Next Article in Journal
Synthesis, Characterization and Evaluation of Supercapacitive Response of Dodecylbenzenesulphonic Acid (DBSA) Doped Polypyrrole/Zirconium Dioxide Composites
Next Article in Special Issue
Temperature Dependence of Conformational Relaxation of Poly(ethylene oxide) Melts
Previous Article in Journal
Hybrid Composite Material Reinforced with Carbon Nanolaminates for Gradient Stiffness: Preparation and Characterization
Previous Article in Special Issue
Compressive Properties and Constitutive Model of Semicrystalline Polyethylene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanocomposite of Fullerenes and Natural Rubbers: MARTINI Force Field Molecular Dynamics Simulations

by
Jiramate Kitjanon
1,2,
Wasinee Khuntawee
1,2,3,
Saree Phongphanphanee
2,3,4,
Thana Sutthibutpong
2,3,5,
Nattaporn Chattham
1,
Mikko Karttunen
6,7,8 and
Jirasak Wong-ekkabut
1,2,3,*
1
Department of Physics, Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
2
Computational Biomodelling Laboratory for Agricultural Science and Technology (CBLAST), Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
3
Thailand Center of Excellence in Physics (ThEP Center), Commission on Higher Education, Bangkok 10400, Thailand
4
Department of Material Science, Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
5
Department of Physics, Faculty of Science, King Mongkut’s University of Technology Thonburi (KMUTT), Bangkok 10140, Thailand
6
Department of Chemistry, The University of Western Ontario, 1151 Richmond Street, London, ON N6A 3K7, Canada
7
Department of Physics and Astronomy, The University of Western Ontario, 1151 Richmond Street, London, ON N6A 3K7, Canada
8
The Center for Advanced Materials and Biomaterials Research, The University of Western Ontario, London, ON N6A 3K7, Canada
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(22), 4044; https://doi.org/10.3390/polym13224044
Submission received: 12 October 2021 / Revised: 12 November 2021 / Accepted: 16 November 2021 / Published: 22 November 2021
(This article belongs to the Special Issue Computational Modeling of Polymers)

Abstract

:
The mechanical properties of natural rubber (NR) composites depend on many factors, including the filler loading, filler size, filler dispersion, and filler-rubber interfacial interactions. Thus, NR composites with nano-sized fillers have attracted a great deal of attention for improving properties such as stiffness, chemical resistance, and high wear resistance. Here, a coarse-grained (CG) model based on the MARTINI force field version 2.1 has been developed and deployed for simulations of cis-1,4-polyisoprene (cis-PI). The model shows qualitative and quantitative agreement with the experiments and atomistic simulations. Interestingly, only a 0.5% difference with respect to the experimental result of the glass transition temperature (Tg) of the cis-PI in the melts was observed. In addition, the mechanical and thermodynamical properties of the cis-PI-fullerene(C60) composites were investigated. Coarse-grained molecular dynamics (MD) simulations of cis-PI-C60 composites with varying fullerene concentrations (0–32 parts per hundred of rubber; phr) were performed over 200 microseconds. The structural, mechanical, and thermal properties of the composites were determined. The density, bulk modulus, thermal expansion, heat capacity, and Tg of the NR composites were found to increase with increasing C60 concentration. The presence of C60 resulted in a slight increasing of the end-to-end distance and radius of the gyration of the cis-PI chains. The contribution of C60 and cis-PI interfacial interactions led to an enhancement of the bulk moduli of the composites. This model should be helpful in the investigations and design of effective fillers of NR-C60 composites for improving their properties.

Graphical Abstract

1. Introduction

Natural rubber (NR), mainly consisting of cis-1,4polyisoprene (cis-PI) with high elasticity, is acknowledged to be virtually irreplaceable in applications such as tires and seals. However, the neat NR usually suffers from poor mechanical strength [1,2]. To alleviate this problem, reinforcing fillers are widely used to improve the mechanical properties of NR materials. Impressive enhancements in the properties, including mechanical [3,4], optical [3,5,6], self-healing [7,8], electrical [4,6,9], thermal [4,6], rheological [4,9], and glass transition [10,11], have been reported. The mechanical properties of NR composites depend on many factors, such as the types, dimensions, concentrations, dispersion states, and alignments of the fillers [4,12,13].
Among the most important methods of tuning the properties of NR are the vulcanization and reinforcement by composites, in particular carbon black (CB) [14,15] Here, the focus is on the latter. The reinforcement of NR composites by carbon black (CB) has been intensively studied and broadly applied in truck and car tires [16,17]. CB may improve tensile properties [18], fatigue resistance [19], stiffness, and wear resistance [20]. The CB concentration has a significant influence on the rubber composite properties; e.g., increasing the amount of CB has been shown to lead to enhanced tear strength, tensile strength, and elastic modulus [21,22]. The study of Wang et al. [23] revealed that the piezoresistivity of CB-NR composites monotonically decreases with decreasing CB concentration. However, CB-reinforced nanocomposites at high loading have shown the decay of the storage modulus at high deformation, indicating high hysteresis [24,25]. This is called the Payne effect, and it has limited the application of CB-reinforced nanocomposites to tires. Moreover, CB is produced from petroleum feedstock, which has a negative impact on the environment [26].
NR composited with carbon-based nanoparticles, e.g., carbon nanotubes (CNT) or graphene, offers the possibility to produce lightweight composites with enhanced electrical and magnetic properties and thermal conductivity [4,27,28]. For example, it has been shown that a small amount of multiwalled carbon nanotubes (MWCNTs) at 0.5 parts per hundred (phr) in NR-MWCNTs composite could enhance the tensile strength, tensile modulus, and tear strength by 61%, 75%, and 59%, respectively [29]. Fullerene(C60) has been extensively used as a filler in polymer composites, especially for poly(3-hexylthiophene) in polymer-based solar cells for low-cost power production [30,31,32], and, in combination with epoxy, it has been shown to result in enhancements in the mechanical properties, electrical conductivity, flame resistance, and anti-corrosion [33]. In addition, the presence of fullerenes in NR composites has been shown to increase the modulus at 100% (M100), the modulus at 300% (M300) of elongation, and the shore hardness [34]. It has also been reported that the addition of 0.75 phr of fullerene in NR increases the M100, M200, M300, and hardness by 178%, 137%, 113%, and 20.5%, respectively, compared to composites at 0.0065 phr of fullerene [35]. After aging (heating at 70° C for 168 h), Jurkowska et al. found an increase in the elastic modulus for [C60] < 0.5 phr, and, interestingly, the elastic modulus decreased at [C60] = 0.75 phr. That could substantially reduce the degradation effects of rubber composites [35].
Atomistic molecular dynamics (MD) simulations can provide nanoscale information, such as the dispersion and aggregation of fillers [36], segmental dynamics [37], polymer chain alignment [38], molecular interactions [36,37,39], and filler size effects [40]. Guseva et al. [37] investigated the macroscopic properties of non-crosslinked cis-PI films confined with silica substrate using atomistic MD simulations. Liu et al. [41] reported that the glass transition temperature (Tg) of the cis-PI film increased when the film thickness was decreased. They also reported that, compared to a pure polymer system, the Tg increased when the filler was added. Raffaini et al. [42] studied rubber–fullerene conformations and observed fullerene aggregation when rubber chains were modified with COOH-termination. Although atomistic MD has been widely used in investigations of reinforced polymer nanocomposites, such simulations are computationally expensive. To overcome this limitation, coarse-grained (CG) models have been developed to produce simplified representations of polymer molecules to access much larger time and length scales.
One of the earlier works on the simulations of polymer melts was the seminal paper by Kremer and Grest focusing on melts well above the glass transition temperature [43]. An important step toward coarse-graining was taken in the 1995 article of Forrest and Suter, who showed how softer pairwise interactions emerge from an atomistic model upon averaging over the pair distribution [44]. The following years were very active for the development of CG models for polymers and polymer melts, including the works of Akkermans and Briels, who constructed a CG model with a focus on the preservation of thermodynamic properties [45], followed by a number of studies on the systematic coarse-graining of polymers and melts [46,47], modification of the dissipative particle dynamics (DPD) method for melts [48], and fluctuating soft-sphere models [49] with the aim of improving the accuracy of CG polymer melt modeling. Regarding the interactions of fullerenes and polymers, Huang et al. [50] developed a polymer-fullerene CG model for polymer-based solar cells, and Volgin et al. [51] investigated the diffusion of fullerenes using a CG model. Although there is no lack of CG models, the MARTINI force field [52,53] has become by far the most dominant one, and it has been used for simulations of amino acids, water, phospholipid membranes, fullerenes, some polymers, and RNA [54,55,56,57,58,59,60,61].
Recently, we introduced a MARTINI CG model for cis-PI to study natural rubber [62]. The interactions of the cis-PI and fullerene, however, were not included in the model. That is done in the current work by combining the CG models of cis-PI and fullerene to study cis-PI-C60 composites. The parameters are based on the MARTINI force field version 2.1 [52]. The thermodynamic and mechanical properties of the cis-PI-C60 composites were calculated and compered with the experiments and atomistic MD simulations. The influence of the C60 concentrations on the structural, mechanical, and thermal properties of the composites was analyzed. The results show that the interfacial interactions between the C60 and cis-PI play an essential role in the composites’ mechanical properties.

2. Methodology

2.1. Molecular Dynamics Simulations

In this study, cis-PI in melts and cis-PI composites at different C60 concentrations were investigated by coarse-grained (CG) molecular dynamics (CGMD) simulations. The CG model for cis-PI was developed based on the MARTINI force field version 2.1 (University of Groninge, Groningen, The Netherlands) and is described in detail in Table S1. A coarse-grained or superatom bead was used to represent an isoprene monomer. CG mapping of stretching and bending parameters for the cis-PI beads from united-atom MD trajectories were published in our previous work [62]. All systems consisted of 500 chains of 32-mers cis-PI, which is equivalent to 16,000 CG beads in total (an isoprene monomer was mapped into one bead in the CG model).
C60 molecules were randomly placed into the cis-PI melts with the number of 53, 107, 213, and 427 molecules corresponding to concentrations of 4, 8, 16, and 32 parts per hundred of rubber (phr), respectively. The details of all simulations are shown in Table S2. The CG fullerene model for the C60 molecule called F16 was taken from Ref. [63]. After steepest descent energy minimization, CGMD simulations were performed using the GROMACS 5.1.1 (University of Groningen, Royal Institute of Technology, Groningen, The Netherlands) package [64]. The number of particles, temperature, and pressure were kept constant (NPT ensemble). cis-PI and C60 molecules were thermostatted separately at 300 K using the Parrinello-Donadio-Bussi velocity rescale algorithm with a time constant of 30 ps [65,66]. Pressure was held constant at 1 bar by the Parrinello-Rahman barostat [67] with a time constant of 1 ps and compressibility of 4.5 × 10−5 bar−1. The reaction-field method was used for the long-range electrostatic interactions [68,69], Lennard-Jones interactions were cut off at 1.1 nm, and periodic boundary conditions were applied in all directions. The simulation protocol had been tested and used in several prior studies [38,70,71,72,73]. All production simulations were run with time step of 30 ps for 21 µs. To confirm that equilibration had been reached, time evolution for the end-to-end distance and the radius of gyration were calculated (see Figures S1 and S2). The last 5 μs were used for data analysis and errors were estimated using standard deviation (SD). Molecular visualizations were done using the Visual Molecular Dynamics (VMD) software (University of Illinois Urbana-Champaign, Urbana and Champaign, IL, USA) [74].

2.2. Calculation of Thermal and Mechanical Properties

Glass Transition Temperature (Tg)
The final states of the simulations after 21 µs were used as the initial structures to calculate the glass transition temperatures (Tg). The systems were cooled from 300 K to 100 K with 10 K interval. The cooling rate was 0.1 K.ns−1. To equilibrate the systems, additional 500 ns were simulated at each temperature in the NPT ensemble. Tg was calculated using the intersection of the two lines describing the behavior at low and high temperatures on density-temperature (ρ-T) curves [75,76,77]. Density versus temperature curves for cis-PI in melts and composites are shown in Figure S3.
The thermal volume expansion coefficient (γ), specific heat capacity (cp), and bulk modulus (κ) were calculated from the fluctuations of the simulation box volumes (V) and enthalpy (H) using:
γ = V H V H N A k B T 2 V
c p = ( H H ) 2 m N A k B T 2
κ = k B T V ( V V ) 2
where kB is the Boltzmann constant, NA Avogadro’s number, and m is the total mass of the system in atomic mass units. The angular brackets refer to averaging over simulation time.

2.3. Calculation of Solvation Free Energy

Thermodynamic integration (TI) [78] was applied to calculate solvation free energies of C60 in water, pure cis-PI, and cis-PI-C60 composites. The last frame from each of the 21 µs simulations was used as the initial structure. A C60 was added into the system and used as a coupling molecule. After steepest descent energy minimization, the MD simulations with the coupling parameter (λ) between 0 and 1 with interval of 0.05 were performed under the NPT ensemble. All systems were run for 300 ns, and the last 100 ns was used for analysis. The Bennett acceptance ratio (BAR) was applied to estimate the solvation free energy [79]. The solvation free energies are shown in Table S3. The effect of C60 concentration is discussed in Section 3.4 (Figure 5).

3. Results and Discussion

3.1. Coarse-Grained Model Based on MARTINI Force Field of cis-1,4-Polyisoprene

The macroscopic and structural properties of the cis-PI in the melts were analyzed and compared to the experiments and atomistic simulations in order to validate the CG model (Table 1). The system density was found to be in good agreement, with the values being about 19% and 27% higher than in the experiments [80] and previous united-atom MD (UAMD) simulations [38], respectively. The thermal expansion coefficient was found to be 52% and 19% higher than in the experiments [81,82] and UAMD [38], respectively. In addition, the squared end-to-end distance (R0), the squared radius of gyration (Rg), and polymer expansion factor (< R 0 2 >/< R g 2 >) of the cis-PI in the melts are within 5% from the UAMD data [38]. R0 is defined as the average distance between the first and the last CG bead of the cis-PI chain. Rg was computed using.
R g 2 = 1 N i = 1 N ( r i r c m ) 2
where ri is the position of particle i and rcm is the position of the center of mass of the cis-PI chain. However, the bulk modulus was 3.8-fold lower than in the experiments [83] and 2.6-fold below the previous UAMD simulations [38]. One of the major contributors to this decrease of bulk modulus is the reduction in the degrees of freedom upon coarse-graining monomers into superatom beads, which diminishes the friction between branching atoms and changes entropy [84,85]. Higher mobility of the CG chains causes large fluctuations of the simulation box volume compared to the UAMD simulations [38].

3.2. Effect of C60 Concentration on Macroscopic and Structural Properties of cis-PI Composites

Adding C60 into the cis-PI composites leads to a linear increase in density (Figure 1a) due to the mass of the C60 filler. Figure 1b shows the enhancement of the bulk modulus as a function of the C60 concentration. A slow increase was found for [C60] < 16 phr (2.17 × 10−4 GPa per 1 phr), followed by a rapid increase at [C60] = 32 phr (4.81 × 10−4 GPa per 1 phr). The increases in the density and bulk moduli as a function of the C60 concentration is similar to previous UA simulations [86]. In addition, the thermal expansion coefficient (γ) and specific heat capacity (cp) at 300 K were computed using Equations (1) and (2). In the composites, both quantities increased when the amount of C60 increased (Figure 1c,d), similar to what has been observed in previous MD [87] and experimental studies [88] of polymer carbon-based nanocomposites.
The structural changes of the cis-PI chains at different C60 concentrations were also investigated by monitoring the R0 and Rg. The averages of the R0 and Rg for the last 5 μs are provided in Table 2. For the cis-PI chains, they increased slightly upon increasing the amount of C60. The polymer chain expansion factor (< R02 >/< Rg2 >) of the cis-PI in the melts was 6.11, which is in good agreement with previous UAMD simulations (6.20 [38] and 5.44 [77]) and refers to a random coil structure of a cis-PI chain [89].

3.3. Effect of C60 Concentration on Microscopic Properties of cis-PI Composites

3.3.1. Interaction of C60-C60 and C60-cis-PI

To investigate the interaction free energy of C60-C60 and C60-cis-PI in the composites at different C60 concentrations, the potential of mean force (PMF) was calculated from the radial distribution function (RDF; g(r)) [90,91] as:
PMF = k B T ln ( g ( r ) )
where k B is the Boltzmann constant and T is the absolute temperature.
The dimerization free energies of C60 in cis-PI composites at different C60 concentrations are shown in Figure 2a. The first, second, and third minima of the C60-C60 PMF corresponding to the distances between the neighboring C60 molecules are at 1.0 nm, 1.5 nm, and 1.9 nm, respectively. The second minimum is at the lowest free energy, suggesting energetic preference of C60 dimerization with the cis-PI chain insertion. That the polymer chains fill the space between the C60 molecules has also been observed in previous studies [38,59,60,92]. The PMF for C60 and cis-PI interactions is shown in Figure 2b. This result demonstrates the C60-cis-PI composites’ preference to a dispersed structure corresponding to the snapshots in (Figure 3), which show the dispersed structure for all the C60 concentrations.
In Figure 2a,b, the first and second local minima are at 0.8 nm and 1.2 nm, respectively. Regarding the first local minima in Figure 2a,b, the free energy for C60-cis-PI is lower than for C60-C60, indicating a preference for the C60-cis-PI interaction. Moreover, in the case of C60-C60 interaction, a concentration dependence was observed in which the free energy decreased slowly upon increasing the C60 concentration from 4 to 8 phr. For the C60-cis-PI, only a minor concentration dependence is present.

3.3.2. Diffusion of C60 and cis-PI in the Composites

The diffusion coefficients of the C60 and cis-PI in the melts and in the composites were analyzed at different C60 concentrations using the mean squared displacement (MSD)
M S D = r 2 ( t )   ~   6 D t  
where D is the diffusion coefficient and t is time.
In agreement with previous studies [38,87], the presence of C60 caused a dramatic decrease in both the C60 and cis-PI diffusion coefficients (Figure 4). The slow movement at high C60 concentration led to confinement, with the C60 molecules being more mobile than the cis-PI ones. The difference between the diffusion coefficients of C60 and cis-PI decreased upon increasing the C60 concentration; the increased interactions between C60 and cis-PI cause a reduction in the diffusive motions of C60 and cis-PI, which is the reason why the bulk modulus of the composites at high C60 concentrations becomes enhanced. Note that the diffusion coefficients, extracted from the CG simulations, were one order of magnitude faster than those obtained from the UA simulations [38]. This is due to the reduction of degrees of freedom in the CG model [93].

3.4. Effect of C60 Concentration on C60 Solvation Free Energy

The solvation free energies of C60 in water, cis-PI melt, and cis-PI-C60 matrix composites were estimated by thermodynamic integration (TI) [78]. The solvation free energy of C60 in water was −99.53 ± 0.03 kJ/mol, which is in good agreement with our previous CGMD calculation using umbrella sampling (−92.6 kJ/mol) [94]. Figure 5 shows the concentration dependence of the solvation free energy. The lowest value (~−115.53 kJ/mol) was found when an added C60 was solvated in a cis-PI melt. It indicates that the cis-PI melt is more favorable than water. Increasing the amount of C60 concentration led to lesser solvation in the composites. Our results suggest that the C60 prefers to interact with the cis-PI chains rather than the other C60 molecules in the cis-PI-C60 composites (Figure 2a,b).

3.5. Effect of C60 Concentration on Glass Transition Temperature (Tg)

As discussed above, the CGMD simulations of cis-PI-C60 composites using our model are able to reproduce the results from previous experiments and atomistic MD simulations [50,95]. The glass transition temperature (Tg) is one of the most important parameters determining the physical state and final mechanical properties of polymer composites [96]. To gain more understanding on the influence of the filler concentration on composites’ properties, Tg was estimated using the change of slope in density (ρ) versus temperature (T) curve [75,76,77]. The ρ-T curves of the cis-PI in the melts and composites are shown in Figure S3. The calculated Tg was 201 K, in good agreement with the all-atom (209 K) [77] and united-atom models (223 K) of Sharma et al., [77], and experimental data (200 K) of Brandrup et al. [82] Finally, the concentration dependence of the Tg in the composites was determined. The data in Figure 6 show a linear increase with a slope of 0.63 K/phr. This same trend has been reported in previous experiments [34] and simulations [41,87].

4. Conclusions

We have developed a MARTINI force field version 2.1-based [52] parameterization and performed systematic studies of cis-PI in melts and cis-PI-C60 composites using it. We validated the thermodynamic, macroscopic, and microscopic properties of the cis-PI in the melts, and the results show good agreement with prior experimental [80,81,82,83] and computational studies [38,77,94]. Perhaps most surprisingly, only a 0.5% difference with respect to the experimentally found glass transition temperature (Tg) [82] in the cis-PI in the melts was found.
The properties of the NR-fullerene composites at different C60 concentrations (0, 4, 8, 16, and 32 phr) were studied by using CGMD simulations over 200 microseconds. The density, bulk modulus, thermal expansion, heat capacity, and glass transition temperature increased upon increasing the C60 concentration. A slight increase in the R0 and Rg of the cis-PI chains was found for the NR-C60 composites. The interaction energies between the C60 and cis-PI decreased when the amount of C60 increased. The diffusion of the cis-PI and C60 was slowed down by increasing the C60 concentration. The decrease of the interaction free energies and the diffusion coefficients resulted in an enhancement of the bulk modulus of the composites. Moreover, the solvation free energies of the C60 in the cis-PI matrix composites increased upon increasing the C60 concentration. These results suggest that C60-cis-PI interactions are more preferable than C60-C60 self-interactions.
The CG model introduced here allows simulations of large systems over long periods of time to study properties such as the impact of the filler concentration on rubber composites at the molecular level. This validated cis-PI-C60 CG model based on the MARTINI force field will also enable simulations of the advanced rubber materials and their interactions with biological molecules.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/polym13224044/s1, Table S1: Details of the CG force field for cis-PI, Table S2: Number of molecules and simulation time, Table S3: Solvation free energy of an added fullerene in water, cis-PI in melt and composite at different C60 concentrations, Figure S1: Time evolution of end-to-end distance (R0) with varies C60 concentration from 0 to 32 phr, Figure S2: Time evolution of radius of gyration (Rg) with varies C60 concentration from 0 to 32 phr, Figure S3: Density versus temperature of cis-PI-C60 composites.

Author Contributions

Conceptualization, J.W.-e; methodology, J.K. and J.W.-e.; software, J.K., T.S. and J.W.-e.; validation, J.K., W.K. and J.W.-e.; formal analysis, J.K.; investigation, J.K., W.K. and J.W.-e.; resources, J.W.-e.; data curation, J.K., W.K.; writing—original draft preparation, J.K., W.K. and J.W.-e.; writing—review and editing, J.W.-e., M.K., S.P., T.S., and N.C.; visualization, J.K.; supervision, J.W.-e. and M.K.; project administration, J.W.-e.; funding acquisition, J.W.-e. and N.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Council of Thailand (NRCT), Thailand Science Research and Innovation (TSRI), the Kasetsart University Research and Development Institute (KURDI) and the Faculty of Science, Kasetsart University through the Research Grants for Talented Mid-Career Researchers Grant No. N41A640080, the Royal Golden Jubilee Ph.D. Program Grant No. PHD/0134/2561, and the International Frontier Research Network (IFRN) research fund. The APC was funded by the Royal Golden Jubilee Ph.D. Program Grant No. PHD/0134/2561.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This work was financially supported by the National Research Council of Thailand (NRCT), Thailand Science Research and Innovation (TSRI) through the Research Grants for Talented Mid-Career Researchers Grant No. N41A640080, and the Royal Golden Jubilee Ph.D. Program Grant No. PHD/0134/2561 (J.W.-e. and J.K.). J.W.-e., N.C., and W.K. acknowledged the Kasetsart University Research and Development Institute (KURDI) and the Faculty of Science, Kasetsart University for supporting the post-doctoral scholarship and the International Frontier Research Network (IFRN) research fund. M.K. thanks the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Canada Research Chairs Program for financial support. Computing facilities were provided by SHARCNET (www.sharcnet.ca), Compute Canada (www.computecanada.ca), and the Department of Physics, Faculty of Science, Kasetsart University.

Conflicts of Interest

The authors declare no competing interest.

References

  1. Park, C.Y. Cure characteristics and dynamic mechanical properties of acrylic rubber and epoxidized natural rubber blend. J. Ind. Eng. Chem. 2001, 7, 212–217. [Google Scholar]
  2. Chollakup, R.; Suethao, S.; Suwanruji, P.; Boonyarit, J.; Smitthipong, W. Mechanical properties and dissipation energy of carbon black/rubber composites. Compos. Adv. Mater. 2021, 30, 1–6. [Google Scholar] [CrossRef]
  3. Ray, S.S.; Okamoto, M. Polymer/layered silicate nanocomposites: A review from preparation to processing. Prog. Polym. Sci. 2003, 28, 1539–1641. [Google Scholar]
  4. Moniruzzaman, M.; Winey, K.I. Polymer nanocomposites containing carbon nanotubes. Macromolecules 2006, 39, 5194–5205. [Google Scholar] [CrossRef]
  5. Caseri, W. Nanocomposites of polymers and metals or semiconductors: Historical background and optical properties. Macromol. Rapid Commun. 2000, 21, 705–722. [Google Scholar] [CrossRef]
  6. Xie, X.-L.; Mai, Y.-W.; Zhou, X.-P. Dispersion and alignment of carbon nanotubes in polymer matrix: A review. Mater. Sci. Eng. R Rep. 2005, 49, 89–112. [Google Scholar] [CrossRef]
  7. Balazs, A.C.; Emrick, T.; Russell, T.P. Nanoparticle polymer composites: Where two small worlds meet. Science 2006, 314, 1107–1110. [Google Scholar] [CrossRef]
  8. Tee, B.C.K.; Wang, C.; Allen, R.; Bao, Z. An Electrically and Mechanically Self-Healing Composite with Pressure- and Flexion-Sensitive Properties for Electronic Skin Applications. Nat. Nanotechnol. 2012, 7, 825–832. [Google Scholar] [CrossRef] [PubMed]
  9. Du, F.; Scogna, R.C.; Zhou, W.; Brand, S.; Fischer, J.E.; Winey, K.I. Nanotube networks in polymer nanocomposites: Rheology and electrical conductivity. Macromolecules 2004, 37, 9048–9055. [Google Scholar] [CrossRef]
  10. Oh, H.; Green, P.F. Polymer chain dynamics and glass transition in athermal polymer/nanoparticle mixtures. Nat. Mater. 2009, 8, 139–143. [Google Scholar] [CrossRef] [PubMed]
  11. Rittigstein, P.; Priestley, R.D.; Broadbelt, L.J.; Torkelson, J.M. Model polymer nanocomposites provide an understanding of confinement effects in real nanocomposites. Nat. Mater. 2007, 6, 278–282. [Google Scholar] [CrossRef]
  12. Suethao, S.; Phongphanphanee, S.; Wong-Ekkabut, J.; Smitthipong, W. The Relationship between the Morphology and Elasticity of Natural Rubber Foam Based on the Concentration of the Chemical Blowing Agent. Polymers 2021, 13, 1091. [Google Scholar] [CrossRef]
  13. Suethao, S.; Ponloa, W.; Phongphanphanee, S.; Wong-Ekkabut, J.; Smitthipong, W. Current challenges in thermodynamic aspects of rubber foam. Sci. Rep. 2021, 11, 1–12. [Google Scholar]
  14. Wichaita, W.; Promlok, D.; Sudjaipraparat, N.; Sripraphot, S.; Suteewong, T.; Tangboriboonrat, P. A Concise Review on Design and Control of Structured Natural Rubber Latex Particles as Engineering Nanocomposites. Eur. Polym. J. 2021, 159, 110740. [Google Scholar] [CrossRef]
  15. Calabrese, M.A.; Chan, W.Y.; Av-Ron, S.H.; Olsen, B.D. Development of a Rubber Recycling Process Based on a Single-Component Interfacial Adhesive. ACS Appl. Polym. Mater. 2021, 3, 4849–4860. [Google Scholar] [CrossRef]
  16. Gielen, D.; Bennaceur, K.; Kerr, T.; Tam, C.; Tanaka, K.; Taylor, M.; Taylor, P. Tracking Industrial Energy Efficiency and CO2 Emissions; International Energy Agency: Paris, France, 2007. [Google Scholar]
  17. Phuhiangpa, N.; Ponloa, W.; Phongphanphanee, S.; Smitthipong, W. Performance of nano- and microcalcium carbonate in uncrosslinked natural rubber composites: New results of structure–properties relationship. Polymers 2020, 12, 2002. [Google Scholar] [CrossRef]
  18. Maiti, M.; Sadhu, S.; Bhowmick, A.K. Effect of carbon black on properties of rubber nanocomposites. J. Appl. Polym. Sci. 2005, 96, 443–451. [Google Scholar] [CrossRef]
  19. Araby, S.; Meng, Q.; Zhang, L.; Zaman, I.; Majewski, P.; Ma, J. Elastomeric composites based on carbon nanomaterials. Nanotechnology 2015, 26, 112001. [Google Scholar] [CrossRef] [PubMed]
  20. Sisanth, K.; Thomas, M.; Abraham, J.; Thomas, S. General introduction to rubber compounding. In Progress in Rubber Nanocomposites; Thomas, S., Maria, H.J., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–39. [Google Scholar]
  21. Zhang, H.; Zhang, Z.; Zhao, G.; Liu, Y. Influence of carbon black with different concentration on dynamic properties and heat buildup of semi-efficient natural rubber composites. Micro Nano Lett. 2016, 11, 402–406. [Google Scholar] [CrossRef]
  22. Savetlana, S.; Zulhendri; Sukmana, I.; Saputra, F. The effect of carbon black loading and structure on tensile property of natural rubber composite. IOP Conf. Ser. Mater. Sci. Eng. 2017, 223, 012009. [Google Scholar] [CrossRef]
  23. Luheng, W.; Tianhuai, D.; Peng, W. Influence of carbon black concentration on piezoresistivity for carbon-black-filled silicone rubber composite. Carbon 2009, 47, 3151–3157. [Google Scholar] [CrossRef]
  24. Scagliusi, S.R.; Cardoso, E.C.L.; Parra, D.F.; Lima, L.F.C.P.; Lugão, A.B. Evaluation of “Payne effect” in radiation-induced modification of chlorobutyl rubber. Radiat. Phys. Chem. 2013, 84, 42–46. [Google Scholar] [CrossRef]
  25. Payne, A.R. The dynamic properties of carbon black-loaded natural rubber vulcanizates. Part I. J. Appl. Polym. Sci. 1962, 6, 57–63. [Google Scholar] [CrossRef]
  26. United States Environmental Protection Agency. Compilation of Air Pollutant Emission Factors, 5th ed.; Office of Air Quality Planning and Standards: Durham, NC, USA, 1995. [Google Scholar]
  27. Yang, M.; Koutsos, V.; Zaiser, M. Interactions between polymers and carbon nanotubes: A molecular dynamics study. J. Phys. Chem. B 2005, 109, 10009–10014. [Google Scholar] [CrossRef]
  28. Dresselhaus, M.S.; Dresselhaus, G.; Eklund, P.C. Science of Fullerenes and Carbon Nanotubes: Their Properties and Applications; Elsevier: Amsterdam, The Netherlands, 1996. [Google Scholar]
  29. George, N.; Chandra, J.; Mathiazhagan, A.; Joseph, R. High performance natural rubber composites with conductive segregated network of multiwalled carbon nanotubes. Compos. Sci. Technol. 2015, 116, 33–40. [Google Scholar] [CrossRef]
  30. Chirvase, D.; Parisi, J.; Hummelen, J.C.; Dyakonov, V. Influence of nanomorphology on the photovoltaic action of polymer–fullerene composites. Nanotechnology 2004, 15, 1317–1323. [Google Scholar] [CrossRef] [Green Version]
  31. Thompson, B.C.; Fréchet, J.M.J. Polymer-fullerene composite solar cells. Angew. Chem. Int. Ed. 2008, 47, 58–77. [Google Scholar] [CrossRef]
  32. Dennler, G.; Scharber, M.C.; Brabec, C.J. Polymer-fullerene bulk-heterojunction solar cells. Adv. Mater. 2009, 21, 1323–1338. [Google Scholar] [CrossRef]
  33. Kausar, A. Fullerene nanofiller reinforced epoxy nanocomposites—Developments, progress and challenges. Mater. Res. Innov. 2020, 25, 175–185. [Google Scholar] [CrossRef]
  34. Al-Hartomy, O.A.; Al-Ghamdi, A.A.; Al-Salamy, F.; Dishovsky, N.; Slavcheva, D.; El-Tantawy, F. Properties of natural rubber-based composites containing fullerene. Int. J. Polym. Sci. 2012, 2012, 967276. [Google Scholar] [CrossRef] [Green Version]
  35. Jurkowska, B.; Jurkowski, B.; Kamrowski, P.; Pesetskii, S.S.; Koval, V.N.; Pinchuk, L.S.; Olkhov, Y.A. Properties of fullerene-containing natural rubber. J. Appl. Polym. Sci. 2006, 100, 390–398. [Google Scholar] [CrossRef]
  36. Liu, J.; Gao, Y.; Cao, D.; Zhang, L.; Guo, Z. Nanoparticle dispersion and aggregation in polymer nanocomposites: Insights from molecular dynamics simulation. Langmuir 2011, 27, 7926–7933. [Google Scholar] [CrossRef]
  37. Guseva, D.V.; Komarov, P.V.; Lyulin, A.V. Molecular-dynamics simulations of thin polyisoprene films confined between amorphous silica substrates. J. Chem. Phys. 2014, 140, 114903. [Google Scholar] [CrossRef] [Green Version]
  38. Khuntawee, W.; Sutthibutpong, T.; Phongphanphanee, S.; Karttunen, M.; Wong-Ekkabut, J. Molecular dynamics study of natural rubber-fullerene composites: Connecting microscopic properties to macroscopic behavior. Phys. Chem. Chem. Phys. 2019, 21, 19403–19413. [Google Scholar] [CrossRef]
  39. Ghanbari, A.; Rahimi, M.; Dehghany, J. Influence of surface grafted polymers on the polymer dynamics in a silica-polystyrene nanocomposite: A coarse-grained molecular dynamics investigation. J. Phys. Chem. C 2013, 117, 25069–25076. [Google Scholar] [CrossRef]
  40. Davris, T.; Mermet-Guyennet, M.R.B.; Bonn, D.; Lyulin, A.V. Filler size effects on reinforcement in elastomer-based nanocomposites: Experimental and simulational insights into physical mechanisms. Macromolecules 2016, 49, 7077–7087. [Google Scholar] [CrossRef]
  41. Liu, J.; Wu, Y.; Shen, J.; Gao, Y.; Zhang, L.; Cao, D. Polymer–nanoparticle interfacial behavior revisited: A molecular dynamics study. Phys. Chem. Chem. Phys. 2011, 13, 13058–13069. [Google Scholar] [CrossRef] [PubMed]
  42. Raffaini, G.; Citterio, A.; Galimberti, M.; Catauro, M. A Molecular dynamics study of noncovalent interactions between rubber and fullerenes. Macromol. Symp. 2021, 395, 2000198. [Google Scholar] [CrossRef]
  43. Kremer, K.; Grest, G.S. Dynamics of entangled linear polymer melts: A molecular-dynamics simulation. J. Chem. Phys. 1990, 92, 5057–5086. [Google Scholar] [CrossRef]
  44. Forrest, B.M.; Suter, U.W. Accelerated equilibration of polymer melts by time-coarse-graining. J. Chem. Phys. 1995, 102, 7256–7266. [Google Scholar] [CrossRef]
  45. Akkermans, R.L.C.; Briels, W.J. A structure-based coarse-grained model for polymer melts. J. Chem. Phys. 2001, 114, 1020–1031. [Google Scholar] [CrossRef]
  46. Faller, R. Automatic coarse graining of polymers. Polymer 2004, 45, 3869–3876. [Google Scholar] [CrossRef] [Green Version]
  47. Reith, D.; Meyer, H.; Müller-Plathe, F. Mapping atomistic to coarse-grained polymer models using automatic simplex optimization to fit structural properties. Macromolecules 2001, 34, 2335–2345. [Google Scholar] [CrossRef] [Green Version]
  48. Nikunen, P.; Vattulainen, I.; Karttunen, M. Reptational dynamics in dissipative particle dynamics simulations of polymer melts. Phys. Rev. E 2007, 75, 036713. [Google Scholar] [CrossRef] [Green Version]
  49. Vettorel, T.; Besold, G.; Kremer, K. Fluctuating soft-sphere approach to coarse-graining of polymer models. Soft Matter 2010, 6, 2282–2292. [Google Scholar] [CrossRef]
  50. Huang, D.M.; Faller, R.; Do, K.; Moulé, A.J. Coarse-grained computer simulations of polymer/fullerene bulk heterojunctions for organic photovoltaic applications. J. Chem. Theory Comput. 2010, 6, 526–537. [Google Scholar] [CrossRef]
  51. Volgin, I.V.; Larin, S.V.; Lyulin, A.V.; Lyulin, S.V. Coarse-grained molecular-dynamics simulations of nanoparticle diffusion in polymer nanocomposites. Polymer 2018, 145, 80–87. [Google Scholar] [CrossRef] [Green Version]
  52. Marrink, S.J.; Risselada, H.J.; Yefimov, S.; Tieleman, D.P.; de Vries, A.H. The MARTINI force field: Coarse grained model for biomolecular simulations. J. Phys. Chem. B 2007, 111, 7812–7824. [Google Scholar] [CrossRef] [Green Version]
  53. Marrink, S.J.; Tieleman, D.P. Perspective on the Martini model. Chem. Soc. Rev. 2013, 42, 6801–6822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Monticelli, L.; Kandasamy, S.K.; Periole, X.; Larson, R.G.; Tieleman, D.P.; Marrink, S.J. The MARTINI coarse-grained force field: Extension to proteins. J. Chem. Theory Comput. 2008, 4, 819–834. [Google Scholar] [CrossRef]
  55. Marrink, S.J.; de Vries, A.H.; Mark, A.E. Coarse grained model for semiquantitative lipid simulations. J. Phys. Chem. B 2004, 108, 750–760. [Google Scholar] [CrossRef] [Green Version]
  56. Wong-Ekkabut, J.; Baoukina, S.; Triampo, W.; Tang, I.-M.; Tieleman, D.P.; Monticelli, L. Computer simulation study of fullerene translocation through lipid membranes. Nat. Nanotechnol. 2008, 3, 363. [Google Scholar] [CrossRef]
  57. Rossi, G.; Monticelli, L.; Puisto, S.R.; Vattulainen, I.; Ala-Nissila, T. Coarse-graining polymers with the MARTINI force-field: Polystyrene as a benchmark case. Soft Matter 2011, 7, 698–708. [Google Scholar] [CrossRef] [Green Version]
  58. Uusitalo, J.J.; Ingólfsson, H.I.; Marrink, S.J.; Faustino, I. Martini coarse-grained force field: Extension to RNA. Biophys. J. 2017, 113, 246–256. [Google Scholar] [CrossRef] [Green Version]
  59. Nisoh, N.; Jarerattanachat, V.; Karttunen, M.; Wong-Ekkabut, J. Formation of aggregates, icosahedral structures and percolation clusters of fullerenes in lipids bilayers: The key role of lipid saturation. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183328. [Google Scholar] [CrossRef]
  60. Nalakarn, P.; Boonnoy, P.; Nisoh, N.; Karttunen, M.; Wong-Ekkabut, J. Dependence of fullerene aggregation on lipid saturation due to a balance between entropy and enthalpy. Sci. Rep. 2019, 9, 1037. [Google Scholar] [CrossRef] [PubMed]
  61. Ingólfsson, H.I.; Melo, M.N.; van Eerden, F.J.; Arnarez, C.; Lopez, C.A.; Wassenaar, T.A.; Periole, X.; de Vries, A.H.; Tieleman, D.P.; Marrink, S.J. Lipid organization of the plasma membrane. J. Am. Chem. Soc. 2014, 136, 14554–14559. [Google Scholar] [CrossRef] [PubMed]
  62. Kitjanon, J.; Khuntawee, W.; Sutthibutpong, T.; Boonnoy, P.; Phongphanphanee, S.; Wong-ekkabut, J. Transferability of Polymer Chain Properties between Coarse-Grained and Atomistic Models of Natural Rubber Molecule Validated by Molecular Dynamics Simulations. J. Phys. Conf. Ser. 2017, 901, 012096. [Google Scholar] [CrossRef]
  63. Monticelli, L. On atomistic and coarse-grained models for C60 fullerene. J. Chem. Theory Comput. 2012, 8, 1370–1378. [Google Scholar] [CrossRef]
  64. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 19–25. [Google Scholar] [CrossRef] [Green Version]
  65. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101. [Google Scholar] [CrossRef] [Green Version]
  66. Bussi, G.; Zykova-Timan, T.; Parrinello, M. Isothermal-isobaric molecular dynamics using stochastic velocity rescaling. J. Chem. Phys. 2009, 130, 074101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Parrinello, M.; Rahman, A. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52, 7182–7190. [Google Scholar] [CrossRef]
  68. Wassenaar, T.A.; Ingólfsson, H.I.; Prieß, M.; Marrink, S.J.; Schäfer, L.V. Mixing MARTINI: Electrostatic coupling in hybrid atomistic-coarse-grained biomolecular simulations. J. Phys. Chem. B 2013, 117, 3516–3530. [Google Scholar] [CrossRef]
  69. Beu, T.A.; Ailenei, A.E.; Costinaş, R.I. Martini Force Field for Protonated Polyethyleneimine. J. Comput. Chem. 2020, 41, 349–361. [Google Scholar] [CrossRef] [PubMed]
  70. Wong-Ekkabut, J.; Karttunen, M. The good, the bad and the user in soft matter simulations. Biochim. Biophys. Acta Biomembr. 2016, 1858, 2529–2538. [Google Scholar] [CrossRef]
  71. Wong-Ekkabut, J.; Karttunen, M. Molecular dynamics simulation of water permeation through the alpha-hemolysin channel. J. Biol. Phys. 2016, 42, 133–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Boonnoy, P.; Karttunen, M.; Wong-Ekkabut, J. Does α-tocopherol flip-flop help to protect membranes against oxidation? J. Phys. Chem. B 2018, 122, 10362–10370. [Google Scholar] [CrossRef] [PubMed]
  73. Enkavi, G.; Javanainen, M.; Kulig, W.; Róg, T.; Vattulainen, I. Multiscale simulations of biological membranes: The challenge to understand biological phenomena in a living substance. Chem. Rev. 2019, 119, 5607–5774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  75. Buchholz, J.; Paul, W.; Varnik, F.; Binder, K. Cooling rate dependence of the glass transition temperature of polymer melts: Molecular dynamics study. J. Chem. Phys. 2002, 117, 7364–7372. [Google Scholar] [CrossRef]
  76. Glova, A.D.; Falkovich, S.G.; Dmitrienko, D.I.; Lyulin, A.V.; Larin, S.V.; Nazarychev, V.M.; Karttunen, M.; Lyulin, S.V. Scale-dependent miscibility of polylactide and polyhydroxybutyrate: Molecular dynamics simulations. Macromolecules 2018, 51, 552–563. [Google Scholar] [CrossRef] [Green Version]
  77. Sharma, P.; Roy, S.; Karimi-Varzaneh, H.A. Validation of force fields of rubber through glass-transition temperature calculation by microsecond atomic-scale molecular dynamics simulation. J. Phys. Chem. B 2016, 120, 1367–1379. [Google Scholar] [CrossRef]
  78. Van Gunsteren, W.F.; Weiner, P.K.; Wilkinson, A.J. Computer Simulation of Biomolecular Systems: Theoretical and Experimental Applications; Springer Science & Business Media: Berlin, Germany, 2013; Volume 3. [Google Scholar]
  79. Bennett, C.H. Efficient estimation of free energy differences from Monte Carlo data. J. Comput. Phys. 1976, 22, 245–268. [Google Scholar] [CrossRef]
  80. Fetters, L.J.; Lohse, D.J.; Graessley, W.W. Chain dimensions and entanglement spacings in dense macromolecular systems. J. Polym. Sci. Part B Polym. Phys. 1999, 37, 1023–1033. [Google Scholar] [CrossRef]
  81. Tobolsky, A.V.; Mark, H.F. Polymer Science and Materials; John Wiley and Sons: Hoboken, NJ, USA, 1971. [Google Scholar]
  82. Brandrup, J.; Immergut, E.H.; Grulke, E.A.; Abe, A.; Bloch, D.R. Polymer Handbook; John Wiley and Sons: Hoboken, NJ, USA, 1999. [Google Scholar]
  83. Holownia, B.P. Effect of carbon black on Poisson’s ratio of elastomers. Rubber Chem. Technol. 1975, 48, 246–253. [Google Scholar] [CrossRef]
  84. Guenza, M. Structural and thermodynamic consistency in coarse-grained models of macromolecules. J. Phys. Conf. Ser. 2015, 640, 012009. [Google Scholar] [CrossRef]
  85. Potestio, R.; Peter, C.; Kremer, K. Computer simulations of soft matter: Linking the scales. Entropy 2014, 16, 4199–4245. [Google Scholar] [CrossRef] [Green Version]
  86. Van der Wal, C.W.; Bree, H.W.; Schwarzl, F.R. Mechanical properties of highly filled elastomers. II. Relationship between filler characteristics, thermal expansion, and bulk moduli. J. Appl. Polym. Sci. 1965, 9, 2143–2166. [Google Scholar] [CrossRef]
  87. Wei, C.; Srivastava, D.; Cho, K. Thermal expansion and diffusion coefficients of carbon nanotube-polymer composites. Nano Lett. 2002, 2, 647–650. [Google Scholar] [CrossRef] [Green Version]
  88. Wang, S.; Tambraparni, M.; Qiu, J.; Tipton, J.; Dean, D. Thermal expansion of graphene composites. Macromolecules 2009, 42, 5251–5255. [Google Scholar] [CrossRef]
  89. Vettorel, T.; Kremer, K. Development of entanglements in a fully disentangled polymer melt. Macromol. Theory Simul. 2010, 19, 44–56. [Google Scholar] [CrossRef]
  90. Mcquarrie, D. Statistical Mechanics; Harper & Row: New York, NY, USA, 1976. [Google Scholar]
  91. Trzesniak, D.; Kunz, A.P.E.; van Gunsteren, W.F. A comparison of methods to compute the potential of mean force. Chemphyschem 2007, 8, 162–169. [Google Scholar] [CrossRef] [PubMed]
  92. Cao, Z.; Peng, Y.; Li, S.; Liu, L.; Yan, T. Molecular dynamics simulation of fullerene C60 in ethanol solution. J. Phys. Chem. C 2009, 113, 3096–3104. [Google Scholar] [CrossRef]
  93. Al-Qattan, M.N.; Deb, P.K.; Tekade, R.K. Molecular dynamics simulation strategies for designing carbon-nanotube-based targeted drug delivery. Drug Discov. Today 2018, 23, 235–250. [Google Scholar] [CrossRef]
  94. Nisoh, N.; Karttunen, M.; Monticelli, L.; Wong-Ekkabut, J. Lipid monolayer disruption caused by aggregated carbon nanoparticles. RSC Adv. 2015, 5, 11676–11685. [Google Scholar] [CrossRef]
  95. Chakrabarty, A.; Cagin, T. Coarse grain modeling of polyimide copolymers. Polymer 2010, 51, 2786–2794. [Google Scholar] [CrossRef]
  96. Sillescu, H. Heterogeneity at the glass transition: A review. J. Non -Cryst. Solids 1999, 243, 81–108. [Google Scholar] [CrossRef]
Figure 1. (a) Density, (b) bulk modulus (Equation (3)), (c) thermal expansion (Equation (1)), and (d) heat capacity (Equation (2)) of cis-PI and C60 composites as a function of C60 concentration.
Figure 1. (a) Density, (b) bulk modulus (Equation (3)), (c) thermal expansion (Equation (1)), and (d) heat capacity (Equation (2)) of cis-PI and C60 composites as a function of C60 concentration.
Polymers 13 04044 g001
Figure 2. Potential of mean force (PMF) of (a) C60-C60 and (b) C60-cis-PI in the composites at different C60 concentrations.
Figure 2. Potential of mean force (PMF) of (a) C60-C60 and (b) C60-cis-PI in the composites at different C60 concentrations.
Polymers 13 04044 g002
Figure 3. Snapshots of cis-PI-C60 composites at different C60 concentrations from 0 (melt) to 32 phr. Cyan and red represent cis-PI and C60, respectively.
Figure 3. Snapshots of cis-PI-C60 composites at different C60 concentrations from 0 (melt) to 32 phr. Cyan and red represent cis-PI and C60, respectively.
Polymers 13 04044 g003
Figure 4. Black line: the diffusion coefficients of rubber in melt (0 phr) and composites (phr ≠ 0). Red line: the diffusion coefficients of C60 in composites as a function of C60 concentration.
Figure 4. Black line: the diffusion coefficients of rubber in melt (0 phr) and composites (phr ≠ 0). Red line: the diffusion coefficients of C60 in composites as a function of C60 concentration.
Polymers 13 04044 g004
Figure 5. Solvation free energies of C60 in cis-PI melt (0 phr) and cis-PI-C60 composites at different C60 concentrations. The dashed line is a guide to the eye.
Figure 5. Solvation free energies of C60 in cis-PI melt (0 phr) and cis-PI-C60 composites at different C60 concentrations. The dashed line is a guide to the eye.
Polymers 13 04044 g005
Figure 6. Glass transition temperature (Tg) of cis-PI-C60 composites at different C60 concentrations.
Figure 6. Glass transition temperature (Tg) of cis-PI-C60 composites at different C60 concentrations.
Polymers 13 04044 g006
Table 1. Macroscopic and structural properties of cis-PI in melts. R0, Rg, and < R 0 2 > / < R g 2 > are the end-to-end distance, the radius of gyration, and polymer expansion factor, respectively.
Table 1. Macroscopic and structural properties of cis-PI in melts. R0, Rg, and < R 0 2 > / < R g 2 > are the end-to-end distance, the radius of gyration, and polymer expansion factor, respectively.
PropertiesExp. United Atom [38]CG
Density; kg/m3910 [80]853.9 ± 1.61087 ± 2.30
Bulk modulus; GPa2.02 [83]1.37 ± 0.020.53 ± 0.00
Thermal expansion; (10−4/K)6.1 [81,82]7.80 ± 0.159.25 ± 0.02
< R 0 2 >-12.85 ± 0.4113.28 ± 0.06
< R g 2 >-2.07 ± 0.042.17 ± 0.01
< R 0 2 > / < R g 2 > -6.206.11
Table 2. Average end-to-end distance (R0), radius of gyration (Rg) of cis-PI chains, and their polymer chain expansion factor ( < R 0 2 > / < R g 2 > ) as a function of C60 concentration.
Table 2. Average end-to-end distance (R0), radius of gyration (Rg) of cis-PI chains, and their polymer chain expansion factor ( < R 0 2 > / < R g 2 > ) as a function of C60 concentration.
C60
Concentration
(phr)
< R0 >
(nm)
< Rg >
(nm)
< R 0 2 > / < R g 2 >
03.64 ± 0.061.47 ± 0.016.11
43.65 ± 0.061.48 ± 0.016.11
83.66 ± 0.061.48 ± 0.016.12
163.67 ± 0.061.48 ± 0.016.12
323.71 ± 0.061.50 ± 0.016.15
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kitjanon, J.; Khuntawee, W.; Phongphanphanee, S.; Sutthibutpong, T.; Chattham, N.; Karttunen, M.; Wong-ekkabut, J. Nanocomposite of Fullerenes and Natural Rubbers: MARTINI Force Field Molecular Dynamics Simulations. Polymers 2021, 13, 4044. https://doi.org/10.3390/polym13224044

AMA Style

Kitjanon J, Khuntawee W, Phongphanphanee S, Sutthibutpong T, Chattham N, Karttunen M, Wong-ekkabut J. Nanocomposite of Fullerenes and Natural Rubbers: MARTINI Force Field Molecular Dynamics Simulations. Polymers. 2021; 13(22):4044. https://doi.org/10.3390/polym13224044

Chicago/Turabian Style

Kitjanon, Jiramate, Wasinee Khuntawee, Saree Phongphanphanee, Thana Sutthibutpong, Nattaporn Chattham, Mikko Karttunen, and Jirasak Wong-ekkabut. 2021. "Nanocomposite of Fullerenes and Natural Rubbers: MARTINI Force Field Molecular Dynamics Simulations" Polymers 13, no. 22: 4044. https://doi.org/10.3390/polym13224044

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop