You are on page 1of 11

pubs.acs.

org/JPCB Article

Constitutive Model of Radiation Aging Effects in Filled Silicone


Elastomers under Strain
A. Maiti,* W. Small,§ M. P. Kroonblawd,§ J. P. Lewicki, N. Goldman, T. S. Wilson, and A. P. Saab

Cite This: J. Phys. Chem. B 2021, 125, 10047−10057 Read Online

ACCESS
Downloaded via BEIJING COMPUTATIONAL SCI RSRCH CTR on November 9, 2021 at 02:09:12 (UTC).

Metrics & More Article Recommendations


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Filled silicone elastomers, an essential component in many technological


applications, are often subjected to controlled or unintended radiation for a variety of reasons.
Radiation exposure can lead to permanent mechanical and structural changes in the material,
which is manifested as altered mechanical response, and in some cases, a permanent set. For
unfilled elastomers, network theories developed and refined over decades can explain these
effects in terms of chain-scission and cross-link formation and a hypothesis involving
independent networks formed at different strain levels of the material. Here, we expose a filled
silicone rubber to gamma radiation while being under finite elongational strain and show that
the observed mechanical and structural changes can be quantitatively modeled within the same
theoretical framework developed for unfilled elastomers as long as nuances associated with the
Mullins effect are accounted for in a consistent manner. In this work, we employ Ogden’s incompressible hyperelastic model within
the framework of Tobolsky’s two-network scheme to describe the observed permanent set and mechanical modulus changes as a
function of radiation dosage. In the process, we conclude that gamma radiation induces both direct cross-linking at chain crossings
(H-links) and main-chain-scission followed by cross-linking (Y-links). We provide an estimate of the ratio of chain-scission to cross-
linking rates, which is in reasonable agreement with previous experimental estimate from Charlesby−Pinner analysis. We use density
functional theory (DFT)-based quantum mechanical calculations to explore the stability of −Si and −SiO radicals that form upon a
radiation-induced chain-scission event, which sheds light on the relative rates of Y-linking and H-linking processes.

■ INTRODUCTION
Filled and cross-linked elastomers and foams are materials of
not only guides the construction of a radiation-age-aware model
but also puts to test several important network theories and
tremendous technological importance1−3 with applications concepts developed over decades and provides useful insights
spanning support cushions and pads, coatings, adhesives, into the chemical effects of radiation at the molecular level.
interconnects, acoustic/seismic-isolation, thermal/electrical Most of the fundamental theories have been developed for
insulation, and bioimplants.4−10 Among the diverse choice of unfilled rubber networks. The earliest important result states
polymers, silicones constitute a special class, possessing extreme that the mechanical response of rubber is primarily entropic in
thermal and electrical stability and a high degree of nature, with the increase in free energy arising out of loss in
biocompatibility, which makes them ideal for applications configurational entropy16−18 when a networked rubber speci-
such as artificial organs/implants and biomedical devices.11−14
Given that in most applications, silicone materials are men is deformed from its equilibrium state. A landmark result of
expected to possess long lifetimes (years to decades), it is this analysis is the fact that if the cross-links (i.e., polymer
important to develop age-aware models that account for the segments between two junctions) are assumed to be Gaussian
interaction of the rubber or foam component with various chains, and each junction is assumed to move affinely under
environmental stresses like temperature, humidity, oxidative deformation, then the shear modulus μ of the rubber is
degradation, and radiation. The latter, which could be in the proportional to its number-density ν of cross-links (per unit
form of high-energy X-rays, gamma rays, or an electron beam, volume) in the form μ = νkBT, where kB is the Boltzmann
often arises in intended exposure, e.g., to perform sterilization of constant and T is the absolute temperature.
a bioimplant component prior to placement within the subject’s
body. Radiation, especially penetrating gamma rays, is known to
induce permanent chemical changes in rubber, filled or unfilled, Received: June 5, 2021
which can lead to the scission of main polymer chains and the Revised: August 6, 2021
formation of new cross-links.15 Published: August 27, 2021
It is particularly interesting to explore the effect of cross-
linking and main-chain-scission on the structural and mechan-
ical response properties of elastomeric networks. Such a study

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcb.1c04958


10047 J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

There have been many refinements of the above picture, Following exposure to controlled duration (and therefore
including (but not limited to) accounting for the effects of controlled cumulative dosages) of radiation, each sample was
junction fluctuations,17,18 contribution of physical entangle- removed from the irradiation chamber, released from the λa
ments,19,20 non-Gaussian behavior at a larger deformation,21−26 strain, and allowed to relax under ambient conditions for 24 h.
and the effect of fillers27,28 that are used for enhancing the The relaxed samples were then subjected to measurements of
mechanical modulus. For lightly cross-linked (or un-cross- the new equilibrium length, called the recovered length λs. Two
linked but entangled) materials, there is a complex time- sets of experiments were performed: (i) in the first, samples
dependent response known as viscoelasticity. Theories for this under λa= 1.20, 1.47, 1.67, and 1.84 were subjected to
are still undergoing refinement,26,29−32 although some basic cumulative dosages of 1, 3, 7, and 17 Mrad of radiation and
concepts involving reptation tube dynamics were established in (ii) in the second, samples under λa= 1.67, 2.00, 2.33, and 2.67
the 1970s33 and more fully developed in the 1980s.34 were subjected to cumulative dosages of 5, 10, and 17 Mrad of
Given that incorporating all the above theories to account for radiation. After several weeks of further equilibration, tensile
the observed structural and mechanical response of radiation- stress−strain analysis was carried out on a few of the specimens,
aged, filled rubber is a tall order, a few simplifying including specimens that were not stretched (i.e., λa= 1.00)
approximations are needed to generate a practical, working when exposed to radiation. The stress−strain analysis was
model. The first (well-justified) approximation for our system is performed on rectangular specimens (∼3 mm wide by ∼40 mm
that the rubber is cross-linked enough that viscoelasticity can be long) cut from the irradiated samples, as well as a pristine unaged
neglected, and the resulting (energy-conserving) response is specimen, using an Instron 5565 dual-column electromechanical
hyperelastic. This means that any stress−strain response is test system with an initial grip separation of ∼20 mm and a
governed by a strain energy functional.35−37 stretching rate of 20 mm/min.
Ideally, we would like to employ a hyperelastic model that
quantitatively describes the observed stress−strain behavior of
unaged and radiation-aged rubber. However, as we will see
■ MULLINS EFFECT
One of the challenges in modeling stress−strain response in a
below, the picture gets complicated for filled rubber because of filled rubber with a hyperelastic strain function is the omni-
the significant presence of the Mullins effect. 38,39 To presence of the so-called Mullins effect, as illustrated in Figure
consistently extract a history-independent hyperelastic stress− 1(a). A few common characteristics of the Mullins effect
strain response for all samples, we focus on the last loading cycle
only up to strain levels beyond which the curve bends upward
due to prior maximum-strain history. Once we focus on this part
of the curve, we show that the response (for both aged and
unaged rubber) can be described well by the commonly
employed hyperelastic model due to Ogden. In fact, for our
material of interest, we show that the stress−strain response can
be modeled by just a single parameter, which can be directly
interpreted as the shear modulus of the material.
We note that the present work is intended as a more careful
theoretical treatment compared to our previous attempts to Figure 1. (a) Engineering stress vs strain for unaged TR-55 silicone
analyze similar data.40,41 The previous work, although a useful rubber displaying strong Mullins effect; (b) modeling the last loading
first step, employed the simple neo-Hookean model,16 which is curve from the left figure along with the best fit using the N = 1 Ogden
justified only to describe very small-strain behavior. The present incompressible hyperelastic model. Only the part of the curve
work accounts for the Mullins effect in a consistent manner, independent of strain history (i.e., prior to upward swing) is modeled.
employs a more accurate response model (i.e., Ogden), carefully
considers structural relaxation effects due to main-chain-scission include38,39,42,43 (1) significant softening upon the first
(Fricker’s transfer function), provides a range of estimates of the unloading cycle; (2) an increase in softening with an increase
ratio of Y-link versus H-link formation rates, and delves deeper in the maximum strain in the first cycle; (3) subsequent
into the radiation-induced chemistry, both at the network- unloading showing much less softening as long as the previous
topology level and at the molecular level.


maximum strain is not exceeded; and (4) a small but noticeable
permanent set at the end of the first unloading curve, which
EXPERIMENTAL PROCEDURE remains stable through subsequent cycles. The permanent set
Rectangular samples of commercial rubber TR-55 manufactured typically increases upon unloading from an increased maximum
by Dow Corning were used for the radiation aging experiments. strain, although in some cases, it can recover after a long resting
The material is primarily polydimethylsiloxane (PDMS) and time.44 There is an extensive literature study that attempts to tie
incorporates roughly 30 wt % of silica filler. Thin samples (∼1 the Mullins effect to various causes, primarily involving quasi-
mm thick) were stretched and held at a fixed elongation strain irreversible structural changes in filler configuration, chain-
while exposed to γ radiation from a Co-60 source (1.2 MeV, slippage, and network damage.42 In this work, we do not delve
∼0.1 Mrad/h dose rate), which provided a controllable into such details but rather attempt to focus on the last loading
degradation pathway. To avoid oxidative degradation, the cycle (see Figure 1(b)), which represents a quasi-steady-state
samples were irradiated under vacuum until the desired stress−strain response largely free of energy loss due to
cumulative dosage (D) was reached. In the following, we hysteresis or viscoelasticity.
express elongational strain in terms of the stretch ratio λ = 1 + ε, All radiation exposure experiments in this work were
where ε is the engineering strain, and we use λa to denote the performed after the material underwent four or five loading−
stretch ratio corresponding to the constant “aging strain” the unloading cycles; the aim was to minimize any coupling of
system is under when exposed to radiation. Mullins cycles with radiation effects.
10048 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B


pubs.acs.org/JPCB Article

OGDEN HYPERELASTIC AND TOBOLSKY W2net(λ1, λ 2 , λ3 , λa|μ1 , α1 , μ2 , α2)


2-NETWORK MODELS
2μ 2μ
As mentioned in the foregoing, we focus on the hyperelastic = 21 (λ1α1 + λ 2α1 + λ3α1 − 3) + 22
response of the material; that is, any hysteretic loss due to α1 α2
viscoelasticity is neglected in our analysis. With this caveat, we {(λ1λa−1)α2 + (λ 2λa1/2)α2 + (λ3λa1/2)α2 − 3}
chose a commonly employed hyperelastic strain energy model
due to Ogden, given by eq 1:45,46 − p(λ1λ 2λ3 − 1) (6)
N 2μj By using the same argument that led to eqs 3 and 4, it is
α α α
W (λ1 , λ 2 , λ3) = ∑ 2
[λ1 j + λ 2 j + λ3 j − 3] straightforward to obtain the following expression for Cauchy
j = 1 αj and Engineering stress (in direction 1) for the two-network
system (in the limit α1, α2 → 0):
− p(λ1λ 2λ3 − 1) (1)
σ2net = 3μ1ln λ1 + 3μ2 ln(λ1/λa); P2net = σ2net /λ1 (7)
where λ1, λ2, and λ3 are the eigenvalues of the deformation tensor
F, and p is a Lagrange multiplier that imposes the constraint that A primary effect analyzed in this work is an irreversible change in
the material is incompressible, i.e., detF = λ1λ2λ3 = 1. The length, as defined by the residual strain εs, or equivalently the
parameters {μj} are related to the shear modulus μ by the permanent stretch ratio λs = 1 + εs. The quantity λs is simply the
N
relation μ = ∑ j = 1 μj , and the parameters {αj} are exponents value of λ1 at which σ2net = 0, which from eq 7 is obtained as
follows:
that can be positive or negative.
If we focus on the stress−strain curve of the last loading cycle, λs = 1 + εs = λa1 − p (8)
we find that in all cases, it is described well by just a single term,
with
i.e., N = 1 Ogden model (eq 2):
p = μ1 /(μ1 + μ2 ) (9)
W1net(λ1, λ 2 , λ3| μ , α)
2μ Since the shear moduli μ1 and μ2 are functions of the dosage D
= 2 [λ1α + λ 2α + λ3α − 3] − p(λ1λ 2λ3 − 1) (as made clearer in the sections below), the residual strain εs is a
α (2)
function of both the aging strain and the cumulative radiation
where μ is the shear modulus, and α is a scalar exponent. The dosage. Occasionally, for brevity, we will make loose reference to
true (Cauchy) stress in the principal direction i is given by σi = λs and εs as the “permanent set.”
λi∂W1net/∂λi. Under a uniaxial engineering stretch εa along
direction 1, we have λ1 = 1 + εa and λ2 = λ3 = λ−1/2
1 (assuming an
isotropic and incompressible material). Using the boundary
■ RADIATION AGING UNDER ZERO STRAIN
Figure 2(a) displays the results for stress−strain response upon
condition σ2 = σ3 = 0, we obtain the following expression for the radiation exposure under zero strain for four different
Cauchy stress along direction 1:
2μ α
σ1net = (λ1 − λ1−α /2)
α (3)
where index subscript “1” in σ (to indicate the principal stretch
direction) has been omitted for simplicity. When eq 3 is fitted to
the last loading cycle of Figure 1, we find that for all our samples,
the best fit occurs for α → 0 in which limit eq 3 simplifies to
σ1net = 3μlnλ1 (4)
The corresponding engineering stress P is obtained simply as Figure 2. (a) Modeling stress−strain response (last loading cycle) of
σ1net/λ1, i.e., TR-55 rubber aged under various radiation dosages at zero strain; (b)
corresponding shear modulus as a function of radiation dosage,
P1net = 3μλ1−1ln λ1 (5) showing an approximately linear behavior. The vertical error bars in
figure (b) indicate standard error due to sample-to-sample variation.
In our experiments, the material is subjected to a cumulative
radiation dose while being under an elongation engineering
strain εa (i.e., stretch ratio λa = 1 + εa). Any radiation-induced cumulative dosages. Each curve represents the stress response
cross-link formation under this condition happens in equili- averaged over two samples. Figure 2(b) plots the corresponding
brium with the strained state, while main-chain-scission can mean shear modulus μ by fitting eq 5 to the stress−strain
happen both to the original cross-links formed under zero strain response. To eliminate dependence on strain history (in this
and to the new cross-links formed in the strained state. To model case, memory of the previously attained maximum strain), only
such effects, we need a hyperelastic strain energy expression for a the part of the response curve prior to upward swing was
two-network system, i.e., the original network (formed under modeled. From Figure 2(b), we see that the shear modulus
zero strain) and the induced network (formed under stretch increases linearly as a function of dosage D, i.e.,
ratio λa). For an unfilled network, Tobolsky argued that the net μ(D) = μ0 + kD (10)
effect of these two networks is linear and additive.47 Assuming
Tobolsky’s picture also holds true for a filled system, we express From the data in Figure 2(b), we obtained a mean value of 0.57
the net effect of these two networks as the following strain MPa for the intercept μ0, and a mean value of 0.074 MPa/Mrad
energy expression (eq 6): for the slope k. From uncertainty analysis, the 90% confidence
10049 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

interval for the marginal distribution of the intercept μ0 was pristine cross-link density. It is used in eq 13 below, where
(0.49,0.65), while that for the slope k was (0.069, 0.079), with a we develop the two-network constitutive model. In
negative correlation (−0.70) between the two quantities. particular, the case C′ > C (i.e., β < 1) can be interpreted
We note that all the analysis in this work is based on as an increase in the filler-enhancement factor induced by
experiments performed with our specific Co-60 radiation source radiation (relative to the filler-enhancement factor in
that has a fixed dosage rate (see the Experimental Procedure pristine cross-linked rubber), which could be attributed to
section). As long as the γ photon absorption events are spatially much stronger radical activation in silica fillers,48,49 as
and temporally separated, one expects the overall model and the compared to radiation-induced cross-linking within the
conclusions to be largely independent of the dosage rate and bulk elastomer.
depend only on the cumulative dosage D. However, since such
absorption events involve a complex cascade of chemical (3) When the material is radiation-aged while under a state of
processes involving multielectronic excitations, a more extensive strain, the resulting material behaves like an additive
experimental study with several different radiation sources is combination of two independent networks, following the
necessary to probe possible dose-rate dependence of our model model developed by Tobolsky47 for unfilled networks. In
parameters. fact, we have already used this assumption to derive eqs 8


and 9.
RADIATION AGING UNDER FINITE STRAIN: (4) If radiation induces main-chain-scission of the original
PERMANENT SET cross-links, there is a feedback effect in which a fraction of
Figure 3 (symbols) displays results for the observed residual the network 2 cross-links generated prior to that scission
strain εs for various values of dosage D and aging strain εa (note: event is effectively transferred to the original network (i.e.,
network 1). Such analysis was first carried out by Flory50
for a Gaussian-chain network and by Fricker51 for a
phantom network of Gaussian chains. Assuming a
phantom network model, in the following analysis, we
use Fricker’s formula to model such behavior, although
alternative approaches have recently been suggested in
the literature.52

■ RADIATION-INDUCED SCISSION AND


CROSS-LINKING RATES
Figure 3. Residual strain (variable εs in the text, expressed here in High-energy electromagnetic or electron-beam radiation is well
percent), a measure of the permanent set, in radiation-aged TR-55 known to produce permanent chemical changes in silicone
samples while being under different aging strain levels (εa, indicated in polymers.15,53−59 Two of the chemical changes particularly
the legend in percent). Points indicate sample-averaged measured relevant to change in network topology and mechanical moduli
values. The average experimental standard error is 2.2% (not shown for are scission and cross-linking processes. Scission processes could
clarity). Solid curves are the best model fit under the consideration of involve either main-chain-scission (i.e., breaking of Si−O
both H-linking and Y-linking processes. bonds) or side-chain-scission (i.e., breaking of Si−C or C−H
bonds). Any scission process typically produces radicals, which,
aging stretch ratio λa = 1 + εa). To model such data with eqs 8 in the absence of radical scavengers, could persist long enough
and 9, one needs to first determine the evolution of the shear until another active site with a radical is found and a new
stresses of the original and the induced networks, i.e., μ1 and μ2, chemical bond is formed by radical-pair annihilation. Using this
as a function of radiation dosage D. To make progress on this picture as a good approximation to reality (i.e., ignoring other
front, we borrow polymer network concepts developed for more complex processes, some of which are briefly mentioned in
unfilled rubber networks and make the following implicit the Insights from QM Modeling Section), let us introduce rates
assumptions about our filled system: of two broad classes of processes: (i) direct cross-linking of
(1) Despite all the structural complexities in a filled elastomer, active side chains belonging to two different load-bearing chains,
there is a linear relationship between the shear modulus with a rate of kxl processes per radiation dosage, and (ii) main-
(μ) and the chemical cross-link density (ν) in the bulk chain-scission followed by the bonding of a fraction (say, δ) of
elastomer for the pristine cross-linked polymer (i.e., prior the resulting active −Si and/or −SiO radicals with an active side-
to radiation exposure), i.e., μ = CkBTν, where C is a chain site of a different load-bearing chain, with a rate of ksci
material-dependent dimensionless constant that accounts scission processes per radiation dosage. Process (i) produces a
for effects like filler enhancement, departure from affine fourfold-coordinated junction, which we refer to below as an H-
behavior, entanglements, and so on. link, while process (ii) yields a threefold-coordinated junction,
(2) Cross-linking effects due to radiation is in general which we refer to below as a Y-link.
different at the filler surface than in the bulk polymer, Mechanical changes due to only H-link formation are simpler
to analyze, and we do that first in the following section.


and one expects a different proportionality constant for
the radiation-induced change in the modulus, i.e., Δμ =
C′kBTΔν, where Δμ is the radiation-induced change in CONSIDERING ONLY H-LINK PROCESSES
the shear modulus of the material and Δν is the radiation- In this section, we analyze how the formation of H-links only
induced change in cross-link density in the bulk (i.e., no main-chain-scission) affects the mechanical properties
elastomer. The ratio β = C/C′ is important in expressing of the system and the permanent set. In the analysis below, we
radiation-induced bulk cross-link density in terms of make use of our heuristic understanding that shear moduli of the
10050 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

two networks (μ1 and μ2) are proportional to the corresponding chain-scission of network 1 cross-links, eq 11 gets modified to
cross-link densities, even for the filled rubber material. the following:
Let us consider the material radiation-exposed under an aging
strain εa (i.e., stretch ratio λa = 1 + εa). Prior to radiation dμ1 /dD = −ksciφ(r ); dμ2 /dD = 4δksci − ksci(1 − φ(r ))
exposure, i.e., at D = 0, we have μ1 = μ0 and μ2 = 0 (where μ0 is
+ 2kxl = ksciφ(r ) + k (14)
the intercept in eq 10).
Let us now consider an H-linking process. In this process, two with
cross-links become four cross-links. The question is the strained
state status of the four new links. To address this point, we use k = (4δ − 1)ksci + 2kxl (15)
the important result from bead-spring simulations,60,61 which Equation 14 is derived based on the assumption that each
showed that such cross-linking processes under a state of strain monomer in the system has the same probability of being the site
do not alter the cross-link density of the original network. This of main-chain-scission.
simply translates to the result that if the two chains involved in However, as mentioned earlier, there is an additional effect
the H-linking process belong to the original network (i.e., that needs to be accounted for. This involves the fact that when
network 1), then after cross-linking two of the resulting four cross-linking is followed by scission of some of the network 1
chains effectively remain in the strained state of the original linkages, then a fraction f of network 2 cross-links effectively
network, while the other two chains belong to the induced becomes part of network 1.50,51,62 When this effective transfer of
network (i.e., network 2). The above argument can be easily cross-links is taken into account, eq 14 gets modified as follows:
extended to the case where either one or both the chain
segments involved in H-linking belong to network 2. In all such dμ1 /dD = −ksciφ(r ) + fμ2 ; dμ2 /dD = ksciφ(r ) + k − fμ2
cases, we see that H-linking leads to a net increase of two chains (16)
in network 2, without any change in the number of chains (and
hence the chain density) of network 1. Thus, we can express For a phantom network of Gaussian chains, Fricker showed51,62
radiation-induced changes in shear moduli μ1 and μ2 by the that f is approximately given by the ratio of the number of
following equations: network 1 cross-links that get scissioned and the sum of two
quantities, i.e., (i) number of network 2 cross-links that exist
dμ1 /dD = 0; dμ2 /dD = 2kxl (11) prior to the scission events and (ii) number of network 1 cross-
links of the pristine system (i.e., prior to radiation exposure). In
which can be easily solved as follows (eq 12): the present case, the number of network 1 cross-links that
μ1(D) = μ0 ; μ2 (D) = 2kxlD undergo scission between cumulative dosages D and D + dD is
(12)
proportional to βksciφ(r)dD, while the denominator is propor-
If this model is correct, then to have consistency with Figure 2, tional to μ0 + βμ2. Thus, according to the Fricker formula, in eq
we need to have 2kxl = k, where k is the slope in eq 10. From the 16, we identify the fraction f as follows:
90% CI values (0.069 and 0.079) for k, this implies that kxl
f ≈ βk sci φ(r )/(μ0 + βμ2 )
should be ∼0.037 MPa/Mrad with an error margin of 0.0025. (17)
However, we find that the value of kxl that fits the observed The initial conditions to eq 16 are
permanent set data (symbols in Figure 3) is ∼0.052 MPa/Mrad,
which is inconsistent with the confidence interval obtained from μ1(D = 0) = μ0 ; μ2 (D = 0) = 0 (18)
the mechanical modulus results of Figure 2. Additionally, even if
one uses the higher value of kxl, the fit is not as accurate as the where μ0 (in eq 18) and k (in eqs 14, 15, and 16) are the
solid curves in Figure 3, which is obtained from a model that intercept and slope, as previously defined in eq 10 with mean
incorporates both Y-linking and H-linking processes, as values of 0.57 MPa and 0.074 MPa/Mrad, respectively.
described in the section below. Determining the monomer fraction of network φ(r) in eqs 14,


16, and 17 is difficult. To build a practical working model, we use
CONSIDERING BOTH H-LINKING AND Y-LINKING a simple approximation based on the following argument. At
PROCESSES small times (i.e., small cumulative dosage D), we have r ≈ 1 and
φ(r) ≈ 1; i.e., almost all scissioned chains belong to network 1.
In the presence of both main-chain-scission (followed by Y-link However, since the average length of the induced chains
formation) and H-link processes, eq 11 needs to be modified. (following scission/cross-linking processes) is smaller than that
Let kxl and ksci be H-linking and main-chain-scission rates, of the original cross-links, for any D > 0, one expects r < 1 and
respectively. Let r be the fraction of cross-links belonging to φ(r) > r. There is no simple way to derive a closed-form
network 1, and φ(r) be the fraction of monomers belonging to expression for φ(r) as it likely depends on the molecular weight
network 1. From our model assumptions, we can relate r to the distribution of the material at D = 0. To avoid such complexities,
shear moduli μ1 and μ2 by the following equation: we assume a tractable approximation φ(r) ≈ rγ, where 0 < γ < 1;
r = {μ0 + β(μ1 − μ0 )}/{μ0 + β(μ1 + μ2 − μ0 )} (13)
this is a simple formula that satisfies the desired relations φ(1) =
1, and φ(r) > r for 0 < r < 1, and φ(r) → 0 for r → 0. Assuming
The above equation follows from our model that the pristine that initially (i.e., at low dosages D ≪ 1) the average lengths of
(i.e., preradiation) cross-link density is proportional to μ0/C, network 2 cross-links are roughly half of those of the network 1
and the radiation-induced change in network 1 cross-link density cross-links, we expect γ to be around 0.5.
is proportional to (μ1 − μ0)/C′, etc., and note that β = C/C′, by If we use φ(r) = rγ with γ = 0.5 in eq 16, numerically integrate
definition. Let us also assume that following main-chain-scission, to solve for μ1 and μ2 as a function of D, and then use eqs 8 and 9
only a fraction δ of the broken links forms active (i.e., load- to compute the residual strain εs, the best agreement with the
bearing) cross-links through the Y-linking process. Then, experimental data in Figure 3 happens for ksci ≈ 0.016 MPa/
without the consideration of any feedback effect due to main- Mrad and β ≈ 0.50. The corresponding computed results for εs
10051 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B


pubs.acs.org/JPCB Article

as a function of the cumulative dosage D (for various values of INSIGHTS FROM QM MODELING
aging strain levels) are indicated by the solid lines in Figure 3
From the discussion in the previous section, it is clear that the
showing good agreement between the observed and model
ratio rYH of Y-linking to H-linking rates depends strongly on the
results.
fraction (δ) of scissioned ends that form stable, load-bearing Y-
We point out that the above value of ksci corresponds to a
links. To estimate δ, one needs to understand the stability of
scission-to-cross-linking ratio of ksci/k = 0.22. This ratio is in
radicals formed following a scission event. In this section, we
good semi-quantitative agreement with experimental esti-
explore such radical stability by computing the energetics of
mates63 from Charlesby−Pinner analysis64 upon radiation
aging of the same filled rubber system (i.e., TR-55) as used in possible pathways using DFT-based first-principles QM
the present study. The optimized value of β ≈ 0.50 indicates that calculations65,66 with the CP2K code.67
radiation-induced filler-enhancement effect is a factor of β−1 ≈ 2 It is understood that there are number of complex reaction
larger than the filler-enhancement effect in pristine TR-55. channels in irradiated PDMS58 and silica fillers,48,49 including
We have experimented with other values of γ, e.g., between 0.5 ones not directly tied to cross-linking. In this section, we explore
and 0.7, and found the optimized values of ksci and β and the some representative pathways using clusters of atoms in the gas
quality of agreement in Figure 3 (in terms of RMSE deviation) phase. The DFT calculations reported below employed the
to be largely unchanged. This provides confidence in the generalized gradient approximation functional due to Perdew−
robustness of the model and the optimized parameters. Burke−Ernzerhof68 with Grimme D3 dispersion corrections69
Additionally, the model results for the permanent set are and the core electrons represented by the Goedecker−Teter−
independent of the parameter δ (the fraction of scissioned chain Hutter pseudopotentials.70 CP2K is a mixed Gaussian/plane-
ends that form load-bearing Y-links, as introduced just above eq wave code; we used a double-ζ plus polarization Gaussian basis
14) as long as it is chosen within reasonable physical bounds set (DZVP) and a 200 Ha plane wave energy cutoff for
(e.g., between 0.5 and 1.0), but it depends on the choice of k expansion of the charge density. The electronic structure was
(which we chose as 0.074 MPa/Mrad, following the fit to Figure evaluated with spin polarization at the Γ-point only. Reactions
2(b), see discussion following eq 10). On the other hand, from considered here involved migration of a single radical, so the
eq 15, the parameter kxl depends on the choice of δ. Additionally, spin multiplicity was taken as a doublet. We used the
the ratio of the number of Y-linking to H-linking events (rYH), nonperiodic Martyna−Tuckerman Poisson electrostatic solv-
which according to the present model is rYH = 2δksci/kxl, also has er71 and a cubic simulation cell with a side length of 30 Å, which
a strong dependence on δ as well. As we argue in the Insights is more than twice the maximum cluster size.
from QM Modeling Section below, we expect δ to be between Chemical reaction pathways were sampled using a con-
0.5 and 1. For δ= 0.5, 0.75, and 1.0, the Y-linking/H-linking ratio strained “bond scan” optimization approach in which the
rYH is 0.55, 1.1, and 2.5, respectively (using ksci = 0.016 and k = separation distance between two chosen atoms was incremented
0.074 MPa/Mrad). Thus, without knowing the parameter δ, we through set values. This was implemented in CP2K by applying
cannot reliably predict the Y-linking/H-linking ratio rYH. It is a large harmonic bias potential to a single collective variable
likely that δ varies significantly depending on the polymer defined as the separation distance between the two chosen
system, filler distribution, cross-linking history prior to atoms. Each step in the scan was optimized subject to the biased
irradiation, and many other factors. This could explain the collective variable using the BFGS algorithm with a force
lack of agreement in the literature on the relative importance of tolerance of 10−3 Ha/Bohr. These scans were performed starting
the Y-linking and H-linking rates.56,58 from a configuration optimized without constraints and were
Finally, using the above model, we can compute the stress− iterated on a uniformly spaced grid with a resolution of 0.1 Å.
strain response of the radiation-aged system relative to the new The initial configuration for each subsequent iteration of the
equilibrium. To this end, we define the longitudinal stretch ratio scan was taken as the optimized configuration from the previous
(in direction 1) in the aged system as λ′1 = λ1/λs, where λ1 is the iteration. It should be noted that while this approach effectively
stretch ratio in the unaged system and λs is the permanent explores the potential energy surface, it does not necessarily fully
stretch ratio defined previously. Then, eq 7 can be rewritten as optimize local minima or transition states that would be
eq 19: obtained without constraints.
′ = 3μ1ln(λ1′λs) + 3μ2 ln(λ1′λs /λa) = 3(μ1 + μ2 )ln λ1′ Radiation damage in PDMS is understood to result from the
σ2net
formation of radicals (see Figure 4) that arise from either main-
(19) chain-scission (species A and B) or excitation of the side groups
where the last step was simplified by using the result that σ2net as (species C and D).15 Recombination reactions involving
defined in eq 7 is 0 at λ1 = λs. In other words, the aged material is biradical pairs are energetically favorable and result in the
predicted to exhibit exactly the same stress−strain behavior as formation of H or Y cross-links. More specifically, recombina-
the unaged material with a modulus of μ1 + μ2. Adding the two tion reactions involving main-chain-scission products A and B
subequations of eq 16 and integrating, we obtain eq 20: can lead to either a healed chain (e.g., A + B) or the formation of
μ1 + μ2 = μ0 + kD Y-links (e.g., A + C). In contrast, reactions involving side group
(20)
radical species C and D can result in both types of cross-links. A
which is identical to eq 10. In other words, the stress−strain Y-link is formed if C or D reacts with A or B, whereas an H-link
behavior and shear modulus of the radiation-aged system is forms if C or D reacts with another radical of type C or D. If
predicted to be independent of the aging strain λa (with strain in biradical reactions are the only chemical pathways considered,
the aged sample being defined relative to its new equilibrium then the only cross-link-producing outcome of main-chain-
length that incorporates the permanent set). Such predicted scission is Y-link formation. However, given that the timescale
behavior largely agrees with our measured stress−strain curves for cross-linking is governed by chain dynamics in which two
of the aged samples, within the uncertainty of sample-to-sample radicals must migrate to reach each other and react, it is
variations. important to address the stability and lifetime of −Si (species A)
10052 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

is unstable even if formed. Based on the potential energy


landscape, this radical/H transfer reaction is plausible but not
likely to be a dominate pathway.
Thus, according to the above analysis, the Si terminus is
expected to be stable until it forms a Y-link with another side-
group monoradical.
The energy landscape explored in the third case is more
complicated than the previous two. There is a very small barrier
(5 kcal/mol) to proton transfer between a radical oxygen
terminus and methyl side group (4 → 5). This reaction produces
a silanol (Si−OH) terminus and methyl radical side group
identical to species C in Figure 4. Compared to state 4, there is
also a modest energetic driver (ΔE ∼ −5 kcal/mol) to reach
state 5. Thus, the radical oxygen terminus is substantially less
stable than the radical silicon terminus and can react to transfer
the radical to a nearby side group. The back-reaction barrier is
also comparatively modest (9 kcal/mol), indicating the
possibility for a non-negligible equilibrium concentration of
state 4. However, this equilibrium will be sensitive to entropic
contributions arising from chain motions that cause the silanol
and methyl radical side group to diffuse apart, which would favor
Figure 4. Initial radicals formed in irradiated PDMS and example Y-
state 5. Following the formation of a silanol, constrained
and H-type cross-links formed from biradical reactions. Species A and B
are formed as a pair from scission of the main chain, and species C and optimization uncovers another local minimum that corresponds
D are formed in separate events from excitation of the side groups. to a conformational change (5 → 5′) before a substantial 30+
Atom centers with an initial radical are rendered in color. kcal/mol barrier to further forward the chemical reaction. While
state 6 is a local minimum, it is unlikely that reaction 5 → 6
and/or −SiO radicals (species B) formed upon chain-scission, would occur through thermally activated chemistry.
especially given the high reactivity (and the resulting instability) Large barriers to hydrogen abstraction by a radical silicon
of most radicals. terminus are consistent with radiolysis experiments coupled to
To explore the stability of −Si and −SiO radicals, we used electron spin resonance spectroscopy, which found that silicon
DFT calculations to explore reaction energetics of such radicals radicals persist under cryogenic conditions.72 Recent cryogenic
with neighboring side groups in bulk PDMS. Three cases were radiolysis experiments sensitive to methyl radicals found an
considered, including (a) reaction between a radical silicon increase in the concentration of radical side group species C that
terminus (species A) and a methyl side group forming a cross- was correlated with the decay of free methyl radical (·CH3)
link; (b) reaction between a radical silicon terminus and the concentration,73 although these changes in concentration were
hydrogen on a methyl side group leading to an H-transfer; and not equivalent. One possible contributor to this trend is that free
(c) reaction between a radical oxygen terminus (species B) and a methyl radicals may scavenge hydrogen from nearby side groups
methyl side group. Constrained bond scan optimizations were producing species C.73 The buildup of species C concentration
used to sample these reactions and are shown in Figure 5, panels is also consistent with the proposed monoradical reactions in
(a), (b), and (c), respectively. Distinct chemical configurations Figure 5(b,c). Formation of Si−H bonds is consistent with
sampled in these reactions are shown to the right of each energy infrared and NMR measurements of irradiated PDMS
scan plot in the figure. materials.53,58
In the first case, there is a significant barrier to reaction Silanols are known to form in irradiated PDMS in the
between a silicon radical and a methyl side group to form a cross- presence of large concentrations of strong hydrogen donors such
link. The constrained optimization predicts an energetic penalty as alkylthiols.56 Small quantities of silanol products were found
of ∼35 kcal/mol to induce a reaction from state 1 to 2. This in hexamethyldisiloxane irradiated under an inert N2 atmos-
reaction produces a Y-link and a free hydrogen radical that are phere.74 Water is a common environmental antagonist known to
energetically unfavorable relative to the starting point (ΔE ∼ be coupled to silanol production. Experimental NMR measure-
+15 kcal/mol). The back-reaction barrier is also quite large ments found that thermal and radiative aging of PDMS in the
(∼20 kcal/mol), indicating that state 2 would be kinetically presence of 17O-labled water led to 17O uptake in silanol groups
stable once formed. Owing to both the large barrier and lack of and the main chain.75 Recent DFT calculations indicate that a
energetic drivers for reaction 1 → 2, it is reasonable to expect radical oxygen terminus can readily abstract hydrogen from
that the radical silicon terminus will not directly form a cross-link water, while the barrier for proton transfer between water and a
unless it interacts with a more reactive site, such as another radical silicon terminus is substantially higher.76 Water can
radical located on the polymer or an unsaturated bond found in a couple with radiation effects to produce silanols77 and is
vinyl or phenyl side group. predicted to destabilize highly strained chains leading to
The second energy profile shows an alternative reaction hydrolytic scission.78 Many mechanisms generate silanol
channel with a radical silicon terminus (1 → 3). Compared to functionalities under a range of conditions in PDMS, which
case (a), there is a more modest energy penalty (15 kcal/mol) to complicates its use as a signature to conclusively validate the
radical/H transfer resulting in a stable Si−H terminus and proposed mechanism through experiments.
methyl radical side group. Similar to case (a), there is no energy To summarize, the principal consequence for PDMS
driver for this reaction (ΔE ∼ +12 kcal/mol), but the back- radiolysis cross-linking chemistry inferred from Figure 5 is that
reaction barrier is very small (3 kcal/mol), indicating that state 3 main-chain-scission can promote both Y- and H-link formation.
10053 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 5. Constrained optimization scans of the potential energy surface for monoradical reactions between (a) radical silicon terminus and methyl
side group, (b) radical silicon terminus and hydrogen on methyl side group, and (c) radical oxygen terminus and methyl side group. Distinct chemical
configurations sampled in the optimization scans are indicated in the figure panels and shown at the bottom. Biased atoms are rendered in color.

This effect arises because radical side groups such as species C 0.75. In general, we expect δ to be between 0.5 and 1.0. A
are necessary intermediates for both types of cross-linking. comparatively large δ is consistent with the NMR radiolysis
While the radical silicon terminus produced from scission is experiments by Hill et al.58 that found Y-linking to be the
quite stable and would primarily engage in Y-linking through the dominant route to cross-linking in PDMS over more well-
biradical reaction, the radical oxygen terminus supports reaction
established H-linking mechanisms.53,55,74 Additional environ-
pathways for both Y- and H-linking through radical migration to
mental factors, such as large gas concentrations, may have other
side groups. The value of the fraction δ (see the beginning of this
section for definition) is governed primarily by the relative rates synergistic interactions with the radical silicon and oxygen
of the two types of linking processes resulting from the −SiO termini. High-throughput computational screening of reactivity
terminus. For instance, if these rates are nearly equal, and if we under many conditions using graph descriptors offers a
assume almost all cross-links to be load-bearing, we obtain δ≈ promising route to elucidate the complex and often inseparable
10054 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

couplings of different atomic-scale aging mechanisms in PDMS and Figure 3) is ∼22%, which is similar to earlier experimental
and other polymer materials.79 estimates using Charlseby−Pinner analysis on a similar
Before leaving this section, we would like to note that the system;63 (4) it is difficult to make a definite statement about
present analysis is relevant to radiation-induced scission/cross- the relative rates of Y-linking versus H-linking events: depending
linking reactions in the bulk of the elastomer only. This is in upon processing history and material details, the ratio of these
spirit of the previous sections where fillers are not explicitly two types of cross-linking can be more than or less than 1, as we
considered but assumed to influence mechanical properties and illustrated for various reasonable values of the parameter δ; (5)
radiation-induced net cross-linking through multiplicative the filler-enhancement factor of the mechanical modulus in the
constants C, C′. A more accurate analysis of radiation-induced presence of radiation is a factor of ∼2 stronger as compared to
chemical changes in filled rubber should take account of radical that in the absence of radiation; and (6) the −Si termini of
formation and cross-linking chemistry at the filler surface. Given scissioned chains should lead to primarily Y-links, while the
the significant presence of SiOH termini at the silica filler −SiO termini likely lead to significant fractions of both Y-links
surfaces, such analysis should be similar to Figure 5(c), although and −SiOH passivated ends.
effects of other molecules, e.g., H2O at the filler surface need to While the above results are interesting, we realize that some of
be explicitly modeled. In future studies, we have plans to explore the assumptions made are perhaps too simplistic. For instance,
such reactions in more detail. in the network theory for unfilled rubber, it is well known that

■ SUMMARY AND DISCUSSION


In this work, we show that radiation-induced permanent set and
physical entanglements lead to a different stress−strain response
than chemical cross-links,18−20 while in this work, we lump such
effect (along with filler enhancement) within a contribution
mechanical property changes of a filled silicone elastomer can be proportional to the chemical cross-link density. In fact, to
explained in terms of known concepts from network theory address this, we have explored the possibility of splitting the μ0
originally developed for unfilled rubber materials. The term into contributions from entanglement and chemical cross-
applicability of network theory concepts to filled systems hinges links present prior to radiation exposure and found that the best
on several assumptions, the most important of which involves a fits to the data occur when the contribution due to entanglement
direct proportionality between the shear modulus of the filled is zero. This is likely not due to the absence of entanglements but
system and the cross-link density of the underlying network, and rather reflects inconsistency of the isolated contribution of
a direct proportionality (with a different proportionality entanglements to the hyperelastic stress−strain response of the
constant) between the radiation-induced modulus increase filled system, at least for the data set analyzed here. Another
and the increase in the cross-link density in the bulk elastomeric aspect of the network theory neglected in this work is
network. We use such assumptions in conjunction with consideration of different functionalities of the cross-link
Tobolsky’s two-network model to quantitatively interpret the junctions. For instance, one could borrow ideas from the results
observed permanent set when filled-rubber samples are exposed of a phantom network model and weight the contribution of
to radiation while being under an applied tensile strain. For a each cross-link segment by a factor ≈(1 − ϕ−1 −1
1 − ϕ2 ) (where ϕ1
consistent treatment of stress−strain response under the and ϕ2 are functionalities of junctions at the two ends).
omnipresent Mullins effect, we used the last loading cycle and However, that significantly complicates the analysis and
fitted it using Ogden’s incompressible hyperelastic model. We warrants knowing the starting functionality distribution of the
found that for all our samples, the Ogden N = 1 model with zero pristine network prior to radiation exposure. In addition, it is not
α exponent led to the best fit, which led to a simple clear how a weighting factor such as the above gets modified in
interpretation of the single Ogden parameter as the shear the presence of fillers. During our analysis, we also made the tacit
modulus of the material. When used in conjunction with the assumption that the ratio of non-load-bearing to load-bearing
Tobolsky two-network model, such a framework was able to cross-link formation does not change with cumulative dosage.
quantitatively explain the observed permanent set data. Finally, Given that the average end-to-end distance of a network chain
through DFT calculations, we explored the stability of radicals at between its two end-junctions is proportional to the square root
the −Si and −SiO termini that are generated following a chain- of the number of monomers in the chain, a strain of 2.67 is not
scission event and its effects on the relative rates of Y-linking enough to strain any chemical bond in the siloxane system and
versus H-linking processes. We found that the -Si terminus thus does not alter the chemical reactivity of the system.
radicals are relatively stable and highly probable to create Y-links However, at much larger macroscopic strain levels, where a
(if not regenerating the prescissioned chain through recombi- significant number of network chains are stretched to near
nation with the −SiO terminus), while the −SiO terminus maximum-stretchable limits, one needs to account for stretches
radicals can potentially form either Y-links or get passivated in the chemical bonds in the system, which can result in a higher
through hydrogen abstraction (in which case, an H-link is likely susceptibility to radiation-induced chemistry.
to form). Finally, we would like to note that the rates of cross-linking
Our analysis led to several interesting results, including the (both Y and H) are kinetically governed by chain-diffusion rates
following: (1) the radiation-induced modulus enhancement and as well as factors such as the density of vinyl side groups,
the permanent set data cannot be explained by H-linking peroxide additives, and various radical scavengers present in the
processes alone, and incorporating contributions of Y-linking system. A more complete model needs to account for all the
above factors at an appropriate phenomenological level.


processes, which follow main-chain-scission, becomes neces-
sary; (2) the feedback effect due to structural relaxation
following main-chain-scission of network 1 cross-links can be AUTHOR INFORMATION
effectively modeled by Fricker’s transfer function, originally Corresponding Author
developed for an unfilled phantom network of Gaussian chains; A. Maiti − Lawrence Livermore National Laboratory,
(3) the predicted scission to cross-linking ratio that best fits the Livermore, California 94550, United States; orcid.org/
observed stress−strain and permanent set data (i.e., Figure 2(b) 0000-0003-0831-0700; Email: amaiti@llnl.gov
10055 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Authors (13) Brook, M.A. Silicon in a biological environment. In Silicon in


W. Small − Lawrence Livermore National Laboratory, Organic, Organometallic and Polymer Chemistry; Brook, M. A. (ed), John
Livermore, California 94550, United States Wiley and Sons: New York, NY, 2000, 459−479.
M. P. Kroonblawd − Lawrence Livermore National Laboratory, (14) Van Dyke, M. A.; Clarson, S. J.; Arshady, R. Silicone biomaterials,
in Introduction to Biomaterials; Arshady, R. (ed), Citus Books, London,
Livermore, California 94550, United States; orcid.org/
UK, 2003, 109−135.
0000-0002-5009-5998 (15) Palsule, A. S.; Clarson, S. J.; Widenhouse, C. W. Gamma
J. P. Lewicki − Lawrence Livermore National Laboratory, Irradiation of Silicones. J. Inorg. Organomet. Polym. 2008, 18, 207−221.
Livermore, California 94550, United States; orcid.org/ (16) Treloar, L. R. G. The Physics of Rubber Elasticity; Clarendon
0000-0002-2467-702X Press: Oxford, 1975.
N. Goldman − Lawrence Livermore National Laboratory, (17) Boyd, R. H.; Phillips, P. J. The Science of Polymer Molecules;
Livermore, California 94550, United States; orcid.org/ Cambridge University Press, 1996.
0000-0003-3052-2128 (18) Rubinstein, M.; Colby, R. H. Polymer Physics; Oxford University
T. S. Wilson − Lawrence Livermore National Laboratory, Press, 2003.
Livermore, California 94550, United States (19) Edwards, S. F.; Vilgis, T. The effect of entanglements in rubber
A. P. Saab − Lawrence Livermore National Laboratory, elasticity. Polymer 1986, 27, 483−492.
Livermore, California 94550, United States (20) Rubinstein, M.; Panyukov, S. Elasticity of Polymer Networks.
Macromolecules 2002, 35, 6670−6686.
Complete contact information is available at: (21) Arruda, E. M.; Boyce, M. C. A three-dimensional constitutive
https://pubs.acs.org/10.1021/acs.jpcb.1c04958 model for the large stretch behavior of rubber elastic materials. J. Mech.
Phys. Solids 1993, 41, 389−412.
Author Contributions (22) Bergström, J.; Boyce, M. C. Constitutive modeling of the large
§ strain time−dependent behavior of elastomers. J. Mech. Phys. Solids
W.S and M.P.K. contributed equally to this paper.
1998, 46, 931−954.
Notes (23) Itskov, M.; Ehret, A. E.; Dargazany, R. A Full-Network Rubber
The authors declare no competing financial interest. Elasticity Model based on Analytical Integration. Math. Mech. Solids

■ ACKNOWLEDGMENTS
This work was performed under the auspices of the U.S.
2010, 15, 655−671.
(24) Davidson, J. D.; Goulbourne, N. C. A nonaffine network model
for elastomers undergoing finite deformations. J. Mech. Phys. Solids
2013, 61, 1784−1797.
Department of Energy by Lawrence Livermore National (25) Ehret, A. E. On a molecular statistical basis for Ogden’s model of
Laboratory under Contract DE-AC52-07NA27344. M.P.K. rubber elasticity. J. Mech. Phys. Solids 2015, 78, 249−268.
thanks I-Feng W. Kuo for technical assistance with the CP2K (26) Li, Y.; Tang, S.; Kröger, M.; Liu, W. K. Molecular simulation
code. guided constitutive modeling on finite strain viscoelasticity of


elastomers. J. Mech. Phys. Solids 2016, 88, 204−226.
REFERENCES (27) Wang, M.-J. Effect of Polymer-Filler and Filler-Filler Interactions
on Dynamic Properties of Filled Vulcanizates. Rubber Chem. Tech.
(1) Mark, J. E.; Erman, B.; Eirich, F. R. Science and Technology of 1998, 71, 520−589.
Rubber; Academic Press: New York, 2005. (28) Vilgis, T. A.; Heinrich, G.; Klüppel, M. Reinforcement of Polymer
(2) Harper, C. A. Handbook of Plastics, Elastomers, & Composites; Nano-composites; Cambridge University Press: Cambridge, UK, 2009).
MacGraw-Hill: New York, 2002. (29) McLeish, T. Tube theory of entangled polymer dynamics. Adv.
(3) Ciesielski, A. An Introduction to Rubber Technology; Rapra Phys. 2002, 51, 1379−1527.
Technology Ltd.: UK, 1999. (30) Boyd, R. H.; Smith, G. D. Polymer Dynamics and Relaxation;
(4) Owen, M. J. Siloxane polymers, 1st ed. Englewood Cliffs, NJ: Cambridge University Press, 2007.
Prentice Hall, 1993. (31) Tang, S.; Greene, M. S.; Liu, W. K. Two-scale mechanism-based
(5) Morent, R.; De Geyter, N.; Axisa, F.; De Smet, N.; Gengembre, L.;
theory of nonlinear viscoelasticity. J. Mech. Phys. Solids 2012, 60, 199−
De Leersnyder, E.; Leys, C.; Vanfleteren, J.; Rymarczyk-Machal, M.;
226.
Schacht, E.; et al. Adhesion enhancement by a dielectric barrier
(32) Dealy, J. M.; Read, D. J.; Larson, R. G. Structure and Rheology of
discharge of PDMS used for flexible and stretchable electronics. J. Phys.
D Appl. Phys. 2007, 40, 7392−7401. Molten Polymers; Hanser Publishers: Munich, 2018.
(6) Andersson, L. H. U.; Hjertberg, T. Silicone Elastomers for (33) de Gennes, P. G. Reptation of a Polymer Chain in the Presence of
electronic applications. I. Analyses of the noncross-linked fractions. J. Fixed Obstacles. J. Chem. Phys. 1971, 55, 572−579.
Appl. Polym. Sci. 2003, 88, 2073−2081. (34) Doi, M.; Edwards, S. F. The Theory of Polymer Dynamics;
(7) Wu, K. H.; Chao, C. M.; Yeh, T. F.; Chang, T. C. Thermal Stability Clarendon Press: Oxford, 1986.
and Corrosion Resistance of Polysiloxane Coatings on 2024-T3 and (35) Ogden, R. W. Nonlinear Elastic Deformations; Dover
6061-T6 Aluminum Alloy. Surf. Coat. Technol. 2007, 201, 5782−5788. Publications: New York, 1984.
(8) Buyl, F. Silicone Sealants and Structural Adhesives. Int. J. Adhes. (36) Marckmann, G.; Verron, E. Comparison of Hyperelastic Models
2001, 21, 411−422. for Rubber-Like Materials. Rubber Chem. Tech. 2006, 79, 835−858.
(9) Ikada, Y. Surface modification of polymers for medical (37) Holzapfel, G. A. Nonlinear Solid Mechanics; John Wiley and Sons:
applications. Biomaterials 1994, 15, 725−736. Chichester, UK, 2000.
(10) Abbasi, F.; Mirzadeh, H.; Katbab, A. Bulk and surface (38) Mullins, L. Effect of Stretching on the Properties of Rubber.
modification of silicone rubber for biomedical applications. Polym. Rubber Chem. Technol. 1948, 21, 281−300.
Int. 2002, 51, 882−888. (39) Mullins, L. Softening of Rubber by Deformation. Rubber Chem.
(11) Colas, A.; Curtis, J. Silicone biomaterials: history, chemistry and Technol. 1969, 42, 339−362.
major applications of silicones, printed in Biomaterials Science: An (40) Maiti, A.; Weisgraber, T. H.; Dinh, L. N.; Gee, R. H.; Wilson, T.
Introduction to Materials in Medicine; 2nd edn., Elsevier Publications, S.; Chinn, S. C.; Maxwell, R. S. Controlled manipulation of elastomers
1996 80−86. with radiation: Insights from multiquantum nuclear-magnetic-
(12) Arkles, B. Look what you can make out of silicones (Biomedical resonance data and mechanical measurements. Phys. Rev. E 2011, 83,
applications of silicones). Chem 1983, 13, 542−555. No. 031802.

10056 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(41) Maiti, A.; Weisgraber, T. H.; Gee, R. H.; Small, W.; Alviso, C. T.; (64) Charlesby, A.; Pinner, S. H. Analysis of the solubility behaviour of
Chinn, S. C.; Maxwell, R. S. Radiation-induced mechanical property irradiated polyethylene and other polymers. Proc. R. Soc. A 1959, 249,
changes in filled rubber. Phys. Rev. E 2011, 83, No. 062801. 367−386.
(42) Diani, J.; Fayolle, B.; Gilormini, P. A review on the Mullins effect. (65) Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys.
Eur. Polym. J. 2009, 45, 601−612. Ther. Rev. 1964, 136, B864−B871.
(43) Maiti, A.; Small, W.; Gee, R. H.; Weisgraber, T. H.; Chinn, S. C.; (66) Kohn, W.; Sham, L. Self-Consistent Equations Including
Wilson, R. S.; Maxwell, R. S. Mullins effect in a filled elastomer under Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133−A1138.
uniaxial tension. Phys. Rev. E 2014, 89, No. 012602. (67) Kühne, T. D.; Iannuzzi, M.; Del Ben, M.; Rybkin, V. V.; Seewald,
(44) Itskov, M.; Haberstroh, E.; Ehret, A. E.; Voehringer, M. C. P.; Stein, F.; Laino, T.; Khaliullin, R. Z.; Schütt, O.; Schiffmann, F.; et al.
Experimental observation of the deformation induced anisotropy of the CP2K: An electronic structure and molecular dynamics software
Mullins effect in rubber. KGK, Kautsch. Gummi Kunstst. 2006, 59, 93− package - Quickstep: Efficient and accurate electronic structure
calculations. J. Chem. Phys. 2020, 152, 194103.
96.
(68) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
(45) Ogden, R. W. Large deformation isotropic elasticity − on the
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
correlation of theory and experiment for incompressible rubberlike
(69) Grimmea, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and
solids. Proc. R. Soc. London A 1972, 326, 565−584. accurate ab initio parametrization of density functional dispersion
(46) Wikipedia page: https://en.wikipedia.org/wiki/Ogden_ correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,
(hyperelastic_model). 132, 154104.
(47) Tobolsky, A. V. Properties and Structure of Polymers; Wiley: New (70) Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space
York, 1960. Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703−1710.
(48) Dondi, D.; Buttafava, A.; Zeffiro, A.; Conzatti, L.; Faucitano, A. (71) Martyna, G. J.; Tuckerman, M. E. A reciprocal space based
The role of silica in radiation induced grafting and crosslinking of silica/ method for treating long range interactions in ab initio and force-field-
elastomers blends. Polymer 2012, 53, 4579−4584. based calculations in clusters. J. Chem. Phys. 1999, 110, 2810−2821.
(49) Dondi, D.; Buttafava, A.; Zeffiro, A.; Bracco, S.; Sozzani, P.; (72) Ormerod, M. G.; Charlesby, A. The radiation chemistry of some
Faucitano, A. Reaction Mechanisms in Irradiated, Precipitated, and polysiloxanes: An electron spin resonance study. Polymer 1963, 4, 459−
Mesoporous Silica. J. Phys. Chem. A 2013, 117, 3304−3318. 470.
(50) Flory, P. J. Elasticity of polymer networks cross-linked in states of (73) Pankratova, L. N.; Saenko, E. V.; Bugaenko, V. L.; Makhlyarchuk,
strain. Trans. Faraday Soc. 1960, 56, 722−743. V. V. Kinetics of decay of methyl radicals in irradiated dimethylsila-
(51) Fricker, H. S. The effects on rubber elasticity of the addition and nediol and polymethylsiloxanes. Mendeleev Commun. 2011, 21, 157−
scission of cross-links under strain. Proc. R. Soc. London A 1973, 335, 159.
267−287. (74) Dewhurst, H. A.; St. Pierre, L. E. Radiation chemistry of
(52) Yu, Y.; Yan, S.; Fang, Y.; He, Q.; Wang, H.; Qiu, Y.; Wan, Q. hexamethyldisiloxane, a polydimethyl-siloxane model. J. Phys. Chem.
Theoretical Model of Polymer Network Chain Formation under Strain. 1960, 64, 1063−1065.
ACS Omega 2018, 3, 15615−15622. (75) Patel, M.; Morrell, P.; Cunningham, J.; Khan, N.; Maxwell, R. S.;
(53) Charlesby, A. Changes in silicone polymeric fluids due to high- Chinn, S. C. Complexities associated with moisture in foamed
energy radiation. Proc. Roy. Soc. A 1955, 230, 120−135. polysiloxane composites. Polym. Degrad. Stab. 2008, 93, 513−519.
(54) Bueche, A. M. An investigation of the theory of rubber elasticity (76) Wang, P.-C.; Yang, N.; Liu, D.; Qin, Z.-M.; An, Y.; Chen, H.-B.
using irradiated polydimethylsiloxanes. J. Polym. Sci. A 1956, 19, 297− Coupling effects of gamma irradiation and absorbed moisture on
306. silicone foam. Mater. Des. 2020, 195, No. 108998.
(55) Miller, A. A. Radiation chemistry of Polydimethylsiloxane: I. (77) Kroonblawd, M. P.; Goldman, N.; Lewicki, J. P. Chemical
Degradation Pathways in Siloxane Polymers Following Phenyl
Cross-linking and gas yields. J. Am. Chem. Soc. 1960, 82, 3519−3523.
Excitations. J. Phys. Chem. B 2018, 122, 12201−12210.
(56) Miller, A. A. Radiation chemistry of Polydimethylsiloxane: II.
(78) Kroonblawd, M. P.; Goldman, N.; Lewicki, J. P. Anisotropic
Effect of Additives. J. Am. Chem. Soc. 1961, 83, 31−36.
Hydrolysis Susceptibility in Deformed Polydimethylsiloxanes. J. Phys.
(57) Hill, D. J. T.; Preston, C. M. L.; Whittaker, A. K.; Hunt, S. M. The
Chem. B 2019, 123, 7926−7935.
radiation chemistry of poly(dimethyl-siloxane). Macromol. Symp. 2000, (79) Kroonblawd, M. P.; Goldman, N.; Maiti, A.; Lewicki, J. P. A
195, 95−102. Quantum-Based Approach to Predict Primary Radiation Damage in
(58) Hill, D. J. T.; Preston, C. M. L.; Whittaker, A. K. NMR study of Polymeric Networks. J. Chem. Theory Comput. 2021, 17, 463−473.
the gamma radiolysis of poly(dimethyl siloxane) under vacuum at 303
K. Polymer 2002, 43, 1051−1059.
(59) Hill, D. J. T.; Preston, C. M. L.; Salisbury, D. J.; Whittaker, A. K.
Molecular weight changes and scission and cross-linking in poly
(dimethyl siloxane) on gamma radiolysis. Rad. Phys. Chem. 2001, 62,
11−17.
(60) Rottach, D. R.; Curro, J. G.; Grest, G. S.; Thomson, A. P. Effect of
strain history on stress and permanent set in cross-linking networks: A
molecular dynamics study. Macromolecules 2004, 37, 5468−5473.
(61) Rottach, D. R.; Curro, J. G.; Budzien, J.; Grest, G. S.; Svaneborg,
C.; Everaers, R. Permanent set of cross-linking networks: Comparison
of theory with molecular dynamics simulations. Macromolecules 2006,
39, 5521−5530.
(62) Rottach, D. R.; Curro, J. G.; Budzien, J.; Grest, G. S.; Svaneborg,
C.; Everaers, R. Molecular dynamics simulations of polymer networks
undergoing sequential cross-linking and scission reactions. Macro-
molecules 2007, 40, 131−139.
(63) Maxwell, R. S.; Chinn, S. C.; Alviso, C. T.; Harvey, C. A.; Giuliani,
J. R.; Wilson, T. S.; Cohenour, R. Quantification of radiation induced
cross-linking in a commercial, toughened silicone rubber, TR55 by 1H
MQ-NMR. Polym. Degrad. Stab. 2009, 94, 456−464.

10057 https://doi.org/10.1021/acs.jpcb.1c04958
J. Phys. Chem. B 2021, 125, 10047−10057

You might also like