You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/226922098

Gamma Irradiation of Silicones

Article in Journal of Inorganic and Organometallic Polymers and Materials · January 2008
DOI: 10.1007/s10904-008-9205-0

CITATIONS READS

69 11,135

3 authors, including:

Stephen J. Clarson
University of Cincinnati
210 PUBLICATIONS 5,708 CITATIONS

SEE PROFILE

All content following this page was uploaded by Stephen J. Clarson on 09 March 2016.

The user has requested enhancement of the downloaded file.


J Inorg Organomet Polym (2008) 18:207–221
DOI 10.1007/s10904-008-9205-0

REVIEW

Gamma Irradiation of Silicones


Aniruddha S. Palsule Æ Stephen J. Clarson Æ
Christopher W. Widenhouse

Accepted: 5 February 2008 / Published online: 1 April 2008


Ó Springer Science+Business Media, LLC 2008

Aniruddha S. Palsule completed his Bachelor of Engineering degree


from the University of Pune, India in the field of Polymer Engineering in
2004. He commenced his doctoral thesis work with Dr. Steve Clarson at
the University of Cincinnati in 2004. His research work has focused on the
evaluation of permeation and biocompatibility of biomedical grade sili-
cone elastomers. This work has been supported by the NSF Membrane
Applied Science and Technology (MAST) center. He is also working on
the development of enzyme based ‘green’ catalytic routes to silicone
polymers.

Stephen J. Clarson obtained his doctoral degree in Chemistry at the Uni-


versity of York, England in 1985. He then spent the summer of 1985 at the
Institute of Macromolecular Chemistry in Prague, Czechoslovakia before tak-
ing up a postdoctoral appointment with James Mark. In 1988 he joined the
faculty of the Department of Materials Science and Engineering at the Uni-
versity of Cincinnati. He is currently the Director of the NSF I/UCRC
Membrane Science and Technology Center. He is the North American Editor of
European Polymer Journal. In January, 2000, he was elected Fellow of the Royal
Society of Chemistry (FRSC). Dr. Clarson has published over 200 technical
articles and co-authored four books. The research carried out in his group has led
to several inventions and he holds five US patents and two Japanese patents. His
current scientific research interests include materials chemistry, polymer syn-
thesis, silicon-based chemistry, biomaterials, biomineralization, biocatalysis,
optoelectronic materials, thin films and surface science.

A. S. Palsule
Department of Chemical and Materials Engineering,
University of Cincinnati, 400 Rhodes Hall, ML 0012,
Cincinnati, OH 45221, USA

S. J. Clarson (&)
Department of Chemical and Materials Engineering,
University of Cincinnati, 601B ERC, ML 0012, Cincinnati,
OH 45221, USA
e-mail: Stephen.Clarson@UC.Edu

C. W. Widenhouse
Ethicon Endo-Surgery, 4545 Creek Road, Cincinnati,
OH 45242-2839, USA

123
208 J Inorg Organomet Polym (2008) 18:207–221

Christopher W. Widenhouse received his doctoral degree from the


Materials Science and Engineering Department of the University of
Florida at Gainsville. He worked as Staff Engineer-New Product Devel-
opment at Cordis Corporation. He is currently based in Cincinnati at
Ethicon Endo-Surgery, a Johnson & Johnson company, where his work
focuses on developing surgical tools and techniques.

Abstract Silicones in various forms have been utilized in about 10 billion dollars per year [10]. Their excellent
biomedical applications over the past several decades. thermal and electrical stability has resulted in their use in
While silicone elastomers are currently the most common the aerospace industry and as insulating materials. Sili-
silicone biomaterial, low molar mass cyclic silicones, low cones have excellent low and high temperature properties.
molar mass linear silicones, silicone oligomers, silicone They can be used as sealants, adhesives and for waterproof
polymers and silicone gels have each found applications as coatings due to their durability and low surface tension.
biomaterials. For the biomedical applications of silicones, Another major application of the silicones is in the medical
it is required that the materials be sterilized prior to field where their excellent biocompatibility makes them
application or implantation. Common methods of sterili- ideal for use in numerous personal care products and
zation include steam sterilization, ethylene oxide medical devices [11–19]. Silicones in the form of low
sterilization, electron beam sterilization and gamma ster- molar mass cyclic silicones, low molar mass linear sili-
ilization. In this review we will describe the behavior of cones, silicone oligomers, silicone polymers, silicone gels
various silicones upon exposure to gamma radiation. and silicone elastomers have all found applications.
The biocompatibility and the biodurability of silicones is
Keywords Silicones  Polysiloxanes  PDMS  a result of other material properties such as hydrophobicity,
Gamma irradiation  Gamma sterilization low surface tension, elastic nature and chemical and thermal
stability. The range of use of silicones in the medical field
covers applications such as orthopedics, catheters, heart
1 Introduction valves, pacemakers and breast, chin and testicular implants
[11–19]. A couple of familiar examples are the use of sili-
Silicones are synthetic polymers whose skeletal backbone cone fluids for lubricating hypodermic syringes and silicone
is made up of silicon–oxygen bonds. The basic repeating elastomers as catheters. Silicone elastomers can also be fil-
unit of a linear polysiloxane molecule has the structure— led with silica to enhance the mechanical properties. Some
(RR0 SiO)n–. The most common member of the siloxane implants also utilize swollen and lightly crosslinked sili-
family is polydimethylsiloxane (PDMS) i.e. when both R cones, referred to as silicone gels, for soft tissue mimicry.
and R’ are methyl groups. Since their development in the The implantable materials must be sterilized before use.
early 1940s [1–8], dimethylsiloxane has been the work- Sterilization can be defined as ‘‘any process which results
horse of the silicone industry. Siloxane units are also in the total destruction or death of all forms of micro-
utilized that have various functional and non functional bial life (bacteria, spores, fungi and viruses) on the
groups attached to silicon. The silicon atoms can be bonded devices.’’ [20]. The commonly used techniques for steril-
to organic groups such as phenyl and vinyl groups. Other izing implants are steam autoclaving, treatment with
important materials include those where silicones are ethylene oxide and exposure to high energy ionizing radi-
copolymerized with a wide variety of organic repeat units. ation [21]. A comparison of the sterilization methods for
The presence of organic groups attached to an inorganic implants has been presented by Park and Lakes and by
backbone gives silicones a set of properties which are Kowalski and Morrissey [22, 23]. They also include ster-
different from the corresponding organic compounds. Sil- ilization technologies such as application of gluteraldehyde
icones can be grouped into compounds based on fluids, solutions, gaseous chlorine dioxide, low-temperature gas
resins and elastomers with the elastomer field accounting plasma, vapor-phase hydrogen peroxide and machine-
for about 50% of the applications in extremely diverse generated X-rays. Sterilization by irradiation can be
fields [9]. Current sales of silicones are estimated to be achieved by irradiating the implantable material either by

123
J Inorg Organomet Polym (2008) 18:207–221 209

gamma rays (Co-60 source) or by a high energy beam of total dose, the dose rate, the presence of additives (solvents,
electrons. Broadly speaking, sterilization by irradiation is fillers, antioxidants, pigments, etc.), the ‘equation of state’
an efficient method and offers the advantage of being able variables (temperature, pressure, etc.) and the atmosphere
to sterilize a packaged article with little or no thermal (air, oxygen, nitrogen, hydrogen, argon, vacuum, etc.).
effect on the device. The penetrating power of gamma
radiation is higher than that of electrons but it is more
costly to install, and sterilization typically takes a longer 2 Linear Polysiloxanes
time [20]. However certain plastics degrade when irradi-
ated, and chemical changes can be induced in the article on Linear polysiloxanes are available as fluids whose viscosity
a molecular level depending on the chemical structure depends on their molecular weight. They consist of Si–O
involved. bonds in the main chain of the polymer. Therefore, it is
Sterilization of polymers can induce either crosslinking or easier to determine the difference between main chain and
degradation in the polymer depending upon the chemical side chain chemistry when the polymer is exposed to
structure. Crosslinking is usually manifested by an increase radiation. In the case of main chain fracture, fractions
in the molecular weight of the polymer with radiation dose, containing silicon or oxygen may be obtained. Thus, it is
giving rise to a branched structure; and finally, a network is possible to study the effects of radiation by irradiating
formed. On the other hand degradation can give rise to main siloxane liquids of different viscosities and monitoring the
chain scission and a drop in the molecular weight. In actu- molecular weight changes for an indication of crosslinking.
ality, both crosslinking and degradation are known to occur Polymers containing carbon in the main chain are
in most polymeric systems with one or the other effects known to crosslink when irradiated. However, the presence
predominating. Examples of the systems where degradation of carbon in the backbone of the chain was not a pre-
is predominant are PMMA and polyisobutylene, while requisite for crosslinking to occur. PDMS can also be
polyolefins are known to crosslink when irradiated [20]. crosslinked when exposed to high energy gamma radiation
Silicones or polysiloxanes and polyurethanes are elas- and electrons, giving rise to a material with elastomeric
tomers that are approved by the Food and Drug properties. The extent of crosslinking is a function of the
Administration (FDA) for use in the form of biomedical radiation dose and can be deduced from solubility studies
devices. Thus, they are used in a wide variety of biomed- carried out on the crosslinked elastomer [24–26]. Linear
ical devices as stated above. All these devices need to be silicones are readily soluble in solvents such as benzene
sterilized before implantation. Sterilization by gamma and xylene. In a typical crosslinking process some of the
radiation is fast becoming the preferred technique in the molecules get tied together with covalent bonds between
United States, replacing ethylene oxide or autoclaving [21]. chains. This results in the increase in viscosity of the fluid.
Thus, as this sterilization technique grows, it is extremely However the solubility is still retained. Adjacent chain
important to have knowledge about the effects of the segments come together and connect to form a network
exposure of silicones to gamma irradiation. structure which is of an infinite extent. All the initial
The purpose of this review is to compile the available molecules taking part in this reaction now form one single
literature on this subject and present an overview in a molecule which loses solubility in any of the aforemen-
concise fashion. The effects of gamma irradiation on the tioned organic solvents. Some molecules remain which are
three different forms of silicones not crosslinked to any portion of the network. These mol-
ecules can be dissolved out of the gel upon swelling in a
(1) Linear chains of polysiloxanes,
solvent and are referred to as the sol fraction. The cross-
(2) Crosslinked polysiloxane elastomers (filled and
linking efficiency should therefore be inversely related to
unfilled), and
the sol fraction. Thus solubility studies on the gel are a
(3) Cyclic siloxanes
good indication of the extent of crosslinking occurring as a
are described in detail. The effects of irradiation on the function of different doses of radiation.
crosslinking and curing of the linear polymer, changes in Early studies by Charlesby and Miller established that the
the mechanical properties of elastomers and polymeriza- degree of crosslinking was a function of the radiation dose.
tion of cyclic species have all been given due consideration Charlesby obtained this by extracting the sol fraction for the
in this review. crosslinked polymer [24]. Miller observed the increasing gel
While the various investigations of gamma irradiation of content with electron irradiation dose [26] (Fig. 1).
silicones reported in the literature and discussed herein have Gel formation commences when there is on average one
been carried out under a variety of experimental conditions, crosslinked monomer unit per weight average molecule
some of the important variables have been identified and and further crosslinking will just increase the amount of gel
evaluated. Of particular importance are the exposure times or produced. Charlesby measured the radiation dose required

123
210 J Inorg Organomet Polym (2008) 18:207–221

energy absorption per crosslink was 32 eV [24]. Based on


the same criteria, Miller calculated a crosslinking yield of
3.0% for irradiation by electrons [26].
Thus irradiating a linear silicone polymer by gamma rays
results in the formation of a crosslinked network polymer
and the degree of crosslinking is related to the radiation dose.
Irradiation of polymers results in the evolution of gases. The
total gas yield should correspond to the crosslinking
occurring in the polymer as long as there are no intramo-
lecular crosslinking or backbiting reactions occurring.
PDMS structure precludes the formation of unsaturation in
the main chain. This is a major difference between the
behavior of silicones and that of the corresponding paraffins.
Sterilization of a silicone article packaged in a polyolefin
film results in the packaging material darkening after irra-
diation. However the silicone itself remains clear and
transparent after sterilization. This is attributed to the ability
of the paraffins to form C=C double bonds in its backbone.
Thus we can expect the total gas yield to match with the
degree of crosslinking in the structure. Miller observed that
this correlation was obeyed up to 100 °C. But for higher
temperatures there was divergence between the gas
yields and the degree of crosslinking [26]. He attributed
this to main chain scission occurring at the elevated
temperatures.
Fig. 1 (a) plot showing the drop in the sol fraction with radiation Mass spectrometric analysis of the evolved gases
dose (gamma) for different viscosities of silicone fluids [24] and (b) revealed that hydrogen, methane and ethane gases were
plot showing the increase in gel fraction with radiation dose
(electrons) for 30,000 cs PDMS oil (Mw = 85,000) irradiated at
produced [24, 26]. This led to the idea that fracture
25 °C and 13.8 Mr/min [26] occurred at the C–Si or C–H bonds giving rise to methyl
radicals. These radicals could then combine with hydrogen
for incipient gelation for linear polymers of different gas or other methyl radicals to give methane or ethane.
molecular weights and viscosities and used the data to Ethane was observed in a very low yield which is rea-
estimate the percentage of monomers crosslinked due to sonable (Fig. 2).
radiation (Table 1). He concluded that the density of
crosslinking is independent of the starting molecular 2.1 Radical Formation Upon Irradiation
weight of the polymer. We see that it varies within the
range of experimental error around the mean value of Radiation induced crosslinking reactions in most organic
2.2%. However the radiation dose required for incipient polymers proceed through the formation of radicals.
gelation is a function of the starting molecular weights.
Charlesby deduced a crosslinking yield of 2.2% and the

Table 1 Calculations for the crosslink density [24]


Bulk Weight No. of Minimum Rgn 100/Rgn
viscosity average monomers per radiation
(cS, 25 °C) mol. Wt. weight dose for gel
(Mw) average formation
molecule (n) (Rg)

1,00,000 1,06,000 1,430 .037 53 1.9


30,000 80,000 1,080 .051 55 1.8
12,500 62,000 840 .07 59 1.7
1,000 26,400 360 .12 43 2.3
50 3,900 53 .72 38 2.6 Fig. 2 Proposed crosslinking mechanism based on gas yields after
irradiation [24, 26]

123
J Inorg Organomet Polym (2008) 18:207–221 211

Knowledge about the nature and the yield of these radi- Si–OH functionalities, for irradiation in the presence of
cals can provide an insight into the mechanism of mercaptan, by the infrared analysis, led Miller to speculate
crosslinking involved. It is possible to react these radicals that main chain scission was occurring and the H atom was
with some radical scavengers such as iodine in very small abstracted by the Si–O functionality [29]. These observa-
concentrations. The scavenger will stabilize the radicals. tions were then corroborated by an electron spin resonance
Thus these radicals will no longer be available for (ESR) study of PDMS in which the –(CH3)2–O–Si* radicals
crosslinking. Consequently the dose of radiation required were clearly identified at room temperature and at low
for incipient gelation will increase in that proportion. This temperature such as 77 K [30]. The radicals at low tem-
increased dose can then be correlated with the additive peratures disappeared after subsequent annealing. However,
concentration. Radical scavengers used this way can the –Si* radical at room temperature was positively iden-
protect the linear silicone from crosslinking. Iodine, sul- tified. This verified Miller’s conclusion that crosslinking
phur and benzophenone have been shown to prevent was occurring through –Si* and –Si–CH2* radicals.
crosslinking in linear silicones either by reacting with the The direct measurement of the trapped radicals at low
polymer radical and stabilizing it or by reacting with the temperature is not a very efficient technique. These radicals
hydrogen atom which would otherwise be available for disappeared at slightly higher temperatures. ESR tech-
abstraction [27, 28]. niques using spintrap measurements indicated the presence
Miller was able to identify and determine the yields of of H3C* and –Si* radicals at room temperature. Keeping in
–SiH, –SiCH2Si–, and –SiCH2CH2Si– groups through FTIR mind the nature of the observed radicals it is possible to
study of the irradiated samples. He also established the propose a reaction mechanism explaining their formation
presence of Si–Si crosslinks [26]. The observation of [31] (Fig. 3).

Fig. 3 Proposed reaction


mechanism to explain the
observed radicals by Miller and
Menhofer [31]

123
212 J Inorg Organomet Polym (2008) 18:207–221

2.2 Effect of Oxygen chain scission has not been elaborated upon in any of the
studies described above. Based on some indirect evidence in
Charlesby observed that the gelation dose required for cases where the gas yield did not agree fully with the cal-
PDMS exposed to the atmosphere was approximately culated degrees of crosslinking, Miller alluded to the idea of
double that of the dose required for degassed samples of main chain scission occurring in irradiated PDMS.
the same viscosity and molecular weight [27]. For electron In a series of subsequent studies involving solid state
irradiation, Miller observed that the crosslinking yield and solution state 29Si–NMR, 13C –NMR, solvent extrac-
decreased from G(X) = 3 to G(X) = 1 in the presence of tion studies and GPC studies for molecular weight
oxygen [29]. He postulated that, in the presence of oxygen, determination, Hill et al. provided the first direct evidence
the –Si–CH2* and the –Si* polymer radicals produced of the occurrence of main chain scission [36–38]. These
upon irradiation are converted into their peroxy derivatives studies were carried out at 303 K in an inert atmosphere.
which are unable to directly participate in crosslinking. With the use of 29Si NMR they were able to identify the
Oxygen interferes with the formed radicals which is presence of three membered, four membered and five
manifested by the decrease in the yield of the dimers membered cyclic structures of PDMS. The only way these
formed upon irradiation and an increase in the oxidation structures could occur was through main chain scission of
products [32]. The oxygenated products were identified as the chain upon irradiation. Backbiting was the postulated
carboxylic acid and two peroxides of the type Si–OOC–. In mechanism for the formation of these cyclics (Fig. 4).
the case of polyolefins, the presence of oxygen during After assignment of the various structures formed after
irradiation inhibits the crosslinking process. However, in irradiation to NMR peaks, the authors were successful in
the case of polysiloxanes, it was not possible to completely verifying the existence of new types of crosslinks formed
inhibit crosslinking. The residual crosslinking is attributed due to main chain scission along with the H-type crosslink
to the existence of small amounts of dimers and also to the described by Charlesby and Miller which was based on
rearrangement of the oxygenated products [29, 32]. Evo- side chain scission. The existence of two different types of
lution of carbon monoxide and carbon dioxide is also crosslinks, i.e. one based on main chain scission (Y-type)
observed upon irradiation in the presence of oxygen [33]. and the other based on side scission (H-type), was proven
for the first time. Figure 5 shows the different types of
2.3 Protective Effect of the Phenyl Group crosslinks as postulated by Hill.
Calculation of the degrees of crosslinking for these new
Oxygen acts as an external agent for the prevention of types of crosslinks revealed that crosslinking occurred
crosslinks in siloxane polymers exposed to radiation. How- principally through the Y-type crosslinks described by Hill.
ever, this protection can be internal as well with the The G values (yield, gaseous or otherwise, per 100 eV of
incorporation of phenyl groups on the PDMS chains by energy absorbed) calculated by Hill were G(Y) = 1.7 and
copolymerizing with small amounts of phenylmethylsilox- G(H) = 1.45 along with a G value for scission of
ane. The phenyl group is capable of protecting as many as six G(S) = 1.15. Thus, crosslinking is predominant for irra-
adjacent siloxane units on the main backbone chain. The diation under vacuum as well as in air, although the net
phenyl group is a resonant structure. An electron knocked out yield of crosslinking in air is slightly lower [38].
of the side chain by ionization or irradiation can be easily
replaced by another electron from the Si–O backbone. The
energy in this process can be dissipated in the various reso-
nant forms of the phenyl group which has a delocalized p
electron structure. The extent of protection increases for
biphenyl groups and so on. However, saturation in protection
is observed for a phenyl content of more than 40 mole %.
Details of copolymers are beyond the scope of this review
work; however readers are referred to more comprehensive
studies on the same subject [34, 35].
The crosslinking mechanism proposed by Charlesby and
Miller (Fig. 2) considers that crosslinking occurs through
C–Si and C–H scission with essentially no main chain scis-
sion. This was corroborated by the positive identification of –
CH2, –CH3 and –Si radicals both at low temperatures and
high temperatures by spin trapping studies. However, the Fig. 4 Mechanism for the formation of low molecular cyclics upon
existence of other structures occurring primarily due to main irradiation of PDMS [36]

123
J Inorg Organomet Polym (2008) 18:207–221 213

Fig. 5 (a) Silmethylene


crosslink (H-type) formed by a
combination of a side chain
methylene radical and an in-
chain silicon radical [37] (b)
Y-type crosslink formed by the
reaction of methylene radicals
and silicon radicals formed by
main chain scission [37]

3 Silicone Elastomers Sheets of similar formulation were also exposed to


e-beam radiation. The effect on the tensile properties, shore
Silicone polymers can be easily transformed into three hardness and percent elongation are shown in the table
dimensional networks through a number of different below. In this case, based on the tensile properties, the
crosslinking reactions. They include the use of a tri or optimum cure seems to be around 10 Mrep (Table 3).
tetra functional siloxane comonomer, using thermal initi- With an increase in the crosslinking, the hardness of the
ators such as a peroxide which abstract a hydrogen atom cured polymer also goes up. It is safe to assume that the
from the siloxane or by the use of high energy radiation. hardness increase is proportional to the radiation dose.
Most of the commercially useful applications of the Comprehensive data for Durometer A type hardness, pro-
polymer are in its elastomeric form which makes it vided by Basfar for different formulations of curing agent
extremely vital to understand the effect of gamma radia- in silicone rubber, is shown in Fig. 6. Different formula-
tion on the properties of these materials. Most of the tions are represented as 0/10 (parts of curing agent/parts of
elastomers are filled with silica or titania filler for rein- silicone rubber), 0.5/10, 0.25/10 and 1/10 for Si, Si-1, Si-2
forcement purposes. The effect of filler in the interactions and SiC respectively [40]. A 300–400 kGy dose was
with radiation is also of interest. required for optimum cure.
The improvements in the tensile properties, modulus
and ultimate properties for PDMS elastomers blended with
3.1 Enhancement in Mechanical Properties

For most polymers, crosslinking results in the formation of Table 2 Effect of gamma radiation on silicone rubber [39]
a dense network with an enhancement in mechanical Dose (Mrep) Shore Tensile (psi) % Elongation
properties such as modulus, hardness and tensile properties.
25 53 916 158
Thus, these properties, especially the modulus, can be
5 27 1,180 750
monitored as an indication of crosslinking in a sample.
1.25 18 135 550
Warrick, in 1955 [39], provided a comparative study of
the effects of gamma ray and electron-beam (e-beam)
Table 3 Effect of e-beam radiation on silicone rubber [39]
irradiation on the mechanical properties of a silicone
elastomer filled with 35% SiO2 filler loading. The vulca- Dose (Mrep) Shore Tensile (psi) % Elongation
nization of silicone rubber was carried out by gamma rays
2 15 153 780
from a Cobalt-60 source, and the tensile strength, %
6 26 742 605
elongation and shore hardness were studied as functions of
10 29 876 580
the radiation dose. Based on the tensile properties and the
20 43 679 250
% elongation values, the optimum cure for the silicone
40 52 561 117
sample was somewhat above 5 Mrep (Table 2).

123
214 J Inorg Organomet Polym (2008) 18:207–221

Fig. 6 Shore A hardness measurements as a function of the dose for


formulations of silicone rubber [40]

silica filler and cured by gamma radiation have been


studied as a function of the radiation dose [41]. Swelling
studies revealed that the presence of filler slightly
improved the crosslinking efficiency in PDMS and caused
a drop in the soluble fraction in the elastomer after
extraction in toluene. Mooney–Rivlin plots for the elasto-
mer revealed an increase in the tensile properties with
higher doses (Fig. 7).
For most polymeric systems, an increase in the ultimate
strength of the material at rupture is synonymous with a
Fig. 7 (a) Soluble fraction after extraction v/s radiation dose for
decrease in the ultimate elongation. However, this was not
filled and unfilled systems and (b) Mooney–Rivlin stress–strain plots
observed for unfilled PDMS cured by gamma radiation, for filled and unfilled elastomers for different radiation doses [41]
where the ultimate strength increased with an increase in the
radiation dose along with a concurrent increase in the ulti-
mate elongation. For filled systems the reinforcing effect of the actual crosslink density of the elastomer. Studies
the colloidal silica completely masks that of radiation. revealed that silica reinforcement contributed about 67% to
However, both the filler as well as the radiation have a the apparent crosslink density of the filled elastomer [42].
cumulative effect on the strength enhancement (Fig. 8). It was also observed that filled silicone copolymer samples
irradiated in air showed an initial drop in the apparent
3.2 Effects of Filler Loading crosslink density due to a disruption of the hydrogen bonds
on the filler surface. There was no drop observed for
Silica filler plays an important role in PDMS by contrib- samples irradiated in a vacuum and the crosslink density
uting to the overall increase in crosslink density by forming showed steady increase (Fig. 9).
hydrogen bonds between the silicone polymer chain and NMR and GC-MS studies also revealed the formation of
the surface silanol groups on the silica filler. These are degradation products in the form of low molecular weight
often termed as apparent crosslinks and the resultant cyclics formed as a result of backbiting [42]. The yield of
increase in the crosslink density is the apparent crosslink these degradation products was higher in the vacuum
density. It is possible to disrupt this hydrogen bonding on environment as opposed to those irradiated in atmospheric
the surface of the filler by swelling in ammonia to obtain conditions, thus highlighting the role of oxygen as a radical

123
J Inorg Organomet Polym (2008) 18:207–221 215

scavenger. This is analogous to some of the studies


described for linear polysiloxanes described above
involving main chain scission and subsequent degradation
products [36–38].
Dynamic mechanical analysis of filled PDMS-PDPS
copolymer systems has revealed the apparent hardening and
the decrease in the segmental mobility of the elastomer by
monitoring the equilibrium storage modulus with increasing
radiation dose. As expected, a direct dependence is
observed due to a predominant crosslinking reaction [43].

3.3 Effects on Crystallization Behavior

Crosslinking on irradiation leads to a dramatic change in


the segmental dynamics of the elastomer, reducing the free
motion of the polymer chains. This has to manifest itself in
the crystallization behavior of the material because seg-
mental motion is vital for the polymer chains to reach the
crystallization nucleus and contribute to the overall degree
of crystallinity in the material.
Crystallization studies on irradiated silicone rubbers
indicated that the degree of crystallinity and the rate of
crystallization decreases with an increasing radiation dose
and a dose of 100 Mrads almost entirely eliminates crys-
tallization. There was observed to be a drop of about 20 °C
in the melting point for doses up to 40 Mrads. However,
changes in Tg were insignificant [44]. This impairment in
crystallization is to be expected on account of the observed
crosslinking in the rubber and was corroborated by several
later studies [43, 45, 46].

3.4 Toxicity Studies


Fig. 8 (a) Ultimate strength as a function of radiation dose for filled
and unfilled elastomeric systems and (b) Ultimate elongation as a
function of the radiation dose for filled and unfilled systems [41] It is clear from this discussion so far that silicones upon
irradiation primarily undergo crosslinking with minor
amounts of degradation products being formed as a result
of main chain scission. However the degradation products
happen to be cyclic siloxanes of low molecular weight in
extremely small amounts, trapped inside the matrix. Thus
they are expected to be equally benign to a cellular envi-
ronment. According to the Silicones Environmental, Health
and Safety Council of North America (SEHSC), there is no
evidence of toxicity issues or adverse health effects when
D4 and D5 are used as intended in personal care products.
In vitro cytotoxicity studies on irradiated silicone rubbers
further confirmed this observation [45].
Cells were cultured in the presence of irradiated silicone
samples and viability was measured by the neutral red
uptake assay in the presence of positive and negative con-
trols. Phenol was used as a positive control and PVC pellets
were used as a negative control. The behavior of the silicone
Fig. 9 Plot of percent change in crosslink density against radiation samples mimicked that of the negative control indicating
dose (crosslink density increases in the downward direction) [42] that no cytotoxicity occurs upon irradiation (Fig. 10).

123
216 J Inorg Organomet Polym (2008) 18:207–221

of polymerization in the solid state are complex, making


this an over-simplified assumption.
Attempts made to polymerize the cyclic tetramer by
gamma radiation led to a mixture of dimeric products and
high molecular weight linear and cyclics components [48].
The dimers were produced in very small yields along with
gases like hydrogen, methane and ethane. These dimers were
produced through Si–Si, Si–CH2–Si and Si–CH2–CH2–Si
linkages. There was correspondence between the gas yields
and the yield of dimers which confirmed the mechanism.
However for the polymer to form, siloxane ring rear-
rangement had to occur. This would involve the opening of
a ring which would lead to subsequent opening of other
rings to be incorporated into the polymer chain. Since most
Fig. 10 Cell viability curves by neutral red uptake assay for silicones ring opening catalysts are ionic, it was speculated that
irradiated at 50 kGy [45] radiation produced an ion of the nature shown in Fig. 11.
Starting from the silicenium ion, a reaction sequence for
4 Cyclic Polysiloxanes the formation of linear PDMS would then be represented as,
M ! Mþ þ CH3 þ e
This class of compounds has the general structural formula
(RR0 SiO)n, where values of n up to 500 have been observed in Mþ þ C ðinhibitiorÞ ! P (unreactive species)
well equilibrated mixtures. When R=R0 =CH3, the rings can
Mþ ! Rþ (rearrangement to radical ion)
simply be labeled as D (D5, D6 and D7). Some of the most
commonly prepared compounds in this series include cyclic Rþ þ M ! RMþ
poly(dimethylsiloxane) [(CH3)2SiO]n, cyclic poly(vinyl-
RMþ þ M ! RMþ
2 ðpropagation stepÞ
methylsiloxane) [CH2=CH(CH3)SiO]n and cyclic poly
(phenylmethylsiloxane) [C6H5(CH3)SiO]n. The primary use RMn1 þ M ! RMþ
n
of cyclic siloxanes is in the synthesis of linear polysiloxanes
RMþ
n þ e ! RMn ðcharge neutralization termination stepÞ
through ring-chain equilibration reactions and ring-opening
polymerization. Ring-opening polymerization requires a Chawla and St. Pierre reported that gamma irradiation
break in the backbone Si–O bonds of the cyclic structure. initiated liquid state polymerization of both D3 and D4 [49].
Thus, cyclic structures such as D3 and D4, with inherent The previously known polymerization of D3 was in the
strain in them due to strained bond angles, are preferred for solid state. Appreciable yields were obtained when both the
polymerization. monomers were in the ‘ultra pure’ and ‘ultra dry’ states.
To determine the nature of the propagating species
4.1 Radiation Polymerization of Cyclic Siloxanes (anionic or cationic), it is necessary to use electron scav-
engers. If the propagating species is cationic, then it will be
Polymerization of cyclic siloxanes requires ionic initiators terminated by an anionic species in the reaction. Thus, if we
for siloxane bond rearrangement reactions to occur. High
energy radiation can generate both ions as well as radicals
in the monomeric materials and thus there was a possi-
bility of initiating polymerization in cyclic siloxanes by
ionizing radiation. However, early attempts to polymerize
the cyclic trimer (D3) by high energy electrons gave poor
yields [47]. Also, the polymerization could be carried out
only in the solid state and as soon as the monomer
transformed to above its melting point, the polymerization
yield reduced to zero. This hinted towards a terminating
agent in the liquid state or the absence of proper propa-
gating species in the liquid state. Also the mechanism
could not be established and the dependence of the yield
on the radiation dose rate led to speculation that the Fig. 11 Silicenium ion produced as a result of gamma irradiation of
mechanism was free radical based. However, the kinetics the cyclic tetramer [48]

123
J Inorg Organomet Polym (2008) 18:207–221 217

add a cationic species to the reaction, which will act as a


scavenger to all the anions present in the reaction, the rate of
polymerization will increase [50, 51]. Work was carried out
using nitrous oxide as a potential electron scavenger. This
led to a 4-fold increase in the rate of polymerization which,
consequently, indicates that the polymerization reaction
propagates through a cationic species [50] (Fig. 12).
Rapid advances have since been reported in a number of
other studies. With the use of ultra dry (obtained by contact
with a series of sodium mirrors) and ultra pure monomers, Fig. 13 Reaction sequence for ROP of gamma radiation initiated D4
polymerizations could be carried out both in liquid as well [54]
as in solid states with different rate dependence for both.
Molecular weight measurements indicated that linear
PDMS with molecular weight going up to several million
can be formed [51]. Polymerization of D5 in the liquid state
has also been successfully reported [52].
In all these cases, the product obtained contains a small
percentage of cyclics which are presumed to be formed by
backbiting reactions. These reactions compete with the
primary propagation reaction and originate from the same
cationic active center generated by gamma irradiation [52–
54]. The same type of active center was formed for all the
three monomers i.e. D3, D4 and D5. The initial rates, the
final leveled off conversions, the concentration of cyclics
formed by backbiting and the activation energies were
consequently the same for all three monomers [54]. These
results are completely different from chemically-initiated
species. Thus, a possible reaction sequence for polymeri-
zation of cyclic PDMS is shown in (Fig. 13).
The plot of the rate of polymerization for both D4 and
D5 and for the square root of the radiation dose rate was
linear and passed through the origin, as is the case for ionic
polymerizations in the absence of impurities. Addition of
ammonia and triethylamine inhibited the polymerization
process, further corroborating the cationic mechanism for
propagation [53] (Fig. 14).

Fig. 14 Plot of the rate of polymerization vs. square root of radiation


dose [54]

The same mechanisms of silicenium ion formation by


gamma irradiation and the subsequent ring opening poly-
merization has also been applied to co-polymerizations of
D3 with D4 and D4 with D5. For further details regarding
the same, readers are directed towards texts describing the
process in detail [55].

4.2 Irradiation of Cyclic Poly(dimethylsiloxane)

We have discussed above the investigations that have been


reported in the literature on the gamma irradiation of low
Fig. 12 Plot showing the percent yield of the polymer against molar mass cyclic siloxanes. We have also performed
concentration of nitrous oxide [50] studies on the effects of irradiation of well characterized

123
218 J Inorg Organomet Polym (2008) 18:207–221

cyclic poly(dimethylsiloxane) sharp fractions (S. J. Clarson would cause a drop in the apparent crosslinking
et al., Unpublished results). One of the reasons for these density and small changes in the modulus.
studies was the idea that networks formed from large rings
The phenomena described above are occurring concur-
can have unique physical properties [56, 57]. In particular,
rently when a device is exposed to sterilizing radiation. As
these ‘model networks’ would perhaps be unique in that if
a result, the effects may compensate for each other causing
the cross-linking chemistry is solely via hydrogen
negligible noticeable changes in the structure and in the
abstraction (and therefore Si–CH2–CH2–Si linkages), then
mechanical properties of the product.
such networks would contain no dangling chain ends.
Recently, we have investigated a variety of aspects of
Another point worth noting is the susceptibility of the
the effects of gamma irradiation on the properties of
Si–H linkage in various organosilicon compounds to radi-
commercial biomedical grade silicone elastomers before
ation damage. After studying the gas yields upon exposure
and after sterilization. These investigations have included
to gamma radiation, Warrick observed that the Si–H bond
sol-fraction determination and analysis, thermal aging
is 30 times more prone to radiation rupture than the C–H or
studies, and dynamic mechanical modulus measurements
Si–C bonds. Thus, hydride terminated silicones and other
and some results from these investigations are presented
materials containing the silyl hydride functionality are
here.
particularly sensitive to radiation damage [58].
Storage modulus measurements were conducted on a
Mettler Toledo 861E Dynamic Mechanical Analyzer in a
tensile clamping mode as a function of the temperature.
5 Observations The experiments revealed an increase of about 8.3% in the
storage modulus of the gamma sterilized silicone rubber as
Silicone medical devices are sterilized in the final product shown in Fig. 15 (A.S. Palsule et al., 2006, Unpublished
form after packaging by exposing them to gamma radiation results).
or to a high energy beam of electrons. Many of the prod- Modulus measurements were accompanied by the
ucts are in a covalently crosslinked elastomeric form and quantification of extractable sol content in the elastomer.
some of them may contain a lightly crosslinked silicone The silicone sample was swollen in toluene to open up the
gel. A small fraction of uncrosslinked oligomeric fractions diffusion paths in the network. This allowed the low
of linear and cyclic polysiloxane are also known to occur in molecular weight uncrosslinked oligomeric fractions to
the elastomeric part of the device. They are trapped in the leach out of the network. The sample was then shrunk in
network elastomeric phase and several weight percent of
such materials can often be extracted from the device by
swelling in toluene or other appropriate solvents.
On the basis of the studies described above, it is rea-
sonable to make the following observations regarding the
effect of sterilizing radiation on a silicone medical implant.
(1) The implant is made up of a pre-cured elastomer with
the optimum cure having been decided by the balance
of mechanical properties. For a peroxide cured
material, additional crosslinking may occur during
sterilization. This may result in the incorporation of
the oligomeric, extractable, previously uncrosslinked
fractions in the network structure.
(2) Crosslinking dominates over main chain scission
when silicones are exposed to gamma radiation.
However, small amounts of chain scission can occur
which leads to backbiting and formation of small
amounts of low molecular weight cyclic and linear
polysiloxanes. These products are again trapped in the
network and can be subsequently crosslinked in to the
network structure.
(3) Radiation may cause breakage of the hydrogen bonds
Fig. 15 Storage modulus of gamma sterilized and unsterilized
at the filler interface between the surface silanol silicone implantable elastomer obtained by DMA measurements
groups of the filler and the siloxane polymer. This (A.S. Palsule et al., 2006, Unpublished results)

123
J Inorg Organomet Polym (2008) 18:207–221 219

Sterilized Silicone Thus, we can say that,


Unsterilized silicone 1 Rad = 0.01 Gy or 1 centigray and 1 Rep = 0.93 Rad
or 9.3 milligrays.
TGA of sterilized and unsterilized Silicone
4.5

6 Concluding Remarks
4
Based on the review of the existing literature on the gamma
irradiation of silicones, it is reasonable to suggest that
Weight(mg)

sterilization of silicones by gamma irradiation has the


3.5 potential to be a more efficient and more effective tech-
nique when compared to thermal or chemical treatments.
We certainly hope that those of you who that have
reached this point in the text have found the material
3
presented herein and the associated bibliography compiled
to be useful to you. As the use of silicones as biomedical
materials continues to expand, we fully anticipate that we
2.5 will revisit this topic at some suitable point in the future. At
0 200 400 600 800 1000
that time, we will either prepare an updated article or
Temperature (C)
provide an appendix to this one. We therefore openly
Fig. 16 Thermogravimetric data of gamma sterilized and unsterilized welcome any comments on this text and also any sugges-
silicone (A.S. Palsule et al., 2006, Unpublished results) tions that you may have for material to be included in the
future.

methanol and the sol fraction was calculated by monitoring


the weight loss in the sample. The procedure was repeated References
for both the sterilized and unsterilized samples. This
analysis revealed an average drop of 17.5% in the sol 1. E.G. Rochow, An Introduction to the Chemistry of the Silicones
(John Wiley & Sons, New York, NY, USA, 1946)
fraction of the sterilized sample, which indicated a post 2. W. Noll, Chemistry and Technology of Silicones (Academic
sterilization increase in the crosslinking of the elastomer as Press, Orlando, FL, USA, 1968)
was evident from the increase in the modulus. 3. H.A. Liebhafsky, Silicones Under the Monogram: A Story of
Finally, thermogravimetric degradation studies of the Industrial Research (John Wiley & Sons, New York, NY, 1978)
4. E.L. Warrick, Forty Years of Firsts: The Recollections of a Dow
material did not reveal any significant changes in the Corning Pioneer (McGraw-Hill, New York, NY, USA, 1990)
degradation characteristics of the material post steriliza- 5. E.G. Rochow, Silicon and Silicones (Springer-Verlag, Berlin,
tion. There was no observable evidence of the formation of Germany, 1987)
low molecular weight degradation products upon gamma 6. S.J. Clarson, J.A. Semylen (eds.), Siloxane Polymers (Prentice
Hall, Englewood Cliffs, NJ, USA, 1993)
irradiation of the silicone and, thus, it is safe to assume that 7. B. Knaap, Silicon (Grolier Educational, Danbury, CT, USA,
the net effect of gamma radiation on silicones is to increase 1996)
the crosslinking in the elastomer as has been proposed 8. J. Thomas, The Elements: Silicon (Benchmark Books, Marshall
above (A.S. Palsule et al., 2006, Unpublished results) Cavendish Corporation, Tarrytown, NY, USA, 2002)
9. S.J. Clarson, J.E. Mark, Silicone elastomers, in Polymeric
(Fig. 16). Materials Encyclopedia, ed. by J. Salamone, vol 10, (1996) p.
7663
5.1 Different Units of Radiation Dose Used in the 10. S.J. Clarson, Silicones and silicone-modified materials: a concise
Review and Their Inter-Conversions overview, in Synthesis and Properties of Silicones and Silicone–
Modified Materials, ed. by S.J. Clarson, J.J. Fitzgerald, M.J.
(1) Gray (Gy, MGy) is the SI unit of absorbed dose and Owen, S.D. Smith, M.E. Van Dyke, ACS Symposium Series
Volume 838, (Oxford University Press, New York, NY, USA,
one gray is equal to an absorbed dose of 1 joule/Kg
2003), p. 2
(2) Rad (Mrad) is also a unit of the absorbed dose and 11. A. Colas, J. Curtis, Silicone biomaterials: history, chemistry and
one rad dose is equal to an absorbed dose of 100 ergs/ major applications of silicones, printed in Biomaterials Science:
gram An Introduction to Materials in Medicine, 2nd edn., (Elsevier
Publications, 1996) pp. 80–86 and pp. 697–707
(3) Rep (Mrep) is an obsolete unit of radiation dose and it
12. R.R. Mcgregor, Physiological Response to Silicones, in Silicones
is an acronym for Roentgen equivalent physical. It is and Their Uses (McGraw-Hill, New York, NY, 1954), pp.
equal to an absorbed energy dose of 93 ergs/gram 188–198

123
220 J Inorg Organomet Polym (2008) 18:207–221

13. W. Noll, Physiological Behaviour, in Chemistry and Technology 35. G.C. Corfield, D.T. Astill, D.W. Clegg, Radiation stability of
of Silicones (Academic Press, Orlando, Florida, USA, 1968), pp. silicon elastomers, Amer. Chem. Soc. 24, 473 (1984)
516–527 36. D.J.T. Hill, C.M.L. Preston, A.K. Whittaker, S.M. Hunt, The
14. B. Arkles, Look what you can make from silicones. Chemtech, radiation chemistry of poly(dimethylsiloxane), Macromol. Symp.
13, 542–555 (1983) 195, 95 (2000)
15. E.L. Warrick, Center for aid to medical research, in Forty Years 37. D.J.T. Hill C.M.L. Preston A.K. Whittaker, NMR study of the
of Firsts: The Recollections of a Dow Corning Pioneer, ed. by gamma radiolysis of poly(dimethylsiloxane) under vacuum at
E.L. Warwick, (McGraw-Hill, New York, NY, USA, 1990), pp. 303 K, Polymer 43(4), 1051 (2002)
184–187 38. D.J.T. Hill, C.M.L. Preston, D.J. Salisbury, A.K. Whittaker,
16. E.L. Warrick, Medical products business, in Forty Years of Molecular weight changes and scission and crosslinking in
Firsts: The Recollections of a Dow Corning Pioneer, ed. by E.L. poly(dimethylsiloxane) on gamma radiolysis, Rad. Phys. Chem.
Warwick (McGraw-Hill, New York, NY, USA, 1990) 62, 11–17 (2001)
17. S.A. Visser, R.W. Hergenrother, S.L. Cooper, Polymers, in Bio- 39. E.L. Warrick, Effect of radiation on organopolysiloxanes, Ind.
materials Science: An Introduction to Materials in Medicine, ed. Engg. Chem. 24(7), 842 (1955)
by B.D. Ratner, A.S. Hoffman, F.J. Schoen, J.E. Lemons (Aca- 40. A.A. Basfar, Hardness measurements of silicone rubber and
demic Press, San Diego, CA, USA, 1996), p. 59 polyurethane rubber cured by ionizing radiation, Radiat. Phys.
18. M.A. Brook, Silicon in a biological environment, in Silicon in Chem. 50(6), 607 (1997)
Organic, Organometallic and Polymer Chemistry, ed. by M.A. 41. J.E. Mark, D.W. McCarthy, Poly(dimethylsiloxane) elastomers
Brook (John Wiley and Sons, New York, NY, 2000), pp. 459–479 from aqueous emulsions: III effects of blended silica fillers and
19. M.A. Van Dyke, S.J. Clarson, R. Arshady, Silicone biomaterials, gamma-radiation induced crosslinking, Rub. Chem. Tech. 71,
in Introduction to Biomaterials, ed. by R. Arshady (Citus Books, 941 (1998)
London, UK, 2003), pp. 109–135 42. A. Chien, R. Maxwell, D. Chambers, B. Balazs, J. LeMay,
20. S.D. Bruck, Sterilization problems of synthetic biocompatible Characterization of radiation-induced aging in silica-reinforced
materials. J. Biomed. Mater. Res. 5(3), 139 (1971) polysiloxane composites, Rad. Phys. Chem. 59, 493–500 (2000)
21. S.D. Bruck, E.P. Mueller, Radiation sterilization of polymeric 43. R. Maxwell, R. Cohenour, W. Sung, D. Solyom, M. Patel, The
implant materials, J. Biomed. Mater. Res. Appl. Biomater. effects of gamma-radiation on the thermal, mechanical and
22(A2), 133–144 (1988) segmental dynamics of a silica filled, room temperature vulca-
22. J.B. Park, R.S. Lakes, Sterilization effects in biomaterials: an nized polysiloxane rubber, Poly. Degrad. Stab. 80, 443–450
introduction by Plenum Press, New York, NY, USA, (1992), p. (2003)
162 44. I.S. Lazurkin, G.P. Ushakov, The effect of radiation on properties
23. J.B. Kowalski, R.F. Morrissey, Sterilization of implants, in Bio- of silicone rubber, Atom. Energy 4(3), 365 (1958)
materials Science: An Introduction to Materials in Medicine, ed. 45. S.O. Rogero, J.S. Sousa, D. Alario, L. Lopergolo, A.B. Lugao,
by B.D. Ratner, A.S. Hoffman, F.J. Schoen, J.E. Lemons (Aca- Silicone crosslinked by ionizing radiation as a potential poly-
demic Press, San Diego, CA, USA, 1996), pp. 415–420 meric matrix for drug delivery, Nucl. Instrum. Methods Phy. Res.
24. A. Charlesby, Changes in silicone polymeric fluids due to high- B, 236, 521 (2005)
energy radiation, Proc. Roy. Soc. A230(1180), 120, (1955) 46. I. Stevenson, L. David, C. Gauthier, L. Arambourg, J. Davenas,
25. A.M. Bueche, An investigation of the theory of rubber elasticity G. Vigier, Influence of silica fillers on the irradiation ageing of
using irradiated Polydimethylsiloxane, J. Polym. Sci. 19, silicone rubbers, Polymer, 42, 9287–9292 (2001)
297–306 (1956) 47. E.J. Lawton, W.T. Grubb, J.S. Balwit, A solid state polymeri-
26. A.A. Miller, Radiation chemistry of polydimethylsiloxane, zation initiated by high energy electrons, J. Polym. Sci. 19, 455
crosslinking and gas yields, J. Amer. Chem. Soc. 82, 3519 (1960) (1956)
27. A. Charlesby, P.G. Garratt, Radiation protection in irradiated 48. C.J. Wolf, A.C. Stewart, Radiation chemistry of octa-
dimethylsiloxane polymers, Proc. Roy. Soc. Lond. A273, methylcyclotetrasiloxane, J. Phys. Chem. 66, 1119 (1962)
117–132 (1963) 49. A.S. Chawla, L.E. St. Pierre, Gamma ray-induced liquid–state
28. H.A. Dewhurst, L.E. St. Pierre, Radiation chemistry of hexame- polymerization of hexamethylcyclotrisiloxane and octa-
thyldisiloxane, A Polydimethylsiloxane model, J. Phys. Chem. methylcyclotetrasiloxane, J. App. Polym. Sci. 16, 1887–1891
64, 1063 (1960) (1972)
29. A.A. Miller, Radiation chemistry of polydimethylsiloxane II 50. A.S. Chawla, L.E. St. Pierre, Solid state polymerization of hex-
effects of additives, J. Am. Chem. Soc. 83, 31–36 (1961) amethylcyclotrisiloxane II, J. Polym. Sci. A1, 10(9), 2691 (1972)
30. M.G. Ormerod, A. Charlesby, The radiation chemistry of some 51. J.J. Lebrun, A. Deffieux, P. Sigwalt, A. Wang, V. Stannett, The
polysiloxanes: an electron spin resonance study, Polymer 4, radiation induced polymerization of hexamethylcyclotrisiloxane
459–470 (1963) and octamethylcyclotetrasiloxane in solid and liquid states,
31. H. Menhofer, H. Heusinger, Radical formation in poly- Radiat. Phys. Chem. 24(2), 239 (1984)
dimethylsiloxanes and polydimethylphenylsiloxanes studies by the 52. D.M. Naylor, V. Stannett, A study of the ionic polymerization of
ESR spintrap technique, Radiat. Phys. Chem. 29(4), 243–251 (1987) decamethylcyclopentasiloxane induced by gamma radiation, J.
32. H.A. Dewhurst, L.E. St. Pierre, The effect of oxygen on the Polym. Sci: Part C: Polym. Lett. 24, 319–326 (1986)
radiolysis of silicones, J. Phys. Chem. 64, 1060 (1960) 53. D.M. Naylor, V. Stannett, A. Deffieux, P. Sigwalt, Radiation
33. H. Menhofer, J. Zluticky, H. Heusinger, The influence of irradi- induced polymerization of dimethylcyclosiloxanes in the liquid
ation temperature and oxygen on crosslink formation and state: I. Influence of drying and the nature of the propagation
segment mobility in gamma irradiated polydimethylsiloxanes, mechanism, Polymer 31, 954 (1990)
Radiat. Phys. Chem. 33(6), 561–566 (1986) 54. D.M. Naylor, V. Stannett, A. Deffieux, P. Sigwalt, Radiation
34. C.G. Delides, The protective effect of phenyl group on the induced polymerization of dimethylcyclosiloxanes in the liquid
crosslinking of irradiated dimethyldiphenylsiloxane, Radiat. state: 2. Polymerization kinetics of D3, D4 and D5 and the dis-
Phys. Chem. 16(5), 345–352 (1980) tribution of products, Polymer 32(6), 1084 (1991)

123
J Inorg Organomet Polym (2008) 18:207–221 221

55. D.M. Naylor, V. Stannett, A. Deffieux, P. Sigwalt, Radiation 57. P.G. de Gennes, Cornell University Press, Ithaca, NY, USA, pp.
induced polymerization of dimethylcyclosiloxanes in the liquid 132–133
state: 3. Copolymerization of D3 with D4 and D4 with D5, reac- 58. J.F. Zack, E.L. Warrick, G. Knoll, Radiation stability of or-
tivities and interpretation, Polymer 35(8), 1764 (1994) ganosilicon compounds, J. Chem. Eng. Data 6(2), 279 (1961)
56. J.A. Semlyen, Polymers based on long chain and large ring
molecules, in Large Ring Molecules, ed. by J.A. Semlyen (John
Wiley & Sons, Chichester, England), p. 3

123

View publication stats

You might also like