Next Article in Journal
Ciprofloxacin-Modified Degradable Hybrid Polyurethane-Polylactide Porous Scaffolds Developed for Potential Use as an Antibacterial Scaffold for Regeneration of Skin
Next Article in Special Issue
Effects of Zirconium Silicide on the Vulcanization, Mechanical and Ablation Resistance Properties of Ceramifiable Silicone Rubber Composites
Previous Article in Journal
Improving Mildew Resistance of Soy Meal by Nano-Ag/TiO2, Zinc Pyrithione and 4-Cumylphenol
Previous Article in Special Issue
Improved Processing and Properties for Polyphenylene Oxide Modified by Diallyl Orthophthalate Prepolymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Surface Modification Design for Improving the Strength and Water Vapor Permeability of Waterborne Polymer/SiO2 Composites: Molecular Simulation and Experimental Analyses

1
School of Materials Science & Engineering, Shaanxi University of Science and Technology, Xi’an 710021, China
2
Key Laboratory of Leather Cleaner Production, China National Light Industry, College of Bioresources Chemical and Materials Engineering, Shaanxi University of Science and Technology, Xi’an 710021, China
3
Shaanxi Collaborative Innovation Center of Industrial Auxiliary Chemistry and Technology, Xi’an 710021, China
4
Department of Applied Chemistry, School of Science, Northwestern Polytechnical University, Xi’an 710129, China
*
Authors to whom correspondence should be addressed.
Polymers 2020, 12(1), 170; https://doi.org/10.3390/polym12010170
Submission received: 31 July 2019 / Revised: 4 January 2020 / Accepted: 7 January 2020 / Published: 9 January 2020

Abstract

:
Polymer-based nanocomposites properties are greatly affected by interfacial interaction. Polyacrylate nanocomposites have been widely studied, but few studies have been conducted on their interface mechanism. Therefore, there was an urgent demand for providing a thorough understanding of the polymethyl acrylate/SiO2 (PMA/SiO2) nanocomposites to obtain the desired macro-performance. In this paper, a methodology, which combined molecular dynamics simulation with experimental researches, was established to expound the effect of the surface structure of SiO2 particles which were treated with KH550, KH560 or KH570 (KH550-SiO2, KH560-SiO2 and KH570-SiO2) on the mechanical characteristic and water vapor permeability of polymethyl acrylate/SiO2 nanocomposites. The polymethyl acrylate/SiO2 nanocomposites were analyzed in binding energy and mean square displacement. The results indicate that PMA/KH570-SiO2 had the highest tensile strength, while PMA/KH550-SiO2 had the highest elongation at break at the same filler content; KH550-SiO2 spheres can significantly improve water vapor permeability of polyacrylate film.

Graphical Abstract

1. Introduction

Nanoparticles have received significant attention from researchers in the fields of electronics, metallurgy, aerospace, chemical engineering, biology and medicine due to their unique properties [1,2,3,4,5,6,7,8,9,10]. However, on account of the high surface energy of nanoparticles and the fact that they usually do not contain active functional groups, the dispersion of nanoparticles in polymers is very poor, thus leading to weak reinforcing effects [11,12]. A number of studies have shown that most properties of composite materials depend on the nanoparticles dispersion in polymer matrix and their interfacial interaction [13,14,15]. Therefore, there are numerous examples have been proposed for improving the dispersion of nanoparticles in polymer matrix thus improving the interfacial interactions by grafting or surface modified materials on the surface of nanoparticles for functionalization [16,17,18]. Meanwhile, effects of surface structure of nanomaterials on the properties of composites are still the focus of research [19,20,21].
Polyacrylate due to its excellent film forming properties, light stability, weather resistance and compatibility is widely used as a coating material for leather, textile materials, wood, metal surfaces, etc. However, the linear molecular structure and side chain polar groups always cause negative effects on mechanical properties, water resistance and water vapor permeability of the as-obtained films, which limits its scope of use. With the development of nanotechnology, many nanoparticles have been attempted to introduce into polyacrylate to improve its film-forming properties. Chen et al. synthesized polyacrylate/modified-TiO2 coating film [22]. The results prove that the Modification of TiO2 particles can build up their dispersibility in polyacrylate coating and fall off water absorption of coating. Zhao et al. prepared an organic nano-SiO2/fluorinated polyacrylate composite latex [23]. The results showed that the films exhibit strong hydrophobicity. From our previous studies, Pickering emulsion steadied by silica sol has fine effects on pigment printing adhesives [24]. The improvement of latex film performance is due to the addition of nano-TiO2 [25]. We have done a lot of research on polyacrylate/silica [26,27,28,29]. The results show that the introduction of silica improves the mechanical properties, thermal properties and water vapor permeability of polyacrylate film. However, intensive studies on the interfacial interaction between polyacrylate and SiO2 particles have rarely been reported. The interaction between polymers and SiO2 particles originate from molecules, atoms, and their underlying quantum mechanical arguments (nanoscopic and/or sub-nanometer), which interactions are hard to observe and examine by traditional experimental tests.
A strong supporting tool for studying the interfacial interactions between the polymer matrix and nanoparticles is molecular dynamics (MD) simulations, which has been diffusely used to view the interaction machine-processed between different materials [30,31,32,33,34,35,36,37]. Rissanou et al. analyzed several graphene/polymer nanocomposites by atomistic molecular dynamics simulations [38]. The results indicate that chain segmental dynamics is slower at the PE/graphene interface than the bulk one. Wang et al. have performed the mechanical properties of PET/silica composites by MD simulations in detail [39]. The simulation results shown that nanocomposites have the higher mechanical properties in comparison with those in pure PET system, ascribing a stronger interaction between the modified silica and polymer chains. We have successfully employed MD simulation to check on the presence of p-p stacking interaction between poly (styrene-butyl acrylate) latex (P(St-BA)) and sulfonated graphene nanosheet (S-GNS) [40].
In this work, we study the effects of interfacial structure between SiO2 particle and polymer matrix on the properties of their composites. SiO2 was modified by polysiloxane (KH550, KH560 and KH570) with similar chain length to enhance its interfacial compatibility and binding with PA. The properties of its composites were studied through experiments and MD to explore its influence rule. This study is expected to provide a theoretical basis for the structural design of nanoparticles and their applications in functional composites.

2. Materials and Methods

2.1. Materials

All chemicals were of analytical grade and used without any further purification. Octyltrimethoxysilane (OTMS, 97%), ammonia, ethanol, tetraethoxysilane (TEOS), 3-aminopropyl triethoxy silane (KH550, 97%), γ-glycidoxypropyl trimethoxy silane (KH560, 97%), γ-methacryloxypropyl trimethoxy silane (KH570, 97%), xylene, methyl acrylate (MA), sodium dodecyl sulfate (SDS) and potassium persulfate (KPS) were all purchased from Tianjin Fuchen Chemical Reagent Factory.

2.2. Preparation of SiO2 Nanoparticles and Modified-SiO2 Nanoparticles

The procedure for preparing SiO2 nanoparticles was as follows: firstly, 5 mL of ammonia, 100 mL of ethanol and 5mL of TEOS were added into 250 mL three-necked flask. Then, the mixture was stirred for 5 h at 60 °C. Finally, the product was centrifuged and washed by deionized water and ethanol for several times, then dried for 12 h at 60 °C to produce a white SiO2 powder.
The procedure for preparing SiO2 nanoparticles modified by KH570 as follows:
About 5 mL ethanol solution of KH570 was gradually added to 100 mL SiO2 nanoparticles ethanol dispersion under a magnetic stirrer. Then, the mixed solution was stirred at room temperature for 72 h. The mixed solution was centrifuged and washed with xylene, ethanol and deionized water several times, then SiO2 nanoparticles modified by KH570 were attained via drying for 12 h at 60 °C (KH570-SiO2).
The procedure for preparing SiO2 nanoparticles modified by KH550, which is defined as KH550-SiO2 (SiO2 modified by KH560, which is defined as KH560-SiO2) as follows: SiO2 nanoparticles were ultrasonic dispersed into 100 mL of ethyl alcohol at 25 °C for 30 min, and the pH of the mixed solution dispersion reached 6 with glacial acetic acid. Then 5 mL of KH550 (or KH560) was added into the solution under the stirring at 60 °C for 6 h. The obtained solution was then centrifuged and washed by xylene, ethanol and deionized water for several times, and dried at 60 °C for 6 h to obtain KH550-SiO2 (or KH560-SiO2) nanoparticles.

2.3. Preparation of PMA/Modified-SiO2 Nanocomposite and its Composite Film

Some modified-SiO2 (2 wt% of PMA mass) and SDS were ultrasonic dispersed for 10 min at 75 °C. Then KPS and MA were added to the emulsion drop wisely, while the polymerization was conducted at 75 °C for 6 h.
Finally, the PMA/modified-SiO2 composite emulsion was poured into polytetrafluoroethylene (PTFE) mold and laid on the horizontal surface until dried completely at room temperature to obtain PMA/modified-SiO2 composite film. PMA/SiO2 composite film was prepared in the same method.

2.4. Characterization and Measurements

The microstructures and morphology of all samples were measured by scanning electron microscope (SEM, S4800, Rigaku) and transmission electron microscope (TEM, Tecnai G2 F20, FEI). The chemical structures of all materials were analyzed by fourier transform infrared spectrum (FT-IR, VECTOR-22, Brucker) and Ultraviolet-visible-near infrared spectrophotometer (Cary 5000, Agilent). The thermal stability of the samples was studied by thermogravimetric analysis (TGA, STA409PC, Netzsch) at the nitrogen atmosphere from room temperature to 600 °C with a heating rate of 5 °C/min. The glass transition process of samples was characterized using a differential scanning calorimeter (DSC, Q5000 IR) The mechanical properties were tested by a servo material multi-functional high and low temperature control testing machine (AI-7000-NGD, Goodtechwill) at a loading rate of 100 mm/min according to QB/T 1331-1998. The water vapor transmission rate (WVP) was tested by a Water vapor transmittance tester (W3/060, Labthink) according to QB/T 1279-2012.

2.5. Simulation Methodologies

The reactant molecules were built in the Visualizer module of Material Studio 8.0 software (Accelrys Inc., San Diego, CA, USA). For the MD simulations, the Forcite and Amorphous cell modules of the Materials Studio suite of software were used. All the theoretical calculations were performed using the Condensed-Phase Optimized Molecular Potentials for Atomistic Simulation Studies COMPASS force field [41,42].

2.5.1. Construction of SiO2 Nanoparticles

The xsd molecular model of SiO2 is imported from MS software material library.
After the unit cell model of silica is obtained, it is geometrically optimized to obtain a lower energy structure. Next, the silica particles with a radius of 1 nm (10 Å) are constructed, and a spherical silica nanoparticle with a surface saturated with unsaturated bonds between Si atoms and O atoms is obtained. Select the broken bond on the Si atom on the surface of the SiO2 unit cell and combine it with -OH, and combine the broken bond on the surface O atom with the H atom. Increase the reliability, and optimize the structure to obtain a spherical SiO2 model (Figure 1).
Three different modifiers were grafted on the silica surface, and the three modifiers were KH550 (2a), KH560 (2b) and KH570 (2c). The structure of the modified surface was optimized, and the energy converged to 1 × 10−4 kca1/mol. Figure 2 shows the structure of three modifiers, and Figure 3 shows the surface of modified-SiO2. In this paper, four identical silane coupling agent chains were grafted onto SiO2 sphere, and the grafted microspheres were optimized by the Smart method to optimize the energy to 1 × 10−4 kca1/mol. In order to search for the optimal structure, the cell is then annealed at 0.1 MPa from the low temperature of 300 K to the upper temperature of 500 K for 200 ps to prevent the system to form being trapped at a local high energy minimum. Subsequently, 200 ps of NVT (constant number of particles, volume, and temperature) simulation is performed at 298 K.
Since the double bond on KH570-SiO2 is polymerized with MA, in this paper a model of polymerizing one double bond on the surface of KH570-SiO2 with MA (PMA-KH570-SiO2) is constructed. PMA polymer chain has 20 repeat units, as shown in the Figure 4.

2.5.2. Construct the Composite System Model

PMA/SiO2 (KH550-SiO2, KH560-SiO2): Amorphous cells containing composites of PMA polymer chains with 20 repeat units and a SiO2 (or modified-SiO2) nanoparticle (diameter 20 nm) were constructed, and periodic boundary conditions were applied.
PMA/KH570-SiO2: Amorphous cells containing composites of PMA polymer chains with 19 repeat units and one PMA-KH570-SiO2 (Figure 4) were constructed, and periodic boundary conditions were applied.
To study the diffusion coefficient of H2O in composite systems, the MSDs of H2O in composite systems were analyzed. Some composite systems containing water molecules were constructed as follows (Supplementary Materials, Figure S1):
PMA/SiO2 (KH550-SiO2, KH560-SiO2)/H2O: Amorphous cells containing composites of PMA polymer chains with 20 repeat units, one SiO2 (or modified-SiO2) nanoparticle (diameter 20 nm) and 10 H2O molecules were constructed, and periodic boundary conditions were applied.
PMA/KH570-SiO2/H2O: Amorphous cells containing composites of PMA polymer chains with 19 repeat units, one PMA-KH570-SiO2 (Figure 4) and 10 H2O molecules were constructed, and periodic boundary conditions were applied.

2.5.3. Molecular Dynamics Simulation Process

After building PMA/SiO2 (or modified-SiO2) composite systems, the energy of each generated cell is minimized to a convergence value of 1.0 × 10−4 kcal mol−1 by using the Smart Minimizer method to relax the state of minimal potential energy. Whereafter, 200 ps of NVT (constant number of particles, volume, and temperature) simulation is performed at 298 K. The cell is then annealed at 0.1 MPa from the low temperature of 300 K to the upper temperature of 500 K for 200 ps to prevent the system to form being trapped at a local high energy minimum [43]. Subsequently, 200 ps of NVT (constant number of particles, volume, and temperature) simulation is performed at 298 K, 500 ps of NPT (constant number of particles, pressure, and temperature) simulation is performed at 0.1 MPa and 2 ns of NVE (constant number of particles, volume, and energy) simulation is performed to further relax the polymer structure by using the Andersen Thermostat for temperature control and the Berendsen Barostat for pressure control [44,45] (Figure 5).
At last, the cell can be used to analyze properties of the system. In order to further verify the effect of the number of polymer chains on the properties of composites, a composite system with 30 polymer chains was studied (Figure S2). The results show that the binding energy between the polymer and SiO2 is not significantly different from the data of the composite system of 20 polymer chains in the article, which is reasonable (Tables S1 and S2). This result indicates that the composite system of 20 polymer chains may basically match the experiment. With the increasing of the number of polymer chains, the performance of the composite system has not changed significantly.
System equilibrium is judged by temperature and energy balance. Figure 6 shows the trajectory temperature and energy fluctuation chart of 200 ps NVT in the MD equilibrium stage. From the Figure 5, the trajectory energy fluctuation of each frame is gentle, indicating that the system energy has reached equilibrium.
The above two criteria showed that the PMA/SiO2 composite system has indeed reached equilibrium through MD simulation, and the subsequent analysis results are reliable. The remaining PMA and modified-SiO2 interaction systems could all reach the same conclusion.

3. Results

3.1. Morphological and Structural Characterization of Modified-SiO2

Surface modification is essential for the synthesis and functionality of composites. FT-IR is often used to characterize surface modification. FT-IR spectra of SiO2 and modified-SiO2 are shown in Figure 7a. The characteristic peaks assigned to the stretching vibration of Si-O-Si at 1101 cm−1 are observed in the spectrum of SiO2. After modification with silane coupling agent, there appeared absorption peaks at 1705 cm−1 (C=C stretching vibration) in the spectrum of KH570-SiO2, as well as peaks at 1730 cm−1 (–C–H– asymmetric stretching vibration), 1623 cm−1 (N–H in-plane bending vibration) in the spectrum of KH550-SiO2. The peak at 2977 cm−1 reveals the existence of –CH3 or –CH2– on SiO2 surface. These may suggest that silane coupling agent is successfully grafted onto the SiO2 surface.
To reveal the chemically bonding between SiO2 and the silane coupling agent, UV absorption spectrum of the modified-SiO2 was characterized. It can be seen in Figure 7b that there is a small shift in the position of the absorption peak, which may be attributed to changes in the surface structure of SiO2.
SEM images (Figure 8) present that the as-prepared SiO2 and modified-SiO2 samples are uniform in size with spherical shape, and the average size was about 80 nm. Meanwhile, SiO2 are well dispersed (See Figure 8a). Nevertheless, the modified-SiO2 are not particularly well dispersed (Figure 8b–d), which may be caused by self-polymerization of the silane coupling agent.

3.2. Morphological and Structural Characterization of PMA/SiO2 Composite Emulsion and Film

Figure 9 shows the TEM image of PMA and PMA/SiO2 composite latex particles. As shown in Figure 9a, the pristine PMA latex particles display a well-defined spherical morphology and the latex particle size is 100–200 nm. The black phase which are SiO2 particles are on the surface of latex particles (the fuzzy layers). SiO2 is located onto the surface of latex particles (Figure 9b), KH550-SiO2 is onto the surface of latex particles (Figure 9c), and KH560-SiO2 is also on the surface of latex particles (Figure 9d). There are more KH550-SiO2 and 560-SiO2 particles are on the surface of latex particles, this is mainly because that stronger hydrogen bonds or electrostatic interactions is formed between modified-SiO2 nanoparticles and latex particles, compared with SiO2 nanoparticles. KH570-SiO2 nanoparticles which enter the interior of PMA latex particles seem less clear (Figure 9e), and other nanoparticles appear on the latex particle surface as black phase [46]. Different surface modification between the above four SiO2 sources explains the difference in the distribution of SiO2 nanoparticles.
The dispersion of SiO2 particles in the PMA film before and after modification can be observed by SEM, as shown in Figure 10. The untreated SiO2 particles aggregated severely in PMA film with the size equivalenting about to 500 nm (Figure 10b). The KH560-SiO2 particles have better dispersibility in PMA, and good interfacial adhesion with PMA film. Which is better than those of the unmodified particles (Figure 10c). Nevertheless, there are still some aggregates in PMA film. KH550-SiO2 particles are well dispersed as small aggregates (Figure 10d). When PMA film was filled with KH570-SiO2 particles, the nanoparticles are uniformly dispersed in PMA film, and it is difficult to see the aggregate of the nanoparticles (Figure 10e). The interfacial compatibility between KH570-SiO2 nanoparticles and PMA film is well.
The unmodified-SiO2 particles aggregate in PMA film on account of their high polar surface energy. The poor compatibility of aggregated SiO2 particles with the PMA film is due to the hydrophilic surface. When SiO2 particles are modified with KH560, their surface is covered by long alkyl chain, which gives the particles a well interfacial compatibility to PMA film [47]. The interface compatibility between KH550-SiO2 particles and PMA film is a bit better, which introduce the amidogen group. The KH570-SiO2 particles can form chemically bond with PMA matrix through double bond polymerization. Good compatibility makes better dispersion of SiO2 particles in PMA film.

3.3. Properties of PMA/SiO2 and PMA/Modified-SiO2 Composite Films

3.3.1. Mechanical Properties of PMA/SiO2 and PMA/Modified-SiO2 Composite Films

Figure 11 reveals the tensile strength and elongation at break of PMA/SiO2 nanocomposites, in which SiO2 is modified by different silane coupling agents. As everyone knows, the interfacial interaction between polymer and nanoparticles has a greater influence on the tensile strength of composites [20]. The weak interfacial interaction between polymer and nanoparticles results in less stress being transferred from polymer to nanoparticles [47]. The stronger the interfacial interaction between polymer and nanoparticles, the greater the stress transmitted by the polymer to the nanoparticles, resulting in higher tensile strength. As can be seen in Figure 11a, the tensile strength of PMA/modified-SiO2 film is higher than that of PMA/SiO2 film. And, the tensile strength of PMA/KH570-SiO2 film have the highest value than that of other films.
The dispersion and interfacial interaction between polymer and nanoparticles have a great influence on elongation at break of nanocomposites [48,49]. Good dispersibility and proper interfacial interaction can enhance the value of elongation at break, while excessive strong interfacial interaction can reduce it. The elongation at break of composite films is shown in Figure 11b, compared with KH550-SiO2, the addition of KH570-SiO2 reduced the elongation at break of the film, which is consistent with the strong interfacial interaction between PMA and KH570-SiO2.

3.3.2. Water Vapor Permeability of PMA/SiO2 and PMA/Modified-SiO2 Composite Films

Water vapor permeability is an important index when emulsion products are used in coatings requiring air permeability. The water vapor permeation rate is primarily assumed by the diffusion process and adsorption/desorption process, which are influenced by the composition and structure of the polymer chains. The soft polymer chains of polyacrylate can give more free volume for the passage of water vapor molecules, although hydrophobic segments of polyacrylate are detrimental to the adsorption process [50].
The effects of functionalized SiO2 on water vapor permeability of as-obtained films are shown in Figure 12a. Compared with pure PMA film (Figure 12a), water vapor transmission rate of composite films is improved. It is evident that by blending the KH550-SiO2 or KH560-SiO2 nanoparticles in the PMA film, water vapor permeability of PMA/SiO2 nanocomposite films increases significantly. This is mainly attributed to the fact that KH550-SiO2 and KH560-SiO2 contain a hydrophilic amino group or an epoxy group to facilitate water vapor transmission through the film. It can be seen that the water vapor transmission rate of PMA/SiO2 and PMA/KH570-SiO2 composite films are higher than that of PMA composite film, which is mainly result from that: On the one hand, there is a large number of interfacial pores between SiO2 nanoparticles and PMA film, which provides a good channel for water vapor molecules. On the other hand, an enhancement in the amount of hydrophilic groups in film leads to an increase in water vapor permeability [51]. The surface of SiO2 nanoparticles contains a large amount of hydroxyl groups, which increases the number of hydrophilic groups inside the film.

3.3.3. Water Resistance of PMA/SiO2 and PMA/Modified-SiO2 Composite Films

Under normal circumstances, water absorption of the composite film is used to reflect its water resistance, and the higher the water absorption rate, the worse the water resistance. It can be seen from Figure 12b that compared with PMA film, water absorption of PMA/KH550-SiO2 and PMA/KH560-SiO2 composite films show higher value than that of PMA. While water absorption of PMA/SiO2 and PMA/KH570-SiO2 composite films show lower value than those of PMA, and the PMA/KH570-SiO2 composite film has the lowest water absorption. This is mainly due to the fact that KH550-SiO2 and KH560-SiO2 contain a hydrophilic amino group and an epoxy group to facilitate water vapor transmission through the film, while PMA/KH570-SiO2 contains a hydrophobic C=C which is not favorable for water vapor transmission through the film.

3.3.4. Thermal Properties of PMA/SiO2 Composite Films

TGA curves of as-prepared films are shown in Figure 13. Compared with PMA film, the T5 of PMA/SiO2 and PMA/modified-SiO2 increased.
The corresponding characteristic heat data for all samples is shown in Table 1. By comparing, heat-resistance index of the PMA/SiO2 and PMA/modified-SiO2 composite films obviously increased. This is mainly because that the addition of SiO2 nanoparticles causes entanglement of polymer chain, which slows down the decomposition of the molecular chain. The heat-resistance index of PMA/modified-SiO2 composite films is higher than that of PMA/SiO2. In addition, the heat-resistance index of PMA/KH570-SiO2 composite film is the highest. This is mainly ascribed to the stronger interface interaction between PMA and KH570-SiO2 [52,53].

3.4. Molecular Dynamics Simulation

3.4.1. Binding Energy Analysis

Molecular dynamics simulations are currently effective methods for verifying the strength of interfacial interactions. The strength of the interaction between the PMA film and the SiO2 particles can be reflected by the amount of binding energy between them. Generally, the greater the binding energy, the stronger the interaction force between PMA film and SiO2 particles. As a result, the simulation of the binding energy between PMA film and SiO2 (or modified-SiO2) particles can be used to study the interaction mechanism. The binding energies of PMA/SiO2 and PMA/modified-SiO2 composite materials can be used by:
E b i n g d i n g = E inter = ( E t o t a l E P M A E S i O 2 ) , E b i n g d i n g = E inter = ( E t o t a l E P M A E m o d i f i e d S i O 2 )
where Etotal is the energy of the PMA/SiO2 or PMA/modified-SiO2, EPMA is the energy of PMA, and ESiO2 is the energy of SiO2 particles, Emodified-SiO2 is the energy of modified-SiO2 particles. The binding energies between PMA and SiO2 (or modified-SiO2) are given in Table 2.
The total energy of the PMA/SiO2 (modified-SiO2) system, the energy of PMA, and the energy of SiO2 (modified-SiO2) are presented in Table 2.
The binding energy of PMA/KH550-SiO2 reaches a higher value than PMA/SiO2, showing the strongest interfacial interaction between PMA film and KH550-SiO2 particles. A higher binding energy shows good compatibility between PMA film and KH550-SiO2 particles. As the KH560-SiO2 was added into PMA, the binding energy of the PMA/KH560-SiO2 film is better than that of PMA/KH550-SiO2 film, a sign of well compatibility of PMA film with KH560-SiO2 particles [54].

3.4.2. MSD (Mean Square Displacement) and Diffusion Coefficient (D) of Water in Composite System

The diffusion coefficient (D) of water molecules in nanocomposite reflects the water vapor permeability of nanocomposite. The higher the diffusion coefficient is, the better the water vapor permeability is. To study the diffusion coefficient of H2O in films, the MSDs of H2O in the films were analyzed (Figure 14). Diffusivity was calculated by using the slope of MSD diagram (Figure 14a) [55].
The results show that the diffusivity changes as the change of fillers (Figure 14b). The addition of SiO2 (modified-SiO2) nanoparticles improves the diffusivity of H2O molecules in PMA films. The addition of SiO2 nanoparticles results in bigger voids at the interface of PMA/SiO2 (modified-SiO2) film, which facilitates the rapid passage of H2O molecules. The diffusion coefficient of water molecules in PMA/KH550-SiO2 composite system is the best. And after that, the order of diffusion coefficient of water molecules in composite systems from high to low is PMA/KH560-SiO2, PMA/KH570, PMA and PMA/SiO2, respectively. This is consistent with the results of the water vapor permeability of the previous films.

4. Conclusions

In this study, experimental methods combined with molecular simulation ways have been successfully applied to study the microstructure-property relationship in various polyacrylate/modified-SiO2 composite systems. To investigate the effect of different surface structures on mechanical property and water vapor permeability of resultant PMA/SiO2 nanocomposite films, SiO2 particles were treated with polysiloxane (KH550, KH560 and KH570). Different functional groups on the surface of SiO2 lead to different interfacial interactions with PMA, which give different mechanical properties of composite films. At the same time, due to the different hydrophilic and hydrophobic surface of SiO2, water vapor permeability of composite film is different. It can be used to study the interfacial interaction and microstructure-property relationships of polyacrylate-based nanocomposites, thus guiding the design of high performance polyacrylate-based nanocomposites.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/12/1/170/s1, Figure S1: Models for water diffusion in composite system, Figure S2. Models for MD simulation of composite system, Table S1: Binding energy of PMA/SiO2 and PMA/KH560-SiO2 composites, Table S2: Binding energy of composites system.

Author Contributions

Conceptualization, J.M., C.L. and Y.W.; methodology, Y.W.; software, Y.W.; validation, J.M. and C.L.; resources, J.M. and C.L.; data curation, Y.W.; writing—original draft preparation, Y.W.; writing—review and editing, J.M., C.L. and Y.W.; supervision, J.M., C.L., H.Y. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2017YFB0308602), the Natural Science Basic Research Plan in Shaanxi Province of China (2018JQ5211) and the Special Scientific Research Program Founded by Shaanxi Provincial Education Department (18JK0102).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, Y.; Ji, H.; Zhu, K.; Jing, W.; Qiu, J. Dramatically improved piezoelectric properties of poly(vinylidene fluoride) composites by incorporating aligned TiO2 @MWCNTs. Compos. Sci. Technol. 2016, 123, 259–267. [Google Scholar]
  2. Wang, H.; Hsu, J.H.; Yi, S.I.; Kim, S.L.; Choi, K.; Yang, G.; Yu, C. Thermally Driven Large N-Type Voltage Responses from Hybrids of Carbon Nanotubes and Poly(3,4-ethylenedioxythiophene) with Tetrakis(dimethylamino)ethylene. Adv. Mater. 2016, 27, 6855–6861. [Google Scholar] [CrossRef] [PubMed]
  3. Yu, C.; Zhang, H.B.; Yang, Y.; Mu, W.; Cao, A.; Yu, Z.Z. High-Performance Epoxy Nanocomposites Reinforced with Three-Dimensional Carbon Nanotube Sponge for Electromagnetic Interference Shielding. Adv. Funct. Mater. 2016, 26, 447–455. [Google Scholar]
  4. Hou, G.; Tao, W.; Liu, J.; Gao, Y.; Zhang, L.; Li, Y. Tailoring the dispersion of nanoparticles and the mechanical behavior of polymer nanocomposites by designing the chain architecture. Phys. Chem. Chem. Phys. 2017, 19, 32024–32037. [Google Scholar] [CrossRef] [PubMed]
  5. Shojaeiarani, J.; Bajwa, D.; Stark, N.M. Spin-coating: A new approach for improving dispersion of cellulose nanocrystals and mechanical properties of poly (lactic acid) composites. Carbohyd. Polym. 2018, 190, 139–147. [Google Scholar] [CrossRef] [PubMed]
  6. Yue, Y.; Xu, M.; Hai, P.; Ren, D.; Li, K.; Liu, X. Secondary dispersion of BaTiO3 for the enhanced mechanical properties of the Poly (arylene ether nitrile)-based composite laminates. Polym. Test. 2018, 66, 164–171. [Google Scholar]
  7. Kazemi, Y.; Kakroodi, A.R.; Ameli, A.; Filleter, T.; Park, C.B. Highly stretchable conductive thermoplastic vulcanizate/carbon nanotube nanocomposites with segregated structure, low percolation threshold and improved cyclic electromechanical performance. J. Mater. Chem. C 2018, 6, 350–359. [Google Scholar] [CrossRef]
  8. Arjmand, M.; Sadeghi, S.; Khajehpour, M.; Sundararaj, U. Carbon nanotube/graphene nanoribbon/polyvinylidene fluoride hybrid nanocomposites: Rheological and dielectric properties. J. Phys. Chem. C 2016, 121, 169–181. [Google Scholar] [CrossRef]
  9. Kou, Z.; Zhang, L.; Ma, Y.; Liu, X.; Zang, W.; Zhang, J.; Huang, S.; Du, Y.; Cheetham, A.K.; Wang, J. 2D carbide nanomeshes and their assembling into 3D microflowers for efficient water splitting. Appl. Catal. B Environ. 2019, 243, 678–685. [Google Scholar] [CrossRef]
  10. Si, Y.; Huang, T.; Li, Q.; Huang, Y.; Gao, S.; Chen, M.; Wu, L. Hierarchical Macro–Mesoporous Polymeric Carbon Nitride Microspheres with Narrow Bandgap for Enhanced Photocatalytic Hydrogen Production. Adv. Mater. Interfaces 2018, 5, 1801241. [Google Scholar] [CrossRef]
  11. Senses, E.; Akcora, P. An Interface-Driven Stiffening Mechanism in Polymer Nanocomposites. Macromolecules 2013, 46, 1868–1874. [Google Scholar] [CrossRef]
  12. Yang, J.; Han, C.R.; Duan, J.F.; Xu, F.; Sun, R.C. Interaction of Silica Nanoparticle/Polymer Nanocomposite Cluster Network Structure: Revisiting the Reinforcement Mechanism. J. Phys. Chem. C 2013, 117, 8223–8230. [Google Scholar] [CrossRef]
  13. Petcu, C.; Purcar, V.; Spătaru, C.; Alexandrescu, E.; Şomoghi, R.; Trică, B.; Niţu, S.; Panaitescu, D.; Dan, D.; Jecu, M.L. The Influence of New Hydrophobic Silica Nanoparticles on the Surface Properties of the Films Obtained from Bilayer Hybrids. Nanomaterials 2017, 7, 47. [Google Scholar] [CrossRef] [PubMed]
  14. Sabzi, M.; Mirabedini, S.M.; Zohuriaan-Mehr, J.; Atai, M. Surface modification of TiO2 nano-particles with silane coupling agent and investigation of its effect on the properties of polyurethane composite coating. Prog. Org. Coat. 2009, 65, 222–228. [Google Scholar] [CrossRef]
  15. Kumar, S.K.; Jouault, N.; Benicewicz, B.; Neely, T. Nanocomposites with Polymer Grafted Nanoparticles. Macromolecules 2013, 46, 3199–3214. [Google Scholar] [CrossRef]
  16. Zhang, X.; Zheng, J.; Fang, H.; Zhang, Y.; Bai, S.; He, G. High dimensional stability and low viscous response solid propellant binder based on graphene oxide nanosheets and dual cross-linked polyurethane. Compos. Sci. Technol. 2018, 161, 124–134. [Google Scholar] [CrossRef]
  17. Zou, Y.; Sun, Y.; He, J.; Tang, Z.; Zhu, L.; Luo, Y.; Liu, F. Enhancing mechanical properties of styrene-butadiene rubber/silica nanocomposites by in situ interfacial modification with a novel rare-earth complex. Compos. Part A Appl. Sci. Manuf. 2016, 87, 297–309. [Google Scholar] [CrossRef]
  18. Fang, H.; Zhao, Y.; Zhang, Y.; Ren, Y.; Bai, S. Three-dimensional graphene foam-filled elastomer composites with high thermal and mechanical properties. ACS Appl. Mater. Inter. 2017, 9, 26447–26459. [Google Scholar] [CrossRef]
  19. Zhang, S.; Meng, G.; Chen, Z.; Qiu, H.L.; Liu, X. Grafting photosensitive polyurethane onto colloidal silica for use in UV-curing polyurethane nanocomposites. Colloids Surf. A Physicochem. Eng. Asp. 2014, 443, 525–534. [Google Scholar] [CrossRef]
  20. Yu, Y.; Shu, Y.; Ye, L. In situ crosslinking of poly (vinyl alcohol)/graphene oxide-glutamic acid nano-composite hydrogel as microbial carrier: Intercalation structure and its wastewater treatment performance. Chem. Eng. J. 2018, 336, 306–314. [Google Scholar] [CrossRef]
  21. Huang, Q.; Liu, M.; Chen, J.; Wan, Q.; Tian, J.; Huang, L.; Jiang, R.; Deng, F.; Wen, Y.; Zhang, X.; et al. Marrying the mussel inspired chemistry and Kabachnik-Fields reaction for preparation of SiO2 polymer composites and enhancement removal of methylene blue. Appl. Surf. Sci. 2017, 422, 17–27. [Google Scholar] [CrossRef]
  22. Chen, Y.; Lin, A.; Gan, F. Improvement of polyacrylate coating by filling modified nano-TiO2. Appl. Surf. Sci. 2006, 252, 8635–8640. [Google Scholar] [CrossRef]
  23. Zhao, F.; Zeng, X.; Li, H.; Zhang, J. Preparation and characterization of nano-SiO2/fluorinated polyacrylate composite latex via nano-SiO2/acrylate dispersion. Colloids Surf. A Physicochem. Eng. Asp. 2012, 396, 328–335. [Google Scholar] [CrossRef]
  24. Gao, D.; Rui, C.; Lyu, B.; Ma, J.; Duan, X. Preparation of epoxy-acrylate copolymer/nano-silica via Pickering emulsion polymerization and its application as printing binder. Appl. Surf. Sci. 2018, 435. [Google Scholar] [CrossRef]
  25. Yan, B.; Shi, C.; Ma, J.; Bing, W.; Zhang, Y. Double in-situ synthesis of polyacrylate/nano-TiO2 composite latex. Prog. Org. Coat. 2015, 85, 101–108. [Google Scholar]
  26. Gao, D.; Chang, R.; Lyu, B.; Ma, J. Growth from spherical to rod-like SiO2: Impact on microstructure and performance of nanocomposite. J. Alloys Compd. 2019, 151814. [Google Scholar] [CrossRef]
  27. Gao, D.; Chang, R.; Lyu, B.; Ma, J.; Liu, J.; Li, Q. Synthesis of raspberry-like SiO2/polyacrylate nanocomposite latexes via a one-step miniemulsion polymerization and its film properties. J. Sol-Gel Sci. Technol. 2019, 92, 695–705. [Google Scholar] [CrossRef]
  28. Ma, J.; Hu, J.; Zhang, Z. Polyacrylate/silica nanocomposite materials prepared by sol–gel process. Eur. Polym. J. 2007, 43, 4169–4177. [Google Scholar] [CrossRef]
  29. Bao, Y.; Shi, C.; Yang, Y.; Ma, J.; Sha, R. Effect of hollow silica spheres on water vapor permeability of polyacrylate film. RSC Adv. 2015, 5, 11485–11493. [Google Scholar] [CrossRef]
  30. Lu, J.; Liu, D.; Yang, X.; Zhao, Y.; Liu, H.; Tang, H.; Cui, F. Molecular dynamics simulations of interfacial interactions between small nanoparticles during diffusion-limited aggregation. Appl. Surf. Sci. 2015, 357, 1114–1121. [Google Scholar] [CrossRef]
  31. Roussou, R.; Karatasos, K. Graphene/poly(ethylene glycol) nanocomposites as studied by molecular dynamics simulations. Mater. Des. 2016, 97, 163–174. [Google Scholar] [CrossRef]
  32. 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]
  33. Sadeghi, S.; Arjmand, M.; Otero Navas, I.; Zehtab Yazdi, A.; Sundararaj, U. Effect of nanofiller geometry on network formation in polymeric nanocomposites: Comparison of rheological and electrical properties of multiwalled carbon nanotube and graphene nanoribbon. Macromolecules 2017, 50, 3954–3967. [Google Scholar] [CrossRef]
  34. Gooneie, A.; Schuschnigg, S.; Holzer, C. Dissipative Particle Dynamics Models of Orientation of Weakly-Interacting Anisometric Silicate Particles in Polymer Melts under Shear Flow: Comparison with the Standard Orientation Models. Macromol. Theor. Simul. 2016, 25, 287–302. [Google Scholar] [CrossRef]
  35. Gooneie, A.; Hufenus, R. Hybrid carbon nanoparticles in polymer matrix for efficient connected networks: Self-assembly and continuous pathways. Macromolecules 2018, 51, 3547–3562. [Google Scholar] [CrossRef]
  36. Majidian, M.; Grimaldi, C.; Forró, L.; Magrez, A. Role of the particle size polydispersity in the electrical conductivity of carbon nanotube-epoxy composites. Sci. Rep. 2017, 7, 12553. [Google Scholar] [CrossRef] [Green Version]
  37. Fujishiro, S.; Kan, K.; Akashi, M.; Ajiro, H. Stability of adhesive interfaces by stereocomplex formation of polylactides and hybridization with nanoparticles. Polym. Degrad. Stab. 2017, 141, 69–76. [Google Scholar] [CrossRef]
  38. Rissanou, N.A.; Power, J.A.; Harmandaris, V. Structural and Dynamical Properties of Polyethylene/Graphene Nanocomposites through Molecular Dynamics Simulations. Polymers 2015, 7, 390–417. [Google Scholar] [CrossRef] [Green Version]
  39. Wang, Y.; Wang, W.; 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]
  40. Wei, L.; Zhang, W.; Ma, J.; Bai, S.; Ren, Y.; Liu, C.; Simion, D.; Qin, J. π-π stacking interface design for improving the strength and electromagnetic interference shielding of ultrathin and flexible water-borne polymer/sulfonated graphene composite. Carbon 2019, 149, 679–692. [Google Scholar] [CrossRef]
  41. Dai, S.; Yan, L.; Huang, Z.; Zhao, X. Molecular dynamics simulations on the interaction between microsphere and water in nanosilica/crosslinked polyacrylamide microsphere aqueous solution with a core–shell structure and its swelling behavior. Compos. Interface 2017, 25, 69–92. [Google Scholar] [CrossRef]
  42. Dai, S.; Yan, L.; Zhang, J.; Zhang, T.; Huang, Z.; Zhao, X. Molecular dynamic simulation of core–shell structure: Study of the interaction between modified surface of nano-SiO2 and PAMAA in vacuum and aqueous solution. Compos. Interface 2017, 24, 897–914. [Google Scholar] [CrossRef]
  43. Colmenero, J.; Alvarez, F.; Arbe, A. Self-motion and the alpha relaxation in a simulated glass-forming polymer: Crossover from Gaussian to non-Gaussian dynamic behavior. Phys. Rev. E Stat. Nonlin. Soft Matter. Phys. 2002, 65, 41804. [Google Scholar] [CrossRef] [PubMed]
  44. Basconi, J.E.; Shirts, M.R. Effects of Temperature Control Algorithms on Transport Properties and Kinetics in Molecular Dynamics Simulations. J. Chem. Theory Comput. 2013, 9, 2887–2899. [Google Scholar] [CrossRef]
  45. Davoodi, J.; Ahmadi, M. Molecular dynamics simulation of elastic properties of CuPd nanowire. Compos. Part B Eng. 2012, 43, 10–14. [Google Scholar] [CrossRef]
  46. Martín-Fabiani, I.; Koh, M.L.; Dalmas, F.; Elidottir, K.L.; Hinder, S.J.; Jurewicz, I.; Lansalot, M.; Bourgeat-Lami, E.; Keddie, J.L. Design of waterborne nanoceria/polymer nanocomposite UV-absorbing coatings: Pickering versus blended particles. ACS Appl. Nano Mater. 2018, 1, 3956–3968. [Google Scholar] [CrossRef]
  47. Zhang, L.; Luo, M.; Sun, S.; Ma, J.; Li, C. Effect of Surface Structure of Nano-CaCO3 Particles on Mechanical and Rheological Properties of PVC Composites. J. Macromol. Sci. Part B 2010, 49, 970–982. [Google Scholar] [CrossRef]
  48. Meijer, H.E.H.; Govaert, L.E. Mechanical performance of polymer systems: The relation between structure and properties. Prog. Polym. Sci. 2005, 30, 915–938. [Google Scholar] [CrossRef]
  49. Jancar, J.; Kucera, J. Yield behavior of polypropylene filled with CaCO3 and Mg(OH)2. I. ‘Zero’ interfacial adhesion. Polym. Eng. Sci. 2004, 30, 707–713. [Google Scholar] [CrossRef]
  50. Shen, J.; Sun, J.W.; Hu, Y.; Kan, C.Y. Polysiloxane/polyacrylate composite latexes with balanced mechanical property and breathability: Effect of core/shell mass ratio. J. Appl. Polym. Sci. 2017, 134, 45258. [Google Scholar] [CrossRef]
  51. Yan, B.; Yang, Y.; Ma, J. Fabrication of monodisperse hollow silica spheres and effect on water vapor permeability of polyacrylate membrane. J. Colloid Interface Sci. 2013, 407, 155–163. [Google Scholar]
  52. Huangfu, Y.; Ruan, K.; Qiu, H.; Lu, Y.; Liang, C.; Kong, J.; Gu, J. Fabrication and investigation on the PANI/MWCNT/thermally annealed graphene aerogel/epoxy electromagnetic interference shielding nanocomposites. Compos. Part A: Appl. Sci. Manuf. 2019, 121, 265–272. [Google Scholar] [CrossRef]
  53. Yang, X.; Tang, L.; Guo, Y.; Liang, C.; Zhang, Q.; Kou, K.; Gu, J. Improvement of thermal conductivities for PPS dielectric nanocomposites via incorporating NH2-POSS functionalized nBN fillers. Compos. Part A Appl. Sci. Manuf. 2017, 101, 237–242. [Google Scholar] [CrossRef]
  54. Song, M.; Zhao, X.; Li, Y.; Hu, S.; Zhang, L.; Wu, S. Molecular dynamics simulations and microscopic analysis of the damping performance of hindered phenol AO-60/nitrile-butadiene rubber composites. RSC Adv. 2014, 4, 6719–6729. [Google Scholar] [CrossRef]
  55. Khosravanian, A.; Dehghani, M.; Pazirofteh, M.; Asghari, M.; Mohammadi, A.H.; Shahsavari, D. Grand canonical Monte Carlo and molecular dynamics simulations of the structural properties, diffusion and adsorption of hydrogen molecules through poly (benzimidazoles)/nanoparticle oxides composites. Int. J. Hydrogen Energy 2018, 43, 2803–2816. [Google Scholar] [CrossRef]
Figure 1. SiO2 nanoparticle without broken bonds on the surface.
Figure 1. SiO2 nanoparticle without broken bonds on the surface.
Polymers 12 00170 g001
Figure 2. Molecular model of modifiers: (a). KH550, (b). KH560, (c). KH570 (Gray represents C atom, yellow represents Si atom, white represents H atom, red represents O atom and blue represents N atom).
Figure 2. Molecular model of modifiers: (a). KH550, (b). KH560, (c). KH570 (Gray represents C atom, yellow represents Si atom, white represents H atom, red represents O atom and blue represents N atom).
Polymers 12 00170 g002
Figure 3. Model of modified-SiO2: (a). KH550-SiO2, (b). KH560- SiO2, (c). KH570-SiO2.
Figure 3. Model of modified-SiO2: (a). KH550-SiO2, (b). KH560- SiO2, (c). KH570-SiO2.
Polymers 12 00170 g003
Figure 4. Model of PMA-KH570-SiO2.
Figure 4. Model of PMA-KH570-SiO2.
Polymers 12 00170 g004
Figure 5. Models for MD simulation of composite system: (a). PMA, (b). PMA/SiO2, (c). PMA/KH550-SiO2, (d). PMA/KH560-SiO2, and (e). PMA/KH570-SiO2.
Figure 5. Models for MD simulation of composite system: (a). PMA, (b). PMA/SiO2, (c). PMA/KH550-SiO2, (d). PMA/KH560-SiO2, and (e). PMA/KH570-SiO2.
Polymers 12 00170 g005
Figure 6. Temperature (a) and energy (b) of PMA/SiO2 nanocomposite at a temperature of 298K during the MD simulation.
Figure 6. Temperature (a) and energy (b) of PMA/SiO2 nanocomposite at a temperature of 298K during the MD simulation.
Polymers 12 00170 g006
Figure 7. (a) FT-IR spectrum of SiO2 and modified-SiO2, (b) UV curve of SiO2 and modified-SiO2.
Figure 7. (a) FT-IR spectrum of SiO2 and modified-SiO2, (b) UV curve of SiO2 and modified-SiO2.
Polymers 12 00170 g007
Figure 8. SEM images of nano-SiO2 (a). SiO2, (b). KH550-SiO2, (c). KH560-SiO2, (d). KH570-SiO2.
Figure 8. SEM images of nano-SiO2 (a). SiO2, (b). KH550-SiO2, (c). KH560-SiO2, (d). KH570-SiO2.
Polymers 12 00170 g008
Figure 9. TEM images: (a) PMA latex, (b) PMA/SiO2 composite, (c) PMA/KH550-SiO2 composite, (d) PMA/KH560-SiO2 composite, and (e) PMA/KH570-SiO2 composite (Red circles refer to latex particles, yellow circles refer to SiO2 particles).
Figure 9. TEM images: (a) PMA latex, (b) PMA/SiO2 composite, (c) PMA/KH550-SiO2 composite, (d) PMA/KH560-SiO2 composite, and (e) PMA/KH570-SiO2 composite (Red circles refer to latex particles, yellow circles refer to SiO2 particles).
Polymers 12 00170 g009
Figure 10. SEM of composite films: (a) PMA, (b) PMA/SiO2, (c) PMA/KH550-SiO2, (d) PMA/KH560-SiO2 and (e) PMA/KH570-SiO2).
Figure 10. SEM of composite films: (a) PMA, (b) PMA/SiO2, (c) PMA/KH550-SiO2, (d) PMA/KH560-SiO2 and (e) PMA/KH570-SiO2).
Polymers 12 00170 g010
Figure 11. Effect of fillers on mechanical properties of composite films: (a) Tensile strength, (b) Elongation at break (0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2 and 4. PMA/KH570-SiO2).
Figure 11. Effect of fillers on mechanical properties of composite films: (a) Tensile strength, (b) Elongation at break (0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2 and 4. PMA/KH570-SiO2).
Polymers 12 00170 g011
Figure 12. (a) Effects of fillers on water vapor permeability and (b) water resistance of composite films (0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2 and 4. PMA/KH570-SiO2).
Figure 12. (a) Effects of fillers on water vapor permeability and (b) water resistance of composite films (0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2 and 4. PMA/KH570-SiO2).
Polymers 12 00170 g012
Figure 13. TGA curve of composite films.
Figure 13. TGA curve of composite films.
Polymers 12 00170 g013
Figure 14. (a) MSD diagrams of H2O in the PMA, PMA/SiO2, PMA/KH550-SiO2, PMA/KH560-SiO2, and PMA/KH570-SiO2 films during the 40-ps MD duration, (b) The diffusion coefficient (D) of H2O in composite system: 0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2, and 4. PMA/KH570-SiO2.
Figure 14. (a) MSD diagrams of H2O in the PMA, PMA/SiO2, PMA/KH550-SiO2, PMA/KH560-SiO2, and PMA/KH570-SiO2 films during the 40-ps MD duration, (b) The diffusion coefficient (D) of H2O in composite system: 0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2, and 4. PMA/KH570-SiO2.
Polymers 12 00170 g014
Table 1. Thermal data of the composite films from TGA analysis.
Table 1. Thermal data of the composite films from TGA analysis.
SamplesWeight Loss Temperature (°C)Heat-Resistance Index a (°C)
T5T30
0318.43375.43172.79
1325.42374.92174.01
2327.88375.88174.77
3332.99374.49175.37
4329.98376.98175.51
0. PMA, 1. PMA/SiO2, 2. PMA/KH550-SiO2, 3. PMA/KH560-SiO2 and 4. PMA/KH570-SiO2. a Heat resistance index = 0.49[T5 + 0.6(T30 − T5)]; T5, T30 is the decomposing temperature at 5%, 30% weight loss, respectively.
Table 2. Binding energy of PMA/SiO2 and PMA/modified-SiO2 composites.
Table 2. Binding energy of PMA/SiO2 and PMA/modified-SiO2 composites.
SystemsEtotal (kcal/mol)EPMA (kcal/mol)ESiO2 (or Emodified-SiO2) (kcal/mol)Einter (kcal/mol)Ebinding (kcal/mol)
PMA10,043.9710,043.97---
PMA/SiO2−1812.3812,722.97−14,260.52−274.83274.83
PMA/KH550-SiO2−2323.9712,545.98−14,552.63−317.32317.32
PMA/KH560-SiO2−381.2013,785.56−13,810.49−356.27356.27
PMA/KH570-SiO2−1646.98----

Share and Cite

MDPI and ACS Style

Wu, Y.; Ma, J.; Liu, C.; Yan, H. Surface Modification Design for Improving the Strength and Water Vapor Permeability of Waterborne Polymer/SiO2 Composites: Molecular Simulation and Experimental Analyses. Polymers 2020, 12, 170. https://doi.org/10.3390/polym12010170

AMA Style

Wu Y, Ma J, Liu C, Yan H. Surface Modification Design for Improving the Strength and Water Vapor Permeability of Waterborne Polymer/SiO2 Composites: Molecular Simulation and Experimental Analyses. Polymers. 2020; 12(1):170. https://doi.org/10.3390/polym12010170

Chicago/Turabian Style

Wu, Yingke, Jianzhong Ma, Chao Liu, and Hongxia Yan. 2020. "Surface Modification Design for Improving the Strength and Water Vapor Permeability of Waterborne Polymer/SiO2 Composites: Molecular Simulation and Experimental Analyses" Polymers 12, no. 1: 170. https://doi.org/10.3390/polym12010170

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