Anda di halaman 1dari 9

Phase morphology, physical properties, and biodegradation behavior of novel PLA/PHBHHx blends

Qiang Zhao,1 Shufang Wang,1 Meimei Kong,1 Weitao Geng,2 Robert K. Y. Li,3 Cunjiang Song,2 Deling Kong1
1

Key Laboratory of Bioactive Materials, Ministry of Education, College of Life Science, Nankai University, Tianjin 300071, Peoples Republic of China 2 Department of Microbiology, College of Life Science, Nankai University, Tianjin 300071, Peoples Republic of China 3 Department of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, Peoples Republic of China Received 21 April 2011; revised 12 June 2011; accepted 16 June 2011 Published online 26 September 2011 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/jbm.b.31915 Abstract: In this study, two biodegradable polyesters [i.e., polylactic acid (PLA) and poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx)] with complementarity in terms of mechanical performance have been combined, and a series of blends with a broad range of compositions has been prepared by thermal compounding. The evolution of phase morphologies with the variation of compositions has been characterized by using Fourier transform infrared spectroscopic imaging together with scanning electron microscope analyses. Thermal, mechanical, and biodegradation properties of the PLA/PHBHHx blends were systematically investigated. Mechanical properties were further analyzed by using theoretical models and correlated with the results of the morphology/structure and compatibility of the blends. Results indicate that PLA/PHBHHx blends are immiscible but can be compatible to some extent at certain compositions (e.g., PLA/PHBHHx (w/w) 80/20 and 20/80). The physical properties of the blend could be ne tuned by adjusting the C blend composition. V 2011 Wiley Periodicals, Inc. J Biomed Mater
Res Part B: Appl Biomater 100B: 2331, 2012.

Key Words: biodegradable, blends, phase behavior, poly(3hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx), polylactic acid (PLA)

How to cite this article: Zhao Q, Wang S, Kong M, Geng W, Li RKY, Song C, Kong D. 2012. Phase morphology, physical properties, and biodegradation behavior of novel PLA/PHBHHx blends. J Biomed Mater Res Part B 2012:100B:2331.

INTRODUCTION

Recently, studies of biodegradable polymers have attracted increasing attention due to the great demand in biomedical eld, including resorbable sutures, drug delivery systems, and tissue scaffolds.1,2 Currently, commercially available biodegradable polymers cannot fully satisfy the requirements of medical devices due to the limitation in performance, including physiochemical and biological properties. Polylactic acid (PLA) is a biodegradable polyester synthesized through chemical polymerization of monomers that are converted from renewable agricultural feedstock, such as corn or sugar beets.2 Since the 1980s, PLA has been studied extensively for biomedical usages because of its biodegradability and biocompatibility. PLA also has other desirable properties, including high mechanical strength, good transparent property, and fabricability.3 However, commodity products made of PLA have not gained widespread usage because of their relatively high cost and slow biodegradation rate for the environment.4 In addition, PLA is a

hard material with relatively low fracture toughness, which often limits its usage to some extent, especially in the biomedical eld. Polyhydroxyalkanoates (PHAs) represent another type of biodegradable polymers, which are biosynthesized by some microorganisms as intracellular carbon and energy storage compounds.58 Because of the bacterial origins, this class of polyesters shows good degradability. PHAs have been used in number of biomedical applications, including tissue engineering scaffold, drug delivery system, resorbable surgical sutures, and so on.6,9 Poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx) is a new member of the PHA family.1012 Due to the longer alkyl side chain, PHBHHx exhibits low crystallinity and a broad processing window compared with conventional poly(3-hydroxybutyrate) (PHB). The mechanical performance is utterly different from that of PHB, that is, PHBHHx is a soft and exible polymer with a relatively higher break elongation but lower tensile strength and modulus.7

Correspondence to: S. Wang; e-mail: wangshufang@nankai.edu.cn or C. Song; e-mail: songcj@nankai.edu.cn Contract grant sponsor: National Natural Science Foundation of China; contract grant numbers: 50830104, 51073081, 81000680 Contract grant sponsor: Tianjin Committee of Science and Technology, PR China (Key Projects); contract grant numbers: 09JCZDJC18400, 09ZCKFSH00800)

C V 2011 WILEY PERIODICALS, INC.

23

Therefore, these two polymers (i.e., PLA and PHBHHx) are both biodegradable and hold complementary potential in terms of mechanical properties. Blending is a cost effective strategy for combining the advantages of different polymers. In addition, some additional desirable properties that neither component possesses alone may also be obtained.13,14,15 In comparison with the studies on commodity polymers, the pertinent literature has paid very little attention to the phase behavior of biodegradable polymer blends, which is a key factor in determining physical performance as well as potential applications of the nal product. Fourier transform infrared (FTIR) spectroscopic imaging provides an efcient tool for characterizing the phase structure and morphology of polymer blends.16,17 This method allows spectral and spatial analyses in a single measurement, so the spatial distribution of different components in a blend can be visualized by identifying their characteristic IR bands. In this study, the evolution of the phase morphology of PLA/PHBHHx blends over a broad range of compositions has been characterized by using FTIR spectroscopic imaging. The physical properties and degradation behavior also have been investigated systematically. With thorough understanding of the structure-property relationship, it aims at designing and manufacturing PLA/PHBHHx blends with desirable properties for biomedical applications, such as articial vascular graft.
EXPERIMENTAL

The weight of the remaining material was measured to determine the relative biodegradation.19 Characterization FTIR spectroscopic imaging analysis was performed by using a Perkin-Elmer Spotlight 100 system in attenuated total reectance (ATR) mode (diamond crystal). The size of one detector pixel was 6.25 6.25 lm2. The spectral resolution was 4 cm1, and four scans were co-added for each spectrum. Differential scanning calorimetry (DSC) measurements were conducted using a Netzsch 204 thermal analysis system in a nitrogen atmosphere. For the DSC measurements, samples of approximately 10 mg in weight were rst heated from 25 to 200 C at a heating rate of 10 C/min (rst heating scan) and were kept at 200 C for 1 min to eliminate thermal history. The samples were then rapidly cooled to 100 C at a cooling rate of 50 C/min (cooling scan). Finally, the samples were heated from 100 to 200 C at the rate of 10 C/min (second heating scan). The glass transition temperature (Tg) was taken as the midpoint of the heat capacity change. The clod crystallization temperature (Tcc), and melting temperature and corresponding enthalpy (Tm and DHm) were determined from exothermic and endothermic peaks in the second heating scans, respectively. Thermogravimetric analysis (TGA) measurements were conducted on a Netzsch TG 209 analyzer. Samples of approximately 10 mg in weight were used for each measurement. The samples were heated at a heating rate of 10 C/min from room temperature to 600 C, under a nitrogen atmosphere. Static tensile tests were performed by a Testometric universal tensile tester (United Kingdom) with a crosshead speed of 10 mm/min at ambient temperature. For the tensile tests, dumbbell-shaped specimens were used. Each testing was repeated on ve specimens, and the mean values as well as standard deviation (SD) were reported. The sample of PLA/PHBHHx (80/20) was fractured in liquid nitrogen, and the fractured surface was immersed in acetone and heated for 15 min to selectively extract the PHBHHx component. Then, the solvent was replaced, and the sample remained in the solvent overnight at room temperature. The solvent-etched sample was dried in a vacuum to completely evaporate the solvent. After gold coating, the morphology of the fractured surface was analyzed by using a scanning electron microscope (SEM) (JEOL JSM-820) with an accelerating voltage of 20 kV.
RESULTS

Materials The PLA (Mn 47,000; Mw 192, 000) was obtained from Shimadzu Co. (Japan). PHBHHx (Mn 292,000; Mw 624, 000) was provided by Kanaka Co. (Japan), and the mole ratio of 3HHx is 20%. Sample preparation The PLA/PHBHHx blends were prepared by melt mixing with a twin-screw extruder (Kurimoto (Japan) (/25 mm, L/D 10.2) at 190 C and 100 rpm. The extrudates were granulated into pellets. A series of blend specimens was prepared with weight ratios (PLA/PHBHHx) of 100/0, 80/20, 60/40, 50/50, 40/60, 20/80, and 0/100. The melt-compounded blends were compression molded into sheets of 0.2 mm thickness at 130190 C. In vitro biodegradation Each of PLA/PHBHHx blends (100/0, 80/20, 60/40, 40/60, 20/80, and 0/100) was submerged in 30 mL of 0.1M phosphate-buffered saline (PBS) solution in a 50 mL test tube at physiological temperature (37 6 2  C). The ionic concentration was adjusted to the physiological range as described in the ASTM Designation: F 1635-04a.18 Sodium azide (0.1% (w/w)), penicillin (100 U/mL), and streptomycin (100 lg/mL) were added to the solution as the antimicrobials to prevent bacterial growth. The specimens were immersed in the physiological solution, and removed at certain time intervals, washed with 75% ethanol and distilled water, and dried to constant weight under vacuum.

Phase morphology In this study, the phase morphologies of PLA/PHBHHx blends with a broad range of compositions were investigated by using FTIR spectroscopic imaging analysis in ATR mode. FTIR spectra of neat PLA and PHBHHx were characterized by their intense absorption bands at 1750 and 1720 cm1, respectively, which are due to the stretching of the carbonyl group (C (Figure 1). By mapping the relative intensity of O)

24

ZHAO ET AL.

BIODEGRADABLE PLA/PHBHHX BLENDS

ORIGINAL RESEARCH REPORT

FIGURE 1. FTIR spectra for neat PLA and PHBHHx. [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

these two bands, the PLA-specic and PHBHHx-specic FTIR spectroscopic images for the blends can be plotted, and the result is shown in Figure 2. The red and blue colors represent the higher and the lower concentration of the corresponding polymer, respectively. PLA/PHBHHx blends with intermediate compositions of PLA/PHBHHx (with blend ratios that range from 40/60 to 60/40) show macroscopic phase separation, and PHBHHxspecic and PLA-specic images were perfectly complementary [Figure 2(AC)]. The 40/60 PLA/PHBHHx blend exhibits a dual-phase continuous morphology. Segregated PHBHHx (e.g., spot 15 in Figure 2) and PLA (e.g., spot 17 in Figure 2) phases had spectra that were typical of the neat component and could be clearly distinguished. The dimensions of these domains approach about 100 lm. In addition, an inter-phase region (green-colored region, spot 16 in Figure 2) was evident, where two components coexist with comparable peak intensities. To acquire further insight about phase morphology, we did additional analyses on the system with equal composition (i.e., PLA/PHBHHx (50/50)). The corresponding FITRmicroscopic images indicated that the distribution characteristics of two phases apparently were different. In the PHBHHx-rich phase, there is a certain degree of mixing between the two components, which resulted in a decrease in the size of the separated PHBHHx domains (by several tens of microns) compared with those for PLA/PHBHHx (40/60). In contrast, The PLA-rich phase was homogeneous in composition with no signs of mixing. Based on these results, it is reasonable to conclude that the PHBHHx-rich phase is more compatible than the PLA-rich phase. Further decreasing the PHBHHx content leads to a ner phase morphology with decreased domain size in PLA/ PHBHHx (60/40). The continuous phase is PLA (e.g., spot 7 in Figure 2), and the dispersed phases are those irregularshaped domains that consist of both components with various compositions, which were reected by the intensity ratios of two characteristic peaks (e.g., spots 5 and 6 in Figure 2). Again, this morphology reects the high level of compatibility in the PHBHHx-rich phase. Some fraction of PLA mixed with predominant PHBHHx phase with certain

degree of compatibility, while residual PLA has been segregated and presented as dispersed domains due to the limited compatibilization capability of PHBHHx matrix. For the blends in which one of the components has a relatively minor concentration (e.g., PLA/PHBHHx (20/80) and PLA/PHBHHx (80/20)), both PHBHHx-specic and PLAspecic images show a homogenous pattern, and no signs of phase separation can be identied. Only slight, regular streaks were observed, the appearance of which is due to the relatively narrow numerical scale over the color scale bar. The spectra of the spots selected from different regions had two characteristic peaks with the intensity ratios remaining almost unchanged. Despite the homogenous pattern reected by the FTIRmicroscopic analysis, it would be arbitrary to draw the conclusion that these blends (i.e., PLA/PHBHHx (20/80) and PLA/PHBHHx (80/20)) are miscible, one-phase systems. As has been mentioned earlier, the resolution of the FITRmicroscope is 6.25 lm; therefore, the phase separation with domain size similar to and/or below this dimension cannot be effectively differentiated. We further investigated the phase structure by using higher magnication SEM analyses in the case of PLA/PHBHHx (80/20). Before analysis, the minor PHBHHx phase was selectively etched by acetone, because acetone is an effective solvent for PHBHHx, whereas it will not dissolve PLA. The SEM image demonstrated that there were craters (arrow indicated) in the surviving matrix, which have been left by the solvent extraction of the PHBHHx phase. The domain size can be approximately estimated within the range of several microns (Figure 3). In contrast with the macroscopic phase-separated system (such as PLA/PHBHHx (40/60)), the blend of PLA/PHBHHx (80/20) shows a higher degree of compatibility, that is, the dispersed phase has a ner distribution with remarkably reduced domain size. Furthermore, some desirable properties (e.g., improved mechanical performance) can be anticipated from the compatible blends.15,20 Thermal properties The thermal properties of PLA/PHBHHx blends were obtained from the second heating scan of the DSC (Figure 4 and Table I). The rst heating scan was used to eliminate the thermal history of the specimens that were prepared by thermal processing. Neat PHBHHx is a partially crystalline polymer with low crystallinity owing to the bulky side group. The melting of PHBHHx shows a double peak, and the shoulder peak at the lower temperature corresponds to the melting of the imperfect crystalline PHBHHx and the subsequent re-crystallization. After the incorporation of the PLA component, the crystallization of the PHBHHx phase was further suppressed, and no detectable melting peak was identied. By comparing the heating scans with those of neat PLA and PHBHHx, the exothermal and endothermal peaks evident in the thermograms of PLA/PHBHHx blends were ascribed to cold crystallization of the PLA component, followed by subsequent melting. The variation of cold crystallization temperatures (Tcc) with blend compositions is somewhat complex, that is, the Tcc values of PLA/PHBHHx

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | JAN 2012 VOL 100B, ISSUE 1

25

FIGURE 2. FTIR-microscopic images for the PLA/PHBHHx blends and the corresponding spectra for the labeled spots. The size of the image is 250 250 lm. Column 1: PHBHHx-specic image; Column 2: PLA-specic image; Column 3: Spectra at certain spots. A: PLA/PHBHHx (40/60); B: PLA/PHBHHx (50/50); C: PLA/PHBHHx (60/40); D: PLA/PHBHHx (80/20); and E: PLA/PHBHHx (20/80). [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

(60/40) and PLA/PHBHHx (40/60) are higher than that of neat PLA, whereas that of PLA/PHBHHx (80/20) is slightly lowered. PLA/PHBHHx (20/80) shows a small, broad exothermal peak, and the peak temperature cannot be precisely

determined. The possible reason explaining the depressed Tcc in PLA/PHBHHx (80/20) is that, during the cooling process, the minor PHBHHx component forms ne spherulites, which can act as nucleating sites to accelerate the

26

ZHAO ET AL.

BIODEGRADABLE PLA/PHBHHX BLENDS

ORIGINAL RESEARCH REPORT

FIGURE 3. Scanning electron micrograph (SEM) of fractured surface after solvent etching for PLA/PHBHHx blend (80/20).

nonisothermal crystallization of the PLA phase in the subsequent heating process.21 The glass transition can also be identied from the DSC thermograms. Generally, the blends show two distinctive glass transition temperatures (Tg) that correspond to the two components, respectively. Of these two Tg values, the lower value belonged to the PHBHHx component, which is independent of composition and remains almost unchanged in all systems; the higher Tg value, which is associated with PLA, showed a small, but still evident, shift toward low temperature in the cases of PLA/PHBHHx (80/20) and PLA/PHBHHx (20/80). It has been accepted that the glass transition is a kinetic phenomenon that is strongly inuenced by the local material environment. Previous phase morphology analyses have indicated that these two blends have more compatible characteristics, which, in turn, results in the shift of Tg.22 The thermal degradation behavior of the PLA/PHBHHx blends was evaluated by using TGA (Figure 5). The blends showed two, well-dened decomposition peaks in the DTG curves, which were attributed to the thermal degradations of the PLA and PHBHHx components by comparing them with the decomposition of neat polymers without compounding. Detailed results are summarized in Table I. The decomposition temperatures of the PHBHHx component (Tp(I)) in all blends were increased compared with the neat

FIGURE 4. DSC thermograms of PLA/PHBHHx blends during the second heating scan (10 C/min). [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

PHBHHx, and this enhancement is more signicant in compatible systems, such as PLA/PHBHHx (80/20) and PLA/ PHBHHx (20/80). Published data have indicated that one prominent disadvantage of PHA polymers (e.g., PHB) is their narrow processing windows. For the PLA/PHBHHx blends investigated in this study, the ranges of the processing temperature (i.e., T0(I)Tm(max)) are relatively broad with values larger than 70 C; therefore, the processability of this kind of materials is acceptable for large-scale industrial applications. Mechanical properties The mechanical performances of PLA/PHBHHx blends were evaluated by tensile testing (Figure 6 and Table II). Generally, neat PHBHHx is a soft, exible polyester with relatively low tensile stress and Youngs modulus (less than 1 GPa), while PLA has higher strength and hardness but lower toughness and ductility.23,24 Previous studies have gained success in toughening the PLA matrix by the incorporation of second ductile polymers, such as PBAT (poly(butylene adipate-co-terephthalate)25 or PCL.22,26 On the basis of the mechanical characteristic of the

TABLE I. Thermal Transition Data of PLA/PHBHHx Blends Obtained from DSC and TGA Analyses PLA Sample Code PLA PLA/PHBHHx PLA/PHBHHx PLA/PHBHHx PLA/PHBHHx PHBHHx
a b

PHBHHx DHm (J/g) 43.69 37.36 19.44 17.40 9.19 Tp ( C) 372.7 371.2 366.6 357.5 337.5 Tg ( C) 5.4 3.3 3.9 3.0 3.8 Tm ( C) 108.0 (115.8)b Tp ( C) 289.7 281.2 274.2 284.0 272.8

Blend Processing Windowa 144.4 92.4 72.0 71.1 77.7 124.2

Tg ( C) (80/20) (60/40) (40/60) (20/80) 60.7 55.1 59.1 59.6 55.6

Tcc ( C) 117.4 110.4 118.2 119.1

Tm ( C) 168.9 (173.1)b 173.8 173.0 173.9 172.3

T0(I) Tm(max). Double peaks. Tg: glass transition temperature; Tcc: cold crystallization temperature; Tm: melting temperature; Tp: peak temperature of the DTG curve;DHm: melting enthalpy; T0: temperature at which weight loss started to occur; Tm(max): the higher one of the two Tms.

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | JAN 2012 VOL 100B, ISSUE 1

27

(i.e., PLA/PHBHHx (80/20)), the elongation at break (eb) of which is almost 600% higher than that of neat PLA even taking the relatively large SD into account. The toughening effect becomes less effective on further increasing the PHBHHx loading (i.e., PLA/PHBHHx (60/40) and PLA/ PHBHHx (50/50)) due to macroscopic phase-separation morphology illustrated by the FTIR-microscope image, which may induce severe stress concentration, and, therefore, signicantly degrade fracture toughness. On the other hand, the tensile strength and Youngs modulus were adversely declined with the PHBHHx addition. At the other composition extreme, the introduction of minor amount of PLA effectively reinforced the soft and weak PHBHHx matrix, that is, the tensile stress and Youngs modulus were apparently enhanced in case of PLA/PHBHHx (20/80). In addition, an evident yielding point was identied in the stress-strain curve, and the elongation at break (eb) was increased compared with the neat PHBHHx. This reinforcing effect was ascribed to the ne distribution of stiff PLA particles of small size within the PHBHHx continuous phase. The variation tendency of Youngs modulus with the blend composition was further tted by using theoretical models of rule of mixture and foam model, respectively.26,27 The rule of mixture [Eq. (1)] assumes the perfect inter-phase adhesion between the matrix and the dispersed phase and perfect dispersion of the spherical inclusion in the matrix:  Eb
FIGURE 5. TGA curves of PLA/PHBHHx blends at a heating rate of 10 C/min: (a) TGA curves; (b) DTG curves. [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

  Ed 1 /d 1 Em Em

(1)

PHBHHx, the incorporation of it can improve the ductility and exibility of the PLA matrix. Evident toughening effect has been observed with the addition of 20 wt % of PHBHHx

where /d is the volume ratio of the dispersed phase which has been determined according to the density of the two polymers (DPLA 1.24 g/cm3; DPHBHHx 1.21 g/cm3), Eb, Ed, and Em are the modulus of the blend, dispersed phase, and matrix. In the Foam model [Eq. (2)], dispersed phase is considered as noninteracting phase equivalent to a void or pores: Eb 1 /d Em
2=3

(2)

The plot of Youngs modulus against blend ratio follows the rule of mixture more closely and is apparently higher than the prediction based on foam model, revealing a

TABLE II. Mechanical Properties of PLA/PHBHHx Blends Determined from Stress-Strain Curves Sample Code PLA PLA/PHBHHx PLA/PHBHHx PLA/PHBHHx PLA/PHBHHx PLA/PHBHHx PHBHHx (80/20) (60/40) (50/50) (40/60) (20/80) Youngs Stress at Elongation Modulus (GPa) Max (MPa) at Break (%) 1.39 1.32 1.24 0.91 1.25 0.59 0.37 6 6 6 6 6 6 6 0.09 0.04 0.16 0.07 0.15 0.18 0.12 36.4 29.5 33.5 22.1 27.7 23.6 17.6 6 6 6 6 6 6 6 3.9 0.9 4.0 0.9 1.8 0.7 1.4 13.8 99.6 7.68 7.26 11.5 83.5 19.3 6 6 6 6 6 6 6 5.7 69.4 1.60 2.72 4.3 64.2 1.4

FIGURE 6. Stress-strain curves of PLA/PHBHHx blends obtained from static tensile testing. [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

28

ZHAO ET AL.

BIODEGRADABLE PLA/PHBHHX BLENDS

ORIGINAL RESEARCH REPORT

FIGURE 7. Plot of Youngs modulus against blend ratio based on experimental data and theoretical prediction. [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

certain degree of interfacial adhesion (Figure 7). Direct evidence for the interfacial adhesion has already been demonstrated by FTIR microscope image [Figure 2(A,B)]. In vitro biodegradation behavior The in vitro biodegradation behavior of PLA/PHBHHx blends with various compositions was evaluated in the simulated body uid (PBS buffer, pH 7.4). Figure 8 shows the relationship between weight loss of PLA/PHBHHx-blend lms and the degradation time. In general, all blends degraded slowly as well as the neat polymers up to 5 months, which agreed with the reported data under similar conditions.28 It has been reported PLA began to experience signicant weight loss after a time period of 16 months; therefore, the degradation remained in the introduction period during the time period investigated in this study. It is reasonable to predicate that the in vivo degradation of these materials should be faster than in vitro process because some enzymes present in human body can catalyze the hydrolysis of the polymer chains. Despite the relative low absolute value for weight loss, they still show a regular variation trend with blend composition, that is, the weight retention decreased with the decrease of PHBHHx loading in general, which is the inverse of that reected by the simulated soil medium (data not shown). It is worth noting that no signicant change was observed for the neat PLA and blend with PLA as predominant phase (i.e., PLA/PHBHHx (80/20)) during the rst month. This phenomenon could be attributed to the relatively high crystallinity of PLA phase compared with PHBHHx (as shown in Table I), which can inhibit the diffusion of degraded product from the bulky material.
DISCUSSION

properties. The two polymers were combined via blending with the aim to fabricate novel material with comprehensively optimal performance. As far as we know, the nal physical properties of the polymer blend are determined by the compatibility and phase structure/ morphology, and macroscopic phase separation often results in the poor mechanical performance. In essence, the blend developed in this study is thermodynamically immiscible, because the two components (i.e., PLA and PHBHHx) are both semicrystalline polyesters and cannot undergo cocrystallization. The FTIR-microscopic and SEM analyses indicate that the phase morphology evolves continuously with the variation of compositions, which leads to phase domains that gradually vary in size. Furthermore, it is well established that, even in the case of immiscible blend systems, molecular chains of different components can interpenetrate at the inter-phase locations to some extent.29 In addition, the constituents of polyester blends readily undergo interchange reactions near or above the melting points of the blends. For the processing conditions used in this study, some transesterications may take place at the interface of two phases, and block copolymers may thus be formed. These copolymers can improve the local miscibility, which in turn promotes further contact between the two components.14 As a result, certain degree of compatibility has been observed in the blends with relatively minor second component (i.e., PLA/PHBHHx (w/w) 80/20 and 20/80), due to the ne distribution of separated phase with reduced size. Enhanced compatibility was also supported by the shift in glass transition temperature demonstrated in DSC thermogram. The mechanical properties are generally in agreement with the described phase structure and morphology and conrm the complementarity of these two polymers. In detail, the incorporation of a minor amount of the second component can effectively overcome the drawbacks of the predominant matrix due to the better compatibility

PLA and PHBHHx are two biodegradable polyesters with evident complementary potential in terms of mechanical

FIGURE 8. Changes of weight remaining for PLA/PHBHHx blends in PBS buffer at 37 C. [Color gure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | JAN 2012 VOL 100B, ISSUE 1

29

and ne phase distribution. Evident toughening and reinforcing effect have been observed in the case of PLA/ PHBHHx (w/w) 80/20 and 20/80, respectively. The mechanical performance of as-prepared blends could satisfy a wide range of biomedical usages through ne tuning the blend composition. Interfacial behavior of polymer blends (e.g., perfect adhesion and no adhesion) is very important in determining the nal properties. Various approaches have been used to characterize the interfacial behavior, including SEM, dynamic viscoelasticity, and so forth. In addition, prediction models offer an expedient alternative for recognizing and quantifying to some extent the interfacial behavior of polymer blends.22,26,27 In this study, the variation tendency of mechanical performance with the blend composition has also been tted by using theoretical models of rule of mixture and foam model. Results indicate there is a certain degree of interfacial adhesion in PLA/PHBHHx blends, which has been supported by FTIR microscopy analysis. Biodegradation behavior of PLA/PHBHHx blends evaluated in simulated body uid environment show a decreasing trend with the increase of PHBHHx loading. The dominant mechanism for in vitro degradation of the polyester is the scission of ester bonds, and macromolecular chains were hydrolyzed into water-soluble oligomers and then released from the bulky lm into the surrounding medium, which results in the weight loss. The acidic degradation product can catalyze and therefore accelerate the hydrolysis process.30 It has been accepted that the hydrolysis rate is proportional to the concentration of carboxyl groups of the degraded products. Hence, PLA with shorter repeating unit (C3) is more liable to be hydrolyzed than PHBHHx which consists of six carbon atoms in each repeating unit.

REFERENCES
1. Grifth LG. Polymeric biomaterials. Acta Mater 2000;48:263277. 2. Gross RA, Kalra B. Biodegradable polymers for the environment. Science 2002;297:803807. 3. Engelberg I, Kohn J. Physico-mechanical properties of degradable polymers used in medical applications: A comparative study. Biomaterials 1991;12:292304. 4. Ramsay BA, Langlade V, Carreau PJ, Ramsay JA. Biodegradability and mechanical properties of poly-(b-hydroxybutyrate-co-bhydroxyvalerate)-starch blends. Appl Environ Microbiol 1993;59: 12421246. 5. Anderson AJ, Dawes EA. Occurrence, metabolism, metabolic role, and industrial uses of bacterial polyhydroxyalkanoates. Microbiol Rev 1990;54:450472. 6. Doi Y. Microbial Polyesters. New York: VCH Publishers; 1990. 7. Gao Y, Kong LJ, Zhang L, Gong YD, Cheng GG, Zhao NM, Zhang XF. Improvement of mechanical properties of poly(DL-lactide) lms by blending of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Eur Polym J 2006;42:764775. 8. Inoue Y, Yoshie N. Structure and physical properties of bacterially synthesized polyesters. Prog Polym Sci 1992;17:571610. 9. Misra SK, Valappil SP, Roy I, Boccaccini AR. Polyhydroxyalkanoate (PHA)/inorganic phase composites for tissue engineering applications. Biomacromolecules 2006;7:22492258. 10. Chen GQ, Wu Q. Polyhydroxyalkanoates as tissue engineering materials. Biomaterials 2005;26:65656578. 11. Qu XH, Wu Q, Liang J, Qu X, Wang SG, Chen GQ. Enhanced vascular-related cellular afnity on surface modied copolyesters of 3-hydroxybutyrate and 3-hydroxyhexanoate (PHBHHx). Biomaterials 2005;26:69917001. 12. Cheng S, Chen GQ, Leski M, Zou B, Wang Y, Wu Q. The effect of D,L-b-hydroxybutyric acid on cell death and proliferation in L929 cells. Biomaterials 2006;27:37583765. 13. Wang XL, Du FG, Jiao J, Meng YZ, Li RKY. Preparation and properties of biodegradable polymeric blends from poly(propylene carbonate) and poly(ethylene-co-vinyl alcohol). J Biomed Mater Res 2007;83B:373379. 14. Groeninckx G, Sarkissova M, Thomas S. Chemical reactions in blends based on condensation polymers: Transreactions and molecular and morphological characterization, Chapter 14. In: Paul DR, Bucknall CB, editors. Polymer Blends, Vol. 1, Formulation, USA: Wiley; 1999. 15. Hobbs SY, Watkins VH. Morphology characterization by microscopy techniques, Chapter 9. In: Paul DR, Bucknall CB, editors. Polymer Blends, vol.1, Formulation. USA: Wiley; 1999. 16. Vogel C, Wessel E, Siesler HW. FT-IR imaging spectroscopy of phase separation in blends of poly(3-hydroxybutyrate) with poly(L-lactic acid) and poly(e-caprolactone). Biomacromolecules 2008;9:523527. 17. Malheiro VN, Caridade SG, Alves NM, Mano JF. New poly(e-caprolactone)/chitosan blend bers for tissue engineering applications. Acta Biomater 2010;6:418428. 18. ASTM Designation: F 1635-04a, Standard test method for in vitro degradation testing of hydrolytically degradable polymer resins and fabricated forms for surgical implants. 19. Tao J, Song CJ, Cao MF, Hu D, Liu L, Liu N, Wang SF. Thermal properties and degradability of poly(propylene carbonate)/poly(bhydroxybutyrate-co-b-hydroxyvalerate) (PPC/PHBV) blends. Polym Degrad Stab 2009;94:575583. 20. Zhang L, Xiong C, Deng X. Miscibility, crystallization and morphology of poly(b-hydroxybutyrate)/poly(d,l-lactide) blends. Polymer 1996;37:235241. 21. Furukawa T, Sato H, Murakami R, Zhang J, Duan YX, Noda I, Ochiai S, Ozaki Y. Structure, dispersibility, and crystallinity of poly(hydroxybutyrate)/poly(L-lactic acid) blends studied by FT-IR microspectroscopy and differential calorimetry. Macromolecules 2005;38:64456454. 22. Broz ME, VanderHart DL, Washburn NR. Structure and mechanical properties of poly(D,L-lactic acid)/poly(e-caprolactone) blends. Biomaterials 2003;24:41814190. 23. Takayama T, Todo M. Improvement of impact fracture properties of PLA/PCL polymer blend due to LTI addition. J Mater Sci 2006; 41:49894992.

CONCLUSIONS

In this study, we have prepared a type of biodegradable PLA/PHBHHx blends by melt mixing. The phase structures and morphologies of these blends were characterized by using FTIR-spectroscopic imaging together with SEM analyses. Thermal, mechanical, and in vitro biodegradation properties of the PLA/PHBHHx blends were systematically characterized. The variation tendency of mechanical performance was further tted by using theoretical models to explore the inherent structure-property relationship. The results indicate that PLA/PHBHHx blends are immiscible but can be compatible to some extent at certain compositions. The phase structure and morphology evolve continuously with the variation of compositions, leading to gradually varied sizes of the phase domains. The two comprising polymers hold complementary capability in terms of mechanical properties. Incorporation of a minor amount of the second component can improve the mechanical performance remarkably by providing a toughening or a reinforcing effect in compatible systems, such as PLA/PHBHHx (80/20) and PLA/PHBHHx (20/80).

30

ZHAO ET AL.

BIODEGRADABLE PLA/PHBHHX BLENDS

ORIGINAL RESEARCH REPORT

24. Todo M, Park SD, Takayama T, Arakawa K. Fracture micromechanisms of bioabsorbable PLLA/PCL polymer blends. Eng Fract Mech 2007;74:18721883. 25. Jiang L, Wolcott MP, Zhang J. Study of biodegradable polylactide/poly(butylene adipate-co-terephthalate) blends. Biomacromolecules 2006;7:199207. 26. Simoes CL, Viana JC, Cunha AM. Mechanical properties of poly(e-caprolactone) and poly(lactic acid) blends. J Appl Polym Sci 2009;112:345352. 27. Tomar N, Maiti SN. Mechanical properties of PBT/ABAS blends. J Appl Polym Sci 2007;104:18071817.

28. Tsuji H. In vitro hydrolysis of blends from enantiomeric poly(lactide)s Part 1. Well-stereo-complexed blend and non-blended lms. Polymer 2000;41:36213630. 29. Sakai F, Nishikawa K, Inoue Y, Yazawa K. Nucleation enhancement effect in poly(L-lactide) (PLLA)/poly(e-caprolactone) (PCL) blend induced by locally activated chain mobility resulting from limited miscibility. Macromolecules 2009;42: 83358342. 30. Cha Y, Pitt CG. The biodegradability of polyester blends. Biomaterials 1990;11:108112.

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | JAN 2012 VOL 100B, ISSUE 1

31

Anda mungkin juga menyukai