NC State
BioResources
Zhang, K., Lin, J., Hao, C., Hong, G., Chen, Z., Chen, Z., Zhang, S., and Song, W. (2019). "Effect of nano-hydroxyapatite modification of bamboo fiber on the properties of bamboo fiber/polylactic acid composites," BioRes. 14(1), 1694-1707.

Abstract

An investigation was conducted of the modification of bamboo fibers (BF) by nano-hydroxyapatite (N-HA) and its impact on the mechanical and thermal properties of BF/PLA composites. The functional groups, and crystallinity of the N-HA modified BF were investigated with Fourier transform infrared (FTIR) spectroscopy and X-ray diffraction (XRD), respectively. The effects of different N-HA contents on the properties of the BF/PLA composites were evaluated using mechanical testing (bending, tensile, and impact properties), differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), and scanning electron microscopy (SEM). The results showed that the most suitable N-HA content was 7.5 wt.%. Treating BF with an optimum concentration of N-HA decreased the polarity of the bamboo fiber, while maintaining the crystal structure of the cellulose. Compared with the control group, the mechanical properties and the crystallinity of the modified BF/PLA composites were improved, and the flexural, tensile, and impact strengths increased by 10.2%, 11.6%, and 28.1%, respectively. The thermogravimetric analysis results indicated modified BF/PLA composites has higher thermal stability.


Download PDF

Full Article

Effect of Nano-hydroxyapatite Modification of Bamboo Fiber on the Properties of Bamboo Fiber/Polylactic Acid Composites

Kaiqiang Zhang,a,b,c Jianyong Lin,a,b,c Chengyi Hao,a,b,c Gonghua Hong,a,b,c Zhenghao Chen,a,b,c Zhangjing Chen,d Shuangbao Zhang,a,b,c,* and Wei Song a,b,c,*

An investigation was conducted of the modification of bamboo fibers (BF) by nano-hydroxyapatite (N-HA) and its impact on the mechanical and thermal properties of BF/PLA composites. The functional groups, and crystallinity of the N-HA modified BF were investigated with Fourier transform infrared (FTIR) spectroscopy and X-ray diffraction (XRD), respectively. The effects of different N-HA contents on the properties of the BF/PLA composites were evaluated using mechanical testing (bending, tensile, and impact properties), differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), and scanning electron microscopy (SEM). The results showed that the most suitable N-HA content was 7.5 wt.%. Treating BF with an optimum concentration of N-HA decreased the polarity of the bamboo fiber, while maintaining the crystal structure of the cellulose. Compared with the control group, the mechanical properties and the crystallinity of the modified BF/PLA composites were improved, and the flexural, tensile, and impact strengths increased by 10.2%, 11.6%, and 28.1%, respectively. The thermogravimetric analysis results indicated modified BF/PLA composites has higher thermal stability.

Keywords: Bamboo fiber; Polylactic acid; Nano-hydroxyapatite

Contact information: a: MOE Key Laboratory of Wooden Material Science and Application, College of Materials Science and Technology, Beijing Forestry University, Beijing 100083, China; b: Beijing Key Laboratory of Wood Science and Engineering, College of Materials Science and Technology, Beijing Forestry University, Beijing 100083, China; c: MOE Engineering Research Center of Forestry Biomass Materials and Bioenergy, College of Materials Science and Technology, Beijing Forestry University, Beijing 100083, China; d: Department of Sustainable Biomaterials, Virginia Tech, Blacksburg VA 24061, USA; *Corresponding authors: shuangbaozhangj5@163.com; j5international@163.com

INTRODUCTION

Polylactic acid (PLA) is a polymeric material that exhibits excellent biocompatibility and biodegradation. Thus, it is widely used in the medical field and has received increasing attention (Singh et al. 2004; Tiainen et al. 2006). It is used as a new type of orthopedic internal fixation material for use in medical products, such as bone nails and bone plates. Although PLA has many excellent properties, it cannot meet the fixation requirements of many functional bones because of the low mechanical strength of pure PLA (Wan et al. 2004). At present, the performance of PLA is mainly enhanced by self-reinforcement or by compounding it with other materials, such as natural and carbon fibers. In addition to biodegradable and environmentally friendly nature, bamboo fiber (BF) has a unique structure, good mechanical properties, and natural bacteriostatic properties. Thus, it has attracted attention for enhancing the performance of PLA (Abdul Khalil et al. 2012; Liu et al. 2012; Mishra et al. 2012; Yu et al. 2014). However, BF contains a large number of polar hydrophilic hydroxyl groups, which lead to poor compatibility with non-polar PLA and poor performance of composites; in contrast, the degradation products of PLA are acidic, which easily cause an inflammatory reaction in the human body (Sawpan et al. 2011). These problems plague the application of PLA in composites.

It is necessary to modify the surface of BF to enhance the performance of its composites. Shi et al. (2011) and Wang et al. (2017) reported that the modification of natural fibers by inorganic nanoparticles is an effective method that remarkably improves the interface between the fibers and polymer matrix. The inorganic nanoparticles can adhere to the surface of cell walls. As a nucleation site, it affects the crystallization behavior of the polymer matrix and produces a robust electrostatic attraction to the surface of the non-polar polymer. Therefore, the interfacial compatibility of the composite may be improved (Wang et al. 2016). Nano-hydroxyapatite (N-HA) has good bioactivity and osteoconductivity, and it can form a direct bond with bone, so it is widely used in bone replacement implant materials. Moreover, the acidic degradation products of PLA can be buffered by hydroxyapatite (HA). Additionally, the osteoinductivity of N-HA can provide a suitable environment for bone cell adhesion and growth, which conforms to the requirements of bone tissue engineering (Verheyen et al. 1992; Shikinami and Okuno 2001; Zhao et al. 2002).

In this study, the BF were modified with a series of concentrations of N-HA, and then the treated BF was introduced into the PLA matrix. The purpose of this study was to explore the interfacial enhancement effect of N-HA on composite materials, which may provide new theoretical and practical bases for the practical application of inorganic particle-reinforced bamboo plastic composites.

EXPERIMENTAL

Materials

Bamboo fibers ranging from 60 mesh to 80 mesh in size were purchased from Sentai Wood Plastic Composites Material Co., Ltd. (Huzhou, China). The PLA (grade 4032D, melting point = 160 °C, and density = 1.24 g/cm3) was provided by NatureWorks Ind. Co. Ltd. (Northford, CT, USA). The N-HA was purchased from Nanjing Emperor Nano Material Co., Ltd. (Nanjing, China) in sizes ranging between 60 nm and 80 nm. The deionized water was provided by the laboratory of Beijing Forestry University.

Table 1. Process Parameters for N-HA Impregnation

Modification process of the bamboo fiber

The BF was dried for 8 h at 105 °C in a drying oven. Subsequently, 100 g of BF were dispersed in 1 L of deionized water for 30 min at 300 rpm with a mechanical stirring apparatus at room temperature. Next, different N-HA contents (2.5 %, 5%, 7.5%, 10% and 12.5%) were added to the mixed solution, respectively, and the mixtures were stirred for 30 min. After this, the modified BF was washed with deionized water and placed in an air-dry oven and dried at 105 °C for 8 h. The process parameters are shown in Table 1.

Preparation of the BF/PLA composites

The route of the preparation of BF/PLA composites was similar to our group’s previous research (Lin et al. 2018). The weight ratio of PLA to BF was maintained at 3:2 during preparation of the BF/PLA composites. Firstly, the PLA and BF were mixed uniformly in the blending machine, and then the compound was extruded via a co-rotating twin-screw extruder (KESUN KS-20, Kunshan, China). The twin-screw rotation speed was kept at 100 rpm, and the temperatures of each part of the twin screw were 165 °C, 170 °C, 175 °C, 165 °C, and 160 °C. Then, the mixture was pulverized into small particles with 4-mm diameters. Next, the composites were produced after these particles underwent a hot-pressing process (HAPCO, BY 602 × 2/2 150 T Testing Press, Suzhou, China), which was performed with a pressing time of 6 min, pressing pressure of 4 MPa, and hot-pressing temperature of 175 °C. Finally, the sample was pressed at a pressure of 4 MPa for 10 min at room temperature. The dimensions of the PLA composites were 270 mm × 270 mm × 4 mm, and the density was 1.2 g/cm3.

Characterization of the Bamboo Fiber

Fourier transform infrared spectroscopy

The Fourier transform infrared (FTIR) analysis of the BF was performed using a Vertex 70v (BRUKER, Karlsruhe, Germany), with a resolution of 4 cm-1 and average of 40 scans over a scanning range from 4000 cm-1 to 400 cm-1. The BF sample was mixed with potassium bromide, and then the mixture was compressed into tablets for recording the spectra. Each group was measured three times to minimize error.

X-ray diffraction procedure and analysis

The crystallinities of the BF were measured by X-ray diffraction (XRD) using a D8 Advanced X-ray diffractometer (BRUKER, AXS, Karlsrushe, Germany) with Cu Kα radiation (λ = 0.154060 nm), a radiation tube voltage of 40 kV, and a current of 40 mA. The scanning rate was 2°/min at 2θ angles of 5° to 40°. The cellulose crystallinity index (IXRD) was calculated with Eq. 1 (Le Troedec et al. 2008; Pickering et al. 2011):

 (1)

where I002 is the maximum intensity of the (002) lattice diffraction plane at a 2θ angle between 22° and 23° (22° ≤ 2θ ≤ 23°), and Iamp is the diffraction intensity at a 2θ angle close to 18°, which represents amorphous materials in the cellulosic fibers.

Characterization of the Bamboo Fiber/PLA Composites

Mechanical characterization

The tensile property of the composite was measured according to ASTM D638-03 (2003) using an AG-IS (Shimadzu, Kyoto, Japan). The dimensions of the specimens were 165 mm × 20 mm × 4 mm (dumbbell-shape, with a radius of arc of 76 mm, gauge length of 50 mm, and middle part width of 13 mm), and each group had eight specimens that were measured at a stretching speed of 5 mm/min.

The flexural property of the composite was measured according to ASTM D790-03 (2003) using the AG-IS. The dimensions of the specimens were 80 mm × 10 mm × 4 mm, and each group had eight specimens that were measured at a strain speed of 20 mm/min.

The impact property of the composite was measured according to ASTM D256-03 (2003) using a CREE-1002C (XJJ5, KERUI, Dongguan, China). The dimensions of the specimens were 80 mm × 10 mm × 4 mm, and each group had eight specimens that were measured at an impact speed of 2.9 m/s with a 1 J hammer. These tests were performed at room temperature.

Thermogravimetric analysis

The thermal stability of the composites was measured using a Q5000 apparatus (TA Instrument, New Castle, USA) at a heating rate of 10 °C/min from 40 °C to 600 °C under an N2 gas flow rate of 50 mL/min.

Differential scanning calorimetry

Differential scanning calorimetry (DSC) analysis was performed using a Netzsch DSC-204 (Bavaria, Germany). All of the data obtained were taken within a temperature range of 40 °C to 200 °C at a heating rate of 10 °C/min; the samples were then cooled at a rate of 10 °C/min and heated again to 200 °C at the same rate. The crystallinity (XDSC, %) was calculated with Eq. 2 (Fischer et al. 1973; Rana et al. 2000; Joseph et al. 2003),

 (2)

where f is the mass fraction of the PLA matrix in the composites, ΔHm is the melting enthalpy of the composites, and ΔH0m is the melting enthalpy of the 100% crystalline PLA (93.7 J/g).

Interfacial morphology analysis

The impact fracture morphology analysis was performed using an FEI Quanta FEG 650 scanning electron microscopy (SEM, Thermo Fisher Scientific Inc., Waltham, MA, USA) with an acceleration voltage of 15 kV. All of the specimens were sputter-coated with gold before observation.

RESULTS AND DISCUSSION

FTIR Analysis of the Bamboo Fiber

The FTIR spectra of the untreated (Control) and treated BF (HA-BF-2 and HA-BF-5) are shown in Fig. 1. As shown, there were characteristic peaks near 3430 cm-1, which according to Fourier transform infrared (FTIR) spectroscopy were designated as stretching vibration of hydroxyl (–OH) groups. The characteristic peaks at 2910 cm-1 were attributed to the symmetric and asymmetric stretching vibration absorption peaks of methyl (–CH3) and methylene (–CH2) groups that belong to cellulose (Sgriccia et al. 2008). The characteristic peaks located at 1740 cm-1, 1425 cm-1, 1370 cm-1, and 1160 cm-1 were assigned to C=O stretching, -CH2– shear, and C-H bending vibrations (Zhang et al. 2012).

Fig. 1. FTIR spectra of the untreated and N-HA-treated BF

It was clear that the intensity of the absorption peaks of the three curves showed a distinct difference at 3430 cm-1. This was explained by the fact that during the BF treatment with N-HA, the -OH groups on the N-HA molecules formed hydrogen bonds with the hydroxyl groups on the surface of the BF. As the N-HA content was increased, the amount of hydrogen bonds formed increased, which resulted in a decrease in the number of -OH groups. Compared with the untreated BF, the curves of the treated BF showed weak peaks at 603 cm-1 and 564 cm-1, which were assigned to PO43- stretching vibration. This indicated that N-HA has potential to improve the interfacial properties of BF and plastic composites.

XRD Analysis of the Bamboo Fiber

Figure 2 shows the results of the analysis of the crystalline performances of the treated and untreated BF by XRD. The crystalline structure of the natural plant fiber is reflected by the two crystalline areas of (101) and (002) in the XRD pattern, whose 2θ angles are 17.5° and 22.3°, respectively. The position of the characteristic peak of the treated BF was barely changed compared with that of the untreated BF, which indicated that N-HA had almost no influence on the crystalline region of the cellulose. The structure of the crystalline areas of the cellulose chain was barely destroyed in the treatment process.

As shown in Fig. 2, the treated BF showed characteristic peaks at 25.7° and 31.8°, which were designated as characteristic bands of HA. When the content of the N-HA was increased from 0% to 7.5%, the crystallinity of cellulose increased gradually. And when the content of N-HA was 7.5%, the crystallinity of cellulose reached the maximum value of 44.0%. Then, with the increase of N-HA content, the crystallinity of the cellulose gradually decreased and reached a minimum of 41% when the N-HA content was 12.5%. This indicates that the low content of N-HA induced the nucleation of the cellulose of the BF and then initiated crystallization. When the content of N-HA exceeds a certain amount, the crystallization of cellulose was hindered when the N-HA content exceeded 7.5 wt.%, agglomeration occurred and the crystal nucleus failed to increase (Qian et al. 2018).

Fig. 2. XRD analysis of the treated and untreated BF

Effect of the BF Treatments on the Mechanical Properties of the BF/PLA Composites

Table 2 shows the effects of the different N-HA contents on the mechanical properties of the BF/PLA composites. Compared with the control group, the flexural strength, flexural modulus, tensile strength, tensile modulus, and impact strength of the treated BF all increased. The main reasons were that the N-HA particles gradually filled in the micropores, grooves, etc. on the surface of the fiber, which reduced the generation of void defects during the compounding process. The N-HA formed strong hydrogen bonds with the fiber molecules. Thus, the hydrophilicity of the fiber surface was reduced and the bonding strength with PLA was enhanced. Thereby, the interface properties of the composites were increased.

Table 2. Mechanical Properties of BF/PLA Composites

Note: The data were analyzed by one-way ANOVA based on a 95% confidence interval. The standard deviations of the test results are shown in parentheses; the letter markers show the statistical differences. The groups do not differ significantly from one another when they have the same letter(s), and vice versa.

With the increase of N-HA content, the mechanical properties of treated BF/PLA composites first increased and then decreased. When the N-HA content was 7.5 wt.%, the mechanical properties of the composites were superior in this study. However, when the N-HA content exceeded this value, the mechanical properties of the composites were adversely affected. This was because the to high concentration N-HA was unevenly distributed on the surface of the fiber, which aggregated into large particles and caused the cross-link density of the modified fiber and PLA matrix to not be uniform. This resulted in poor stress transfer between the interfaces, which reduced the interfacial compatibility between the BF and PLA (Qian et al.2015b; Qian et al. 2016). Therefore, the optimum N-HA content in this study was 7.5 wt.%, which caused the flexural strength, flexural modulus, tensile strength, tensile modulus, and impact strength to increase by 10.2%, 5%, 11.6%, 22.7%, and 28.1%, respectively, in comparison with those of the untreated composites.

Thermogravimetric Analysis

Figure 3 shows the thermogravimetric (TGA) and differential thermogravimetric (DTG) curves based on the PLA matrix composites enhanced with untreated and treated BF. The initial thermal degradation temperature (T1) and maximum rate temperature (T2) of the composite material are listed in Table 3.

Table 3. Thermal Parameters of the BF/PLA Composites

Note:Twas the 5% mass loss rate of the material. T2 were the first peaks in the DTG curve of the material.

Figure 3 shows that weight loss in the composites can be divided into three stages. In the first stage (from 30 °C to 230 °C), the mass loss was associated with the evaporation of moisture in the BF and depolymerization of macromolecular compounds. In the second stage (230 °C to 400 °C), the decline was obvious within the range of 230 °C to 350 °C, which was mainly because of the thermal decomposition of hemicellulose.

Fig. 3. TGA curves (a) and DTG curves (b)

From 350 °C to 400 °C, which was because of the thermal decomposition of the cellulose . The T1 and T2 of the untreated BF/PLA composites were 291 °C and 312 °C, respectively. With the addition of N-HA, the T1 and T2 showed an increasing trend. It was observed that the thermal stability of the treated BF/PLA was significantly higher than that of the control. This was explained by the following reasons. The first reason was that the characteristics of N-HA prevented heat transfer; and the second reason was that the N-HA improved the cross-linking density of the PLA matrix and strengthened the molecular structure, so that the thermal stability of the PLA matrix-based composites was improved. In the third stage (from 400 °C to 600 °C), the N-HA treated group had a lower mass loss compared with that of the control group. Consequently, the N-HA treatment resulted in a higher thermal stability.

Differential Scanning Calorimetry

Figure 4 and Table 4 show the DSC thermograms and relevant thermal parameters of the composite. Compared with the untreated BF/PLA composites, the glass transition temperature (Tg) of the treated BF-reinforced composites shifted to a higher temperature. The main reason was that the N-HA-modified BF had more hydrogen bonds, which hindered the molecular chain movement of the PLA. Therefore, the molecules required more energy to start moving and needed a higher temperature to convert the PLA from a glassy state to a rubbery state (Qian et al. 2015a). However, the Tg showed a downward trend when the N-HA content was above 7.5 wt.%. The reason was the agglomeration of HA, which led to the uneven cross-linking density of the interfacial region and resulted in a decrease in the total crosslinking density of the PLA matrix (Qian et al. 2018).

Table 4. Typical Thermal Parameters of the DSC Analysis of the Various BF/PLA Composites

Tm – The melting temperature of the PLA matrix; and Xc – the crystallinity of PLA matrix

Fig. 4. DSC curves of the untreated and treated BF/PLA composites

Surface Morphology

The microscopic morphology of the impact cross-section of the composites is shown in Fig. 5.

Fig. 5. SEM micrographs of the impact cross-section of the untreated and treated BF/PLA: (a) control group; (b) HA-BF-1; (c) HA-BF-2; (d) HA-BF-3; (e) HA-BF-5; (f) HA-BF-5

Figure 5a shows that the untreated-BF composite group had a high surface roughness and numerous large diameter voids, which were produced when the BF was pulled out. This showed that the interface between the BF and PLA was poor. This was attributed to the presence of free hydroxyl groups in the BF, which made them agglomerate and inhibited the dispersion of the BF in the PLA matrix. Thus, the interaction between the BF and PLA was dramatically weakened (Hong et al. 2017). Figure 5b and 5c show that the number of voids and agglomerated BF on the impact-crushed surface decreased rapidly, which indicated that the N-HA treatment improved the interfacial compatibility between the filler and matrix. Figure 5b (HA-BF-1) shows that the lower N-HA content improved the interfacial properties of the BF and PLA to a limited extent. Figures 5c and 5d show that the interfacial bonding of the composites increased remarkably with an increase in the N-HA content (HA-BF-2 = 5 wt.%; and HA-BF-3 = 7.5 wt.%). Figure 5c shows that the single fiber formed a tight interface with the PLA matrix and the excellent interfacial compatibility allowed for a uniform stress transmission. Figure 5d shows that the dispersion of BF in the PLA matrix was enhanced. However, Fig. 5e shows that when the N-HA content continued to increase (HA-BF-4 = 10 wt.%), a void developed between the BF and PLA matrix. This phenomenon indicated that an excessively high concentration N-HA caused uneven distribution on the fiber surface and agglomeration into large particles, which generated voids between the BF and PLA. This thereby weakened the interface interaction between the BF and PLA matrix. Also, when the N-HA content increased to the highest level used in this study (HA-BF-5 = 12.5 wt.%), agglomeration of N-HA might be more pronounced, which led to the weakest interface bonding in the experimental group.

CONCLUSIONS

  1. The present research investigated the effect of the N-HA modification on the properties of BF-reinforced PLA composites. The enhancement of the interfacial adhesion between the BF and PLA matrix was highly dependent upon a solid connection established by the N-HA, which improved the mechanical properties of the BF/PLA composites.
  2. The XRD results showed that after the N-HA treatment the crystallinity of the cellulose was increased.
  3. The TGA and DTG curves indicated that the N-HA treatment improved the thermal stability of the composite. Additionally, the N-HA treatment also improved the thermal properties of the BF/PLA composites, such as the TgTm, and Xc.
  4. The SEM micrographs of the impact-fractured specimens demonstrated that the HA-BF-3/PLA composite had a better fiber-matrix adhesion than the other composites. The improved adhesion implied that the N-HA treatment was efficient in terms of the mechanical and thermal properties.

ACKNOWLEDGMENTS

This work was supported by the National Natural Science Foundation of China (31670571), the National Key R & D Program of China (2017YFD0600804), the Co-Built Foundation with Zhejiang Province (“R & D on Natural Fiber Composites and Environmentally Friendly Adhesives”, CZXC201410), and the Zhejiang Chengzhu Advanced Material Technology Co., Ltd.

REFERENCES CITED

Abdul Khalil, H. P. S., Bhat, I. U. H., Jawaid, M., Zaidon, A., Hermawan, D., and Hadi, Y. S. (2012). “Bamboo fibre reinforced biocomposites: A review,” Mater. Design 42, 353-368. DOI: 10.1016/j.matdes.2012.06.015

ASTM D790-03 (2003). “Standard test methods for flexural properties of unreinforced and reinforced plastics and electrical insulating materials,” ASTM International, West Conshohocken, USA. DOI: 10.1520/D0790-03

ASTM D638-03 (2003). “Standard test method for tensile properties of plastics,” ASTM International, West Conshohocken, USA. DOI: 10.1520/D0638-03

ASTM D256-03 (2003). “Standard test methods for determining the lzod pendulum impact resistance of plastics,” ASTM International, West Conshohocken, USA. DOI: 10.1520/D0256-03

Fischer, E. W., Sterzel, H. J., and Wegner, G. (1973). “Investigation of the structure of solution grown crystals of lactide copolymers by means of chemical reactions,” Colloid Polym. Sci. 251(11), 980-990. DOI: 10.1007/BF01498927

Hong, G., Meng, Y., Yang, Z., Cheng, H., Zhang, S., and Song, W. (2017). “Mussel-inspired polydopamine modification of bamboo fiber and its effect on the properties of bamboo fiber/polybutylene succinate composites,” BioResources 12(4), 8419-8442. DOI: 10.15376/biores.12.4.8419-8442

Joseph, P. V., Mathew, G., Joseph, K., Groeninckx, G., and Thomas, S. (2003). “Dynamic mechanical properties of short sisal fibre reinforced polypropylene composites,” Compos. Part A-Appl. S. 34(3), 275-290. DOI: 10.1016/S1359-835X(02)00020-9

Le Troedec, M., Sedan, D., Peyratout, C., Bonnet, J. P., Smith, A., Guinebretiere, R., Gloaguen, V., and Krausz, P. (2008). “Influence of various chemical treatments on the composition and structure of hemp fibres,” Compos. Part A-Appl. S. 39(3), 514-522. DOI: 10.1016/j.compositesa.2007.12.001

Lin, Y. C. , Cho, J. , Tompsett, G. A. , Westmoreland, P. R. , and Huber, G. W. (2009). “Kinetics and mechanism of cellulose pyrolysis,” Journal of Physical Chemistry C 113(46), 20097-20107.DOI: 10.1021/jp906702p

Lin, J., Yang, Z., Hu, X., Hong, G., Zhang, S., and Song, W. (2018). “The effect of alkali treatment on properties of dopamine modification of bamboo fiber/polylactic acid composites,” Polymers-Basel 10(4), 403. DOI: 10.3390/polym10040403

Liu, D., Song, J., Anderson, D. P., Chang, P. R., and Hua, Y. (2012). “Bamboo fiber and its reinforced composites: Structure and properties,” Cellulose 19(5), 1449-1480. DOI: 10.1007/s10570-012-9741-1

Mishra, R., Behera, B. K., and Pal, B. P. (2012). “Novelty of bamboo fabric,” J. Text. I. 103(3), 320-329. DOI: 10.1080/00405000.2011.576467

Pickering, K. L., Sawpan, M. A., Jayaraman, J., and Fernyhough, A. (2011). “Influence of loading rate, alkali fibre treatment and crystallinity on fracture toughness of random short hemp fibre reinforced polylactide bio-composites,” Compos. Part A-Appl. S. 42(9), 1148-1156. DOI: 10.1016/j.compositesa.2011.04.020

Qian, S., Mao, H., Zarei, E., and Sheng, K. (2015a). “Preparation and characterization of maleic anhydride compatibilized poly(lactic acid)/bamboo particles biocomposites,” J. Polym.Environ. 23(3), 341-347. DOI: 10.1007/s10924-015-0715-x

Qian, S., Wang, H., Zarei, E., and Sheng, K. (2015b). “Effect of hydrothermal pretreatment on the properties of moso bamboo particles reinforced polyvinyl chloride composites,” Compos. Part B – Eng. 82, 23-29. DOI: 10.1016/j.compositesb.2015.08.007

Qian, S., Sheng, K., Yao, W., and Yu, H. (2016). “Poly(lactic acid) biocomposites reinforced with ultrafine bamboo-char: Morphology, mechanical, thermal, and water absorption properties,” J. Appl. Polym. Sci. 133(20). DOI: 10.1002/app.43425

Qian, S., Zhang, H., Yao, W., and Sheng, K. (2018). “Effects of bamboo cellulose nanowhisker content on the morphology, crystallization, mechanical, and thermal properties of PLA matrix biocomposites,” Compos. Part B – Eng. 133, 203-209. DOI: 10.1016/j.compositesb.2017.09.040

Rana, D., Kim, H. L., Kwag, H., and Choe, S. (2000). “Hybrid blends of similar ethylene 1-octene copolymers,” Polymer 41(19), 7067-7082. DOI: 10.1016/S0032-3861(00)00066-5

Sawpan, M. A., Pickering, K. L., and Fernyhough, A. (2011). “Effect of various chemical treatments on the fibre structure and tensile properties of industrial hemp fibres,” Compos. Part A-Appl. S. 42(8), 888-895. DOI: 10.1016/j.compositesa.2011.03.008

Sgriccia, N., Hawley, M. C., and Misra, M. (2008). “Characterization of natural fiber surfaces and natural fiber composites,” Compos. Part A-Appl. S. 39(10), 1632-1637. DOI: 10.1016/j.compositesa.2008.07.007

Shi, J., Shi, S. Q., Barnes, H. M., Horstemeyer, M. F., and Wang, G. (2011). “Kenaf bast fibers-Part II: Inorganic nanoparticle impregnation for polymer composites,” Int. J. Polym. Sci. 2011. DOI: 10.1155/2011/736474

Shikinami, Y., and Okuno, M. (2001). “Bioresorbable devices made of forged composites of hydroxyapatite (HA) particles and poly L-lactide (PLLA). Part II: Practical properties of miniscrews and miniplates,” Biomaterials 22(23), 3197-211. DOI: 10.1016/S0142-9612(01)00072-2

Singh, L., Kumar, V., and Ratner, B. D. (2004). “Generation of porous microcellular 85/15 poly (DL-lactide-co-glycolide) foams for biomedical applications,” Biomaterials 25(13), 2611-2617. DOI: 10.1016/j.biomaterials.2003.09.040

Tiainen, J., Soini, Y., Suokas, E., Veiranto, M., Törmälä, P., Waris, T., and Ashammakhi, N. (2006). “Tissue reactions to bioabsorbable ciprofloxacin-releasing polylactide-polyglycolide 80/20 screws in rabbits’ cranial bone,” J. Mater. Sci.-Mater. M. 17(12), 1315-1322. DOI: 10.1007/s10856-006-0606-8

Verheyen, C. C. P. M., De Wijn, J. R., Van Blitterswijk, C. A., and De Groot, K. (1992). “Evaluation of hydroxylapatite/poly(l‐lactide) composites: Mechanical behavior,” J. Biomed. Mater. Res. 26(10), 1277-1296. DOI: 10.1002/jbm.820261003

Wan, Y., Qu, X., Lu, J., Zhu, C., Wan, L., Yang, J., Bei, J., and Wang, S. (2004). “Characterization of surface property of poly (lactide-co-glycolide) after oxygen plasma treatment,” Biomaterials 25(19), 4777-4783. DOI: 10.1016/j.biomaterials.2003.11.051

Wang, C., Song, W., Cheng, H., Yu, X., Li, W., and Zhang, S. (2016). “Dipping modification with nano-CaCO3 to improve tensile properties of individual bamboo fiber for developing bamboo-plastic composite,” J. Nat. Fibers 13(6), 737-748. DOI: 10.1080/15440478.2015.1137526

Wang, C., Yu, X., Smith, L. M., Wang, G., Cheng, H., and Zhang, S. (2017). “Interfacial properties of bamboo fiber-reinforced high-density polyethylene composites by different methods for adding nano calcium carbonate,” Polymers-Basel 9(11), 587. DOI: 10.3390/polym9110587

Yu, Y., Huang, X., and Yu, W. (2014). “A novel process to improve yield and mechanical performance of bamboo fiber reinforced composite via mechanical treatments,” Compos. Part B-Eng. 56, 48-53. DOI: 10.1016/j.compositesb.2013.08.007

Zhang, Y., Yu, C., Chu, P. K., Lv, F., Zhang, C., Ji, J., Zhang, R., and Wang, H. (2012). “Mechanical and thermal properties of basalt fiber reinforced poly(butylene succinate) composites,”Mater. Chem. Phys. 133(2-3), 845-849. DOI: 10.1016/j.matchemphys.2012.01.105

Zhao, J., Liao, W., Wang, Y., Pan, J., and Liu, F. (2002). “Preparation and degradation characteristic study of bone repair composite of DL-polylactic acid/ hydroxyapatite/decalcifying bone matrix,” Chin. J. Traumatol. 5(6), 369-373.

Article submitted: September 6, 2018; Peer review completed: November 23, 2018; Revised version received: December 21, 2018; Accepted: December 28, 2018; Published: January 14, 2019.

DOI: 10.15376/biores.14.1.1694-1707