You are on page 1of 9

Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

Contents lists available at ScienceDirect

Journal of the Mechanical Behavior of


Biomedical Materials
journal homepage: www.elsevier.com/locate/jmbbm

Investigating the properties and interaction mechanism of nano-silica in T


polyvinyl alcohol/polyacrylamide blends at an atomic level
⁎ ⁎
Qinghua Weia,b, Yanen Wanga, , Shuzhi Wanga, Yingfeng Zhanga, Xiongbiao Chenb,
a
Department of Industry Engineering, College of Mechanical Engineering, Northwestern Polytechnical University, Xi'an 710072, PR China
b
Department of Mechanical Engineering, College of Engineering, University of Saskatchewan, Saskatoon, Canada, S7N5A9

A R T I C L E I N F O A B S T R A C T

Keywords: The nano-silica can be incorporated into polymers for improved mechanical properties. Notably, the interaction
Nano-silica between nano-silica and polymer is of a microscopic phenomenon and thus, hard to observe and study by using
PVA/PAM blends experimental methods. Based on molecular dynamics, this paper presents a study on the properties and the
Nanocomposite interaction mechanism of nano-silica in the polyvinyl alcohol (PVA)/polyacrylamide (PAM) blends at an atomic
Molecular dynamics simulation
level. Specifically, six blends of PVA/PAM with varying concentrations of nano-silica (0–13 wt%) and two in-
Interaction mechanism
terfacial interaction models of polymers on the silica surface were designed and analyzed at an atomic level in
terms of concentration profile, mechanical properties, fractional free volume (FFV), dynamic properties of
polymers and X-ray diffraction patterns. The concentration profile results and micromorphologies of equilibrium
models suggest PAM molecular chains are easier to be adsorbed on the silica surface than PVA molecular chains
in blends. The incorporation of nano-silica into the PVA/PAM blends can increase the blend mechanical prop-
erties, densities, and semicrystalline character. Meanwhile, the FFV and the mobility of polymer chain decrease
with the silica concentration, which agrees with the results of mechanical properties, densities, and semi-
crystalline character. Our results also illustrate that an analysis of binding energies and pair correlation functions
(PCF) allows for the discovery of the interaction mechanism of nano-silica in PVA/PAM blends; and that hy-
drogen bond interactions between polar functional groups of polymer molecular chains and the hydroxyl groups
of the silica surface are involved in adsorption of the polymers on the silica surface, thus affecting the interaction
mechanism of nano-silica in PVA/PAM blend systems.

1. Introduction Sureshkumar, 2014; Wei et al., 2017a) so as to obtain the combined


advantages of PVA and PAM. Meanwhile, it is noted that due to the
Polyvinyl alcohol (PVA) is a water-soluble synthetic polymer, which brittleness and poor mechanical properties of PAM, the addition of PAM
is commercially available in a form of partial or complete hydrolysis of can weaken the original mechanical properties of PVA hydrogel (Wei
poly(vinyl acetate). Due to its hydrophilicity, biodegradability, bio- et al., 2017a; El-Zawawy et al., 2012), which limits the application of
compatibility, and high mechanical strength, PVA has been widely used PVA/PAM blends in some tissue engineering application, such as bone
in potential drug delivery systems and scaffolds for various tissue en- and cartilage repair, where the mechanical property is a critical issue
gineering applications (Anastasia et al., 2016; Morales-Hurtado et al., (Olubamiji et al., 2016; Chen, 2014; Little et al., 2011).
2015; Jiang et al., 2011) PVA, however, has some demerits, such as lack A large number of studies show that adding the nanoparticles into
of biological activity, inability to withstand complex loads, and un- polymers can significantly improve the mechanical properties (e. g.,
favorable properties for cell adhesion and growth (Choi et al., 2012), strength, stiffness, elastic modulus) of polymers (Qiao et al., 2016;
thus limiting its application to tissue engineering. Polyacrylamide Gomez et al., 2016; Zhao et al., 2016). Due to its properties of non-
(PAM) is a type of water-soluble polymer formed from acryl amide toxicity, nano-silica has been widely used as an inorganic nanomaterial
monomers; and its hydrogel possesses favorable characteristics for to improve the mechanical properties as well as thermodynamic prop-
tissue engineering, such as three-dimensional network structure, non- erties, chemical stability and water resistance (Dil et al., 2016; Zhou
toxic side effects, good biocompatibility and biological activity (Leng et al., 2016; Griffin et al., 2016). Taheri-Behrooz et al. (Taheri-Behrooz
et al., 2011; Suriano et al., 2014; Labarre et al., 2002). Therefore, PAM et al., 2015) investigated the addition of nano-silica in phenolic resin,
has been always used for the modification of PVA (Patel and with the results illustrating that the resultant Young's modulus and


Corresponding authors.
E-mail addresses: wangyanen@126.com (Y. Wang), xbc719@mail.usask.ca (X. Chen).

http://dx.doi.org/10.1016/j.jmbbm.2017.08.027
Received 19 July 2017; Received in revised form 18 August 2017; Accepted 22 August 2017
Available online 24 August 2017
1751-6161/ © 2017 Elsevier Ltd. All rights reserved.
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

hardness steadily increased as the weight percent of silica synthesizing nano-silica/polymers composites for various applications.
nanoparticles. Malaki et al. (2016) studied the erosion resistance, ad-
hesive strength, micro-hardness and weathering stability of the acrylic- 2. Materials models and simulation methods
based polyurethane nano-composites and found that they could be re-
inforced with the addition of nano-silica particles. Sowjanya et al. 2.1. Polymer model
(2013) fabricated biocomposite scaffolds from the blends of chitosan,
alginate and nano-silica by using a freeze drying method; and they An appropriate molecular chain length in polymers is critically
found the scaffolds possessed a well-defined porous architecture with important to the model development, thus contributing the accuracy of
pore sizes varying from 20 to 100 µm suitable for cell infiltration and the simulation results. A short molecular chain length may not re-
that the presence of nano-silica in the scaffolds increased the protein present the real polymers appropriately, while a long molecular chain
adsorption and reduced the swelling occurred. Besides, the silica pro- length may lead to difficulties for computer simulations. In the present
cessed in this scaffold also assisted with osteoblast cell proliferation, study, each polymer molecular chain with the same molecular weight
and the collagen fiber network interconnecting the ß-TCP/silica gran- was constructed, so that the results of models in simulations can com-
ules yielded a rigid, fleece-like mold (Wilmowsky et al., 2008; Ghanaati pare with each other. In this study a PVA molecular chain with 50 re-
et al., 2010). peat units and a PAM molecular chain with 31 repeat units were con-
As discussed above, adding nano-silica particles into PVA/PAM structed according to the molecular structural formula (Fig. S1) as per
blend composite can not only improve mechanical properties, but also our previous study (Wei et al., 2016b, 2016c). The all simulation
promote proliferation of cells in blend. These advantages make PVA/ models were constructed by using the Materials Studio software
PAM/nano-silica composite suitable for bone tissue engineering. To this (Accelrys Inc., 2010).
end, the knowledge on the micro properties and interaction mechanism
of nano-silica in PVA/PAM blend is of importance to prepare the PVA/
2.2. Nano-silica model
PAM/nano-silica composites. Notably, the interaction between nano-
silica and polymer is of a microscopic phenomenon and thus, is hard to
The crystal structure of a-quartz (silica, space group P3321) was
observe and study by using experimental methods. As a result, the in-
adopted from the Cambridge Structural Database with the lattice
teraction mechanism of nano-silica in PVA/PAM blend has been rarely
parameters as follows: a = b = 0.491 nm, c = 0.5402 nm, α = β =
investigated and elucidated.
90°, γ = 120°, and its density is 2.648 g/cm3 (Fig. S2a). Considering the
Molecular dynamics (MD) simulations has been illustrated pro-
computer efficiency and the specific surface area, a spherical silica
mising for investigating material properties and interfacial interactions nanoparticle with a radius of 6 Å was constructed, without considera-
of inorganic fillers and polymer matrix in an atomic scale (Lammers tion of surface functionalization or covalent grafting (Fig. S2b). The
et al., 2017; Roussou and Karatasos, 2016; Hagita et al., 2016), thus unsaturated boundary effect was avoided by adding hydrogen atoms to
providing the microscopic information or details of molecular interac- the unsaturated oxygen atoms and hydroxyl groups to the unsaturated
tions inside materials. Base on MD simulations, Lai et al. (2014) in- silicon atoms of the silica particle surface (Fig. S2c).
vestigated the mechanical characteristics of the interfaces between os-
teopontin and hydroxyapatite and they found that the interfacial
2.3. PVA/PAM/silica blend models
mechanical behavior was governed by the electrostatic attraction be-
tween acidic amino acid residues in OPN and calcium in HA. Boulet and
Before constructing the blend models, the PVA chain, PAM chain
Coveney (2004) and Gardebien et al. (2005) also employed the ex-
and the spherical silica nanoparticle were all optimized by using the
perimental and MD methods to study the properties of the montmor-
geometry optimization method (as detailed below), to ensure the en-
illonite polymer nanocomposite, which helped understand different
ergies of the models built were minimized and following the stable
aspects of structures and properties of polymer nanocomposite. Sahu
configurations of PVA, PAM and silica nanoparticle were obtained. To
and Anup (2016) applied MD simulations to investigate the effect of the
eliminate the effects of polymer composition ratio on calculation re-
structural arrangement of reinforcements on the mechanical properties
sults, the amount ratio of PVA and PAM chains in each model was 1:1.
of a nano-composite under transverse loading. In this regard, we have
Furthermore, to purely study the interaction mechanism of nano-silica
also successfully employed MD simulations to investigate the properties
and polymers and prevent any finite size effect of nanocomposites and
of biopolymer blends (Wei et al., 2016a, 2016b, 2017a; Wang et al.,
account for any particle effect at the bulk level, only one silica nano-
2017) and interfacial interaction (Wei et al., 2016c, 2017b).
particle was embedded at the center of the blend unit cell and the
Inspired by the above studies, we employed the MD simulation
periodic boundary condition was used. Thus, all the simulation models
method to investigate into the micro structures and properties of PVA/
used here were constructed under an ideal condition without con-
PAM/nano-silica blend composites, as presented in this paper. We also
sidering the agglomeration of nano-silica, and the blends with different
present the discovery of interaction mechanism between nano-silica
silica contents were constructed by changing the number of polymer
and polymers in blend system from the perspective of molecular in-
chains in the models. In our research, six blend models with different
teraction. This work not only elucidates the reinforcing mechanism of
compositions of PVA/PAM/silica were constructed, and the ratios of
nano-silica in polymeric matrix, but also provides theoretical basis for
molecular number in models were 7/7/0, 7/7/1, 6/6/1, 5/5/1, 4/4/1,

Table 1
Parameters of simulated blend models.

Cell component Weight fraction (wt%) Cell after refinement

PVA PAM Silica Cell length (nm) Density (g/cm3) 2θ (°) d-spacing (nm)

7PVA/7PAM/0silica 50.0 50.0 0.0 3.412 1.295 18.65 0.476


7PVA/7PAM/1silica 47.0 47.0 6.0 3.459 1.317 18.95 0.468
6PVA/6PAM/1silica 46.5 46.5 7.0 3.288 1.329 19.15 0.463
5PVA/5PAM/1silica 45.9 45.9 8.2 3.091 1.350 19.35 0.459
4PVA/4PAM/1silica 45.0 45.0 10.0 2.876 1.368 19.45 0.456
3PVA/3PAM/1silica 43.5 43.5 13.0 2.624 1.397 19.65 0.452

530
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

Fig. 1. Final equilibrium structures of PVA/PAM blend models with different silica contents. (a) 7PVA/7PAM (0.0 wt% silica content); (b) 7PVA/7PAM/1silica (6.0 wt% silica content);
(c) 6PVA/6PAM/1silica (6.9 wt% silica content); (d) 5PVA/5PAM/1silica (8.2 wt% silica content); (e) 4PVA/4PAM/1silica (10.0 wt% silica content); (f) 3PVA/3PAM/1silica (13.0 wt%
silica content).

3/3/1. To ensure the molecular chains have sufficient space for re- the equilibrium systems under the NPT ensemble (P = 1 bar) until the
laxation and to avoid chain winding and overlapping, the initial den- density and energy of the system no longer changed (Fig. S3 shows the
sities of the blends were all set to 0.3 g/cm3. Table 1 lists the detailed time evolution density profiles during MD simulation under the NPT
parameters of all simulated blend amorphous cells at the temperature of ensemble of 6PVA/6PAM/1silica blend model); and last an additional
298 K. MD simulation under the NVE ensemble was carried out on the above
systems for 40 ps, and the trajectory frames of the last 30 ps were used
for results analysis. For the single polymer molecular chains and nano-
2.4. MD simulations silica, only the first step of MD simulation above needs to be performed.
All the parameters used of MD simulations were selected as follows.
The geometry optimization was conducted on all the initial models The Verlet velocity integration method was used with a time step of
of polymer chain, silica nanoparticle, and blend models prior to their 1 fs, the Anderson thermostat and barostat (Andersen, 1980) were used
use in simulations. Geometry optimizations were performed by utilizing to maintain the temperature and pressure for all NVT and NPT simu-
the smart minimization method, the convergence tolerance was applied lations. The non-bonded van der Waals and electrostatic interactions
using customized quality with an energy convergence of 4.186 × were truncated using a group based cut off distance of 1.25 nm. The
10−4 kJ/mol and displacement of 5 × 10−6 nm. The number of spline width was 0.1 nm, and the buffer width was 0.05 nm. The co-
iterations was set as 5000 and the non-bonded van der Waals and ordinates of the system were collected every 500 time steps. Trajectory
electrostatic interactions were truncated using atom-based at the fine frames were captured during the production run, and the data from the
level of quality were applied for simulations. The energies of the final 30 ps was used for analysis.
amorphous cells were minimized by using the smart minimization al-
gorithm to remove unfavorable interactions and attain the lowest en-
ergy state. 2.5. System balance
All the MD simulations were performed after geometry optimization
and under the COMPASS (Condensed-Phase Optimized Molecular For a molecular dynamics simulation, an equilibrium calculation
Potentials for Atomistic Simulation Studies) force field. This force field model is of vital importance, which determines the accuracy and re-
is the first ab initio force field based on the Polymer Consistent Force liability of the simulation results. Generally, there are mainly two
Field (PCFF) (Sun, 1998), its parameterization and validation have been methods to determine the equilibrium: (1) the temperature equilibrium,
done using condensed-phase properties and various ab initio and em- at which the standard deviation is less than 10%; (2) the energy equi-
pirical data of various molecules. Thus this force field can accurately librium, at which the energy is constant or fluctuates up and down
predict the various properties for a wide range of materials especially along the constant value (Gupta et al., 2011). In the present study, all
for polymer molecules (Wescott et al., 2006). The MD simulations of the results were calculated under the condition of system equilibrium.
blend amorphous cells involve the following three steps. First, MD si- As showed in Fig. S4, the temperature and energy of the system (6PVA/
mulation was conducted under the NVT ensemble for 100 ps to release 6PAM/1silica) met the above equilibrium standard, indicating that the
any possible tension; secondly, the MD simulation was performed on system reached equilibrium after the MD simulation.

531
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

After the MD simulation, the final equilibrium models of PVA/PAM


blends with different silica contents at a temperature of 298 K were
obtained, as shown in Fig. 1 and their final densities were present in
Table 1. To better distinguish the components in blend models, the PVA
molecular chains were represented in green, and the PAM molecular
chains were represented in purple.

2.6. Model development and simulations of the interfacial interaction

To further investigate into the interaction mechanisms of polymer


molecular chains and silica in PVA/PAM/silica blend, here two inter-
facial interaction models of PVA and PAM molecular chains on silica
crystal surface were constructed and simulated. The most re-
presentative silica (110) surface was selected to build the super cell,
whose dimensional parameters were 2.161 × 3.402 × 1.436 nm, α =
β = γ = 90°. Besides, the unsaturated atoms on the surface of silica
super cell were treated by adding hydrogen atoms to the unsaturated
oxygen atoms and hydroxyl groups to the unsaturated silicon atoms. In
each interaction model of polymer/silica, only one molecular chain
built above was contained. A vacuum slab with a thickness of 3 nm was
added to the polymer layer to ensure the polymer chain only interacts
with one side of the silica surface. Geometry optimizations and MD
simulations were conducted on the interaction models once con-
structed. In this section, MD simulation involved two steps: (1) MD
simulation under NVT ensemble for 60 ps at 298 K to obtain the equi-
librium models (Fig. S5). (2) An additional 30 ps MD simulation under
NVE ensemble at 298 K was performed.

3. Results and discussion


Fig. 3. Relative concentration distributions of different components in (a) 7PVA/7PAM
(0.0 wt% silica content) and (b) 7PVA/7PAM/1silica (6.0 wt% silica content) blend
3.1. Concentration distributions of compositions models along Z axis direction.

After reaching equilibrium, an interesting phenomenon can be ob-


served from each equilibrium blend model in Fig. 1. The polymer chains density decreases outward therefrom. The polymer density in the area
tend to gather around the nano-silica particle, and relative to the PVA around the spherical silica particle is significantly higher than further
molecular chains (the green), the probability of PAM molecular chains away from the silica nanoparticle. Moreover, the concentration profile
(the purple) appearing around the silica nanoparticle is higher. From map of PAM (Fig. 2b) also shows the density decreases outward
the distribution of PVA and PAM molecular chains in blend models therefrom, and the greatest density appears near the surface of nano-
(Fig. 1), the nano-silica particle was almost wrapped by the PAM mo- silica particle, while the opposite pattern appears in the concentration
lecular chains, and the larger number of polymer chains in blend profile map of PVA (Fig. 2c). This result is consistent with the adsorp-
system, the more obvious of this phenomenon. To further investigate tion behavior of polymer molecular chains on the surface of the sphe-
density distributions of different components in blends, the blends of rical nano-silica particle in the final equilibrium models of PVA/PAM
7PVA/7PAM and 7PVA/7PAM/1silica were chosen to analyze the blends (Fig. 1). Additionally, the comparison of the relative con-
concentration profile. Fig. 2 shows the concentration profile maps of (a) centration distributions among polymers in 7PVA/7PAM and 7PVA/
the whole nanocomposite, (b) PAM molecules and (c) PVA molecules in 7PAM/1silica blend models illustrates that the addition of nano-silica
7PVA/7PAM/1silica blend model. And Fig. Fig. 3 show the con- has a great influence on the distribution of the polymers. The relative
centration profiles of different components in blend models of (a) concentration distributions of PVA and PAM in PVA/PAM blend model
7PVA/7PAM (0.0 wt% silica content) and (b) 7PVA/7PAM/1silica (Fig. 3a) indicate the polymers in this system are distributed evenly,
(6.0 wt% silica content) along Z axis direction. which shows a agreement with their good miscibility. But after adding
In Fig. 2a, the blue in the center represents the silica particle and the the nano-silica into the PVA/PAM blend, the relative concentration

Fig. 2. Concentration profile maps of different components in 7PVA/7PAM/1silica blend system, and each color band corresponds to a range of field values. (a) The whole nanocomposite
which contains all components in system; (b) PAM component without PVA and nano-silica; (c) PVA component without PAM and nano-silica.

532
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

distributions of polymers in system (Fig. 3b) changed greatly. Closed to fracture toughness of a material is determined by K, with a higher value
the surface of the nano-silica particle, the relative concentration of PAM indicating greater fracture resistance (Jin et al., 2000). Cauchy pressure
is greater than that of PVA. This means a stronger interaction force (C12-C44) is often used to evaluate the ductility of a material, a more
exists between the PAM molecules versus PVA molecules and the nano- positive Cauchy pressure corresponds to a better ductility (Weiner,
silica, which is the primary reason affecting their distribution. 2002).
The previous study (Sakineh et al., 2015) illustrates that the addi-
3.2. Mechanical properties tion of nano-silica can improve the mechanical properties of polymer
blends. As inspired by this, the present study the nano-silica was used to
It is noted that the mechanical properties mentioned here refers to improve the mechanical properties defect of PVA/PAM blends. Notably,
the micro mechanical properties, which are obtained by analyzing the the calculated Poisson's ratios of blends here consistently fluctuate be-
trajectories of atoms. Besides, they are different from the macro ex- tween 0.29 and 0.32, this is within the range of Poisson's ratio of plastic
perimental mechanical properties. Here, the micro mechanical prop- (0.2–0.4), suggesting that the blends possess the properties similar to
erties were calculated by means of the static model, in which the fol- plastics. However, the Cauchy pressure value of the blends decreases
lowing equation was used to evaluate the internal stress tensor (σ ) at with the increase of silica content, but it is always positive, suggesting
the atomic level (Swenson, 1983). that the blends with nano-silica generally have a good ductility al-
N
though less than that of pure PVA/PAM blend and that the addition of
1 ⎡
σ=− ∑ mi (Vi , ViT ) ⎤⎥
V0 ⎢ i = 1
nano-silica weakens the ductility of blends. Finally, comparing the two
⎣ ⎦ (1) group elastic coefficient values of C11, C22, C33 and C44, C55, C66 shows
that the addition of nano-silica can change the properties of isotropy,
where mi and Vi represent the quality and speed of atoms, respectively,
with the anisotropy of the system becoming correspondingly more ob-
and V0 is volume without distortion.
vious.
The elastic coefficient matrix is obtained by solving the first deri-
To verify the reliability of the computed results here, PVA (17–99,
vative of the corresponding stress and strain. The blend systems used in
MW 80,000), PAM (Analytical reagent, MW 200,000) and spherical
the present study are assumed to be the isotropic material, for which
nano-silica (diameter 10–15 nm) were used to prepare the PVA/PAM
the static elastic properties, i.e., tensile (Young's) modulus (E), bulk
blend membranes with different nano-silica contents by the method of
modulus (K), and shear modulus (G), and Poisson's ratio (γ ) were cal-
vacuum evaporation, then their tensile strength were obtained by the
culated as follows:
electronic universal testing machine. Fig. 4 shows the results of ex-
μ (3λ + 2μ) 2 λ perimental tensile strength and simulation tensile modulus, from which
E= , K = λ + μ, G = μ, γ =
λ+μ 3 2(λ + μ) (2) we can know both the experimental tensile strength and simulation
tensile modulus increase with the silica content, the different tendency
where μ and λ are the Lamé coefficients, which are calculated from the
was mainly caused by the small scale effect and the aggregation be-
elastic coefficients according to the statistical mechanics of elasticity
havior of nano-silica. Anyway, this conclusion can verify the reliability
(Watt et al., 1976).
of simulation results to some extent.
1 2
λ= (C11 + C22 + C33) − (C44 + C55 + C66)
3 3 (3)
3.3. Fractional free volume
1
μ= (C44 + C55 + C66)
3 (4) The void space within the molecules is termed as the free volume,
which can be measured by fractional free volume (FFV). In our simu-
According to the principles of static mechanics, the mechanical
lations, atoms were represented by the hard spheres with van der Waals
properties of PVA/PAM blends with different silica contents were ob-
radius as per the previous study (Pan et al., 2008), and the method of
tained by calculating the trajectories of the equilibrium models at the
Connolly surface was adopted for use. The FFV is evaluated by the
temperature of 298 K. The elastic stiffness constants and the en-
following equation (Peng et al., 2011).
gineering modulus are listed in Table 2.
The elastic modulus of a material is a measurement of its rigidity. Vf
Generally, the hardness of a material is closely related to E and G; and FFV = × 100%
Vf + Vo (5)
larger values of E and G suggests a stronger rigidity and hardness. The
where Vf is the free volume and Vo is the volume occupied by the
Table 2
Mechanical properties of PAM/PVA blends with different compositions at 298 K (GPa).

Systems 7/7/0 7/7/1 6/6/1 5/5/1 4/4/1 3/3/1

C11 10.60 12.38 13.41 14.09 15.39 16.22


C22 11.65 12.64 14.28 16.50 13.23 19.91
C33 10.26 11.88 12.47 13.55 17.19 14.78
C44 2.42 2.59 3.07 3.27 3.90 5.23
C55 3.41 3.74 3.84 3.98 4.26 4.11
C66 3.66 3.90 4.15 4.26 4.02 4.94
C12 5.25 5.20 5.19 5.25 5.76 6.44
C13 5.20 5.33 6.91 7.67 6.16 5.98
C23 5.39 5.70 6.21 7.11 10.25 9.65
C15 0.27 0.32 −0.14 −1.65 −0.27 −1.42
C25 0.13 −0.48 0.62 0.20 −0.62 0.42
C35 −0.16 0.36 −0.71 −1.14 0.12 0.36

Tensile modulus 8.16 8.92 9.67 10.12 10.71 12.42


Bulk modulus 6.56 7.75 8.46 9.27 9.86 10.62
Shear Modulus 3.16 3.41 3.69 3.84 4.06 4.76
Poisson ratio 0.29 0.31 0.31 0.32 0.32 0.31
Fig. 4. Results of experimental tensile strength and simulation tensile modulus for PVA/
C12-C44 2.83 2.61 2.12 1.98 1.86 1.21
PAM/silica with different silica contents.

533
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

As seen from Fig. 6, the MSD-t curves exhibit linear as expected, and
the slopes of the MSD-t curves decrease with the addition of silica
content, which indicates the mobility of the PAM polymer chains be-
come worse. This probably attributes to the interaction force between
polymer and the nano-silica particle increasing with the addition of
silica content. This result is also in agreement with the FFV results.
To further examine the effect of nano-silica on the flexibility of the
polymer chains, the PAM chain radii of gyration (Rg) in 7PVA/7PAM
and 7PVA/7PAM/1silica blends were analyzed and compared to each
other. The radius of gyration is a parameter commonly used to describe
the size of a flexible molecular chain by the following equation (Wang
et al., 2015).

∑i mi ri2
Rg =
∑i mi (7)
Fig. 5. FFVs of PVA/PAM blends at different probe radii.
where mi is the mass of Atom i, and ri is the position of Atom i with
respect to the center of mass of the molecule. In general, the more
polymeric material. Fig. 5 shows the distributions of the FFVs for the flexible of the polymer chain, the smaller Rg of the polymer chain is.
different blends, in which the value of probe radius, r, varies from 0.00 Fig. S6 shows the radii of gyration for PAM and PVA chains in the pure
to 0.30 nm and Δr is set as 0.02 nm. PVA/PAM and PVA/PAM/silica blends, in which there were seven PAM
It is seen from Fig. 5 that the FFVs in blends with different silica and seven PVA molecular chains, respectively. It is noted that the radii
contents are all decreased with the increase of probe radius. In other of gyration for the most of PAM chains in PVA/PAM/silica blend (Fig.
words, the bigger the solvent molecule causes the smaller FFV. It is also S6b) are much bigger than those in PVA/PAM blend (Fig. S6a) at the
seen that for the probe radius, the FFVs decrease with the addition of same temperature. This indicates that nano-silica particle can sig-
silica content, suggesting that the addition of silica can enhance the nificantly weakened the flexibility of PAM molecular chains, which can
intermolecular interaction, thus decreasing the value of FFV. In other also be found in the radius of gyration results of PVA molecular chains.
words, the addition of silica makes the system denser, which is con- This result is also in agreement with the analytical results of MSD and
sistent with the observation from both the calculated density results in FFV. Therefore, the interaction between nano-silica particle and poly-
Table 1 and the mechanical property results in Table 2. The main mers plays a fundamental role in the structural morphology of polymer
reason behind this is that with the enhanced interaction between silica systems.
and polymers, the distance between molecules is reduced, thus making
the system denser.
3.5. Simulated X-ray diffraction patterns

3.4. Dynamic properties of polymer chains X-ray diffraction (XRD) is a method to investigate material micro-
structure, and is usually used to evaluate the crystal characteristics of
The mean square displacement (MSD), which is associated with the systems. The simulated scattering intensity curve, I(Q), for each system
temperature and pressure, is one of the critical parameters to determine is related to the radial distribution function through a Fourier transform
the mobility of the polymer chains. Therefore, the MSD can used to operation by (Jawalkar and Aminabhavi, 2007)
characterize the effect of silica particle on the mobility of the polymer
f j fk (sin Qr jk ) 4π sin θ
chains in blends, as per the previous study (Luo and Jiang, 2010). I (Q) = ∑∑ , Q=
j k
Qr jk λ (8)
MSD = < (ri (t ) − ri (0))2> (6)
where λ is the wavelength, θ is the scattering angle, rjk is the distance
where ri(t) and ri(0) are the positions of Atom i at time t and 0, re-
between two adjacent atoms, and f is the atomic diffraction factor. In
spectively. The bracket denotes the ensemble average, which was ob-
general, the maximum peak in the diffraction pattern is noteworthy
tained from averaging over all molecules and all time origins t = 0.
because it is associated with the Bragg equation and is represented by
Here, all the atoms of a PAM chain were selected to study the mobility
of the PAM chains in different blends at 298 K, and the MSDs of PAM λ
d=
molecular chains are presented in Fig. 6. 2 sin θ (9)
The d-spacing value is representative of inter-chain distances be-
tween polymer backbones. The XRD pattern was obtained by employing
the Forcite module. The wavelength was set to 0.15418 nm and the
angle of diffraction was varied from 0 to 45° using a step size of 0.05°.
The XRD patterns of the blends are shown in Fig. 7.
By comparing the integrated intensity of the background pattern to
that of the sharp peaks, the crystallinity of the system can be de-
termined. In contrast to a crystalline pattern consisting of a series of
sharp peaks, amorphous materials produce a broad background signal
(Shariatinia et al., 2016). For polymers, an ordered crystallite can be
formed by folding the molecule, possesses a semicrystalline character.
The XRD patterns in Fig. 7 display some relatively sharp peaks with
moderate intensity below 20° for all of the blends, and a very sharp
peak is observed near 2θ = 3° and 2θ = 7° in the XRD patterns of all the
blends, which demonstrates that all of the blends are of a semicrystal-
line structure. The intensity of peaks around 2θ=3° and 2θ=7° also
Fig. 6. MSDs of PAM molecular chains in different blends at 298 K.
increase markedly with increasing silica content, which indicates the

534
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

molecular chain in Fig. S5a folds on the silica surface and cannot fully
interact with each other. This result can well explain why the prob-
ability of PAM molecular chains appearing around the silica nano-
particle is higher than PVA. Additionally, the hydrogen bond energy
and interaction energy of two polymer repeat units are nearly equal,
which indicates the interactions between polymers and silica mainly
attribute to hydrogen bonds.

3.6.2. Interfacial interaction mechanism


The binding energy results suggest the polymers interact with silica
in blends mainly by hydrogen bonds and that the hydrogen bonds may
exist between functional groups or polar atoms in polymers and atoms
on the silica surface. Therefore, to further discover the interaction
mechanism between PVA, PAM, and the silica surface, the pair corre-
lation functions (PCF) of PVA/silica and PAM/silica interaction models
Fig. 7. Calculated X-ray intensity versus 2θ of PVA/PAM blends with different silica were studied by analyzing their balancing trajectory files. For PCF, the
content.
probability of finding another particle at a distance r from a specific
particle characterizes the microstructure and the interaction between
addition of silica can enhance the semicrystalline character of the non-bonding atoms. This probability, denoted by g(r), is given by
blends. A pure PVA/PAM blend exhibits a semicrystalline structure
nB N
with three sharp peaks at 2θ = 16.8°, 18.5° and 21.5°. Therefore, the gA − B (r ) = ⎛ ⎞/⎛ B ⎞
peak near 2θ = 16–22° present in the XRD patterns of all of our systems ⎝ 4πr 2dr ⎠ ⎝ v ⎠ (11)
was used to calculate the inter-chain distance with Eq. (9). If there where nB is the number of B atoms that surround A atoms at a distance
existed more than one peak, the average of the peaks was used. The r, NB is the total number of B atoms, and V is the volume of the entire
calculated results of d-spacing (Table 1) show that the distance between system.
polymer molecular chains decreases with the addition of silica. The Figs. 8 and 9 show the PCF g(r) values of functional groups and
decrease in d-spacing can be attributed to the enhanced intermolecular polar atoms that exert stronger interaction forces in PVA/silica and
interactions between the nano-silica and the polymers, which has the PAM/silica interfacial models, respectively. In these PCFs, OH(PVA)
same pattern with the FFV and mechanical property analyses. denotes the hydroxyl groups (-OH) of the PVA molecular chain and H
(PVA) represents the hydrogen atoms connecting with the carbon
3.6. Interfacial analysis backbone. Similarly, in a PAM molecular chain, the oxygen of the
carbonyl groups (-C˭O), amino groups (-NH2), and hydrogen atoms
3.6.1. Interfacial binding energy directly connecting with the carbon atoms were marked as O (PAM),
The size of binding energies between polymers and silica crystal- NH2(PAM), and H(PAM), respectively. For the silica model, OH(SiO2),
lographic surfaces can reflect their interaction strength. Generally, the O(SiO2), and Si(SiO2) represent the hydroxyl groups (-OH) on the SiO2
larger the binding energy, the stronger the interaction force between surface and the oxygen atoms and silicon atoms of the SiO2 cell, re-
polymers and silica is. Therefore, comparing the binding energies be- spectively.
tween different compositions will help us understand the interaction Generally, intermolecular interactions involve chemical bonding,
mechanism and better explain the density distribution on the silica hydrogen bonding, and van der Waals forces. A peak value below
surface. The average binding energies of each polymer repeat unit for 0.35 nm indicates chemical bonds and hydrogen bonds, and values
PVA/silica and PAM/silica interaction models can be calculated by: higher than 0.35 nm indicate van der Waals interactions (Yang et al.,
2015).
Etotal − EPolymer − Esilica
Ebind = −Eint er = − Fig. 8 shows the PCFs for the PVA/silica interfacial interaction.
N (10)
There are two peak values of g(r) for OH(PVA)~OH(SiO2) appearing at
where Etotal is the energy of the polymer/silica, Epolymer is the energy of r = 0.2 nm and r = 0.3 nm, which predict that the hydroxyl groups
polymer, and Esilica is the energy of the silica surface, N is the number of (-OH) in PVA interact with the hydroxyl groups on the silica surface
repeat units for a polymer chain. All the energies in this research were mainly by the strong hydrogen bonding. Two peak values of g(r) for OH
calculated by using the Dreiding force field, and by this force field the (PVA)~Si(SiO2) occur at r = 0.4 nm and r = 0.65 nm, which are
hydrogen bond energies of each part can be obtained. The average within the effective range of van der Waals forces. Similarly, the g(r) for
binding energies and hydrogen bond energies between each PVA and H(PVA)~OH(SiO2) features two peak values at r = 0.25 nm and r =
PAM repeat units and the silica surface are given in Table 3. 0.48 nm; the value at r = 0.25 nm is the smaller of the two, which
It is seen that the average binding energy of each PVA repeat unit is indicates only a small portion of the H(PVA) atoms can interact with the
smaller than that of PAM repeat unit, which suggests the interfacial hydroxyl groups on the silica surface by hydrogen bonding, and most of
interaction force between a PVA repeat unit and silica is smaller than them form van der Waals interactions. The other interactions of func-
that between a PAM repeat unit and silica. This conclusion can also be tional groups and atoms in the PVA/silica interfacial model are all weak
reflected by observing their equilibrium interfacial interaction models van der Waals interactions because their peak g(r) values are > 0.5 nm.
(Fig. S5). The PAM molecular chain in Fig. S5b spreads out on the silica These results above are consistent with the FTIR measurement results
surface and they fully interact with each other, while the PVA obtained by Chrissafis et al. (2008).

Table 3
Binding energies and hydrogen bond energies for each PVA and PAM repeat unit on silica surface (kJ/mol).

System Etotal Epolymer Esilica Einter Ebind Ehydrogen bond

PVA/silica − 206,872.08 −3388.73 − 202,731.62 −15.03 15.03 −14.55


PAM/silica − 207,789.48 −4219.15 − 202,706.59 −27.86 27.86 −26.80

535
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

PCFs above, we can determine the strength of each type of the hydrogen
bonds has the order of O(PAM)…OH(SiO2) > NH2(PAM)…OH
(SiO2) > OH(PVA)…OH(SiO2) > H(PAM)…OH(SiO2) > H(PVA)…OH
(SiO2). This conclusion shows a good agreement with the analysis re-
sults of binding energies between polymers and silica interface and
hence is why PAM molecular chains more easily absorb on the silica
surface.

4. Conclusions

In this research, six models of PVA/PAM/nano-silica blends of


varying concentrations were constructed and used to investigate into
the effects of nano-silica on the properties of, and the interaction in-
volved in, the blends. At each concentration, the mechanical properties,
FFVs, dynamic properties of polymers and XRD patterns were system-
Fig. 8. PCFs of the PVA/silica interfacial interaction.
atically examined. The results of micromorphologies from the equili-
brium models showed PAM molecular chains are easier to be adsorbed
on the silica surface than PVA molecular chains in blends. The ex-
amination of mechanical properties suggested that the addition of silica
into the polymer blend can significantly improve the stiffness of
polymer blend, meanwhile affecting the properties of ductility and
anisotropy. Due to the interaction between the polymers and nano-si-
lica, the FFV and the mobility of polymer chain decreased as the con-
centration of silica increased, suggesting that the systems became
denser. Also, the results of XRD patterns illustrated that the addition of
nano-silica enhanced the semicrystalline character of the blends and
decreased the distance between polymer molecular chains, which
concurred with the results of FFV and mechanical property. Finally, two
interfacial interaction models featuring PVA/silica and PAM/silica were
designed to explore the interaction mechanisms and explain how nano-
silica particles influence the properties of PVA/PAM blends. The results
of binding energies and PCFs indicated hydrogen bond interactions
Fig. 9. PCFs of PAM/silica interfacial interaction. between polar functional groups of polymer molecular chains and the
hydroxyl groups of silica surface are implicated in polymer adsorption
onto the nano-silica surface. The interaction between PAM and the si-
As Fig. 9 indicates, the PFC g(r) values for O(PAM)~OH(SiO2)
lica surface is stronger than that between PVA and the silica surface,
contain two obvious peaks at r = 0.19 nm and r = 0.28 nm, which
which agrees with the observed concentration profiles of different
means stronger hydrogen bonds form between the oxygen from the
compositions in blend systems. The results of this research not only aid
carbonyl groups in PAM and the hydroxyl groups on the silica surface.
in understanding the microscopic structure and interaction mechanism
Moreover, the g(r) for O(PAM)~Si(SiO2) reaches two peak values at r =
of nano-silica in PVA/PAM blends, but also provide theoretical gui-
0.36 nm and r = 0.48 nm, which is within the range of van der Waals
dance for the in-depth study and design of PVA/PAM/silica blend na-
interactions. This demonstrates a strong van der Waals interaction can
nocomposites.
form between the oxygen of the carbonyl groups of the PAM and the
silicon atoms on the silica surface. The value of g(r) for NH2(PAM)~OH
Acknowledgements
(SiO2) reaches a significant peak value close to r = 0.29–0.32 nm,
which means the amino groups of the PAM interact with the hydroxyl
This project was sponsored by the China Scholarship Council (CSC,
groups of the silica surface mainly by hydrogen bonds. Two peak values
201606290095), the National Natural Science Foundation of China,
of g(r) for NH2(PAM)~Si(SiO2) occur at r = 0.45 nm and r = 0.65 nm,
(Grant No. 51175432), the Innovation Platform of Biofabrication
which are within the effective range of van der Waals forces. Similarly,
(Grant No. 17SF0002), the Key Industrial Science and technology pro-
The PCF g(r) for H(PAM)~OH(SiO2) reaches peak values at r =
jects of Shaanxi (Grant No. 2015GY047), and Natural Sciences and
0.28 nm and r = 0.58 nm, and the peak value at r = 0.58 nm is larger
Engineering Research Council of Canada (NSERC RGPIN-2014-05648).
than that at r = 0.28 nm. This means a small portion of the H(PAM)
atoms can interact with the hydroxyl groups on the silica surface by
Appendix A. Supplementary material
hydrogen bonding, and the remainder participate in van der Waals
interactions. Finally, other peak values of g(r) between functional
Supplementary data associated with this article can be found in the
groups and atoms in the PVA/silica interfacial model are not particu-
online version at http://dx.doi.org/10.1016/j.jmbbm.2017.08.027.
larly evident, indicating they only interact with each other by weak van
der Waals interactions.
References
In brief, analyzing the PCFs of PVA and PAM at the interface of silica
revealed the interaction mechanism of nano-silica in PVA/PAM blends.
Anastasia, V.P., Steve, F.A. Acquah, Dmitrenko, Maria E., Sokolova, Maria P., et al., 2016.
Hydrogen bond interactions between polar functional groups of the Improvement of pervaporation PVA membranes by the controlled incorporation of
polymer molecular chains and the hydroxyl groups on the silica surface fullerenol nanoparticles. Mater. Des. 96, 416–423.
explain how the polymer adsorbs on the silica surface. Moreover, by Andersen, H.C., 1980. Molecular dynamics simulations at constant pressure and/or
temperature. J. Chem. Phys. 72, 2384–2393.
analyzing the peak values size of PCFs, the strength of hydrogen bond
Boulet, P., Coveney, P.V., Stackhouse, Stephen, 2004. Simulation of hydrated Li+-, Na+-
formation can be compared, the larger the peak value, the stronger the and K+-montmorillonite/polymer nanocomposites using large-scale molecular dy-
hydrogen bonds are to be formed. Thus, from the peak values of the namics. Chem. Phys. Let. 389, 261–267.

536
Q. Wei et al. Journal of the Mechanical Behavior of Biomedical Materials 75 (2017) 529–537

Chen, X.B., 2014. Dispensed-based bio-manufacturing scaffolds for tissue engineering solution-cast technique and its characterization: a spectroscopic study. Iran. Polym. J.
applications. Int. J. Eng. Appl. 2 (1). 23, 153–162.
Choi, Jeeyoung, Kung, Hsiang J., Macias, Celia E., Muratoglu, Orhun K., 2012. Highly Peng, Fubing, Jiang, Zhongyi, Hoek, Eric M.V., 2011. Tuning the molecular structure,
lubricious poly (vinyl alcohol)-poly (acrylic acid) hydrogels. J. Biomed. Mater. Res. separation performance and interfacial properties of poly(vinyl alcohol)-polysulfone
Part B: Appl. Biomater. 100 (2), 524–532. interfacial composite membranes. J. Membr. Sci. 368, 26–33.
Chrissafis, K., Paraskevopoulos, K.M., Papageorgiou, G.Z., Bikiaris, D.N., 2008. Thermal Qiao, Bing, Liang, Yong, Wang, Ting-Jie, Jiang, Yanping, 2016. Surface modification to
and dynamic mechanical behavior of bionanocomposites: fumed silica nanoparticles produce hydrophobic nano-silica particles using sodium dodecyl sulfate as a modifier.
dispersed in poly(vinyl pyrrolidone), chitosan, and poly(vinyl alcohol). J. Appl. Appl. Surf. Sci. 364, 103–109.
Polym. Sci. 110, 1739–1749. Roussou, R.E., Karatasos, K., 2016. Graphene/poly(ethylene glycol) nanocomposites as
Dil, E.J., Virgilio, Nick, Favis, B.D., 2016. The effect of the interfacial assembly of nano- studied by molecular dynamics simulations. Mater. Des. 97, 163–174.
silica in poly(lactic acid)/poly(butylene adipate-co-terephthalate) blends on mor- Sahu, R., Anup, S., 2016. Molecular dynamics study of toughening mechanisms in nano-
phology, rheology and mechanical properties. Eur. Polym. J. 85, 635–646. composites as a function of structural arrangement of reinforcement. Mater. Des. 100,
El-Zawawy, Waleed K., Ibrahim, Maha M., 2012. Preparation and characterization of 132–140.
novel polymer hydrogel from industrial waste and copolymerization of poly(vinyl Sakineh, A., Mehdi, H., Mohammad, A., 2015. Preparation and investigation of novel
alcohol) and polyacrylamide. J. Appl. Polym. Sci. 124, 4362–4370. PVA/silica nanocomposites with potential application in NLO. Polym. Plast. Technol.
Gardebien, F., Brédas, J.L., Lazzaroni, R., 2005. Molecular dynamics simulations of na- Eng. 54, 192–201.
nocomposites based on poly(ε-caprolactone) grafted on montmorillonite clay. J. Shariatinia, Z., Jalali, A.M., Taromi, F.A., 2016. Molecular dynamics simulations on de-
Phys. Chem. B 109, 12287–12296. sulfurization of n-octane/thiophene mixture using silica filled polydimethylsiloxane
Ghanaati, S.M., Thimm, B.W., Unger, R.E., Orth, C., et al., 2010. Collagen-embedded nanocomposite membranes. Model. Simul. Mater. Sci. Eng. 24, 035002.
hydroxylapatite-beta-tricalcium phosphate–silicon dioxide bone substitute granules Sowjanya, J.A., Singh, J., Mohita, T., Sarvanan, S., et al., 2013. Biocomposite scaffolds
assist rapid vascularization and promote cell growth. Biomed. Mater. 5, 025004. containing chitosan/alginate/nano-silica for bone tissue engineering. Colloids Surf.
Gomez, J.C., Chen, X.B., Yang, Q.Q., 2016. Effect of nanoparticle incorporation and B: Biointerfaces 109, 294–300.
surface coating on mechanical properties of bone scaffolds: a brief review. J. Funct. Sun, H., 1998. COMPASS: an ab initio force-field optimized for condensed-phase appli-
Biomater. 7, 18. cations overview with details on alkane and benzene compounds. J. Phys. Chem. B
Griffin, M., Nayyer, L., Butler, Peter, E., Palgrave, R.G., 2016. Development of mechano- 102, 7338–7364.
responsive polymeric scaffolds using functionalized silica nano-fillers for the control Suriano, R., Griffini, G., Chiari, M., Levi, M., et al., 2014. Rheological and mechanical
of cellular functions. Nanomed.-Nanotechnol. Biol. Med. 12, 1725–1733. behavior of polyacrylamide hydrogels chemically crosslinked with allyl agarose for
Gupta, J., Nunes, C., Vyas, S., Jonnalagadda, S., 2011. Prediction of solubility parameters two-dimensional gel electrophoresis. J. Mech. Behav. Biomed. Mater. 30, 339–346.
and miscibility of pharmaceutical compounds by molecular dynamics simulation. J. Swenson, R.J., 1983. Comments on virial theorems for bounded systems. Am. J. Phys. 51,
Phys. Chem. B 115, 2014–2023. 940–942.
Hagita, K., Morita, H., Takano, H., 2016. Molecular dynamics simulation study of a Taheri-Behrooz, F., Memar Maher, T.B., Shokrieh, M.M., 2015. Mechanical properties
fracture of filler-filled polymer nanocomposites. Polymer 99, 368–375. modification of a thin film phenolic resin filled with nano silica particles. Comput.
Jawalkar, S.S., Aminabhavi, T.M., 2007. Molecular dynamics simulations to compute Mater. Sci. 96, 411–415.
diffusion coefficients of gases into polydimethylsiloxane and poly{(1,5-naphthalene)- Wang, Jinjian, Zhu, Xiaolei, Lu, Xiaohua, Zhou, Zhou, et al., 2015. On structures and
co-[1,4-durene-2,2’-bis(3,4-dicarboxyl phenyl)hexafluoropropanediimide]}. Polym. properties of polyethylene during heating and cooling processes based on molecular
Int. 56, 928–934. dynamics simulations. Comput. Theor. Chem. 1052, 26–34.
Jiang, Shan, Liu, Sha, Feng, Wenhao, 2011. PVA hydrogel properties for biomedical ap- Wang, Yanen, Wei, Qinghua, Wang, Shuzhi, Chai, Weihong, et al., 2017. Structural and
plication. J. Mech. Behav. Biomed. Mater. 4, 1228–1233. water diffusion of poly(acryl amide)/poly(vinyl alcohol) blend films: experiment and
Jin, Rihua, Hua, Youqing, 2000. Polymer Physics, second edition. Chemical Industry molecular dynamics simulations. J. Mol. Gr. Model. 71, 40–49.
Press, Beijing. Watt, J.P., Davies, G.F., O’Connell, R.J., 1976. The elastic properties of composite ma-
Labarre, D., Laurent, A., Lau, Tier A., et al., 2002. Complement activation by substituted terials. Rev. Geophys. 14, 541–563.
polyacrylamide hydrogels for embolisation and implantation. Biomaterials 23, Wei, Qinghua, Zhang, Yingfeng, Wang, Yanen, Chai, Weihong, et al., 2016a.
2319–2327. Measurement and modeling of the effect of composition ratios on the properties of
Lai, Z.B., Wang, M.C., Yan, C., Oloyede, Adekunle, 2014. Molecular dynamics simulation poly(vinyl alcohol)/poly(vinyl pyrrolidone) membranes. Mater. Des. 103, 249–258.
of mechanical behavior of osteopontin-hydroxyapatite interfaces. J. Mech. Behav. Wei, Qinghua, Wang, Yanen, Chai, Weihong, Wang, Tao, et al., 2016b. Effects of com-
Biomed. Mater. 36, 12–20. position ratio on the properties of poly(vinyl alcohol)/poly (acrylic acid) blend
Lammers, L.N., Bourg, Ian C., Okumura, M., Kolluri, K., et al., 2017. Molecular dynamics membrane: a molecular dynamics simulation study. Mater. Des. 89, 848–855.
simulations of cesium adsorption on illite nanoparticles. J. Colloid Interface Sci. 490, Wei, Qinghua, Wang, Yanen, Li, Xipei, Yang, Mingming, et al., 2016c. Study the bonding
608–620. mechanism of binders on hydroxyapatite surface and mechanical properties for 3DP
Leng, Qing, Lin, Jianming, Wu, Jihuai, et al., 2011. Synthesis and properties of poly(AM/ fabrication bone scaffolds. J. Mech. Behav. Biomed. Mater. 57, 190–200.
AAK) porous IPN hydrogels. Mater. Rev. B 25, 28–31. Wei, Qinghua, Wang, Yanen, Che, Yu, Yang, Mingming, et al., 2017a. Molecular me-
Little, Christopher James, Bawolin, N.K., Chen, X.B., 2011. Mechanical properties of chanisms in compatibility and mechanical properties of polyacrylamide/polyvinyl
natural cartilage and tissue-engineered constructs. Tissue Eng. Part B Rev. 17, alcohol blends. J. Mech. Behav. Biomed. Mater. 65, 565–573.
213–227. Wei, Qinghua, Zhang, Yingfeng, Wang, Yanen, et al., 2017b. A molecular dynamic si-
Luo, Zhonglin, Jiang, Jianwen, 2010. Molecular dynamics and dissipative particle dy- mulation method to elucidate the interaction mechanism of nano-SiO2 in polymer
namics simulations for the miscibility of poly(ethylene oxide)/poly(vinyl chloride) blends. J. Mater. Sci. 52, 12889–12901.
blends. Polymer 51, 91–299. Weiner, J.H., 2002. Statistical Mechanics of Elasticity. Dover Publications.Com, New
Malaki, Massoud, Hashemzadeh, Yasser, Karevan, Mehdi, 2016. Effect of nano-silica on York.
the mechanical properties of acrylic polyurethane coatings. Prog. Org. Coat. 101, Wescott, J.T., Qi, Y., Subramanian, L., Capehart, T.W., 2006. Mesoscale simulation of
477–485. morphology in hydrated perfluorosulfonic acid membranes. J. Chem. Phys. 124,
Materials Studio Version 5.5, 2010. Accelrys Inc., San Diego, CA. 134702–134716.
Morales-Hurtado, M., Zeng, X., Gonzalez-Rodriguez, P., Ten Elshof, J.E., et al., 2015. A Wilmowsky, Cornelius von, Vairaktaris, Elftherios, Pohle, Dirk, Rechtenwald, Thomas,
new water absorbable mechanical epidermal skin equivalent: the combination of et al., 2008. Effects of bioactive glass and beta-TCP containing three-dimensional
hydrophobic PDMS and hydrophilic PVA hydrogel. J. Mech. Behav. Biomed. Mater. laser sintered polyetheretherketone composites on osteoblasts in vitro. J. Biomed.
46, 305–317. Mater. Res. 87A, 896–902.
Olubamiji, Adeola D., Izadifar, Zohreh, Si, Jennifer L., Cooper, David M., Eames, B.F., Yang, Junqing, Gong, Xuedong, Wang, Guixiang, 2015. Compatibility and mechanical
Chen, D.X., 2016. Modulating mechanical behaviour of 3D-printed cartilage-mimetic properties of BAMO–AMMO/DIANP composites: a molecular dynamics simulation.
PCL scaffolds: influence of molecular weight and pore geometry. Biofabrication 8, Comput. Mater. Sci. 102, 1–6.
025020. Zhao, Yucheng, Wang, Changan, 2016. Nano-network MnO2/polyaniline composites with
Pan, Fusheng, Peng, Fubing, Lu, Lianyu, Wang, Jingtao, et al., 2008. Molecular simulation enhanced electrochemical properties for supercapacitors. Mater. Des. 97, 512–518.
on penetrants diffusion at the interface region of organic–inorganic hybrid mem- Zhou, H.Z., Liu, H.Y., Zhou, H.M., Zhang, Y., et al., 2016. On adhesive properties of nano-
branes. Chem. Eng. Sci. 63, 1072–1080. silica/epoxy bonded single-lap joints. Mater. Des. 95, 212–218.
Patel, Gaurang, Sureshkumar, M.B., 2014. Preparation of PAM/PVA blending films by

537

You might also like