ArticlePDF Available

A Martini 3 coarse-grain model for the simulation of the photopolymerizable organic phase in dental composites

Royal Society of Chemistry
RSC Advances
Authors:

Abstract and Figures

Light-hardening dental composites can be used in a large number of applications in restorative dentistry. They are based on photopolymerizable resins, which are highly relevant also in other industries like 3D printing. Much effort is therefore being put into developing and optimizing photopolymers. Currently used photopolymers still have limitations regarding mechanical properties, shrinkage and leaching of uncured monomers. These issues are strongly linked to the network structure of the polymer and are usually addressed using trial and error methods. Therefore, it is of interest to have a model for the network structure of such materials and to have a tool to facilitate scientific progress and the development of high-performance photopolymers. This work presents a coarse grain model of Bis-GMA/TEGDMA formulations and their corresponding networks, following the Martini 3 guidelines and using a simulated polymerization algorithm. The model proved to reproduce the densities and volumetric shrinkage values found in the literature well. Furthermore, it was possible to estimate the final double bond conversion of the polymer material. Martini's building block-like design makes it easy to extend the model to other monomers in the future.
This content is subject to copyright. Terms and conditions apply.
A Martini 3 coarse-grain model for the simulation of
the photopolymerizable organic phase in dental
composites
Alexander Hochwallner *and J¨
urgen Stamp
Light-hardening dental composites can be used in a large number of applications in restorative dentistry.
They are based on photopolymerizable resins, which are highly relevant also in other industries like 3D
printing. Much eort is therefore being put into developing and optimizing photopolymers. Currently
used photopolymers still have limitations regarding mechanical properties, shrinkage and leaching of
uncured monomers. These issues are strongly linked to the network structure of the polymer and are
usually addressed using trial and error methods. Therefore, it is of interest to have a model for the
network structure of such materials and to have a tool to facilitate scientic progress and the
development of high-performance photopolymers. This work presents a coarse grain model of Bis-
GMA/TEGDMA formulations and their corresponding networks, following the Martini 3 guidelines and
using a simulated polymerization algorithm. The model proved to reproduce the densities and
volumetric shrinkage values found in the literature well. Furthermore, it was possible to estimate the nal
double bond conversion of the polymer material. Martini's building block-like design makes it easy to
extend the model to other monomers in the future.
Introduction
Light-hardening composites as restorative materials in
dentistry are very popular. They are used to replace amalgam as
the default choice for dentists, and patients.
1
Dental composites
consist of an organic matrix, which solidies upon photo-
chemical initiation, and inorganic llers.
2
The organic phase
most commonly contains Bis-GMA diluted by TEGDMA.
3
Both
are bifunctional methacrylate substances. The molecular
structures are shown in Scheme 1.
For a composite to be used in restorative dentistry, it must
full several challenging requirements. The mechanical prop-
erties but also chemical stability are of high importance. Addi-
tionally, polymerization shrinkage poses a limitation to the
applicability of light-cured polymer networks.
4,5
In order to establish safe and reliable curing schemes for the
application of dental compounds, the properties of the organic
phase need to be investigated thoroughly regarding the
composition of resins and the properties in relation to the
polymerization reaction.
Photopolymer resins are not only being used in dental
applications. They are also used extensively in additive
manufacturing, which has the potential to become a key
technology in the future, provided the material properties
meet the requirements.
6
Much eort is therefore being put
into synthesizing new 3D printing materials for a variety of
applications. Polymerization shrinkage is also an issue in
additive manufacturing as it can lead to warpage and residual
stresses.
7
Both of which compromise the usability of 3D
printed parts.
Developing and optimizing photopolymer resins is a time-
consuming task, which includes mixing of the components,
photochemical curing, testing and characterization. In order to
optimize a resin composition, the outlined sequence of steps
has to be repeated with dierent ratios of the ingredients and
dierent curing schemes. This leads to a high number of
samples to be analyzed. Hence, it is of interest to establish
a method that can screen a large part of the compound space
automatically or at least semi-automatically.
Scheme 1
Institute of Materials Science and Technology, TU Wien, Getreidemarkt 9, 1060
Vienna, Austria. E-mail: alexander.hochwallner@tuwien.ac.at
Cite this: RSC Adv., 2022, 12,12053
Received 3rd February 2022
Accepted 14th April 2022
DOI: 10.1039/d2ra00732k
rsc.li/rsc-advances
© 2022 The Author(s). Published by the Royal Society of Chemistry RSC Adv., 2022, 12,1205312059 | 12053
RSC Advances
PAPER
Another limitation to the purely experimental approach is
that no method is available to get direct information on the
network structure. It is evident that the constitution of the cured
polymer network is integral to the properties of the nal
material. Therefore, establishing structureproperty relations is
of high relevance. Additional information on the network
structure may make it easier to relate material properties to the
molecular network structure obtained with dierent curing
conditions, giving way to optimizing dental composites and 3D
printing resins. This work aims to explore the possible usage of
molecular simulations for the investigation of photopolymer
resins. The technologically highly relevant mixture of Bis-GMA
with TEGDMA was chosen as a test system. The transferability
of the modeling approach to molecules relevant in additive
manufacturing is a key motivation for this work.
Molecular dynamics simulation is a method to investigate
a material on the molecular level with possibilities to automate
the tasks.
8
To get accurate results from molecular dynamics
simulations, the force eld must be adequate, and system sizes
and time scales must be large enough.
9
To satisfy these condi-
tions, coarse grain force elds for molecular systems are under
constant development. A coarse grain model maps groups of
individual atoms to one coarse grain bead. The underlying
chemical group then denes the inter-particle interactions as
eective interactions.
10
The reduction of the degrees of freedom
by grouping atoms to beads leads to computationally less
expensive simulations. This enables the simulations of larger
systems over a longer period of time compared to all-atom
simulations.
One widely used coarse grain potential that provides rec-
ommended mapping schemes and interparticle potentials is
Martini.
11
The strength of the Martini force eld is, on the one
hand, its easy applicability, and on the other hand, its veried
accuracy in numerous chemical problems.
12
The Martini beads
are designed to be used as building blocks and can be applied to
a variety of dierent molecules.
12,13
This building block
approach makes it perfectly suitable to eciently explore the
space of photopolymer materials.
Compared to previous Martini models, Martini 3 oers more
bead types, and the mapping scheme was changed from centre
of mass mapping to centre of geometry mapping. This change
leads to better reproduction of packing densities.
14
This work presents a molecular dynamics model following
the Martini 3 mapping scheme to simulate the polymer network
of the organic phase in dental composites. Next to proper
representations of the monomers, this requires a close to reality
network structure. To achieve this, a simulated polymerization
algorithm called Polymatic
15
is applied. It is used to establish
chemical bonds between beads that represent the reactive
groups of the monomers. This is done by employing a cut-o
criterion, which means that a bond is formed if a bead
labelled as active and a bead labelled as reactive are within
a certain distance to each other. The active label is then passed
on to the reaction partner to allow for chain-growth
polymerization.
The model is veried by comparing monomer densities and
curing shrinkage to values reported in literature. A comparison
to experimental double bond conversions is also presented to
give an outlook on the predictive value of a reactive model to
study photopolymer resins. The present model should serve as
a basis for future extensions with arbitrary monomers of
interest.
Computational details
For all-atom simulations, GROMACS
16
was used. The run
parameters were obtained from the Martini web page
17
and are
in accordance with ref. 13. Only the constraints setting was
changed from all-bonds to h-bonds. All-atom OPLS force eld
les were generated using LigParGen
1820
following ref. 17. The
simulation box was prepared to contain a single molecule of
interest, diluted in 500 molecules of TEGDMA. Energy mini-
mization, equilibration, and production run were done as
described in ref. 17. The mapping les necessary to map the
trajectories from the all-atom representation to the coarse-grain
representation were generated using CGbuilder.
17,21
The coarse
grain runs needed to obtain the coarse grain distributions were
also done with GROMACS
16
and run les taken from ref. 17. The
time step was set to 20 fs. The remaining parameters were used
as provided. All distributions were evaluated using the les
provided in ref. 17.
The simulated polymerization was done using Polymatic
15
which sequentially checks the dened bonding criterion,
updates the topology if a bond was formed, and calls LAMMPS
22
for energy minimization and molecular dynamics run. The
simulation box for the polymerization simulation was prepared
with the packing algorithm that is delivered with Polymatic.
15
First, the active monomers that acted as initiators were
randomly added to the box. Then Bis-GMA molecules were
added. The remaining TEGDMA molecules were added last.
Then energy minimization was done, and the system was
equilibrated at 298 K and hydrostatic pressure of 1 bar for
80 000 timesteps with 20 fs. Polymatic
15
was set to call
LAMMPS
22
for energy minimization, and molecular dynamics
run for 40 000 timesteps aer ten bonds were successfully
formed. A 20 000 timesteps run is performed if no bond was
formed, and the bond formation is attempted again. The cut-o
for bond formation was set to 5.5 ˚
A. The simulation was
terminated if no bond was formed aer 40 attempts. During all
runs, a temperature of 298 K and a pressure of 1 bar were
maintained. The Shake algorithm was used because LAMMPS
does not have the Lincs algorithm implemented. For that, one
angle and two bonds are constrained in every aromatic ring in
the system. The number of molecules in each formulation was
(0;559), (200;447), (300;391), (400;335), (500;279), (600;223),
(700;168), (800;112), (1000;0). The number before the semicolon
represents TEGDMA. Temperature sweeps were performed over
one million timesteps from 150 K to 500 K.
Coarse graining
Bead assignment
The mapping of the atoms into coarse grain beads was done as
indicated in Fig. 1. Care was taken to have two, three, or four
12054 |RSC Adv., 2022, 12,1205312059 © 2022 The Author(s). Published by the Royal Society of Chemistry
RSC Advances Paper
heavy atoms grouped into tiny, small, and regular beads,
respectively. The bead type of bead number 4 was chosen to be
TN2a in accordance with methoxybenzene as proposed in ref.
23. The remaining bead types were selected to be in accordance
with the default choice of bead types as presented in ref. 14 and
24. The bead choices are summarized in Table 1.
Bonded interactions
The coecients for the bonded interactions were set to best
match the distributions of the all-atom simulation. The values
for the harmonic bond coecients are given in Table 2. The
harmonic bond potential has the following form:
V¼1
2kðrr0Þ2
where kis the force constant, ris the distance between bonded
beads and r
0
is the equilibrium distance. Note that the bonds
between the beads forming the aromatic structure are modelled
using constraints as the bond length distribution is narrow.
23
Two dierent bond lengths were used between type 5 beads. A
longer bond length was used between type 5 beads bonded to
bead type 4. Table 3 shows the coecients for the harmonic
angles. To reproduce the shape of the all-atom simulation,
a relatively stiangle was necessary in Bis-GMA for the angle
with a type 6 bead as the central bead. The harmonic angle
potential looks as follows:
V¼1
2kðqq0Þ2
where qis the three-body angle between bonded beads and q
0
is
the equilibrium three-body angle.
The values for the dihedral coecients used are presented in
Table 4. Special care was taken to reproduce the angular
arrangement of the aromatic rings with respect to each other
(see Fig. 2). The all-atom simulation showed that the distribu-
tion of the dihedrals that describe the aromatic ring's rotation
are bimodal and not independent of each other. One of the
rings in a Bis-GMA molecule resides in one of two possible
Fig. 1 Mapping scheme. Choices for Martini 3 bead types are given in
Table 1.
Table 1 Choices of standard Martini 3 bead types for mapping Bis-
GMA and TEGDMA
Bead number Martini 3 bead type
1 SC4
2 N4a
3 TP1
4 TN2a
5 TC5
6 SC2
7 SN3a
Table 2 Harmonic bond coecients
Beadbead Distance (nm)
Force constant
(kJ mol
1
nm
2
)
12 0.355 35 000
23 0.23 5000
34 0.24 28 000
45 0.3 15 000
55*
constrained
0.229
55*
constrained
0.197
56 0.271 35 000
27 0.345 6000
77 0.37 9000 Fig. 2 Dihedral rotations of the aromatic rings.
Table 3 Harmonic angle coecients
Beadbeadbead Angle (deg)
Force constant
(kJ mol
1
rad
2
)
123 108 110
234 76 120
345 130 40
455 120 80
556 140 130
565 70 700
127 115 100
277 124 100
Table 4 Dihedral coecients
Angle (deg)
Energy
(kJ mol
1
) Multiplicity
1234 0 35 1
5578 313 8.46 2
12770 2 3
27720 2 3
© 2022 The Author(s). Published by the Royal Society of Chemistry RSC Adv., 2022, 12,1205312059 | 12055
Paper RSC Advances
angular arrangements, while the other one is present in both
arrangements with equal probability. This behavior could be
reproduced in the coarse grain representation using two dihe-
dral angles per aromatic ring and carefully adjusting the force
constant. A too high force constant would entrap leand right
aromatic dihedrals into one of the possible congurations. A
force constant that is too low would result in equal probability
of the congurations. Fig. 3 shows the dihedral distributions of
the central dihedrals in a Bis-GMA molecule with equal proba-
bility of both congurations. Fig. 4 shows the dihedrals at the
second aromatic ring. Here one can see that only one congu-
ration is occupied. In the case of the all-atom simulation, the
distributions are located around a single angular value. In
comparison, the coarse grain simulation also shows some
probability to be present in the other conguration. However,
the maximum is clearly located around a single angular value
which is a good approximation of the all-atom simulation.
In order to keep the aromatic rings planar, improper dihe-
drals were used. The values for the dihedral coecients are
given in Table 5. The following functional form was used for the
dihedral potential and improper potential:
V¼k(1 + cos(nff
0
))
where k, in this case, is the energy constant, fis the dihedral or
improper angle and nis the multiplicity.
Simulated polymerization
In order to obtain a realistic network structure of the polymer-
ized material, a simulated polymerization approach was
chosen. To do this, Polymatic
15
was used. This code uses
a simple cut-ocriterion to decide whether or not a bond will be
established between two beads. All beads in a simulation rep-
resenting a reactive group are labelled as such. The beads that
currently bear the radical character are labelled as active. If two
beads labelled as active and reactive are within the cut-o
distance to each other, a bond is established between them.
The active label is then passed on to the reaction partner for the
chain polymerization to proceed. Aer a successful bond
formation, the backbone potential was updated. To evaluate the
additional interactions necessary for the polymer, a TEGDMA 4-
mer was parametrized using the same method as described
above. For the sake of numerical stability, no backbone dihe-
drals were set. The backbone angle was set to 350 kJ mol
1
rad
2
at an equilibrium angle of 125 deg. The backbone-
sidechain angle was set to 50 kJ mol
1
rad
2
at an equilib-
rium angle of 80 deg, and the backbone bond was set to
35 000 kJ mol
1
nm
2
and bond length of 0.31 nm.
Results and discussion
Simulated density
The simulated densities for the respective compounds are pre-
sented in Fig. 5. The values for the mixtures with little Bis-GMA
Fig. 3 Dihedral angles in Bis-GMA, which show two peaks in the
distribution.
Fig. 4 Dihedral angles in Bis-GMA, which show one peak in the
distribution.
Table 5 Improper dihedral coecients
Angle (deg)
Energy (kJ
mol
1
) Multiplicity
4555 0 65 1
5556 180 100 1
Fig. 5 Densities as measured by Dewaele et al.
25
compared to the
presented model over the Bis-GMA content.
12056 |RSC Adv., 2022, 12,1205312059 © 2022 The Author(s). Published by the Royal Society of Chemistry
RSC Advances Paper
content are in good agreement with the values reported by
Dewaele et al.
25
. With higher Bis-GMA content, the deviation
increases. A possible explanation for this is that Bis-GMA was
parametrized with respect to a system of a single Bis-GMA
molecule in a box of TEGDMA as a solvent. As a result, the
Bis-GMA Bis-GMA interactions are not accurately captured,
whereas Bis-GMA TEGDMA and TEGDMA TEGDMA inter-
actions are properly modelled. Another possible source of
deviation is the choice of MARTINI 3 bead types. Denition of
custom bead types backed by free energy calculations could
potentially improve the overall model.
26
For the sake of simple
addition of other molecules later on, only standard beads were
used in this work. Note that for calculating the model densities,
the actual molecular weights of Bis-GMA and TEGDMA were
used, not the standard MARTINI masses. For all simulation
runs, standard MARTINI masses were used.
Shrinkage of simulated material
The shrinkage was determined as the volumetric decrease of the
simulation box aer the degree of conversion has reached the
values reported by Dewaele et al.
25
The values are presented in
Fig. 6 and are in excellent agreement with the experimental
values. The model is within the reported standard deviation
except for the compound with 20 wt% Bis-GMA and 50 wt% Bis-
GMA content. The deviation in densities of the high Bis-GMA
compounds does not seem to inuence the precision of the
shrinkage values negatively.
Double bond conversion
The simulated polymerization method oers the possibility of
learning about the investigated material's reaction kinetics
prior to material synthesis. This can be a valuable information
in resin development because the degree of polymerization
greatly inuences the material properties and possibly limits
the applicability of a photopolymer. It was investigated if the
presented simulated polymerization procedure allows for an
estimation of the nal double bond conversion of various Bis-
GMA TEGDMA compositions. The results of the simulated
polymerization using 20 active sites are shown in Fig. 7,
together with the experimental values reported by Dewaele
et al.
25
It can be seen that the simulation yields too high values
for the case of 20 active propagation sites. Only at a Bis-GMA
content of 70 wt% is the agreement reasonable. The agree-
ment for pure Bis-GMA is not good. Generally, the simulated
double bond conversion values are expected to inherit more
uncertainties, coming from the fact that the condition for
polymerization termination was chosen more or less arbitrarily.
However, the trend of the values from the experiment is re-
ected in the simulation, showing that the model can reason-
ably accurately simulate the mobility of the reaction partners
during curing of the resin.
This series of simulations were repeated using only ten active
sites. Fewer active sites yield a lower probability of nding
a reaction partner, especially when the mobility is limited by the
onset of gelation in a later stage of the curing process. The
Fig. 6 Polymerization shrinkage as measured by Dewaele et al.
25
compared to the simulated values over the Bis-GMA content.
Fig. 7 Double bond conversion as measured by Dewaele et al.
25
compared to the double bond conversion of the model with 20 active
sites after the termination criterion was reached over the Bis-GMA
content.
Fig. 8 Double bond conversion as measured by Dewaele et al.
25
compared to the double bond conversion of the model with ten active
sites after the termination criterion was reached over the Bis-GMA
content.
© 2022 The Author(s). Published by the Royal Society of Chemistry RSC Adv., 2022, 12,1205312059 | 12057
Paper RSC Advances
results are presented in Fig. 8. It can be seen that the double bond
conversion is generally lower than in the case of 20 active sites
and is in excellent agreement with the experimental values. Only
for 60 wt% and 70 wt% Bis-GMA content, the values are too high.
Note that the presented method does not account for any
termination reactions that might lower the degree of polymer-
ization. This model only takes propagation into account.
Termination only happens when the mobility is reduced so
much that no reaction occurs in the specied amount of time.
From the excellent agreement of the simulation with ten active
sites, it can be concluded that this is the number of eective
active sites in the system. To relate the eectivenumber of
sites to the concentration of initiator in the real system,
a precise knowledge of the kinetics of the termination reactions
would be necessary.
To better judge the mobility change induced by the cross-
linking, temperature sweeps were performed on the 50 wt% Bis-
GMA resin and its crosslinked version to observe changes in the
density slopes. This is a common method to estimate the glass
transition temperature
27
and was also previously performed
using the Martini force eld.
28
The results are shown in Fig. 9.
For estimating the glass transition temperature, the data was
piecewise tted using linear functions. The temperature at the
intersection is 310 K for the polymer and 232 K for the resin.
This indicates that during the polymerization simulation the
resin transitions to a glassy polymer as the simulation
temperature is below the estimated glass transition for the
polymerized material.
Conclusions
The goal of this work was to explore the use of the Martini
coarse grain model for simulating photopolymerizable resins,
which are used in multiple dierent industries. One major eld
of application is restorative dentistry. Thus, Bis-GMA
TEGDMA mixtures were chosen as a logical starting point. Both
molecules were parametrized as outlined on the Martini web
page.
17
The bonded interactions were parametrized to t the
distributions of the all-atom simulations best. In Bis-GMA, the
asymmetric dihedral angle conguration observed in the all-
atom simulation was also reproduced in the coarse grain
model. This was achieved while keeping the symmetry of all
bonded interaction parameters.
Atom to bead mapping was done as recommended by the
Martini authors.
14,23,24
The obtained model reproduces resin
densities reported in literature well. The use of standard
Martini building blocks makes it easy to add additional mole-
cules of interest later on and expand the set of possible resin
formulations in the future.
Critical for the use of a photopolymer is the polymerization
shrinkage and double bond conversion of the polymerized
material. This is the case in restorative dentistry as well as in
additive manufacturing. Our simulations show that the pre-
sented model can successfully predict polymerization shrinkage
of various molecular mixtures. This information is valuable for
the development of dental composites to assess the risk of
debonding of lling and tooth
4,5
or for 3D printing materials to
avoid warpage or shrinkage stresses in additively manufactured
parts.
7
The use of a simulated polymerization also allows for an
estimation of double bond conversions for dierent resin
mixtures and dierent initiator concentrations. The presented
method well reproduces double bond conversions for Bis-GMA
TEGDMA mixtures reported by Dewaele et al.
25
Author contributions
Alexander Hochwallner: conceptualization; methodology; so-
ware; formal analysis; investigation; writing original dra;
writing review and editing; visualization. J¨
urgen Stamp:
conceptualization; writing review and editing; supervision.
Conicts of interest
There are no conicts to declare.
Acknowledgements
Funding by the Christian Doppler Research Association within
the framework of a Christian Doppler Laboratory for Advanced
Polymers for Biomaterials and 3D Printingand the nancial
support by the Austrian Federal Ministry for Digital and
Economic Aairs and the National foundation for Research,
Technology and Development are gratefully acknowledged.
Notes and references
1 M. Goldberg, In Vitro and in Vivo Studies on the Toxicity of
Dental Resin Components: A Review, Clin. Oral Investig.,
2008, 12(1), 18, DOI: 10.1007/s00784-007-0162-8.
2 A. Peutzfeldt, Resin Composites in Dentistry: The Monomer
Systems, Eur. J. Oral Sci., 1997, 105(2), 97116, DOI: 10.1111/
j.1600-0722.1997.tb00188.x.
3 J. L. Ferracane, Resin CompositeState of the Art, Dent.
Mater., 2011, 27(1), 2938, DOI: 10.1016/
j.dental.2010.10.020.
Fig. 9 Temperature sweep results of 50 wt% Bis-GMA resin and
polymer. The density is shown over the temperature together with
linear t functions.
12058 |RSC Adv., 2022, 12,1205312059 © 2022 The Author(s). Published by the Royal Society of Chemistry
RSC Advances Paper
4 R. Labella, P. Lambrechts, B. Van Meerbeek and G. Vanherle,
Polymerization Shrinkage and Elasticity of Flowable
Composites and Filled Adhesives, Dent. Mater., 1999, 15(2),
128137, DOI: 10.1016/S0109-5641(99)00022-6.
5I. A. Mj
¨
or, Minimum Requirements for New Dental
Materials, J. Oral Rehabil., 2007, 34(12), 907912, DOI:
10.1111/j.1365-2842.2007.01726.x.
6 S. C. Ligon, R. Liska, J. Stamp, M. Gurr and R. M¨
ulhaupt,
Polymers for 3D Printing and Customized Additive
Manufacturing, Chem. Rev., 2017, 117(15), 1021210290,
DOI: 10.1021/acs.chemrev.7b00074.
7 P. S. Bychkov, V. M. Kozintsev, A. V. Manzhirov and
A. L. Popov, Determination of Residual Stresses in
Products in Additive Production by the Layer-by-Layer
Photopolymerization Method, Mech. Solids, 2017, 52(5),
524529, DOI: 10.3103/S0025654417050077.
8 M. A. F. Afzal, A. R. Browning, A. Goldberg, M. D. Halls,
J. L. Gavartin, T. Morisato, T. F. Hughes, D. J. Giesen and
J. E. Goose, High-Throughput Molecular Dynamics
Simulations and Validation of Thermophysical Properties
of Polymers for Various Applications, ACS Appl. Polym.
Mater., 2021, 3(2), 620630, DOI: 10.1021/acsapm.0c00524.
9 T. E. Gartner and A. Jayaraman, Modeling and Simulations of
Polymers: A Roadmap, Macromolecules, 2019, 52(3), 755786,
DOI: 10.1021/acs.macromol.8b01836.
10 F. M¨
uller-Plathe, Coarse-Graining in Polymer Simulation:
From the Atomistic to the Mesoscopic Scale and Back,
ChemPhysChem, 2002, 3, 9, 754769, DOI: 10.1002/1439-
7641(20020916)3:9<754::AID-CPHC754>3.0.CO;2-U.
11 S. J. Marrink, H. J. Risselada, S. Yemov, D. P. Tieleman and
A. H. de Vries, The MARTINI Force Field: Coarse Grained
Model for Biomolecular Simulations, J. Phys. Chem. B,
2007, 111(27), 78127824, DOI: 10.1021/jp071097f.
12 S. J. Marrink and D. P. Tieleman, Perspective on the Martini
Model, Chem. Soc. Rev., 2013, 42(16), 68016822, DOI:
10.1039/C3CS60093A.
13 R. Alessandri, P. C. T. Souza, S. Thallmair, M. N. Melo,
A. H. de Vries and S. J. Marrink, Pitfalls of the Martini
Model, J. Chem. Theory Comput., 2019, 15(10), 54485460,
DOI: 10.1021/acs.jctc.9b00473.
14 P. C. T. Souza, R. Alessandri, J. Barnoud, S. Thallmair,
I. Faustino, F. Gr¨
unewald, I. Patmanidis, H. Abdizadeh,
B. M. H. Bruininks, T. A. Wassenaar, P. C. Kroon, J. Melcr,
V. Nieto, V. Corradi, H. M. Khan, J. Doma´
nski,
M. Javanainen, H. Martinez-Seara, N. Reuter, R. B. Best,
I. Vattulainen, L. Monticelli, X. Periole, D. P. Tieleman,
A. H. de Vries and S. J. Marrink, Martini 3: A General
Purpose Force Field for Coarse-Grained Molecular
Dynamics, Nat. Methods, 2021, 18(4), 382388, DOI:
10.1038/s41592-021-01098-3.
15 L. J. Abbott, K. E. Hart and C. M. Colina, Polymatic: A
Generalized Simulated Polymerization Algorithm for
Amorphous Polymers, Theor. Chem. Acc., 2013, 132(3),
1334, DOI: 10.1007/s00214-013-1334-z.
16 M. J. Abraham, T. Murtola, R. Schulz, S. P´
all, J. C. Smith,
B. Hess and E. Lindahl, GROMACS: High Performance
Molecular Simulations through Multi-Level Parallelism
from Laptops to Supercomputers, SowareX, 2015, 12,19
25, DOI: 10.1016/j.sox.2015.06.001.
17 Parameterizing a new small molecule,https://cgmartini.nl/
index.php/martini-3-tutorials/parameterizing-a-new-small-
molecule, accessed 2021-12-22.
18 W. L. Jorgensen, J. Tirado-Rives and B. J. Berne, Potential
Energy Functions for Atomic-Level Simulations of Water
and Organic and Biomolecular Systems, Proc. Natl. Acad.
Sci. U. S. A., 2005, 102(19), 66656670.
19 L. S. Dodda, J. Z. Vilseck, J. Tirado-Rives and W. L. Jorgensen,
1.14*CM1A-LBCC: Localized Bond-Charge Corrected CM1A
Charges for Condensed-Phase Simulations, J. Phys. Chem.
B, 2017, 121(15), 38643870, DOI: 10.1021/acs.jpcb.7b00272.
20L.S.Dodda,I.CabezadeVaca,J.Tirado-Rivesand
W. L. Jorgensen, LigParGen Web Server: An Automatic OPLS-
AA Parameter Generator for Organic Ligands, Nucleic Acids
Res., 2017, 45(W1), W331W336, DOI: 10.1093/nar/gkx312.
21 CG Builder,https://jbarnoud.github.io/cgbuilder/, accessed
2021-12-22.
22 A. P. Thompson, H. M. Aktulga, R. Berger,
D. S. Bolintineanu, W. M. Brown, P. S. Crozier, P. J. in t
Veld, A. Kohlmeyer, S. G. Moore, T. D. Nguyen, R. Shan,
M. J. Stevens, J. Tranchida, C. Trott and S. J. Plimpton,
LAMMPS - a Flexible Simulation Tool for Particle-Based
Materials Modeling at the Atomic, Meso, and Continuum
Scales, Comput. Phys. Commun., 2022, 271, 108171, DOI:
10.1016/j.cpc.2021.108171.
23 R. Alessandri, J. Barnoud, A. S. Gertsen, I. Patmanidis,
A. H. de Vries, P. C. T. Souza and S. J. Marrink Martini 3
Coarse-Grained Force Field: Small Molecules. 2021, DOI:
10.33774/chemrxiv-2021-1qmq9.
24 P. C. T. Souza, R. Alessandri, J. Barnoud, S. Thallmair,
I. Faustino, F. Grunewald, I. Patmanidis, H. Abdizadeh,
T. A. Wassenaar, P. C. Kroon, J. Melcr, V. Nieto, V. Corradi,
H. M. Khan, J. Doma´
nski, M. Javanainen, N. Reuter,
R. B. Best, I. Vattulainen, L. Monticelli, X. Periole,
D. P. Tieleman, A. H. de Vries and S. J. Marrink, Supporting
Information for: Martini 3: A General Purpose Force Field
for Coarse-Grained Molecular Dynamics, Nat. Methods, 2021,
18(4), 382388, DOI: 10.1038/s41592-021-01098-3.
25 M. Dewaele, D. Truer-Boutry, J. Devaux and G. Leloup,
Volume Contraction in Photocured Dental Resins: The
Shrinkage-Conversion Relationship Revisited, Dent. Mater.,
2006, 22(4), 359365, DOI: 10.1016/j.dental.2005.03.014.
26 F. Grunewald, G. Rossi, A. H. de Vries, S. J. Marrink and
L. Monticelli, Transferable MARTINI Model of
Poly(Ethylene Oxide), J. Phys. Chem. B, 2018, 122(29), 7436
7449, DOI: 10.1021/acs.jpcb.8b04760.
27 J. Han, R. H. Gee and R. H. Boyd, Glass Transition
Temperatures of Polymers from Molecular Dynamics
Simulations, Macromolecules, 1994, 27(26), 77817784, DOI:
10.1021/ma00104a036.
28 G. Rossi, I. Giannakopoulos, L. Monticelli, N. K. J. Rostedt,
S. R. Puisto, C. Lowe, A. C. Taylor, I. Vattulainen and
T. Ala-Nissila, A MARTINI Coarse-Grained Model of
a Thermoset Polyester Coating, Macromolecules, 2011,
44(15), 61986208, DOI: 10.1021/ma200788a.
© 2022 The Author(s). Published by the Royal Society of Chemistry RSC Adv., 2022, 12,1205312059 | 12059
Paper RSC Advances
ResearchGate has not been able to resolve any citations for this publication.
Preprint
Full-text available
The recent re-parametrization of the Martini coarse-grained force field, Martini 3, improved the accuracy of the model in predicting molecular packing and interactions in molecular dynamics simulations. Here, we describe how small molecules can be accurately parametrized within the Martini 3 framework and present a database of validated small molecule models (available at https://github.com/ricalessandri/ Martini3-small-molecules and http://cgmartini.nl). We pay particular attention to the description of aliphatic and aromatic ring-like structures, which are ubiquitous in small molecules such as solvents and drugs or in building blocks constituting macromolecules such as proteins and synthetic polymers. In Martini 3, ring-like structures are described by models that use higher resolution coarse-grained particles (small and tiny particles). As such, the present database constitutes one of the cornerstones of the calibration of the new Martini 3 small and tiny particle sizes. The models show excellent partitioning behavior and solvent properties. Miscibility trends between different bulk phases are also captured, completing the set of thermodynamic properties considered during the parametrization. We also show how the new bead sizes allow for a good representation of molecular volume, which translates into better structural properties such as stacking distances. We further present design strategies to build Martini 3 models for small molecules of increased complexity. The present database, along with the outlined design strategies and the modularity of the Martini force field, constitute a resource of models that 1) can be used “as is” in (bio)molecular simulations; 2) gives reference points for the construction of other small molecule models; 3) provides guidelines and reference data for automated topology builders; and 4) can be used as building blocks for the construction of more complex (macro)molecules, hence enabling investigations of complex biomolecular and soft material systems.
Article
Full-text available
The coarse-grained Martini force field is widely used in biomolecular simulations. Here we present the refined model, Martini 3 (http://cgmartini.nl), with an improved interaction balance, new bead types and expanded ability to include specific interactions representing, for example, hydrogen bonding and electronic polarizability. The updated model allows more accurate predictions of molecular packing and interactions in general, which is exemplified with a vast and diverse set of applications, ranging from oil/water partitioning and miscibility data to complex molecular systems, involving protein–protein and protein–lipid interactions and material science applications as ionic liquids and aedamers.
Article
Full-text available
Recent advances in graphics processing unit (GPU) hardware and improved efficiencies of atomistic simulation programs allow for the screening of a large number of polymers to predict properties that require running and analyzing long molecular dynamics (MD) trajectories. This paper outlines a MD simulation workflow based on GPU MD simulation and the refined optimized potentials for liquid simulation (OPLS) OPLS3e force field to calculate glass transition temperatures (Tgs) of 315 polymers for which Bicerano reported experimental values [Bicerano, J. Prediction of Polymer Properties; Marcel Dekker Inc.: New York, 1996]. Applying the workflow across this large set of polymers allowed for a comprehensive evaluation of the protocol performance and helped in understanding its merits and limitations. We observe a consistent trend between predicted Tg values and empirical observation across several subsets of polymers. Thus, the protocol established in this work is promising for exploring targeted chemical spaces and aids in the evaluation of polymers for various applications, including composites, coatings, electrical casings, etc. During the stepwise cooling simulation for the calculation of Tg, a subset of polymers clearly showed an ordered structure developing as the temperature decreased. Such polymers have a point of discontinuity on the specific volume vs temperature plot, which we associate with the melting temperature (Tm). We demonstrate the distinction between crystallized and amorphous polymers by examining polyethylene. Linear polyethylene shows a discontinuity in the specific volume vs temperature plot, but we do not observe the discontinuity for branched polyethylene simulations.
Article
Full-text available
The computational and conceptual simplifications realized by coarse-grain (CG) models make them a ubiquitous tool in the current computational modeling landscape. Building block based CG models, such as the Martini model, possess the key advantage of allowing for a broad range of applications without the need to reparametrize the force field each time. However, there are certain inherent limitations to this approach, which we investigate in detail in this work. We first study the consequences of the absence of specific cross Lennard-Jones parameters between different particle sizes. We show that this lack may lead to artificially high free energy barriers in dimerization profiles. We then look at the effect of deviating too far from the standard bonded parameters, both in terms of solute partitioning behavior and solvent properties. Moreover, we show that too weak bonded force constants entail the risk of artificially inducing clustering, which has to be taken into account when designing elastic network models for proteins. These results have implications for the current use of the Martini CG model and provide clear directions for the reparametrization of the Martini model. Moreover, our findings are generally relevant for the parametrization of any other building block based force field.
Article
Full-text available
Motivated by the deficiencies of the previous MARTINI models of polyethylene oxide (PEO), we present a new model featuring a high degree of transferability. The model parameterization is based on (a) a set of 8 free energies of transfer of dimethoxyethane (PEO dimer) from water to solvents of varying polarity; (b) the radius of gyration of the polymer in water at high dilution; and (c) matching angle and dihedral distributions from atomistic simulations. We demonstrate that our model behaves well in five different areas of application: (1) it produces accurate densities and phase behavior or small PEO oligomers and water mixtures; (2) it yields chain dimensions in good agreement with experiment in three different solvents (water, diglyme and benzene) over a broad range of molecular weights (~1.2 kg/mol to 21 kg/mol); (3) it reproduces qualitatively the structural features of lipid bilayers containing PEGylated lipids in the brush and mushroom regime; (4) it is able to reproduce the phase behavior of several PEO-based non-ionic surfactants in water; (5) it can be combined with the existing MARTINI PS to model PS-PEO block-copolymers. Overall, the new PEO model outperforms previous models and features a high degree of transferability.
Article
Full-text available
Additive manufacturing (AM) alias 3D printing translates computer-aided design (CAD) virtual 3D models into physical objects. By digital slicing of CAD, 3D scan, or tomography data, AM builds objects layer by layer without the need for molds or machining. AM enables decentralized fabrication of customized objects on demand by exploiting digital information storage and retrieval via the Internet. The ongoing transition from rapid prototyping to rapid manufacturing prompts new challenges for mechanical engineers and materials scientists alike. Because polymers are by far the most utilized class of materials for AM, this Review focuses on polymer processing and the development of polymers and advanced polymer systems specifically for AM. AM techniques covered include vat photopolymerization (stereolithography), powder bed fusion (SLS), material and binder jetting (inkjet and aerosol 3D printing), sheet lamination (LOM), extrusion (FDM, 3D dispensing, 3D fiber deposition, and 3D plotting), and 3D bioprinting. The range of polymers used in AM encompasses thermoplastics, thermosets, elastomers, hydrogels, functional polymers, polymer blends, composites, and biological systems. Aspects of polymer design, additives, and processing parameters as they relate to enhancing build speed and improving accuracy, functionality, surface finish, stability, mechanical properties, and porosity are addressed. Selected applications demonstrate how polymer-based AM is being exploited in lightweight engineering, architecture, food processing, optics, energy technology, dentistry, drug delivery, and personalized medicine. Unparalleled by metals and ceramics, polymer-based AM plays a key role in the emerging AM of advanced multifunctional and multimaterial systems including living biological systems as well as life-like synthetic systems.
Article
Full-text available
The accurate calculation of protein/nucleic acid-ligand interactions or condensed phase properties by force field-based methods require a precise description of the energetics of intermolecular interactions. Despite the progress made in force fields, small molecule parameterization remains an open problem due to the magnitude of the chemical space; the most critical issue is the estimation of a balanced set of atomic charges with the ability to reproduce experimental properties. The LigParGen web server provides an intuitive interface for generating OPLS-AA/1.14*CM1A(-LBCC) force field parameters for organic ligands, in the formats of commonly used molecular dynamics and Monte Carlo simulation packages. This server has high value for researchers interested in studying any phenomena based on intermolecular interactions with ligands via molecular mechanics simulations. It is free and open to all at jorgensenresearch.com/ligpargen, and has no login requirements.
Article
Since the classical molecular dynamics simulator LAMMPS was released as an open source code in 2004, it has become a widely-used tool for particle-based modeling of materials at length scales ranging from atomic to mesoscale to continuum. Reasons for its popularity are that it provides a wide variety of particle interaction models for different materials, that it runs on any platform from a single CPU core to the largest supercomputers with accelerators, and that it gives users control over simulation details, either via the input script or by adding code for new interatomic potentials, constraints, diagnostics, or other features needed for their models. As a result, hundreds of people have contributed new capabilities to LAMMPS and it has grown from fifty thousand lines of code in 2004 to a million lines today. In this paper several of the fundamental algorithms used in LAMMPS are described along with the design strategies which have made it flexible for both users and developers. We also highlight some capabilities recently added to the code which were enabled by this flexibility, including dynamic load balancing, on-the-fly visualization, magnetic spin dynamics models, and quantum-accuracy machine learning interatomic potentials. Program Summary Program Title: Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) CPC Library link to program files: https://doi.org/10.17632/cxbxs9btsv.1 Developer's repository link: https://github.com/lammps/lammps Licensing provisions: GPLv2 Programming language: C++, Python, C, Fortran Supplementary material: https://www.lammps.org Nature of problem: Many science applications in physics, chemistry, materials science, and related fields require parallel, scalable, and efficient generation of long, stable classical particle dynamics trajectories. Within this common problem definition, there lies a great diversity of use cases, distinguished by different particle interaction models, external constraints, as well as timescales and lengthscales ranging from atomic to mesoscale to macroscopic. Solution method: The LAMMPS code uses neighbor lists, parallel spatial decomposition, and parallel FFTs for long-range Coulombic interactions [1]. The time integration algorithm is based on the Størmer-Verlet symplectic integrator [2], which provides better stability than higher-order non-symplectic methods. In addition, LAMMPS supports a wide range of interatomic potentials, constraints, diagnostics, software interfaces, and pre- and post-processing features. Additional comments including restrictions and unusual features: This paper serves as the definitive reference for the LAMMPS code. References [1]S. Plimpton. Fast parallel algorithms for short-range molecular dynamics. J. Comp. Phys., 117:1–19, 1995. [2]L. Verlet. Computer experiments on classical fluids: I. Thermodynamical properties of Lennard–Jones molecules. Phys. Rev., 159:98–103, 1967.
Article
Molecular modeling and simulations are invaluable tools for the polymer science and engineering community. These computational approaches enable predictions and provide explanations of experimentally observed macromolecular structure, dynamics, thermodynamics, and microscopic and macroscopic material properties. With recent advances in computing power, polymer simulations can synergistically inform, guide, and complement in vitro macromolecular materials design and discovery efforts. To ensure that this growing power of simulations is harnessed correctly, and meaningful results are achieved, care must be taken to ensure the validity and reproducibility of these simulations. With these considerations in mind, in this Perspective we discuss our philosophy for carefully developing or selecting appropriate models, performing, and analyzing polymer simulations. We highlight best practices, key challenges, and important advances in model development/selection, computational method choices, advanced sampling methods, and data analysis, with the goal of educating potential polymer simulators about ways to improve the validity, usefulness, and impact of their polymer computational research.
Article
A computational-experimental model for identifying the residual (shrinkage) stresses arising in parts after their additive manufacturing by the layer-by-layer photopolymerizationmethod is given. The model is based on an analysis of flexible shrinkage strains of a series of specimens shaped as plate-strips with the same parameters but with different polymerization time which is prescribed in their manufacturing on the 3d-printer.