Barrat 2010

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

View Article Online / Journal Homepage / Table of Contents for this issue

REVIEW www.rsc.org/softmatter | Soft Matter

Molecular dynamics simulations of glassy polymers


Jean-Louis Barrat,a J€
org Baschnagelb and Alexey Lyulincd
Received 22nd December 2009, Accepted 10th February 2010
DOI: 10.1039/b927044b

We review recent results from computer simulation studies of polymer glasses, from the chain dynamics
around the glass transition temperature Tg to the mechanical behaviour below Tg. These results clearly
show that modern computer simulations are able to address and give clear answers to some important
issues in the field, in spite of the obvious limitations in terms of length and time scales. In the present
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

review we discuss the cooling rate effects, and the dynamic slowing down of different relaxation
processes when approaching Tg for both model and chemistry-specific polymer glasses. The impact of
geometric confinement on the glass transition is discussed in detail. We also show that computer
simulations are very useful tools to study structure and mechanical response of glassy polymers. The
influence of large deformations on mechanical behaviour of polymer glasses in general, and strain
hardening effect in particular are reviewed. Finally, we suggest some directions for future research,
which we believe will be soon within the capabilities of state of the art computer simulations, and
correspond to problems of fundamental interest.

1 Introduction pervaded by other chains in the melt.4 The strong interpenetra-


tion of the chains has important consequences for the properties
A polymer is a macromolecular chain resulting from the of the melt. For instance, intrachain excluded volume interac-
connection of a large number of monomeric units. The number tions, which swell the polymer in dilute solution,1,4 are almost
of monomers (the chain length N) typically ranges between 103 screened so that a chain behaves on large length scales approx-
and 105 in experimental studies of polymer melts.1 Recently, this imately as a random coil.3 Furthermore, chain interpenetration
range of chain lengths has also become accessible in simula- also impacts the polymer dynamics by creating a temporary
tions.2,3 Such a long chain has an open structure which is strongly network of entanglements. Entanglements strongly slow down
the chain relaxation and make the melt viscoelastic already at
high temperature.5,6
a
Universite de Lyon, Univ. Lyon I, Laboratoire de Physique de la Mati
ere On cooling toward low temperature the polymer melt even-
Condens ee et des Nanostructures, CNRS, UMR 5586, 43 Bvd. du 11 Nov.
1918, 69622 Villeurbanne Cedex, France tually transforms into a solid. Polymeric solids are either semi-
b
Institut Charles Sadron, Universite de Strasbourg, CNRS UPR 22, 23 Rue crystalline or glassy (Fig. 1).9 In the semicrystalline state
du Loess, BP 84047, 67034 Strasbourg Cedex 2, France amorphous regions are intercalated between crystalline lamellar
c
Group Theory of Polymers and Soft Matter, Eindhoven Polymer sheets. The sheets consist of chains which are folded back on
Laboratories, Technische Universiteit Eindhoven, P.O. Box 513, 5600
MB Eindhoven, The Netherlands
themselves so that chain sections align parallel to each other.9,10
d
Dutch Polymer Institute, P.O. Box 902, 5600 AX Eindhoven, The The ability to form crystals crucially depends on the polymer
Netherlands microstructure. Only chains with regular configurations, e.g.

Jean-Louis Barrat obtained his J€org Baschnagel received his


PhD in 1987 from the University Habilitation in physics in 1999
of Paris. He has worked in from the University of Mainz
Munich, Santa Barbara and (Germany) and was appointed
Lyon, where he holds a professor professor in the same year at
position since 1994 and directs the University of Strasbourg
the condensed matter physics (France), where he has stayed
laboratory. He applies statis- ever since. He is heading the
tical physics and simulation to ‘‘Theory and Simulation group’’
various problems in soft matter at the Institut Charles Sadron in
and materials science (glassy Strasbourg. His research inter-
systems, confined fluids, poly- ests are the physics of polymers,
mers). He is a senior member currently in particular, the
Jean-Louis Barrat of ‘‘Institut Universitaire de J€org Baschnagel dynamics of polymers, polymer
France’’. glass transition, and the proper-
ties of thin polymer films.

3430 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

transition, solid polymers are also integral components in many


modern applications.18 For instance, glassy polymeric materials
are appreciated because of their spectacular mechanical proper-
ties.19–21 Instead of failing abruptly when subject to strong
deformations, some polymers, such as polycarbonate, may
harden for large strains, leading to a tough mechanical response.
A microscopic understanding, elucidating the structure–property
relationship, of this behaviour is still elusive. A further example is
provided by thin polymer films, extensively used in technological
applications as protective coatings, optical coatings, adhesives,
Fig. 1 Volume–temperature diagram of a coarse-grained model for
etc. In the ongoing quest for progressively smaller structures and
poly(vinyl alcohol).7,8 The volume per monomer (v) and the temperature
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

(T) are given in Lennard-Jones units. In the liquid phase the chains have
devices these films may attain nanoscopic dimensions, where
random-coil-like configurations and the structure of the melt is amor- deviations from the bulk behaviour should be expected.22 Indeed,
phous. The amorphous structure is preserved on rapid cooling and the many recent studies suggest that the Tg of thin films is shifted by
melt eventually undergoes a glass transition at low T. For slow cooling spatial confinement, a striking observation which is not well
the melt transforms into a semicrystalline material in which sections of understood yet.23,24
folded chains order in lamellar sheets that coexist with amorphous This wide array of challenging questions—from the glass
regions. When the crystal is slowly heated up (dashed line), it melts at transition to the impact of external stimuli—bestows the theo-
a higher T than the temperature where crystallization occurs. This retical understanding of glassy polymers with particular signifi-
hysteresis is typical of first-order phase transitions. cance. Molecular simulations can contribute to this research.
Over the past two decades, computational models and methods
isotatic or syndiotatic orientations of the sidegroups or chains have been developed for simulating these glassy systems. The
without sidegroups, can fold into crystalline lamellae. However, progress made in the field has regularly been the subject of
even in these favorable cases full crystallization is hard to topical reviews. For instance, detailed reports of chemically
achieve.10 realistic modeling approaches may be found in ref. 25–27, while
This intrinsic difficulty to crystallize favors glass formation.11–13 work on coarse-grained models is reviewed in ref. 27–32. The
Polymer melts either can be easily supercooled (Fig. 1) or, due to purpose of the present article is to give a brief account of recent
the irregular chain structure, do not crystallize at all. An example work.
for the former case is bisphenol-A polycarbonate (PC),14 while We begin our survey by a short introduction to the modeling
examples for the latter involve homopolymers with atactic (bulky) of polymers in simulations (section 2). The glass transition
sidegroups, such as atactic polystyrene (PS), or random copoly- temperature naturally splits the following discussion into two
mers, such as cis-trans-1,4-polybutadiene (PBD). These poly- parts, a part devoted to the properties of the model glass formers
meric glass formers share with other (intermediate and fragile) above Tg (section 3) and a part addressing sub-Tg phenomena
glass-forming liquids the key characteristic feature of the glass (section 4). In section 3, our discussion is mainly concerned with
transition; that is, little change of the amorphous structure, but the dynamics of weakly supercooled polymer melts and the
a huge non-Arrhenius-like slowing down of the dynamics on impact of spatial confinement on their behaviour. Section 4
cooling toward the glass transition temperature Tg.12,13,15 focuses on the response of polymeric glasses to both weak and
Understanding the molecular origin of this disproportionate large external deformation, where respectively the linear
behaviour represents a great scientific challenge.12,13,15–17 In mechanical behaviour of the glassy melt and its approach to
addition to this fundamental interest in the study of the glass material failure will be explored. The article concludes with
a short summary of the presented results and an outlook on
possible future research directions.
Alexey Lyulin obtained his PhD
in 1992 from the Institute of 2 Computer simulations: models and computational
Macromolecular Compounds of
aspects
Russian Academy of Sciences,
on computer simulation of poly- Molecular simulations of glass-forming polymers face the
mer liquid crystals. From the important challenge that the structure and dynamics of these
year 2000 he is a senior systems are governed by a large spread of length and time
researcher of the Dutch Polymer scales.33,34 The relevant length scales extend from the atomic
Institute and assistant professor diameter (1010 m) to the chain dimension (107 m for N 
at the Eindhoven University 104). The spread of time scales is even larger; it ranges from bond
of Technology. His research vibrations (1013 s) to the slow structural relaxation close to Tg
focuses mainly on computational (102 s).
studies of mechanical properties From the inspection of these scales, it is clear that simulation
Alexey Lyulin of polymer glasses, non-new- approaches have some limitations and also require simplifica-
tonian rheology of hyper- tions. An obvious limitation are the accessible time scales. With
branched polymers and stability the currently available computer power the longest time that can
and toxicity of complexes of charged dendrimers. be reached in molecular dynamics (MD) simulations is roughly

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3431
View Article Online

a few ms. This time scale is 8 orders of magnitude smaller than the comparison between simulation and experiment, should involve
relaxation time at the experimental Tg, implying that the simu- information about both structural and dynamic properties.26
lated Tg is shifted by about 25 degrees to higher temperature (T)
relative to experiments (where we used the rule of thumb: 1 Generic models
decade in time z 3 K). The simplifications that are currently
necessary concern the simulation models which are obtained by Atomistic simulations are ideally suited to study specific poly-
some kind of coarse-graining procedure, designed to eliminate mers, including their glass transition and the properties of the
(some) fast degrees of freedom by incorporating them in effective glassy state. On the other hand, the strong increase of the
potentials. This model-building step can take different levels of relaxation time, which eventually leads to vitrification on cooling
complexity,2,35–37 but roughly speaking, there are two families, through Tg, is common to all glass-forming polymers, irrespec-
atomistic models and coarse-grained models. In the following, tive of their chemical composition and architecture. This
universal aspect suggests to employ simplified simulation models
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

we briefly introduce these models and discuss their strengths and


weaknesses as we go. which only retain generic features of a polymer. Various such
Before doing so, however, we want to point out some general generic models—on a lattice28,30,31 or in the spatial con-
appealing features of the simulations to balance the reservations tinuum31,32—have been studied for glass-forming polymers. In
expressed above. The simulations offer full control over the the continuum, the simplest model are highly flexible bead-spring
perfectly defined system under study. For instance, it is possible models, where spherical monomers are tethered together by
to vary in a systematic manner parameters, such as chain stiff- springs and have nonbonded, Lennard-Jones (LJ) interac-
ness, chain length, etc., while keeping all other defining proper- tions.31,32 More chemical realism can be introduced by making
ties of the model, so as to single out the impact of each parameter the chains semiflexible, through the addition of a bond-angle
on the properties of the system. Furthermore, the simulations potential58 and, possibly also, of a torsional potential.59–61 Due to
enable one to study local properties or to explore correlations their simplicity, these generic models allow for an efficient
which are hard (or even impossible) to access in experiments. simulation. This is important if one wants to vary the cooling
This provides microscopic insight and a means to test theoretical rate62 or the strain rate27 over decades, to explore systematically
concepts. Examples from the recent literature involve investiga- the impact of model parameters, such as chain length or chain
tions of the potential energy landscape of glass-forming mate- flexibility,58–61 or to obtain good statistics for comparison with
rials,38,39 of single-molecule diffusion in polymeric matrices40,41 or theory.63 Computational expedience is also important for
of fracture in glassy polymers.27 exploratory studies of more complex systems, such as inhomo-
geneous polymer systems (e.g., polymer films, nanocomposites,
semi-crystalline polymers) or glassy polymer mixtures (e.g.,
dynamically asymmetric polymer-polymer mixtures or polymer-
Atomistic models solvent systems). Some examples will be discussed in Section 3
Atomistic models replace electronic degrees of freedom by force and 4.
fields, i.e., empirical potentials for bond-length, bond-angle,
torsional-angle, and nonbonded interactions, whose parameters Hierarchical models
are determined from quantum-chemical calculations and exper- For generic models, the gain in the accessible length and time
iments.26,35 During the past decades, force fields have been scales is obtained at the expense of a loss of correlation to the
developed for both explicit atom (EA) models, treating every atomistic conformation of the polymer. For many problems in
atom present in nature as a separate interaction site, and united materials research, this loss is undesirable. Therefore, much
atom (UA) models which represent a small number of real atoms research efforts currently goes into the development of hierar-
(e.g., CH2 or CH3) by a single site. The reduction of the number chical approaches consisting of interconnected levels of modeling
of interaction sites in the UA model has the computational (atomistic, generic, macroscopic).2,37,64,65 The idea is that each
advantage of allowing longer simulation times. With a time step level treats phenomena on its specific length and time scales and
of 1015 s a few thousand united atoms can be simulated over then passes on the results as input to the next, more coarse-
several 100 ns, about an order of magnitude longer than an EA grained level, until the desired materials properties can be pre-
simulation of comparable system size. dicted. Such multiscale simulation methods represent a powerful
Both EA and UA models were employed to study the prop- approach whose potential for the modeling of glassy polymers is
erties of glassy polymers (see e.g. refs. 25–27 for reviews). Recent beginning to be explored (see e.g. ref. 66).
examples include polyisoprene (EA model42,43), atactic PS (UA
model44–48), bisphenol-A PC (UA model47,48) 1,4-PBD (UA and
Remarks on simulation methods
EA models26,49–55), and poly(ethylene oxide) or atactic poly-
(propylene oxide) (EA models56). Certainly, the main strength of Sections 3 and 4 will present results from molecular dynamics
these modeling efforts is that the simulation results allow for (MD) simulations, a numerical method to integrate the classical
a direct comparison with experiments. In some cases, such equations of motion for a many-body system in a given
a comparison is possible with commercial software packages,55 thermodynamic ensemble.67,68 Therefore, MD is the natural
while a careful fine-tuning of the force field is often required.49,53 simulation technique to address dynamical problems, such as the
Experience from those modeling approaches—not only for glass transition. However, the realistic MD dynamics carries an
polymers but also e.g. for amorphous SiO231,57—suggests that the obvious price: the equilibration time for the system under
design of a chemically realistic model, aiming at a parameter-free consideration—for instance, for a long-chain polymer melt close

3432 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

to its Tg—will exceed the maximum time of a few microseconds


one is currently able to simulate.
Here Monte Carlo (MC) techniques may provide a promising
avenue because of large freedom to design MC moves.67,69 The
hope is to find an efficient algorithm allowing one to decorrelate
the configurations of glassy polymer melts rapidly. For long-
chain polymer melts, this demand on the algorithm implies,
already at high T, that the MC move should be nonlocal, i.e., it
should modify the chain conformation at large scales, and it
should not require empty space because the melt is a dense liquid.
A promising algorithm satisfying these requirements employs
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

double-bridging moves which alter the connectivity between two Fig. 2 Panel (a): Glass transition temperature (Tg in Kelvin) versus
neighbouring chains while preserving the monodispersity of the cooling rate (GT in K/ps) for an atomistic model of atactic PS.46 The
chains (for recent reviews see e.g. refs. 2 and 3). Such a connec- simulation box contains 8 chains with N ¼ 80 monomers each. The solid
tivity-altering move drastically changes the conformation of the line shows a fit to eqn (1) with T0 ¼ 371 K, B ¼ 110 K, and A ¼ 0.23 ps/K.
two chains involved and thus relaxes the length scales on the Panel (b): Tg versus chain length (N) for a fully flexible and a semiflexible
order of the chain dimension efficiently. However, it does not (with angular potential) bead-spring model.75 The simulation box
contains at least 192 chains and the cooling rate is GT ¼ 2  105. The
alter the local packing of the monomers. An inherent hazard of
solid lines are results of a fit to eqn (2) with TgN ¼ 0.432, K ¼ 0.145
the algorithm therefore is that, if the move is attempted repeat-
(flexible model) and TgN ¼ 0.525, K ¼ 0.264 (semiflexible model). All data
edly, a successful double-bridging event is likely to annihilate one are given in Lennard-Jones units for the bead–spring models.
of its predecessors by performing the transition between two
chains in the reverse direction. To avoid this inefficiency the
nonlocal chain updating should be complemented by a move Similar extrapolation procedures are also applied in simula-
which efficiently mixes up the local structure of the melt. At low tion studies.26,30,46,47,62,75–79 The resulting Tg values have depen-
T, efficient relaxation of the liquid structure calls for a method dences comparable to experimental ones despite the much larger
which alleviates the glassy slowing down in general. Thus, any cooling rate employed—typically 1012 K min1 in simulations
algorithm achieving this aim in nonpolymeric liquids should also and 10 K min1 in experiments. For instance, Soldera and
accelerate the equilibration of glassy polymer melts, provided Metatla find a linear relationship between numerical and
that it can be generalized to respect chain connectivity. At experimental Tg values from atomistic simulations of various
present, no technique has been established to solve this problem. vinylic polymers.76 Fig. 2 reveals that Tg decreases nonlinearly
However, possible candidates could be parallel tempering,70–72 with the logarithm of the cooling rate (GT)26,30,46,62
Wang–Landau sampling69 or variants thereof,73 or transition
path sampling methods.74 B
Tg ðGT Þ ¼ Tg0  (1)
lnðAGT Þ
in accordance with experimental observation.80 Also in agree-
3 Glass transition and properties of the supercooled
ment with experiment,11,81–84 Tg increases with chain rigidity and
polymer liquid chain length.46,59,85 The chain length dependence can be fitted to
3.1 Bulk properties the empirical Fox–Flory equation
K
3.1.1 Glass transition temperature. In polymer melts, the Tg ðNÞ ¼ TgN  (2)
transition from the glass to the liquid is accompanied by a strong N
decrease in the shear modulus, typically of three to four orders of which usually describes experimental data well (if the molecular
magnitude.21 It is thus clear that the glass transition temperature weight is not too small11), although other forms have recently
is a characteristic of high engineering relevance. Classical been discussed in the literature83,84 and can be rationalized
methods for the experimental determination of Tg are calori- theoretically.86
metry and dilatometry. Both methods hint at the kinetic features
of the glass transition. The transition occurs when the relaxation 3.1.2 Dynamics of the supercooled melt. The precursor of the
time for volume recovery (dilatometry) or enthalpy recovery glass transition is the strong slowing down of structural relaxa-
(calorimetry) becomes longer than the time scale of the experi- tion processes on approach to the transition from the liquid. This
ment (i.e., than the cooling or heating rate).11–13 Upon cooling the dynamical feature is a hallmark of strongly interacting disor-
polymer liquid falls out of equilibrium close to Tg and freezes dered matter, including dense colloids,87 granular materials,88
below Tg in a glassy state, the properties of which depend on the and supercooled liquids.12,16 In simulated polymer melts, it is
details of the cooling process and tend to age physically during observed in all dynamic correlation functions, e.g., in dynamic
further isothermal equilibration.11–13 The glass transition structure factors, conformational correlation functions, or
temperature is located in the T interval where the polymer melt dielectric relaxation.26
smoothly evolves from the liquid to the solid state, and can be As an example, Fig. 3 shows the mean-square displacement
defined operationally through some prescription, for instance, as (MSD) gM(t), averaged over all monomers of a chain, for
the intersection of straight-line extrapolations from the glassy a chemically realistic model of 1,4-PBD.52 At high temperature,
and liquid branches of the volume–temperature curve.11,13 the MSD directly crosses over from ballistic motion (t2) at short

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3433
View Article Online

of mass increases sublinearly for intermediate times. The origin


of these deviations is not fully understood. However, for short
chains the relaxation time of a chain is not well separated from
the (local) a-relaxation so that finite-N corrections to the Rouse
behaviour must be expected.63 Moreover, even for long chains
intermolecular interactions between the polymers are not
completely screened on mesoscopic length scales, which may
cause subdiffusive center-of-mass motion91 or deviations from
reptation theory,6 and lead to corrections to chain ideality.3,92
On cooling towards Tg the Rouse-like motion shifts to
Fig. 3 Mean-square displacement (MSD) gM(t) versus time t for a united progressively longer times due to the appearance of a plateau
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

atom model of 1,4-PBD (results adapted from ref. 52 with permission). regime. In this regime, the MSD increases only very slowly with
The MSD is averaged over all united atoms in the melt time and is of the order of 10% of the monomer diameter,
containing 40 chains of 30 repeats units. The temperatures shown are reflecting the temporary localization of a monomer in the cold
(from left to right): T ¼ 400 K, 323 K, 273 K, 240 K, 225 K, 213 K, and
melt. For polymer models with intramolecular rotational
198 K (Tc z 214 K). The horizontal dashed lines indicate the radius of
2, the average Lennard-Jones diameter of the united barriers, such as PBD, this intermittence of large scale motion
gyration Rg2 z 218 A

atoms s z 3.8 A, and an estimate for the Lindemann localization length
stems from two mechanisms of dynamical arrest:26,50,54 monomer
 42 The dotted lines for T ¼ 400 K and 213 K show the power
rsc z 0.45 A. caging by near neighbours in the dense melt and the slowing
0.61
law t , characteristic of Rouse-like motion. down of intrachain conformational transitions occurring
through (correlated) torsional motion.
Recent work by Smith and Bedrov53 suggests that the interplay
times to subdiffusive motion (tx with x0 z 0.61) at interme-
0
of both mechanisms can explain the Johari–Goldstein b relaxa-
diate times where the MSD is bound between the monomer size tion in polymers.13 Fig. 4 reproduces one of their results. Since
(s) and the end-to-end distance Re of a chain. This subdiffusive experiments reveal that the separation of the a and b processes
motion does not depend on the strength of the torsional barrier occurs only on time scales significantly longer than the multiple
(cf. Fig. 4), is present even if the torsional potential is absent,51 microsecond trajectories generated for the chemically realistic
and thus reflects the universal Rouse-like dynamics of non- (CR) PBD model, Smith and Bedrov employ a PBD model with
entangled chains in a polymer melt.4 The term ‘‘Rouse-like’’ reduced torsional barriers (LB model) but otherwise identical
stresses the fact that simulations of nonentangled chains26,63,89,90 interactions as for the CR model. While structural properties of
find both accord with Rouse predictions—e.g., the Rouse modes the PBD melt remain unaffected,50,51 the reduction of the barriers
are (nearly) orthogonal for all t—and deviations from them— accelerates the dynamics and shifts the a–b bifurcation into the
e.g., the Rouse modes are stretched with stretching exponents simulation time window. Fig. 4 compares gM(t) with the decay of
depending on the mode index and the MSD of the chain’s center the torsional autocorrelation function (TACF). Similar to gM(t),
the TACF relaxes in two steps. To both steps can be associated
relaxation times, sb and sa, which display an Arrhenius (sb) and
a non-Arrhenius (sa) increase with decreasing T, characteristic of
b and a processes, respectively. At low T, both processes are well
separated from each other, the beta process corresponding to
times for which the monomers displace, on average, by about
10% of their diameter. This implies that the cages imposed by the
polymer matrix remain largely intact during the b process and
create a potential energy landscape for the underlying confor-
mational transitions. Indeed, detailed analysis reveals that
(nearly) all dihedrals visit all torsional states during the b relax-
ation. However, these visits do not occur with the equilibrium
probability because the propensity of a dihedral to return to
a preferred conformational state increases on cooling, due to the
Fig. 4 Monomer MSD gM(t) and torsional autocorrelation function stiffening of the matrix. Equilibrium occupancy of conforma-
(TACF) versus t for 1,4-PBD. The data for T ¼ 130 K and 140 K (lines) are tional states is only achieved for times comparable to the a time
obtained from a model with reduced torsional barriers (LB), but which is scale when the ‘‘cage effect’’ of the matrix fully decays (see also
otherwise the same as in Fig. 3 (results adapted from ref. 53 with refs. 55 and 56 for similar results).
permission). For comparison the symbols show the MSDs at T ¼ 198 K In glass physics, the term ‘‘cage effect’’ is intimately connected
and 213 K for the chemically realistic (CR) model with the full torsional
to the mode-coupling theory (MCT) of the glass transition,
potential. These temperatures are approximately at the same distance to
which has certainly been one of the most influential theoretical
Tg (z 170 K) as for the LB model (Tg z 102 K).53 The horizontal dashed
lines indicate the average Lennard-Jones diameter of the united atoms s z
approaches in the field during the last twenty years. A compre-
and an estimate for the Lindemann localization length rsc z 0.45 A.
3.8 A  42 hensive review of its foundations and applications was published
The solid line for T ¼ 140 K shows the power law t0.61, characteristic of recently.17 By correlating the equilibrium structure to the
Rouse-like motion. The vertical dotted lines represent the b-relaxation dynamics of a glass-forming liquid MCT provides a framework
times (sb ¼ 716.5 ps for T ¼ 130 K, sb ¼ 319.4 ps for T ¼ 140 K). for interpreting spatio-temporal correlations measured in

3434 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

experiment or simulation. Analytical predictions are derived in


the vicinity of an ideal glass transition which occurs at a critical
temperature Tc and is driven by the mutual blocking of a particle
and its neighbours (‘‘cage effect’’). Extensive tests by experiments
and simulations indicate that, although (the extrapolated) Tc lies
above Tg and thus no structural arrest is observed at Tc (cf.
Fig. 5), MCT still describes many dynamical features well.17,94
Why this is so, represents a great challenge for the theoretical
understanding.87,95–99
A key prediction of MCT is that structural relaxation func-
tions should obey a factorization property in the plateau regime
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

(also called b regime in MCT). For the self (Gs) and distinct (Gd) Fig. 5 Temperature dependence of the a relaxation time (sa) for united
parts of the van Hove correlation function100 this reads atom models of atactic PS and bisphenol-A PC. sa is normalized by its
value at T ¼ 600 K for both polymers (PS: s(600 K) ¼ 1.4 ps, PC:
Gx(r, t) ¼ Fx(r) + Hx(r)G(t) (x ¼ s, d) (3) s(600 K) ¼ 1.3 ps) and is obtained from the monomer MSD in simula-
tions of 1 chain with N ¼ 80 for PS and 64 chains with N ¼ 10 for PC. For
The factorization property refers to the fact that the correction both polymers, the increase of sa at high T is compatible with the MCT
to the nonergodicity parameter Fx(r) splits into two factors, of prediction sa  (T  Tc)g (PS: Tcx375 K, gx3; PC: Tcx450 K,
which G(t) depends only on time (and temperature) and Hx(r) gx2.2). However, the divergence at Tc is avoided and a crossover to an
only on r.17,94 Therefore, the ratio (r0 ¼ constant)101,102 Arrhenius behaviour (PS: Eax30 kJ mol1; PC: Eax16.8 kJ mol1)
occurs close to Tg (derived from dilatometry. PS: Tgx375 K; PC:
Gx ðr; tÞ  Gx ðr; t0 Þ Hx ðrÞ Tgx433 K). This crossover indicates the transition from the cooperative
Rx ðr; tÞ ¼ ¼ (4)
Gx ðr0 ; tÞ  Gx ðr0 ; t0 Þ Hx ðr0 Þ translational dynamics above Tg to activated (b process like) hoping
below Tg. Results adapted from ref. 47.
should be independent of t. For a flexible bead–spring model93
Fig. 6 confirms this prediction directly from the simulation data
(no fit) and additionally reveals the length scales pertinent for the simulations of a bead-spring model with N ¼ 10.63,109 The
dynamics in the plateau regime because distances for which Hx(r) comparison gives semi-quantitative agreement between simula-
is zero will not contribute to the relaxation. For the self part of tion and theory (for models with large rotational barriers further
the van Hove function the dynamics involves displacements up to complications might arise61). As an example, Fig. 7 shows
the monomer diameter, whereas for the distinct part it includes various MSDs, revealing the strengths and weaknesses of the
monomers up to about the forth neighbour shell. This local approach.63 Certainly, the agreement for the monomer dynamics
character of the relaxation is a direct evidence for the cage effect. is very good. MCT describes the dynamics in the plateau regime,
The agreement between MCT and simulation, demonstrated the following polymer-specific subdiffusive increase, gM  t0.63
in Fig. 6, is not limited to the flexible bead–spring model.
Simulations of semiflexible bead–spring models with rotational
barriers60,61 and of chemically realistic models, including 1,4
PBD,49,52,54 PS and PC,47 generally find that many features of the
spatiotemporal relaxation in weakly supercooled polymer melts
(i.e., T T Tc) are well described by MCT (see however ref. 56).
The main difference between these models and the flexible bead–
spring model is that two coexisting arrest mechanisms—intra-
molecular barriers and monomer caging—determine the struc-
tural relaxation of the former, whereas only caging operates for
the flexible model. This interpretation is a key result of ref. 50,51
and of the recent work by Bernabei et al.60,61 who also make the
interesting conjecture that, within MCT, the pertinent theoretical
framework for polymer models with internal rotational barriers
are (so-called) higher-order transition scenarios,17 as it appears to Fig. 6 Simulation results for a flexible bead-spring model: Rx(r, t) versus
be the case for other systems with distinct arrest mechanisms, r for different times from the plateau regime at T ¼ 0.48 (Tcx0.45). Panel
such as dense colloidal suspensions with short-range attrac- (a) shows Rx(r, t) for the self-part of the van Hove function and panel (b)
tion103,104 or polymer mixtures with strong dynamic asymme- for its distinct part. In panel (b), the pair-distribution function g(r) is also
shown (dotted line; rescaled to fit into the figure). The first peak of g(r)
try.105–107
reflects the bond length (z 0.97) and the minimum position of the
A distinctive feature of MCT is that the dynamics (beyond the
Lennard-Jones potential (z 1.12). In panel (a), rl (¼ 0.2323) denotes the
short time regime) is fully specified in terms of the liquid struc- zero of Rs(r, t) and the circles represent a Gaussian approximation which
ture. This opens the way for an ‘‘ab initio’’ prediction of the pffiffiffi pffiffiffiffiffi
has a zero at 6rsc x0:2327 and a minimum around 10rsc x0:3. Here,
simulated dynamics, solely based on static input obtained from rsc (x0.095) is the Lindemann localization length. In both panels, the
an independent simulation of the studied glass former. Such an dash-dotted lines correspond to the time closest to t0 which is the least
atomistic theory for the slow relaxation of (nonentangled) precise because Rx(r, t) is undetermined for t ¼ t0 . Adapted from ref.
polymer melts has been developed108 and compared to 32,93.

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3435
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45. View Article Online

Fig. 8 Simulation results for a flexible bead–spring model: Probability


distribution P(ln r;t) of the monomer displacements (left scale) at T ¼
Fig. 7 Simulation results for a flexible bead–spring model: Log–log plot
0.43 (Tcx0.415) for t ¼ 10, 100, 250, 500, 1000, 1500, 2500, 5000, 10 000,
of the monomer MSD (labeled M, left scale) and the MSD of the chain’s
and 25 000 (from left to right). The thick dotted line shows P(ln r;t) at t ¼
center of mass (labeled C, right scale) versus Dt with D being the diffusion
t0 2, the peak time of a0 2(t). The dotted curve for t ¼ 25000 represents the
coefficient of a chain. The inset shows the ratio g1(t)/g5(t) (end monomer
Gaussian approximation for P(ln r;t). The vertical dashed line indicates
MSD over middle monomer MSD). The circle refer to the MD results at T
the displacement corresponding to the Lindemann localization length
¼ 0.47 (Tcx0.45), the solid lines to the MCT a master curve, and the pffiffiffi
6rsc ¼ 0:232. Right scale: Monomer MSD gM(t) and non-Gaussian
dashed lines to the MCT predictions including the MCT b process. The
parameter a0 2(t) at T ¼ 0.43. The vertical dotted line indicates the peak
dash-dotted lines indicate the diffusive motion 6Dt and the dotted line
time t0 2 of a0 2(t) and the horizontal dashed line the ‘‘plateau value’’ 6rsc2.
shows the power law  t0.63 of Rouse-like motion. Figure taken from ref. 63.
Figure adapted from ref. 122.

which is identified as a finite-N deviation from Rouse behaviour, where Gs(r, t) is the self-part of the van Hove function, as before.
and the faster motion of the end monomer (g1) relative to the This probability is shown in Fig. 8, together with gM(t) and the
central monomer (g5) of the chain. On the other hand, the theory non-Gaussian parameter110,111
is not so satisfactory for the MSD of the chain’s center of mass * +
(gC). Besides underestimating the plateau height it does not 0 gM ðtÞ 1
a2 ðtÞ ¼  1 (6)
reproduce the subdiffusive center-of-mass motion91 (gC  t z 0.8) 3 ~ rð0Þj2
rðtÞ  ~
between the plateau and diffusive regimes.
The representation in Fig. 7—that is, plotting the data versus where ~ r(t) denotes the position of a monomer at time t. In the
Dt with D being the chain’s diffusion coefficient—facilitates the cold melt there are clear deviations from Gaussian behaviour at
comparison of the time dependence in the b and early a-regimes, all but the shortest and longest times. The non-Gaussian
but camouflages a systematic deviation between theory and parameter is positive and has a maximum at time t0 2 in the late-
simulation. MCT predicts that the a relaxation time sa(q) has the a regime. At this time the distribution P(ln r;t) is very broad,
pffiffiffi
same T dependence for all wave vectors q, whereas simulations, exhibiting small (rT 6rsc ) and large (r T 1) displacements. This
not only for flexible bead–spring models32 but also for other glass hints at large, non-Gaussian fluctuations in particle mobility
formers,110–114 find that sa for wave vectors smaller than the when a0 2 peaks. On cooling toward Tc the broad displacement
position q* of the first maximum of static structure factor distribution develops into a double-peak structure, indicative of
increases on cooling more weakly than for q T q*. This difference large disparities in particle mobility. Apparently, two pop-
also implies a violation of the Stokes–Einstein relation,115–117 i.e., ulations of monomers coexist, ‘‘slow’’ ones which have not
the product of the diffusion coefficient (corresponding to the moved much farther than 10% of their diameter in time t0 2, and
limit q / 0) and sa(q*) is not independent of T. ‘‘fast’’ ones which have left their cage and covered a distance of
This decoupling has been interpreted as a signature of about their diameter or more. This bimodal character of the
increasingly heterogeneous dynamics in the liquid near its glass structural relaxation is hard to predict from MCT,111 could be
transition116–119). The term ‘‘heterogeneous dynamics’’ means responsible for the absence of the divergence of the relaxation
that a glass former near Tg contains subensembles of particles time at Tc, and appears to be a general feature of materials close
with enhanced or reduced mobility relative to the average. To to glass or jamming transitions.114
reveal this dynamic heterogeneity various methods were
deployed,118,119 including filtering techniques to track slow or fast
3.2 Effects of confinement on glassy polymers
particles, analysis of ensemble-averaged three- or four-point
correlation functions,120,121 or scrutiny of the self-part of the van 3.2.1 Brief overview of some experimental results. During the
Hove correlation function.110,111,114 past fifteen years the impact of geometric confinement on the glass
Here we briefly discuss the approach of Ref. 110,111 which has transition has received considerable attention. The progress in the
recently been applied to simulation data of a flexible bead–spring field is described in several comprehensive reviews.23,24,123–127 The
model.122 Following Ref. 110,111 we define the probability picture emerging from these studies is that glass formers confined
distribution P(ln r;t) of the logarithm of monomer displacements to nanoscopic dimensions may exhibit deviations from bulk
in time t by behaviour due to the interplay of spatial restrictions and interfa-
cial effects. The latter (nonuniversal) effects can result from
P(ln r;t) ¼ 4pr3Gs(r, t) (5) particle–substrate interactions, confinement-induced changes of

3436 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

the liquid structure or polymer conformations, density variations,


etc.,13,24 and often appear to dominate the behaviour of the
confined glass former.126
These interfacial effects have an important impact on the glass
transition of thin polymer films. For films supported on
a substrate, many studies,128–142 though not all,143,144 find reduc-
tions in Tg with decreasing film thickness (h) if the polymer–
substrate attraction is weak. This is also the case for one of the
most extensively studied systems, polystyrene on a variety of
substrates.128,130,131,133,134–136,138–142 Here measurements of surface
Fig. 9 Tc(h)/Tc and Tg(h)/Tg (Tc and Tg denote the bulk values) versus
relaxation after nanodeformation145 or of the positional depen- rescaled film thickness h/h0. MD results for supported films (open circles),
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

dence of Tg138 further indicate that the depression of Tg is related free-standing films (shaded circles), and films confined between two
to a ‘‘free-surface effect’’: monomers near the free surface are smooth repulsive walls (filled circles) are compared to the glass transition
expected to be more mobile because they feel less steric contraint temperatures Tg(h) of three studies: (i) Monte Carlo simulations of
than their peers in the bulk. This enhanced mobility should lead a lattice model for free-standing atactic polypropylene (PP) films176
to reductions in Tg. (crosses). (ii) Experiments of supported atactic PS films of low molecular
Measurements of (the average) Tg provide information on the weight (open squares).131 (iii) Experiments of supported, high-molecular
dynamic response of the polymer films on the time scale associ- weight PS films128 (stars). The solid and dashed line show eqn (7) and eqn
ated with Tg which depends on experimental conditions, such as (8), respectively. The vertical dotted line roughly indicates a film thick-
ness of 10 nm. Figure adapted from ref. 170.
the cooling rate.146 Additionally, the full structural relaxation has
also been explored by several techniques,24,125 in particular by
dielectric spectroscopy.123,129,133,134,140–142,147–151 Dielectric spec-
completely smooth walls which are either purely repulsive or
troscopy allows for a simultaneous measurement of Tg and the
weakly attractive. Despite the obvious differences between
relaxation spectrum, even for thin films. A key finding of these
experimental and computational systems—different polymers,
studies is that the a process is broadened relative to the bulk. It is
absence or presence of a substrate, etc.—a master curve for the
possible to interpret this broadening as a consequence of
reduction of Tg can be constructed, if Tg(h) is scaled by the
spatially heterogeneous dynamics in nanoconfinement. At the
bulk value and h by a characteristic thickness h0 that depends
interfaces the segmental dynamics can be enhanced (e.g. at the
on the nature of the system, but only (very) weakly on
free surface) or slowed down (e.g. at an attractive substrate)
molecular weight. This master curve allows us to compare the
relative to the bulk. Chain segments in layers adjoining these
different systems, which is instructive in several respects. First,
interfacial layers should also have their dynamics perturbed,
the close agreement of the PS data for low and high molecular
albeit to a lesser extent, which will lead to a still weaker pertur-
weight suggests that possible changes of the entanglement
bation for the next layer, and so on. This interpretation implies
density136,181 or chain conformation182,181 in thin films are
that there is a smooth transition from interface-induced pertur-
probably not responsible for the Tg depression. Furthermore,
bations of the dynamics to bulk behaviour with increasing
the experimentally observed Tg shifts have stimulated several
distance from the interfaces. Recently, this idea has been
attempts to model this phenomenon theoretically.131,183–192 For
exploited to analyze dielectric spectra of nanostructured diblock
instance, based on a percolation model for slowly relaxing
copolymer melts,149 and it is also consistent with the positional
domains Long and Lequeux190 derive the formula (solid line in
dependence of Tg found in ref. 138.
Fig. 9)
"  d #
3.2.2 Simulation work on geometrically confined glass-form- h0
ing polymer systems. Simulation studies of confined glass formers Tg ðhÞ ¼ Tg 1  (7)
h
support this view of an interface-induced gradient in relaxation.32
Many of these studies utilize simple models, e.g. binary originally suggested by Keddie et al. as an empirical parametri-
liquids152–155 or bead–spring polymer models156–173 (see however zation in their seminal study on supported PS films.128 An
ref. 174–178 for work on chemically realistic models), to explore alternative parameterization was proposed by Kim et al.135
the impact of confinement in thin films,152,154,156–173,176–178 pores153 (dashed line in Fig. 9)
or systems containing nanofillers.179,180,155 The simulations reveal
Tg
a complex relaxation behaviour on approach to the glass transi- Tg ðhÞ ¼ (8)
1 þ h0 =h
tion of the confined glass former and also report shifts of Tg,
qualitatively similar to the trends sketched above for experiments. and an attempt was undertaken to justify this formula by
In the following we illustrate these results by some examples. a viscoelastic capillary waves model.131,183 Fig. 9 shows both
Fig. 9 compares the reduction of Tg with film thickness formulas and reveals that, for a typical range of experimental
found in experiments on supported PS films of low131 and high film thicknesses, 15 ( h/h0 ( 300, it is hard to decide whether
molecular weight128 with the Tg shifts obtained from simula- eqn (7) or eqn (8) is more accurate. However, including the
tions of free-standing films,169,170,176 supported films,169,170 and simulation data for h ( 10 nm, eqn (8) appears to provide the
films confined between two substrates.165,166 The simulations better description of the Tg shift. Therefore, eqn (8) will be
study a chemically realistic model of polypropylene176 or flex- employed later (Fig. 12) in an analysis of the local Tg of simu-
ible bead–spring models,165,166,169,170 and use as a substrate lated polymer films.

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3437
View Article Online

The explanation for the Tg shift in the simulations rests upon


the impact that the boundaries exert, because the average
behaviour of the film—and so its Tg—aggregates contributions
from all layers in the film, and these layers have distinct prop-
erties. This can be illustrated by an analysis resolving structure
and dynamics as a function of distance from the boundaries. For
flexible bead–spring models Fig. 10 shows two examples of such
an analysis: the layer-resolved incoherent scattering function
fqs(t, z) at q* (maximum of S(q); cf. Fig. 11) for a polymer melt
surrounding a highly faceted, but nearly spherical filler particle
with a structured and attractive surface (panel (a)),179 and
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

fq s(t,z) for a polymer film supported by a smooth, weakly


*

attractive substrate (panel (b)).169


For the polymer melt surrounding the nanofiller the scattering
function, averaged over all monomers (crosses in Fig. 10(a)),
displays features unfamiliar from the bulk. The a-relaxation
seems to occur in two steps, as if there were two distinct
processes, a fast one corresponding to a bulk-like phase far away
Fig. 10 (a) Layer-resolved incoherent scattering function fqs(t, z) at T ¼ from the filler and a slow one associated with interfacial relax-
0.4 and q ¼ 7.08 (¼ maximum of S(q)) for a bead-spring polymer melt ation. The layer-resolved analysis reveals that this interpretation
surrounding an icosahedral filler particle. The chain length is N ¼ 20, the is misleading. The strongly stretched tail of the average corre-
melt density is r ¼ 1 (dashed horizontal line in the inset), and the filler lator results from the smooth gradient in the decay of fqs(t, z)
attracts the monomers more strongly than they attract each other in the which slows down on approach to the filler particle. Ref. 179,180
bulk. The solid lines and the squares show fqs(t, z) for different distances show that the amplitude of this tail can be tuned by the mono-
z from the surface of the filler particle. The location of the first two layers mer-filler interaction. Strong attraction leads to a more
is illustrated in the inset which depicts the monomer density profile pronounced tail; vanishing attraction suppresses the tail. In the
r(z/Rg) (Rgx2.17). The crosses indicate the average over all layers.
latter case, the shape of the scattering function is bulk-like.
Figure adapted from ref. 32. (b) fqs(t, z) at T ¼ 0.42 and q ¼ 6.9 (z
Additional insight into the slow relaxation of particles in
maximum of S(q)) for a bead–spring polymer melt in a supported film of
thickness h ¼ 20.3 (Tc z 0.392). z denotes the distance from the (left) contact with a structured wall was obtained from studies of the
wall. fqs(t, z) is obtained as an average over all monomers which remain self-part of the van Hove function.154,193,194 For times in the
for all times shown in a layer of width Dz ¼ 2 and centered at z. The a regime and T [ Tc, Gs(r, t) has a clear two-peak structure. The
average behaviour of the film (average over all layers) is indicated by first peak reflects particles that remain trapped in their cages,
crosses and the bulk data by filled circles. Inset: Monomer density profile while the second peak, located on a length scale corresponding to
r(z) versus z. The layers for which fqs(t, z) is shown in the main figure, are the wall structure, reveals that particles have migrated to
indicated. Figure adapted from ref. 169. neighbouring wells. Evidence for this kind of ‘‘hopping motion’’
is found for a binary LJ mixture,154 for a polymer melt adsorbed
on a structured surface193 or for model of liquid toluene confined
in cylindrical mesopores.194 This relaxation mechanism could be
generic when a liquid may lock into registry with the surface
topography, leading to a mechanism of structural slowing down
which coexists with glassy arrest at low T.195,196
A similar locking is not possible at the free surface or
a smooth, repulsive or weakly attractive wall. In thin films, one
therefore expects enhanced dynamics relative to the bulk.
Fig. 10(b) shows that this expectation is borne out for flexible
bead–spring models. For these models the free and smooth
interfaces also create environments for nearby monomers which
tend to reduce the cage effect: local spatial correlations on the
scale q* of the maximum of S(q) are weaker in the films than in
the bulk at the same temperature (Fig. 11). This is an important
Fig. 11 Static structure factor S(q) for a bead–spring melt (N ¼ 10) in contributing factor to the enhanced monomer dynamics found at
the bulk (circles) and in supported film geometry (lines) at T ¼ 0.44 (Tc z both interfaces.
0.405 in the bulk; Tc(hx7) z 0.361 in the film169). For the film S(q) is
Fig. 10(b) also shows that the interface-induced deviations
shown respectively for layers in film center, at wall and at the free surface.
from bulk dynamics continuously turn into bulk-like relaxation
At the free surface the steep rise of S(q) for small moduli q of the wave
vector is due to capillary waves. For the bulk the compressibility plateau
with increasing distance from the boundaries. The range of this
kBTrkT is indicated by a horizontal dotted line (kT denotes the isothermal crossover grows on cooling. Fig. 12 illustrates this point by an
compressibility). Inset: Same data as in the main figure, but in a log–log analysis of the temperature dependence of the layer-resolved
representation. The small-q behaviour expected from capillary wave a relaxation time, sa(z, T). At high temperature sa(z, T) slightly
theory is depicted by a dashed line (g is the surface tension). deviates from the bulk value near the interfaces. With decreasing

3438 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

can enhance or retard the relaxation relative to the bulk.


Enhanced relaxation may be expected for smooth or free inter-
faces, whereas strong particle–substrate attraction or particle
caging in cavities of the substrate tend to slow down the
dynamics. These interface-induced perturbations smoothly
transition from the boundaries to the interior of the confined
liquid. The range of this gradient grows on cooling so that the
perturbations can propagate across the entire liquid for suffi-
ciently strong confinement or low T. Similar interfacial effects
may also be important for the analysis of other problems, for
Fig. 12 Temperature dependence of the layer-resolved a relaxation time instance, for solvent evaporation from (spincoated) polymer
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

sa(z, T) in a supported polymer film (N ¼ 10; Tc z 0.405 in the bulk). The films172,173 or for the modeling of the hydrodynamic boundary
simulated system is the same as in Fig. 10(b). The solid lines at the free conditions in microfluidic devices.197 Therefore, it appears that
surface show eqn (11) (no adjustable parameter). The horizontal dashed those theoretical approaches, which treat the interplay of
lines indicate the bulk values for sa at the respective T. Figure adapted
boundary effects and glassy slowing-down of the dynamics on
from ref. 170.
the same (microscopic) footing, are most likely to advance our
understanding in this field. Folding in boundary effects is
T the interface-induced enhancement of the dynamics increas- certainly a major challenge.195,196,198 Simulations of model
ingly penetrates into the film and eventually propagates across systems—as those reported here—should be helpful for the
the entire system for sufficiently low T. This spatial dependence development of such theories.
of sa(z, T) can be modeled if two assumptions are made: (i) the
average Tc(h) of the film is given by eqn (8) and can be written as 4 Studies of the glassy state
a ‘‘democratic average’’192 of the local Tc(z). That is,
Molecular simulations are intrinsically limited in terms of time
ð
h=2
scales. Therefore studying the glassy state, in which the relax-
Tc 2
Tc ðhÞ ¼ ¼ dz Tc ðzÞ (9) ation time scales are by definition extremely large, could seem to
1 þ h0 =h h
0 be out of reach for this kind of numerical approach. Paradoxi-
cally, however, the fact that the intrinsic time scales of the system
which gives
are large brings the simulations very close to actual experiments.
Tc ð1 þ h0 =zÞ The important fact is that, deep in the glassy state, segmental
Tc ðzÞ ¼ (10)
ð1 þ h0 =2zÞ2 relaxation times199 greatly exceed both the experimental and the
simulation time scales. For both laboratory and computer
(ii) the second assumption is that the sole effect of the interface experiments, a time window is probed where segmental polymer
is to shift Tc from the bulk value to Tc(z), whereas all other motions are frozen and hence non-equilibrium phenomena
parameters determining the T dependence of sa remain the same associated with the glassy state are observed. The main issue is
as in the bulk. Here we model this T dependence by the MCT then rather a matter of sample preparation, which is usually done
power law with much faster quenching rates in computer experiments (cf.
sbulk Sect. 3.1.1). As we will see below, it turns out that the mechanical
0
sa ðz; TÞ ¼ (11) behaviour is not strongly affected, at least at a qualitative level,
ðT  Tc ðzÞÞgbulk
and that simulations can therefore be used to analyse the struc-
Fig. 12 demonstrates that eqn (11) yields a reasonable ture and mechanical response of the glassy state with some
description of the increase of sa from the free surface to the confidence.
center of the film (solid lines in the figure).170 The decrease of the
a relaxation time on approach to the free surface therefore
4.1 Small deformations
corresponds to a decrease of the local glass transition tempera-
ture, in qualitative agreement with the experimental results of ref. We start our discussion by considering the small strain, elastic
138. part of the mechanical response of a polymer glass, which can be
The propagation of enhanced or reduced mobility from the described at the macroscopic scale by linear elasticity. It is now
boundary toward the interior of the film has also been observed well recognized that, in spite of their homogeneity in density,
in other simulations on freely-standing157 and supported polymer glassy materials are heterogeneous at the nanoscale in terms of
films.163 These studies carried out a cluster analysis, of highly their elastic properties. This heterogeneity is reflected indirectly
mobile monomers in the case of the freely-standing film and of in some vibrational properties (e.g., the Boson peak200), and is
immobile monomers for the supported film. In both cases, it was most obviously revealed by simulation, that allows one to
found that clusters start at the interface and penetrate into the compute elastic constants and study elastic response at various
film. scales.201 For polymer glasses, the first study of that kind was
In summary, simulation studies suggest that confined (poly- presented in ref. 202,203, following earlier studies that indicated
meric) liquids display complex relaxation behaviour on approach size dependent results for the elastic constants of nanometric
to the glass transition because of the interplay of bulk-like systems.204 This study, which uses a simple definition of local
slowing down of the dynamics and interfacial effects. Interfaces elastic constants based on a local calculation of the usual

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3439
View Article Online

fluctuation formulae,203,205 shows clearly that the material is Further studies of weak deformation include the influence of
inhomogeneous at scales of the order of 5 to 10 monomeric sizes. an external strain on the microscopic dynamics214 (in fact the
Small regions displaying negative elastic constants and therefore latter study also extends to large, uniaxial deformations), and the
would be unstable if they were not surrounded by regions with evolution of the creep compliance upon aging.215 It is found
large local moduli. These aspects are not specific to polymer a finite strain rate accelerates the segmental dynamics—
systems, but can be observed in simple molecular or metallic measured by the bond orientation relaxation time—whether
glass formers,206,201 and using more sophisticated definitions of the corresponding strain is positive (dilation) or negative
the local elastic moduli, which points to their universal character. (compression). This symmetry between compression and dilation
However, they are particularly important when the polymer glass is indicative of a stress induced modification of the potential
is cast in the form of a nanostructure, as the mechanical failure of energy landscape, with the applied stress lowering the local
such structures will be strongly affected by the presence of elastic barriers and accelerating the caging dynamics. Free volume
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

heterogeneities. considerations, on the other hand, would not explain such


While the response of a polymer glass for deformations smaller a symmetry.214 Acceleration of the dynamics under load (some-
than a few percent can be described as elastic from a macroscopic times described as ‘‘mechanical rejuvenation’’) is also observed in
viewpoint, a detailed microscopic studies show that irreversible the aging study of ref. 215, in which the creep compliance J(t, tw)
rearrangements take place at low temperature even for very small of a simulated polymer is also shown to be well accounted for by
deformations.207 This irreversible behaviour is responsible for the classical description of Struik,216 in which the creep compli-
a dissipative part or the mechanical response, sometimes ance is described as a scaling function J(teff/twm) where tw is the
described as anelasticity. Ultimately, this dissipation is associ- Ðt  tw m
aging time, and teff ¼ dt. This result again shows
ated with strongly localized plastic events that involve only a few 0 tw þ t
0

monomers, and can be thought as the precursors of the dissi- that, in spite of the wide difference in time scales, the pheno-
pative processes that take place under larger deformation. menology of glassy polymers is reproduced by simulation work.
Two specificities of polymer glasses, already mentioned in It should also be noticed217 that the relation between segmental
Sect. 3.2.2, are the interest for thin films and the ability to mobility and strain rate is also dependent on the general strain
introduce various types of additives, from small molecules to history, so that, in contrast to the usual assumptions of Eyring’s
nanoparticle fillers. Relatively few studies are concerned with the theory, no simple mechanical variable can relate to the local
thin film modifications of the glassy state itself. Jain and de mobility. Recent theoretical approaches,99 based on nonlinear
Pablo208 showed that the vibrational density of states is modified Langevin equation description of segmental dynamics, go
compared to the same glass in the bulk. In particular, the usual beyond the simple Eyring description and account for this aging
excess of low frequency modes observed in glasses (the so called behaviour.
Boson peak) is enhanced in free standing thin films. This increase Finally, a rich and promising domain for simulation is the
of the soft modes, low frequency part of the spectrum is consis- study of the complex effects of small molecules on the glass
tent with a negative shift of the glass transition temperature. As transition and on the glassy state. Such small molecules can be
the strength of the Boson peak is in general, connected to described as solvent, plasticizers or antiplasticizers, and in
a decreased fragility (in the sense of Angell’s classification of general their presence leads to an acceleration of the dynamics
glasses, see ref. 12, this points to a reduction in fragility with film and a decrease in the glass transition More precisely, plasticiza-
thickness. tion of the polymer means that the solvent does not only decrease
Many studies on the other hand have been devoted to the Tg, but also softens the polymer glass by reducing the elastic
modification of the glassy state under the influence of nano- moduli. Besides this normally encountered case, there are also
metric filler particles.179,180,209 As discussed in Sect. 3.2.2 (see systems in which the solvent antiplasticizes the polymer. That is,
Fig. 10), the vicinity of fillers was shown to modify the dynamics the solvent decreases Tg, but increases the elastic moduli of the
of the polymer, with a slowing down and a shift of the glass polymer glass. The origin of this antiplasticizing effect has
transition temperature in the case of attractive interactions.
From the mechanical point of view, it was also shown that the
presence of attractive fillers results in the presence of a layer
with enhanced mechanical properties (higher local moduli)209,210
and modifies the structure of the entanglement network.211 The
spatial extension of this layer is of the order of a few monomer
diameters and does not depend on the particle radius as soon as
the latter exceeds 2 to 3 monomer size. As a result the
mechanical properties of the composite are enhanced, however
for a well dispersed composite this enhancement remains
moderate. Note that the existence of a glassy layer around the
nanoparticles is expected to affect the properties of the nano-
Fig. 13 Snapshots of three configurations of a polymer glass: initial
composite mostly in the region of the glass transition, where the undeformed state (up), uniaxially deformed configuration (left), and
contrast between this layer and the rest of the polymer is triaxially deformed configuration (right). Both deformed configurations
maximal.212,213 As for polymer films, the fragility (inferred from correspond to a strain of 50% along the z axis of the simulation box. The
Boson peak strength and position) appears to be decreased with simulation results are obtained from a flexible bead-spring model (N ¼
respect to the pure polymer. 100) at T ¼ 0.2 (Tg z 0.43). Figure taken from Ref. 75.

3440 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

recently been studied by simulations,218,219 and appears to be due The first part is a constant dissipative stress sY, due to the
to a more efficient packing associated with solvent molecules presence of energy barriers. The second is the strain-dependent
smaller than the monomers. Small molecule fit into the ‘‘holes’’ of part, which is thought to be described by rubber-elasticity theory
the polymer melt, and increase the elastic stiffness. However, and represents the strain-hardening effect. In this description the
their mobility also facilitates segmental mobility. Both for anti- strain-hardening modulus is not affected by thermally activated
plasticizers and plasticizers, a strongly heterogeneous dynamics processes.
of the solvent molecules has been reported.122,219 The corres- However, recent experiments demonstrate that such
ponding dynamical correlation length is however decreased in a description for the strain-hardening part is invalid; the strain-
the presence of antiplasticizres, with a fragility of the system that hardening modulus has characteristics of a thermally-activated
is also decreased. process, and decreases for higher temperatures.228 Secondly, at
higher strain rate the strain-hardening modulus increases,229
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

although for some polymers the dependency on strain rate is


rather weak.230 This increase can be interpreted in a very
4.2 Large deformations and strain hardening
elementary manner within a barrier-crossing, Eyring like picture,
The mechanical properties of glassy polymers at large deforma- which however, as discussed above, is known to be over-
tions were first investigated in the pioneering work of R€ ottler and simplified. Finally, the external pressure affects the strain-hard-
Robbins,220–223 and their results were subsequently reproduced in ening modulus as well. A higher external pressure leads to an
a number of studies.75,224 Beyond the peak stress, whose actual increase in the strain-hardening modulus.231 Again, this behav-
amplitude depends on the strain rate in a logarithmic way, as in iour is typical of thermally-activated processes. All these three
simple glasses, the stress strain behaviour depends on the type of observations on the strain-hardening modulus are not present
sollicitation. In pure shear or in simulations of a tensile test under within the classical rubber theory. The failure of rubber-elasticity
triaxial conditions, the plastic flow proceeds through a cavitation theory is due to the essential difference between a rubbery state,
of cavities and subsequent formation of fibrils (Fig. 13). The with many possible chain conformations between cross-links,
drawing of these fibrils is essentially at constant stress. This and a non-ergodic glassy state where chain conformations are
phenomenon is discussed in detail by R€ ottler.27 Under uniaxial, practically frozen, and transitions between different conforma-
almost volume conserving conditions (Poisson ratio close to 1/2), tions are not possible. Deformation of the polymer glass facili-
a marked strain hardening is observed, i.e. the plastic stress tates these transitions and is accompanied with the energy
increases with strain with a dependency that is close to linear. dissipation. The work for this irreversible energy dissipation is
The corresponding deformation, however, is very far from reflected in a dissipative stress, which is absent in the rubber-
elastic, and essentially non recoverable. elasticity theory. The dissipative nature of a polymer strain-
This peculiar property of polymer glasses, which strongly hardening is confirmed by experiments. It is found that for PC
contributes to their practical applications, is the existence of the and PS at large (>15–30%) strains20,232,233 more work is dissipated
strain hardening regime at large deformation. Large strain- through heat than converted into internal energy.
hardening effect prevents the strain localization and leads to The increase in stress for larger strains in combination with the
a tough response of the polymeric material; the material breaks dissipative nature of the stress in the strain-hardening regime
only after a significant plastic strain, as in the case for poly- suggest that there is an increase in the rate of energy dissipation,
carbonate. For brittle polymers, as atactic polystyrene, the i.e., more energy per unit of strain is needed for more stretched
strain-hardening effect is usually too weak, which leads to samples to stretch them further. This picture is supported by the
a brittle fracture due to crazes formation within a few percent of computer simulations of Hoy and Robbins,234 who showed that
strain. Understanding the microscopic origin of the strain- the dissipative stress increases with larger strain and that at zero
hardening effect provides a strategy for new tailor-made poly- temperature the stress was directly correlated to the rate of
meric materials. changes in Lennard-Jones (LJ) binding. Their more recent
The microscopic origin of this regime has been clarified only simulations of polymer toy models also demonstrated that most
recently, thanks—to a large extent—to numerical simulations. of the stress at large strains is due to dissipation.234,235
Both the generic aspects of this phenomenon, and the way strain- If the rubber elasticity theory is invalid, what is then the
hard and strain-soft polymer materials differ in terms of polymer-specific part of the strain-hardening modulus? Simula-
segmental mobility and energetics, have been investigated by tions of the mechanical deformation of simple molecular glasses
extensive MD simulations. show no strain hardening.236 Hence, for short polymer chains the
Before we discuss simulation results, let us recall the popular amount of strain hardening is expected to be small as well.
rubber elasticity models225 which predicts that the hardening Moreover, experiments227 and simulations237 show that the strain
modulus Gh varies linearly with temperature T and entanglement hardening modulus is positively correlated with the entanglement
density re, Gh ¼ rekBT. Experimental values, however, are two density. Therefore one expects the strain hardening phenomenon
orders of magnitude larger, and, more important, show decrease to disappear progressively as the chain length falls below the
with increasing T.226,227 To take the dissipative nature of the entanglement threshold. Simulations show that this is indeed the
plastic deformation into account the rubber elasticity model case. In a model polymer glass, Hoy and Robbins234 demon-
describes the stress-strain relation in the strain-hardening regime strated that there is a gradual increase in the strain-hardening
with the help of two contributions.225 modulus as a function of chain length for chains up to about the
entanglement length. For longer chains saturation in the
s ¼ sY + c2(l2  l1) (12) modulus occurs. Similarly, in MD simulations of atactic

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3441
View Article Online

polystyrene47 only a weak strain hardening was observed for strain, and that the increase is larger for PC. The non-affine
chain lengths of 80 monomers, below the experimentally displacements of particles are due to restrictions and hindrances,
observed entanglement length of about 128–139 monomers.238,239 in particular from covalent and steric interactions. These
In the simulations of ref. 234 it was also observed that the restrictions are extremely important at the scale of the covalent
strain hardening is more correlated with the change in the end-to- bond. If particles would displace affinely, then the equilibrium
end distance of the polymer chains than with the change in the value of this chemical bond would be excessively disturbed. To
global sample size. If the sample size is decreasing while the end- circumvent the bond stretch, the bond vector will not move
to-end distance does not, then the stress does not increase. affinely with the deformation. At much larger length scales the
After the fundamental aspects of strain hardening have been situation is different. For a long chain the internal conformation
clarified using simulations of simple, coarse-grained models, can be adjusted, while still obeying to the equilibrium length of
more detailed studies using realistic models are required to the covalent bond. The effective spring constant associated with
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

understand the influence of chain architecture. For example, it is the end-to-end distance is much less stiff. Moreover, due to the
known that polymers with a larger persistence length often have glassy state the relaxation time of spontaneous rearrangements at
a higher strain-hardening modulus.225 It was even shown by the scale of the whole chain greatly exceeds experimental and
molecular-dynamics (MD) simulations that if the persistence simulation time scales. Therefore, the end-to-end distance cannot
length of a polyethylene-like polymer is artificially increased by adjust back towards the equilibrium value by means of sponta-
changing the trans-to-gauche ratio, the strain-hardening modulus neous relaxations. Hence the end-to-end distance will follow the
of the resulting material increases as well.240 That example imposed deformation much more affinely. So we expect a more
illustrates that the strain-hardening modulus depends on the affine response, as we probe larger length scales.
conformation of the chain, which is frozen in the glassy state. The This expectation is borne out by a detailed study of the affine
changes in this conformation are associated with plastic events, character of the deformation of the polymer chain, as a function
that involve a collective nanoscale segmental dynamics of indi- of the internal distance along the chain.242 At the scale of 100
vidual or neighbouring chains. Understanding the differences in chemical bonds, it is found that the deformation is essentially
local mobility and dynamics for chemically different polymers affine up to about 15% strain for both PS and PC, while it is
under large deformations, and how it relates to the different strongly non affine already at this strain at the scale of 30 bonds,
mechanical characteristics, is therefore an important question,
which has been studied in relatively few cases.
As an example of such studies, and of the possibilities offered
by simulation, we consider the extensive studies in ref.
47,231,2412,242 on two glassy polymers that vary greatly in their
strain-hardening moduli, viz. polystyrene (PS) and poly-
carbonate (PC, of which the modulus is more than a factor of two
higher243). Molecular-dynamics simulations47,231,241 have repro-
duced these experimental findings qualitatively, with a strain-
hardening modulus of polystyrene that is much lower than that
of polycarbonate, as shown in Fig. 14.
They also allow one to identify the microscopic mechanisms
responsible for the observed difference in the strain-hardening
modulus. In particular, they show the rate of non-affine
displacements (or local plastic events) increases with larger

Fig. 15 The normalized characteristic ratio C0 n(3eng) ¼ Cn(3eng)/Cn(0),


which gives the deviation from affine deformation as a function of strain,
Fig. 14 The von Mises equivalent true stress s vs. strain 3eng for PS and for a fixed curvilinear distance n along the chain backbone. Upper panel:
PC. Solid lines are fits to eqn (12). Fit range is 3eng ¼ 0.3–0.8. Note that PS. Lower panel: PC. Solid lines with filled symbols are simulation
the strain-hardening modulus Gh for PC (19 MPa) is almost twice that of results, while open symbols are results as if the sample would deform in
PS (11 MPa), while their extrapolated offset yield values sY are about the an affine way. The black curve is the affine approximation (lx2 + ly2 +
same (PC: 88 MPa; PS: 86 MPa). Adapted from ref. 231. lz2)/3. Adapted from ref. 231.

3442 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

see Fig. 15. For larger strains the relative deviation from affinity existence of side groups, to torsional modes, and other inter-
becomes larger for even long enough (100 backbone bonds) pretations that have been put forward.244 Molecular simulation,
chains. This effect is present in both PS and PC, the difference by the flexibility it offers to block specific motions and to control
being in the magnitude of the effect. For an internal distance of 30 chain properties at various scales of coarse graining, should also
monomeric units, the deformation of the polycarbonate chain is allow one to investigate the putative relations between these
only 30% of the affine value at a strain of 50%, while in poly- secondary relaxations, fragility, and mechanical properties.
styrene under the same conditions the chain deformation is 50% Nanocomposites and thin films have already been the subject
of the affine value. This is because at the scale of the Kuhn length of a number of simulation studies (see Sect. 3.2.2). Still, the
a chain cannot be stretched any further. As the Kuhn length of precise mechanisms that give rise to reinforcement and nonlinear
polycarbonate is already larger than that of PS and the total non- behaviour in nanocomposites, and to speed up of the dynamics in
affine displacement of PC is to a large extent determined by the thin films, remain to be elucidated in detail. In particular, for
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

backbone atoms, the increase in the effective stiffness length leads entangled polymers it can be expected that the presence of
to an increase in non-affine displacement, more energy dissipa- interfaces modifies both the characteristics of the entanglement
tion and hence a higher strain-hardening effect. For polystyrene network and the monomeric friction coefficients. The interplay
the Kuhn length is small and the major part of non-affine between these different aspects could, in principle, be elucidated
displacement is not caused by the backbone. Hence the expected in simulations of entangled systems. Again, the simulation times
increase in the effective stiffness length during deformation does needed to explore such phenomena are extremely large, but, with
not lead to a substantial increase in plastic flow, so that at the development of advanced simulation algorithms—as those
moderate strains PS behaves more like a simple glass without alluded to in Sect. 2—and the increase of computer power they
strain hardening, as opposed to polycarbonate. are starting to be within the attainable range.
Finally, we note that many polymer materials of practical
interest display a mixed structure, being either partially crystal-
5 Perspectives
line with an amorphous fraction, or chemically inhomogeneous
The present review has attempted to describe some aspects of the with phases (or microphases in the case of block copolymers)
recent progresses in our understanding of glassy polymers that displaying different mechanical properties, e.g. glassy and
have been obtained from simulation work, with a particular rubbery. The properties of the resulting nanocomposites have, up
focus on microscopic dynamic, glass transition and mechanical to now, received little attention from the standpoint of molecular
properties. The results obtained in the past ten years show that simulation (see however ref. 106,107,245). They are likely to be
the simulations have reached a state of maturity that allows them strongly influenced, especially for large deformations, by the
to address and clarify important issues in the field, in spite of the interplay between chain architecture and material nanostructure.
obvious limitations in terms of length and time scales. While the Another interesting phenomenon of that kind is strain induced
illustrative examples we have chosen were relatively simple, they crystallisation, which plays an important role in the strain
show that the general phenomenology of the glass transition and hardening of some specific polymer. The capability of molecular
the properties of the glassy state can be accounted for using simulation to describe polymer crystallization has been recently
molecular simulations, in spite of the relatively small time scale demonstrated,7,8,246 so that the study of strain induced crystal-
that can be considered in such simulations. We would like to lisation—as well as studies of semi crystalline phases appears to
close the discussion by suggesting some directions for future be an interesting goal for the future.
research, which we believe will be soon — or are already —
within the capabilities of state of the art simulations, and
correspond to problems of practical or fundamental interest. Acknowledgements
A major ingredient of the physics of glassy polymers—and The work reported in this article was carried out in fruitful
more generally of glassy systems—is the so called ‘‘time collaboration with M. Aichele, N. Balabaev, C. Bennemann, K.
temperature superposition principle’’ which assumes that relax- Binder, S.-H. Chong, C. Donati, M. Fuchs, Y. Gebremichael, S.
ation processes can be rescaled by using a single, temperature C. Glotzer, O. Lame, A. Makke, M. Mareschal, H. Meyer, M. A.
dependant relaxation time sa(T). While this ‘‘principle’’ is widely J. Michels, S. Napolitano, J. J. de Pablo, G. Papakon-
used to produce frequency dependent relaxation function from stantopoulos, W. Paul, M. Perez, S. Peter, R. Riggleman, B.
data taken at various temperature, it is also well known to be Schnell, R. Seemann, F. W. Starr, F. Varnik, T. Vettorel, B.
applicable at best in the frequency region corresponding to the Vorselaars, and M. W€ ubbenhorst. It is a great pleasure to thank
a peak, presumably only above Tc. It is a challenge to understand all of them. We presented results from the research of W. Paul, G.
the molecular mechanisms underlying this superposition prin- D. Smith, D. Bedrov and coworkers. We are grateful that they
ciple and the deviations thereof.117 A prominent feature of quickly provided the figures requested. Our simulations were
polymeric glasses is the observation of secondary relaxations at made possible by generous grants of computer time at the IDRIS
higher frequency, the Johari–Goldstein b processes (see Sect. in Orsay and at the NCF in Amsterdam, JB gratefully acknowl-
3.1.2). The associated relaxation times merge with the a relaxa- edges financial support by the IRTG ‘‘Soft condensed matter’’.
tion times typically at temperatures of 0.8Tg, and the associated
time scales are of the order of 106–107 s at such temperatures.
Such time scales are beginning to be in the range attainable by References
MD simulations, a fact that should allow one to clarify the 1 T. P. Lodge and M. Muthukumar, J. Phys. Chem., 1996, 100, 13275.
molecular origins of secondary relaxations, their relation to the 2 D. N. Theodorou, Chem. Eng. Sci., 2007, 67, 5697–5714.

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3443
View Article Online

3 J. P. Wittmer, P. Beckrich, H. Meyer, A. Cavallo, A. Johner and 43 J. Colmenero, A. Arbe, F. Alvarez, M. Monkenbusch, D. Richter,
J. Baschnagel, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., B. Farago and B. Frick, J. Phys.: Condens. Matter, 2003, 15, S1127.
2007, 76, 011803. 44 A. V. Lyulin, N. K. Balabaev and M. A. J. Michels, Macromolecules,
4 M. Rubinstein and R. H. Colby, Polymer Physics, Oxford University 2002, 35, 9595–9604.
Press, Oxford, 2003. 45 A. V. Lyulin and M. A. J. Michels, Macromolecules, 2002, 35,
5 T. C. B. McLeish, Adv. Phys., 2002, 51, 1379. 1463–1472.
6 A. E. Likhtman, J. Non-Newtonian Fluid Mech., 2009, 157, 158–161. 46 A. V. Lyulin, N. K. Balabaev and M. A. J. Michels, Macromolecules,
7 H. Meyer and F. M€ uller-Plathe, J. Chem. Phys., 2001, 115, 2003, 36, 8574–8575.
7807–7810. 47 A. V. Lyulin, B. Vorselaars, M. A. Mazo, N. K. Balabaev and
8 H. Meyer and F. M€ uller-Plathe, Macromolecules, 2002, 35, M. A. J. Michels, Europhys. Lett., 2005, 71, 618–624.
1241–1252. 48 A. V. Lyulin and M. A. J. Michels, Phys. Rev. Lett., 2007, 99,
9 G. Strobl, The Physics of Polymers: Concepts for Understanding their 085504.
Structures and Behavior, Springer, Berlin–Heidelberg, 1997. 49 G. D. Smith, D. Bedrov and W. Paul, J. Chem. Phys., 2004, 121,
10 M. Muthukumar, Adv. Chem. Phys., 2003, 128, 1–63. 4961.
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

11 G. B. McKenna, Comprehensive Polymer Science, Pergamon, New 50 S. Krushev and W. Paul, Phys. Rev. E: Stat., Nonlinear, Soft Matter
York, 1986, vol. 2, pp. 311–362. Phys., 2003, 67, 021806.
12 C. A. Angell, K. L. Ngai, G. B. McKenna, P. F. McMillan and 51 S. Krushev, W. Paul and G. D. Smith, Macromolecules, 2002, 35,
S. W. Martin, J. Appl. Phys., 2000, 88, 3113. 4198.
13 E. Donth, The Glass Transition, Springer, Berlin–Heidelberg, 2001. 52 W. Paul, D. Bedrov and G. D. Smith, Phys. Rev. E: Stat., Nonlinear,
14 P. R. Sundararajan, Can. J. Chem., 1985, 63, 103–110. Soft Matter Phys., 2006, 74, 021501.
15 K. Binder and W. Kob, Glassy Materials and Disordered Solids, 53 G. D. Smith and D. Bedrov, J. Polym. Sci., Part B: Polym. Phys.,
World Scientific, New Jersey–London, 2005. 2007, 45, 627.
16 P. G. Debenedetti and F. H. Stillinger, Nature, 2001, 410, 259. 54 J. Colmenero, A. Narros, F. Alvarez, A. Arbe and A. J. Moreno,
17 W. G€ otze, Complex Dynamics of Glass-Forming Liquids: A Mode- J. Phys.: Condens. Matter, 2007, 19, 205127.
Coupling Theory, Oxford University Press, Oxford, 2009. 55 A. Narros, A. Arbe, F. Alvarez, J. Colmenero and D. Richter,
18 G. H. Fredrickson, Nat. Mater., 2008, 7, 261. J. Chem. Phys., 2008, 128, 224905.
19 J. Ferry, Viscoelastic Properties of Polymers, Wiley, New York, 56 M. Vogel, Macromolecules, 2008, 41, 2949.
1980. 57 W. Kob, J. Phys.: Condens. Matter, 1999, 11, R85–R115.
20 R. N. Haward and R. J. Young, The physics of glassy polymers, 58 T. C. Dotson, J. V. Heffernan, J. Budzien, K. T. Dotson, F. Avila,
Chapman & Hall, London, 2nd edn, 1997. D. T. Limmer, D. T. McCoy, J. D. McCoy and D. B. Adolf,
21 I. M. Ward and D. W. Hadley, An Introduction to the Mechanical J. Chem. Phys., 2008, 128, 184905.
Properties of Solid Polymers, Wiley, Chichester–New York, 1993. 59 M. Bulacu and E. van der Giessen, Phys. Rev. E: Stat., Nonlinear,
22 S. Granick, S. K. Kumar, E. J. Amis, M. Antonietti, A. C. Balazs, Soft Matter Phys., 2007, 76, 011807.
A. K. Chakraborty, G. S. Grest, C. Hawker, P. Janmey, 60 M. Bernabei, A. J. Moreno and J. Colmenero, Phys. Rev. Lett., 2008,
E. J. Kramer, R. Nuzzo, T. P. Russell and C. R. Safinya, 101, 255701.
J. Polym. Sci., Part B: Polym. Phys., 2003, 41, 2755–2793. 61 M. Bernabei, A. J. Moreno and J. Colmenero, J. Chem. Phys., 2009,
23 J. A. Forrest and K. Dalnoki-Veress, Adv. Colloid Interface Sci., 131, 204502.
2001, 94, 167. 62 J. Buchholz, W. Paul, F. Varnik and K. Binder, J. Chem. Phys.,
24 M. Alcoutlabi and G. B. McKenna, J. Phys.: Condens. Matter, 2005, 2002, 117, 7364–7372.
17, R461. 63 S. H. Chong, M. Aichele, H. Meyer, M. Fuchs and J. Baschnagel,
25 J. H. R. Clarke, Monte Carlo and Molecular Dynamics Simulations in Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2007, 76, 051806.
Polymer Science, Oxford Univ. Press, New York, 1995, pp. 272–306. 64 J. McCarty, I. Y. Lyubimov and M. G. Guenza, J. Phys. Chem. B,
26 W. Paul and G. D. Smith, Rep. Prog. Phys., 2004, 67, 1117. 2009, 113, 11876–11886.
27 J. Rottler, J. Phys.: Condens. Matter, 2009, 21, 463101. 65 T. Murtola, A. Bunker, I. Vattulainen, M. Deserno and
28 K. Binder, J. Baschnagel, W. Kob and W. Paul, Bridging Time M. Karttunen, Phys. Chem. Chem. Phys., 2009, 11, 1869–1892.
Scales: Molecular Simulations For The Next Decade, 2002, pp. 66 T. Strauch, L. Yelash and W. Paul, Phys. Chem. Chem. Phys., 2009,
199–228. 11, 1942–1948.
29 K. Binder, J. Non-Cryst. Solids, 2002, 307–310, 1–8. 67 D. Frenkel and B. Smit, Understanding Molecular Simulation,
30 K. Binder, J. Baschnagel and W. Paul, Prog. Polym. Sci., 2003, 28, Academic Press, London, 2nd edn, 2002.
115–172. 68 K. Binder, J. Horbach, W. Kob, W. Paul and F. Varnik, J. Phys.:
31 K. Binder, Phys. Complex Systems (new Adv. Perspectives), 2004, Condens. Matter, 2004, 16, S429–S453.
155, 17–82. 69 D. P. Landau and K. Binder, A Guide to Monte Carlo Simulations in
32 J. Baschnagel and F. Varnik, J. Phys.: Condens. Matter, 2005, 17, Statistical Physics, Cambridge University Press, Cambridge, 2009.
R851–R953. 70 A. Bunker and B. D€ unweg, Phys. Rev. E: Stat. Phys., Plasmas,
33 K. Binder, Monte Carlo and Molecular Dynamics Simulations in Fluids, Relat. Interdiscip. Top., 2000, 63, 016701.
Polymer Science, Oxford University Press, New York, 1995, pp. 71 R. Yamamoto and W. Kob, Phys. Rev. E: Stat. Phys., Plasmas,
3–46. Fluids, Relat. Interdiscip. Top., 2000, 61, 5473.
34 C. Abrams, L. Delle Site and K. Kremer, Bridging Time Scales: 72 C. De Michele and F. Sciortiono, Phys. Rev. E: Stat., Nonlinear, Soft
Molecular Simulations for the Next Decade, Springer, Berlin, 2002, Matter Phys., 2002, 65, 051202.
pp. 143–164. 73 Q. Yan, T. S. Jain and J. J. de Pablo, Phys. Rev. Lett., 2004, 92,
35 J. Baschnagel, K. Binder, P. Doruker, A. A. Gusev, O. Hahn, 235701.
K. Kremer, W. L. Mattice, F. M€ uller-Plathe, M. Murat, W. Paul, 74 M. Chopra, R. Malshe, A. S. Reddy and J. J. de Pablo, J. Chem.
S. Santos, U. W. Suter and V. Tries, Adv. Polym. Sci., 2000, 152, Phys., 2008, 128, 144104.
41–154. 75 B. Schnell, Ph.D. thesis, University of Strasbourg, 2006.
36 F. M€ uller-Plathe, ChemPhysChem, 2002, 3, 754. 76 A. Soldera and N. Metatla, Phys. Rev. E: Stat., Nonlinear, Soft
37 C. Peter and K. Kremer, Soft Matter, 2009, 5, 4357–4366. Matter Phys., 2006, 74, 061803.
38 D. G. Tsalikis, N. Lempesis, G. C. Boulougouris and 77 N. Metatla and A. Soldera, Macromolecules, 2007, 40, 9680.
D. N. Theodorou, J. Phys. Chem. B, 2008, 112, 10619–10627. 78 K. Vollmayr, W. Kob and K. Binder, J. Chem. Phys., 1996, 105,
39 A. Heuer, J. Phys.: Condens. Matter, 2008, 20, 373101. 4714.
40 R. A. L. Vallee, W. Paul and K. Binder, J. Chem. Phys., 2007, 127, 79 K. Vollmayr, W. Kob and K. Binder, Phys. Rev. B, 1996, 54, 15808.
154903. 80 R. Br€uning and K. Samwer, Phys. Rev. B, 1992, 46, 11318.
41 R. A. L. Vallee, M. V. der Auweraer, W. Paul and K. Binder, Phys. 81 Y. Ding, A. Kisliuk and A. P. Sokolov, Macromolecules, 2004, 37,
Rev. Lett., 2006, 97, 217801. 161–166.
42 J. Colmenero, F. Alvarez and A. Arbe, Phys. Rev. E: Stat., 82 K. Kunal, C. G. Robertson, S. Pawlus, S. F. Hahn and
Nonlinear, Soft Matter Phys., 2002, 65, 041804. A. P. Sokolov, Macromolecules, 2008, 41, 7232–7238.

3444 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010
View Article Online

83 J. Hintermeyer, A. Herrmann, R. Kahlau, C. Goiceanu and 124 F. Kremer, A. Huwe, A. Sch€ onhals and S. A. Roz_ a
nski, Broadband
E. A. R€osler, Macromolecules, 2008, 41, 9335–9344. Dielectric Spectroscopy, Springer, Berlin, Heidelberg, New York,
84 A. L. Agapov and A. P. Sokolov, Macromolecules, 2009, 42, 2003, pp. 171–224.
2877–2878. 125 C. B. Roth and J. R. Dutcher, Soft Materials: Structure and
85 B. Lobe, J. Baschnagel and K. Binder, J. Non-Cryst. Solids, 1994, Dynamics, Marcel Dekker, New York, 2005, pp. 1–38.
172–174, 384–390. 126 C. Alba-Simionesco, B. Coasne, G. Dosseh, G. Dudziak,
86 J. Dudowicz, K. F. Freed and J. F. Douglas, Adv. Chem. Phys., 2007, K. E. Gubbins, R. Radhakrishnan and M. Sliwinska-Bartkowiak,
137, 125–222. J. Phys.: Condens. Matter, 2006, 18, R15.
87 K. S. Schweizer, Curr. Opin. Colloid Interface Sci., 2007, 12, 297. 127 O. K. C. Tsui, Polymer Thin Films, World Scientific, Singapore,
88 A. S. Keys, A. R. Abate, S. C. Glotzer and D. J. Durian, Nat. Phys., 2008, p. 267.
2007, 3, 260. 128 J. L. Keddie, R. A. L. Jones and R. A. Cory, Europhys. Lett., 1994,
89 C. Bennemann, J. Baschnagel, W. Paul and K. Binder, Comput. 27, 59.
Theor. Polym. Sci., 1999, 9, 217–226. 129 J. S. Sharp and J. A. Forrest, Phys. Rev. E: Stat., Nonlinear, Soft
90 M. Brodeck, F. Alvarez, A. Arbe, F. Juranyi, T. Unruh, Matter Phys., 2003, 67, 031805.
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

O. Holderer, J. Colmenero and D. Richter, J. Chem. Phys., 2009, 130 J. A. Forrest, Eur. Phys. J. E, 2002, 8, 129.
130, 094908. 131 S. Herminghaus, K. Jacobs and R. Seemann, Eur. Phys. J. E, 2001,
91 M. Zamponi, A. Wischnewski, M. Monkenbusch, L. Willner, 5, 531.
D. Richter, P. Falus, B. Farago and M. Guenza, J. Phys. Chem. B, 132 Y. Grohens, L. Hamon, G. Reiter, A. Soldera and Y. Holl, Eur.
2008, 112, 16220–16229. Phys. J. E, 2002, 8, 217.
92 H. Meyer, J. P. Wittmer, T. Kreer, P. Beckrich, A. Johner, J. Farago 133 K. Fukao and Y. Miyamoto, Phys. Rev. E: Stat. Phys., Plasmas,
and J. Baschnagel, Eur. Phys. J. E, 2008, 26, 25–33. Fluids, Relat. Interdiscip. Top., 2000, 61, 1743.
93 M. Aichele and J. Baschnagel, Eur. Phys. J. E, 2001, 5, 229. 134 K. Fukao and Y. Miyamoto, Phys. Rev. E: Stat., Nonlinear, Soft
94 W. G€ otze, J. Phys.: Condens. Matter, 1999, 11, A1. Matter Phys., 2001, 64, 011803.
95 P. Mayer, K. Miyazaki and D. R. Reichman, Phys. Rev. Lett., 2006, 135 J. H. Kim, J. Jang and W.-C. Zin, Langmuir, 2001, 17, 2703.
97, 095702. 136 O. K. C. Tsui and H. F. Zhang, Macromolecules, 2001, 34, 9139.
96 B. Kim and K. Kawasaki, J. Stat. Mech.: Theory Exp., 2008, 2008, 137 D. S. Fryer, R. D. Peters, E. J. Kim, J. E. Tomaszewski, J. J. de
P02004. Pablo, P. F. Nealey, C. C. White and W. L. Wu, Macromolecules,
97 G. Szamel, J. Chem. Phys., 2007, 127, 084515. 2001, 34, 5627–5634.
98 A. Andreanov, G. Biroli and J.-L. Bouchaud, Europhys. Lett., 2009, 138 C. J. Ellison and J. M. Torkelson, Nat. Mater., 2003, 2, 695.
88, 16001. 139 C. Ellison, M. Mundra and J. Torkelson, Macromolecules, 2005, 38,
99 K. Chen, E. J. Saltzman and K. S. Schweizer, J. Phys.: Condens. 1767.
Matter, 2009, 21, 503101. 140 R. D. Priestley, L. J. Broadbelt, J. M. Torkelson and K. Fukao,
100 J. P. Hansen and I. R. McDonald, Theory of Simple Liquids, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2007, 75, 061806.
Academic Press, London, 1986. 141 V. Lupas xcu, S. J. Picken and M. W€ ubbenhorst, J. Non-Cryst. Solids,
101 G. F. Signorini, J.-L. Barrat and M. L. Klein, J. Chem. Phys., 1990, 2006, 352, 5594.
92, 1294–1303. 142 S. Napolitano and M. W€ ubbenhorst, J. Phys. Chem. B, 2007, 111,
102 W. Kob and H. C. Andersen, Phys. Rev. E: Stat. Phys., Plasmas, 9197.
Fluids, Relat. Interdiscip. Top., 1995, 51, 4626–4641. 143 M. Y. Efremov, E. A. Olson, M. Zhang, Z. Zhang and L. H. Allen,
103 M. Sperl, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2003, 68, Macromolecules, 2004, 37, 4607.
031405. 144 H. Huth, A. A. Minakov, A. Serghei, F. Kremer and C. Schick, Eur.
104 E. Zaccarelli and W. C. K. Poon, Proc. Natl. Acad. Sci. U. S. A., Phys. J. Spec. Top., 2007, 141, 153.
2009, 106, 15203–15208. 145 Z. Fakhraai and J. A. Forrest, Science, 2008, 319, 600.
105 A. J. Moreno and J. Colmenero, J. Chem. Phys., 2006, 124, 184906. 146 Z. Fakhraai and J. A. Forrest, Phys. Rev. Lett., 2005, 95, 025701.
106 A. J. Moreno and J. Colmenero, J. Phys.: Condens. Matter, 2007, 19, 147 K. Fukao, Eur. Phys. J. E, 2003, 12, 119–125.
466112. 148 S. Napolitano, V. Lupas xcu and M. W€ ubbenhorst, Macromolecules,
107 J. Colmenero and A. Arbe, Soft Matter, 2007, 3, 1474. 2008, 41, 1061.
108 S.-H. Chong and M. Fuchs, Phys. Rev. Lett., 2002, 88, 185702. 149 R. Lund, L. Willner, A. Alegrı́a, J. Colmenero and D. Richter,
109 M. Aichele, S. H. Chong, J. Baschnagel and M. Fuchs, Phys. Rev. E: Macromolecules, 2008, 41, 511.
Stat., Nonlinear, Soft Matter Phys., 2004, 69, 061801. 150 A. Serghei, M. Tress and F. Kremer, Macromolecules, 2006, 39,
110 E. Flenner and G. Szamel, Phys. Rev. E: Stat., Nonlinear, Soft 9385.
Matter Phys., 2005, 72, 011205. 151 A. Serghei, L. Hartmann and F. Kremer, J. Non-Cryst. Solids, 2007,
111 E. Flenner and G. Szamel, Phys. Rev. E: Stat., Nonlinear, Soft 353, 4330.
Matter Phys., 2005, 72, 031508. 152 P. Scheidler, W. Kob and K. Binder, Europhys. Lett., 2002, 59,
112 G. Foffi, W. G€ otze, F. Sciortino, P. Tartaglia and T. Voigtmann, 701–707.
Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2004, 69, 011505. 153 P. Scheidler, W. Kob and K. Binder, Computer Simulations Surfaces
113 T. Voigtmann, A. M. Puertas and M. Fuchs, Phys. Rev. E: Stat., Interfaces, 2003, 114, 297–312.
Nonlinear, Soft Matter Phys., 2004, 70, 061506. 154 P. Scheidler, W. Kob and K. Binder, J. Phys. Chem. B, 2004, 108,
114 P. Chaudhuri, L. Berthier and W. Kob, Phys. Rev. Lett., 2007, 99, 6673–6686.
060604. 155 P. Gallo, R. Pellarin and M. Rovere, Europhys. Lett., 2002, 57, 212.
115 S. Merabia and D. Long, Eur. Phys. J. E, 2002, 9, 195. 156 J. A. Torres, P. F. Nealey and J. J. de Pablo, Phys. Rev. Lett., 2000,
116 S. K. Kumar, G. Szamel and J. F. Douglas, J. Chem. Phys., 2006, 85, 3221–3224.
124, 214501. 157 T. S. Jain and J. J. de Pablo, Phys. Rev. Lett., 2004, 92, 155505.
117 A. P. Sokolov and K. S. Schweizer, Phys. Rev. Lett., 2009, 102, 158 T. S. Jain and J. J. de Pablo, Macromolecules, 2002, 35, 2167–2176.
248301. 159 T. R. Bohme and J. J. de Pablo, J. Chem. Phys., 2002, 116,
118 M. D. Ediger, Annu. Rev. Phys. Chem., 2000, 51, 99–128. 9939–9951.
119 S. C. Glotzer, J. Non-Cryst. Solids, 2000, 274, 342–355. 160 K. Yoshimoto, T. S. Jain, P. F. Nealey and J. J. de Pablo, J. Chem.
120 L. Berthier, G. Biroli, J.-P. Bouchaud, W. Kob, K. Miyazaki and Phys., 2005, 122, 144712.
D. R. Reichman, J. Chem. Phys., 2007, 126, 184503. 161 R. A. Riggleman, K. Yoshimoto, J. F. Douglas and J. J. de Pablo,
121 L. Berthier, G. Biroli, J.-P. Bouchaud, W. Kob, K. Miyazaki and Phys. Rev. Lett., 2006, 97, 045502.
D. R. Reichman, J. Chem. Phys., 2007, 126, 184504. 162 A. R. C. Baljon, M. H. M. Van Weert, R. Barber DeGraaff and
122 S. Peter, H. Meyer and J. Baschnagel, Eur. Phys. J. E, 2009, 28, R. Khare, Macromolecules, 2005, 38, 2391.
147–158. 163 A. R. C. Baljon, J. Billen and R. Khare, Phys. Rev. Lett., 2004, 93,
123 L. Hartmann, K. Fukao and F. Kremer, Broadband Dielectric 255701.
Spectroscopy, Springer, Berlin, Heidelberg, New York, 2003, pp. 164 H. Morita, K. Tanaka, T. Kajiyama, T. Nishi and M. Doi,
433–473. Macromolecules, 2006, 39, 6233.

This journal is ª The Royal Society of Chemistry 2010 Soft Matter, 2010, 6, 3430–3446 | 3445
View Article Online

165 F. Varnik, J. Baschnagel and K. Binder, Phys. Rev. E: Stat., 209 G. J. Papakonstantopoulos, M. Doxastakis, P. F. Nealey,
Nonlinear, Soft Matter Phys., 2002, 65, 021507. J. L. Barrat and J. J. de Pablo, Phys. Rev. E: Stat., Nonlinear, Soft
166 F. Varnik, J. Baschnagel and K. Binder, Eur. Phys. J. E, 2002, 8, Matter Phys., 2007, 75, 031803.
175–192. 210 G. J. Papakonstantopoulos, K. Yoshimoto, M. Doxastakis,
167 F. Varnik and K. Binder, J. Chem. Phys., 2002, 117, 6336–6349. P. F. Nealey and J. J. de Pablo, Phys. Rev. E: Stat., Nonlinear,
168 F. Varnik, J. Baschnagel, K. Binder and M. Mareschal, Eur. Phys. J. Soft Matter Phys., 2005, 72, 031801.
E, 2003, 12, 167–171. 211 R. A. Riggleman, G. Toepperwein, G. J. Papakonstantopoulos,
169 S. Peter, H. Meyer and J. Baschnagel, J. Polym. Sci., Part B: Polym. J.-L. Barrat and J. J. de Pablo, J. Chem. Phys., 2009, 130, 244903.
Phys., 2006, 44, 2951–2967. 212 J. Berriot, H. Montes, F. Lequeux, D. Long and P. Sotta, Europhys.
170 S. Peter, H. Meyer, J. Baschnagel and R. Seemann, J. Phys.: Lett., 2003, 64, 50–56.
Condens. Matter, 2007, 19, 205119. 213 H. Montes, F. Lequeux and J. Berriot, Macromolecules, 2003, 36,
171 S. Peter, S. Napolitano, H. Meyer, M. Wubbenhorst and 8107–8118.
J. Baschnagel, Macromolecules, 2008, 41, 7729–7743. 214 R. A. Riggleman, H. N. Lee, M. D. Ediger and J. J. D. Pablo, Phys.
172 S. Peter, H. Meyer and J. Baschnagel, J. Chem. Phys., 2009, 131, Rev. Lett., 2007, 99, 215501.
Published on 01 April 2010. Downloaded by Gazi Universitesi on 09/10/2017 08:17:45.

014902. 215 M. Warren and J. Rottler, Phys. Rev. E: Stat., Nonlinear, Soft
173 S. Peter, H. Meyer and J. Baschnagel, J. Chem. Phys., 2009, 131, Matter Phys., 2007, 76, 031802.
014903. 216 L. Struik, Physical aging in amorphous polymers and other materials,
174 D. Hudzinskyy, A. V. Lyulin, A. R. C. Baljon, N. K. Balabaev and Elsevier, Amsterdam, 1978.
M. A. J. Michels, J. Polym. Sci. B, submitted. 217 H.-N. Lee, R. A. Riggleman, J. J. de Pablo and M. D. Ediger,
175 O. Alexiadis, V. G. Mavrantzas, R. Khare, J. Beckers and Macromolecules, 2009, 42, 4328–4336.
A. R. C. Baljon, Macromolecules, 2008, 41, 987–996. 218 R. A. Riggleman, J. F. Douglas and J. J. D. Pablo, Phys. Rev. E:
176 G. Xu and W. L. Mattice, J. Chem. Phys., 2003, 118, 5241. Stat., Nonlinear, Soft Matter Phys., 2007, 76, 011504.
177 K. F. Mansfield and D. N. Theodorou, Macromolecules, 1991, 24, 219 R. A. Riggleman, J. F. Douglas and J. J. de Pablo, J. Chem. Phys.,
6283–6294. 2007, 126, 234903.
178 E. Manias, V. Kuppa, D.-K. Yang and D. B. Zax, Colloids Surf., A, 220 J. Rottler and M. O. Robbins, Phys. Rev. Lett., 2002, 89, 148304.
2001, 187–188, 509. 221 J. Rottler and M. Robbins, Phys. Rev. E: Stat., Nonlinear, Soft
179 F. W. Starr, T. B. Schroder and S. C. Glotzer, Phys. Rev. E: Stat., Matter Phys., 2003, 68, 011507.
Nonlinear, Soft Matter Phys., 2001, 64, 021802. 222 J. Rottler and M. Robbins, Phys. Rev. E: Stat., Nonlinear, Soft
180 F. W. Starr, T. B. Schroder and S. C. Glotzer, Macromolecules, 2002, Matter Phys., 2003, 68, 011801.
35, 4481–4492. 223 J. Rottler and M. Robbins, Phys. Rev. Lett., 2005, 95, 225504.
181 H. Meyer, T. Kreer, A. Cavallo, J. Wittmer and J. Baschnagel, Eur. 224 A. Makke, M. Perez, O. Lame and J.-L. Barrat, J. Chem. Phys.,
Phys. J. Spec. Top., 2007, 141, 167–172. 2009, 131, 014904.
182 A. Cavallo, M. M€ uller, J. P. Wittmer, A. Johner and K. Binder, 225 R. N. Haward, Macromolecules, 1993, 26, 5860–5869.
J. Phys.: Condens. Matter, 2005, 17, S1697. 226 E. J. Kramer, J. Polym. Sci., Part B: Polym. Phys., 2005, 43,
183 S. Herminghaus, Eur. Phys. J. E, 2002, 8, 237. 3369–3371.
184 P. G. de Gennes, Eur. Phys. J. E, 2000, 2, 201. 227 H. G. H. van Melick, L. E. Govaert and H. E. H. Meijer, Polymer,
185 J. D. McCoy and J. G. Curro, J. Chem. Phys., 2002, 116, 9154. 2003, 44, 2493–2502.
186 T. M. Truskett and V. Ganesan, J. Chem. Phys., 2003, 119, 1897. 228 L. E. Govaert and T. A. Tervoort, J. Polym. Sci., Part B: Polym.
187 J. Mittal, P. Shah and T. M. Truskett, J. Phys. Chem. B, 2004, 108, Phys., 2004, 42, 2041–2049.
19769. 229 S. S. Sarva, S. Deschanel, M. C. Boyce and W. N. Chen, Polymer,
188 T. S. Chow, J. Phys.: Condens. Matter, 2002, 14, L333. 2007, 48, 2208–2213.
189 K. L. Ngai, Eur. Phys. J. E, 2003, 12, 93. 230 M. C. Boyce and E. M. Arruda, Polym. Eng. Sci., 1990, 30,
190 D. Long and F. Lequeux, Eur. Phys. J. E, 2001, 4, 371. 1288–1298.
191 S. Merabia, P. Sotta and D. Long, Eur. Phys. J. E, 2004, 15, 189. 231 B. Vorselaars, A. V. Lyulin and M. A. J. Michels, Macromolecules,
192 J. E. G. Lipson and S. T. Milner, Eur. Phys. J. B, 2009, 72, 133. 2009, 42, 5829–5842.
193 G. D. Smith, D. Bedrov and O. Borodin, Phys. Rev. Lett., 2003, 90, 232 O. B. Salamatina, S. I. Nazarenko, S. N. Rudnev and E. F. Oleinik,
226103. Mech. Compos. Mater., 1989, 24, 721–725.
194 V. Teboul and C. Alba-Simionesco, J. Phys.: Condens. Matter, 2002, 233 E. F. Oleinik, S. N. Rudnev, O. B. Salamatina, S. V. Shenogin,
14, 5699. M. I. Kotelyanskii, T. V. Paramzina and S. I. Nazarenko,
195 V. Krakoviack, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., E-Polymers, 2006, 29, 029.
2007, 75, 031503. 234 R. S. Hoy and M. O. Robbins, Phys. Rev. Lett., 2007, 99, 117801.
196 V. Krakoviack, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 235 R. S. Hoy and M. O. Robbins, Phys. Rev. E: Stat., Nonlinear, Soft
2009, 79, 061501. Matter Phys., 2008, 77, 031801.
197 J. Servantie and M. M€ uller, Phys. Rev. Lett., 2008, 101, 026101. 236 M. Utz, P. Debenedetti and F. Stillinger, J. Chem. Phys., 2001, 114,
198 M. Fuchs and K. Kroy, J. Phys.: Condens. Matter, 2002, 14, 9223. 10049–10057.
199 C. A. Angell, Polymer, 1997, 38, 6261. 237 R. S. Hoy and M. O. Robbins, J. Polym. Sci., Part B: Polym. Phys.,
200 A. Tanguy, J. P. Wittmer, F. Leonforte and J.-L. Barrat, Phys. Rev. 2006, 44, 3487–3500.
B: Condens. Matter Mater. Phys., 2002, 66, 174205. 238 L. J. Fetters, D. J. Lohse, D. Richter, T. A. Witten and A. Zirkel,
201 M. Tsamados, A. Tanguy, C. Goldenberg and J.-L. Barrat, Phys. Macromolecules, 1994, 27, 4639–4647.
Rev. E: Stat., Nonlinear, Soft Matter Phys., 2009, 80, 026112. 239 L. J. Fetters, D. J. Lohse, S. T. Milner and W. W. Graessley,
202 K. Yoshimoto, G. Papakonstantopoulos, J. Lutsko and J. de Pablo, Macromolecules, 1999, 32, 6847–6851.
Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 184108. 240 J. I. McKechnie, R. N. Haward, D. Brown and J. H. R. Clarke,
203 K. Yoshimoto, T. Jain, K. Workum, P. Nealey and J. de Pablo, Macromolecules, 1993, 26, 198–202.
Phys. Rev. Lett., 2004, 93, 175501. 241 A. V. Lyulin, N. K. Balabaev, M. A. Mazo and M. A. J. Michels,
204 K. Van Workum and J. de Pablo, Nano Lett., 2003, 3, 1405–1410. Macromolecules, 2004, 37, 8785–8793.
205 J.-L. BarratModeling And Simulation Of New Materials, AIP 242 B. Vorselaars, A. V. Lyulin and M. A. J. Michels, J. Chem. Phys.,
Conference Proceedings 1091, ed. Garrido et al., 2009, vol. 1091, 2009, 130, 074905.
pp. 79–94. 243 H. G. H. van Melick, L. E. Govaert and H. E. H. Meijer, Polymer,
206 S. G. Mayr, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 2003, 44, 3579–3591.
060201. 244 J. Mark, Physical Properties of Polymers Handbook, Springer,
207 G. J. Papakonstantopoulos, R. A. Riggleman, J. L. Barrat and Berlin, 2007.
J. J. de Pablo, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 245 A. J. Moreno and J. Colmenero, Phys. Rev. Lett., 2008, 100, 126001.
2008, 77, 041502. 246 T. Vettorel, H. Meyer, J. Baschnagel and M. Fuchs, Phys. Rev. E:
208 T. S. Jain and J. J. de Pablo, J. Chem. Phys., 2004, 120, 9371–9375. Stat., Nonlinear, Soft Matter Phys., 2007, 75, 041801.

3446 | Soft Matter, 2010, 6, 3430–3446 This journal is ª The Royal Society of Chemistry 2010

You might also like