is often in excess of that expected from consideration of the yield of alkyl radicals leading to the formation of cross-links. As early as 1954, Dole and co-workers (46) determined that up to 80% of the hydrogen gas measured after irradiation of polyethylene arose from the formation of vinylene unsaturation, with cross-linking reactions accounting for the remainder. The presence of hydrogen gas has an important inuence on the kinetics of radical reactions, and in principle on the overall product distribution in irradiated polymers. It has been suggested (47,48) that hydrogen gas is capable of enhancing the rate of diffusion of free radical species via reaction with alkyl radicals as depicted below: R + H H + H R R H H H R R H + H H + R (4)

The importance of the presence of hydrogen gas was rst demonstrated by Dole and Cracco in 1961 (4951) by measuring the rate of hydrogen exchange from irradiated high density polyethylene to gaseous D2 , and explored more completely

RADIATION CHEMISTRY OF POLYMERS

by Dole and co-workers over the next decade (47). Clough (48) much later provided a detailed analysis of the mechanism of migration of radicals through the solid state in aliphatic hydrocarbon polymers, including an analysis of the number of these hydrogen-hopping reactions expected on the production of main-chain radicals. The calculations accounted for localized hydrogendeuterium exchange reactions in linear alkanes irradiated in the presence of D2 gas. It should also be noted that in aromatic polymers, hydrogen atom addition to aromatic rings results in the formation of cyclohexadienyl radicals and, potentially, initiation of cross-linking reactions (22,24,5254). The great importance of the hydrogen-hopping reaction can be appreciated by considering the differences in the radiation chemistry of polyhydrocarbons and polyuorocarbons. In the latter materials, reaction 4 is of course not operative, and the corresponding reaction involving uorine radicals is believed to be unimportant after consideration of the respective C F and F F bond energies (approximately 510 and 155 kJ/mol, respectively). Thus, the reactivity of uorocarbon radicals in irradiated uoropolymers is much more dependent on the inherent mobility of the polymer chains than in hydrogenated polymers. Thus, cross-linking reactions, which require two alkyl radicals to meet each other, are unlikely to occur in uoropolymers irradiated at normal temperatures. Thus, polytetrauoroethylene (PTFE) undergoes chain scission at all temperatures below the melting temperature (55,56). On the other hand the uoroelastomer Kalrez (DuPont), a copolymer of tetrauoroethylene and peruorinated methyl vinyl ether, can be readily cross-linked by exposure to radiation, because of the mobility of the polymer chains and related higher probability of radical recombination reactions (57). A corollary to this is the marked long-term stability of free radicals in irradiated uoropolymers. Indeed, Judeikis and co-workers (58) and others have proposed the use of PTFE as a radiation dosimeter. Of more practical concern is the propensity of these stable radicals to react slowly with dissolved oxygen, a phenomenon leading to the well-known poor radiation stability of uoropolymers. Finally, examination of the volatile products of irradiation of branched polymers reveals the importance of loss of the side chains. For example, irradiation of branched copolymers of ethylene and -olens results in volatile products characteristic of both the frequency and structure of the side chain (59,60). In this work the authors also highlight the inuence of radiation temperature and morphology of the irradiated semicrystalline polymer. Formation of Unsaturated Groups. As indicated above, an important result of the radiolysis of polymer is the formation of unsaturated groups. In 1954, Dole and co-workers (46) reported the formation of main-chain unsaturation in irradiated polyethylene and demonstrated that the yield of these groups was proportional to the yield of hydrogen gas. In these materials the mechanism of formation of unsaturation was thus established at an early time. In this and a subsequent article (46,61), Dole and co-workers also indicated that other types of unsaturated groups, for example vinyl and vinylidene groups, are rather consumed in the radiation-induced reactions. Later, this same group deduced from measurements of gel fractions that the terminal unsaturation was reacting via an end-linking reaction, a result conrmed much later by Randall and co-workers (6264) and Horii and co-workers (65,66).

RADIATION CHEMISTRY OF POLYMERS

Perhaps the most well-known system in which large amounts of unsaturated groups are formed is poly(vinyl chloride). It is well established that irradiation leads to the formation of alkyl radicals, which are the locus for subsequent dehydrochlorination reactions. This results in characteristic changes in the infrared and visible absorption spectra (6769), as well as the formation of stable polyenyl radicals (67,68). Color Centers. An obvious corollary of the formation of polyenyl radicals and conjugated double bonds in irradiated poly(vinyl chloride) is a marked changed in the visible appearance of the polymer (6770). Color changes on irradiation of polymers may have important implications for a number of commercial applications, for example when the appearance of the material is of importance, as in medical applications, or when the optical properties of the material must be maintained in a high dose environment, for example in polymeric ber optics in particle physics detectors. On the other hand, color development for radiation dosimeters has been exploited by a number of workers, most notably the group of McLaughlin (7174). Until recently the phenomenon of color changes in irradiated polymers has not received detailed attention. Clough and co-workers (7577), however, have recently reviewed studies of color changes in a number of important synthetic polymers. They report that there are two classes of color centers in polymers, annealable and permanent. The former class of color center appears to be associated with free radicals trapped within the polymer matrix, whereas permanent color centers are stable chromophores formed as a result of radiochemical reactions. The free radicals can be destroyed by the application of heat, or often through reaction with atmospheric oxygen. Clough and coworkers (77) also reported that whereas the incorporation of aromatic groups provides a degree of protection against radiochemical damage to polymers, it does not necessarily protect against color formation, because presumably the products of reaction in aromatic polymers tend to have high extinction coefcients.

Changes in Molecular Weight of Polymers on Exposure to Radiation


The effects of cross-linking and main-chain scission on the molecular weight and the molecular weight distribution of polymers have been examined in great detail by a number of authors (7889). The exact method of treatment of the problems depends on the molecular weight distribution of the polymer prior to irradiation, and on whether it can be described by a most-probable distribution or recourse must be made to a higher-order distribution. For simplicity, in this article we shall consider the effects of cross-linking and scission on polymer having an initial mostprobable distribution; however, reference will be made to more complete analyses where appropriate. Effects of Chain Scission. In the early 1950s, Charlesby and co-workers were considering theoretical approaches to the description of the changes in molecular weight of polymers during irradiation. In 1954 they reported that the molecular weight of PMMA, measured by solution viscometry, was inversely proportional to the radiation dose (80,90). Since that time a number of

RADIATION CHEMISTRY OF POLYMERS

renements to these initial observation and theoretical expression have been made (81,9194). In summary, the number- and weight-average molecular weights change as described below: 1 1 = + 2k1 G(S)D Mn (D) Mn (0) 1 1 = + k1 G(S)D Mw (D) Mw (0) (5)

(6)

In these expressions the constant k1 is equal to 5.18 10 8 if the dose D is expressed in units of kGy. Most importantly, equations 5 and 6 are applicable to any initial molecular weight distribution, because in the case of degrading polymers, the molecular weight distribution will approach the most-probable distribution. Effects of Simultaneous Cross-Linking. Charlesby (78) was also the rst to describe the relationship between the molecular weight of a polymer and the radiation-induced chain cross-linking. He reported in 1954 that the dose required to produce incipient gel formation, known as the gel dose Dg , was inversely proportional to the initial molecular weight of the irradiated siloxane. 1 1 = + 2k1 [G(S) G(X)]D Mn (D) Mn (0) 1 1 = + k1 [G(S) 4G(X)]D Mw (D) Mw (0) (7)

(8)

In this case, equation 7 applies to all initial molecular weight distributions, whereas equation 8 applies only to polymers having an initial most-probable distribution undergoing H-linking reactions. Expressions for changes in molecular weight in polymers having other initial molecular weight distributions, and also for the case of cross-linking via a Y-linking mechanism, have been developed, and these have been summarized by various authors (8688,93). The above expressions have been derived for the rst two moments of the molecular weight distribution, M n and M w . However, the third moment, the zaverage molecular weight, M z , is experimentally accessible through measurement of sedimentation velocity of polymer chains. ODonnell and co-workers (85,86) have reported the respective relationships between M z and the radiochemical yields for scission and cross-linking. The results of analysis of the changes in M z and M w of irradiated poly(acrylic acid) and poly(methacrylic acid) (85) and polystyrene have also been reported (86). Soluble Fractions. The above descriptions of the change in molecular weight on irradiation presuppose the ability to accurately measure the molecular weight distribution of the polymer after irradiation. A number of authors have pointed out that for polymers for which cross-linking dominates chain scission, measurement of molecular weight is limited to doses well below 50% of the gel

10

RADIATION CHEMISTRY OF POLYMERS

dose. More conveniently measured parameters are therefore the gel dose, that is, the dose to achieve incipient gel formation, and the residual soluble fraction beyond the gel dose. The most widely used expression for analysis of the soluble fraction is also due to Charlesby and Pinner (95). G(S) + k2 /[G(X)Mn (0)D] 2G(X)

S + S0.5 =

(9)

The constant k2 is equal to 4.82 106 if the dose is expressed in kGy. Equation 9 assumes a post-probable initial molecular weight distribution, random scission and cross-linking reactions, and cross-linking by an H-linking mechanism. If crosslinking occurs by a Y-linking mechanism, the relationship between soluble fraction and dose is given by (96) S + S0.5 = 2G(S) + k3 /[G(X)Mn (0)D] G(X) (10)

The constant k3 is equal to 1.93 104 if the dose is expressed in kGy. Clearly, the choice of application of either equation 9 or equation 10 must be directed by additional evidence of the mechanism of cross-linking. Finally, a number of workers have derived expressions for the relationship between the soluble fraction and radiation dose for more general initial molecular weight distributions (81,88, 9193,97).

The Radiation Chemistry of Specic Polymers


The radiation chemistry of polymers has attracted enormous interest since the initial studies by Charlesby, Dole, and Chapiro in the early 1950s. This interest has been a result of the recognition of the tremendous commercial importance of the radiation chemistry of polymers, and of the importance of an understanding of radiation chemistry to other elds. A number of excellent reviews have appeared over the past 45 years, including the seminal texts by Charlesby (1) and Chapiro (2), and the comprehensive review texts edited by Dole (98,99). Since then, a number of broad reviews of the effects of radiation have been published. For example, the Polymer Handbook has previously tabulated lists of the radiochemical yields reported elsewhere (100). The American Chemical Society has published a number of collections of papers presented at radiation chemistry meetings (101,102), and the two journals Radiation Physics and Chemistry and Polymer Degradation and Stability are of much interest. The 11 international meetings on radiation processing, which have been published in Radiation Physics and Chemistry (103 113), and the text by Singh and Silverman (114), provide an excellent overview of the eld. A number of more specic reviews have also been published. For example, Reichmanis has written about the use of radiation chemistry for electronics applications (43,115). Clough (116,117) has written a number of excellent reviews with

RADIATION CHEMISTRY OF POLYMERS

11

an emphasis on commercial applications of radiation chemistry, and the effects of oxygen on radiochemical events. The radiation chemistry of uoropolymers has been reviewed recently by Lyons (118) and Forsythe and Hill (119). Burillo and coworkers (120) have addressed the use of radiation in the polymer recycling industry. Rosiak (121,122) has reviewed the emerging eld of the radiation processing of biomedical polymers. More recently, Pruitt (123) has reviewed the major changes induced in medical polymers on irradiation, and given specic case studies of a number of materials. Hill and Whittaker (124) have detailed the importance of NMR spectroscopy in this eld. Polyethylene. It has been known for many years that polyethylene undergoes cross-linking on exposure to high energy radiation. In 1952 Charlesby (125) reported that exposure of polyethylene to pile radiation resulted in chemical changes consistent with cross-linking reactions. This initial observation prompted extensive work by him and others, most notably Dole, which has laid the foundations for the modern eld of radiation chemistry of materials. The main ndings of these and subsequent workers are summarized below (see Ethylene Polymers). Free-Radical Intermediates. From an early time the participation of free radicals in the radiation chemistry of polymers has been well understood. Charlesby (125) in 1952 invoked a free-radical mechanism of cross-linking of polyethylene, although for other materials the participation of ionic species has also been suggested (126137). The primary radicals observed at 77 K, at which temperature a signicant proportion of the radicals are assumed to be trapped and prevented from further reaction, are the secondary alkyl radicals I (49,138148).

Dole and co-workers have reported yields of alkyl free radicals in polyethylene irradiated at 77 K ranging from 2.7 to 3.7 (141,145,149). Furthermore, Cracco, Arvia, and Dole (49) reported that on warming, alkyl radicals decay by a rst-order process, and they attributed this to reactions between alkyl radicals within isolated spurs. The persistent free radicals on warming to room temperature are the allyl radicals II. The impact of long-term stability of radical species on the stability of polyethylene has been underlined by studies of Jahan and co-workers (150157) of ultrahigh molecular weight polymer used in medical implants. Cross-Linking. In addition to conversion of alkyl radicals to more thermally stable allyl radicals, a signicant proportion can undergo radicalradical recombination reactions (eq. 11).

(11) The product shown on the right of equation 11 is called an H-type crosslink. A number of alternative mechanisms of cross-linking have been proposed,

12

RADIATION CHEMISTRY OF POLYMERS

the most important being the formation of Y-type cross-links through either radical recombination reactions (eq. 12) or end-linking to terminal vinyl groups (eq. 13).

(12)

(13) In the case of polyethylene the initial cleavage of carboncarbon main-chain bonds is believed to be relatively unimportant compared with end-linking reactions. Evidence for this includes the absence of a contribution from chain-end free radicals in the ESR spectra of polyethylene irradiated at low temperatures, and careful analysis of the residual soluble fractions formed on irradiation at higher temperatures. Dole and Katsuura (158) suggested that if main-chain scission does occur, the cage effect is responsible for an overwhelming proportion undergoing recombination reactions. Not long after this, Kang, Saito, and Dole (159) measured the yields of cross-linking and main-chain scission in high density polyethylene at a range of temperatures by analyzing the residual soluble fractions with a CharlesbyPinner equation modied for a Wesslau molecular weight distribution. They found that at each temperature over 308393 K, the ratio of scission to cross-linking reactions was signicant (ca 0.2) and decreased slightly at higher doses, an effect they ascribed to the reactions of vinylene double bonds to form cross-links. Similar results were reported by Dole and co-workers (149,160). In contrast to these articles, Lyons and Fox (161) reported negligible dose-dependence of G(X) and 0.03 as the ratio of the yield of scission to that of cross-linking reactions. The yields of cross-linking reactions in polyethylene have been reported extensively (97,149,159,160,162) and generally fall in the range G(X) = 0.82.5 for samples irradiated at ambient temperatures. The principle reason for this wide range of values of cross-link yields is the variation in crystallinity in the materials studied. It is reasonably well accepted that cross-linking reactions do not occur within the crystalline lamellae of polyethylene; Patel and Keller (163) conrmed this by analyzing the residual molecular weights of the chains within the crystalline regions of irradiated high density polyethylene after the amorphous regions had been removed by exposure to ozone. There was no evidence for crosslinking reactions between the chains previously within the crystalline regions. To further underline this relationship, ODonnell and Whittaker (164) have reported that the value of G(X) decreases progressively with increasing crystalline content for a range of materials having a similar initial molecular weight distribution. In this same article, ODonnell and Whittaker (164) compared the yields of cross-linking and scission, and other radiochemical products, in fully amorphous ethylenepropylene rubber (EPR) with the yields measured by them and reported by other workers for semicrystalline polyethylene and polypropylene. The yields of cross-linking and scission were signicantly lower than those expected from extrapolation of the G-values for the homopolymers to zero crystalline content. In

RADIATION CHEMISTRY OF POLYMERS

13

other words, the yields for the fully amorphous materials were much lower than expected from considering yields within the amorphous regions of the semicrystalline polymers. It was concluded that the radiation-induced reactions must be favored at the crystallineamorphous interface and that this might be a result of the stabilization of free-radical intermediates at the interface. The relative importance of reactions 1113 in the radiation cross-linking of polyethylene has been resolved to a large extent through a number of elegant studies of the structures of polymer irradiated to doses below the gelation dose, using high resolution solution-state 13 C NMR spectroscopy. The initial use of 13 C NMR to study radiochemical reactions was by Bennett and co-workers (165), who reported structural changes in irradiated liquid model compounds of polyethylene, n-hexadecane, and n-eicosane. These authors reported the formation of H-link structures apparently via recombination of two secondary main-chain alkyl radicals. Not long after this, Bovey and co-workers (166) reported the identication of both H- and Y-type links in n-C44 H90 irradiated in the melt using solutionstate 13 C NMR. Irradiation in the crystalline state produced only linear dimers, apparently through the end-linking of molecules at the crystal surfaces. Several years after this, Randall and co-workers (6264) studied irradiated high density polyethylenes using 13 C NMR, and identied a number of new structures, including internal double bonds and Y-type cross-links. The materials under consideration initially contained a high concentration of vinyl end groups, which react readily with secondary alkyl radicals (reaction 12 above) to form Y-links. The authors were unable to identify H-links in their irradiated HDPE samples, and suggested that for these materials irradiated to low doses, the H-linking mechanism is relatively unimportant. More recently, Horii and co-workers (65,66) examined changes in the solution-state 13 C NMR spectra of relatively low molecular weight PE samples irradiated under a range of experimental conditions. The lower viscosity of the solutions as a result of the low molecular weight was manifested in narrow line widths in the NMR spectra, and thus the authors were able to observe resonances in the spectra due to H-type cross-links in all samples. The yields of H-links and Y-links were measured as a function of temperature (66), and it was found that H-links were the dominant product when PE was irradiated in the molten state, whereas Y-links were more commonly formed on irradiation at ambient temperatures. It was suggested that at higher temperatures the probability of reaction of the primary radicals (in reaction 12 above) with hydrogen radicals is higher in the molten state, and thus recombination of the longer-lived secondary radicals (reaction 11) dominates. These conclusions are supported by the study of ODonnell and Whittaker (167) in which the formation of H-type cross-links in fully saturated ethylenepropylene rubber (EPR) irradiated well above T g was conrmed. In addition, ODonnell and Whittaker (168) have reported the observation of resonances due to H-type cross-links in LLDPE irradiated to very high doses. Volatile Products. As discussed above, Dole and co-workers were the rst to observe that hydrogen gas was a major product formed on the irradiation of polyethylene (46). Gaseous hydrogen was formed by recombination of hydrogen atoms formed by cleavage of C H bonds, as illustrated in reaction 4 above, or by abstraction of hydrogen atom from the polymer main-chain by this ejected,

14

RADIATION CHEMISTRY OF POLYMERS

perhaps thermally activated, hydrogen atom itself. The role of the hydrogen gas so formed in radical migration was fully described by Dole and Cracco in 1961 and 1962 (50,51), and later through observation of the catalytic effect of hydrogen gas on the kinetics of decay of alkyl free radicals by Waterman and Dole (169) and Johnson and co-workers (145). The yield of hydrogen gas in irradiated polyethylene has been reported by a number of workers to fall in the range of G(H2 ) = 34, and shown by Kang and co-workers (159) to increase with increasing irradiation temperature. There is also evidence that the yield does depend to some extent on radiation dose (159) and polymer crystallinity (164). In addition to hydrogen gas, small quantities of hydrocarbon gas reective of the concentration and identity of side chains are produced on irradiation of polyethylene containing short chain branches. Bowmer and co-workers in 1983 (59) reported that the primary scission process leading to the formation of these small alkanes must occur at the tertiary carbon on the main chain. They also reported that the relative yields of volatile saturated products were essentially constant with increasing temperature, but that the total overall yield of gases did increase. On the other hand the formation of unsaturated products, principally ethylene and butylene, attributed to thermal depolymerization, increased markedly at elevated temperatures. Unsaturation. The formation and reaction of unsaturated groups in polyethylene has been extensively studied by Dole and co-workers over a number of years, with a particular emphasis on the relationship between the rate of formation and decay of unsaturation with other radiochemical reactions. The main unsaturated product formed on irradiation of polyethylene is internal vinylene unsaturation. Dole and co-workers (46) suggested that up to 80% of the hydrogen gas evolved on irradiation resulted from processes leading to formation of vinylene groups. Dole and co-workers (61) further reported that the rate of formation of vinylene unsaturation was scarcely affected by irradiation at 77 K, at which temperature migration reactions are presumably retarded. This led them to suggest that molecular detachment of hydrogen molecules may be involved. Later, this group (159) reported that the G-value for the formation of vinylene groups was close to 2.4 at all temperatures in the solid state. In addition to the formation of vinylene unsaturation, evidence for the formation of conjugated double bonds can be found in the appearance of new absorption bands in the UV spectrum of irradiated polyethylene (170172). The rate of formation of dienes is signicantly higher than that expected from consideration of the statistics of formation of vinylene unsaturation; Fallgatter and Dole (170) speculate on possible cooperative mechanisms of formation. The formation of more highly conjugated species is suggested to result from secondary reaction (171,172). The kinetics of decay of double bonds initially present in the material have also been studied in detail. Dole and co-workers (61) reported that the decay of terminal vinyl or vinylene groups follows rst-order kinetics until the formation of vinylene groups, mentioned immediately above, becomes signicant. The exact mechanism of reaction of the vinyl end groups remains unclear; however, the participation of these groups in end-linking reactions (reaction 12) has been conrmed by Randall and co-workers (6264) and Horii and co-workers (65,66).

RADIATION CHEMISTRY OF POLYMERS

15

Polypropylene. Polypropylene undergoes net cross-linking when exposed to ionizing radiation in the absence of oxygen. However, the radiation chemistry of polypropylene differs markedly from that of polyethylene as a consequence of the presence of tertiary carbons along the polymer backbone. Therefore, polypropylene undergoes signicant levels of main-chain scission on irradiation (see PROPYLENE POLYMERS (PP)). Initial studies of this polymer were mainly concerned with understanding the changes in molecular weight and the rate of gel formation on irradiation. Much of this work was concerned with the development of fundamental relationships between the radiochemical yields and molecular weights. The reported yields of chain scission and cross-linking fall in the ranges G(S) = 0.1980.62 and G(X) = 0.06450.272, respectively (81,82,95,173176). The apparent large variation in values of G(S) and G(X) are in part due to the difculty in analysis of molecular weight or solubility changes in materials having broad molecular weight distributions, and also possibly due to a dose dependence of the G-value for chain scission (81,175,176). Despite the increased probability of main-chain scission in polypropylene compared with polyethylene, there is no evidence of the formation of stable scission radicals on irradiation at low temperatures. Rather extensive ESR evidence has been accumulated that the stable free-radical intermediates are those listed below:

On irradiation at low temperatures, alkyl radicals (III and IV) are formed with a radiochemical yield G(R) of 2.42.6, depending on the crystallinity of the material (99,164). A particularly informative set of experiments which reveal the extent of initial bond cleavage has been conducted by Buseld and co-workers (3538). They have studied the radiation-induced changes in the stereochemistry of isotactic polypropylene (iPP). The changes in the stereochemistry are a result of initial chain scission, isomerization, and rehealing of the broken bond, as shown below in Figure 1. Studies of irradiated iPP (3538,177) have provided evidence that the racemization reactions do not occur as isolated events, but rather that they may occur in clusters. The methyl region of the solution-state 13 C NMR spectrum of iPP can be resolved to the pentad level. The changes in the pentad distribution on irradiation are not consistent with random reactions occurring along the polymer backbone (3538). A number of mechanisms have been suggested to account for the observed changes in tacticity, involving for example an increased probability of reaction after initial reaction, but all schemes have in common the observation that the proportion of syndiotactic sequences formed on irradiation is high and that the rate of equilibration of these sequences is low. It was also found that

16

RADIATION CHEMISTRY OF POLYMERS

Fig. 1. The racemization of an initially isotactic polymer through chain scission, isomerization, and bond reforming.

electron-beam irradiation at high dose rates was much more effective in promoting changes in tacticity than irradiation to comparable doses (37). Polyisobutylene. In 1954 Charlesby (80) noted that the molecular weight of polyisobutylene (PIB) decreased rapidly on exposure to high energy radiation. It has since been well understood that the presence of quaternary carbons in the main chain of polymers predisposes such a material to degradation (see BUTYL RUBBER). In the case of PIB, the primary reactions during radiolysis have been suggested by Alexander and co-workers (178) and Chapiro (179) to be those described in Figure 2. On the other hand Miller and co-workers (180) suggested that the initial chemical changes involve loss of a methyl hydrogen to form radical VII, followed by spontaneous -cleavage (Fig. 3). In addition, Ilicheva and Slovokhotova (181) have suggested that the energetically favoured radical VIII can also undergo -cleavage, as in Figure 4. The ESR spectra of PIB irradiated at low temperature consists of a broad doublet with a hyperne coupling constant of 2 mT, attributed to either radical VIII (182) or a combination of contributions from radicals VII and VIII (183, 184). The primary radicals V and VI (Fig. 2) have not been observed by ESR spectroscopy. Despite extensive studies of the radiochemical changes in PIB by ESR, volatile product analysis, and UV and IR spectroscopy, the mechanism of degradation had been uncertain until the advent of high resolution 13 C NMR methods. In a comprehensive study, Bremner and co-workers (185) and Hill and

Fig. 2. Proposed mechanism of degradation of PIB via initial main-chain cleavage.

RADIATION CHEMISTRY OF POLYMERS

17

Fig. 3. Proposed mechanism of degradation of PIB via scission of the side-chain methylene radical.

co-workers (186) identied the major products of degradation by NMR and other methods. They concluded that the schemes shown in Figures 24 are all possible degradation pathways, and that the major products are structures formed through the reaction scheme shown in Figure 2. Polyethers. The radiation chemistry of the polyethers has received relatively little attention. It is known, however, that all of this class of materials undergoes net cross-linking reactions. A comprehensive study of the radiation chemistry of polyoxymethylene (POM) has been reported by Fischer and Langbein (187). These authors reported that irradiation at the high dose rates of 150 kGy/min with 1-MeV electrons at 287 K resulted in a reduction in the total mass of the sample of approximately 1% per 100 kGy absorbed dose. There was a corresponding increase in the yield of new acetate chain ends determined by IR spectroscopy. The evolved gases consisted of many products, however the main products were hydrogen (G = 1.7), methane (G = 0.14), and CO (G = 0.013). The gases also included formaldehyde and higher molecular weight molecules. These results were combined with 1 H NMR analysis of the polymer after irradiation. Peaks due to new acetate and methoxy chain ends, and H-type cross-links, were observed. The yields of products obtained from the NMR spectra appear to be high, perhaps reecting the difculty in obtaining quantitative NMR spectra of irradiated materials. Despite this, they were able to conrm that this polymer does crosslink on irradiation. The effects of irradiation on the properties of poly(ethylene oxide) (PEO) have been examined by a number of workers (see ETHYLENE OXIDE POLYMERS). A number of these studies have been concerned with irradiation cross-linking of PEO in solution (188201). When irradiated in the solid state, PEO undergoes net cross-linking with the reported gel doses of 3.5 and 1.1 kGy, respectively, for highly crystalline, lower molecular weight and more highly disordered, high molecular weight materials (202). King reported earlier changes in solution viscosity of PEO irradiated in air and in vacuum (203). Other studies have been concerned with changes in the crystalline structure on irradiation; the crystalline order is reduced at moderate radiation doses consistent with cross-linking reactions (202,204,205).

Fig. 4. Proposed mechanism of degradation of PIB via scission of the main-chain methine radical.

18

RADIATION CHEMISTRY OF POLYMERS

Finally, the effects of radiation cross-linking on the chain dynamics have been examined by Schilling and co-workers (206) using solid-state 13 C NMR techniques. Poly(propylene oxide) (PPO) has also been observed to form a gel on irradiation (207). Roberts and co-workers (207) reported G(S) values of 0.22 and 0.55 and G(X) values of 0.15 and 0.31 for atactic and isotactic PPO, respectively. In this article the products of degradation are compared with previous reports of products of irradiation of PP and POM. It was found that the radiation chemistry is broadly consistent with that expected from previous studies. Poly(vinyl alcohol). The radiation chemistry of poly(vinyl alcohol) (PVA) is similar in many respects to the degradation of PVC, described below (see VINYL ALCOHOL POLYMERS). PVA undergoes a progressive darkening on irradiation at elevated temperatures or gets heated subsequent to irradiation due to the formation of sequences of conjugated double bonds (208). The mechanism of degradation proposed by Zhao and co-workers (208), and shown in Figure 5 below, involves the participation of radical cations. Previous studies have indicated that hydrogen gas is a primary product of the degradation of PVA (209,210), leading to the formation of the -carbon radicals, as shown in Figure 5. More recently, using ESR spectroscopy, Zainuddin and co-workers (211) have examined in detail the radical species formed on irradiation of PVA. They identied contributions for three radical species at low temperatures: the -carbon radical (triplet ESR spectrum), a radical anion (doublet spectrum), and an unidentied neutral species (doublet radical). The latter two species contribute 62% of the total radical concentration at 77 K. These results are broadly consistent with the mechanism proposed in Figure 5; however, neither article comments on the expected changes in molecular weight on irradiation. Lu and co-workers (212) have shown that the molecular weight of PVA decreases on irradiation, with the polymer remaining soluble up

Fig. 5. Proposed mechanism of formation of conjugated structures during the radiolysis of poly(vinyl alcohol).

RADIATION CHEMISTRY OF POLYMERS

19

to doses greater than 1000 kGy. It is apparent that chain scission reactions must dominate over cross-linking reactions. Poly(vinyl chloride). The mechanism of radiation chemistry of poly(vinyl chloride) (PVC) in many ways resembles the well-known thermal degradation of this polymer (see VINYL CHLORIDE POLYMERS). The principal results of exposure of PVC to ionizing radiation are evolution of hydrochloric acid and development of intense absorption bands in the UV and visible regions of the electromagnetic spectrum (6769,213). The result of dehydrochlorination is the formation of polyene radicals. For example, absorption peaks at 252, 291, and 330 nm were assigned to allyl, dienyl, and trienyl radicals, respectively. The dehydrochlorination reactions dominate the radiation chemistry; the yield of cross-linking reactions is very low, as reported by Chapiro (69). Salovey and Gebauer (214) later reported from the changes in the polydispersity of irradiated PVC that chain scission and crosslinking are occuring simultaneously on irradiation in air. Aliphatic Polyesters. Over the past 10 years the linear aliphatic polyesters, the polyalkanoates, and the copolymers of lactic and glycolic acids have attracted enormous interest as a result of their widespread use as biomedical materials. The radiation stability of these materials is of importance because of the potential to effect sterilization through exposure to ionizing radiation (215225). Polyhydroxybutyrate. Copolymers of hydroxybutyrate and hydroxyvalerate possess many mechanical properties in common with synthetic polyolens, and so have attracted much attention as replacements for these materials in environments where biodegradability is an important parameter (see POLY(3HYDROXYALKANOATES)). Carswell-Pomerantz and co-workers (217,218) have reported a detailed study of the radicals formed on irradiation of such materials. They found that the yield of radicals at 77 K was G(R) = 1.7 0.2, independent of copolymer composition, but that on irradiation at 300 K, the yield of radicals was reduced for the copolymers because of their lower glass-transition temperatures compared with the homopolymer. At low temperatures a signicant contribution to the ESR spectra from radical anions was noted. These radicals were observed to decay on warming to produce scission radicals. At still higher temperatures, radicals produced by abstraction of a methylene proton adjacent to the carbonyl group were detected. The predominant reaction on radiolysis of polyalkanoates is chain scission. Carswell-Pomerantz and co-workers (218) have shown that the main volatile products of radiolysis are CO, CO2 , and H2 . Scission at the ester unit results in a reduction in the molecular weight of these materials, and the G-value for scission G(S) for PHB was calculated to be 1.3. Using solution NMR, a very similar yield of reactions was determined from measurement of the yield of new chain ends. Cross-linking is not important in the radiation chemistry of PHB; however, analysis of the changes in molecular weights of irradiated PHBV copolymers provided evidence of some cross-linking, with the ratio of scission to cross-linking being 9 and 6.8 for materials having valerate mole fractions of 0.184 and 0.263, respectively. Poly(lactic acid-co-glycolic acid). The degradation of the homopolymers and copolymers of lactic and glycolic acids (PLGA) has been reported by a number of authors (220223). Babanalbandi and co-workers (220) have measured the G-radical values at 77 K for poly(L-lactic acid) (L-PLA) and poly(D,L-lactic acid)

20

RADIATION CHEMISTRY OF POLYMERS

(D,L-PLA) and reported these to be 2.0 and 2.4, respectively. They also reported G(S) = 2.3 and G(X) = 0.0 for D,L-PLA and G(S) = 2.4 and G(X) = 0.28 for L-PLA. These results are in contrast to the values of Collett and co-workers, who suggested that rather than chain scission dominating the radiation chemistry, these polymers form a gel on irradiation. The results of Babanalbandi and co-workers (221), in which new aliphatic chains ends formed by cleavage of the main chain at the ester unit are observed, are in support of a mechanism in which chain scission dominates cross-linking. These authors reported G-values for the formation of chain end structures comparable with earlier study. Furthermore, the main volatile products of radiolysis of PLA and poly(glycolic acid) (GPA) are CO2 and CO, consistent with chain scission being the most important reaction. In addition, small amounts of hydrogen and ethane gas were observed on the radiolysis of PLA. Finally, Montanari and co-workers (226) have examined the effects of radiation sterilization on the stability of PLGA microparticles used for drug delivery. Aliphatic Polysulfones. The class of copolymers of olens and sulfur dioxide are well known for undergoing extensive degradation on exposure to high energy radiation. In 1965 Ayscough and co-workers (227) reported an ESR study of the degradation products of sulfones and poly(olen sulfone)s. They were able to conrm that the relatively weak C S bonds are more susceptible to rupture compared with the C H bonds. Later, Brown and ODonnell (228,229) conrmed that this initial chain breakage leads to loss of SO2 and permanent chain scission and reduction of molecular weight. The yield of scission reactions was found to increase progressively up to the ceiling temperature for the propagationdepropagation equilibrium, at which temperature rapid depropagation occurred. The yield of olen gas evolved on radiolysis was observed to be lower than the yield of sulfur dioxide, because of the cationic homopolymerization of the olens initiated by a polymeric carbonium ion (45,230). Isomerization of the olen via a carbonium ion intermediate was also conrmed. The mechanisms of these reactions are summarized in Figure 6. Polymethacrylates and Polyacrylates. The high energy radiation chemistries of polyacrylates and polymethacrylates were some of the rst studied in detail for any polymer (see ACRYLIC ESTER POLYMERS; METHACRYLIC ESTER POLYMERS). For example, in the 1950s, Charlesby and co-workers reported extensively on the breakdown of poly(methyl methacrylate) (PMMA) on exposure to high energy radiation, and also on studies of coloration and gas formation associated with the loss of the side chains (20,39,80). The molecular weight of PMMA was shown to decrease on irradiation (80,231), with the decrease inversely proportional to the irradiation dose. The latter observation was later rationalized through the famous CharlesbyPinner relationship (95), which allowed G-values of chain scission and cross-linking to be evaluated. Shultz and co-workers (232) reported that for 1-MeV electron irradiation of PMMA under vacuum, G(S) = 0 and G(X) = 1.65, and Wall and Brown (231) examined the yields for the principal gases formed on radiolysis under vacuum: H2 , CH4 , CO, and CO2 , for which the G-values were 0.21, 0.54, 0.45, and 0.32, respectively [G(gas) = 1.6]. ESR studies by Ovenall and co-workers (233236) during the same period identied the natures of the spectra of the radicals formed in PMMA on radiolysis over a temperature range from ambient to 77 K. They identied the multiline

RADIATION CHEMISTRY OF POLYMERS

21

Fig. 6. Proposed mechanism of degradation of poly(olen sulfone)s.

spectra observed at room temperature as being principally associated with chainend radicals, which were characterized by a nine-line spectrum at room temperature, with G(R) = 2.5. Later, ODonnell and co-workers (237) were able to account for the nine lines in terms of two overlapping spectra, one consisting of ve and the other of four lines, arising from two conformations of the chain-end radicals. Bowden and ODonnell (238) later showed that the relative contributions of the two conformations varied with the observation temperature. A consideration of the molecular weight changes and the natures and yields of the gaseous and radical products formed on radiolysis at room temperature allowed a mechanism for the radiolysis of PMMA to be formulated. The principal process associated with the degradation was scission of the ester side chain, forming the principal gaseous products and a main-chain radical which can further undergo -scission to form the propagating radical and a chain-end double bond. It was also thought that some direct main-chain scission may occur at the -carbon. The coloration of PMMA on radiolysis was believed to arise from the formation of oxygen and temperature-sensitive radical and charged species.

22

RADIATION CHEMISTRY OF POLYMERS

David and co-workers (32) conrmed the occurrence of direct main-chain scission and reported that racemization also occurred during radiolysis of syndiotactic PMMA. In the presence of a radical scavenger, no racemization was found to take place, suggesting that radical processes are involved in the racemization process (33). These results therefore indicate that irradiation must induce some temporary main-chain scission, which is followed by recombination of the scission radicals within the cage leading to a concomitant change in the stereochemistry of the polymer. The extent of the racemization observed for a -ray dose of 8 MGy was reported to be approximately 24%. Because PMMA has a low ceiling temperature, the formation of propagating radicals on radiolysis at 453 K, with G(S) = 10, Charlesby and Moore (40) found spontaneous depolymerization of the polymer to monomer at this temperature. In a later article, David and co-workers (239) showed that depolymerization also occurred when PMMA was irradiated at ambient temperature and then heated to 433 K, and that the yield of monomer was independent of the absorbed dose and the initial molecular weight of the PMMA. By contrast with PMMA, poly(methyl acrylate), PMA, and several other aliphatic polyacrylates were found by Shultz and Bovey (240) to undergo crosslinking and gel formation on irradiation with 1-MeV electron beams. They reported G(S) = 0.15 and G(X) = 0.52 for PMA. Graham (241,242) reported that the phenyl, benzyl, and 2-phenyl ethyl acrylate polymers also undergo cross-linking on radiolysis under vacuum. There was evidence of side-chain scission also, and according to Fox and co-workers (243,244) the major volatile products were similar to those observed for PMMA. In these early years, the high energy radiation induced reactions of PMA and PMMA were not the only polyacrylates and polymethacrylates investigated, but they were found to be representative of the radiation chemistries of the two polymer families with different ester side-chains. With improvements in the sensitivities of NMR and ESR spectrometers during the 1970s and the growing importance of microlithography in the production of solid-state devices, extra interest was directed to studies of the radiation-induced reactions of both polymethacrylates and polyacrylates. Because the methacrylate polymers almost entirely undergo scission on exposure to radiation under vacuum and the acrylates undergo cross-linking, they can be utilized as the basis for positive and negative resists, respectively. Thompson and co-workers (42) and Bowden (245) reported studies of the applications of PMMA in resist development, and examined the effect of incorporation of heavy atoms to enhance the sensitivity of PMMA to X-rays. Bowden (245) found that the sensitivity of a resist depends not only on the nature of the polymer but also on its molecular weight and molecular weight distribution, and that it increases with deceasing polydispersity. The development dose was found to decrease with increasing G(S) for positive resists such as the polymethacrylates. In a series of articles, Buseld, ODonnell and co-workers reexamined the radical yields, gaseous products, molecular weight, and material property changes on radiolysis of PMMA and PMA (22,23,27,246248). ODonnell and co-workers (22,248) have reevaluated radical formation on irradiation of PMMA at 77 K and room temperature. At 77 K the spectrum was analyzed in terms of the presence of CHO, CH3 , COOCH3 , and COOCH2 ,

RADIATION CHEMISTRY OF POLYMERS

23

as well as the propagation radical. At room temperature, only the propagating radical spectrum was observed. The G-value for overall radical formation at 77 K was 0.58, which is lower than that found at room temperature, 2.2, indicating the importance of radical recombination within the cage at 77 K. On annealing of the radicals produced at 77 K to room temperature and above, the radical decay is not linear, but reects the onset of thermal transitions in PMMA (249). A wide range of gaseous products have been identied by Buseld and coworkers (246) on radiolysis of PMMA at room temperature, which arise principally from a breakdown of the ester side-chains. The principal products and their yields G are respectively H2 , 0.34; CH4 , 0.66; CO, 1.08; CO2 0.68; CH3 OH, 0.36; HCO2 CH3 , 0.69; dimethoxymethane, 0.11; and CH3 Ac, 0.02. Kudoh and coworkers (250) have measured the yields of H2 , CH4 , CO, and CO2 on radiolysis at 77 K and report that whereas the hydrogen yields are the same at 77 K and room temperature, those for the other gases are lower at 77 K. This is in accord with the temperature dependence of the observed radical yields (248) and with the mechanical properties (251). Buseld and co-workers (246) have examined the molecular weight dependence for radiolysis of PMMA in vacuum at room temperature, and Kudoh and co-workers (250) have also reported the dependence at 77 K. The values of G(S) at room temperature were 1.3 (246) and 1.7 (250) and at 77 K, 0.24 (250). These reect the reported radical yields at the two temperatures. The variation of the molecular weight distribution with dose for e-beam irradiation of PMMA has been studied by Viswanathan (83). He has modeled the dependence of the distributions for use in predicting the sensitivities of positive resists in terms of product solubilities for development following exposure. Buseld and ODonnell (23) have examined the correlations between the changes in the molecular weight and the tensile properties of PMMA following radiolysis at room temperature. Both the exural and compressive strengths decrease with increasing dose, reecting the falling molecular weight of the polymer. Kudoh and co-workers have drawn similar conclusions for radiolysis at 77 K and ambient temperature (251). The racemization of PMMA on radiolysis has been studied by Luo and coworkers (34), Dong and co-workers (31), and Thominette and Verdu (252) by use of NMR. Luo and co-workers (34) reported that for radiolysis of isotactic PMMA in solution, the yield of scission was greater than that for racemization, but the order was reversed for radiolysis in the solid state, indicating an important role for molecular motion of the scissioned chains. Dong and co-workers (31) investigated the radiolysis of both highly isotactic and highly syndiotactic PMMA. They found that their results were consistent with a temporary chain scission G-value of 18.6 at 353 K, and the polymers tended to approach a common average tacticity with increasing absorbed dose, which was the same as that found for PMMA prepared by free-radical polymerization. By comparison with PMMA, the radiolysis of PMA has received much less recent attention. Buseld and co-workers (247) have measured the G-values for gaseous product formation at room temperature and 423 K, and they have also measured the dose to gel. The major volatile products were the same as those observed for the radiolysis of PMMA, but the total gas yield was much smaller: G(gas) = 1.6 compared with 3.9 for PMMA. However, the gas yield increased with

24

RADIATION CHEMISTRY OF POLYMERS

temperature, reaching 3.9 at 423 K. At 195 K the dose to gel was approximately 0.25 MGy, and decreased with increasing temperature. The sensitivities of a wide range of other acrylate and methacrylate polymers to radiation have also been investigated, including poly(acrylic acid) (85,253), poly(methacrylic acid) (85,254263), poly(hydroxyethyl methacrylate) (264,265), polymethacrylates with longer ester side-chains (266,267), and poly(ethylene glycol dimethacrylate) (268). In general, the behaviors of these polymers mirror the behaviors of the related methyl ester polymers, but some observations are worthy of special mention. Hill and co-workers (85) have reported a study of the scission and crosslinking yields on radiolysis of poly(acrylic acid) and poly(methacrylic acid) using sedimentation equilibrium to obtain the weight- and Z-average molecular weights as a function of absorbed dose. The values G(S) = 0, G(X) = 0.44 and G(S) = 6.0, G(X) = 0, respectively, were obtained for the two polymers, highlighting the signicant role of the substituent methyl group in determining the outcome of the radiolysis. Kudoh and co-workers (259) have investigated LET effects for electron radiolysis of PMMA. A comparison of the scission probabilities for radiation (0.5 Gy/s) and high energy electron beams (up to 4 1010 Gy/s) showed no evidence for dose rate effects; thus they found no evidence for spur overlapping at the high electron beam dose rates. Dong and co-workers (267) studied the role of the length of the aliphatic side chains in a series of highly syndiotactic polymethacrylates and found that poly(heptyl methacrylate) indeed undergoes gelation on radiolysis, via the heptyl side chain. Heavy ion irradiation of PMMA has also received some attention, particularly by researchers at the Japan Atomic Energy Research Institute. Kudoh and co-workers (269273) have compared the effects of radiation and electron beams with those for 30 and 45 MeV protons and heavier ion beams on the tensile and molecular weight properties of PMMA. They have found no difference between the sensitivities of the tensile properties of PMMA to -rays and electron and proton beams as a function of dose. However, the molecular weight dependence for PMMA showed a clear LET dependence for heavy ions above a critical LET, indicating overlapping between spurs. For low LET radiation the scission probability remains constant, but the scission probability decreases with increasing LET for high LET radiation. The critical LET level for PMMA is a few hundred MeV cm2 /g. Esters of Other Acids. Poly(tert-butyl crotonate) (PtBC) is an analogue of poly(tert-butyl methacrylate) (PtBMA) but it does not have the disubstituted main-chain carbon atom. Thus, on these grounds, PtBC might potentially undergo cross-linking, but this process may potentially be hampered by the severe steric hindrance involved in the formation of a cross-link at the main-chain estercarbon atom. ODonnell and co-workers (274,275) have reported studies of the yields for radical formation, cross-linking and scission, and the gaseous products on radiolysis of PtBC. Three major radicals were observed on radiolysis at 77 K: tertiary butyl radicals, oxygen centred radicals formed by loss of the tertiary butyl groups, and a third unidentied radical. The value of G(R) was 1.28. The major gaseous radiolysis products were the same as those observed by Ungar and co-workers (276) for PtBMA, and the yield of CO2 (G = 1.0) was greater than that for CO (G = 0.26), in contrast to the observations for PMMA and PMA. This is consistent

RADIATION CHEMISTRY OF POLYMERS

25

with the preferential loss of the labile t-butyl ester group. The values of G(S) = 0.59 and G(X) = 0.66 for PtBC can be compared with those for PtBMA, G(S) = 0.21 and G(X) = 0.17, reported by Shultz and Bovey (240). These values indicate that PtBC indeed undergoes cross-linking, while PtBMA undergoes net scission. However, the nature of the cross-links formed in PtBC have not been identied. Polysiloxanes. Poly(dimethyl siloxane) (PDMS) undergoes cross-linking on exposure to ionizing radiation. As early as 1954, Charlesby (78,277,278) had reported that PDMS can be cross-linked to form a transparent rubbery material, and that the dose to achieve a cross-linked network was inversely proportional to the initial polymer molecular weight (see SILICONES). Early work also reported that the main volatile products of irradiation are hydrogen, methane, and ethane gases (277,279,280). Other chemical changes observed with IR spectroscopy were Si H structures and new, unidentied Si-O structures (280,281). ESR studies have revealed that on irradiation at 77 K CH3 , CH2 , Si, and O radicals are produced, and that, as expected, the radicals react on warming above the glasstransition temperature (127,282,283). G-values between 2.3 and 2.7 for crosslinking have been reported (279,284,285). Whereas some evidence for chain-scission structures has been obtained from these previous studies, not until recently have conclusive measurements of the yields of scission reactions been reported (286288). Hill and co-workers (286,287) have identied new structures formed in PDMS using solid-state 29 Si and 13 C NMR. They were able to measure the rate of formation of a number of chain-end, side-chain, and cross-link structures. The mechanism of cross-linking is dominated by Y-linking, although two minor H-linking products were also observed. On irradiation at 303 K, the radiochemical yields were G(S) = 1.3, G(H) = 0.34, and G(Y) = 1.7.

Fluorinated Polymers. Fully Fluorinated Polymers. The radiation chemistry of fully uorinated
polymers shows remarkable temperature dependence, with all of the uorinated thermoplastics undergoing degradation, ie, chain scission, at ambient temperatures, but with an increasing yield of cross-linking reactions at elevated temperatures. Over the past 10 years, this has led to renewed interest in the radiation chemistry and applications of these materials (see PERFLUORINATED POLYMERS, POLYTETRAFLUOROETHYLENE). As discussed in a number of reviews (1,2,289), the parent of the uoropolymer family, polytetrauoroethylene (PTFE), undergoes degradation on irradiation. Over a number of years, this process has been exploited commercially to reduce the molecular weight of PTFE resins. Despite this recognition, the details of the mechanism of reaction have remained unclear, in part due to the paucity of experimental techniques available to study these materials. The inherent insolubility of PTFE has until recently restricted studies of radiation chemistry of PTFE to measurements of mechanical and thermal properties, and melt viscosity. For example, Bro and co-workers (290) reported in 1963 that the melt viscosity of PTFE decreased on irradiation followed by heating. This effect was more pronounced in the presence of air. In another example, Hedvig (291) in 1969 showed that the mechanical strength and in particular the elongation at break of irradiated PTFE dropped dramatically on irradiation, but that these properties disappeared on irradiation in air.

26

RADIATION CHEMISTRY OF POLYMERS

As mentioned above, the mechanism of degradation of PTFE is unclear, particularly the location of radiation-induced chemical reactions. However, ESR spectroscopy (290293) has conrmed that the stable free-radical species at low temperatures are the chain-end radicals IX (shown below) and that at higher temperatures the secondary radicals X dominate the spectrum. Hedvig (291) has suggested that chain-end radicals terminate by abstraction to form the more stable secondary radical. These radicals persist for very long times, and this leads some authors (58) to suggest the suitability of PTFE as a radiation dosimeter.

The effect of irradiation at elevated temperatures was initially investigated by Lovejoy and co-workers (294), who found that irradiation of PTFE above the glass-transition temperature resulted in a decrease in the molecular weight of the material, as evidenced by a decrease in melt viscosity. In a later series of articles Tutiya (295299) studied the effect of various irradiation conditions on the broadline 19 F NMR spectra of PTFE. He reported in 1972 (295) that on irradiation above the melting temperature there was a signicant decrease in the crystalline content, ascribed in part to the increased melt viscosity of the irradiated polymer. It was not until about 20 years later that these observations were pursued. In 1994 and the early 1990s, Sun and co-workers (55,300) and Oshima and co-workers (56) reported the radiation-cross-linking of PTFE above its melting temperature. Since then, Oshima and co-workers (56,301310) have examined in detail the mechanism of formation and properties of radiation-cross-linked PTFE. It was found that cross-linked PTFE had signicantly enhanced radiation stability compared with the uncross-linked resin despite the radical yield at low temperatures being signicantly higher (304,308,309). PTFE undergoes cross-linking via a Y-type mechanism (reaction 12 above). This has been conrmed by solid-state 19 F NMR measurements both by Katoh and co-workers (307,310) and Fuchs and co-workers (311,312). Initially, the Y-linking mechanism was favored over the H-linking mechanism after considering steric effects. However, it has been conclusively shown that branch points are formed and that the yield of branch points is superior to the yield of new chain ends (312), whereas Oshima (310) has suggested that double bond structures may also participate in the branching reactions. In an early article Lovejoy and co-workers (294,313) reported that the copolymer of tetrauoroethylene and hexauoropropylene (FEP) would undergo degradation on irradiation at room temperature, but that it cross-links at higher temperatures. In 1985 Seguchi and co-workers (8) conrmed that FEP indeed undergoes cross-linking reactions through measurements of the soluble fractions after irradiation. The primary free radicals have been identied by Hill and co-workers (30) to be radicals IX and X above. In addition the yield of free radicals is signicantly enhanced at room temperature compared with irradiation at 77 K. This is a consequence of the lower probability of cage recombination of the primary species at higher temperatures. In a later series of articles

RADIATION CHEMISTRY OF POLYMERS

27

(314316) these authors have measured the yields of new structures by solidstate 19 F NMR spectroscopy and found that, as with PTFE, irradiation below the melting temperature resulted in predominately new chain ends due to chain scission. On the other hand, upon irradiation at 523 K, just above the melting temperature, new branch or cross-link structures were formed in signicant quantities. Similar conclusions have been reached for the third commercially important fully uorinated polymer, namely the copolymer of tetrauoroethylene and peruoropropyl vinyl ether (317320). This is not unexpected, because this material typically contains approximately 2 mol% of the vinyl ether. Irradiation at elevated temperatures by electron beam or with -rays resulted in increased yields of cross-links. Finally, Forsythe and co-workers (321) have also reported the radiation chemistry of members of the Teon AF DuPont family of thermoplastics, and found that degradation proceeded through loss of hexauoroacetone of the dioxole ring, leading to chain-scission reactions. Fluoroelastomers. The above discussion has highlighted the importance of molecular motion on the ultimate radio-chemical products formed in uorinated polymers. This is a consequence of the limited number of mechanisms of termination of radicals in uorinated polymers when compared with protonated polymers. Specically, in uoropolymers there is no mechanism of radical migration equivalent to the hydrogen-hopping mechanism at work in irradiated hydrogencontaining polymers. The three-body intermediate suggested to participate in the hydrogen-hopping mechanism (eq. 4) is unlikely to form on radiolysis of uorocarbons, on purely energetic grounds. Thus, the radicals produced in uoropolymers are particularly long-lived, and when they do react it is usually through combination reactions. This means that the radicals must have sufcient mobility to be able encounter one another; hence the probability of chain cross-linking increases with increasing temperature, and in the semicrystalline materials discussed above, sufcient mobility is only achieved above the crystalline melting temperature. The situation is somewhat similar for uorinated elastomers (see FLUOROCARBON ELASTOMERS). The radiation chemistry of the copolymer of tetrauoroethylene (TFE) and peruoro-methyl vinyl ether (PFMVE), known commercially as Kalrez, has been studied in detail by Hill and co-workers (57,322327) This material is a random copolymer of these two monomers with a TFE to PFMVE molar ratio of 2:1. It is a fully amorphous polymer with a glass-transition temperature close to 273 K. On irradiation at room temperature it undergoes cross-linking, as conrmed by measurement of structural changes by NMR spectroscopy (57), and changes in mechanical and solubility properties (327). When this material is irradiated below its glass transition temperature, chain scission tends to dominate over cross-linking reactions (326). Partially Fluorinated Polymers. The radiation chemistry of uoropolymers is very sensitive to the presence of hydrogen in the chemical structure. The introduction of hydrogen to the backbone generally results in a marked increase in the probability of cross-linking reactions occurring. Presumably, this is a result of the increased mobility of radicals as discussed immediately above. In addition, as pointed out by Wall and co-workers (328), the introduction of hydrogen leads to the formation of HF on irradiation, with concomitant formation of main-chain

28

RADIATION CHEMISTRY OF POLYMERS


Table 2. Radiochemical Yields of Cross-Linking and Scission for Several Partially Fluorinated Polymers Polymera PVDF PVDF PVF PTrFE PTrFE ECTFE
a PVDF,

G(X) 1.0 0.951.05 38 1.1 1.11.2

G(S) 0.3 0.20.4 1.84.0 0.4 0.30.5

p0 /q0 0.15 0.180.21 0.280.35 0.18 0.160.21 0.3

Ref. 319 317 317 319 317 322

poly(vinylidene uoride); PVF, poly(vinyl uoride); PtrFE, poly(triuoroethylene); ECTFE, poly(ethylene-co-triuoroethylene).

double bonds. These double bonds act as efcient sites for subsequent cross-linking reactions. Of this class of polymers, the radiation chemistry of poly(vinylidene uoride) (PVDF) has been examined in most detail. In 1962, Timmerman and Greyson (329) compared the mechanical properties of PVDF with PTFE and found that overall PVDF was more resilient to exposure to radiation. Not long after this, Yoshida and co-workers (330) reported the yields of cross-linking and scission reactions in irradiated PVDF, and that cross-linking dominated. The radiochemical yields of scission and cross-linking of PVDF and a number of other materials are listed in Table 2. The identity of the radical species formed on irradiation has been conrmed by Seguchi and co-workers (331). They found that the chain-end radical XI dominated the ESR spectrum at low temperatures, but that the main-chain radical XII became more important on warming. On long standing the polyenyl radical XIII came to dominate the ESR spectrum, highlighting the importance of loss of HF, and the increased radical mobility compared with fully uorinated polymers.

As mentioned above, all of this class of polymers undergoes cross-linking on irradiation in vacuum. The yields of cross-linking and scission for these polymers are listed in Table 2. Of note also are the reports of cross-linking predominating in the copolymers of ethylene with TFE (332) and of ethylene with chlorotriuoroethylene (333). Polychlorotriuoroethylene. In contrast to the above materials, the homopolymer of polychlorotriuoroethylene (PCTFE) undergoes radiation degradation on exposure to ionizing radiation. Florin and Wall reported in 1961 that PCTFE undergoes chain scission, and calculated a G-value of 0.67 for scission from the dose to achieve zero strength. A relatively low yield of volatile products was reported: G(volatiles) = 0.11. Very recently, Hill and co-workers reexamined the radiation chemistry of this polymer (334,335). They measured the yields of free radicals to be 1.55 at 77 K decreasing to 0.32 at room temperature (335). The

RADIATION CHEMISTRY OF POLYMERS

29

ESR spectra were difcult to identify because of the large anisotropic hyperne coupling to the uorine nuclei. These authors also examined the radio-chemical products using solid-state 19 F NMR and FTIR spectroscopies (334). On irradiation below the melting temperature, there was no evidence of branch/cross-link formation. However, the yields of chain ends, double bonds, and branch points on irradiation at 493 K were 3.6, 0.2, and 0.5, respectively. The structure of the branches could not be conrmed; however, the respective yields of chain ends and branches indicate that scission dominates for this polymer. Polydienes. The polydienes undergo cross-linking on radiolysis in vacuum. The principle reactions occurring are cross-linking, consumption of double bonds, isomerization of double bonds, and evolution of hydrogen gas. There is considerable variation in the reported G-values for these reactions, arising primarily from differences in the experimental techniques used in their determination. For example, physical measurements, usually of soluble fraction or swelling ratio, used to determine yields of cross-linking and scission reactions are particularly susceptible to errors in the case of polydienes. These techniques assume a random distribution of cross-linking and scission reactions; as will be seen below, this assumption may break down in the case of radiolysis of polydienes. Polybutadiene. The primary products of irradiation of polybutadiene (PBD) are allyl free radicals. The yield of radicals has been reported to range from 0.19 to 0.44 (336,337). On heating the radicals decay in a monotonic fashion to low concentrations just above the glass-transition temperature of the polymer (337) (see BUTADIENE POLYMERS). As mentioned above, the net result of irradiation of PBD is cross-linking. Gvalues for cross-linking, determined as described immediately above, range from 3.2 to 6.7 (338343). In one case a G(S) value of 0.52 has been reported, although it is more generally accepted that the G-value for scission in irradiated PBD is much smaller than this (343). As will be discussed below, these G-values contrast with the very high yield of reactions which consume double bonds in these polymers. The reason for this is the highly nonuniform cross-linking reactions in polydienes. The advent of solidstate 13 C NMR allowed direct observation and quantitation of the cross-link structures in irradiated polymers. In the case of poly(1,4-butadiene), ODonnell and Whittaker (344,345) have reported very high G-values for consumption of double bonds, and formation of cross-link structures. The authors have suggested a chain reaction through successive double bonds to form a highly nonuniform network structure. As a result, the effective cross-link density is much lower than the measured chemical cross-link density. Similar results were reported for materials having different initial microstructure. A number of workers have reported a very high rate of disappearance of double bonds on irradiation of PBD. Using IR spectroscopy, Parkinson and Sears (346) determined G(cis) to be 15, G(-trans) 1122, and G(cistrans) 7. The last reaction is the isomerization of a cis double bond to a trans double bond via an allylic radical intermediate. The rate of consumption of 1,2-vinyl groups was found to depend on the initial vinyl content; values from 0.2 to 40 were reported. Similarly, Golub (347,348) reported a G-value of 7.9 for consumption of double bonds, and 7.2 for isomerization. He suggested the formation of cyclic structures. ODonnell and Whittaker (344) have reported the yields of these reactions for a number

30

RADIATION CHEMISTRY OF POLYMERS

of polybutadienes. The calculated G-value for loss of double bonds was initially 41.8 over a dose range of 01 MGy and decreased to 6.3 at higher doses (410 MGy). For poly(1,2-butadiene) the G-value for loss of double bonds was 237 (00.5 MGy), because of, it was suggested, an intramolecular chain reaction of the vinyl groups. Polyisoprene. As with polybutadiene, polyisoprene (PIP) and natural rubber are cross-linking polymers. The radical intermediates have been identied as allyl radicals (336,337,342). Radical yields at 77 K ranging from 0.28 to 0.73 have been reported. It is also clear that the mechanism of reaction of these radicals is similar to that discussed above for polybutadiene. A G-value of 6.7 for the decrease in unsaturation has been reported by Golub and Danon (347). Much later, using solid-state NMR, Whittaker (349) obtained a G(d.b.) value of 45.8 (01 MGy). Signicantly lower yields of cross-links have been reported; Boehm (350) has compiled a list of G(X) values ranging from 0.4 to 3.5 as determined by measurement of physical properties. Polychloroprene. In contrast to PBD and PIP, the radiation cross-linking of polychloroprene does not appear to proceed by a chain-reaction mechanism (see CHLOROPRENE POLYMERS). The yields of cross-links [G(X)] were measured by a number of methods and determined to be 3.9 (NMR), 4.8 (swelling ratio), and 3.2 (sol. fraction) (351). This close correlation, and the observation of a narrow resonance in the 13 C NMR spectrum assigned to cross-link structures, indicated that cross-linking in this material occurs randomly throughout the material. Peaks are observed in the NMR spectra because of new chain structures formed on irradiation.

Aromatic Polymers. Polystyrene. Polystyrene undergoes net cross-linking on irradiation (see


STYRENE POLYMERS). In 1953 Charlesby (352) reported that the polymer undergoes gelation on irradiation, and that the yield of permanent chain scission reactions is small or negligible. Furthermore, he reported that despite cross-link reactions dominating the radiation chemistry, the overall yield of cross-linking is very small compared with nonaromatic polymers, because of the presence of the benzene ring. A major study of the changes in molecular weight (viscosity) and formation of volatile products was subsequently undertaken by Wall and Brown in 1957 (231). The major volatile product, hydrogen gas, is produced in low yields [G(H2 ) = 0.0220.026] compared with nonaromatic polymers. Since these early reports, a number of further studies have appeared (6,231,339) and these have been summarized in 1973 by Parkinson and Keyser (353). Since that date, efforts have been made to understand the relatively wide discrepancies in reported yields of radiochemical reactions, and also the effect of temperature on the radiolysis products. For example, values of the yield of radicals, G(R), measured at 77 K after irradiation at that temperature range from 0.04 to 0.50 (22,100,354357). ODonnell and co-workers (358) have proposed that the variation in reported yields of radical formation, and of chain scission and cross-linking, may be in part due to the presence of trace amounts of residual oxygen. Using ESR spectroscopy, the mechanism of radiation-induced cross-linking of polystyrene has been revealed to a large extent by careful studies of radical intermediates. The major free-radical products formed on irradiation are the

RADIATION CHEMISTRY OF POLYMERS

31

-carbon radical XIV and the cyclohexadienyl radical XV shown below. The contribution of radical anions to the ESR spectra was conrmed by Garrett (357) and Hill and co-workers (24). On heating these samples toward the glass-transition temperature (ca. 373 K), free-radical reactions occur until at room temperature the ESR spectrum resembles that observed for a sample irradiated at the elevated temperature. The radical concentration has decayed to negligible values at the glass-transition temperature. The mechanism of cross-linking has not been conrmed, but is suggested to involve recombination reactions between radicals X and XI. Measurements of gel doses for materials having different initial molecular weight distributions have suggested that H-linking rather than end-linking is the major mechanism of cross-linking (359).

The effect of irradiation temperature has been examined in some detail. A number of workers have reported that the yield of cross-links decreases progressively above room temperature with a corresponding increase in the yield of chain-scission reactions (302,360362). Burlant and co-workers (363) initially suggested that as the temperature increases, an increasing proportion of mainchain radicals XIV undergo a -cleavage reaction to form terminal unsaturation and chain-end radical. On the other hand, from analysis of changes in molecular weights, Bowmer and co-workers (360) reported that the ratio of yields of scission to cross-linking reactions, G(S)/G(X), increases from 0.02 at 303 K to 2.8 at 423 K. These authors suggest that competition between combination and disproportionation reactions between temporary scission radicals is responsible for this trend, with disproportionation, and hence permanent chain scission, becoming more likely at higher temperatures. A comparison was made with the mechanism of termination in the free-radical polymerization of styrene. Seguchi and co-workers (302,362) have also examined in detail the effects of temperature and tacticity on the relative yields of cross-linking and scission reactions in irradiated polystyrene, and have arrived at similar conclusions. Poly(-methyl styrene). Poly(-methyl styrene) (PMSTY) is one of the class of degrading polymers, having along its backbone a quaternary carbon, at which a stable scission radical can be formed. In addition to this, PMSTY has a low polymerization ceiling temperature of 334 K (364). At and above this temperature, the polymer will undergo depolymerization to its monomers. Garrett and co-workers (365) measured the yield of chain scission G(S) and the rate of depolymerization at temperatures of 298, 353, and 536 K. At the two lower temperatures the weight loss due to depropagation was negligible and the authors were able to calculate yields of chain scission from changes in the number-average molecular

32

RADIATION CHEMISTRY OF POLYMERS

weights (see eq. 5 above). The values of G(S) were 0.29 0.01 and 0.48 0.06 at 298 K and 353 K, respectively. However, at the highest temperature, extensive and rapid rst-order weight loss occurred. The G-value for depolymerization [G(M)] was found to be 320 45. The results of a Monte Carlo simulation of this process resulted in an estimate of the yield of chain scission [G(S)] of 1.8 and the average unzipping length of 400 (365). This same group has also reported an ESR study of the radicals formed on irradiation at 77 K and room temperature (54). The authors identied a number of radical species including a propagation radical formed by direct chain scission, and radicals associated with the side chain and the aromatic ring. Main-Chain Aromatic Polymers. The above discussion highlights the relative stability of polymers containing aromatic rings, due to the ability of these structures to dissipate excess energy through resonance effects (1). For example the yields of cross-links and free radicals in irradiated polystyrene are approximately 50 times lower in polyethylene. It has also been noted from comparisons of the radiation sensitivity of polymers having similar structures that those in which the structures are incorporated into the main chain are less sensitive to degradation than those with similar side-chain structures. This is partly due to the more effective cage effect in main-chain polymers, which tend to have higher glass-transition temperatures than those of side-chain polymers, but also due to more effective resonance coupling to the aromatic rings. It follows therefore that main-chain aromatic polymers represent some of the most radiation-resistant organic materials discovered so far. Aromatic Polysulfones. The presence of the relatively labile sulfone linkage in Polysulfones (qv) results in these materials having a tendency to undergo degradation reactions. Under many conditions, however, aromatic sulfones are relatively stable and undergo net cross-linking. Brown and ODonnell (366) initially reported the radiation chemistry of copolymers of bisphenol A and 4,4 dichlorodiphenyl (Table 3). They found that on irradiation in vacuum at ambient temperatures, this material undergoes cross-linking with relative yields of crosslinking and scission reactions being respectively 0.051 and 0.012. These studies lead to a large body of work considering the effect of chemical structure on the radiation sensitivity of aromatic polymers. In general for all aromatic polysulfones, on irradiation at ambient temperature, cross-linking predominates. In addition, as aliphatic groups are introduced into the main chain, the radiation sensitivity is increased. Lewis and co-workers (367) have reported a detailed study of the effect of polymer structure on the yields of chemical reactions for a series of polysulfone terpolymers containing varying amounts of bisphenol A and biphenyl units. They found a close to linear relationship between the yields of radicals and volatile sulfur dioxide with the ratio of the bisphenol A and biphenyl units. They report that the biphenyl unit imparts signicantly higher radiation stability. As with other polymers, the details of the radiation chemistry are affected by the irradiation temperature, with Hill and co-workers (368) providing evidence of enhanced chain scission on irradiation of polysulfone copolymers (Table 3) at temperatures approaching the glass-transition temperature. The ESR spectra of irradiated aromatic polysulfones, and aromatic polymers in general, are often poorly resolved, because of the effect of delocalization of

Table 3. Radiochemical Yieldsa of Formation of Free Radicals in Aromatic Main-Chain Polymers Unit 2 PEEK 1.0 Name Radical Yield 1020 spins/kg Ref. 379

Unit 1

PEEKK

1.1

379

Kapton

1.7

371

RADIATION CHEMISTRY OF POLYMERS

Ultem

2.2

371

33

34

Table 3. (Continued) Unit 2 Bis-A PEPO 2.5 Name Radical Yield 1020 spins/kg Ref. 371

Unit 1

Bis-A PSO

2.8

371

Kevlar

5.6

371

RADIATION CHEMISTRY OF POLYMERS

Nomex

5.6

371

PPO

5.7

371

a The

yield of free radicals is expressed in the units of spins/kg after irradiation to a dose of 10 kGy, rather than a G value, since the increase in radical concentration with dose is highly nonlinear with dose in the initial stages. To obtain a single-point G value at 10 kGy, divide the radical yield by 6.24 1020 .

RADIATION CHEMISTRY OF POLYMERS

35

the electron spin across aromatic structures, and the contribution from radical anions to the spectra at low temperatures. For example, Lewis and co-workers (367) reported that up to 75% of the total area of the ESR spectra of bisphenol A polysulfone irradiated to 10 kGy arose from radical anions, which were susceptible to photobleaching with UV radiation. These relatively uninformative spectra offer little information to assign reaction mechanisms. On the other hand, Hill and coworkers (96) have deduced, from measurements of changes in molecular weights and the formation of new structures detected on irradiation using solution-state 13 C NMR, that the main cross-linking mechanism involves Y-linking. Polyimides. The two most important commercial Polyimides (qv) are Kapton and Ultem (structures shown in Table 3). Whereas Kapton is insoluble in most convenient solvents, Ultem is soluble and hence amenable to detailed study by standard methods. In 1985 Basheer and Dole (369) reported the yields of crosslinking and scission in irradiated Ultem to be respectively 0.014 and 0.005 on irradiation at room temperature. The gel dose of commercial Ultem lm is of the order of 10 MGy. Devasahayam and co-workers (370) have reported a more detailed analysis of the changes in molecular weights on irradiation, and calculated G-values for scission and cross-linking for either H- or Y-linking mechanisms. Basheer and Dole (369) reported that the ESR spectrum of irradiated Ultem was composed of contributions from radicals from scission at the isopropylidene unit and aromatic ether group. Recently, Devasahayam and co-workers (370) reported a detailed study of the free-radical yields and radical reactions in irradiated Ultem. At 77 K, the spectra contain signicant contributions from radical anions, which were susceptible to thermal or photolytic annealing. Potential scission sites were identied from the radical products. Similarly, the formation of methylene radicals through loss of hydrogen atoms from the isopropylidene unit was suggested as a precursor to cross-linking reactions. Associated NMR analysis (371) of the radiolysis products identied a large number of products, the majority of which were identied with chain-scission products. Kapton polyimide is an intractable material under usual conditions, and hence studies of its radiation chemistry are conned to measurements of volatile and free-radical products. The ESR spectrum of irradiated Kapton is a featureless singlet, and the yield of free radicals (1.7 1020 spins/kg after 10 kGy irradiation) was the one of the lowest of a large number of aromatic polymers compared in this study (372). Earlier, Hegazy and colleagues (373,374) conrmed the very high radiation stability of Kapton by measurement of the yields of volatile products of this polymer and other aromatic materials, including several aromatic polyimides. For Kapton, the evolved gases, in decreasing yields, were CO2 , CO, nitrogen, and hydrogen. Again Kapton was the most stable material studied. A number of other polyimides have been studied by Hill and co-workers Devasahayam and co-workers (375) reported changes in the optical properties, and formation of free radicals in a series of uorinated polyimides, and found that all materials were less stable than Kapton lm. The uorinated groups are incorporated in the polyimides to reduce coloration, and to provide oxygen atom resistance for space applications. Alexander and co-workers (376) have measured the radical yields for two uorinated polyimide containing phosphine oxide groups attached to the aromatic main chain. The phosphine oxide unit also provides protection for the polymer against degradation from atomic oxygen. The relatively

36

RADIATION CHEMISTRY OF POLYMERS

high yields of radicals at 77 K, G(R) = 0.50 and 0.42 indicate, however, that this protection has been gained at the expense of radiation stability. Prior to this, Hopewell and co-workers (129) measured the radical yields and the solution viscosity of a number of phosphine oxide containing polyimides, and found that for these materials chain scission and cross-linking occur simultaneously on irradiation, but that scission is predominant on irradiation at room temperature. Polyarylene Ether Ketones. The polyarylene ether ketone, known commercially as PEEK, is a semicrystalline thermoplastic of high melting temperature, and is insoluble in common organic solvent. The fully aromatic structure of PEEK imparts very high radiation resistance to this material. In a series of articles, Sasuga and co-workers (377,378) and Hegazy and co-workers (373,374,379) have measured a number of properties of PEEK and compared its radiation stability to a number of materials. The main volatile products, in decreasing yields at 15 MGy, are methane, hydrogen, CO, and CO2 . The source of the methane is not discussed. However, the overall yield of gases is very low compared to other aromatic polymers. The effect of crystallinity on the radiation chemistry is also demonstrated through the signicantly higher yields of gases obtained from irradiation of a fully amorphous polymer. In a related study, Hegazy and co-workers (379) examined the changes in the thermal properties of irradiated PEEK. An increase in glass-transition temperature and a small decrease in the heat of crystallization are both evidence of cross-linking formation. Similarly, the decrease and broadening of the recrystallization transition is due to cross-links. Heiland and co-workers (380) have examined the radiation sensitivity of a series of polyarylene ether ketones using ESR spectroscopy and gas chromatography. The yields of radicals formed in all materials at 77 K were very low (approximately 12 1020 spins/kg after 10 kGy irradiation). Introduction of methyl groups on the main-chain aromatic rings decreased the radiation stability, and further substitution of isopropylidene units into the main chain reduced the radiation stability even further. Polycarbonates. The most commercially important Polycarbonates (qv) is bisphenol A polycarbonate (Bis-A PC). The effects of radiation on this material have been studied for a number of years, with a particular interest in the identity of the species responsible for a pronounced color change in Bis-A PC after irradiation. The radiation-induced green coloration has been assigned variously to trapped electrons (381) or a combination of trapped electron and radical ions (382,383). Golden and co-workers (384386) have also in some detail examined the evolution of gases on irradiation (386), and found that the major volatile products are carbon dioxide and oxygen, with small amounts of hydrogen, methane, and benzene. Bis-A PC undergoes net chain scission during irradiation in vacuum, presumably because of the high sensitivity of the carbonate group to chain scission. Evidence for this is the decrease in limiting viscosity after irradiation (384,386,387). More recently, further insight into the radiation degradation of Bis-A PC has been gained in a study by Babanalbandi and co-workers (388). In this the authors were able to conrm that the G-value for radical formation at 77 K is 0.5 0.02, and that approximately 50% of the observed radicals are able to be photobleached with visible light, and hence are radical anions. NMR analysis of the soluble polymer after irradiation revealed new peaks, assigned to phenol-type chain end structures,

RADIATION CHEMISTRY OF POLYMERS

37

with a G-value for formation of 0.7. This observation conrmed that the main site of radiation degradation of Bis-A PC is the carbonate unit. Aromatic Polyesters. The degradation of aromatic polyester has been investigated by a number of authors (7,53,389398). The most commercially important material of this class is poly(ethylene terephthalate), a semicrystalline material of high melting temperature (T m 538 K) and moderate glass-transition temperature (T g 343 K) (see POLYESTERS, THERMOPLASTIC). On irradiation at low temperatures the observed ESR signal is primarily attributed to the secondary alkyl radicals formed by loss of a hydrogen atom from ethylene segments (53,391,394). The yield of radicals at room temperature was 0.025 (394). Since then, Choi and co-workers (132,398401) have reported the radical yields of a number of allied main-chain aromatic polyesters. They found that for series of polymers containing uorinated isopropylidene units the radiation stability was enhanced with higher content of the uorinated residues (400). The yields of free radicals at 77 K were found to fall in the range 0.380.46. Later, this same group reported the radical yields for a number of linear halogenated aromatic polyesters [G(R) = 0.410.81] (401), terephthalate-based polyesters [G(R) = 0.190.41] (398), and naphthalene-containing polyesters [G(R) = 0.080.12] (132,399). In the last study the yield of radicals was found to increase, as expected, with the length of the aliphatic chain in the polyester, with, for example, G(R) at 77 K increasing from 0.19 to 0.27 to 0.41 for terephthalate polyesters with the increasing length of the alkyl chain (n) of 2, 4, and 10, respectively (394). Limited data is available on the respective rates of cross-linking and scission of the aromatic polyesters under radiolysis. Turner (394) has collected the data available from measurements of residual soluble fractions of irradiated poly(ethylene terephthalate), and reported values of G(S) and G(X) ranging from 0.07 to 0.17 and 0.07 to 0.14, respectively. The large range reects the experimental difculties and the range of irradiation conditions employed. Babanalbandi and co-workers (388) have used ESR and NMR to determine the yields of radical intermediates and nal products in irradiated U-polymer, a polyester of bisphenol A and terephthalic acid. The radical yield G(R) = 0.5 at 77 K was consistent with previous studies of similar polyesters. The NMR spectra of polymer irradiated to very high doses at room temperature revealed that phenoltype chain ends were the main stable products. The results are consistent with the observation of CO and CO2 being the main gaseous products of the degradation of polyesters. For this material no evidence of cross-linking reactions was found. Copolymers. The properties of polymers can be enhanced extensively through Copolymerization (qv) or blending of homopolymers. The effects of radiation on copolymers and blends might be expected to show behaviors that are related to the behaviors of the homopolymer constituents, modied by the polymer mole fraction composition and the electron densities of the components. For simple polymers, eg, those composed of carbon, hydrogen, and oxygen, random copolymers show a linear relationship between the properties of the homopolymers and the weight fraction of the components, eg, for copolymers of aliphatic esters (402). Although this behavior has been found to apply for many copolymer and blend systems, major differences from linear relationships are also common. Aromatic components in mixtures of low molecular weight materials have long been known (18) to exhibit a protective effect on other components. This

38

RADIATION CHEMISTRY OF POLYMERS

protective effect is also found in random copolymers, such as the copolymers of styrene and methyl methacrylate, which have been studied by Buseld and co-workers (246), where the presence of styrene lowers the yields for radical intermediates, volatile products, and chain scission below those expected from the linear additivity rule. Similar observations have been made for radical formation in copolymers of glutamic acid and tyrosine (403), and several other systems. On the other hand, positive deviations from the additivity rule for copolymers have also been reported. For example, Fox and co-workers (404) have reported that the radical yields at 77 K for radiolysis of the copolymers of methacrylic acid and acrylonitrile exhibit a positive deviation from the rule. They have attributed this to the breakup of the hydrogen bonding between methacrylic acid units by the acrylonitrile units in the copolymer. Polymer Blends (qv) can be classied as being either compatible or incompatible. Most blends fall into the latter category. If this is the case, the additivity rule might be expected to apply, even for blends of homopolymers such as PS and PMMA, because the two components are phase separated. Garrett (357) has shown that this is indeed the case when homopolymers of comparable molecular weight are blended. However, if PS of low molecular weight is blended with PMMA of high molecular weight, a level of compatibility is achieved, and the PS then provides protection of the PMMA against degradation (357). Many other instances of deviations from the additive rule and of protective effects in copolymers and blends have been reported in some of the studies discussed elsewhere in this article. The use of aromatic groups to provide protection against degradation has been utilized in some commercial materials. One example of this has been the incorporation of aromatic groups into polyester medical sutures (405) in order to protect the sutures against damage during radiation sterilization.

The Inuence of Oxygen


Oxygen is a diradical, and so it can readily react at diffusion-controlled rates with free radicals formed in a polymer during radiolysis. This process has been well researched because it also occurs as a result of the thermal degradation of polymers, and consequently occurs during polymer processing. The mechanisms of these oxidation processes have been well documented in the literature and involve propagation, termination, and branching steps as part of a chain reaction. These mechanisms are summarized in Figure 7 (P = polymer). Methods for stabilizing polymers by the addition of antioxidants and hydroperoxide decomposers are also well known, and are generally used with polymers which undergo radiolysis. For example, so called antirads are added to polypropylene to allow medical goods manufactured from it to be radiationsterilized without discoloration or degradation. The mechanism and kinetics of the reactions of oxygen with polymer radicals produced during radiolysis have been studied and extensively reviewed by Gillen, Clough, and co-workers (117,406409), who have contributed greatly to this eld. The roles of added polymer antioxidants and stabilizers have also been reviewed

RADIATION CHEMISTRY OF POLYMERS

39

Fig. 7. A summary of the reaction pathways in the oxidative degradation of polymers.

in a treatise edited by Clough, Billingham, and Gillen (410). Therefore, a detailed review will not be presented herein, and only some recent developments will be discussed. In many polymers, radicals can have very long lifetimes, particularly if they are located in crystalline regions where they are protected from any added stabilizer. Jahan and co-workers (153), for example, have reported that long-lived radicals formed during sterilization of PE articial joint components are at least partially responsible for their long-term deterioration. In addition, Young and

40

RADIATION CHEMISTRY OF POLYMERS

Slemp (411) have found that polymers exposed to space irradiation can exhibit a slow deterioration of properties on their return to earth atmosphere. These processes are just two examples of slow polymer aging due to polymer oxidation in air. Clough and co-workers (409,412,413) have recently studied the structural changes induced in polymers and model compounds by oxidation through the use of 17 O-labeled oxygen and by the analysis of the polymer using 17 O NMR. This approach has provided a new insight of the oxidation and subsequent aging of irradiated polymers.

The Inuence of Irradiation Temperature


The chemical changes that occur when polymers are irradiated include the formation of free-radical intermediates, the scission of main-chain bonds, the loss or modication of side-chains, the cross-linking of chains, the formation of new polymer structures, and the production low molecular weight products. One of the most important factors determining the extent to which these processes can occur is the prevailing temperature. For example, the crystalline melting and the -transition temperatures, along with the temperatures of other secondary transitions in a polymer, play an important role in controlling chain motions, and hence the rates of reactions. In addition, the ceiling temperature determines the overall thermodynamic stability of a polymer chain when propagating radicals are present. The temperature dependences of the rates of radiolysis reactions are determined by their activation energies, and can be predicted through the Arrhenius equation. In polymers, the activation energies for radiolysis reactions are positive, so that the rates of these reactions show a positive dependence on temperature. However, the activation energies for radiolysis reactions can show a considerable variation depending on both chemical and physical factors. For example, the activation energies for reactions can be quite different in the crystalline and amorphous regions of a polymer. Thus, the overall rates for formation of radiolysis products depend on the morphology of the polymer, which can vary from sample to sample. In addition, the rates of radiolysis reactions are likely to exhibit discontinuities at any rst- and second-order polymer transition temperatures, which are characteristic of a polymer. Thermodynamic considerations must also be considered. Negative enthalpies and negative entropies of reaction characterize free-radical polymerization reactions. Thus, whereas the free energy change for polymerization of a vinyl monomer may be negative at a low temperature, it will become zero at a higher temperature, the ceiling temperature. Above this temperature, spontaneous depolymerization of the polymer is thus thermodynamically possible. Some representative examples of the effects of temperature drawn from reported work are presented below. Secondary Transitions. The secondary transitions in polymers characterize the onset chiey of side-chain motions and the motions of chain ends. The secondary transitions for PMMA and PS are given in Table 4. In these polymers, the transitions occur below room temperature, and are associated with the

RADIATION CHEMISTRY OF POLYMERS


Table 4. Transition Temperatures for PMMA and PS Transition T T T T PMMA, K 378 K 270 K 103 K 4K PS, K 370 K 300 K 138 K 48 K

41

large-scale motions of the side chains. The and transitions occur at much lower temperatures. At 77 K the polymer chain motions in PMMA and PS are very restricted, but as the temperature is increased to room temperature, the side chains become mobile. Thus, the polymer radicals formed in PMMA and PS on radiolysis at 77 K are relatively stable, but they decompose on annealing to higher temperatures. Garrett (357) has reported a careful study of the reactions of the radicals formed on irradiation of PMMA and PS at 77 K when they are annealed to higher temperatures. For PS a small decrease in the radical concentration was observed at approximately 140 K and a further decrease occurs at about 270 K. The temperatures at which these changes occur coincide closely with the - and -transition temperatures, respectively, for PS. The radical concentration for PMMA decreased by approximately 45% at about 190 K and is associated chiey with the loss of the anion formed on radiolysis (357). The temperature of 190 K corresponds to the temperature at which a change in the spin-lattice relaxation time has been reported to occur in s-PMMA (414), which is believed to be associated with the superposition of the motions of the -methyl group and rapid uctuations of the adjacent main chain. Wundrich and co-workers (415) has examined the variation in G(S) for PMMA following radiolysis from 77 K to room temperature. From 77 K to approximately 140 K, G(S) was found to be constant, but above 140 K the values increased with temperature. He observed a second discontinuity at approximately 268 K. His reported values of G(S) in these regions are as follows: 0.3, 77140 K; 0.30.9, 140268 K; and 0.91.15, 268293 K. Although there is no reported chain transition at 140 K, 268 K corresponds to the transition in PMMA. It has been proposed that the loss of the ester side group is the precursor for main-chain scission in PMMA, and so the lower G-values for scission below the -transition temperature should be associated with lower G-values of volatile side-chain products. Kudoh and co-workers (250) found that the yield for hydrogen for PMMA was not sensitive to the radiolysis temperature. However, the yields of the products of ester side-chain scission and decomposition, principally CO and CO2 , were depressed by a factor of about 5 to 10 times when compared with those found for room temperature radiolysis. Kudoh and co-workers (260) also found a large difference in the tensile properties consistent with less chain scission at 77 K. Transition. The transition in polymers, T g , is associated with the onset of signicant main-chain motion spread over several adjacent carbon atoms. Above T g the free volume of the polymer increases, and so the rates of radical and other reactions may increase signicantly. The effect of the glass-transition temperature

42

RADIATION CHEMISTRY OF POLYMERS

on radical reactions was clearly demonstrated by the work of Hill and co-workers (186) on polyisobutene (PIB). Annealing of PIB beyond its glass-transition temperature, 203 K, results in a rapid decrease in the radical concentration. At room temperature the radical concentration is very small and cannot be measured effectively by ESR spectroscopy. Many other polymers have been reported to behave similarly. However, Garrett and co-workers (416) found no evidence for a discontinuity in G(S) of PMMA at its transition, 378 K. They reported that the value of G(S) increases continuously with temperature between 268 and 400 K and rises smoothly from 1.3 at 303 K to 3.9 at 413 K. A similar observation has been reported by Bowmer and co-workers (360) for PS, which undergoes cross-linking on radiolysis at room temperature, G(S) = 0.0086 and G(X) = 0.043. However, as the temperature increases, G(S) increases and G(X) decreases. As a result, although no signicant discontinuity was observed in the scission and cross-linking yields, in the neighborhood of the transition, the behavior of PS changes, and it becomes a polymer which undergoes net scission on radiolysis [at 423 K G(S) = 0.074, G(X) = 0.027]. So, although the transition does not appear to have a signicant inuence on the scission and cross-linking in PMMA or PS, it does play an important role in PIB and several other polymers. For example, the temperature dependence of G(S) has been found to increase signicantly above this transition temperature for PIB (186), and for Kalrez, Forsythe and co-workers (57,383) have demonstrated that cross-linking only occurs on radiolysis above T g . Although no discontinuity has been found for G(S) at the -transition in PMMA, Dong and co-workers (31) have shown conclusively that tacticity changes do occur in i-PMMA and s-PMMA with high yield above T g . Similar changes in the tacticity of isotactic polypropylene have also been observed by Buseld and co-workers (35) for radiolysis at room temperature, which is above its -transition temperature. The G-value for the loss of isotactic triads was estimated to be 64 at 298 K. Crystalline Melting. Many of the radiolysis reactions of polymers are believed to occur principally in the amorphous regions or interphase regions at the crystalline boundaries. A strong inuence of morphology has been reported by Buseld and Hanna (38) for the radiolysis of i-PP. They found that above the crystalline melting temperature the values of G(S) and G(X) at low dose are 11.6 and 3.5, respectively, whereas below the crystalline melting temperature the values are approximately equal and fall in the range 0.10.9. They also report that in the molten state the G-values for inversion of isotactic sequences in i-PP are also much higher: G(inversion) = 106. Another recently reported difference in radiation chemistry above the crystalline melting temperature has been for polytetrauoroethylene (PTFE). In the solid state PTFE is known to undergo scission readily. However, Oshima and coworkers (56) have found that just above the melting temperature of 600 K, PTFE undergoes cross-linking on e-beam irradiation, with a concomitant enhancement in the mechanical and chemical properties of the polymer. Ceiling Temperature. For some polymers, for example PTFE, the ceiling temperature, T c , is very high (693 K), whereas for others it may be below room temperature. Polymers with quaternary carbon atoms in the main chain, such as PMMA, PIB, and poly(-methyl styrene), PMS, are particularly susceptible

RADIATION CHEMISTRY OF POLYMERS


Table 5. The Temperature Dependence for G(M) for Poly(-methyl styrene) T, K 303 373 383 393 453 G(M) 0.3 95 153 182 529

43

to radiation-induced depolymerization at elevated temperatures, because they have relatively low ceiling temperatures. For example, the ceiling temperature for PMS is only slightly above room temperature, 339 K (for the pure liquid monomer standard state). Hill and co-workers (417) have reported the G-values for loss of monomer, G(M), on radiolysis of PMS over a range of temperatures (see Table 5). Clearly, G(M) increases signicantly for radiolysis above T c where spontaneous depolymerization can occur. Property discontinuities which can be associated with temperature transitions in many polymers have been observed, and the cases discussed above serve only as a few examples. Attention has been drawn to many other cases elsewhere in the article.

Summary and Major Conclusions


The radiation chemistry of polymers is an extremely rich eld of study, both from an intellectual and a practical viewpoint. The very large body of research devoted to this eld is testament to this fact. The overall outcome of this work is a clear understanding of the effect of radiation on this class of materials, and the factors which inuence the rates of reaction and the product distribution. These include factors intrinsic to the polymers, and extrinsic factors more readily under the control of the experimental scientist or engineer. These are summarized below.

Intrinsic Factors. Aliphatic Polymers. The main reactions occurring on radiolysis of aliphatic
polymers are cross-linking, chain scission, evolution of gaseous molecules, formation of unsaturation, and reaction of unsaturated groups. Changes in physical properties of polymers on irradiation depend on the balance and overall yields of cross-linking and chain-scission reactions. For polyolens, the presence of secondary, tertiary, and quaternary carbons in the main-chain leads to a progressively higher probability of chain scission. The presence of main-chain units such as sulfone, ester, and amide groups results in a higher probability of main-chain scission. Cross-linking can proceed through H- or Y-type linking. The rate of gel formation depends on the mechanism of cross-linking. The introduction of double bonds, either as terminal double bonds, or main-chain double bonds in polydienes, results in an increased probability of cross-linking reactions. Fully uorinated thermoplastics tend to undergo chain scission reactions on irradiation at ambient temperature.

44

RADIATION CHEMISTRY OF POLYMERS

Aromatic Polymers. The introduction of aromatic groups results in significantly higher stability to radiation due to the resonant structure of the aromatic rings. Main-chain aromatic groups provide the highest stability to degradation. Morphology. In semicrystalline polymers, chemical reaction is largely conned to the disordered, amorphous regions. Chemical reactions appear to be enhanced by the presence of an interface between crystalline and amorphous phases. Extrinsic Factors. The temperature of irradiation can dramatically affect the radiation stability of a polymer. Irradiation of degrading polymers above their ceiling temperatures leads to extensive depolymerization. Unsaturated polyolens tend to undergo increasing amounts of H-linking as opposed to Y-linking as the temperature is raised. Irradiation of uoroplastics immediately above their melting temperature, or of uoroelastomers above T g , increases the probability of cross-linking reactions. Irradiation in the presence of oxygen increases the probability of long-term degradation reactions.

BIBLIOGRAPHY
Cross-linking with Radiation in EPST 1st ed., Vol. 4, pp. 398414, by Allan R. Schultz, General Electric Co.; in EPSE 2nd ed., Vol. 4, pp. 418449, by Vincent D. McGinniss, Battelle Columbus Laboratories; Radiation-Induced Reactions in EPST 1st ed., Vol. 11, pp. 702755, by A. Chapiro, Laboratoire de Chimie des Radiations du CNRS, Bellevue, France; Radiation-Resistant Polymers in EPST 1st ed., Vol. 11, pp. 783809, by W. W. Parkinson, Oak Ridge National Laboratory; in EPSE 2nd ed., Vol 13, pp. 667708, by Roger Clough, Sandia National Laboratories. 1. A. Charlesby, Atomic Radiation and Polymers, Vol. 1, Permagon Press, Oxford, 1960. 2. A. Chapiro, ed. Radiation Chemistry of Polymeric Systems, Vol. XV, Interscience, New York, 1962. 3. F. Williams, Radiat. Chem. Macromol. 1, 723 (1972). 4. J. H. ODonnell and D. F. Sangster, Principles of Radiation Chemistry, Edward Arnold, London, 1970, p. 176. 5. J. W. T. Spinks and R. J. Woods, An Introduction to Radiation Chemistry, 3rd ed., John Wiley & Sons, Inc., New York, 1990. 6. W. W. Parkinson, C. D. Bopp, D. Binder, and J. E. White, J. Phys. Chem. 69, 828833 (1965). 7. R. H. Boyett and K. Becker, J. Appl. Polym. Sci. 14, 16541656 (1970). 8. T. Seguchi, N. Hayakawa, K. Yoshida, N. Tamura, Y. Katsumura, and Y. Tabata, Radiat. Phys. Chem. 26, 221225 (1985). 9. T. Sasuga, S. Kawanishi, T. Seguchi, and I. Kohno, Polymer 30, 20542059 (1989). 10. T. Sasuga, H. Kudoh, and T. Seguchi, Polymer 40, 50955102 (1999). 11. D. J. T. Hill and J. L. Hopewell, Radiat. Phys. Chem. 48, 533537 (1996). 12. T. Sasuga, S. Kawanishi, M. Nishii, T. Seguchi, and I. Kohno, Radiat. Phys. Chem. 37, 135140 (1991). 13. S. Seki, K. Kanzaki, Y. Kunimi, S. Tagawa, Y. Yoshida, H. Kudoh, M. Sugimoto, T. Sasuga, T. Seguchi, and H. Shibata, Radiat. Phys. Chem. 50, 423427 (1997). 14. T. Seguchi, H. Kudoh, M. Sugimoto, and Y. Hama, Nucl. Instrum. Methods Phys. Res., Sect. B 151, 154160 (1999). 15. S. Seki, K. Maeda, Y. Kunimi, S. Tagawa, Y. Yoshida, H. Kudoh, M. Sugimoto, Y. Morita, T. Seguchi, T. Iwai, H. Shibata, K. Asai, and K. Ishigure, J. Phys. Chem. B 103, 30433048 (1999).

RADIATION CHEMISTRY OF POLYMERS

45

16. Y. Hama, T. Oka, K. Inoue, M. Kitoh, M. Washio, H. Kudoh, M. Sugimoto, and T. Seguchi, Res. Chem. Intermed. 27, 469473 (2001). 17. A. Oshima and M. Washio, Nucl. Instrum. Methods Phys. Res., Sect. B 208, 380384 (2003). 18. J. P. Manion and M. Burton, J. Phys. Chem. 56, 560569 (1952). 19. M. Burton and S. Lipsky, J. Phys. Chem. 61, 14611467 (1957). 20. P. Alexander and A. Charlesby, Nature (London) 173, 578579 (1954). 21. P. Alexander and A. Charlesby, Proc. R. Soc. London, Ser. A 230, 136145 (1955). 22. J. H. ODonnell and P. J. Pomery, J. Polym. Sci., Polym. Symp. 55, 269278 (1976). 23. W. K. Buseld and J. H. ODonnell, J. Polym. Sci., Polym. Symp. 49, 227237 (1975). 24. D. J. T. Hill, A. P. Lang, J. H. ODonnell, and P. J. Pomery, Polym. Degrad. Stab. 38, 205218 (1992). 25. L. Dong, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Polym. Degrad. Stab. 48, 7177 (1995). 26. A. Zeman and H. Heusinger, J. Phys. Chem. 77, 33743376 (1966). 27. R. W. Garrett, J. H. ODonnell, P. J. Pomery, and E. C. Shum, J. Appl. Polym. Sci. 24, 24152418 (1979). 28. J. Kroh and S. Karolczak, Radiat. Res. Rev. 1, 411470 (1969). 29. J. H. ODonnell, ACS Symp. Ser. 475, 402413 (1991). 30. D. J. T. Hill, S. Mohajerani, P. J. Pomery, and A. K. Whittaker, Radiat. Phys. Chem. 59, 295302 (2000). 31. L. Dong, D. J. T. Hill, J. H. ODonnell, T. G. Carswell-Pomerantz, P. J. Pomery, A. K. Whittaker, and K. Hatada, Macromolecules 28, 36813691 (1995). 32. C. David, A. Verhasselt, and G. Geuskens, Polymer 9, 287288 (1968). 33. G. Geuskens, D. Fuld, and C. David, Makromol. Chem. 160, 347348 (1972). 34. Y. Luo, H. Yang, S. Li, B. Yu, M. Ding, and B. Jiang, Radiat. Phys. Chem. 42, 233236 (1993). 35. W. K. Buseld, J. V. Hanna, J. H. ODonnell, and A. K. Whittaker, Br. Polym. J. 19, 223226 (1987). 36. P. F. Barron, W. K. Buseld, and J. V. Hanna, J. Polym. Sci., Part C: Polym. Lett. 26, 225228 (1988). 37. P. F. Barron, W. K. Buseld, and J. V. Hanna, Polym. Commun. 29, 7072 (1988). 38. W. K. Buseld and J. V. Hanna, Polym. J. (Tokyo) 23, 12531263 (1991). 39. A. Charlesby and M. Ross, Nature (London) 171, 1153 (1953). 40. A. Charlesby and N. Moore, Int. J. Appl. Radiat. Isot. 15, 703708 (1964). 41. A. Charlesby and D. K. Thomas, Proc. R. Soc. London, Ser. A 269, 104124 (1962). 42. L. F. Thompson, E. D. Feit, M. J. Bowden, P. V. Lenzo, and E. G. Spencer, J. Electrochem. Soc. 121, 15001503 (1974). 43. E. Reichmanis, O. Nalamasu, F. M. Houlihan, and A. E. Novembre, Polym. Int. 48, 10531059 (1999). 44. M. J. Bowden, Crit. Rev. Solid State Mater. Sci. 8, 223264 (1979). 45. T. N. Bowmer and J. H. ODonnell, J. Polym. Sci., Polym. Chem. Ed. 19, 4550 (1981). 46. M. Dole, C. D. Keeling, and D. G. Rose, J. Am. Chem. Soc. 76, 43044311 (1954). 47. M. Dole, Radiat. Chem. Macromol. 1, 335347 (1972). 48. R. L. Clough, J. Chem. Phys. 87, 15881595 (1987). 49. F. Cracco, A. J. Arvia, and M. Dole, J. Chem. Phys. 37, 24492457 (1962). 50. M. Dole and F. Cracco, J. Am. Chem. Soc. 83, 25842585 (1961). 51. M. Dole and F. Cracco, J. Phys. Chem. 66, 193201 (1962). 52. D. J. T. Hill, D. S. Hunter, D. A. Lewis, J. H. ODonnell, and P. J. Pomery, Radiat. Phys. Chem. 36, 559563 (1990). 53. K. Araki, D. Campbell, and D. T. Turner, J. Polym. Sci. 3, 993996 (1965).

46

RADIATION CHEMISTRY OF POLYMERS

54. R. W. Garrett, D. J. T. Hill, T. T. Le, J. H. ODonnell, and P. J. Pomery, Radiat. Phys. Chem. 39, 215221 (1992). 55. J. Sun, Y. Zhang, and X. Zhong, Polymer 35, 28812883 (1994). 56. A. Oshima, Y. Tabata, H. Kudoh, and T. Seguchi, Radiat. Phys. Chem. 45, 269273 (1995). 57. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, T. Seguchi, and A. K. Whittaker, Macromolecules 30, 81018108 (1997). 58. H. S. Judeikis, H. Hedgpeth, and S. Siegel, Radiat. Res. 35, 247262 (1968). 59. T. N. Bowmer, S. Y. Ho, and J. H. ODonnell, Polym. Degrad. Stab. 5, 449456 (1983). 60. T. N. Bowmer, S. Y. Ho, and J. H. ODonnell, G. S. Park, and M. Saleem, Eur. Polym. J. 18, 6166 (1982). 61. M. Dole, D. C. Milner, and T. F. Williams, J. Am. Chem. Soc. 80, 15801588 (1958). 62. J. C. Randall, F. J. Zoep, and J. Silverman, Radiat. Phys. Chem. 22, 183192 (1983). 63. J. C. Randall, F. J. Zoep, and J. Silverman, ACS Symp. Ser. 247, 245267 (1984). 64. J. C. Randall, F. J. Zoep, and J. Silverman, Makromol. Chem., Rapid Commun. 4, 149157 (1983). 65. F. Horii, Q. Zhu, R. Kitamaru, and H. Yamaoka, Macromolecules 23, 977981 (1990). 66. Q. Zhu, F. Horii, R. Kitamaru, and H. Yamaoka, J. Polym. Sci., Part A: Polym. Chem. 28, 27412751 (1990). 67. R. Salovey, R. V. Albarino, and J. P. Luongo, Macromolecules 3, 314318 (1970). 68. R. Salovey, J. P. Luongo, and W. A. Yager, Macromolecules 2, 198200 (1969). 69. A. Chapiro, J. Chim. Phys. 53, 895902 (1956). 70. R. V. Albarino and E. P. Otocka, J. Appl. Polym. Sci. 16, 6168 (1972). 71. M. F. Desrosiers, J. M. Puhl, and W. L. McLaughlin, Appl. Radiat. Isot. 44, 325326 (1993). 72. W. L. McLaughlin, Radiat. Phys. Chem. 41, 4556 (1993). 73. W. L. McLaughlin and M. F. Desrosiers, Radiat. Phys. Chem. 46, 11631174 (1995). 74. W. L. McLaughlin, J. Silverman, M. Al-Sheikhly, W. J. Chappas, L. Zhan-Jun, A. Miller, and W. Batsberg-Pedersen, Radiat. Phys. Chem. 56, 503508 (1999). 75. K. T. Gillen, J. S. Wallace, and R. L. Clough, Radiat. Phys. Chem. 41, 101113 (1993). 76. R. L. Clough, G. M. Malone, K. T. Gillen, J. S. Wallace, and M. B. Sinclair, Polym. Degrad. Stab. 49, 305313 (1995). 77. R. L. Clough, K. T. Gillen, G. M. Malone, and J. S. Wallace, Radiat. Phys. Chem. 48, 583594 (1996). 78. A. Charlesby, Nature (London) 173, 679680 (1954). 79. A. Charlesby, J. Polym. Sci. 14, 547553 (1954). 80. A. Charlesby, Proc. R. Soc. London, Ser. A 224, 120128 (1954). 81. M. Inokuchi and M. Dole, J. Chem. Phys. 38, 30063009 (1963). 82. N. S. Marans and L. J. Zapas, J. Appl. Polym. Sci. 11, 705718 (1967). 83. N. S. Viswanathan, J. Polym. Sci., Polym. Chem. Ed. 14, 15531555 (1976). 84. J. H. ODonnell, C. A. Smith, and D. J. Winzor, J. Polym. Sci., Polym. Phys. Ed. 16, 15151518 (1978). 85. D. J. T. Hill, J. H. ODonnell, C. L. Winzor, and D. J. Winzor, Polymer 31, 538542 (1990). 86. J. H. ODonnell, C. L. Winzor, and D. J. Winzor, Macromolecules 23, 167172 (1990). 87. W. T. Bremner, K. A. Milne, and J. ODonnell, Polym. Bull. (Berlin) 35, 525532 (1995). 88. C. L. Moad and D. J. Winzor, Prog. Polym. Sci. 23, 759813 (1998). 89. J. H. ODonnell, N. P. Rahman, C. A. Smith, and D. J. Winzor, Macromolecules 12, 113119 (1979). 90. P. Alexander, A. Charlesby, and M. Ross, Proc. R. Soc. London, Ser. A 223, 392404 (1954). 91. O. Saito, H. Y. Kang, and M. Dole, J. Chem. Phys. 46, 36073616 (1967).

RADIATION CHEMISTRY OF POLYMERS


92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133.

47

H. Y. Kang, O. Saito, and M. Dole, J. Polym. Sci., Polym. Symp. 25, 123130 (1968). O. Saito, Radiat. Chem. Macromol. 1, 223261 (1972). M. Tsuda and S. Oikawa, J. Polym. Sci., Polym. Chem. Ed. 17, 37593773 (1979). A. Charlesby and S. H. Pinner, Proc. R. Soc. London, Ser. A 249, 367386 (1959). D. J. T. Hill, D. A. Lewis, J. H. ODonnell, and A. K. Whittaker, Polym. Adv. Technol. 9, 4551 (1998). V. N. Handlos and K. A. Singer, J. Appl. Polym. Sci. 20, 33753386 (1976). M. Dole, ed. Radiat. Chem. Macromol. 1, (1972). M. Dole, ed. Radiat. Chem. Macromol. 2, (1973). J. Brandrup and E. H. Immergut, eds., Polymer Handbook, 3rd. ed., Vol. 3, John Wiley & Sons, Inc., New York, 1989, Chapt. 2. R. L. Clough and S. W. Shalaby, Irradiation of Polymers: Fundamentals and Technological Applications, Vol. 620, American Chemical Society, Washington, D.C., 1996. R. L. Clough and S. W. Shalaby, Radiation Effects on Polymers, Vol. 475, American Chemical Society, Washington, D.C., 1991. Radiat. Phys. Chem. 7, 1 (1977). Radiat. Phys. Chem. 57, 203 (2002). Radiat. Phys. Chem. 18, 1 (1981). Radiat. Phys. Chem. 22, 1 (1983). Radiat. Phys. Chem. 25, 1 (1985). Radiat. Phys. Chem. 31, 1 (1988). Radiat. Phys. Chem. 35, 1 (1990). Radiat. Phys. Chem. 42, 1 (1993). Radiat. Phys. Chem. 46, 399 (1995). Radiat. Phys. Chem. 52, 1 (1998). Radiat. Phys. Chem. 14, 1 (1979). A. Singh and J. Silverman, Radiation Processing of Polymers, Hanser, Munich, 1992. E. Reichmanis, ACS Symp. Ser. 381, 132154 (1989). R. L. Clough, Nucl. Instrum. Methods Phys. Res., Sect. B 185, 833 (2001). R. L. Clough and K. T. Gillen, and M. Dole, Irradiat. Eff. Polym. 79156 (1991). B. J. Lyons, Mod. Fluoropolym. 335347 (1997). J. S. Forsythe and D. J. T. Hill, Prog. Polym. Sci. 25, 101136 (2000). G. Burillo, R. L. Clough, T. Czvikovszky, O. Guven, A. Le Moel, W. Liu, A. Singh, J. Yang, and T. Zaharescu, Radiat. Phys. Chem. 64, 4151 (2002). J. M. Rosiak and P. Ulanski, Radiat. Phys. Chem. 55, 139151 (1999). J. M. Rosiak and F. Yoshii, Nucl. Instrum. Methods Phys. Res., Sect. B 151, 5664 (1999). L. A. Pruitt, Adv. Polym. Sci. 162, 6393 (2003). D. J. T. Hill and A. K. Whittaker, Annu. Rep. NMR Spectrosc. 46, 135 (2002). A. Charlesby, Proc. R. Soc. London 215A, 187214 (1952). A. Charlesby, Polym. J. (Tokyo) 19, 649657 (1987). H. Menhofer and H. Heusinger, Radiat. Phys. Chem. 29, 243251 (1987). D. J. T. Hill, J. H. ODonnell, P. J. Pomery, and C. L. Winzor, Radiat. Phys. Chem. 39, 237241 (1992). J. L. Hopewell, D. J. T. Hill, J. H. ODonnell, P. J. Pomery, J. E. McGrath, D. B. Priddy Jr., and C. D. Smith, Polym. Degrad. Stab. 45, 293299 (1994). A. Faucitano, A. Buttafava, V. Patruno, P. A. Guadra, and G. Marchionni, Radiat. Phys. Chem. 45, 2330 (1995). D. J. T. Hill, J. L. Hopewell, J. H. ODonnell, and P. J. Pomery, JAERI-Conf. 95003, 115119 (1995). D. J. T. Hill, B.-K. Choi, H.-K. Ahn, and E. J. Choi, Polym. Int. 48, 956962 (1999). S. Y. Ho and J. H. ODonnell, Eur. Polym. J. 20, 421424 (1984).

48

RADIATION CHEMISTRY OF POLYMERS

134. R. Basheer and M. Dole, J. Polym. Sci., Polym. Phys. Ed. 22, 13131329 (1984). 135. T. N. Bowmer, J. H. ODonnell, and P. R. Wells, Makromol. Chem., Rapid Commun. 1, 16 (1980). 136. M. G. Ormerod and A. Charlesby, Polymer 5, 6788 (1964). 137. A. Charlesby and J. Morris, J. Polym. Sci. 4, 11271133 (1964); discussion 11321123. 138. C. F. Albert, W. K. Buseld, and P. J. Pomery, Polym. Int. 27, 285291 (1992). 139. V. I. Feldman, Appl. Radiat. Isot. 47, 14971501 (1996). 140. R. Basheer and M. Dole, J. Polym. Sci., Polym. Phys. Ed. 21, 957967 (1983). 141. N. Gvozdic and M. Dole, Radiat. Phys. Chem. 15, 435440 (1980). 142. S. Shimada, M. Maeda, Y. Hori, and H. Kashiwabara, Polymer 18, 1924 (1977). 143. Y. Hori, S. Shimada, and H. Kashiwabara, Polymer 18, 151154 (1977). 144. Y. Hori, S. Shimada, and H. Kashiwabara, Polymer 18, 567572 (1977). 145. D. R. Johnson, W. Y. Wen, and M. Dole, J. Phys. Chem. 77, 21742179 (1973). 146. D. C. Waterman and M. Dole, J. Phys. Chem. 74, 19061912 (1970). 147. M. Iwasaki, T. Ichikawa, and T. Ohmori, J. Chem. Phys. 50, 19841990 (1969). 148. M. Dole, G. G. A. Boehm, and D. C. Waterman, Eur. Polym. J. 93104 (1969). 149. M. Dole and V. M. Patel, Radiat. Phys. Chem. 9, 433444 (1977). 150. H. H. Trieu, D. E. Thomas, M. Shah Jahan, W. O. Haggard, R. L. Conta, and J. E. Parr, ASTM Spec. Tech. Publ. STP 1307, 109119 (1998). 151. M. S. Jahan, D. E. Thomas, and M. D. Ridley, Mater. Sci. Forum 426432, 31393144 (2003). 152. N. Naheed, M. S. Jahan, and M. Ridley, Nucl. Instrum. Methods Phys. Res., Sect. B 208, 204209 (2003). 153. M. S. Jahan, M. C. King, W. O. Haggard, K. L. Sevo, and J. E. Parr, Radiat. Phys. Chem. 62, 141144 (2001). 154. M. C. Buncick, D. E. Thomas, K. S. McKinny, and M. S. Jahan, Appl. Surf. Sci. 156, 97109 (2000). 155. M. S. Jahan, K. S. McKinny, Nucl. Instrum. Methods Phys. Res., Sect. B 151, 207212 (1999). 156. M. S. Jahan, D. E. Thomas, K. Banerjee, H. H. Trieu, W. O. Haggard, and J. E. Parr, Radiat. Phys. Chem. 51, 593594 (1998). 157. M. S. Jahan, J. C. Stovall, J. A. Davidson, and G. Hines, Appl. Radiat. Isot. 46, 637638 (1995). 158. M. Dole and K. Katsuura, J. Polym. Sci., Part B: Polym. Lett. 3, 467472 (1965). 159. H. Y. Kang, O. Saito, and M. Dole, J. Am. Chem. Soc. 89, 19801986 (1967). 160. R. Basheer and M. Dole, J. Polym. Sci., Polym. Phys. Ed. 21, 949956 (1983). 161. B. J. Lyons and A. S. Fox, J. Polym. Sci., Polym. Symp. 21, 159170 (1968). 162. H. Mitsui and Y. Shimizu, J. Polym. Sci., Polym. Chem. Ed. 17, 28052813 (1979). 163. G. N. Patel and A. Keller, J. Polym. Sci., Polym. Phys. Ed. 13, 323331 (1975). 164. J. H. ODonnell and A. K. Whittaker, J. Macromol. Sci., Chem. 29, 19 (1991). 165. R. L. Bennett, A. Keller, J. Stejny, and M. Murray, J. Polym. Sci., Polym. Chem. Ed. 14, 30273033 (1976). 166. F. A. Bovey, F. C. Schilling, and H. N. Cheng, ACS Symp. Ser. 169, 133141 (1978). 167. J. H. ODonnell and A. K. Whittaker, Polymer 33, 6267 (1992). 168. J. H. ODonnell and A. K. Whittaker, Radiat. Phys. Chem. 39, 209214 (1992). 169. D. C. Waterman and M. Dole, J. Phys. Chem. 74, 19131922 (1970). 170. M. B. Fallgatter and M. Dole, J. Phys. Chem. 68, 19881997 (1964). 171. D. M. Bodily and M. Dole, J. Chem. Phys. 45, 14281432 (1966). 172. D. M. Bodily and M. Dole, J. Chem. Phys. 45, 14331439 (1966). 173. D. O. Geymer, Makromol. Chem. 99, 152159 (1966). 174. W. Schnabel and M. Dole, J. Phys. Chem. 67, 295299 (1963). 175. R. W. Keyser, B. Clegg, and M. Dole, J. Phys. Chem. 67, 300303 (1963).

RADIATION CHEMISTRY OF POLYMERS

49

176. M. Inokuti and M. Dole, J. Polym. Sci. 1, 32893302 (1963). 177. T. Asanuma, T. Shiomura, Y. Hirase, T. Matsuyama, H. Yamaoka, A. Tsuchida, M. Ohoka, and M. Yamamoto, Polym. Bullet. (Berlin) 29, 7983 (1992). 178. P. Alexander, R. M. Black, and A. Charlesby, Proc. R. Soc. London, Ser. A 232, 3148 (1955). 179. A. Chapiro, J. Chim. Phys. Phys.-Chim. Biol. 53, 306307 (1956). 180. A. A. Miller, E. J. Lawton, and J. S. Balwit, J. Polym. Sci. 14, 503504 (1954). 181. Z. F. Ilicheva and N. A. Slovokhotova, Khim. Vysokikh Energii 3, 272273 (1969). 182. B. Ranby and P. Carstensen, Adv. Chem. Ser. 66, 256270 (1967). 183. Y. Hori and H. Kashiwabara, J. Polym. Sci., Polym. Phys. Ed. 19, 11411149 (1981). 184. J. Bartos, Collect. Czech. Chem. Commun. 50, 16991713 (1985). 185. T. Bremner, D. J. T. Hill, J. H. ODonnell, M. C. S. Perera, and P. J. Pomery, J. Polym. Sci., Part A: Polym. Chem. 34, 971984 (1996). 186. D. J. T. Hill, J. H. ODonnell, M. C. S. Perera, and P. J. Pomery, ACS Symp. Ser. 620, 139150 (1996). 187. H. Fischer and W. Langbein, Kolloid Z. Z. Polym. 216/217, 329336 (1967). 188. H. Savas and O. Guven, Radiat. Phys. Chem. 64, 3540 (2002). 189. R. A. Van Brederode, F. Rodriguez, and G. G. Cocks, J. Appl. Polym. Sci. 12, 20972104 (1968). 190. U. Borgwardt, W. Schnabel, and A. Henglein, Makromol. Chem. 127, 176184 (1969). 191. W. Schnabel and U. Borgwardt, Makromol. Chem. 123, 7379 (1969). 192. P. A. King and J. A. Ward, J. Polym. Sci., Polym. Chem. Ed. 8, 253262 (1970). 193. J. W. Stafford, Makromol. Chem. 134, 99111 (1970). 194. J. W. Stafford, Makromol. Chem. 134, 8797 (1970). 195. J. W. Stafford, Makromol. Chem. 134, 7186 (1970). 196. J. W. Stafford, Makromol. Chem. 134, 5769 (1970). 197. J. W. Stafford, Makromol. Chem. 134, 113119 (1970). 198. R. A. Van Brederode and F. Rodriguez, J. Appl. Polym. Sci. 14, 979987 (1970). 199. J. A. Ward, J. Polym. Sci., Polym. Chem. Ed. 9, 35553561 (1971). 200. A. Charlesby and M. Byrne, Radiat. Phys. Chem. 12, 129131 (1978). 201. P. Konas, V. Athanassiou, and E. W. Merrill, Biomaterials 17, 15471550 (1996). 202. L. Zhang, W. Zhang, Z. Zhang, L. Yu, H. Zhang, Y. Qi, and D. Chen, Radiat. Phys. Chem. 40, 501505 (1992). 203. P. A. King, Adv. Chem. Ser. 66, 113126 (1967). 204. E. Nedkov and S. Tsvetkova, Radiat. Phys. Chem. 44, 251256 (1994). 205. E. Nedkov and S. Tsvetkova, Radiat. Phys. Chem. 43, 397401 (1994). 206. F. C. Schilling, A. E. Tonelli, and A. L. Cholli, J. Polym. Sci., Part B: Polym. Phys. 30, 9196 (1992). 207. G. P. Roberts, M. Budzol, and M. Dole, J. Polym. Sci., Polym. Phys. Ed. 9, 17291745 (1971). 208. W. Zhao, Y. Yamamoto, and S. Tagawa, J. Polym. Sci., Part A: Polym. Chem. 36, 3089 3095 (1998). 209. J. Pacansky and S. Schneider, J. Phys. Chem. 94, 31663179 (1990). 210. G. Burillo, G. Canizal, and T. Ogawa, Radiat. Phys. Chem. 33, 351354 (1989). 211. Zainuddin, D. J. T. Hill, and T. T. Le, Radiat. Phys. Chem. 62, 283291 (2001). 212. G. Lu, H. Chen, and D. Liu, Radiat. Phys. Chem. 42, 229232 (1993). 213. P. L. Walker Jr. and D. E. Kline, Carbon 6, 739741 (1968). 214. R. Salovey and R. C. Gebauer, J. Polym. Sci., Polym. Chem. Ed. 10, 15331537 (1972). 215. A. Babanalbandil and D. J. Hill, Polym. Int. 48, 963970 (1999). 216. T. Carswell-Pomerantz, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 35, 913, 918 (1994).

50

RADIATION CHEMISTRY OF POLYMERS

217. T. Carswell-Pomerantz, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Radiat. Phys. Chem. 45, 737744 (1995) 218. T. Carswell-Pomerantz, L. Dong, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, ACS Symp. Ser. 620, 1127 (1996) 219. S. Luo and A. N. Netravali, J. Appl. Polym. Sci. 73, 10591067 (1999). 220. A. Babanalbandi, D. J. T. Hill, J. H. ODonnell, P. J. Pomery, and A. Whittaker, Polym. Degrad. Stab. 50, 297304 (1995). 221. A. Babanalbandi, D. J. T. Hill, and A. K. Whittaker, Polym. Degrad. Stab. 58, 203214 (1997). 222. A. Babanalbandi, D. J. T. Hill, D. S. Hunter, and L. Kettle, Polym. Int. 48, 980984 (1999). 223. J. H. Collett, L. Y. Lim, and P. L. Gould, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 30, 468469 (1989). 224. O. Wollersheim, H. Zumaque, J. Hormes, D. Kadereit, J. Langen, L. Haeussling, P. Hoessel, and G. Hoffmann, Nucl. Instrum. Methods Phys. Res., Sect. B 97, 273278 (1995). 225. O. Wollersheim and J. Hormes, Chem. Phys. 204, 129134 (1996). 226. L. Montanari, M. Costantini, E. C. Signoretti, L. Valvo, M. Santucci, M. Bartolomei, P. Fattibene, S. Onori, A. Faucitano, B. Conti, and I. Genta, J. Controlled Release 56, 219229 (1998). 227. P. B. Ayscough, K. J. Ivin, and J. H. ODonnell, Trans. Faraday Soc. 61, 11101117 (1965). 228. J. R. Brown and J. H. ODonnell, Macromolecules 3, 265267 (1970). 229. J. R. Brown and J. H. ODonnell, Macromolecules 5, 109114 (1972). 230. T. N. Bowmer, J. H. ODonnell, and P. R. Wells, Polym. Bull. (Berlin) 2, 103110 (1980). 231. L. A. Wall and D. W. Brown, J. Phys. Chem. 61, 129136 (1957). 232. A. R. Shultz, P. I. Roth, and J. M. Berge, J. Polym. Sci., Part A: Gen. Pap. 1, 16511669 (1963). 233. D. W. Ovenall, Nature (London) 184, 181182 (1959). 234. D. W. Ovenall, J. Polym. Sci. 41, 199211 (1959). 235. D. W. Ovenall, J. Polym. Sci. 39, 2127 (1959). 236. R. J. Abraham, H. W. Melville, D. W. Ovenall, and D. H. Whiffen, Trans. Faraday Soc. 54, 11331139 (1958). 237. J. H. ODonnell, B. McGarvey, and H. Morawetz, J. Am. Chem. Soc. 86, 23222325 (1964). 238. M. J. Bowden and J. H. ODonnell, J. Phys. Chem. 72, 15771582 (1968). 239. C. David, D. Fuld, G. Geuskens, and A. Charlesby, Eur. Polym. J. 5, 641648 (1969). 240. A. R. Shultz and F. A. Bovey, J. Polym. Sci. 22, 485493 (1956). 241. R. K. Graham, J. Polym. Sci. 37, 441444 (1959). 242. R. K. Graham, J. Polym. Sci. 38, 209212 (1959). 243. R. B. Fox, L. G. Isaacs, S. Stokes, and R. E. Kagarise, J. Polym. Sci. 20852092 (1964). 244. R. B. Fox, L. G. Isaacs, R. E. Kagarise, and S. Stokes, Am. Chem. Soc., Div. Org. Coatings, Plastic Chem., Prepr. 22, 205210 (1962). 245. M. J. Bowden, J. Polym. Sci., Polym. Symp. 49, 221226 (1975). 246. W. K. Buseld, J. H. ODonnell, and C. A. Smith, Polymer 23, 431434 (1982). 247. W. K. Buseld, J. H. ODonnell, and C. A. Smith, J. Macromol. Sci., Chem. A17, 1263 1272 (1982). 248. T. G. Carswell, R. W. Garrett, D. J. T. Hill, J. H. ODonnell, P. J. Pomery, and C. L. Winzor, Polym. Spectrosc. 253274 (1996). 249. D. J. T. Hill, Radiat. Chem. (Japan) 59, 2027 (1995). 250. H. Kudoh, N. Kasai, T. Sasuga, and T. Seguchi, Radiat. Phys. Chem. 48, 95100 (1996).

RADIATION CHEMISTRY OF POLYMERS

51

251. H. Kudoh, N. Kasai, T. Sasuga, and T. Seguchi, Radiat. Phys. Chem. 43, 329334 (1994). 252. F. Thominette and J. Verdu, J. Polym. Sci., Part A: Polym. Chem. 34, 32213223 (1996). 253. H. J. Glasel, E. Hartmann, R. Mehnert, D. Hirsch, R. Bottcher, and J. Hormes, Nucl. Instrum. Methods Phys. Res., Sect. B 151, 200206 (1999). 254. N. W. Isaacs, C. H. L. Kennard, and J. H. ODonnell, Nature (London) 216, 1104 (1967). 255. R. P. Kusy, M. J. Katz, and D. T. Turner, Thermochim. Acta 26, 415425 (1978). 256. R. Jenkins and R. S. Porter, J. Polym. Sci., Polym. Lett. Ed. 18, 743750 (1980). 257. D. J. T. Hill and J. H. ODonnell, J. Chem. Educ. 58, 174176 (1981). 258. Y. Katsumura, Y. Tabata, T. Seguchi, N. Hayakawa, K. Yoshida, and N. Tamura, Radiat. Phys. Chem. 26, 211220 (1985). 259. H. Kudoh, M. Celina, M. Malone, R. J. Kaye, K. T. Gillen, and R. L. Clough, Radiat. Phys. Chem. 48, 555562 (1996). 260. H. Kudoh, N. Kasai, T. Sasuga, and T. Seguchi, ACS Symp. Ser. 620, 313322 (1996). 261. O. Schmalz, M. Hess, and R. Kosfeld, Angew. Makromol. Chem. 239, 7991 (1996). 262. N. M. Bolbit, V. B. Taraban, E. R. Klinshpont, I. P. Shelukhov, and V. K. Milinchuk, High Energ. Chem. 34, 229235 (2000). 263. A. A. Koptelov and S. V. Karyazov, Dokl. Chem. 384, 135139 (2002). 264. F. M. Veronese, G. Ceriotti, G. Keller, S. Lora, and M. Carenza, Radiat. Phys. Chem. 35, 8892 (1990). 265. D. J. T. Hill, J. H. ODonnell, P. J. Pomery, and G. Saadat, Radiat. Phys. Chem. 48, 605612 (1996). 266. L. Dong, D. J. T. Hill, J. H. ODonnel, P. J. Pomery, and I. M. Brereton, Macromol. Chem. Phys. 196, 33793390 (1995). 267. L. Dong, D. J. T. Hill, J. H. ODonnell, P. J. Pomery, and K. Hatada, J. Appl. Polym. Sci. 59, 589597 (1996). 268. L. Brown, T. G. Carswell-Pomerantz, J. T. Hill, and P. J. Pomery, Radiat. Phys. Chem. 48, 627630 (1996). 269. H. Kudoh, T. Sasuga, and T. Seguchi, Radiat. Phys. Chem. 48, 545548 (1996). 270. H. Kudoh, T. Sasuga, T. Seguchi, and Y. Katsumura, Polymer 37, 46634665 (1996). 271. H. Kudoh, T. Sasuga, T. Seguchi, and Y. Katsumura, Polymer 37, 29032908 (1996). 272. H. Kudoh, T. Sasuga, and T. Seguchi, ACS Symp. Ser. 620, 210 (1996). 273. H. Kudoh, T. Sasuga, and T. Seguchi, Radiat. Phys. Chem. 50, 299302 (1997). 274. J. H. ODonnell, N. P. Rahman, and D. J. Winzor, J. Polym. Sci., Polym. Chem. Ed. 15, 131139 (1977) 275. J. H. ODonnell, P. J. Pomery, N. P. Rahman, C. A. Smith, and D. J. Winzor, J. Appl. Polym. Sci., Appl. Polym. Symp. 35, 527536 (1979). 276. I. S. Ungar, J. F. Kircher, W. B. Gager, F. A. Sliemers, and R. I. Leininger, J. Polym. Sci., Part A: Polym. Chem. 1, 277288 (1963). 277. A. Charlesby, Proc. R. Soc. London, Ser. A 230, 120135 (1955). 278. A. Charlesby, J. Polym. Sci. 17, 379390 (1955). 279. L. E. St. Pierre, H. A. Dewhurst, and A. M. Bueche, J. Polym. Sci. 36, 105111 (1959). 280. A. A. Miller, J. Am. Chem. Soc. 82, 35193523 (1960). 281. A. M. Bueche, J. Polym. Sci. 19, 297306 (1956). 282. S. Okamura, H. Inagaki, and K. Ohdan, Doitai Hoshasen 1, 214216 (1958). 283. M. G. Ormerod and A. Charlesby, Polymer 4, 459470 (1963). 284. C. G. Delides, Radiat. Phys. Chem. 16, 345352 (1980). 285. A. Charlesby and P. G. Garratt, Proc. R. Soc. London, Ser. A 273, 117132 (1963). 286. D. J. T. Hill, C. M. L. Preston, A. K. Whittaker, and S. M. Hunt, Macromol. Symp. 156, 95102 (2000). 287. D. J. T. Hill, C. M. L. Preston, and A. K. Whittaker, Polymer 43, 10511059 (2001).

52

RADIATION CHEMISTRY OF POLYMERS

288. D. J. T. Hill, C. M. L. Preston, D. J. Salisbury, and A. K. Whittaker, Radiat. Phys. Chem. 62, 1117 (2001). 289. M. Dole, Radiat. Chem. Macromol. 2, 167177 (1973). 290. M. I. Bro, E. R. Lovejoy, and G. R. McKay, J. Appl. Polym. Sci. 7, 21212133 (1963). 291. P. Hedvig, J. Polym. Sci., Polym. Chem. Ed. 7, 11451152 (1969). 292. H. N. Rexroad and W. Gordy, J. Chem. Phys. 30, 399403 (1959). 293. D. W. Ovenall, J. Chem. Phys. 38, 24482454 (1963). 294. E. R. Lovejoy, M. I. Bro, and G. H. Bowers, J. Appl. Polym. Sci. 9, 401410 (1965). 295. M. Tutiya, Jpn. J. Appl. Phys. 11, 15421546 (1972). 296. M. Tutiya, Jpn. J. Appl. Phys. 11, 1219 (1972). 297. M. Tutiya, Jpn. J. Appl. Phys. 9, 12041209 (1970). 298. M. Tutiya, J. Phys. Soc. Japan 25, 1518 (1968). 299. M. Tutiya, Polym. J. (Tokyo) 6, 3944 (1974). 300. J. Sun, Y. Zhang, X. Zhong, and W. Zhang, Radiat. Phys. Chem. 42, 139142 (1993). 301. Y. Ito, H. F. M. Mohamed, T. Seguchi, and A. Oshima, Radiat. Phys. Chem. 48, 775779 (1996). 302. Y. Tabata, A. Oshima, K. Takashika, and T. Seguchi, Radiat. Phys. Chem. 48, 563568 (1996). 303. A. Oshima, S. Ikeda, H. Kudoh, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 50, 611615 (1997). 304. A. Oshima, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 50, 601606 (1997). 305. A. Oshima, S. Ikeda, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 49, 581588 (1997). 306. A. Oshima, S. Ikeda, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 49, 279284 (1997). 307. E. Katoh, H. Sugisawa, A. Oshima, Y. Tabata, T. Seguchi, and T. Yamazaki, Radiat. Phys. Chem. 54, 165171 (1999). 308. A. Oshima, T. Seguchi, and Y. Tabata, Polym. Int. 48, 9961003 (1999). 309. A. Oshima, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 55, 6171 (1999). 310. A. Oshima, S. Ikeda, E. Katoh, and Y. Tabata, Radiat. Phys. Chem. 62, 3945 (2001). 311. B. Fuchs, U. Lappan, K. Lunkwitz, and U. Scheler, Macromolecules 35, 90799082 (2002). 312. B. Fuchs and U. Scheler, Macromolecules 33, 120124 (2000). 313. G. H. Bowers and E. R. Lovejoy, Ind. Eng. Chem. Prod. Res. Dev. 1, 8992 (1962). 314. D. J. T. Hill, S. Mohajerani, A. K. Whittaker, and U. Scheler, Polym. Int. 52, 17251733 (2003). 315. K. K. Gleason, D. J. T. Hill, K. K. S. Lau, S. Mohajerani, and A. K. Whittaker, Nucl. Instrum. Methods Phys. Res., Sect. B 185, 8387 (2001). 316. J. S. Forsythe, D. J. T. Hill, S. Mohajerani, and A. K. Whittaker, Radiat. Phys. Chem. 60, 439444 (2001). 317. U. Lappan, U. Geissler, U. Scheler, and K. Lunkwitz, Radiat. Phys. Chem. 67, 447451 (2003). 318. T. R. Dargaville, D. J. T. Hill, and A. K. Whittaker, Radiat. Phys. Chem. 62, 2531 (2001). 319. T. R. Dargaville, G. A. George, D. J. T. Hill, U. Scheler, and A. K. Whittaker, Macromolecules 35, 55445549 (2002). 320. T. R. Dargaville, G. A. George, D. J. T. Hill, U. Scheler, and A. K. Whittaker, Macromolecules 36, 71387142 (2003). 321. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, and A. K. Whittaker, Polym. Degrad. Stab. 63, 95101 (1998). 322. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, P. J. Pomery, and A. K. Whittaker, J. Appl. Polym. Sci. 75, 14471452 (2000).

RADIATION CHEMISTRY OF POLYMERS

53

323. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, P. J. Pomery, and A. K. Whittaker, J. Appl. Polym. Sci. 73, 169175 (1999). 324. J. S. Forsythe, D. J. T. Hill, A. K. Whittaker, and A. L. Logothetis, Polym. Int. 48, 10041009 (1999). 325. J. S. Forsythe, D. J. T. Hill, N. Calos, A. L. Logothetis, and A. K. Whittaker, J. Appl. Polym. Sci. 73, 807812 (1999). 326. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, T. Seguchi, and A. K. Whittaker, Radiat. Phys. Chem. 53, 611621 (1998). 327. J. S. Forsythe, D. J. T. Hill, A. L. Logothetis, T. Seguchi, and A. K. Whittaker, Radiat. Phys. Chem. 53, 657667 (1998). 328. L. A. Wall, S. Straus, and R. E. Florin, J. Polym. Sci., Polym. Phys. Ed. A 14, 349365 (1966). 329. R. Timmerman and W. Greyson, J. Appl. Polym. Sci. 6, 456460 (1962). 330. T. Yoshida, R. E. Florin, and L. A. Wall, J. Polym. Sci., Part A: Gen. Pap. 3, 16851712 (1965). 331. T. Seguchi, K. Makuuchi, T. Suwa, N. Tamura, T. Abe, and M. Takehisa, Nippon Kagaku Kaishi 13091314 (1974). 332. A. Oshima, S. Ikeda, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 50, 519522 (1997). 333. Y. X. Luo, F. C. Pang, and J. X. Sun, Radiat. Phys. Chem. 18, 445457 (1981). 334. D. J. T. Hill, K. J. Thurecht, and A. K. Whittaker, Radiat. Phys. Chem. 67, 729736 (2003). 335. D. J. T. Hill, K. J. Thurecht, and A. K. Whittaker, Radiat. Phys. Chem. 68, 857864 (2003). 336. V. T. Kozlov and Z. N. Tarasova, Vysokomol. Soedin. 8, 943948 (1966). 337. V. T. Kozlov, Vysokomol. Soedin., Ser. A 9, 515521 (1967). 338. A. S. Kuzminskii, T. S. Nikitina, E. V. Zhuravskaya, L. A. Oksentevich, L. L. Sunitsa, and N. I. Vitushkin, Proc. 2nd Int. Conf. U.N. Peaceful Uses At. Energy, 1958 29, 258 265 (1959). 339. E. Witt, J. Polym. Sci. 41, 507518 (1959). 340. B. Jankowski and J. Kroh, J. Appl. Polym. Sci. 9, 13631366 (1965). 341. V. T. Kozlov, A. G. Evseev, and P. I. Zubov, Vysokomol. Soedin., Ser. A 11, 22302237 (1969). 342. V. T. Kozlov, M. V. Gurev, A. G. Evseev, N. G. Kashevskaya, and P. I. Zubov, Vysokomol. Soedin., Ser. A 12, 592601 (1970). 343. D. S. Pearson, B. J. Skutnik, and G. G. A. Boehm, J. Polym. Sci., Polym. Phys. Ed. 12, 925939 (1974). 344. J. H. ODonnell and A. K. Whittaker, J. Polym. Sci., Part A: Polym. Chem. 30, 185195 (1992). 345. P. F. Barron, J. H. ODonnell, and A. K. Whittaker, Polym. Bull. (Berlin) 14, 339346 (1985). 346. W. W. Parkinson and W. C. Sears, Adv. Chem. Ser. 66, 5770 (1967). 347. M. A. Golub and J. Danon, Can. J. Chem. 43, 27722785 (1965). 348. M. A. Golub, J. Phys. Chem. 69, 26392645 (1965). 349. A. K. Whittaker, Department of Chemistry, University of Queensland, Brisbane, 1986. 350. G. G. A. Boehm, Radiat. Chem. Macromol. 2, 195260 (1973). 351. D. J. T. Hill, J. H. ODonnell, M. C. Senake Perera, and P. J. Pomery, ACS Symp. Ser. 527, 7494 (1993) 352. A. Charlesby, J. Polym. Sci. 11, 521529 (1953). 353. W. W. Parkinson and R. M. Keyser, Radiat. Chem. Macromol. 2, 5796 (1973). 354. R. Basheer and M. Dole, Int. J. Chem. Kinet. 13, 11431150 (1981). 355. R. Basheer and M. Dole, Radiat. Phys. Chem. 18, 10531060 (1981).

54

RADIATION CHEMISTRY OF POLYMERS

356. H. Moenig, Makromol. Chem. 181, 22932303 (1980). 357. R. W. Garrett, Department of Chemistry, University of Queensland, Brisbane, 1984. 358. J. H. ODonnell, N. P. Rahman, C. A. Smith, and D. J. Winzor, J. Polym. Sci., Polym. Chem. Ed. 17, 40814088 (1979). 359. J. G. Spiro and C. A. Winkler, J. Appl. Polym. Sci. 8, 17091717 (1964). 360. T. N. Bowmer, J. H. ODonnell, and D. J. Winzor, J. Polym. Sci., Polym. Chem. Ed. 19, 11671174 (1981). 361. Y. Tabata and A. Oshima, Macromol. Symp. 143, 337358 (1999). 362. K. Takashika, A. Oshima, M. Kuramoto, T. Seguchi, and Y. Tabata, Radiat. Phys. Chem. 55, 399408 (1999). 363. W. Burlant, J. Neerman, and V. Serment, J. Polym. Sci. 58, 491 (1962). 364. J. G. Kilroe and K. E. Wearle, J. Chem. Soc., Abstr. 38493854 (1960). 365. R. W. Garrett, T. Le, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Polym. Microelectron. Proc. Int. Symp. 4762 (1990). 366. J. R. Brown, J. H. ODonnell, J. Polym. Sci., Part B: Polym. Phys. 8, 121126 (1970) 367. D. A. Lewis, J. H. ODonnell, J. L. Hedrick, T. C. Ward, and J. E. McGrath, ACS Symp. Ser. 381, 252261 (1989). 368. D. J. T. Hill, D. A. Lewis, and J. H. ODonnell, J. Macromol. Sci., Chem. 29, 1118 (1991). 369. R. Basheer and M. Dole, Radiat. Phys. Chem. 25, 389398 (1985). 370. S. Devasahayam, D. J. T. Hill, P. J. Pomery, and A. K. Whittaker, Radiat. Phys. Chem. 64, 299308 (2002). 371. I. Brereton, S. Devasahayam, D. J. T. Hill, and A. K. Whittaker, Radiat. Phys. Chem. 69, 6577 (2004). 372. K. Heiland, D. J. T. Hill, J. L. Hopewell, D. A. Lewis, J. H. ODonnell, P. J. Pomery, and A. K. Whittaker, Adv. Chem. Ser. 249, 637649 (1996). 373. E. S. A. Hegazy, T. Sasuga, M. Nishii, and T. Seguchi, Polymer 33, 28972903 (1992). 374. E. S. A. Hegazy, T. Sasuga, M. Nishii, and T. Seguchi, Polymer 33, 29042910 (1992). 375. S. Devasahayam, D. J. T. Hill, and J. W. Connell, Radiat. Phys. Chem. 62, 189194 (2001). 376. M. A. Alexander, D. J. T. Hill, J. W. Connell, and K. A. Watson, High Perf. Polym. 13, 6778 (2001). 377. T. Sasuga and M. Hagiwara, Polymer 28, 19151921 (1987). 378. T. Sasuga, T. Seguchi, H. Sakai, T. Nakakura, and M. Masutani, J. Mater. Sci. 24, 15701574 (1989). 379. E. S. A. Hegazy, T. Sasuga, and T. Seguchi, Polymer 33, 29112914 (1992). 380. K. Heiland, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Polym. Adv. Technol. 5, 116121 (1994). 381. A. Davis and J. H. Golden, J. Macromol. Sci., Rev. Macromol. Chem. 3, 4968 (1969). 382. Y. Hama and K. Shinohara, J. Polym. Sci., Polymer Chem. Ed. 8, 651663 (1970). 383. A. Torikai, T. Murata, and K. Fueki, Polym. Degrad. Stab. 7, 5564 (1984). 384. J. H. Golden, B. L. Hammant, and E. A. Hazell, J. Polym. Sci. 2, 47874794 (1964). 385. J. H. Golden, Makromol. Chem. 66, 7381 (1963). 386. J. H. Golden and E. A. Hazell, J. Polym. Sci., Part A: Gen. Pap. 1, 16711686 (1963). 387. R. C. Giberson, Mod. Plast. 39, 143, 146, 228 (1962). 388. A. Babanalbandi, D. J. T. Hill, and A. K. Whittaker, Polym. Adv. Technol. 9, 6274 (1998). 389. S. P. Burow, D. T. Turner, G. F. Pezdirtz, and G. D. Sands, Am. Chem. Soc., Div. Polymer Chem., Prepr. 5, 396402 (1964). 390. S. D. Burow, D. T. Turner, G. F. Pezdirtz, and G. D. Sands, J. Polym. Sci., Polymer Chem. Ed. 4, 613622 (1966).

RADIATION CHEMISTRY OF POLYMERS

55

391. D. Campbell, K. Araki, and D. T. Turner, J. Polym. Sci., Polym. Chem. Ed. 4, 25972606 (1966). 392. D. T. Turner, G. F. Pezdirtz, and G. D. Sands, J. Polym. Sci., Polym. Chem. Ed. 4, 252254 (1966). 393. D. T. Turner and R. P. Kusy, Macromolecules 4, 337341 (1971). 394. D. T. Turner, Radiat. Chem. Macromol. 2, 137166 (1973). 395. V. Stannett, T. Memetea, M. Dole, and J. Salik, J. Polym. Sci., Polym. Lett. Ed. 16, 6365 (1978). 396. Y. Komaki and T. Seguchi, Polymer 23, 11431146 (1982). 397. V. L. Bell and G. F. Pezdirtz, J. Polym. Sci., Polym. Chem. Ed. 21, 30833092 (1983). 398. E. J. Choi, D. J. T. Hill, K. Y. Kim, J. H. ODonnell, P. J. Pomery, and A. K. Whittaker, Polym. Int. 48, 971979 (1999) 399. D. J. T. Hill, B. K. Choi, H. K. Ahn, and E. J. Choi, Radiat. Phys. Chem. 62, 195201 (2001). 400. E. J. Choi, D. J. T. Hill, K. Y. Kim, J. H. ODonnell, and P. J. Pomery, Polymer 38, 36693676 (1997). 401. E. J. Choi, D. J. T. Hill, K. Y. Kim, J. H. ODonnell, and P. J. Pomery, Polym. Adv. Technol. 9, 5261 (1998) 402. A. Babanalbandi, D. J. T. Hill, J. H. ODonnell, and P. J. Pomery, Polym. Degrad. Stab. 52, 5966 (1996). 403. D. J. T. Hill, S. Y. Ho, J. H. ODonnell, and P. J. Pomery, Radiat. Phys. Chem. 36, 467474 (1990) 404. P. A. Fox, D. J. T. Hill, A. P. Lang, and P. J. Pomery, Polym. Int. 52, 17191724 (2003). 405. D. D. Jamiolkowski, R. S. Bezwada, and S. W. Shalaby, ACS Symp. Ser. 527, 320325 (1993). 406. K. T. Gillen and R. L. Clough, Irradiat. Eff. Polym. 157223 (1991). 407. M. Celina, K. T. Gillen, J. Wise, and R. L. Clough, Radiat. Phys. Chem. 48, 613626 (1996). 408. M. Celina, K. T. Gillen, G. M. Malone, R. L. Clough, and W. H. Nelson, Radiat. Phys. Chem. 62, 153161 (2001). 409. R. A. Assink, M. Celina, T. D. Dunbar, T. M. Alam, R. L. Clough, and K. T. Gillen, Macromolecules 33, 40234029 (2000). 410. R. L. Clough, N. C. Billingham, and K. T. Gillen, Adv. Chem. Ser. 249 (1996). 411. P. R. Young and W. S. Slemp, ACS Symp. Ser. 527, 278304 (1993). 412. T. M. Alam, Radiat. Phys. Chem. 62, 145152 (2001). 413. T. M. Alam, M. Celina, R. A. Assink, R. L. Clough, and K. T. Gillen, Radiat. Phys. Chem. 60, 121127 (2001). 414. B. Gabrys, F. Horii, and R. Kitamaru, Macromolecules 20, 175177 (1987). 415. K. Wundrich, in International Symposium on Radiation Degradation of Polymers and Radiation Resistant Materials, JAERI, Takasaki, Japan, 1989, p. 159. 416. R. W. Garrett, D. J. T. Hill, T. T. Le, K. A. Milne, J. H. ODonnell, S. M. C. Perera, and P. J. Pomery, ACS Symp. Ser. 475, 146155 (1991). 417. D. J. T. Hill, T. T. Le, J. H. ODonnell, M. C. Senake Perera, and P. J. Pomery, ACS Symp. Ser. 527, 5061 (1993).

DAVID J. T. HILL ANDREW K. WHITTAKER The University of Queensland

You might also like