Você está na página 1de 7

Polymer Degradation and Stability 97 (2012) 1822e1828

Contents lists available at SciVerse ScienceDirect

Polymer Degradation and Stability


journal homepage: www.elsevier.com/locate/polydegstab

Thermal, mechanical and morphological characterization of plasticized PLAePHB blends


Mohamed A. Abdelwahab a, b, Allison Flynn b, Bor-Sen Chiou c, Syed Imam c, William Orts c, Emo Chiellini a, *
a

Laboratory of Bioactive Polymeric Materials for Biomedical and Environmental Applications (BIOlab), UdR INSTM, Department of Chemistry & Industrial Chemistry, University of Pisa, Via Risorgimento 35, 56126 Pisa, Italy Lapol, LLC, 55 Castilian Dr., Santa Barbara, CA 93117, USA c USDA, Agricultural Research Services, WRRC-BCE 800 Buchanan Street Albany, CA 94710, USA
b

a r t i c l e i n f o
Article history: Received 23 April 2012 Received in revised form 25 May 2012 Accepted 28 May 2012 Available online 4 June 2012 Keywords: Poly(lactic acid) Poly(3-hydroxybutyrate) Blend Plasticizer Thermal stability Mechanical properties

a b s t r a c t
A blend of poly(lactic acid) (PLA) (75% by weight) and poly(3-hydroxybutyrate) (PHB) (25% by weight) with a polyester plasticizer (Lapol 108) at two different concentrations (5 and 7% by weight per 100 parts of the blends) were investigated by TGA, DSC, XRD, SEM, mechanical testing and biodegradation studies. PLA/PHB blends showed a good distribution of the major components and absence of phase separation. XRD showed that the original crystal structure of PHB in the PLA75/PHB25 blend had been disturbed. DSC curves of PLA or PHB with plasticizer exhibited one Tg value, indicating that both major blend components are miscible. The Tg values also decreased with increased amount of plasticizer and showed good correlation to the Fox Equation, The melting temperature of PLA and PHB blends mostly did not change with an increase in plasticizer content, and the thermal stability of PLA and PHB was not affected. Also, the elongation at break of the PLA/PHB blend was greatly improved with the addition of plasticizer. In addition, in preliminary biodegradation studies carried in natural compost neat PHB showed some biodegradation, whereas the samples containing PLA did not experience a substantial biodegradation. This last aspect is worthy of further investigation in a more comprehensive and detailed approach. 2012 Elsevier Ltd. All rights reserved.

1. Introduction In recent years, sustainability, environmental concerns and green chemistry have played a big role in guiding the development of the next generation of materials, products and processes. The persistence of plastics in the environment, dwindling petroleum resources, shortage of landll space and the concerns over emissions of toxic gases during incineration have fuelled efforts to develop biodegradable polymers from renewable resources. Particularly, renewable agricultural and biomass feedstock have shown much promise for use in eco-efcient packaging to replace petroleum feedstock without competing with food crops [1]. Poly(lactic acid) (PLA) is a biodegradable and biocompatible crystalline polymer that can be produced from renewable resources which is of great interest to the packaging industry due to its

* Corresponding author. Tel.: 39 050 2210301; fax: 39 050 2210332. E-mail addresses: m_abedelwahab@ns.dcci.unipi.it (M.A. Abdelwahab), emochie@dcci.unipi.it (E. Chiellini). URL: http://syed.imam.ars.usda.gov 0141-3910/$ e see front matter 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.polymdegradstab.2012.05.036

availability and low cost [2e5]. The properties of PLA are dependent on the ratio between the two enantiomers D and L. PLA can show crystalline polymorphism, which can lead to different melting peaks [6,7]. Poly(3-hydroxybutyrate) (PHB) is a typical biodegradable thermoplastic polymer produced by microorganisms. PHB is highly crystalline polyester (80%) with a high melting point (173e180  C) and a glass transition temperature (Tg) of approximately 1e5  C [8e10]. Both PLA and PHB are biodegradable polyesters and are used in consumer products by several industrial sectors due to their biocompatibility, biodegradability and sustainability. They have comparable thermal and mechanical properties to those of some conventional polymers and this has generated much interest in exploring their physical and processing properties for potential applications [6,11,12]. It is noteworthy that both PLA and PHB have poor processing properties and are brittle at room temperature [13]. Thus, several modications have been proposed to improve their processing and mechanical properties, which include the preparation of copolymers containing monomeric units derived from hydroxyl alkanoic acids and blending these with PLA and PHB.

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828

1823

Blending is much easier and faster than copolymerization methods [14e27]. Previous studies on PLA/PHB blends showed that the miscibility between these two polymers depended on the molecular weight of the second component [13,17,19,28]. PHB is miscible with low molecular weight PLA (Mw < 18,000) in the melt over the whole composition range, whereas the PHB blends with high molecular weight PLA (Mw > 18,000) show biphasic separation [17,19]. Similarly, the PLA component is miscible with low molecular weight PHB (Mw 9400) in the melt with the PHB content up to 50 wt-%, and is immiscible with high molecular weight PHB and commercial-grade bacterial PHB [13,28]. PLA/PHB blends properties also depend on composition, chemical or physical cross-links and processing conditions. Zhang et al. [27] investigated the morphology, structure, crystallization behaviour, thermal properties, mechanical properties, and biodegradation of PLA/PHB blends, which were prepared by using melt compounding. The PLA/PHB 75/25 blend showed improved mechanical properties and heat distortion temperature and presence of some interactions. Another way to improve blend compatibility is by using a compatibiliser as a third component. Such compatibilisers include glycidyl methacrylate (GMA) [29], poly(ethylene glycol) (PEG) [26] and poly(vinyl acetate) (PVAc) [25]. PLA can be plasticized using oligomeric lactic acid (OLA), citrate ester or low molecular weight additives [7,30,31]. Plasticization increases the chain mobility and favours PLA crystallization. After plasticization, the crystallinity is in the range between 20 and 30%. The drawback of using these low molecular weight additives is that they leach out of the PLA matrix, resulting in a brittle plastic over time. PHB can be plasticized with dioctyl phthalate, 1,4cyclohexanedimethanol dibenzoate, tert-butylphenyl phosphate, PEG, triacetyl glycerol, tributyrin, dodecanol, lauric acid, trilaurin and low molecular weight poly(adipate) [32e34]. Grillo Fernandes et al. [35] blended PHB with polyols containing different amounts and types of hydroxyl groups. Glycerol and PEG200 have been shown to decrease the thermal stability of PHB. In contrast, tri(ethylene glycol) bis(2-ethylhexanoate) (TEGB) and pentaerythritol had no effect on thermal stability of PHB. The mechanical testing showed all samples were brittle, especially with PHB containing 20 wt-% of PEG. DSC results indicated that PEG200 and TEGB are the most compatible with PHB. At the USDA, a new polyester plasticizer has been developed from mostly renewable resources. This plasticizer is in commercial development and goes by the commercial name of Lapol 108 [36]. This plasticizer has high molecular weight (>80,000 Da), is highly branched and is derived from more than 50% renewable resources. In this paper, we studied the effects of Lapol 108 on PHB, PLA and PLA/PHB blends to create a new type of eco-friendly blend material suitable for single-use applications, such as fast-food packaging. 2. Experimental 2.1. Materials The PLA (4032D, NatureWorks Co. LLC, USA) had an average molecular weight of 52,000 g mol1, a polydispersity (Mw/Mn) of 1.9 and a high optical purity with 98% L-lactide content. The poly(3hydroxybutyrate) (PHB) was Biocyle (average molecular weight (Mw) of 425 kDa and polydispersity of 2.51 supplied in kind by PHB Industrial, SA, Serrana, SP, Brazil. Lapol 108 (Mw) of 80 kDa and polydispersity of 26.6) was purchased from Lapol, LLC, Santa Barbara, CA, USA. It was prepared according to the procedure described previously [36].

2.2. Preparation of the blends The materials were pre-treated by drying at 85  C over drierite for 24 h to eliminate possible absorbed water on the surface of the particles. Blends were melt processed at 175  C for 5 min at 45 rpm using a Haake Rheomix 600 with roller type blades. About 5 g of the obtained blend was compression moulded on a laboratory Carver press to obtain lms of w300e500 m. Moulding conditions were set at 180  C for 3 min at 200 Bar. 2.3. Scanning electron microscopy (SEM) Blend materials were fractured in liquid nitrogen to observe the interior of the unstressed composite. A sample was dropped directly into liquid nitrogen and fractured with a pre-chilled razor blade held in a vice-grip. The fractured pieces were picked out of the liquid nitrogen using pre-chilled forceps and placed in a desiccator to thaw and reduce the condensation of water on the surface of the material. All specimens were coated with gold-palladium for 45 s in a Denton Desk II sputter coating unit (Denton Vacuum USA, Moorestown, NJ). Specimens were viewed in a Hitachi S4700 eld emission scanning electron microscope (Hitachi HTA, Japan) at 2 kV. 2.4. Tensile properties Youngs modulus, tensile strength, and elongation at break were determined according to ASTM method D638M under ambient conditions, using an Instron 5500R Universal Testing Machine (Instron Corp., Canton, MA). Prior to testing, samples were equilibrated at 50% relative humidity in a chamber containing saturated solutions of calcium nitrate. The testing conditions used were: cross head speed of 5 mm/min and load cell of 0.1 kN. Dumbbell samples (0.2e0.5 mm thick) were tested with a gauge length of 20 mm. The reported values were the average of at least 6 measurements. Data points were the mean value of each measurement, with error bars in each graph representing 95% condence intervals. 2.5. Thermal properties Thermogravimetric analysis (TGA) was performed under nitrogen atmosphere in a TA Instruments TGA 2950. The samples were heated from room temperature to 510  C at a heating rate of 10  C/min and a nitrogen gas ow rate of 60 mL/min. The derivative of TGA curves (DTG) was obtained using TA analysis software. Differential scanning calorimetry (DSC) was performed with a Perking Elmer Jade DSC 8000 instrument calibrated with indium (T, DH) and zinc (T). DSC samples of 10e15 g were weighed in a 40 l stainless steel pan and measured against an empty pan as reference. Measurements were carried out under 80 ml min1 nitrogen ow rate according to the following protocol: rst and second heating from 20 to 200  C.min1; rst cooling (quenching after rst heating) from 200 to 20  C at 100  C.min1 and 2 min of isotherm at the end. In particular, the rst heating scan was used to erase any prior thermal history of the sample and the second heating scan was used to evaluate melting temperature (Tm) and crystallinity (Xc). The degree of crystallinity Xc was calculated from DSC curves as follows (Eq. (1)):

Xc

DHm 100 1 fDHm

(1)

where f is the weight fraction of the dispersed phase in the blends, DHm is the melting enthalpy (J/g) that was calculated from the fusion peak in the DSC curve and DH* m is the heat of fusion for completely crystallized PLA (93.1 J/g) and for completely crystallized PHB (146 J/g) [37,38].

1824

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828

2.6. X-ray diffraction X-ray diffraction (XRD) analyses were performed with a Philips 1820 diffractometer operated at 45 kV and, 40 mA using Cu-Ka radiation with a graphite diffracted beam monochromator. Data were acquired in a 2q scale from 2 to 40 . 2.7. Biodegradation test A Micro-Oxymax Respirometer System (Columbus Instruments Inc., Columbus, OH) was used to monitor biodegradation of the neat PHB, neat PLA, PLAePHB blend and PLAePHB-7%Lapol samples. Pregelatinized wheat starch was used as a positive control. Each sample (0.2 g) was mixed with 20 g of compost and kept in a 250mL sample chamber connected to a fully computerized, closedcircuit system. Carbon dioxide evolution from each sample was measured every 6 h for a total of 56 days at room temperature. 3. Results and discussion 3.1. Morphological properties The morphologies of the fracture surfaces were investigated by SEM, as shown in Fig. 1. Neat PLA presents a smooth and uniform surface of an amorphous polymer containing some holes [Fig. 1(a)],

whereas neat PHB shows an irregular fracture surface due to its crystalline structure [Fig.1(b)]. PLA and PHB with 7 wt-% Lapol did not show phase separation [Fig.1(c & d)]. Coarseness was observed on the fracture surface of PLA75/PHB25, and the fractures in the blend mainly developed along the same direction, indicating the resistance to the development of fractures was quite low [Fig. 1(e)]. On the other hand, the surfaces of specimens with Lapol were rougher than those without Lapol, and fractures developed in different directions, as shown in Fig. 1(f & g). Shear-yield and plastic deformation formed on the fracture surfaces of blends with Lapol, especially for the blend with high plasticizer content [(Fig. 1(g)]. Plastic deformation and the different fracture directions required more energy and thus the materials with Lapol should have better toughness. 3.2. XRD analysis X-ray diffraction analysis was used to determine crystalline structure and degree of crystallinity in the blends. PHB is a highly ordered polymer and is known to crystallize in an orthorhombic cell [39,40]. The lattice parameters of PHB are: a 0.572 nm and b 1.312 nm with its chain conformation in the left-handed 21 helix. In the orthorhombic cell, the three axis are mutually perpendicular (b g 90 ) and of unequal length (a s b s c) [37]. PLA can be crystallized from the melt as an orthorhombic unit cell with lattice parameters of a 1.037 nm and b 0.598 nm2.

Fig. 1. SEM images of PLA (a), PHB (b), PLA-7%Lapol (c), PHB-7%Lapol (d), PLA75-PHB25 (e), PLA75-PHB25-Lapol5 (f) and PLA75-PHB25-Lapol7 (g).

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828

1825

Neat PLA is usually formed from the -type helix 103. PLA can form two types of crystals, depending on the crystallization conditions. These are the form with a 103 helical shape and the b form with a spiral shape 31 [2,22,41e43]. PHB has reection peaks at 13.5 and 16.8 , corresponding to the (020) and (110) planes, and three weak peaks at 19.1, 22.2 and 25.5 , with the rst two peaks typical of the form orthorhombic structure [2]. The PLA had only one broad diffraction peak at 2 of 15 corresponding to the (110/200) reections (d 0.536) and smaller peaks at 19.1 and 32 corresponding to (010) and (203), respectively. In general, the patterns of PLA and PHB with Lapol blends are very similar to that of neat PLA and neat PHB (Fig. 2). It can be observed that diffraction at 2 13.5 and 16.9 for PHB-7%Lapol had increased in intensity. For the PLA75/PHB25 blend, the peak located at 16.9 is too weak to be observed in the pattern. Furthermore, the presence of amorphous PLA signicantly reduces the crystallinity of the PLA/PHB samples. This is due to the reduction in the PHB content and the PLA interfering with crystallization of PHB. For the PLA75/PHB25 blend, as shown in Fig. 2, the intensity ratio I(020)/I(110) is different from that of neat PHB. This indicates that the crystal structure of PHB in the PLA75/PHB25 blend has been altered from interactions between PLA and PHB. Since the crystallinity and the crystal growth rate of PHB are higher and faster than those of PLA, the reection peaks of PHB are observed even with higher PLA content. PLA75/PHB25/7Lapol did not have a peak (110) at 16.9 and had a weak peak (020) at 13.5 . This implies that the addition of Lapol signicantly improves the crystallinity and crystallization rate of the PLA75/PHB25 blend. 3.3. Thermal and crystallization behaviour The thermal properties of the blends could be signicantly affected by the crystallization characteristics of PHB and PLA. DSC traces recorded during the second heating cycle of the polymer samples are presented in Fig. 3. The measurements were performed immediately after the melt-quenching scans, thus the samples had the same thermal history and could be compared directly. The curves revealed similar thermal events for all samples, as follows: the glasserubber transition (characterized by Tg), cold crystallization process (characterized by Tcc and the cold crystallization enthalpy (DHcc)), and the melting process (characterized by Tm and the melting enthalpy (DHm)). The neat PLA sample showed a Tg of 63  C, an exothermic crystallization peak at 113  C and an endothermic melting peak at 173  C. Similarly, the neat PHB sample showed a Tg of 5.2  C, an exothermic crystallization peak at 48  C

Fig. 3. DSC traces of the 2nd heating scan of PLA, PHB, Lapol and PLAePHB-Lapol blends.

and a melting temperature of 178.5  C. These values were similar to those reported in literature [35]. As summarized in Table 1, Lapol had only one Tg at 0.7  C. In general, the DSC curves of PLA or PHB with plasticizer (Lapol) blends exhibit one Tg value, which indicates that both blend systems are miscible systems [44e47]. The dependence of Tg, blend (glass transition temperature of the blend) on the composition of the blends follows the Fox relationship [44] (Eq. (2)):

1 W W 1 2 Tg;blend Tg1 Tg2

(2)

where W1 is the mass fraction of component 1, W2 is the mass fraction of component 2, Tg1 is the glass transition temperature of component 1 and Tg2 is the glass transition temperature of component 2. Using Tg values of 62.9  C, 5.2  C and 0.7  C for neat PLA, PHB and Lapol, respectively, we calculated theoretical Tg, blend values of 59  C for PLA with 5 wt-% Lapol, 57.5  C for PLA with 7 wt% Lapol, 4.9  C for PHB with 5 wt-% Lapol and 4.78  C for PHB with 7 wt-% Lapol. These values are in good agreement with the experimental values. On the other hand, the DSC curves of PHB75/PLA25 blends exhibit two Tg values, which indicates that these blends are immiscible systems. However, addition of Lapol to the PLA75PHB25 sample resulted in disappearance of the Tg of Lapol and a decrease in Tg values of PHB and PLA. These behaviours suggest that Lapol has a signicant effect on the PLAePHB macromolecules.

Table 1 DSC parameters (2nd heating) of PLAePHB/Lapol Blends.a Code PLA Lapol PLA-5%Lapol PLA-7%Lapol PHB PHB-5%Lapol PHB-7%Lapol PLA75-PHB25 PLA75-PHB25-Lapol 5 PLA75-PHB25-Lapol 7
a

Tg1 ( C) e 0.7 e e 5.2 5.8 4.6 1.7 2.7 0.6

Tg2 ( C) 63 e 60 58 e e e 62 59 58

Tcc ( C) 115 e 110 110 48 49 50 115 125 120

Tm ( C) 173 e 173 172 179 177 178 173 172 173

DHcc
J/g 30 e 29 29 22 25 25 16 24 24

DHm
J/g 33 e 31 31 72 69 63 13 26 24

Xc (%) 35 e 35 37 49 50 48 36 40 38

Fig. 2. XRD of PLA, PHB and PLAePHB-Lapol blends.

DHm is the enthalpy of melting, DHcc is the cold crystallization enthalpy and Xc is the
degree of crystallinity.

Tg, Tm are the glass crystallization and melting peak temperatures, respectively,

1826

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828

The cold crystallization temperature (Tcc) of PLA decreased and that of PHB increased with an increase in plasticizer content. This corresponded to a decrease of DHcc in PLA blends and an increase of DHcc in PHB blends. However, PLA75/PHB25 had higher cold crystallization temperature and its peak became broader with an increase in Lapol concentration. This is possibly due to the small nely dispersed PHB crystals acting as nucleating agents in PLA. The presence of one crystallization temperature suggested that the PLA/ PHB blend has a signicant single homogeneous phase through the heating process. These results agree with previous studies by Hu et al. [20], Zhang et al. [48] and El-Hadi [25]. From Table 1, the shifts in the cold crystallization peaks (by approximately 11  C and 7  C for 5 wt-% and 7 wt-% of Lapol, respectively) is related to the difference in exibility of the chains and their ability to form a crystalline structure. Melting temperature (Tm) of the PLA and PHB blends mostly did not change with increasing amount of plasticizer. It is interesting to note that the PLA/PHB blend exhibits two melting peaks at 169  C and 173  C (Table 1 shows only one melting peak for the blends), which are attributed to recrystallization or two different lamella thickening. These results are in agreement with the results of Gunaratne et al. [49]. However, the presence of double melting temperatures in the PLAePHB-Lapol blends corresponds to the melting of as formed and recrystallized PHB crystallites. Ikehara et al. [50] stated that when miscible crystalline/crystalline blends are crystallized from a homogeneous melt, the crystallization behaviour largely depends on the difference in the melting point of the two components. When the Tm difference is large, the component with higher Tm crystallizes rst, and its spherulites usually ll the whole volume. The lower Tm component crystallizes at a lower temperature in spatially limited regions inside the spherulites of the other component. When the difference in Tm is small enough, both components have a high possibility to crystallize simultaneously. The Tm difference in most of the miscible crystalline/crystalline blends reported so far was approximately 100  C. In these cases, the two components crystallize at different temperatures. In this study, the Tm difference between PHB and PLA is about 5.8  C, indicating that simultaneous crystallization has occurred. The degree of crystallinity (Xc) of PLA and PHB based blend was basically uniform. A higher value of Xc was observed for PLA with 7 wt-%Lapol and for PHB with 5 wt-%Lapol. According to Azuma et al. [51] and Janigova et al. [52], an increase in crystallinity is due to higher chain mobility and a better packing of segments. The incorporation of Lapol in PLA75-PHB25 slightly increased Xc by 4% for the 5 wt-% Lapol sample and by 2% for the 7 wt-% Lapol sample, as shown in Table 1. This behaviour could be dependent on the chain mobility of the polymer. Also, the interface interactions between PHB and PLA with Lapol could inuence nucleation as supported by the XRD results presented in Fig. 2. In addition, the decrease in DHm values of all blends might imply formation of a more stable structure, perhaps due to increased movement of the polymer segments that facilitate polymer chain mobility. 3.4. Thermogravimetric analysis (TGA) Thermal degradation of plasticized PLA and PHB samples is shown in Fig. 4(a & b). In general, the thermal weight-losses of PHB and PLA were quite sensitive to temperature, with narrow decomposition temperature ranges. The temperature corresponding to the onset of decomposition (Tonset) for a polymer is essential for evaluating its thermal stability and is summarized in Table 2. PLA is more thermally stable than PHB. Lapol decreased the Tonset of PLA by 11  C for the 5 wt-% Lapol sample and by 12  C for the 7 wt-% Lapol sample. In contrast, Lapol increased the Tonset of PHB by 4  C

Fig. 4. a) TGA of PLA, PHB, PLAePHB based blends and b) DTG of PLA, PHB and PLAePHB based blends.

for the 5 and 7 wt-% Lapol samples. PHB and PLA thermally degrade to produce polymeric chains terminated with carboxyl and vinyl groups [53]. Carboxyl end groups of polyester catalyse hydrolysis reaction. In the present study, it is feasible that a synergistic effect occurred to accelerate thermal degradation of PLA by moisture and polyol [54,55]. The Tonset of PLA75/PHB25 blend was between those of PLA and PHB. The addition of Lapol led to a slight shift in Tonset of PLA/PHB blend to a higher temperature, as shown in Table 2. This phenomenon was attributed to the higher thermal stability of Lapol. Another important thermal property is the temperature corresponding to the maximum rate of weight loss (Tp), which is dened
Table 2 TGA data of PLA/PHB/Lapol blends.a Code PLA Lapol PLA-5%Lapol PLA-7%Lapol PHB PHB-5%Lapol PHB-7%Lapol PLA75-PHB25 PLA75-PHB25-Lapol 5 PLA75-PHB25-Lapol 7 Tonset ( C) 324 278 313 312 260 264 264 270 273 271 Tp PHB ( C) e e e e 279 281 282 283 284 283 Tp PLA ( C) 357 340 356 357 e e e 356 357 341 R500 (%) 0.12 8.94 0.16 0.86 0.25 0.54 0.53 0.10 0.86 0.93

a Td is the onset temperature; Tp is the rst derivative peak; and R500 is the residual weight at 500  C.

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828

1827

as the peak value of the rst derivative of the TGA curve. The rst and second derivatives curves for neat PLA, neat PHB and PLAePHB based blends are shown in Fig. 4(b) and their Tp values are listed in Table 2. All neat polymers exhibited a single peak, indicating that the PLA and PHB degraded in only one step. In addition, the Tp of PLA and PHB with Lapol shifted slightly to higher temperature. However, the Tp value of PLA75/PHB25 showed a decrease with increasing amounts of Lapol. All samples had less than 1 wt-% residue at 500  C and this value increased with increasing plasticizer content. 3.5. Mechanical properties Youngs modulus, strain at break and stress at break values of melt-mixed PLA, PHB and PLAePHB based blends are presented in Table 3. The mechanical behaviour of the neat PLA showed characteristics of glassy polymers with low deformation at break. In contrast, the neat PHB has a low melt viscosity and is a brittle material with a higher modulus. The blends of PLA and PHB with Lapol showed different behaviour. PLA with 5 wt-% Lapol had a higher Youngs modulus (YM) than that for neat PLA. However, PHB with 5 wt-%Lapol had a lower YM than that of neat PHB. The YM of the PLA75/PHB25 sample was lower than that of neat PLA and addition of plasticizer resulted in even lower values. The tensile strength was signicantly reduced by addition of Lapol, but the elongation at break increased in value. Tensile strength of PLA75/ PHB25 was lower than those of PLA and PHB, with no change in the elongation at break (w7%). The addition of Lapol markedly changed the stress-strain behaviour of the PLA75/PHB25 blend. The PLA75/ PHB25 blend with 5 wt-%Lapol had a higher elongation at break value than neat PLA and neat PHB. However, its tensile strength was lower than neat PLA and neat PHB. The PLA/PHB/Lapol samples showed some characteristics of exible materials. These improvements might be due to the nely dispersed Lapol acting as llers, which enhanced the properties of neat polymer. They can also be due to possible interactions between PHB and PLA, with Lapol causing strong interfacial bonding. 3.6. Biodegradation results The data herewith reported are referred to a preliminary study relevant to the biodegradation of PHB, PLA and two blends of PLA/ PHB. The neat PHB sample showed some biodegradation during the measurement time period, whereas the other polymer samples did not degrade much. This is shown in Fig. 5, which displays carbon dioxide evolution of the samples as a function of time. The starch sample seemed to degrade the fastest since it generated the largest amount of carbon dioxide. Meanwhile, the neat PHB sample exhibited an increase in carbon dioxide content up to approximately 40 days, but this content levelled off thereafter. In
Table 3 Tensile properties of PLA/PHB/Lapol blend lms.a Code PLA PLA-5%Lapol PLA-7%Lapol PHB PHB-5%Lapol PHB-7%Lapol PLA75-PHB25 PLA75-PHB25-5Lapol PLA75-PHB25-7Lapol
a

Fig. 5. Carbon dioxide evolution of the pregelatinized wheat starch, neat PHB, neat PLA, PLAePHB blend and PLAePHB-7%Lapol samples as a function of time.

comparison, the samples containing PLA did not generate much carbon dioxide, indicating these samples did not degrade much under the experimental conditions. This last aspect is worthy of further investigation in a more comprehensive and detailed approach. 4. Conclusions Biodegradable PLAePHB blends have been successfully prepared by melt mixing PLA, PHB and a new type of plasticizer (Lapol 108). This compounding technique is easy and potentially commercially viable. SEM showed that PLA and PHB with plasticizer exhibited a homogeneous distribution of Lapol 108 in the polymer matrix. Samples of ternary blends also showed homogenous distribution. XRD data showed that Lapol 108 was affecting the interlayer space of the blends. Moreover, Lapol 108 essentially decreased the Tg for all of the binary and ternary blends. Also, crystallinity in the PLA75-PHB25 sample slightly increased with increased loading of Lapol 108. On the other hand, thermal decomposition properties of the blends were not affected by Lapol 108. The addition of Lapol caused the elongation at break to increase and the Youngs modulus to somewhat decrease compared with neat PLA and neat PHB. Also, the neat PHB sample showed some biodegradation, whereas the samples containing PLA did not degrade much during the experimental time period. Acknowledgement The work reported here is part of the activities performed by M.A. Abdelwahab for the achievement of the PhD title in Biomaterials as granted by the University of Pisa (Italy). References
[1] Vasile C. Environmentally degradable polymeric materials: denition and background. In: Vasile C, Zaikov G, editors. Environmentally degradable materials based on multicomponent polymeric systems. Netherlands: Koninklijke Brill NV, Leiden; 2009. [2] Kobayashi J, Asahi T, Ichiki M, Oikawa A, Suzuki H, Watanabe T, et al. Structural and optical properties of poly lactic acids. Journal of Applied Physics 1995;77:2957e73.

Youngs modulus (MPa) 1400 1450 1200 1950 1750 1830 1270 1150 1120 130 180 60 140 110 140 110 40 60

Tensile strength (MPa) 42 14 16 31 29 26 16 13 15 18 2 1 4 2 2 3 2 1

Elongation at break (%) 7.2 14.4 13.7 7.3 7.2 5.6 7.1 15.5 15.1 2 1 3 1 1 1 1 2 3

Errors were calculated at 95% of condence of students t-test.

1828

M.A. Abdelwahab et al. / Polymer Degradation and Stability 97 (2012) 1822e1828 [29] Kumar M, Mohanty S, Nayak SK, Rahail Parvaiz M. Effect of glycidyl methacrylate (GMA) on the thermal, mechanical and morphological property of biodegradable PLA/PBAT blend and its nanocomposites. Bioresource Technology 2010;101:8406e15. [30] Labrecque LV, Kumar RA, Dav V, Gross RA, McCarthy SP. Citrate esters as plasticizers for poly(lactic acid). Journal of Applied Polymer Science 1997;66: 1507e13. [31] Jacobsen S, Fritz HG. Plasticizing polylactide-the effect of different plasticizers on the mechanical properties. Polymer Engineering & Science 1999;39: 1303e10. [32] Baltieri RC, Innocentini Mei LH, Bartoli J. Study of the inuence of plasticizers on the thermal and mechanical properties of poly(3-hydroxybutyrate) compounds. Macromolecular Symposia 2003;197:33e44. [33] Yoshie N, Nakasato K, Fujiwara M, Kasuya K, Abe H, Doi Y, et al. Effect of low molecular weight additives on enzymatic degradation of poly(3hydroxybutyrate). Polymer 2000;41:3227e34. [34] Pizzoli M, Scandola M, Ceccorulli G. Crystallization and melting of isotactic poly(3-hydroxy butyrate) in the presence of a low molecular weight diluent. Macromolecules 2002;35:3937e41. [35] Grillo Fernandes E, Pietrini M, Chiellini E. Thermo-mechanical and morphological characterization of plasticized poly[(R)-3-hydroxybutyric acid]. Macromolecular Symposia 2004;218:157e64. [36] Flynn A, Lennard T. Bioderived plasticizer for biopolymers. In: Patent U, editor. USPTO. USA: Lapol, LLC; 2010. [37] Barham PJ, Keller A, Otun EL, Holmes PA. Crystallization and morphology of a bacterial thermoplastic-poly-3-hydroxybutyrate. Journal of Materials Science 1984;19:2781e94. [38] Fischer EW, Sterzel HJ, Wegner G. Investigation of the structure of solution grown crystals of lactide copolymers by means of chemical reactions. Colloid & Polymer Science 1973;251:980e90. [39] Hurrell BL, Cameron RE. A wide-angle X-ray scattering study of the ageing of poly(hydroxybutyrate). Journal of Materials Science 1998;33:1709e13. [40] Bruckner S, Meille SV, Malpezzi L, Cesaro A, Navarini L, Tombolini R. The structure of poly(D-beta-hydroxybutyrate). A renement based on the rietveld method. Macromolecules 1988;21:967e72. [41] Eling B, Gogolewski S, Pennings AJ. Biodegradable materials of poly(l-lactic acid): 1. Melt-spun and solution-spun bres. Polymer 1982;23:1587e93. [42] Cho J, Baratian S, Kim J, Yeh F, Hsiao BS, Runt J. Crystallization and structure formation of poly(l-lactide-co-meso-lactide) random copolymers: a timeresolved wide- and small-angle X-ray scattering study. Polymer 2003;44: 711e7. [43] Cartier L, Okihara T, Lotz B. Triangular polymer single crystals: stereocomplexes, twins, and frustrated structures. Macromolecules 1997;30: 6313e22. [44] Hale A, Bair H. Polymer blends and block copolymers. In: Thermal characterization of polymeric materials. 2nd ed. San Diego: ED. Academic Press; 1997. [45] Anand J, Palaniappan S, Sathyanarayana DN. Spectral, thermal, and electrical properties of poly(o- and m-toluidine)-polystyrene blends prepared by emulsion pathway. Journal of Polymer Science Part A-Polymer Chemistry 1998;36:2291e9. [46] Archondouli PS, Kallitsis JK, Kalfoglou NK. Compatibilization and property characterization of polycarbonate/polyurethane polymeric alloys. Journal of Applied Polymer Science 2003;88:612e26. [47] Ferreira BMP, Zavaglia CAC, Duek EAR. Films of PLLA/PHBV: thermal, morphological, and mechanical characterization. Journal of Applied Polymer Science 2002;86:2898e906. [48] Zhang LL, Xiong CD, Deng XM. Miscibility, crystallization and morphology of poly(beta-hydroxybutyrate)/poly(d, l-lactide) blends. Polymer 1996;37: 235e41. [49] Gunaratne LMWK, Shanks RA. Miscibility, melting, and crystallization behaviour of poly(hydroxybutyrate) and poly(D, L-lactic acid) blends. Polymer Engineering & Science 2008;48:1683e92. [50] Ikehara T, Kimura H, Qiu Z. Penetrating spherulitic growth in poly(butylene adipate-co-butylene succinate)/poly(ethylene oxide) blends. Macromolecules 2005;38:5104e8. [51] Azuma Y, Yoshie N, Sakurai M, Inoue Y, Chujo R. Thermal-behaviour and miscibility of poly(3-hydroxybutyrate)/poly(vinyl alcohol) blends. Polymer 1992;33:4763e7. [52] Janigova I, Lacik I, Chodak I. Thermal degradation of plasticized poly(3hydroxybutyrate) investigated by DSC. Polymer Degradation and Stability 2002;77:35e41. [53] Kopinke F, Remmler M, Mackenzie K. Thermal decomposition of biodegradable polyester -I: poly(b-hyroxybutyric acid). Polymer Degradation and Stability 1996;52:25e38. [54] Zhang H, Ward IM. Kinetics of hydrolytic degradation of poly(ethylene naphthalene-2,6-dicarboxylate). Macromolecules 1995;28:7622e9. [55] Chiellini E, Fernandes E, Pietrini M, Solaro R. Factorial design in optimization of PHAs processing. Macromolecular Symosium 2003;197:45e55.

[3] Hoogsteen W, Postema AR, Pennings AJ, Ten Brinke G, Zugenmaier P. Crystal structure, conformation and morphology of solution-spun poly(L-lactide) bers. Macromolecules 1990;23:634e42. [4] Sinclair RG. The case for polylactic acid as a commodity packaging plastic. Journal of Macromolecular Science, Part A 1996;33:585e97. [5] James L. Large-scale production, properties and commercial applications of polylactic acid polymers. Polymer Degradation and Stability 1998;59:145e52. [6] Cartier L, Okihara T, Ikada Y, Tsuji H, Puiggali J, Lotz B. Epitaxial crystallization and crystalline polymorphism of polylactides. Polymer 2000;41:8909e19. [7] Martin O, Avrous L. Poly(lactic acid): plasticization and properties of biodegradable multiphase systems. Polymer 2001;42:6209e19. [8] Khanna S, Srivastava AK. Recent advances in microbial polyhydroxyalkanoates. Process Biochemistry 2005;40:607e19. [9] Kang IK, Choi SH, Shin DS, Yoon SC. Surface modication of polyhydroxyalkanoate lms and their interaction with human broblasts. International Journal of Biological Macromolecules 2001;28:205e12. [10] Scandola M, Focarete ML, Adamus Gy, Sikorska W, Baranowska I, Swierczek S, et al. Polymer blends of natural poly(3-hydroxybutyrate-co-3hydroxyvalerate) and a synthetic atactic poly(3-hydroxybutyrate). Characterization and biodegradation studies. Macromolecules 1997;30:2568e74. [11] Iwata T, Aoyagi Y, Fujita M, Yamane H, Doi Y, Suzuki Y, et al. Processing of a strong biodegradable poly[(R)-3-hydroxybutyrate] ber and a new ber structure revealed by micro-beam X-ray diffraction with synchrotron radiation. Macromolecular Rapid Communications 2004;25:1100e4. [12] Zhang J, Duan Y, Sato H, Tsuji H, Noda I, Yan S, et al. Crystal modications and thermal behaviour of poly(l-lactic acid) revealed by infrared spectroscopy. Macromolecules 2005;38:8012e21. [13] Park JW, Doi Y, Iwata T. Uniaxial drawing and mechanical properties of poly [(R)-3-hydroxybutyrate]/poly(l-lactic acid) blends. Biomacromolecules 2004; 5:1557e66. [14] Gassner F, Owen AJ. Physical-properties of poly(beta-hydroxybutyrate) poly(epsilon-caprolactone) blends. Polymer 1994;35:2233e6. [15] Nurkhamidah S, Woo EM. Cracks and ring bands of poly(3-hydroxybutyrate) on precrystallized poly(l-lactic acid) template. Industrial & Engineering Chemistry Research 2011;50:4494e503. [16] Focarete ML, Ceccorulli G, Scandola M, Kowalczuk M. Further evidence of crystallinity-induced biodegradation of synthetic atactic poly(3hydroxybutyrate) by PHB-depolymerase a from pseudomonas lemoignei. blends of atactic poly(3-hydroxybutyrate) with crystalline polyesters. Macromolecules 1998;31:8485e92. [17] Koyama N, Doi Y. Miscibility of binary blends of poly[(R)-3-hydroxybutyric acid] and poly[(S)-lactic acid]. Polymer 1997;38:1589e93. [18] Furukawa T, Sato H, Murakami R, Zhang J, Duan Y-X, Noda I, et al. Structure, dispersibility, and crystallinity of poly(hydroxybutyrate)/poly(l-lactic acid) blends studied by FT-IR microspectroscopy and differential scanning calorimetry. Macromolecules 2005;38:6445e54. [19] Blumm E, Owen AJ. Miscibility, crystallization and melting of poly(3hydroxybutyrate)/poly(L-lactide) blends. Polymer 1995;36:4077e81. [20] Hu Y, Sato H, Zhang J, Noda I, Ozaki Y. Crystallization behaviour of poly(l-lactic acid) affected by the addition of a small amount of poly(3-hydroxybutyrate). Polymer 2008;49:4204e10. [21] Zhang J, Sato H, Furukawa T, Tsuji H, Noda I, Ozaki Y. Crystallization behaviours of poly(3-hydroxybutyrate) and poly(l-lactic acid) in their immiscible and miscible blends. The Journal of Physical Chemistry B 2006;110:24463e71. [22] Furukawa T, Sato H, Murakami R, Zhang J, Noda I, Ochiai S, et al. Comparison of miscibility and structure of poly(3-hydroxybutyrate-co-3hydroxyhexanoate)/poly(l-lactic acid) blends with those of poly(3hydroxybutyrate)/poly(l-lactic acid) blends studied by wide angle X-ray diffraction, differential scanning calorimetry, and FTIR microspectroscopy. Polymer 2007;48:1749e55. [23] Noda I, Satkowski MM, Dowrey AE, Marcott C. Polymer alloys of nodax copolymers and poly(lactic Acid). Macromolecular Bioscience 2004;4:269e75. [24] Furukawa T, Sato H, Murakami R, Zhang J, Noda I, Ochiai S, et al. Raman microspectroscopy study of structure, dispersibility, and crystallinity of poly(hydroxybutyrate)/poly(l-lactic acid) blends. Polymer 2006;47:3132e40. [25] El-Hadi AM. Effect of processing conditions on the development of morphological features of banded or nonbanded spherulites of poly(3hydroxybutyrate) (PHB) and polylactic acid (PLLA) blends. Polymer Engineering & Science 2011;51:2191e202. [26] Wang S, Ma P, Wang R, Zhang Y. Mechanical, thermal and degradation properties of poly(D, L-lactide)/poly(hydroxybutyrate-co-hydroxyvalerate)/ poly(ethylene glycol) blend. Polymer Degradation and Stability 2008;93: 1364e9. [27] Zhang M, Thomas NL. Blending polylactic acid with polyhydroxybutyrate: the effect on thermal, mechanical, and biodegradation properties. Advances in Polymer Technology 2011;30:67e79. [28] Ohkoshi I, Abe H, Doi Y. Miscibility and solid-state structures for blends of poly[(S)-lactide] with atactic poly[(R, S)-3-hydroxybutyrate]. Polymer 2000; 41:5985e92.

Você também pode gostar