Next Article in Journal
GC-MS Screening Analysis for the Identification of Potential Migrants in Plastic and Paper-Based Candy Wrappers
Previous Article in Journal
Influence of SiO2/TiO2 Nanocomposite on the Optoelectronic Properties of PFO/MEH-PPV-Based OLED Devices
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Micro-Structure and Thermomechanical Properties of Crosslinked Epoxy Composite Modified by Nano-SiO2: A Molecular Dynamics Simulation

1
State Key Laboratory of Alternate Electrical Power System with Renewable Energy Sources, North China Electric Power University, Baoding 071000, China
2
Department of Power Engineering, North China Electric Power University, Baoding 071003, China
3
State Key Laboratory of Power Transmission Equipment & System Security and New Technology, Chongqing University, Chongqing 400044, China
*
Author to whom correspondence should be addressed.
Polymers 2018, 10(7), 801; https://doi.org/10.3390/polym10070801
Submission received: 3 July 2018 / Revised: 19 July 2018 / Accepted: 19 July 2018 / Published: 20 July 2018

Abstract

:
Establishing the relationship among the composition, structure and property of the associated materials at the molecular level is of great significance to the rational design of high-performance electrical insulating Epoxy Resin (EP) and its composites. In this paper, the molecular models of pure Diglycidyl Ether of Bisphenol A resin/Methyltetrahydrophthalic Anhydride (DGEBA/MTHPA) and their nanocomposites containing nano-SiO2 with different particle sizes were constructed. The effects of nano-SiO2 dopants and the crosslinked structure on the micro-structure and thermomechanical properties were investigated using molecular dynamics simulations. The results show that the increase of crosslinking density enhances the thermal and mechanical properties of pure EP and EP nanocomposites. In addition, doping nano-SiO2 particles into EP can effectively improve the properties, as well, and the effectiveness is closely related to the particle size of nano-SiO2. Moreover, the results indicate that the glass transition temperature (Tg) value increases with the decreasing particle size. Compared with pure EP, the Tg value of the 6.5 Å composite model increases by 6.68%. On the contrary, the variation of the Coefficient of Thermal Expansion (CTE) in the glassy state demonstrates the opposite trend compared with Tg. The CTE of the 10 Å composite model is the lowest, which is 7.70% less than that of pure EP. The mechanical properties first increase and then decrease with the decreasing particle size. Both the Young’s modulus and shear modulus reach the maximum value at 7.6 Å, with noticeable increases by 12.60% and 8.72%, respectively compared to the pure EP. In addition, the thermal and mechanical properties are closely related to the Fraction of Free Volume (FFV) and Mean Squared Displacement (MSD). The crosslinking process and the nano-SiO2 doping reduce the FFV and MSD value in the model, resulting in better thermal and mechanical properties.

Graphical Abstract

1. Introduction

With the growing expansion of electrical equipment applications, the disadvantages of the Epoxy Resin (EP) insulating materials such as Diglycidyl Ether of Bisphenol A resin/Methyltetrahydrophthalic Anhydride (DGEBA/MTHPA) have become increasingly obvious. Some of the properties, including mechanical and thermal properties, are often not able to meet the ever-increasing requirements of electrical equipment [1,2,3,4]; while, the emergence of nanomaterials and the development of nanotechnology in recent years have provided new ways to improve the comprehensive properties of EP [5,6]. Compared with traditional materials, the distinguishing structure of nanomaterial has provided a set of unique features, such as the small size effect, quantum size effect and surface effect [7,8], which further enhance the performance in various aspects of physics and chemistry. The nanocomposites obtained by combining the nanofillers with the polymer possess the advantages of both the nanomaterial and the polymer itself. Numerous scholars have carried out experimental studies on the improvement of EP thermal properties, mechanical properties, electrical properties, etc. [9,10]. The studies emphasized the effects of filler type, particle size, shape and dosage on the properties of the composite EP [11,12]. Despite the progress made by the above researchers, there is no unified understanding of the interface mechanism of EP nanocomposites. Therefore, it is significant to further conduct research on the relationship between the structure and properties of EP nanocomposite, as well as investigating the underlying mechanism of action.
In recent years, molecular simulation technologies have continued to progress and develop, which, at the molecular level, possess obvious advantages in the study of the microstructure and macro-properties of composite polymer materials [13,14,15]. In particularly, the Molecular Dynamics (MD) method has been applied to the study of EP by various scholars [16,17,18,19,20]. In terms of EP composites, some scholars have studied the glass transition and thermoelasticity of SiC/EP nanocomposites by molecular simulation. It is found that the glass transition temperature (Tg) and the Coefficient of Thermal Expansion (CTE) of the composites is lower than that of the pure EP [21]. Chen Shenghui et al. [22] have constructed an uncrosslinked DGEBA and Isophorone Diamine (DGEBA/IPD) epoxy-carbon nanofiber composite interface model, analyzed the distribution, structure and movement characteristics of the monomer molecules and found that carbon nanofibers can significantly alter the concentration and properties of the EP and curing agent during the interfacial phase. Zhang et al. [20] investigated the influence of silane coupling agent on the structure and thermomechanical properties of the nanocomposites through MD simulation and found that incorporating silica nanoparticles into the EP matrix can significantly improve the mechanical and thermal properties of the composites. Furthermore, the thermomechanical properties were further enhanced by silane coupling agent modification on the surface of fillers. R. Zhu et al. [23] carried out MD research on EPON-862 EP (Bisphenol F epoxy resin) reinforced by the carbon nanotube. It was found that the carbon nanotubes can enhance the mechanical properties of the EP. Moreover, with the increase of the aspect ratio, the enhancement effect is more significant. In summary, MD has become an essential instrument in terms of analyzing the performance of composite materials, which is also expected to provide new approaches for the design and preparation of EP composites with enhanced performance.
SiO2 is one of the most common fillers for the EP insulating materials. The EP composite filled with nano-SiO2 has improved mechanical and thermal properties [7,24,25]. Although scholars have studied the properties of SiO2/EP nanocomposites through experiments [26,27,28,29,30], few involved MD simulations.
In this paper, MD simulations were performed on DGEBA/MTHPA EP filled with nano-SiO2. The SiO2 fillers were spherical, with particle radiuses of 6.5 Å, 7.6 Å, 8.8 Å and 10 Å, respectively. Based on the MD simulations, the parameters of the models, including the Fraction of Free Volume (FFV), Mean Square Displacement (MSD), Tg, CTE and mechanical properties, were analyzed. The effects of crosslinking density and nano-SiO2 particle size on the microstructure, thermal and mechanical properties of EP were investigated. In addition, the relationship between the structure and properties of composite EP was analyzed. Furthermore, this paper provided theoretical guidance for the design and preparation of EP nanocomposites with enhanced performance.

2. Model Construction and Simulation Details

In this section, the simulation steps are synthesized as the following three parts: (1) mix DGEBA and MTHPA according to the actual proportion, and fill the models with nano-SiO2; (2) construct the amorphous models of the crosslinked EP/SiO2 nanocomposites based on the practical crosslinking reaction mechanism; (3) perform MD simulations of DGEBA/MTHPA/SiO2 crosslinked models with different crosslinking densities and filler sizes. The models mentioned in this paper were built using Material Studio (developed by Accelrys).

2.1. Reaction Mechanism of EP and Curing Agent

Due to the existence of the numerous hydroxyl groups on the surface of nano-SiO2, nanofillers were also considered during the crosslinking reaction. The reaction mechanism is shown in Figure 1. By repeating the process in Figure 1c,d, the crosslinking reaction was carried out.
Here, the crosslinking reaction between DGEBA, MTHPA and SiO2 was realized through programming. To simplify the process, the following assumptions have been made: (1) The crosslinking density is defined as the ratio of the reacted epoxy groups to the initial epoxy groups. (2) To start the crosslinking process, we introduce the partial product of (a) during the model construction stage (i.e., ignoring Reaction (a) in the crosslinking process) as a crosslinking primer (denoted by DGEBA-MTHPA). As a result, the crosslinking density of the initial model is 10%. (3) The activity of each reaction group is the same. (4) The reaction is diffusion controlled. (5) The reactions are synchronized.

2.2. Construction of Crosslinked EP/SiO2 Composite Models

Epoxy prepolymer is regarded as a mixture of DGBEA with a polymerization degree of zero and one. In this paper, the polymerization degree of DGEBA was set to zero in the modeling process to simplify the composite model [23]. The modeling process of crosslinked EP/SiO2 composite is as follows:
(1)
Construct the molecular models of DGEBA, MTHPA, DGEBA-MTHPA and spherical nano-SiO2 with radiuses of 6.5 Å, 7.6 Å, 8.8 Å and 10 Å, respectively. Simulate the oxidation reaction of SiO2, so that the surface of fillers has massive hydroxyl groups [31,32]. Subsequently, optimize the structure of the above models and obtain the molecular models as shown in Figure 2.
(2)
Considering that the actual molar ratio of DGEBA and MTHPA is about 1:2 [20,33], we constructed amorphous molecular models of pure EP and EP nanocomposite with different filler sizes. Using periodic boundary conditions, the initial temperature of the model was set to be 580 K to facilitate the subsequent annealing process. Each nanocomposite model was filled with a nano-SiO2 particle to simulate the well-dispersed situation. The mass fraction of SiO2 in each system was maintained at 6.5% by adjusting the amount of the ingredients according to Table 1. Further, 10 amorphous molecular models with an initial crosslinking density of 10% were constructed under each system, and their geometric structures were optimized based on the minimum energy principle. The amorphous models with the lowest energy were selected for further simulation and calculation.
(3)
Perform a 200-ps MD simulation on each model under the NPT ensemble (the constant temperature and pressure ensemble). The specific details of the MD process are as follows. The time step was 1 fs. The temperature and pressure were controlled at 580 K and 1.0 × 10−4 GPa (which is one standard atmospheric pressure) respectively by applying the Andersen and Berendsen methodology [20,34]. In addition, the van der Waals interaction and electrostatic interaction were calculated by the atom-based and Eward method, respectively.
(4)
The scriptlet employed for the crosslinking reaction was written according to Section 2.1. The crosslinking was conducted in stages, and the molecular models with different crosslinking densities were obtained by conducting crosslinking reactions multiple times.

2.3. MD and Annealing Simulation

Molecular models with crosslinking densities of 10%, 34%, 67% and 96% were further selected for MD processing. First, MD simulations of 200 ps were performed under the NVT ensemble (constant volume and temperature ensemble) and NPT ensemble successively, in order to eliminate the stress generated during the crosslinking process. The time step was 1 fs, and the temperature was 580 K. In addition, the pressure in the NPT ensemble was controlled at 1.0 × 10−4 GPa.
Subsequently, anneal simulations were performed on the above models from 580 K–280 K to extract the information of density, volume and temperature for thermal property analysis. The annealing rate was set to be 10 K/100 ps, i.e., an NPT ensemble MD simulation of 100 ps was performed after each temperature decrease by 10 K. The models of EP composite at 300 K are shown in Figure 3.

3. Simulation Results and Model Parameters

3.1. Free Volume

The free volume theory [35] holds that the volume of a liquid or solid substance, VT, consists of two parts, i.e., the volume occupied by the molecule denoted as Vo, and the free volume Vf, where:
V T = V o + V f
Vf is the intermolecular space, which disperses in the form of a hole among the materials and provides space for molecule movements. It also enables the movement of the molecular chain. The influence of nanoparticles on the free volume of the polymer was relatively complex in the following two aspects. First, the nanoparticle occupied a considerable part of the volume of the system and increased the distance between chain segments, which hindered the movement and stacking of molecular segments and reduced the free volume. Moreover, the nanoparticle can participate in the crosslinking reaction of the EP system and fix molecular chain segments, therefore expanding the crosslinked network and increasing the free volume. The above impacts existed in all the EP nanocomposite materials. Therefore, the increase or decrease of the free volume significantly depended on the properties of the nanoparticles.
The volumes of different models differed greatly due to the different numbers of molecules. As a result, the free volumes were different, as well. Therefore, to facilitate the comparison of the free volume characteristics of different systems, the fraction of free volume (FFV) was introduced to characterize the relative size of the free volume. The expression is [35]:
F F V = V f V 0 + V f × 100 %
Here, the FFV of EP/nano-SiO2 composites with different particle sizes were calculated at 300 K. The variation trend of the FFV with the crosslinking density is shown in Figure 4.
According to Figure 4, with the increase of crosslinking density, the FFVs of pure EP and composites decrease first and then increased. The reason for the above trend is as follows. In the initial stage of the crosslinking reaction, the free volume in the crosslinking network was occupied by a dangling chain appearing [36,37]; therefore, the FFV of the system decreased slightly. Subsequently, the formation of a crosslinking network and the decrease of the dangling segment led to the increase of the free volume. On the other hand, the FFV of the nanocomposites was lower than that of the pure EP. The underlying cause is that the nanoparticles occupied a considerable part of the space in the system, resulting in a smaller range of segmental motion and a decrease in the FFV. Furthermore, the FFV decreased with the decreasing particle size. This was due to the final crosslinked network being generated by the covalent links between SiO2 and EP matrix in the system, and the interaction tended to enhance as the particle size decreased. Previous studies have suggested that the mechanical properties of the polymers depended on the free volume to a certain extent. The smaller the free volume was, the better the performance could be [35,38]. Therefore, it can be inferred that the mechanical properties of the system containing fillers were better than that of the pure EP system. Moreover, the smaller the particle size was, the better the mechanical property was when the crosslinking density was relatively high.

3.2. Segment Motion

The molecules of the models were constantly moving throughout the MD process. Some research has proven that there is a correlation between the strength of the segment motion and the mechanical properties of the polymer. Specifically, the enhancement of molecular segment motion may reduce the mechanical modulus of EP [39]. Moreover, the mean squared displacement (MSD) is defined as a microscopic parameter to characterize the motion capability of each atom or molecular segment in the system, which is defined as the mean squared displacement of the molecules or segments, respectively. In general, the slope of the MSD curve indicates the strength of the polymer molecule segment motion. The MSD in a system containing N atoms can be described as follows [38]:
M S D = 1 3 N i = 0 N 1 [ | ( R i ) ( t ) ( R i ) ( 0 ) | 2 ]
where R i ( t ) and R i ( 0 ) denote the displacement vector of any atom i at time t and the initial time in the system, respectively. In this paper, the MSD values of systems with different crosslinking densities, particle sizes and temperature in the first 30 ps of MD simulations under the NPT ensemble were investigated.
The MSD values of the 6.5 Å system under different crosslinking densities are shown in Figure 5a. In addition, the trends of MSD and crosslinking density in other systems were similar. As the crosslinking density increased, the MSD values of EP and its nanocomposites gradually decreased, which indicated that the crosslinked structure limited the molecular motion in the model. To analyze the effect of particle size on the segment motion, the MSD values of different particle size models were normalized based on the number of atoms in the pure EP system, as suggested in Table 1. The processed MSD values are shown in Figure 5c. The segment motion was negatively correlated with the particle size. It can be seen that nano-SiO2 dopants can significantly limit the movement of the molecular segment in the material, and the restriction will be more pronounced as the particle size increases. In addition, temperature had a significant effect on segmental motion. Particularly, temperature had an apparent effect on the segmental motion of EP and its composites, as well. With the increasing of temperature, MSD values increased continuously, as shown in Figure 5b. This tendency was due to the thermal motion of the molecules. As the temperature increased, the thermal motion of the molecule increased, which caused the increasing of MSD. Furthermore, in the temperature range of 400 K–410 K, the variation of MSD values was the most conspicuous, which was related to the glass transition of the materials.

3.3. Glass Transition Temperature

The glass transition temperature (Tg) refers to the temperature for the transition from the glassy state to the rubbery state of the amorphous polymer or vice versa, which is a critical thermal performance index of EP materials. When EP transitions from the glassy state to the rubbery state at temperature Tg, its performance completely changes due to the state conversion. Therefore, obtaining the accurate value of Tg is an important prerequisite for thermal performance comparisons across different systems. In this paper, the density-temperature linear fitting and MSD curve method were used to accurately predict the Tg of each system.
1. Density-temperature linear fitting method:
As the temperature increased, the density of the EP showed a linear decrease. However, the density decrease rates in the glassy and rubbery state were different [40,41]. We plotted the density-temperature scatter diagram based on the temperature and density extracted during the annealing process. Figure 6 presents the density-temperature curve of the 67% crosslinked pure EP, where there is an obvious inflection point in each scatter plot. The scatter points on the two sides of the inflection point are linearly fitted, and the turning point of the two lines is the Tg value of the material.
2. MSD curve method:
Studies have shown that as the temperature increases, the MSD gradually increases. When transitioning from the glassy state to the rubbery state, the motion state of the molecules in the system will change drastically. This mutation reflecting on the MSD curve was a sudden jump of the MSD values at the temperature interval above and lower than Tg. Therefore, by investigating the MSD characteristics of the polymers at different temperatures and observing the temperature range of these curves, the temperature range of Tg can be predicted [42].
The MSD of the pure EP and EP nanocomposite systems at various temperatures was calculated at 67% crosslinking density. To stimulate the process, the temperature interval in this paper was set to be 10 K. As shown in Figure 5b, taking the 6.5 Å EP composite system as an example, there was a wider gap between the 400 K and 410 K MSD curves. Therefore, we concluded that the Tg value of the 6.5 Å EP composite system was in the (400 K, 410 K) interval.
Tg values obtained by the density-temperature linear fitting method and the Tg intervals obtained by the MSD curve method are summarized in Table 2. The results of the two methods were highly consistent with each other. In addition, the Tg value of the pure EP with 67% crosslinking density was 382.68 K, which was highly close to the actual Tg value of DGEBA/MTHPA epoxy [43]. It can be assumed that 67% was close to the industrial crosslinking density of the pure DGEBA/MTHPA system.
The Tg values of EP and its composites for different particle sizes and crosslinking densities are summarized in Figure 7. According to Figure 7a, the Tg value of pure EP and EP composite with 6.5 Å nano-SiO2 increased as the crosslinking reaction went on. Moreover, when the crosslinking density was greater than 34%, the increase rate of the Tg value with crosslinking density was higher than that in the initial stage of the crosslinking process. Therefore, increasing the crosslinking density can significantly improve the Tg value of EP and its composites. According to Figure 5a, the segmental motion in the model was limited by the crosslinked structure, and the flexibility of segments decreased. Therefore, it is difficult to alter the conformation of the models. The higher the crosslinking density, the stronger the effect was and, ultimately, the higher the Tg value.
Figure 7b shows the Tg value of pure EP and EP composite with different particle sizes at 67% crosslinking density. Nano-SiO2 can significantly improve the Tg value of EP composites. Particularly, the Tg value of composites doped with 6.5 Å SiO2 increased by 6.68%. Analysis suggested that the excellent thermal properties of nano-SiO2 can increase the Tg value of the composites. On the other hand, the inorganic-organic interface integration between the filler and the EP matrix was an important factor to increase the Tg value of the material. With the decrease of the particle size, the small size effect of nanoparticles was more intense, and the interface effect was strengthened. Therefore, the interfacial bonding between the nano-SiO2 and the matrix was stronger, resulting in higher Tg values.

3.4. Coefficient of Thermal Expansion

Coefficient of Thermal Expansion (CTE) refers to the ratio of the increment in the length of the unit temperature and its length at 300 K, which is a critical index factor to measure the thermal stability of materials. The calculation formula is as follows [21]:
CTE = 1 V 0 ( V T ) P
where V0 is the volume at 300 K and P is taken as the standard atmospheric pressure. The relationship between the volume and temperature of different EP systems was linearly fitted to obtain the values of V / T .
Taking pure EP and the 6.5 Å composite as examples, the CTE at the glassy state and rubbery state obtained by Equation (4) are shown in Figure 8. According to Figure 8, the CTE of the two states decreased significantly with the increase of the crosslinking density. Specifically, when the crosslinking density increased from 10%–96%, the CTE of pure EP and the 6.5 Å composites in the glassy state decreased by 16.95% and 16.99%, respectively. Correspondingly, those in the rubber state decreased by 27.22% and 32.97%. This was attributed to the 3D network structure formed in the models during the crosslinking reaction, resulting in a reduced flexibility of the molecular chain and bound motion of polymer segments. However, the CTE of EP composites was lower than that of pure EP because of the extremely low CTE of the SiO2 crystal itself. In addition, the CTE in the glassy state was smaller than that in the rubbery state due to the different states of free volume, which is consistent with [20,21,32]. In the glassy state, the free volume of the materials was in the frozen state, and the expansion of the occupied volume was the only dominant factor affecting CTE; while, in the rubbery state, both the free volume and the occupied volume contributed to the thermal expansion, which led to the higher CTE in the glassy state.
On the other hand, the CTE in the glassy and rubbery states showed different trends with the variation of the particle size. As shown in Figure 9, with the increase of the particle size of nano-SiO2, the CTE in the glassy state decreased gradually. In the glassy state, the CTE of the 10 Å model was reduced by 5.70% compared to that of the 6.5 Å system and 7.70% compared to that of the pure EP; while the CTE in the rubbery state increased with the particle size. In general, the CTE in the glassy state provided a better referential value considering the fact that the actual working state of the EP materials was the glassy state.

3.5. Elastic Moduli

The static constant strain method [44] was used to analyze the mechanical properties of the EP systems in this paper. A slight strain was applied to the system, which was originally in the state of mechanical equilibrium, to enable the model to generate the uniaxial tension and compression deformation along the x-, y- and z-axis, respectively, i.e., shear deformation in the xy-, xz- and yz-plane, correspondingly. The stress-strain relationship obeys Hooke’s law:
σ = C ε
where σ is the stress vector, ε is strain vector and C is the stiffness matrix. The EP models in the simulation can be regarded as an isotropic material; thus, C can be simplified as follows:
[ λ + 2 μ 0 0 0 0 0 0 λ + 2 μ 0 0 0 0 0 0 λ + 2 μ 0 0 0 0 0 0 μ 0 0 0 0 0 0 μ 0 0 0 0 0 0 μ ]
where λ and μ are elastic constants and can be obtained from the stiffness matrix. The parameters such as Young’s modulus E and shear modulus G of an EP system can be obtained from λ and μ:
E = μ 3 λ + 2 μ λ + μ
G = μ
The static mechanical properties of the pure EP and composites were investigated at 300 K. Figure 10 shows Young’s modulus and the shear modulus of pure EP and 6.5 Å composites. With the increase of the crosslinking density, the two modulus values increased for all systems. The reason is that the stable 3D polymer network structure formed by the curing agent, EP matrix and nano-SiO2 further enhanced the stiffness and mechanical properties of the material. Furthermore, according to the results shown in Figure 8, the crosslinked structure lowered MSD and limited the molecular motion in the model, which may result in higher moduli, as well [38]. Moreover, the static mechanical properties of the nanocomposite systems were obviously better than that of the pure EP system, and the modulus growth rates with the crosslinking density of the composites were higher than that of pure EP. The main reason is the high module of nanoparticles, strengthening the epoxy matrix [7].
The Young’s and shear moduli of the pure EP and composites with different particle sizes at a crosslinking density of 67% are shown in Figure 11. The doping of nano-SiO2 with different particle sizes improved the mechanical properties of the DGEBA/MTHPA matrix to varying degrees. Among them, the 7.6 Å SiO2 had the best performance, where Young’s modulus and the shear modulus increased by 12.60% and 8.72%, respectively. The phenomena were basically consistent with the results of free volume analysis. The smaller the particle size was, the smaller the free volume was, the higher the moduli could be. However, when the particle size of SiO2 was less than 7.6 Å, the Young’s modulus and the shear modulus decreased slightly, which indicated that there were many factors affecting the mechanical properties of the nanocomposites. The interface characteristics produced by different particle sizes may be another factor affecting the mechanical properties, except the free volume.

4. Conclusions

In this paper, an automatic crosslinking method for epoxy resin, anhydride curing agent and nano-SiO2 was developed, which effectively improved the accuracy of modeling and the molecular simulation efficiency of epoxy composites. The microstructure, thermal and mechanical properties of epoxy resin and its composites were studied by molecular dynamics. The results show that the crosslinking densities of epoxy resin and the doping of nano-SiO2 can influence the structure and properties of the materials in many ways. The obtained findings are as follows:
  • The thermal and mechanical properties were improved by increasing the crosslinking density. With the increase of crosslinking density, the glass transition temperature (Tg) increased and the coefficient of thermal expansion (CTE) decreased; in addition, Young’s modulus and the shear modulus of the materials increased, and the mechanical properties were enhanced.
  • Doping nano-SiO2 particles into epoxy resin effectively improved the thermal and mechanical properties, and the effectiveness was closely related to the particle size of nano-SiO2. The Tg value increased with the decreasing particle size. Compared with pure epoxy resin, the Tg of 6.5 Å composite model was increased by 6.68%. The variation of CTE in the glassy state demonstrated opposite trend as compared with Tg value. The CTE of 10 Å composite model was the lowest, which is 7.70% less than that of pure epoxy. In addition, the mechanical properties first increased and then decreased with the decreasing particle size. Both the Young’s modulus and shear modulus reached the maximum at the 7.6 Å, and increased by 12.60% and 8.72% respectively compared with pure epoxy.
  • The thermal and mechanical properties are closely related to the Fraction of Free Volume (FFV) and Mean Squared Displacement (MSD). The crosslinking process and the nano-SiO2 doping reduced the FFV in the model, impeded the deformation and improved the elastic modulus of the system. Moreover, the decreasing FFV reduced the MSD of the model, limited the segment motion of the molecular chains and made it even harder for the glass transition.

Author Contributions

Q.X., X.Y., F.L. and K.F. conceived of and designed the simulations. L.L. and S.L. performed the MD simulations. Q.X. and K.F. analyzed the data. Z.H. contributed the analysis tools. L.L. and B.L. wrote the paper.

Funding

This research was funded by the National Key R&D Program of China (Grant No. 2017YFB0903800), the Fundamental Research Funds for the Central Universities (Grant No. 2016ZZD07), and the Technology Foundation of SGCC (PGKJ2018-151).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xie, Q.; Lin, H.F.; Zhang, S.; Wang, R.X.; Kong, F.; Shao, T. Deposition of SICxHyOz thin film on epoxy resin by nanosecond pulsed appj for improving the surface insulating performance. Plasma Sci. Technol. 2018, 20. [Google Scholar] [CrossRef]
  2. He, X.; Xu, X.; Wan, Q.; Bo, G.; Yan, Y. Synthesis and characterization of dimmer-acid-based nonisocyanate polyurethane and epoxy resin composite. Polymers 2017, 9, 649. [Google Scholar] [CrossRef]
  3. Kumar, S.; Krishnan, S.; Samal, S.K.; Mohanty, S.; Nayak, S.K. Toughening of petroleum based (DGEBA) epoxy resins with various renewable resources based flexible chains for high performance applications: A review. Ind. Eng. Chem. Res. 2018, 57, 2711–2726. [Google Scholar] [CrossRef]
  4. Zhou, X.; Qiu, S.; Xing, W.; Gangireddy, C.S.R.; Gui, Z.; Hu, Y. Hierarchical polyphosphazene@molybdenum disulfide hybrid structure for enhancing the flame retardancy and mechanical property of epoxy resins. ACS Appl. Mater. Interfaces 2017, 9, 29147–29156. [Google Scholar] [CrossRef] [PubMed]
  5. Bobby, S.; Samad, M.A. Enhancement of tribological performance of epoxy bulk composites and composite coatings using micro/nano fillers: A review. Polym. Adv. Technol. 2017, 28, 633–644. [Google Scholar] [CrossRef]
  6. Zabihi, O.; Ahmadi, M.; Nikafshar, S.; Chandrakumar Preyeswary, K.; Naebe, M. A technical review on epoxy-clay nanocomposites: Structure, properties, and their applications in fiber reinforced composites. Compos. Part B Eng. 2018, 135, 1–24. [Google Scholar] [CrossRef]
  7. Nikje, M.M.A.; Garmarudi, A.B.; Tehrani, Z.M.; Haghshenas, M.; Shakhesi, S. Thermal and mechanical evaluation of epoxy resin composites by synthesis of amine-based coupling agent-nano silica complex. Polym.-Plast. Technol. 2011, 50, 646–650. [Google Scholar] [CrossRef]
  8. Faridirad, F.; Ahmadi, S.; Barmar, M. Polyamide/carbon nanoparticles nanocomposites: A review. Polym. Eng. Sci. 2017, 57, 475–494. [Google Scholar] [CrossRef]
  9. Hoyos, M.; García, N.; Navarro, R.; Dardano, A.; Ratto, A.; Guastavino, F.; Tiemblo, P. Electrical strength in ramp voltage ac tests of ldpe and its nanocomposites with silica and fibrous and laminar silicates. J. Polym. Sci. Pol. Phys. 2008, 46, 1301–1311. [Google Scholar] [CrossRef]
  10. Njuguna, J.; Pielichowski, K.; Desai, S. Nanofiller-reinforced polymer nanocomposites. Polym. Adv. Technol. 2008, 19, 947–959. [Google Scholar] [CrossRef] [Green Version]
  11. Shao, T.; Liu, F.; Hai, B.; Ma, Y.F.; Wang, R.X.; Ren, C.Y. Surface modification of epoxy using an atmospheric pressure dielectric barrier discharge to accelerate surface charge dissipation. IEEE Trans. Dielectr. Electr. Insul. 2017, 24, 1557–1565. [Google Scholar] [CrossRef]
  12. Mo, H.; Huang, X.; Liu, F.; Yang, K.; Li, S.; Jiang, P. Nanostructured electrical insulating epoxy thermosets with high thermal conductivity, high thermal stability, high glass transition temperatures and excellent dielectric properties. IEEE Trans. Dielectr. Electr. Insul. 2015, 22, 906–915. [Google Scholar] [CrossRef]
  13. Yang, L.; Qi, C.; Wu, G.; Liao, R.; Wang, Q.; Gong, C.; Gao, J. Molecular dynamics simulation of diffusion behaviour of gas molecules within oil–paper insulation system. Mol. Simul. 2013, 39, 988–999. [Google Scholar] [CrossRef]
  14. Verma, A.; Parashar, A.; Packirisamy, M. Atomistic modeling of graphene/hexagonal boron nitride polymer nanocomposites: A review. WIREs Comput. Mol. Sci. 2018, 8, e1346. [Google Scholar] [CrossRef]
  15. Paquet, E.; Viktor, H.L. Molecular dynamics, monte carlo simulations, and langevin dynamics: A computational review. BioMed Res. Int. 2015, 2015, 183918. [Google Scholar] [CrossRef] [PubMed]
  16. Masoumi, S.; Arab, B.; Valipour, H. A study of thermo-mechanical properties of the cross-linked epoxy: An atomistic simulation. Polymer 2015, 70, 351–360. [Google Scholar] [CrossRef]
  17. Li, K.; Li, Y.; Lian, Q.; Cheng, J.; Zhang, J. Influence of cross-linking density on the structure and properties of the interphase within supported ultrathin epoxy films. J. Mater. Sci. 2016, 51, 9019–9030. [Google Scholar] [CrossRef]
  18. Odegard, G.M.; Jensen, B.D.; Gowtham, S.; Wu, J.; He, J.; Zhang, Z. Predicting mechanical response of crosslinked epoxy using reaxff. Chem. Phys. Lett. 2014, 591, 175–178. [Google Scholar] [CrossRef]
  19. Sirk, T.W.; Khare, K.S.; Karim, M.; Lenhart, J.L.; Andzelm, J.W.; McKenna, G.B.; Khare, R. High strain rate mechanical properties of a cross-linked epoxy across the glass transition. Polymer 2013, 54, 7048–7057. [Google Scholar] [CrossRef]
  20. Zhang, X.; Wen, H.; Wu, Y. Computational thermomechanical properties of silica–epoxy nanocomposites by molecular dynamic simulation. Polymers 2017, 9, 430. [Google Scholar] [CrossRef]
  21. Choi, J.; Yu, S.; Yang, S.; Cho, M. The glass transition and thermoelastic behavior of epoxy-based nanocomposites: A molecular dynamics study. Polymer 2011, 52, 5197–5203. [Google Scholar] [CrossRef]
  22. Chen, S.; Sun, S.; Steven, G.; Li, C.; Wang, X.; Hu, S. Molecular dynamics simulations of the interaction between carbon nanofiber and epoxy resin monomers. Acta Polym. Sin. 2015, 1158–1164. (In Chinese) [Google Scholar] [CrossRef]
  23. Zhu, R.; Pan, E.; Roy, A.K. Molecular dynamics study of the stress–strain behavior of carbon-nanotube reinforced epon 862 composites. Mater. Sci. Eng. A Struct. 2007, 447, 51–57. [Google Scholar] [CrossRef]
  24. Zheng, Y.; Chonung, K.; Wang, G.; Wei, P.; Jiang, P. Epoxy/nano-silica composites: Curing kinetics, glass transition temperatures, dielectric, and thermal-mechanical performances. J. Appl. Polym. Sci. 2009, 111, 917–927. [Google Scholar] [CrossRef]
  25. Li, W.; Feng, W.; Huang, H. High-performance epoxy resin/silica coated flake graphite composites for thermal conductivity and electrical insulation. J. Mater. Sci. Mater. Electron. 2016, 27, 6364–6370. [Google Scholar] [CrossRef]
  26. Hsieh, T.H.; Kinloch, A.J.; Masania, K.; Taylor, A.C.; Sprenger, S. The mechanisms and mechanics of the toughening of epoxy polymers modified with silica nanoparticles. Polymer 2010, 51, 6284–6294. [Google Scholar] [CrossRef] [Green Version]
  27. Liu, F.; Wang, Z.Q.; Liang, W.Y.; Qu, Y.W. Effect of nano-silica on flexural properties of epoxy resin. Adv. Mater. Res. 2011, 383–390, 3845–3848. [Google Scholar] [CrossRef]
  28. Lee, D.H.; Lee, N.; Park, H. Role of silica nanoparticle in multi-component epoxy composites for electrical insulation with high thermal conductivity. J. Am. Ceram. Soc. 2018, 101, 2450–2458. [Google Scholar] [CrossRef]
  29. Park, J.-J.; Lee, J.-Y. Effect of epoxy-modified silicone-treated micro-/nano-silicas on the electrical breakdown strength of epoxy/silica composites. IEEE Trans. Dielectr. Electr. Insul. 2017, 24, 3794–3800. [Google Scholar] [CrossRef]
  30. Li, Y.; Tian, M.; Lei, Z.; Zhang, J. Effect of nano-silica on dielectric properties and space charge behavior of epoxy resin under temperature gradient. J. Phys. D Appl. Phys. 2018, 51, 12530912. [Google Scholar] [CrossRef]
  31. He, J.M.; Huang, Y.D. Effect of silane-coupling agents on interfacial properties of CF/PI composites. J. Appl. Polym. Sci. 2007, 106, 2231–2237. [Google Scholar] [CrossRef]
  32. Zhang, X.; Chen, X.; Xiao, S.; Wen, H.; Wu, Y. Molecular Dynamics Simulation of Thermal-mechanical Properties of Modified SiO2 Reinforced Epoxy Resin. High Volt. Eng. 2018, 44, 740–749. (In Chinese) [Google Scholar] [CrossRef]
  33. Wu, C.; Xu, W. Atomistic molecular modelling of crosslinked epoxy resin. Polymer 2006, 47, 6004–6009. [Google Scholar] [CrossRef]
  34. Lu, Y.-Y.; Shu, Y.-J.; Liu, N.; Shu, Y.; Wang, K.; Wu, Z.-K.; Wang, X.-C.; Ding, X.-Y. Theoretical simulations on the glass transition temperatures and mechanical properties of modified glycidyl azide polymer. Comput. Mater. Sci. 2017, 139, 132–139. [Google Scholar] [CrossRef]
  35. Wei, Q.; Zhang, Y.; Wang, Y.; Yang, M. A molecular dynamic simulation method to elucidate the interaction mechanism of nano-SiO2 in polymer blends. J. Mater. Sci. 2017, 52, 12889–12901. [Google Scholar] [CrossRef]
  36. Huang, Y.Z.; Tian, Y.Z.; Li, Y.Y.; Tan, X.C.; Li, Q.; Cheng, J.; Zhang, J.Y. High mechanical properties of epoxy networks with dangling chains and tunable microphase separation structure. RSC Adv. 2017, 7, 49074–49082. [Google Scholar] [CrossRef] [Green Version]
  37. Kojio, K.; Furukawa, M.; Matsumura, S.; Motokucho, S.; Osajima, T.; Yoshinaga, K. The effect of cross-linking density and dangling chains on surface molecular mobility of network polyurethanes. Polym. Chem. UK 2012, 3, 2287–2292. [Google Scholar] [CrossRef]
  38. Asaad, J.; Gomaa, E.; Bishay, I.K. Free-volume properties of epoxy composites and its relation to macrostructure properties. Mater. Sci. Eng. A Struct. 2008, 490, 151–156. [Google Scholar] [CrossRef]
  39. Zhang, W.; Qing, Y.; Zhong, W.; Sui, G.; Yang, X. Mechanism of modulus improvement for epoxy resin matrices: A molecular dynamics simulation. React. Funct. Polym. 2017, 111, 60–67. [Google Scholar] [CrossRef]
  40. Wang, Y.-H.; Wang, W.-H.; Zhang, Z.; Xu, L.; Li, P. Study of the glass transition temperature and the mechanical properties of pet/modified silica nanocomposite by molecular dynamics simulation. Eur. Polym. J. 2016, 75, 36–45. [Google Scholar] [CrossRef]
  41. Chinkanjanarot, S.; Radue, M.S.; Gowtham, S.; Tomasi, J.M.; Klimek-McDonald, D.R.; King, J.A.; Odegard, G.M. Multiscale thermal modeling of cured cycloaliphatic epoxy/carbon fiber composites. J. Appl. Polym. Sci. 2018, 135. [Google Scholar] [CrossRef]
  42. Xue, Q.; Lv, C.; Shan, M.; Zhang, H.; Ling, C.; Zhou, X.; Jiao, Z. Glass transition temperature of functionalized graphene–polymer composites. Comput. Mater. Sci. 2013, 71, 66–71. [Google Scholar] [CrossRef]
  43. Lin, S.; Huang, Y.; Xie, D.; Min, D.; Wang, W.; Yang, L.; Li, S. Molecular relaxation and glass transition properties of epoxy resin at high temperature. Acta Phys. Sin. 2016, 65, 304–310. (In Chinese) [Google Scholar] [CrossRef]
  44. Shokuhfar, A.; Arab, B. The effect of cross linking density on the mechanical properties and structure of the epoxy polymers: Molecular dynamics simulation. J. Mol. Model. 2013, 19, 3719–3731. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanism of crosslinking reaction between the epoxy, acid anhydride curing agent and nano-SiO2. (a) Under the effect of trace moisture existing in the system, part of the epoxy groups open and anhydrides hydrolyze to further react with each other and produce monoester containing carboxyl; (b) carboxyl reacts with the epoxy groups and generates hydroxyl groups; (c) the hydroxyl groups on the surface of SiO2 and the generated hydroxyl react with the epoxy groups to generate hydroxyl groups; (d) the hydroxyl groups on the surface of SiO2 and the generated hydroxyl react with anhydrides to generate carboxylic acids.
Figure 1. Mechanism of crosslinking reaction between the epoxy, acid anhydride curing agent and nano-SiO2. (a) Under the effect of trace moisture existing in the system, part of the epoxy groups open and anhydrides hydrolyze to further react with each other and produce monoester containing carboxyl; (b) carboxyl reacts with the epoxy groups and generates hydroxyl groups; (c) the hydroxyl groups on the surface of SiO2 and the generated hydroxyl react with the epoxy groups to generate hydroxyl groups; (d) the hydroxyl groups on the surface of SiO2 and the generated hydroxyl react with anhydrides to generate carboxylic acids.
Polymers 10 00801 g001
Figure 2. Molecular models. (a) Diglycidyl Ether of Bisphenol A resin (DGEBA); (b) Methyltetrahydrophthalic Anhydride (MTHPA); (c) DGEBA-MTHPA; (d) SiO2.
Figure 2. Molecular models. (a) Diglycidyl Ether of Bisphenol A resin (DGEBA); (b) Methyltetrahydrophthalic Anhydride (MTHPA); (c) DGEBA-MTHPA; (d) SiO2.
Polymers 10 00801 g002
Figure 3. The crosslinked network in Epoxy Resin (EP) nanocomposite models. (a) For 6.5 Å; (b) for 7.6 Å; (c) for 8.8 Å; (d) for 10 Å.
Figure 3. The crosslinked network in Epoxy Resin (EP) nanocomposite models. (a) For 6.5 Å; (b) for 7.6 Å; (c) for 8.8 Å; (d) for 10 Å.
Polymers 10 00801 g003
Figure 4. The Fraction of Free Volume (FFV) of each system with different crosslinking densities.
Figure 4. The Fraction of Free Volume (FFV) of each system with different crosslinking densities.
Polymers 10 00801 g004
Figure 5. The variation of Mean Squared Displacement (MSD) values with time. (a) MSD values of the 6.5 Å system at 300 K with different crosslinking densities; (b) MSD values of the 67% crosslinking density system with 6.5 Å particle sizes at different temperatures; (c) MSD values of the 67% crosslinking density system at 300 K with different particle sizes.
Figure 5. The variation of Mean Squared Displacement (MSD) values with time. (a) MSD values of the 6.5 Å system at 300 K with different crosslinking densities; (b) MSD values of the 67% crosslinking density system with 6.5 Å particle sizes at different temperatures; (c) MSD values of the 67% crosslinking density system at 300 K with different particle sizes.
Polymers 10 00801 g005aPolymers 10 00801 g005b
Figure 6. Linear fitting of the density-temperature curve in each system.
Figure 6. Linear fitting of the density-temperature curve in each system.
Polymers 10 00801 g006
Figure 7. The Tg value of EP and its composites. (a) The Tg value of pure EP and EP composite with 6.5 Å nano-SiO2 at different crosslinking densities; (b) the Tg value of pure EP and EP composite with different particle sizes at 67% crosslinking density.
Figure 7. The Tg value of EP and its composites. (a) The Tg value of pure EP and EP composite with 6.5 Å nano-SiO2 at different crosslinking densities; (b) the Tg value of pure EP and EP composite with different particle sizes at 67% crosslinking density.
Polymers 10 00801 g007
Figure 8. The Coefficient of Thermal Expansion (CTE) in the glassy state and rubbery state. (a) Glassy state; (b) rubbery state.
Figure 8. The Coefficient of Thermal Expansion (CTE) in the glassy state and rubbery state. (a) Glassy state; (b) rubbery state.
Polymers 10 00801 g008
Figure 9. The CTE with different particle sizes under 67% crosslinking density.
Figure 9. The CTE with different particle sizes under 67% crosslinking density.
Polymers 10 00801 g009
Figure 10. The static mechanical properties of the pure EP and 6.5 Å composites with different crosslinking densities. (a) Young’s modulus; (b) shear modulus.
Figure 10. The static mechanical properties of the pure EP and 6.5 Å composites with different crosslinking densities. (a) Young’s modulus; (b) shear modulus.
Polymers 10 00801 g010
Figure 11. The static mechanical properties of the pure EP and composites with different particle sizes at a crosslinking density of 67%. (a) Young’s modulus; (b) shear modulus.
Figure 11. The static mechanical properties of the pure EP and composites with different particle sizes at a crosslinking density of 67%. (a) Young’s modulus; (b) shear modulus.
Polymers 10 00801 g011
Table 1. Initial components of the molecular model of each system.
Table 1. Initial components of the molecular model of each system.
Particle Size of SiO2Molecular Number ofAtomic Number
(Å)DGEBAMTHPADGEBA-MTHPA
none4090104650
6.54090104838
7.665146167823
8.8881992310,708
101242884015,830
Table 2. Tg value of 67% crosslink density.
Table 2. Tg value of 67% crosslink density.
Particle Size (Å)Tg by Fitting Method (K)Tg by MSD Curve Method (K)
6.5408.23400–410
7.6397.55390–400
8.8395.93390–400
10387.39380–390
Pure382.68380–390

Share and Cite

MDPI and ACS Style

Xie, Q.; Fu, K.; Liang, S.; Liu, B.; Lu, L.; Yang, X.; Huang, Z.; Lü, F. Micro-Structure and Thermomechanical Properties of Crosslinked Epoxy Composite Modified by Nano-SiO2: A Molecular Dynamics Simulation. Polymers 2018, 10, 801. https://doi.org/10.3390/polym10070801

AMA Style

Xie Q, Fu K, Liang S, Liu B, Lu L, Yang X, Huang Z, Lü F. Micro-Structure and Thermomechanical Properties of Crosslinked Epoxy Composite Modified by Nano-SiO2: A Molecular Dynamics Simulation. Polymers. 2018; 10(7):801. https://doi.org/10.3390/polym10070801

Chicago/Turabian Style

Xie, Qing, Kexin Fu, Shaodong Liang, Bowen Liu, Lu Lu, Xueming Yang, Zhengyong Huang, and Fangcheng Lü. 2018. "Micro-Structure and Thermomechanical Properties of Crosslinked Epoxy Composite Modified by Nano-SiO2: A Molecular Dynamics Simulation" Polymers 10, no. 7: 801. https://doi.org/10.3390/polym10070801

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