You are on page 1of 23

PC62CH26-de-Pablo ARI 25 February 2011 16:25

ANNUAL
REVIEWS Further Coarse-Grained Simulations
Click here for quick links to
Annual Reviews content online,
including:
of Macromolecules: From DNA
• Other articles in this volume
• Top cited articles
to Nanocomposites
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

• Top downloaded articles


• Our comprehensive search
Juan J. de Pablo
Department of Chemical and Biological Engineering, University of Wisconsin–Madison,
Madison, Wisconsin 53706; email: depablo@engr.wisc.edu
by University of Leeds on 06/23/13. For personal use only.

Annu. Rev. Phys. Chem. 2011. 62:555–74 Keywords


First published online as a Review in Advance on molecular simulations, polymers, copolymers, nanoparticles
January 10, 2011

The Annual Review of Physical Chemistry is online at Abstract


physchem.annualreviews.org
This review discusses multiscale modeling and simulations of macro-
This article’s doi: molecules and macromolecular systems in the context of two specific exam-
10.1146/annurev-physchem-032210-103458
ples. In the first, recent attempts to develop coarse-grained representations
Copyright  c 2011 by Annual Reviews. of DNA are reviewed, and a discussion of recent predictions of such models
All rights reserved
is presented, particularly in the context of DNA melting and rehybridization.
0066-426X/11/0505-0555$20.00 The second example considers polymer nanocomposites; a review of recent
simulations is presented, with an emphasis on the description of entangle-
ments in such systems and new methods for the study of the segregation of
nanoparticles that arises in copolymers, in which composition heterogene-
ity can be used to control nanoparticle position and develop an increased
understanding of nanoparticle-polymer interactions.

555
PC62CH26-de-Pablo ARI 25 February 2011 16:25

INTRODUCTION
Multiscale modeling has come of age over the past decade. A number of reviews on hierarchical
modeling, coarse graining, and multiscale methods for soft matter have appeared over the past
several years (1–26). A broad range of topics has been covered, including proteins, glasses, mem-
branes, and polymers. The aim of this review is not to reproduce what can be found in existing
reviews, but to focus on polymeric systems whose function arises over relatively large scales or do-
mains and emerges from specific molecular-level considerations. For concreteness, the discussion
is centered on two specific classes of molecules or materials that have not been discussed before
in the context of a multiscale-modeling review and that represent new and promising trends in
the development of coarse-graining strategies. The first system discussed concerns DNA and its
mesoscale description; the main issue here is to capture the functionality of the molecule that
emerges from its sequence. The second system considered concerns composites of nanoparticles
and polymeric molecules—nanocomposites; the issue here is to understand and predict the be-
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

havior, structure, and directed assembly of the resulting nanocomposite that arises over relatively
long length scales. As shown below, as important as these systems are in science and technology,
their modeling remains in its infancy. The aim of the discussion that follows is to stimulate interest
in the modeling and simulation of functional materials, including DNA and nanocomposites, and
by University of Leeds on 06/23/13. For personal use only.

to provide readers with a starting point for further studies.


Beyond their scientific and technical relevance, DNA and nanocomposites represent a partic-
ularly useful choice for a discussion of modeling and simulation of macromolecules because they
each highlight a different set of needs and challenges. In the case of DNA, one seeks to develop
reduced models capable of describing the many functions of DNA. In the case of nanocomposites,
the challenge is to develop methods or approaches that enable the simulation of these materials,
particularly under deformation.
At the outset I emphasize that the review that follows is not comprehensive; the issues and the
examples discussed here are biased by my current interests. Whenever possible, however, I have
made reference to a number of recent, authoritative reviews and texts on multiscale or hierarchical
modeling of soft materials that offer readers a broader perspective than that presented here.

COARSE-GRAINED MODELS OF DNA


Several models of DNA are available in the literature. These range from fully atomistic repre-
sentations, in which all atoms (including those of the solvent) are considered explicitly, to highly
coarse grained, in which collections of several hundred atoms are represented by a few spherical
beads connected by worm-like chain springs. Although a complete description in terms of all
atomic coordinates would at first glance appear desirable, as more chemical detail is included in a
model the computational requirements associated with its solution increase significantly, thereby
restricting severely the length scales and timescales amenable to study. The challenge is therefore
to include just enough detail in a model to capture the physics that are responsible for DNA’s
relevance in biology (27).
Atomistic models based on force fields such as CHARMM (28) and AMBER (29) provide
the highest degree of detail. Recent reviews have presented a useful perspective of the uses and
limitations of such models (30–32). Such reviews and recent calculations show that comprehensive
studies of long DNA molecules, including hybridization and packaging, continue to be well beyond
the reach of fully atomistic representations.
At the other end of the spectrum of length scales, e.g., for study of full genomic DNA, several
models have been proposed and have shed considerable light onto the dynamical behavior of DNA

556 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

in various environments (33–47). The persistence length of DNA is in the vicinity of 50 nm; such
models are relatively coarse in that they represent several dozen persistence lengths with spherical
beads connected to one another by a worm-like chain spring. In spite of their low resolution, they
yield results in excellent agreement with experimental data for diffusion, structural relaxation, and
behavior under different flow fields for bulk and confined DNA.
For many applications of interest, however, the approaches mentioned above, either atomistic
or bead/spring, are inadequate. Let us consider, for example, the formation of structured materials
through DNA-enabled colloidal assembly. In that case linker DNA (48) comprising on the order
of 10 to 30 nucleotides is used to bind nanoparticles; to improve assembly strategies, it is of
interest to understand hybridization for such molecules. Or let us consider the packaging of DNA
in chromatin, a fundamental element of biology in which a DNA of approximately 150 base pairs
must tightly wrap around a small protein complex to form the nucleosome (49). In each of these
applications, relevant phenomena or processes occur on length scales and timescales commensurate
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

with the contour length of the molecules and their longest relaxation time, but essential effects such
as hybridization or melting occur at a local level. The need for an intermediate-level description
of DNA has led to the development of mesoscale models of DNA (47, 50–62). Such models have
been successful in reproducing distinct features of DNA but do not provide a comprehensive
by University of Leeds on 06/23/13. For personal use only.

description of melting, hybridization, mechanical properties, and electrostatic effects within one
representation.
Recently, we have proposed a coarse-grained representation of DNA to do just that. Below
we describe such a model, some of its predictions, and its limitations. As shown in Figure 1, the
model represents each nucleotide using three interaction sites, and so we refer to it as 3SPN (three
sites per nucleotide) (63–65). There are four different base sites, one for each type of base in DNA.
The backbone phosphate and sugar sites are placed at the center of mass of the respective moiety.
For purine bases (adenine and guanine), the site is placed at the N1 position. For pyrimidine bases
(cytosine and thymine), the site is placed at the N3 position (63). The geometrical features of
the model are important in at least two respects. First, they offer a possibility for reverse coarse
graining or inverse mapping of the model for coupling between different length scales. The three
sites provide the necessary scaffolding to selectively reconstruct an atomistic representation on a
local level while keeping the remainder of the molecule coarse grained. Second, such features are
necessary for investigations of protein/DNA association. The ability of binding proteins to identify
the grooves is a key element in their complexation to double-stranded (ds)DNA. The proposed
model displays these motifs and facilitates studies of protein/DNA interactions that cannot be
pursued with coarser models of DNA. The details of the model and the corresponding potential
energy parameters can be found in the literature (64, 65). Here we provide a brief account of its
predictions.
Figure 2 shows melting curves for short DNA strands. Simulations are able to capture both the
width of the transition and the effect of salt concentration. The model is also able to describe the
persistence length of DNA and its dependency on ionic strength. In this regard, it is of interest to
mention that the development of the model was largely enabled by the use of efficient Monte Carlo
simulation algorithms developed over the past decade, in particular, the use of expanded ensemble
(66–68) and density-of-states ideas (69, 70), parallel tempering (71–76), and histogram reweighting
(77); without these, melting calculations would be much more computationally intensive, and the
entire parameterization of the model would have been difficult.
As alluded to above, one of the main goals of coarse-grained models is to describe the func-
tionality of DNA. Much of that functionality results from DNA’s ability to hybridize. Figure 3
shows a free energy profile as a function of the number of native contacts, which was found to be
a convenient order parameter for the calculation of melting curves (64). As seen in the figure, a

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 557


PC62CH26-de-Pablo ARI 25 February 2011 16:25

O
Phosphate
NH
O
HO P O N O
O– O Base

Sugar OH
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org
by University of Leeds on 06/23/13. For personal use only.

Figure 1
Schematic representation of the 3SPN model, in which each nucleotide is represented by three interaction
sites corresponding to the phosphate, the base, and the sugar.

relatively large free energy arises when two single-stranded molecules form two native contacts;
as more contacts form, the free energy decreases steadily until the molecule is fully hybridized. Al-
though the order parameter for Figure 3 is not a true reaction coordinate, the results do suggest
that hybridization is an activated process. More recent calculations, using transition path sam-
pling simulations (78), have revealed an intriguing picture of the actual process of hybridization
with the simple coarse-grained model shown in Figure 1 (79). Figure 4 shows the probability of
forming distinct contacts between specific bases belonging to the two strands of a 30-base-pair
dsDNA molecule in the transition-state ensemble corresponding to the hybridization reaction.
The molecules do not form native contacts at the transition state but instead prefer to be offset
from the native duplex by approximately three to five nucleotides. The details of the contact map
are highly sensitive to sequence, and they depend on temperature and salt concentration. A gen-
eral feature, however, is that random sequences of DNA exhibit sharp free energy barriers toward
hybridization, reminiscent of those shown in Figure 3, whereas repetitive sequences follow a
diffusive pathway toward hybridization (79).
For numerous applications, including genomic analysis or colloidal assembly (48, 80), it is
of interest to consider hybridization when one of the strands is immobilized on a surface (or a

558 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

Simulation (20 mM)


Simulation (50 mM)
Simulation (120 mM)
1.0
Experiment (20 mM)
Experiment (50 mM)
0.8 Experiment (120 mM)

0.6

f
0.4
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

0.2

0.0
by University of Leeds on 06/23/13. For personal use only.

280 300 320 340 360


Temperature (K)
Figure 2
Melting curves for DNA oligonucleotides. The symbols represent experimental data, and the lines show
results of the 3SPN model. In the ordinate axis, f denotes the extent of denaturation of the molecule, with 0
denoting fully hybridized DNA and unity corresponding to fully dehybridized DNA.

297 K, [NaCl] 69 mM

4
βA

275 K, [NaCl] 5 mM
2

0 2 4 6 8 10 12 14
ξ
Figure 3
Free energy profile for hybridization of a 15-base-pair DNA oligonucleotide as a function of the number of
native contacts, denoted by ξ . The blue curve was calculated at 297 K and a salt concentration (NaCl) of
69 mM. The red curve was calculated at T = 275 K and a salt concentration of 5 mM.

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 559


PC62CH26-de-Pablo ARI 25 February 2011 16:25

T
C RAN30
G all nξ
G
A
C
T
T
C
A
A
G
A
G
Chain 1 T
C
T
A
G
G
T
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

C
T
G
G
T 0.0 0.2 0.4 0.6 0.8 1.0
C Probability
by University of Leeds on 06/23/13. For personal use only.

T
G
A
T C AGA C C AGA C C T AGA C T C T T GA AG T C C GA
Chain 2
Figure 4
Contact map showing the probability that distinct bonds are formed between two complementary 30-base
strands of DNA. The contact map was constructed using configurations from the transition-state ensemble
for the hybridization reaction. The color key provides a measure of probability, and the sequence of the two
strands is provided on the corresponding axis.

particle) (81). Figure 5 shows contact maps for a duplex consisting of 15 bases. In one case the
molecule is in the bulk, and in the other, one of the strands is attached to a surface. Two features
of Figure 5 are worth highlighting. First, the melting temperature of the surface-bound dsDNA
molecule is higher than that in bulk solution, by approximately 20◦ to 30◦ , depending on sequence.
Second, the contact map corresponding to the surface-bound molecule is significantly different
from that in solution. On the surface, the contacts formed at the transition state are predominantly
native contacts, and they occur primarily toward the middle of the molecules. As before, these
contact maps are particularly sensitive to temperature, sequence, and solution conditions (e.g.,
ionic strength).
The results shown in Figures 2–5 serve to illustrate that coarse-grained models, as simple
as they might still be, are beginning to enable comprehensive studies of DNA and are starting
to deliver fundamental insights into its behavior that have been lacking. There is considerable
room for improvement, however, and to the extent that models are refined and tested, they might
contribute significantly to emerging, DNA-based technologies. One of the shortcomings of the
model outlined in Figure 1 is that base-stacking interactions are represented with a Go-like
potential (82). For predictions of new, tailor-made structures of DNA, it would be important to
remove such a constraint and include the planar geometry of the bases. Recent efforts in that
direction are promising. In a new model, based on the 3SPN approach, the spherically symmetric
base site is replaced by a rigid-body ellipsoid (83). The model can predict a number of structural
features of dsDNA without the need for potentials that bias the structure in the direction of
B-DNA, but it is unclear whether it can predict melting temperatures and rehybridization. In a

560 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

A a A b
T T
C C
A A
A A
T T
T T
A A
C C
A A
A A
T T
C 0.0 0.2 0.4 0.6 0.8 1.0 C
A Probability A
T T
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

A T G A T T G T A A T T G A T A T G A T T G T A A T T G A T

Figure 5
Contact maps showing the probability that distinct bonds are formed between two complementary 15-base
by University of Leeds on 06/23/13. For personal use only.

strands of DNA. The contact map was constructed using configurations from the transition-state ensemble
for the hybridization reaction. The color key is the same as that used in Figure 4. The results in panel a
correspond to a situation in which the strand on the abscissa was attached to the surface at the blue site, T.
The results in panel b are for the same molecule under the same conditions, but in free solution.

different model, also based on the 3SPN platform, the number of interaction sites is expanded
from three to six, thereby providing a more complete geometrical representation of the bases
(84). It is able to describe the structure of DNA, and preliminary melting temperature predictions
have been reported to be in qualitative agreement with experiment. Such expanded models are
more computationally demanding, and the issue now is to determine whether they can predict the
structure of other forms of DNA, beyond B-DNA, and whether they can be further developed to
describe related molecules, in particular RNA. Another shortcoming of 3SPN and most existing
models based on the same platform is that electrostatic interactions are treated at the Debye-
Huckel level. For numerous situations, it is of interest to determine the role of specific counterions,
and extensions that replace the Debye-Huckel form with explicit ions would be useful. Recent work
has sought to modify 3SPN by including explicit counterions in studies of particle dimerization
(85); additional efforts in that direction are needed.
Future coarse-grained developments should also consider DNA-protein interactions, and the
delicate issues that arise vis-à-vis electrostatic effects in such systems. Recent efforts in that di-
rection are promising (49, 86–88) and suggest that coarse-grained models will be useful, if not
essential, in interpreting emerging experimental data in the biological literature.

COARSE-GRAINED MODELS OF POLYMERIC NANOCOMPOSITES


The central idea of nanocomposites is that, through the addition of a relatively small amount
of nanoparticles, it is possible to improve significantly the properties of a polymer melt. One of
the central challenges of nanocomposite research has been that of dispersing the nanoparticles
in the polymeric matrix (89). It is known that the aggregation state of the particles in the matrix
plays a key role in the behavior of the material; it is now appreciated that the performance of a
nanocomposite depends on whether the particles are uniformly dispersed (90–92). In some cases,
for example, mechanical strength can be increased by particles that aggregate to form higher-level
structures (92). Other properties, such as electrical conductivity, require that particles be dispersed

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 561


PC62CH26-de-Pablo ARI 25 February 2011 16:25

effectively in order to form percolating structures at low loadings (91). With the emergence of
better processing methods to achieve particle dispersion or aggregation within a polymer matrix
(89, 92), one of the central questions that arises is how the particle’s spatial distribution influences
a nanocomposite’s properties.
With the recent development of new experimental protocols, it is becoming possible to create
materials that exhibit remarkable properties. The challenge for theory and simulation is to de-
velop models capable of describing the structure and properties of such materials—such models
are necessary to interpret experimental data and to formulate a theory of nanocomposite behavior.
Only recently has a molecular theory of composites begun to emerge (93); some of the central
contributions of that theory have been to gradually provide a view of the structure and phase be-
havior of composites (94–97) that is comprehensive enough to explain a wide range of experimental
observations. Note, however, that recent theoretical progress based on integral equation, density
functional, and self-consistent mean field methods has been largely limited to equilibrium, and not
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

dynamics. A molecular theory for nanocomposites under deformation remains to be developed.


Simulations could fill an important gap in our understanding of polymer nanocomposites by
providing a molecular view of nanoparticle distribution, of macromolecular conformations around
the particles, and of the entanglement network that arises in such materials and by providing
by University of Leeds on 06/23/13. For personal use only.

insights into their mechanical or rheological response under deformation. A recent review has
discussed past efforts in simulating nanocomposites at the atomistic, mesoscale, and continuum
levels (20). On the one hand, given that the typical size of a nanoparticle is in the range of 5 to
100 nm, atomistic simulations have mainly focused on studies of the organization of molecules in
the immediate vicinity of the particles; several recent examples can be found in References 98–105.
Continuum studies, on the other hand, have focused on the mechanical behavior of the composites
but have not addressed the polymeric, molecular nature of the underlying materials (20). Hybrid
approaches that combine atomistic modeling and continuum methods have been limited, but offer
significant promise in nanocomposite research, particularly with regard to mechanical property
prediction (106). Attempts to describe the structure and properties of nanocomposites solely on
the basis of atomistic simulations face considerable challenges with regard to the size of the system
that can be investigated (107), and in order to describe their intermediate-to-long length scale
structure, one must adopt coarse-grained representations. As mentioned above, the challenge in
this area is not as much developing such representations, as it is developing algorithms capable of
simulating them.
Coarse-grained simulations of polymeric nanocomposites have generally resorted to lattice
models (108, 109) or to bead-spring models (107, 110–117), in which individual beads correspond
to a Kuhn segment of the polymer. In the latter case the particles are represented by spheres or
collections of spherical interaction sites connected by springs. The majority of the simulations
available in the literature have considered relatively short, unentangled molecules (108–112, 116).
Many of the unusual properties of nanocomposites, however, arise from such entanglements, and,
as alluded to above, it is therefore of interest to determine if and how nanoparticles influence the
entanglement network of a composite (118).
Only recently have combinations of molecular dynamics and Monte Carlo simulations been
able to examine the structure and elastic properties of entangled polymer composites (113, 119),
both at rest and under deformation. Advanced Monte Carlo moves, including connectivity-altering
techniques (120, 121), have been essential to ensure that the conformations of the polymers
around the nanoparticles are sampled correctly, thereby providing a faithful representation of
the entanglement network that arises in model systems. Some representative results are shown in
Figure 6 for the strain as a function of time under constant tensile and compressive stresses.
Simulations have shown that, for well-dispersed particles, the elastic moduli of the composite are

562 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

0.8
Pure
0.6

0.4 Nanocomposite
Strain

0.2

0 Nanocomposite

–0.2
Pure
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

–0.4

0 2,000 4,000 6,000 8,000


Time
by University of Leeds on 06/23/13. For personal use only.

Figure 6
Strain as a function of time for constant-stress deformation of a model nanocomposite polymeric material.
The chains comprise 500 monomers, and the volume fraction of particles is 15%. The diameter of the
particles is five times that of the polymer monomer. The upper curves show results under tension, whereas
the bottom ones show result under compression. The results in blue correspond to deformation of the
material without nanoparticles, and those in red correspond to the nanocomposite.

higher than those of the corresponding pure polymer matrix. Consistent with experiment, simu-
lations under deformation have shown that the toughening modulus increases with the addition of
nanoparticles. Some of the central findings of this more recent work on entangled nanocomposites
have been that well-dispersed nanoparticles serve as entanglement attractors or traps. Figure 7
shows an image of the primitive path (122–127) that long polymeric molecules follow in a compos-
ite melt; one can see that long polymers wrap around the particles, thereby increasing the number
of entanglements in the nanocomposite. Furthermore, as the material is deformed, e.g., under

a b

Figure 7
Primitive path followed by polymer molecules in the vicinity of a nanoparticle dispersed in a polymer melt.
(a) A nanocomposite at equilibrium. (b) The same nanoparticle after a tensile deformation of the material, at
constant stress, and a strain of 0.5.

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 563


PC62CH26-de-Pablo ARI 25 February 2011 16:25

elongation or under compression, new entanglements are formed around the nanoparticle (119).
Such an increased number of entanglements leads to significantly different mechanical and rheo-
logical properties, including a higher plateau modulus (119, 128). Recent studies have also found
that particle size and shape can increase the number of entanglements considerably, thereby lead-
ing to composite materials with considerably different rheological properties, particularly under
deformation (107, 111, 115, 129).
Other important issues in nanocomposite research are to understand how nanoparticles are
distributed throughout the material and, perhaps more importantly, to use that understanding
to manipulate and control their positioning. Block copolymer materials have been particularly
instructive in that regard (130–136). Indeed, with an inherent structural inhomogeneity having
characteristic dimensions in the 5- to 100-nm range, they provide an ideal platform in which to
explore and further our understanding of composites. Theoretical work has been limited (137–
139), and simulations of nanoparticles in block copolymers have been scarce. Early simulations
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

employed lattice models to examine how dilute nanoparticles distribute themselves in a copolymer
(109, 137). Although such calculations were useful in helping to explain a number of experimental
features, including the preference of nanoparticles for interfacial or bulk regions, depending on
the nature of chemical interactions and size, they were limited to small systems. Furthermore,
by University of Leeds on 06/23/13. For personal use only.

by virtue of the underlying lattice, the geometrical shape of the particles was generally cubic
based. More recently, new approaches based on well-defined theoretical formalisms have enabled
coarse-grained simulations over extended regions of space. In one approach, based on a field-
theoretic description of the fluid in which particles are treated explicitly in the fields created by
the copolymer molecules (140), the particle coordinates and chemical potential field variables are
simultaneously updated, and the calculations amount to conducting Brownian dynamics simula-
tions of the particles in an implicit solvent (whose local composition is dictated by the copolymer).
Simulations using this approach have so far been performed on a lamellar phase of a copolymer, in
two dimensions; the results have been shown to be in good agreement with experiments conducted
on polystyrene-functionalized gold nanoparticles in a polystyrene-polyvinylprirrolidone diblock
copolymer melt.
In an approach similar in spirit but different in implementation, simulations of the copolymer
and the particles are conducted explicitly, but the interactions between copolymer molecules and
the particles and the copolymer are given by a coarse-grained Hamiltonian akin to that used
in field-theoretic simulations (5, 141) (or in self-consistent field theory). In other words, the
coarse-grained interactions between particles are informed by an underlying theoretical model,
thereby reducing considerably the computational requirements and avoiding the need to evaluate
pair interactions. Theoretically informed simulations of nanoparticles in copolymers have been
performed on both bulk copolymers (141) and on thin films of copolymers (132, 142), in three
dimensions. The results of such simulations also have been shown to be in good agreement with
experiment.
A particularly interesting system, which serves to illustrate the value of using block copolymers
to further our understanding of nanoparticle-polymer interactions, is described schematically in
Figure 8. In that system, a block copolymer thin film is deposited on a patterned substrate. The
role of the pattern is to control the morphology of the copolymer thin film. In the example shown
in Figure 8, which is concerned with lamellae, the pattern on the surface can stretch or compress
individual domains of the copolymer, thereby inducing pronounced conformational changes of
the molecules. When nanoparticles are added to the thin films, they serve as probes of copolymer
structure and exhibit a distinct preference for certain regions of the material. As the copolymer
morphology is perturbed by the underlying substrate pattern, the particles adopt different arrange-
ments, sometimes forming a single row, two parallel rows, or aggregates that can be visualized in

564 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

a L0
b
0% 20% 40% φhom

Ls = L0

Ls > L0

200 nm
Stretching
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

LS

Figure 8
by University of Leeds on 06/23/13. For personal use only.

(a) Schematic representation of a copolymer thin film deposited on a stripe-patterned substrate.


Nanoparticles ( green) are preferentially dispersed in the pink domains of the diblock. The pink block
exhibits a preference for the pink regions of the substrate, and the blue blocks exhibit a preference for the
blue stripes. (b) Scanning electron micrographs of cadmium-selenide nanoparticles after dispersion in
poly-methyl-methacrylate-polystyrene diblock copolymer thin films. Results are shown for different values
of Ls and different homopolymer concentrations.

scanning electron micrographs. Figure 9 shows results of experiments and theoretically informed
simulations of small cadmium-selenide nanoparticles (5 nm in diameter), treated in such a way as to
exhibit a preference for polystyrene, dispersed in a poly-methyl-methacrylate-polystyrene diblock
copolymer thin film. As seen in the figure, the agreement between simulation and experiment is
good, particularly when one considers the relatively coarse nature of the theory.
Theoretically informed simulations of nanocomposites continue to face challenges and could
be improved in a number of respects. In the hard-particle limit (142), it is currently computa-
tionally demanding to simulate large particles and large-particle volume fractions. That problem
could be alleviated by developing new Monte Carlo moves to facilitate trial particle displacements
while lowering rejection rates. Perhaps more importantly, it is of considerable importance now to
develop approaches that will enable studies of deformation and dynamics of nanocomposites, in
the entangled regime, within the framework of theoretically informed or field-theoretic simulation
approaches.

CONCLUSIONS
A striking feature of the results shown in Figures 8 and 9 concerns the length scales that can now be
probed by simulations and experiments. We have entered an age in which molecular simulations of
relatively complex materials are feasible over large domains, encompassing the molecular details
of a macromolecule’s conformation and the broad features of structures or morphologies that
arise over micrometer regions. Experiments can now manipulate and probe matter at those length
scales, thereby enabling direct comparisons to models, and a more refined and direct interpretation
of observed phenomena in terms of well-defined molecular characteristics and processes.
This feature, which in this review has been highlighted in the context of our own experience
with block copolymer directed assembly, is not limited to this class of materials. Multiple areas

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 565


PC62CH26-de-Pablo ARI 25 February 2011 16:25

0% 20% 40% φhom

0.10

Simulation φ(x) 0.05


Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

0.10
Experiment φ(x)
by University of Leeds on 06/23/13. For personal use only.

0.05

–20 –10 0 10 20 –20 –10 0 10 20 –20 –10 0 10 20


x (nm)

Figure 9
Particle concentration profiles as a function of position from simulations (top panels) and from experiments
(lower panels; see Figure 8). Different lines represent different degrees of mismatch (or stretching) between
the characteristic dimensions of the substrate pattern (Ls ) and the characteristic period of the block
copolymer in the bulk (L0 ). The black line corresponds to Ls = L0 , the blue line to Ls = 1.1L0 , and the
green line to Ls = 1.2L0 .

of soft matter, ranging from colloidal systems to protein extension, have highlighted the synergy
between concurrent experimental measurements and coarse-grained simulations. A decade of
single-molecule experiments (143, 144) and coarse-grained simulations with DNA, for example,
has benefited tremendously from the direct observation of molecular features coupled to coarse-
grained simulations (36, 145, 146).
With this increased ability to interrogate matter at nanometer length scales, new opportunities
will arise to further our understanding of macromolecular materials. Simulations of appropriately
coarse-grained models will be particularly important for the interpretation of experimental data.
Two areas of macromolecular simulation that, in spite of important efforts, continue to be in
need of additional developments involve the structure and dynamics of DNA complexes and the
deformation and dynamics of nanocomposites.
As mentioned above, DNA is increasingly used in the form of linkers to form intricate, higher-
order structures. An elementary unit of such structures often consists of a particle to which single
strands of DNA are attached. The pathways through which two DNA functionalized nanoparticles
hybridize, for example, are not fully understood. Multiscale simulations using some of the models
introduced in this review could shed considerable light onto such processes.

566 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

In chromatin, for example, a single molecule is wrapped around thousands of histones in a tight
complex whose structure continues to be a source of much debate. As it folds around these proteins
(87), DNA might partially melt and do so according to local sequence, concentration, and other
factors. Unraveling that structure and the dynamics of the packaging process will require that the
approaches described in this review be tested and expanded within the context of DNA-protein
interactions. Work of this nature might also require that one resorts to hierarchical approaches,
in which different levels of description communicate with each other and are used to address
systematically longer or shorter length scales in a concerted manner.
Going beyond DNA to RNA, which exhibits a rich variety of folded structures (152), it will
be important to develop models that do not have some of the site-specific restrictions of existing
mesoscale models of DNA. As mentioned above, recent attempts to remove such restrictions
in the context of DNA are promising (84), but the resulting models are more computationally
demanding and need to be further explored and extended to address the challenges that arise in
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

RNA research. This is a fertile area of research in which multiscale simulations have the potential
to make major contributions.
Within the realm of nanocomposites, equilibrium simulations have been largely limited to small
particle sizes and relatively low particle volume fractions. As mentioned above, key challenges are
by University of Leeds on 06/23/13. For personal use only.

to simulate large systems (so that many large particles can be incorporated into a sample) and to
develop algorithms that allow one to sample configuration space through major displacements of
relatively large particles in a sea of other particles and long polymeric molecules. Recent algorithms
such as the bijective mapping approach of Uhlherr & Theodorou (153), coupled to the coarse-
graining strategies discussed in the context of theoretically informed techniques, offer promise
on that front. Clearly, simulations that overcome such limitations and that manage to identify
correlations between measurable properties and particle characteristics and organization would
be particularly useful.
In the case of dynamic or nonequilibrium properties, it is important to point out that me-
chanical or rheological experimental data are generated under conditions that are not accessible
to conventional simulations. Relevant experimental timescales and length scales in a rheometer
are orders of magnitude longer than those probed by simulations of many-particle systems in
nanometer-length-scale simulation boxes. In order to sample longer, relevant scales, simulations
and models aiming to reproduce and interpret such data must deal with the issue of entanglements,
their characteristics, and how these are related to the material’s response under stress. Recent work
with pure polymers suggests that the concept of slip links (147–150) might be particularly useful in
this regard, particularly if coupled to a real-space representation of the molecules (151), as is done
in theoretically informed simulations. Future simulations that incorporate such concepts, partic-
ularly in the context of structured polymers and nanocomposites, would help address a number of
longstanding questions in the general area of nanocomposite rheology and processing.

DISCLOSURE STATEMENT
The author is not aware of any affiliations, memberships, funding, or financial holdings that might
be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
The author is grateful for support from the National Science Foundation and the Semiconductor
Research Corporation. The author is also grateful to the European Community for a Marie Curie

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 567


PC62CH26-de-Pablo ARI 25 February 2011 16:25

Fellowship and to Prof. Manuel Laso for helpful discussions and his hospitality during a sabbatical
stay at the Universidad Politecnica de Madrid.

LITERATURE CITED
1. Baeurle SA. 2009. Multiscale modeling of polymer materials using field-theoretic methodologies: a survey
about recent developments. J. Math. Chem. 46:363–426
2. Bennun SV, Hoopes MI, Xing CY, Faller R. 2009. Coarse-grained modeling of lipids. Chem. Phys. Lipids
159:59–66
3. Bouvard JL, Ward DK, Hossain D, Nouranian S, Marin EB, Horstemeyer MF. 2009. Review of hierar-
chical multiscale modeling to describe the mechanical behavior of amorphous polymers. J. Eng. Mater.
Technol. 131:041206
4. de Pablo JJ, Curtin WA. 2007. Multiscale modeling in advanced materials research: challenges, novel
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

methods, and emerging applications. MRS Bull. 32:905–11


5. Detcheverry FA, Pike DQ, Nagpal U, Nealey PF, de Pablo JJ. 2009. Theoretically informed coarse grain
simulations of block copolymer melts: method and applications. Soft Matter 5:4858–65
6. Ghoniem NM, Busso EP, Kioussis N, Huang HC. 2003. Multiscale modelling of nanomechanics and
by University of Leeds on 06/23/13. For personal use only.

micromechanics: an overview. Philos. Mag. 83:3475–528


7. Ghosh J, Wong BY, Sun Q, Pon FR, Faller R. 2006. Simulations of glasses: multiscale modeling and
density of states Monte-Carlo simulations. Mol. Simul. 32:175–84
8. Guenza MG. 2008. Theoretical models for bridging timescales in polymer dynamics. J. Phys. Condens.
Matter 20:033101
9. Hills RD, Brooks CL. 2009. Insights from coarse-grained Go models for protein folding and dynamics.
Int. J. Mol. Sci. 10:889–905
10. Karakasidis TE, Charitidis CA. 2007. Multiscale modeling in nanomaterials science. Mater. Sci. Eng. C
27:1082–89
11. Muller-Plathe F. 2002. Coarse-graining in polymer simulation: from the atomistic to the mesoscopic
scale and back. Chemphyschem 3:754–69
12. Muller-Plathe F. 2003. Scale-hopping in computer simulations of polymers. Soft Mater. 1:1–31
13. Murtola T, Bunker A, Vattulainen I, Deserno M, Karttunen M. 2009. Multiscale modeling of emergent
materials: biological and soft matter. Phys. Chem. Chem. Phys. 11:1869–92
14. Pandit SA, Scott HL. 2009. Multiscale simulations of heterogeneous model membranes. Biochim. Biophys.
Acta Biomembr. 1788:136–48
15. Peter C, Kremer K. 2009. Multiscale simulation of soft matter systems: from the atomistic to the coarse-
grained level and back. Soft Matter 5:4357–66
16. Sherwood P, Brooks BR, Sansom MSP. 2008. Multiscale methods for macromolecular simulations. Curr.
Opin. Struct. Biol. 18:630–40
17. Tozzini V. 2009. Multiscale modeling of proteins. Acc. Chem. Res. 43:220–30
18. Barrat JL, Baschnagel J, Lyulin A. 2010. Molecular dynamics simulations of glassy polymers. Soft Matter
6:3430–46
19. Yang L-W, Chng C-P. 2008. Coarse-grained models reveal functional dynamics—I. Elastic network
models—theories, comparisons and perspectives. Bioinform. Biol. Insights 2:25–45
20. Zeng QH, Yu AB, Lu GQ. 2008. Multiscale modeling and simulation of polymer nanocomposites. Prog.
Polym. Sci. 33:191–269
21. Schlick T, Dill K. 2006. Special section on multiscale modeling in biology. Multiscale Model. Simul.
5:1174
22. Peter C, Kremer K. 2009. Multiscale simulation of soft matter systems: from the atomistic to the coarse-
grained level and back. Soft Matter 5:4357–66
23. Praprotnik M, Delle Site L, Kremer K. 2008. Multiscale simulation of soft matter: from scale bridging
to adaptive resolution. Annu. Rev. Phys. Chem. 59:545–71

568 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

24. Becker NB, Everaers R. 2007. From rigid base pairs to semiflexible polymers: coarse-graining DNA.
Phys. Rev. E 76:021923
25. Gujrati PD, Leonov AL, eds. 2010. Modeling and Simulations in Polymers. Weinheim: Wiley-VCH Verlag
26. Voth G, ed. 2009. Coarse-Graining of Condensed Phase and Biomolecular Systems. Boca Raton, FL: CRC
27. Phillips R, Kondev J, Theriot J. 2009. Physical Biology of the Cell. New York: Garland Sci.
28. Mackerell AD, Wiorkiewiczkuczera J, Karplus M. 1995. An all-atom empirical energy function for the
simulation of nucleic acids. J. Am. Chem. Soc. 117:11946–75
29. Cheatham TE, Cieplak P, Kollman PA. 1999. A modified version of the Cornell et al. force field with
improved sugar pucker phases and helical repeat. J. Biomol. Struct. Dyn. 16:845–62
30. Cheatham TE. 2004. Simulation and modeling of nucleic acid structure, dynamics and interactions.
Curr. Opin. Struct. Biol. 14:360–67
31. Fadrna E, Spackova N, Sarzynska J, Koca J, Orozco M, et al. 2009. Single stranded loops of quadruplex
DNA as key benchmark for testing nucleic acids force fields. J. Chem. Theory Comput. 5:2514–30
32. Orozco M, Noy A, Perez A. 2008. Recent advances in the study of nucleic acid flexibility by molecular
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

dynamics. Curr. Opin. Struct. Biol. 18:185–93


33. Chopra M, Larson RG. 2002. Brownian dynamics simulations of isolated polymer molecules in shear
flow near adsorbing and nonadsorbing surfaces. J. Rheol. 46:831–62
34. Jendrejack RM, Dimalanta ET, Schwartz DC, Graham MD, de Pablo JJ. 2003. DNA dynamics in a
microchannel. Phys. Rev. Lett. 91:038102
by University of Leeds on 06/23/13. For personal use only.

35. Shaqfeh ESG. 2005. The dynamics of single-molecule DNA in flow. J. Non-Newton. Fluid Mech. 130:1–28
36. Chen YL, Graham MD, de Pablo JJ, Jo K, Schwartz DC. 2005. DNA molecules in microfluidic
oscillatory flow. Macromolecules 38:6680–87
37. Chen YL, Graham MD, de Pablo JJ, Randall GC, Gupta M, Doyle PS. 2004. Conformation and dynamics
of single DNA molecules in parallel-plate slit microchannels. Phys. Rev. E 70:060901
38. Watari N, Makino M, Kikuchi N, Larson RG, Doi M. 2007. Simulation of DNA motion in a microchan-
nel using stochastic rotation dynamics. J. Chem. Phys. 126:094902
39. Hernandez-Ortiz JP, Chopra M, Geier S, Pablo JJ. 2009. Hydrodynamic effects on the translocation
rate of a polymer through a pore. J. Chem. Phys. 131:044904
40. Hernandez-Ortiz JP, de Pablo JJ, Graham MD. 2007. Fast computation of many-particle hydrodynamic
and electrostatic interactions in a confined geometry. Phys. Rev. Lett. 98:140602
41. Hernandez-Ortiz JP, Ma HB, de Pablo JJ, Graham MD. 2008. Concentration distributions during flow
of confined flowing polymer solutions at finite concentration: slit and grooved channel. Korea-Aust. Rheol.
J. 20:143–52
42. Zhang Y, de Pablo JJ, Graham MD. 2009. Multiple free energy minima in systems of confined tethered
polymers: toward soft nanomechanical bistable elements. Soft Matter 5:3694–700
43. Zhang Y, Donev A, Weisgraber T, Alder BJ, Graham MD, de Pablo JJ. 2009. Tethered DNA dynamics
in shear flow. J. Chem. Phys. 130:234902
44. Dambal A, Shaqfeh ESG. 2009. The conformational dynamics of λ-DNA in the anti-Brownian elec-
trokinetic trap: Brownian dynamics and Monte Carlo simulation. J. Chem. Phys. 131:224905
45. Lueth CA, Shaqfeh ESG. 2009. Experimental and numerical studies of tethered DNA shear dynamics
in the flow-gradient plane. Macromolecules 42:9170–82
46. Slater GW, Holm C, Chubynsky MV, de Haan HW, Dube A, et al. 2009. Modeling the separation of
macromolecules: a review of current computer simulation methods. Electrophoresis 30:792–818
47. Kenward M, Dorfman KD. 2009. Brownian dynamics simulations of single-stranded DNA hairpins. J.
Chem. Phys. 130:095101
48. Macfarlane RJ, Lee B, Hill HD, Senesi AJ, Seifert S, Mirkin CA. 2009. Assembly and organization
processes in DNA-directed colloidal crystallization. Proc. Natl. Acad. Sci. USA 106:10493–98
49. Schlick T, Perisic O. 2009. Mesoscale simulations of two nucleosome-repeat length oligonucleosomes.
Phys. Chem. Chem. Phys. 11:10729–37
50. Xu F, Olson WK. 2010. DNA architecture, deformability, and nucleosome positioning. J. Biomol. Struct.
Dyn. 27:725–39
51. Lu XJ, Olson WK. 2008. 3DNA: a versatile, integrated software system for the analysis, rebuilding and
visualization of three-dimensional nucleic-acid structures. Nat. Protoc. 3:1213–27

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 569


PC62CH26-de-Pablo ARI 25 February 2011 16:25

52. Olson WK, Zhurkin VB. 2000. Modeling DNA deformations. Curr. Opin. Struct. Biol. 10:286–97
53. Tan RKZ, Petrov AS, Harvey SC. 2006. YUP: a molecular simulation program for coarse-grained and
multiscaled models. J. Chem. Theory Comput. 2:529–40
54. Locker R, Harvey SC. 2005. A simple model for DNA packaging dynamics in bacteriophages. Biophys.
J. 88:A229–30
55. Olson WK, Bansal M, Burley SK, Dickerson RE, Gerstein M, et al. 2001. A standard reference frame
for the description of nucleic acid base-pair geometry. J. Mol. Biol. 313:229–37
56. Du Q, Smith C, Shiffeldrim N, Vologodskaia M, Vologodskii A. 2005. Cyclization of short DNA
fragments and bending fluctuations of the double helix. Proc. Natl. Acad. Sci. USA 102:5397–402
57. Vologodskii A. 2006. Brownian dynamics simulation of knot diffusion along a stretched DNA molecule.
Biophys. J. 90:1594–97
58. Burnier Y, Dorier J, Stasiak A. 2008. DNA supercoiling inhibits DNA knotting. Nucleic Acids Res. 36:4956–
63
59. Drukker K, Schatz GC. 2000. A model for simulating dynamics of DNA denaturation. J. Phys. Chem. B
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

104:6108–11
60. Bruant N, Flatters D, Lavery R, Genest D. 1999. From atomic to mesoscopic descriptions of the internal
dynamics of DNA. Biophys. J. 77:2366–76
61. Tepper H, Ayton G, Voth GA. 2003. Single-molecule dynamics of DNA molecules: a multi-scale mod-
eling approach. Biophys. J. 84:A466
by University of Leeds on 06/23/13. For personal use only.

62. Tepper HL, Voth GA. 2005. A coarse-grained model for double-helix molecules in solution: spontaneous
helix formation and equilibrium properties. J. Chem. Phys. 122:124906
63. Knotts TA, Rathore N, Schwartz DC, de Pablo JJ. 2007. A coarse grain model for DNA. J. Chem. Phys.
126:084901
64. Sambriski EJ, Ortiz V, de Pablo JJ. 2009. Sequence effects in the melting and renaturation of short DNA
oligonucleotides: structure and mechanistic pathways. J. Phys. Condens. Matter 21:034105
65. Sambriski EJ, Schwartz DC, de Pablo JJ. 2009. A mesoscale model of DNA and its renaturation. Biophys.
J. 96:1675–90
66. Lyubartsev AP, Martsinovski AA, Shevkunov SV, Vorontsovvelyaminov PN. 1992. New approach to
Monte-Carlo calculation of the free energy: method of expanded ensembles. J. Chem. Phys. 96:1776–83
67. Escobedo FA, de Pablo JJ. 1995. Monte-Carlo simulation of the chemical potential of polymers in an
expanded ensemble. J. Chem. Phys. 103:2703–10
68. de Pablo JJ, Yan QL, Escobedo FA. 1999. Simulation of phase transitions in fluids. Annu. Rev. Phys.
Chem. 50:377–411
69. Wang FG, Landau DP. 2001. Efficient, multiple-range random walk algorithm to calculate the density
of states. Phys. Rev. Lett. 86:2050–53
70. Yan QL, de Pablo JJ. 2003. Fast calculation of the density of states of a fluid by Monte Carlo simulations.
Phys. Rev. Lett. 90:035701
71. Hansmann UHE. 1997. Parallel tempering algorithm for conformational studies of biological molecules.
Chem. Phys. Lett. 281:140–50
72. Yan QL, de Pablo JJ. 1999. Hyper-parallel tempering Monte Carlo: application to the Lennard-Jones
fluid and the restricted primitive model. J. Chem. Phys. 111:9509–16
73. Yan QL, de Pablo JJ. 2000. Hyperparallel tempering Monte Carlo simulation of polymeric systems. J.
Chem. Phys. 113:1276–82
74. Faller R, Yan QL, de Pablo JJ. 2002. Multicanonical parallel tempering. J. Chem. Phys. 116:5419–23
75. Sugita Y, Okamoto Y. 1999. Replica-exchange molecular dynamics method for protein folding. Chem.
Phys. Lett. 314:141–51
76. Rathore N, Chopra M, de Pablo JJ. 2005. Optimal allocation of replicas in parallel tempering simulations.
J. Chem. Phys. 122:024111
77. Ferrenberg AM, Swendsen RH. 1988. New Monte-Carlo technique for studying phase transitions. Phys.
Rev. Lett. 61:2635–38
78. Dellago C, Bolhuis PG. 2009. Transition path sampling and other advanced simulation techniques for
rare events. In Advanced Computer Simulation Approaches for Soft Matter Sciences III, ed. C Holm, K Kremer,
pp. 167–233. Adv. Polym. Sci. 221. New York: Springer

570 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

79. Sambriski EJ, Schwartz DC, de Pablo JJ. 2009. Uncovering pathways in DNA oligonucleotide hybridiza-
tion via transition state analysis. Proc. Natl. Acad. Sci. USA 106:18125–30
80. McCullagh M, Prytkova T, Tonzani S, Winter ND, Schatz GC. 2008. Modeling self-assembly processes
driven by nonbonded interactions in soft materials. J. Phys. Chem. B 112:10388–98
81. Hoefert MJ, Sambriski EJ, de Pablo JJ. 2011. Molecular pathways in DNA-DNA hybridization of surface-
bound oligonucleotides. Soft Matter. In press; doi: 0.1039/C0SM00729C
82. Taketomi H, Ueda Y, Go N. 1975. Studies on protein folding, unfolding and fluctuations by computer
simulation 0.1. Effect of specific amino-acid sequence represented by specific inter-unit interactions. Int.
J. Pept. Protein Res. 7(6):445–59
83. Morriss-Andrews A, Rottler J, Plotkin SS. 2010. A systematically coarse-grained model for DNA and its
predictions for persistence length, stacking, twist, and chirality. J. Chem. Phys. 132:035105
84. Dans PD, Zeida A, Machado MR, Pantano S. 2010. A coarse grained model for atomic-detailed DNA
simulations with explicit electrostatics. J. Chem. Theory Comput. 6:1711–25
85. Prytkova TR, Eryazici I, Stepp B, Nguyen SB, Schatz GC. 2010. DNA melting in small-molecule-
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

DNA-hybrid dimer structures: experimental characterization and coarse-grained molecular dynamics


simulations. J. Phys. Chem. B 114:2627–34
86. Poulain P, Saladin A, Hartmann B, Prevost C. 2008. Insights on protein-DNA recognition by coarse
grain modelling. J. Comput. Chem. 29:2582–92
by University of Leeds on 06/23/13. For personal use only.

87. Grigoryev SA, Arya G, Correll S, Woodcock CL, Schlick T. 2009. Evidence for heteromorphic chro-
matin fibers from analysis of nucleosome interactions. Proc. Natl. Acad. Sci. USA 106:13317–22
88. Arya G, Schlick T. 2006. Role of histone tails in chromatin folding revealed by a mesoscopic oligonu-
cleosome model. Proc. Natl. Acad. Sci. USA 103:16236–41
89. Mackay ME, Tuteja A, Duxbury PM, Hawker CJ, Van Horn B, et al. 2006. General strategies for
nanoparticle dispersion. Science 311:1740–43
90. Crosby AJ, Lee JY. 2007. Polymer nanocomposites: the “nano” effect on mechanical properties. Polym.
Rev. 47:217–29
91. Kumar SK, Krishnamoorti R. 2010. Nanocomposites: structure, phase behavior and properties. Annu.
Rev. Chem. Biomol. Eng. 1:37–58
92. Akcora P, Liu H, Kumar SK, Moll J, Li Y, et al. 2009. Anisotropic self-assembly of spherical polymer-
grafted nanoparticles. Nat. Mater. 8:354–59
93. Hooper JB, Schweizer KS. 2005. Contact aggregation, bridging, and steric stabilization in dense polymer-
particle mixtures. Macromolecules 38:8858–69
94. Hall LM, Jayaraman A, Schweizer KS. 2010. Molecular theories of polymer nanocomposites. Curr. Opin.
Solid State Mater. Sci. 14:38–48
95. Hall LM, Schweizer KS. 2008. Many body effects on the phase separation and structure of dense polymer-
particle melts. J. Chem. Phys. 128:234901
96. Hall LM, Schweizer KS. 2010. Structure, scattering patterns and phase behavior of polymer nanocom-
posites with nonspherical fillers. Soft Matter 6:1015–25
97. Ganesan V, Ellison CJ, Pryamitsyn V. 2010. Mean-field models of structure and dispersion of polymer-
nanoparticle mixtures. Soft Matter 17: 4010–25
98. Smith JS, Borodin O, Smith GD, Kober EM. 2007. A molecular dynamics simulation and quantum
chemistry study of poly(dimethylsiloxane)-silica nanoparticle interactions. J. Polym. Sci. B 45:1599–615
99. Li CL, Choi P, Williams MC. 2009. Molecular dynamics study of the melt morphology of polyethylene
chains with different branching characteristics adjacent to a clay surface. Langmuir 26:4303–10
100. Lin PH, Khare R. 2009. Molecular simulation of cross-linked epoxy and epoxy-POSS nanocomposite.
Macromolecules 42:4319–27
101. Mo ZL, Qiao LJ, Sun YL, Li HJ. 2009. Atomistic and mesoscale interface simulation of graphite
nanosheet/AgCl/polypyrrole composite. Comput. Mater. Sci. 45:981–85
102. Zhang HP, Lu X, Fang LM, Qu SX, Feng B, Weng J. 2008. Atomic-scale interactions at the interface
of biopolymer/hydroxyapatite. Biomed. Mater. 3:044110
103. Zeng QH, Yu AB. 2008. Molecular dynamics simulations of organoclays and polymer nanocomposites.
Int. J. Nanotechnol. 5:277–90

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 571


PC62CH26-de-Pablo ARI 25 February 2011 16:25

104. Yani Y, Lamm MH. 2009. Molecular dynamics simulation of mixed matrix nanocomposites containing
polyimide and polyhedral oligomeric silsesquioxane (POSS). Polymer 50:1324–32
105. Zhang QG, Liu QL, Wu JY, Chen Y, Zhu AM. 2009. Structure-related diffusion in poly(methyl
methacrylate)/polyhedral oligomeric silsesquioxanes composites: a molecular dynamics simulation study.
J. Membr. Sci. 342:105–12
106. Sheng N, Boyce MC, Parks DM, Rutledge GC, Abes JI, Cohen RE. 2004. Multiscale micromechanical
modeling of polymer/clay nanocomposites and the effective clay particle. Polymer 45:487–506
107. Brown D, Marcadon V, Mele P, Alberola ND. 2008. Effect of filler particle size on the properties of
model nanocomposites. Macromolecules 41:1499–511
108. Ozmusul MS, Picu CR, Sternstein SS, Kumar SK. 2005. Lattice Monte Carlo simulations of chain
conformations in polymer nanocomposites. Macromolecules 38:4495–500
109. Wang Q, Nealey PF, de Pablo JJ. 2003. Behavior of single nanoparticle/homopolymer chain in ordered
structures of diblock copolymers. J. Chem. Phys. 118:11278–85
110. Starr FW, Douglas JF, Glotzer SC. 2003. Origin of particle clustering in a simulated polymer nanocom-
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

posite and its impact on rheology. J. Chem. Phys. 119:1777–88


111. Knauert ST, Douglas JF, Starr FW. 2007. The effect of nanoparticle shape on polymer-nanocomposite
rheology and tensile strength. J. Polym. Sci. B 45:1882–97
by University of Leeds on 06/23/13. For personal use only.

112. Rahedi AJ, Douglas JF, Starr FW. 2008. Model for reversible nanoparticle assembly in a polymer matrix.
J. Chem. Phys. 128:024902
113. Papakonstantopoulos GJ, Doxastakis M, Nealey PF, Barrat JL, de Pablo JJ. 2007. Calculation of local
mechanical properties of filled polymers. Phys. Rev. E 75:031803
114. Papakonstantopoulos GJ, Yoshimoto K, Doxastakis M, Nealey PF, de Pablo JJ. 2005. Local mechanical
properties of polymeric nanocomposites. Phys. Rev. E 72:031801
115. Cho J, Sun CT. 2007. A molecular dynamics simulation study of inclusion size effect on polymeric
nanocomposites. Comput. Mater. Sci. 41:54–62
116. Vacatello M. 2002. Molecular arrangements in polymer-based nanocomposites. Macromol. Theory Simul.
11:757–65
117. Merabia S, Sotta P, Long DR. 2008. A microscopic model for the reinforcement and the nonlinear
behavior of filled elastomers and thermoplastic elastomers (Payne and Mullins effects). Macromolecules
41:8252–66
118. Sternstein SS, Zhu AJ. 2002. Reinforcement mechanism of nanofilled polymer melts as elucidated by
nonlinear viscoelastic behavior. Macromolecules 35:7262–73
119. Riggleman RA, Toepperwein G, Papakonstantopoulos GJ, Barrat JL, de Pablo JJ. 2009. Entanglement
network in nanoparticle reinforced polymers. J. Chem. Phys. 130:244903
120. Karayiannis NC, Mavrantzas VG, Theodorou DN. 2002. A novel Monte Carlo scheme for the rapid
equilibration of atomistic model polymer systems of precisely defined molecular architecture. Phys. Rev.
Lett. 88:105503
121. Banaszak BJ, de Pablo JJ. 2003. A new double-rebridging technique for linear polyethylene. J. Chem.
Phys. 119:2456–62
122. Kroger M, Ramirez J, Ottinger HC. 2002. Projection from an atomistic chain contour to its primitive
path. Polymer 43:477–87
123. Everaers R, Sukumaran SK, Grest GS, Svaneborg C, Sivasubramanian A, Kremer K. 2004. Rheology
and microscopic topology of entangled polymeric liquids. Science 303:823–26
124. Tzoumanekas C, Theodorou DN. 2006. Topological analysis of linear polymer melts: a statistical ap-
proach. Macromolecules 39:4592–604
125. Tzoumanekas C, Theodorou DN. 2006. From atomistic simulations to slip-link models of entangled
polymer melts: hierarchical strategies for the prediction of rheological properties. Curr. Opin. Solid State
Mater. Sci. 10:61–72
126. Karayiannis NC, Kroger M. 2009. Combined molecular algorithms for the generation, equilibration and
topological analysis of entangled polymers: methodology and performance. Int. J. Mol. Sci. 10:5054–89

572 de Pablo
PC62CH26-de-Pablo ARI 25 February 2011 16:25

127. Kroger M. 2005. Shortest multiple disconnected path for the analysis of entanglements in two- and
three-dimensional polymeric systems. Comput. Phys. Commun. 168:209–32
128. Riggleman RA, Toepperwein GN, Papakonstantopoulos GJ, de Pablo JJ. 2009. Dynamics of a glassy
polymer nanocomposite during active deformation. Macromolecules 42:3632–40
129. Toepperwein G, Riggleman RA, Karayiannis NC, Kroger M, De Pablo JJ. 2010. Entanglement network
of polymer nanocomposite with nanorod inclusions. Manuscript submitted
130. Chiu JJ, Kim BJ, Kramer EJ, Pine DJ. 2005. Control of nanoparticle location in block copolymers. J.
Am. Chem. Soc. 127:5036–37
131. Kim BJ, Fredrickson GH, Hawker CJ, Kramer EJ. 2007. Nanoparticle surfactants as a route to bicon-
tinuous block copolymer morphologies. Langmuir 23:7804–9
132. Kang H, Detcheverry FA, Mangham AN, Stoykovich MP, Daoulas KC, et al. 2008. Hierarchical as-
sembly of nanoparticle superstructures from block copolymer-nanoparticle composites. Phys. Rev. Lett.
100:148303
133. La YH, Stoykovich MP, Park SM, Nealey PF. 2007. Directed assembly of cylinder-forming block
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

copolymers into patterned structures to fabricate arrays of spherical domains and nanoparticles. Chem.
Mater. 19:4538–44
134. Son JG, Bae WK, Kang HM, Nealey PF, Char K. 2009. Placement control of nanomaterial arrays on
the surface-reconstructed block copolymer thin films. ACS Nano 3:3927–34
by University of Leeds on 06/23/13. For personal use only.

135. Balazs AC, Emrick T, Russell TP. 2006. Nanoparticle polymer composites: where two small worlds
meet. Science 314:1107–10
136. Lin Y, Boker A, He JB, Sill K, Xiang HQ, et al. 2005. Self-directed self-assembly of nanoparti-
cle/copolymer mixtures. Nature 434:55–59
137. Huh J, Ginzburg VV, Balazs AC. 2000. Thermodynamic behavior of particle/diblock copolymer mix-
tures: simulation and theory. Macromolecules 33:8085–96
138. Thompson RB, Ginzburg VV, Matsen MW, Balazs AC. 2001. Predicting the mesophases of copolymer-
nanoparticle composites. Science 292:2469–72
139. Thompson RB, Ginzburg VV, Matsen MW, Balazs AC. 2002. Block copolymer-directed assembly of
nanoparticles: forming mesoscopically ordered hybrid materials. Macromolecules 35:1060–71
140. Sides SW, Kim BJ, Kramer EJ, Fredrickson GH. 2006. Hybrid particle-field simulations of polymer
nanocomposites. Phys. Rev. Lett. 96:250601
141. Detcheverry FA, Daoulas KC, Muller M, Nealey PF, de Pablo JJ. 2009. Monte Carlo simulations of a
coarse grain model for block copolymer systems. In Coarse Graining of Condensed Phase and Biomolecular
Systems, ed. GA Voth, pp. 361–78. Boca Raton, FL: CRC
142. Detcheverry FA, Kang HM, Daoulas KC, Muller M, Nealey PF, de Pablo JJ. 2008. Monte Carlo
simulations of a coarse grain model for block copolymers and nanocomposites. Macromolecules 41:4989–
5001
143. Gore J, Bryant Z, Nollmann M, Le MU, Cozzarelli NR, Bustamante C. 2006. DNA overwinds when
stretched. Nature 442:836–39
144. Jo K, Dhingra DM, Odijk T, de Pablo JJ, Graham MD, et al. 2007. A single-molecule barcoding system
using nanoslits for DNA analysis. Proc. Natl. Acad. Sci. USA 104:2673–78
145. Teclemariam NP, Beck VA, Shaqfeh ESG, Muller SJ. 2007. Dynamics of DNA polymers in post arrays:
comparison of single molecule experiments and simulations. Macromolecules 40:3848–59
146. Fang L, Larson RG. 2007. Concentration dependence of shear-induced polymer migration in DNA
solutions near a surface. Macromolecules 40:8784–87
147. Sukumaran SK, Likhtman AE. 2009. Modeling entangled dynamics: comparison between stochastic
single-chain and multichain models. Macromolecules 42:4300–9
148. Likhtman AE. 2005. Single-chain slip-link model of entangled polymers: simultaneous description of
neutron spin-echo, rheology, and diffusion. Macromolecules 38:6128–39
149. Khaliullin RN, Schieber JD. 2009. Self-consistent modeling of constraint release in a single-chain mean-
field slip-link model. Macromolecules 42:7504–17
150. Schieber JD, Horio K. 2010. Fluctuation in entanglement positions via elastic slip-links. J. Chem. Phys.
132:074905

www.annualreviews.org • Coarse-Grained Simulations of Macromolecules 573


PC62CH26-de-Pablo ARI 25 February 2011 16:25

151. Muller M, Daoulas KC. 2008. Single-chain dynamics in a homogeneous melt and a lamellar microphase:
a comparison between Smart Monte Carlo dynamics, slithering-snake dynamics, and slip-link dynamics.
J. Chem. Phys. 129:164906
152. Krasilnikov AS, Xiao YH, Pan T, Mondragon A. 2004. Basis for structural diversity in homologous
RNAs. Science 306:104–7
153. Uhlherr A, Theodorou DN. 2006. Accelerating molecular simulations by reversible mapping between
local minima. J. Chem. Phys. 125:084107
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org
by University of Leeds on 06/23/13. For personal use only.

574 de Pablo
PC62FrontMatter ARI 18 February 2011 17:56

Annual Review of
Physical Chemistry

Volume 62, 2011


Contents

Laboring in the Vineyard of Physical Chemistry


Benjamin Widom p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

The Ultrafast Pathway of Photon-Induced Electrocyclic


Ring-Opening Reactions: The Case of 1,3-Cyclohexadiene
Sanghamitra Deb and Peter M. Weber p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p19
by University of Leeds on 06/23/13. For personal use only.

Coarse-Grained (Multiscale) Simulations in Studies of Biophysical


and Chemical Systems
Shina C.L. Kamerlin, Spyridon Vicatos, Anatoly Dryga, and Arieh Warshel p p p p p p p p p p p p41
Dynamics of Nanoconfined Supercooled Liquids
R. Richert p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p65
Ionic Liquids: Structure and Photochemical Reactions
Edward W. Castner Jr., Claudio J. Margulis, Mark Maroncelli,
and James F. Wishart p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p85
Theoretical Study of Negative Molecular Ions
Jack Simons p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 107
Theoretical and Computational Protein Design
Ilan Samish, Christopher M. MacDermaid, Jose Manuel Perez-Aguilar,
and Jeffery G. Saven p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 129
Melting and Freezing of Metal Clusters
Andrés Aguado and Martin F. Jarrold p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 151
Astronomical Chemistry
William Klemperer p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 173
Simulating Chemistry Using Quantum Computers
Ivan Kassal, James D. Whitfield, Alejandro Perdomo-Ortiz, Man-Hong Yung,
and Alán Aspuru-Guzik p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 185
Multiresonant Coherent Multidimensional Spectroscopy
John C. Wright p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 209
Probing Free-Energy Surfaces with Differential Scanning Calorimetry
Jose M. Sanchez-Ruiz p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 231

viii
PC62FrontMatter ARI 18 February 2011 17:56

Role of Solvation Effects in Protein Denaturation: From


Thermodynamics to Single Molecules and Back
Jeremy L. England and Gilad Haran p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 257
Solid-State NMR Studies of Amyloid Fibril Structure
Robert Tycko p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 279
Cooperativity, Local-Nonlocal Coupling, and Nonnative Interactions:
Principles of Protein Folding from Coarse-Grained Models
Hue Sun Chan, Zhuqing Zhang, Stefan Wallin, and Zhirong Liu p p p p p p p p p p p p p p p p p p p p p 301
Hydrated Acid Clusters
Kenneth R. Leopold p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 327
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

Developments in Laboratory Studies of Gas-Phase Reactions for


Atmospheric Chemistry with Applications to Isoprene Oxidation
and Carbonyl Chemistry
Paul W. Seakins and Mark A. Blitz p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 351
by University of Leeds on 06/23/13. For personal use only.

Bonding in Beryllium Clusters


Michael C. Heaven, Jeremy M. Merritt, and Vladimir E. Bondybey p p p p p p p p p p p p p p p p p p p 375
Reorientation and Allied Dynamics in Water and Aqueous Solutions
Damien Laage, Guillaume Stirnemann, Fabio Sterpone, Rossend Rey,
and James T. Hynes p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 395
Detecting Nanodomains in Living Cell Membrane by Fluorescence
Correlation Spectroscopy
Hai-Tao He and Didier Marguet p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 417
Toward a Molecular Theory of Early and Late Events in Monomer
to Amyloid Fibril Formation
John E. Straub and D. Thirumalai p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 437
The Density Matrix Renormalization Group in Quantum Chemistry
Garnet Kin-Lic Chan and Sandeep Sharma p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 465
Thermodynamics and Mechanics of Membrane Curvature Generation
and Sensing by Proteins and Lipids
Tobias Baumgart, Benjamin R. Capraro, Chen Zhu, and Sovan L. Das p p p p p p p p p p p p p p p 483
Coherent Nonlinear Optical Imaging: Beyond Fluorescence
Microscopy
Wei Min, Christian W. Freudiger, Sijia Lu, and X. Sunney Xie p p p p p p p p p p p p p p p p p p p p p p p 507
Roaming Radicals
Joel M. Bowman and Benjamin C. Shepler p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 531
Coarse-Grained Simulations of Macromolecules:
From DNA to Nanocomposites
Juan J. de Pablo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 555

Contents ix
PC62FrontMatter ARI 18 February 2011 17:56

New Developments in the Physical Chemistry of Shock Compression


Dana D. Dlott p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 575
Solvation Dynamics and Proton Transfer in Nanoconfined Liquids
Ward H. Thompson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 599
Nonadiabatic Events and Conical Intersections
Spiridoula Matsika and Pascal Krause p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 621
Lessons in Fluctuation Correlation Spectroscopy
Michelle A. Digman and Enrico Gratton p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 645

Indexes
Annu. Rev. Phys. Chem. 2011.62:555-574. Downloaded from www.annualreviews.org

Cumulative Index of Contributing Authors, Volumes 58–62 p p p p p p p p p p p p p p p p p p p p p p p p p p p 669


Cumulative Index of Chapter Titles, Volumes 58–62 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 672
by University of Leeds on 06/23/13. For personal use only.

Errata

An online log of corrections to Annual Review of Physical Chemistry articles may be


found at http://physchem.annualreviews.org/errata.shtml

x Contents

You might also like