首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Poly(L ‐lactic acid) (PLLA) and poly(D ‐lactic acid) (PDLA) with very different weight‐average molecular weights (Mw) of 4.0 × 103 and 7.0 × 105 g mol?1 (Mw(PDLA)/Mw(PLLA) = 175) were blended at different PDLA weight ratios (XD = PDLA weight/blend weight) and their crystallization from the melt was investigated. The presence of low molecular weight PLLA facilitated the stereocomplexation and thereby lowered the cold crystallization temperature (Tcc) for non‐isothermal crystallization during heating and elevated the radial growth rate of spherulites (G) for isothermal crystallization, irrespective of XD. The orientation of lamellae in the spherulites was higher for the neat PLLA, PDLA and an equimolar blend than for the non‐equimolar blends. It was found that the orientation of lamellae in the blends was maintained by the stereocomplex (SC) crystallites. Although the G values are expected to decrease with an increase in XD or the content of high‐molecular‐weight PDLA with lower chain mobility compared with that of low‐molecular‐weight PLLA, G was highest at XD = 0.5 where the maximum amount of SC crystallites was formed and the G values were very similar for XD = 0.4 and XD = 0.6 with the same enantiomeric excess. This means that the effect of SC crystallites overwhelmed that of chain mobility. The nucleating mechanisms of SC crystallites were identical for XD = 0.1–0.5 in the Tc range 130–180 °C. Copyright © 2011 Society of Chemical Industry  相似文献   

2.
The rheological behavior of blends of poly(styrene-co-acrylonitrile) (SAN) and poly(ε-caprolactone) (PCL) was investigated, using a cone-and-plate rheometer. For the study, blends of various compositions were prepared by melt blending using a twin-screw compounding machine. The rheological properties measured were shear stress (σ12), viscosity (η), and first normal stress difference (N1) as functions of shear rate (γ) in steady shearing flow, and dynamic storage modulus (G′) and loss modulus (G″) as functions of angular frequency (ω) in oscillatory shearing flow, at various temperatures. It has been found that logarithmic plots of N1 versus σ12, and logarithmic plots of G′ versus G″, become virtually independent of temperature but vary regularly with blend composition, and that the zero-shear viscosity of the blends, (ηo)blend, follows the relationship, 1/log(ηo)blend = wA/log η0A + wB/log η0B, where η0A and η0B are the zero-shear viscosities of components A and B, respectively, and wA and wB are the weight fractions of components A and B, respectively. The physical implications of the relationship found are discussed.  相似文献   

3.
Stereo diblock polylactides (SDB‐PLAs) composed of relatively short poly(d ‐lactide) (PDLA) segments and relatively long poly(l ‐lactide) (PLLA) segments were synthesized to have a wide number‐average molecular weight (Mn) range of 2.5 × 104–2.0 × 105 g mol?1 and d ‐lactyl unit content of 0.9–38.6%. The effects of incorporated short PDLA segments (Mn = 2.0 × 103–7.7 × 103 g mol?1) on crystallization behavior of the SDB‐PLAs were first investigated during heating after complete melting and quenching or during slow cooling after complete melting. Stereocomplex (SC) crystallites can be formed at d ‐lactyl unit content as low as 4.3 and 5.8% for heating and slow cooling, respectively, and for Mn of PDLA segments as low as 2.0 × 103 and 3.5 × 103 g mol?1, respectively. With decreasing Mn and increasing d ‐lactyl unit content, the cold crystallization temperature during heating decreased and the crystallization temperature during slow cooling increased. With increasing d ‐lactyl unit content, the melting enthalpy (ΔHm) of SC crystallites during heating and the crystallinity (Xc) of SC crystallites after slow cooling increased, whereas ΔHm of PLLA homo‐crystallites during heating and Xc of PLLA homo‐crystallites after slow cooling decreased. The total ΔHm of SC crystallites and PLLA homo‐crystallites during heating and the total Xc after slow cooling became a minimum at d ‐lactyl unit content of 10–15% and gave a maximum at d ‐lactyl unit content of 0%. Despite the accelerated crystallization of some of SDB‐PLAs, the low values of total ΔHm and Xc at d ‐lactyl unit content of 10–15% are attributable to the formation of two crystalline species of SC crystallites and PLLA homo‐crystallites.  相似文献   

4.
Amorphous‐made poly(L ‐lactide) [i.e., poly(L ‐lactic acid) (PLLA)], poly(L ‐lactide‐co‐D ‐lactide)[P(LLA‐DLA)](77/23), and P(LLA‐DLA)(50/50) films and PLLA films with different crystallinity (Xc) values were prepared, and the effects of molecular weight, D ‐lactide unit content (tacticity and optical purity), and crystallinity of poly(lactide) [i.e., poly(lactic acid) (PLA)] on the water vapor permeability was investigated. The changes in number‐average molecular weight (Mn) of PLLA films in the range of 9 × 104–5 × 105 g mol?1 and D ‐lactide unit content of PLA films in the range of 0–50% have insignificant effects on their water vapor transmission rate (WVTR). In contrast, the WVTR of PLLA films decreased monotonically with increasing Xc from 0 to 20%, while leveled off for Xc exceeding 30%. This is probably due to the higher resistance of “restricted” amorphous regions to water vapor permeation compared with that of the “free” amorphous regions. The free and restricted amorphous regions are major amorphous components of PLLA films for Xc ranges of 0–20% and exceeding 30%, respectively, resulting in the aforementioned dependence of WVTR on Xc. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  相似文献   

5.
Blend films were prepared from poly(L -lactide) (PLLA) and poly(E-caprolactone) (PCL) with different PLLA contents [XPLLA (w/w) = PLLA/(PCL + PLLA)] by the solution-casting method and their hydrolysis behaviors were investigated up to 20 months in a phosphate-buffered solution of pH 7.4 at 37°C by gel permeation chromatography, tensile testing, differential scanning calorimetry, and gravimetry. Polarizing microscopic observation and dynamic mechanical analysis revealed that PCL and PLLA were phase-separated in blend films before hydrolysis. The mass remaining, molecular weight, and tensile strength of the blend films with XPLLA of 0.5 and 0.75 decreased more rapidly by hydrolysis than those of the nonblended PLLA, while the elongation at break of the blend film of XPLLA = 0.25 decreased the slowest. The rate constant for hydrolysis (k) calculated from the Mn change during hydrolysis was higher for blend films of XPLLA = 0.5 and 0.75 than those expected from k of nonblended PLLA and PCL. The melting temperature (Tm) of PLLA in the blend and nonblended films of XPLLA = 0.5, 0.75, and 1 decreased from 179 to 161, 160, and 175°C upon hydrolysis for 20 months, respectively, while that for XPLLA = 0.25 slightly increased from 176 to 177°C. On the other hand, Tm and the crystallinity of PCL was significantly increased by hydrolysis for 20 months, irrespective of XPLLA. © 1998 John Wiley & Sons, Inc. J Appl Polym Sci 67: 405–415, 1998  相似文献   

6.
Softness parameters σM for cations and σX for anions, have been calculated as dimensionless quantities for approx. 90 cations and 18 anions. They are given by σM = [σA (Mm+) - σA(H+)]/σA(H+) and σX = [σB(Xa?) - σB(OH?)]/σA(H+) where σA = [σIi(M) + ΔH0h(Mm+)]/m and σB = [-Ea(X) + ΔH0h(Xa?)]/a are Ahrland's parameters. The new normalized and comparative (to the test ions H+ and OH?) softness parameters are positive for soft ions and negative for hard ones. These parameters, obtained independently, are used with a four-coefficient equation to calculate coordinate bond energies for metal halides with acceptable accuracy. Considerations of the average coordination in reciprocal molten salt mixture lead to an expression for the metathesis energy change as proportional to the product of the differences in softness parameters of the two cations and the two anions. An empirical one-coefficient equation involving the softness parameters is proposed to deal with next-nearest-neighbor interactions in binary common-ion molten salt mixtures. These relationships are then used with Blander and Topol's equation to predict the occurrence of irascibility gaps in uni-univalent reciprocal salt mixtures. The gaps found in other systems are also discussed in terms of the softness of the constituent ions.  相似文献   

7.
Conformational characteristics of poly(lactide)s have been investigated by density functional theory and ab initio molecular orbital (MO) calculations and NMR experiments on model compounds. Characteristic ratios, configurational entropies, and internal energies of poly(L-lactide) and poly(DL-lactide), whose stereosequences were generated by Bernoulli and Markov stochastic processes, were calculated under the refined rotational isomeric state scheme with conformational energies and geometrical parameters derived from the MO calculations. In terms of the conformational characteristics thus revealed, we have elucidated the reason why unperturbed chain dimensions determined experimentally for poly(L-lactide) are scattered considerably and, furthermore, discussed crystallization and crystal structures of poly(L-lactide) and molecular characteristics of poly(DL-lactide) synthesized from rac-lactide with stereospecific polymerization catalysts.  相似文献   

8.
Effects of the mixing ratio of poly(DL-lactide) (PDLLA) and poly (ε-caprolactone) (PCL) on the thermal and mechanical properties and morphologies of the solution-cast blends were investigated by differential scanning calorimetry (DSC), polarizing microscopy, tensile tests, and dynamic mechanical analysis. The presence of amorphous PDLLA did not disturb crystallization of PCL over the PDLLA content [XPDLLA = PDLLA/(PCL + PDLLA)] from 0.1 to 0.9 and allowed PCL to form spherulites over XPDLLA ranging from 0.1 to 0.6. The spherulite radius was larger for the blends than for the nonblended PCL. Phase separation occurred for the blends with XPDLLA between 0.1 and 0.9 Tm of PCL remained unchanged in the XPDLLA range up to 0.6 but decreased at XPDLLA above 0.6, whereas the crystallinity of PCL was constant around 60%, irrespective of XPDLLA. The tensile strength (σB, the yield stress (σY), the Young's modulus(E), and the storage modulus (G′) of the blends increased monotonously with XPDLLA if σB at XPDLLA = 0 and 0.6 and σY at XPDLLA = 0.6 were excluded. Elongation-at-break (εB) of PDLLA increased dramatically, while εB of PCL decreased remarkably when a small amount of the other component was added. Equations and parameters predicting σY, E, and G′ of the PCL-PDLLA blends were proposed as a function of XPDLLA. © 1996 John Wiley & Sons, Inc.  相似文献   

9.
Phase-separated blend films were prepared with the solution casting method from poly(L -lactide) (PLLA) and poly(ε-caprolactone) (PCL) with different PLLA contents [XPLLA (w/w) = PLLA/(PCL + PLLA)] and their biodegradation was investigated in soil up to 20 months by gravimetry, gel permeation chromatography, tensile testing, differential scanning calorimetry, and scanning electron microscopy. The nonblended PCL film and the blend film with XPLLA = 0.25 disappeared in 4 and 12 months, respectively, while most of the initial mass remained for the blend film of XPLLA = 0.75 and the nonblended PLLA film. The decrease in weight remaining, molecular weight, tensile strength, and elongation-at-break was higher for blend films of low XPLLA. The melting temperature of PLLA in blend films of XPLLA = 0.5 and 0.75, and of nonblended film, remained around 179°C upon biodegradation in soil for 20 months. The preferred biodegradation of PCL in blend films resulted in formation of microspheres of a PLLA-rich phase at the surface for the blend film of XPLLA = 0.25 and the porous structure for blend films of XPLLA = 0.5 and 0.75. Comparison of the weight loss of blend films in biodegradation in soil with that of the nonenzymatic hydrolysis in phosphate-buffered solution revealed preferred enzymatic degradation of PCL and insignificant attack to PLLA in the blends. © 1998 John Wiley & Sons, Inc. J Appl Polym Sci 70: 2259–2268, 1998  相似文献   

10.
Blended films of poly(L ‐lactide) [ie poly(L ‐lactic acid)] (PLLA) and poly(?‐caprolactone) (PCL) without or mixed with 10 wt% poly(L ‐lactide‐co‐?‐caprolactone) (PLLA‐CL) were prepared by solution‐casting. The effects of PLLA‐CL on the morphology, phase structure, crystallization, and mechanical properties of films have been investigated using polarization optical microscopy, scanning electron microscopy, differential scanning calorimetry and tensile testing. Addition of PLLA‐CL decreased number densities of spherulites in PLLA and PCL films, and improved the observability of spherulites and the smoothness of cross‐section of the PLLA/PCL blend film. The melting temperatures (Tm) of PLLA and PCL in the films remained unchanged upon addition of PLLA‐CL, while the crystallinities of PLLA and PCL increased at PLLA contents [XPLLA = weight of PLLA/(weight of PLLA and PCL)] of 0.4–0.7 and at most of the XPLLA values, respectively. The addition of PLLA‐CL improved the tensile strength and the Young modulus of the films at XPLLA of 0.5–0.8 and of 0–0.1 and 0.5–0.8, respectively, and the elongation at break of the films at all the XPLLA values. These findings strongly suggest that PLLA‐CL was miscible with PLLA and PCL, and that the dissolved PLLA‐CL in PLLA‐rich and PCL‐rich phases increased the compatibility between these two phases. © 2003 Society of Chemical Industry  相似文献   

11.
Effect of Poly(l ‐lactide)/Poly(d ‐lactide) (PLLA/PDLA) block length ratio on the crystallization behavior of star‐shaped poly(propylene oxide) block poly(d ‐lactide) block poly (l ‐lactide) (PPO–PDLA–PLLA) stereoblock copolymers with molecular weights (Mn) ranging from 6.2 × 104 to 1.4 × 105 g mol?1 was investigated. Crystallization behaviors were studied utilizing differential scanning calorimetry (DSC), polarized optical microscopy (POM), and wide‐angle X‐ray diffraction (WAXD). Only stereocomplex crystallites formed in isothermal crystallization at 140 to 156°C for all samples. On one hand, the overall crystallization rate decreased as PLLA/PDLA block length ratio increased. As PLLA/PDLA block length ratio increased from 7:7 to 28:7, the value of half time of crystallization (t1/2) delayed form 2.85 to 5.31 min at 140°C. On the other hand, according to the Lauritzen–Hoffman theory, the fold‐surface energy (σe) was calculated. σe decreased from 77.7 to 73.3 erg/cm2 with an increase in PLLA/PDLA block length ratio. Correspondingly increase in nucleation density was observed by the polarized optical microscope. Results indicated that the PLLA/PDLA block length ratio had a significant impact on the crystallization behavior of PPO–PDLA–PLLA copolymers. POLYM. ENG. SCI., 55:2534–2541, 2015. © 2015 Society of Plastics Engineers  相似文献   

12.
Blend films were prepared from hydrophobic poly(L -lactide) (PLLA) and hydrophilic poly(vinyl alcohol) (PVA) with different PLLA contents [XPLLA (w/w) = PLLA/(PVA + PLLA)] by solution casting and melt quenching. Their morphology, swelling behavior, and surface and bulk properties were investigated. Polarizing optical microscopy, scanning electron microscopy, differential scanning calorimetry, X-ray diffractometry, and tensile testing revealed that PLLA and PVA were phase separated in these blend films and the PLLA-rich and PVA-rich phases both formed a continuous domain in the blend film of XPLLA = 0.5. The water absorption of the blend films was higher for the blend films of low XPLLA values when compared at the same immersion time, and it was larger than expected from those of nonblended PLLA and PVA films. The dynamic contact angles of the blend films were linearly increased with an increase in XPLLA. The tensile strength and Young's modulus of the dry blend films decreased with a rise in XPLLA, but this dependence was reversed because of the large decreases in tensile strength and Young's modulus for the blend films having high XPLLA values after immersion in water. The elongation at break was higher for the wet blend film than for the dry blend film when compared at the same XPLLA and that of the dry and wet blend films decreased with an increase in XPLLA. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 81: 2151–2160, 2001  相似文献   

13.
Poly(L -lactide) (PLLA) and poly(DL -lactide) (PDLLA) blends were crystallized from the melt and their crystallization behaviors and morphologies were investigated using differential scanning calorimetry and polarizing microscopy. PLLA could crystallize from the melt in the presence of PDLLA when the PLLA content in the blend (XPLLA) was higher than 0.2. Spherulites were formed when XPLLA was between 0.6 and 1. The radius of the spherulites formed in the presence of PDLLA was larger than that of nonblended PLLA, probably because the coexisting PDLLA reduced the density of nuclei for the spherulites. Small crystallite assemblies were formed when XPLLA was between 0.2 and 0.5, and grew spherically from the nuclei when XPLLA was 0.5. Isotactic PLLA and atactic PDLLA seem to be miscible prior to crystallization, but PDLLA will be trapped between the lamellae of the spherulites or spherulitic assemblies during crystallization. PDLLA delayed the induction for crystallization of the blends. In spite of different morphology, the melting temperature and crystallinity of PLLA remained virtually constant when annealing was conducted for 600 min at XPLLA between 0.5 and 1. © 1995 John Wiley & Sons, Inc.  相似文献   

14.
The molecular and crystal structure of naturally occurring optically active poly(β-hydroxy-butyrate) was analysed by X-ray diffraction. The unit cell is orthorhombic, P212121-D42, with a = 5·76 A?, b = 13·20 A?, and c(fibre period)=5·96 A? and two molecules pass through the unit cell. The molecule has a conformation shown by the Fischer projection for the case of a rectus polymer, giving a left-handed (21) helix.
The crystalline synthesized racemic polymer gave the same X-ray diffraction pattern as the naturally occurring optically active polymer. This result indicates that the racemic polymer has an isotactic configuration and consists of two kinds of crystallites, each composed only of the left-handed helices of rectus polymer chains, or only of the right-handed helices of sinister polymer chains.  相似文献   

15.
The effects of l-lactide dimer as additives on the crystallization behavior of poly(l-lactide) (PLLA) films were studied. Hence, neat PLLA films and PLLA containing l-lactide (5 % w/w) (PLLA/La) were prepared in dichloromethane at room temperature via solution casting. The non-isothermal cold crystallization of PLLA films were studied using differential scanning calorimetry at various heating rates including 2.5, 5, 7.5, 10 and 15 °C/min. However, the X C% was increased for PLLA/La films in comparison with neat PLLA films. The crystallization kinetics was then analyzed by the Avrami, Jeziorny, Ozawa and Mo kinetic models. It is found that all the kinetic models were established to describe the experimental data fairly well except the Ozawa model. The values of t 1/2, Z C and F(T) indicated that the crystallization rate increased with increase in heating rates for PLLA and PLLA/La films. However, l-lactide dimer incorporated in PLLA films accelerates the crystallization process of PLLA at the high heating rate. The nucleation constant (K g) and the surface free energy (σ e) based on Lauritzen–Hoffman theory indicated that these parameters for PLLA/La films is lower than neat PLLA.  相似文献   

16.
17.
The miscibility, crystallization behavior, and component interactions of two binary blends, poly(L ‐lactide) (L ‐PLA)/poly(vinylpyrrolidone) (PVP) and poly(D ,L ‐lactide) (DL ‐PLA)/PVP, were studied with differential scanning calorimetry and Fourier transform infrared (FTIR) spectroscopy. The composition‐dependent changes of the glass‐transition temperature (Tg) and degree of crystallinity (Xc) of the L ‐PLA phase indicated that L ‐PLA and PVP were immiscible over the composition range investigated. However, the sharp decrease of Xc with increasing PVP content in the second heating run demonstrated that the cold crystallization process of L ‐PLA was remarkably restricted by PVP. In DL ‐PLA/PVP blends, the existence of two series of isolated Tg's indicated that DL ‐PLA and PVP were phase‐separated, but evidence showed that there was some degree of interaction at the interface of the two phase, especially for the blends with low DL ‐PLA contents. FTIR measurements showed that there was no appreciable change in the spectra of L ‐PLA/PVP with respect to the coaddition of each component spectrum, implying the immiscibility of the two polymers. In contrast to L ‐PLA, the intermolecular interaction between DL ‐PLA and PVP was detected by FTIR; this was evidenced by the observation of a high‐frequency shift of the C?O stretching vibration band of PVP with increasing DL ‐PLA content, which suggested some degree of miscibility. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 973–979, 2003  相似文献   

18.
Several samples of poly(lactic acid) with different molecular weights and tacticity have been prepared, and some PLLA injection moulded specimens have been annealed to promote their crystallization. From the characterization data, poly(L -lactide) showed more interesting mechanical properties than poly(D, L -lactide), and its behavior significantly improves with crystallization. In fact, annealed specimens possess higher values of tensional and flexural modulus of elasticity, Izod impact strength, and heat resistance. The plateau region of flexural strength as a function of molecular weights appears around Mv = 35,000 for PDLLA and amorphous PLLA and at higher molecular weight, around Mv = 55,000, for crystalline PLLA. The study of temperature effect shows that at 56°C only crystalline PLLA still exhibits useful mechanical properties. © 1996 John Wiley & Sons, Inc.  相似文献   

19.
A series of biodegradable poly (glycerol-sebacate-l-lactide) (PGSLA) copolymers, with variable PLLA length, were synthesized and characterized. The copolymers comprised PGS backbone chain with a nominal molecular weight of 2,800 g/mol. The length of each PLLA side chain covered the 800–14,000 range, while the length of the PLLA was easily controlled by the feed molar ratio of the l-lactide to the PGS. The structure of the copolymer was studied by nuclear magnetic resonance spectroscopy and gel permeation chromatography. Differential scanning calorimetric measurements and thermal gravimetric analysis had been performed to indicate the glass transition temperature (T g), melting point (T m), and the degree of crystallinity (χ c). It was also found that the onset decomposition temperature (T d) of the copolymers was lower than those of the linear polylactide (LPLLA). After solution casting and solvent evaporation, porous structures were found in the copolymer films by scanning electron microscope (SEM). Water contact angle results showed that the hydrophilicity of the copolymers was much higher than that of linear PLLA. In vivo, PGSLA copolymer demonstrated a favorable tissue response profile compared to PGS/LPLLA blend. There was also significantly less inflammation and fibrosis during degradation. PGSLA might therefore serve as an excellent candidate material for medical applications, given its minimal in vivo tissue response.  相似文献   

20.
Poly(L-lactide) (PLLA) oligo-esters with α-hydroxyl-ω-alkyl (alkyl = −CH2−[CH2−CH2]m−CH3, where m = 1, 2, 4, 5, 6, 7, 8, 9, and 10) end groups were synthesized by ring-opening polymerization of L-lactide (L-LA) catalyzed by tin(II) 2-ethylhexanoate Sn(Oct)2 in the presence of aliphatic alcohols as initiators (HO−CH2−[CH2−CH2]m−CH3, where m = 1, 2, 4, 5, 6, 7, 8, 9, and 10). High yields (~ 62 to 71%) and M n(NMR) in the range of 2120–2450 Da (PLLA) were obtained. Effects of alkyl end groups on thermal properties of the oligo-esters were analyzed by DSC, TGA and SAXS. Glass transition temperature (T g) gradually decreases with increase in the percent of−CH2−[CH2−CH2]m−CH3 end group, as results alkyl end group provides most flexibility to PLLA. An important effect of alkyl end group on a double cold crystallization (T c1 and T c2) was observed, and is directly related with the segregation phase between alkyl end group and PLLA. TGA analysis revealed that PLLA oligo-esters are more thermally stable with docosyl (−C22H45) respect to the butyl (−C4H9) end group, probably is due to steric hindrance of the end group (docosyl respect to butyl) toward intermolecular and intramolecular transesterification. SAXS analysis showed that alkyl end group as docosyl restricted the growth of lamellae thickness (D) due to steric hindrance. Characterization of hydroxyl and alkyl end groups in the PLLA oligo-esters was determined by MALDI-TOF, GPC, FT-IR and 1 H and 13 C NMR.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号