Shin-Pon Ju*a,
Wei-Chun Huanga,
Ken-Huang Lina,
Hui-Lung Chen*b,
Jenn-Sen Linc and
Jin-Yuan Hsieh*d
aDepartment of Mechanical and Electro-Mechanical Engineering, National Sun Yat-sen University, Kaohsiung 804, Taiwan. E-mail: jushin-pon@mail.nsysu.edu.tw
bDepartment of Chemistry and Institute of Applied Chemistry, Chinese Culture University, Taipei 111, Taiwan. E-mail: chl3@faculty.pccu.edu.tw
cDepartment of Mechanical Engineering, National United University, Miao-Li 360, Taiwan. E-mail: jsenlin@nuu.edu.tw
dDepartment of Mechanical Engineering, Minghsin Institute of Technology, Hsin-Chu 304, Taiwan. E-mail: jyhsieh@must.edu.tw
First published on 14th February 2014
The mechanical properties of polyglycolic acid (PGA) of different water weight fractions (1.7%, 2.9%, and 5%) were investigated by molecular dynamics (MD) simulation through a tensile test. The variation in the degree of crystallinity with water content was also investigated using XRD profiles. The Young's modulus, mechanical strength, and fracture mechanism of all PGA–water systems were drawn from the corresponding stress–strain profiles. Furthermore, water diffusion behavior within the PGA matrix both before and after the tension test was also studied. The water diffusion coefficients were derived from the mean square displacements of all water molecules within the PGA matrix.
Starting several decades ago, these biodegradable polymers began to be used inside the human body for medical treatments.15 PGA is exposed to physiological conditions; it will be randomly degraded by hydrolysis into shorter PGA fragments. Glycolic acid, which is nontoxic to the human body, is the final degradation product. It can easily go through the tricarboxylic acid cycle and be removed as water and carbon dioxide from the human body by breathing. Part of the glycolic acid can be also excreted via kidney and urine.16 PGA exhibits a degree of crystallinity around 46–52% (ref. 17) and has a glass transition temperature (Tg) between 35 and 40 °C,18 resulting in good mechanical properties with flexibility when used as a biomaterial inside the human body. Because of its high degree of crystallisation, it is not soluble in most organic solvents; the exceptions are highly fluorinated organics such as hexafluoroisopropanol. In Montes' study,19 they investigated the crystallinity of PGA materials from different annealing processes and found that the Tg value and crystallinity of PGA can be adjusted by thermal treatment, and the time period of PGA degradation can also be controlled, which makes the application of PGA broader.
The strength of shearing, tension and compression of PGA are usually adopted to observe its mechanical properties because Young's modulus, bulk modulus, and shear modulus20 are important factors influencing degradation time. Since direct investigation of molecular behaviour of PGA under mechanical deformation is difficult due to the very small scales involved, numerical methods are generally preferred. These methods can assist in clarifying the physical insights gained from experimental results. Furthermore, they can be used to predict possible results in order to obtain a preliminary analysis of the influence of process parameters prior to conducting an expensive experiment. Molecular dynamics (MD) simulation is a powerful tool for the investigation of molecular behaviour at an atomic level. Adopting this technique, Sekine conducted a PGA fiber tensile simulation to monitor the variation of PGA chain conformation.21 They found the PGA torsion angles distribute within a small range after the PGA molecules were stretched to a certain degree of elongation. In Ding's study,22 both molecular dynamics and Monte Carlo method were conducted to simulate the PGA system under tension by using a united-atom potential. They found the Young's modulus of both amorphous PGA and those with partial crystallinity become smaller at higher temperatures (Table 1).
Water contents (wt%) | H2O (M.W.) | PGA chains | Atoms | Density (kg m−3) |
---|---|---|---|---|
0% | 0(0) | 7 | 8414 | 1.4842 |
1.7% | 80(1440) | 7 | 8661 | 1.492 |
2.9% | 135(2430) | 7 | 8826 | 1.461 |
5% | 240(4320) | 7 | 9141 | 1.465 |
Hurrell et al. found that decomposition behaviour of PGA was active after absorbing water molecules, and mass decreases quickly with an increase in the absorption capacity.23 You et al.24 demonstrate by DSC and WAXD analysis that the crystallinity increases significantly in the initial period of PGA decomposition, and gradually decrease in the later period. In particular, Lyu et al.25 investigated the decomposition process of PLA, which is separated into four stages. They found that the degradation of PLA must pass the water saturation process in the first stage. These studies demonstrate the presence of water absorption and the increase of PGA crystallinity in the decomposition reaction. Since PGA is generally applied in the production of surgical sutures, cartilage, bone screws, and other biomaterials, the different stresses and effect of water are crucial considerations. Therefore, the stress and diffusion behaviours are here investigated at different fractions of water molecules in the PGA system. Furthermore, the amount of crystallinity at different stress levels in the decomposition of PGA is elucidated by XRD analysis.
In previous studies, only experimental results were used to determine mechanical properties. However, the detailed changes of microstructures are difficult to observe. In this work, MD simulation was proposed to simulate the PGA strength variation during the degradation process at different temperatures. The Young's modulus, mechanical strength, and fracture mechanism of all PGA–water systems were obtained by the corresponding stress–strain profiles. Furthermore, the water diffusion behaviours within PGA matrix were also studied.
![]() | ||
Fig. 1 (a) PGA chemical formula. Equilibrium configurations of (b) pure PGA and (c) polymer and water composite material. (d) Uniaxial tension model. |
For the uniaxial tension, the model was then relaxed at 300 K and 1 atm for another 50 ps and this model can be seen in Fig. 1(d). During the tensile process, the cell length in the z-dimension was elongated by 0.25 Å, and then the system was relaxed for 2 ps before imposing the next tensile increment. During the tension simulation, the stress σmn on the m plane and in the n-direction is calculated by30
![]() | (1) |
Furthermore, the normal strain in the axial direction ε of the PGA is calculated as
![]() | (2) |
In the crystallinity analysis, X-ray diffraction (XRD) is commonly used as direct evidence of the periodic crystal structure by the relative diffraction intensity at different X-ray incident angles. All XRD profiles shown in the current study were conducted by the REFLEX module in Materials Studio package.31 In REFLEX, Bragg's law is used to obtain the constructive interference intensity for X-rays scattered by materials, and the formula is listed as follows:
![]() | (3) |
We calculate the X-ray diffraction patterns after the equilibrium configurations of the procedure. In the calculations, the diffract meter range 2θ was set from 1° to 45° with the step size of 0.05 degree. The X-ray source anode is copper, and the radiation wavelength is set to 1.540562 Å. We calculate the degree of crystallinity after comparing the experimental XRD patterns with the patterns of the structure generated from the MD simulations.
For the degree of crystallinity calculations, the mixture diffraction pattern is composed of the crystalline phase, the amorphous phase and a background contribution. The intensity of input mixture diffraction pattern can be defined as:
Im(2θ) = Icr(2θ) + Iam(2θ) + Ib(2θ) | (4) |
To obtain the relative weight of the crystalline phase in the decomposition of the mixture diffraction pattern, we can calculate the degree of crystallinity (xcr) from quantitative phase analysis (QPA) theory. This can be defined as the scattering intensity of a powder sample of a single pure phase:
Icr(2θ) = pcrInormcr(2θ) | (5) |
Iam(2θ) = pamInormam(2θ) | (6) |
![]() | (7) |
Fig. 4 shows the mean-square displacement (MSD) plots of pure water and water molecules within PGA at different water contents. The MSD is defined as:
![]() | (8) |
![]() | (9) |
![]() | ||
Fig. 4 The mean-square displacement (MSD) plots of water molecules within PGA at water contents of 1.7%, 2.9%, and 5%. |
In order to use the optimal simulation parameters for Forcite package on the water diffusion behaviour, the diffusion coefficient of bulk water at 300 K was calculated first in the NVT ensemble by eqn (10) with time step length of 0.4 fs. The value of 2.35 × 10−9 m2 s−1 from our MD simulation is very close to the experimental value of 2.3 × 10−9 m2 s−1,33 indicating the reliability of the PCFF force field to predict the diffusion behaviour of water molecules. The water diffusion coefficients within PGA by eqn (8) are listed in Table 2 for water contents of 1.7%, 2.9%, and 5%.
Water contents | Diffusion coefficients (m2 s−1) |
---|---|
1.7% | 2.38 × 10−10 |
2.9% | 3.46 × 10−10 |
5% | 3.07 × 10−10 |
Pure water32 | 2.3 × 10−9 |
From Fig. 4, it is clear that the MSD value of water content of 2.9% is the largest and the MSD value of 5% water content is smaller than that of 2.9% water content, but larger than that of 1.7% water content. From Fig. 3 and 4, one can see the influence that degree of PGA crystallinity has influence on water diffusion behaviour, such that PGA with a higher degree of crystallinity has a larger water diffusion coefficient. In the comparison between pure water and water-content PGA, the diffusion coefficient of the pure water is significantly larger than that of water within the PGA, indicating the flowed difficulty of water in the polymer structure.
The stress–strain profiles for PGA with 2.9% and 5% water contents under tension are shown in Fig. 5, and the corresponding morphologies at different strains, labelled by (a)–(d) in Fig. 5, are displayed in Fig. 6(a)–(d) for PGA with 5% water content. The stress values for both plots display an abrupt linear increase with strain from 0 to 0.02, indicating an elastic characteristic over this strain range. At strain between 0.02 and 0.2, the stresses fluctuate around a constant value and then display a decreasing trend when the strain continuously increases. In Fig. 6(c), for the morphology at strain of 0.3, there are many voids that have appeared within the PGA matrix and lead to a decrease in stress. In Fig. 6(d), the areas of voids have further expanded, causing the fracture of PGA material. It should be noted the yielding stress is larger than 20 GPa, which is much larger than experimental values. The main reason is the PBC models used in our current study lack free surfaces because of the limitation of computational power. Although the smaller model yields much higher yielding stress for PGA, the tension deformation characteristics obtained are still reasonable when compared to experimental observation.
This journal is © The Royal Society of Chemistry 2014 |