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 343 毫秒
1.
Linear 1-arm and 2-arm poly(l-lactide) [i.e., poly(l-lactic acid) (PLLA)] polymers having relatively low number-average molecular weights (Mn) (≤5 × 104 g mol−1) were synthesized by ring-opening polymerization of l-lactide initiated with tin(II) 2-ethylhexanoate (i.e., stannous octoate) and coinitiators of l-lactic acid, 1-dodecanol (i.e., lauryl alcohol), and ethylene glycol (these PLLA polymers are abbreviated as LA, DN, and EG, respectively). For Mn below 1.5 × 104 g mol−1, non-isothermal crystallization during heating and isothermal spherulite growth were disturbed in linear 2-arm PLLA (EG) compared to those in linear 1-arm PLLA (LA and DN). This finding indicates that the chain directional change, the incorporation of the coinitiator moiety as an impurity in the middle of the molecule, and their mixed effect disturbed the crystallization of linear 2-arm PLLA compared to that of linear 1-arm PLLA, in which the chain direction is unvaried and the coinitiator moiety is incorporated in the chain terminal. Also, the finding strongly suggests that the reported low crystallizability of multi-arm PLLA (arm number ≥ 3) compared to that of linear 1-arm PLLA is caused not only by the presence of branching points but also by the chain directional change, the incorporation of the coinitiator moiety in the middle of the molecule, and their mixed effect. The effects of the chain directional change and the position of the incorporated coinitiator moiety on the crystallization and physical properties of linear 1-arm and 2-arm PLLA decreased with an increase in Mn.  相似文献   

2.
The non-isothermally and isothermally crystallized stereodiblock copolymers of poly(l-lactide) (PLLA) and poly(d-lactide) (PDLA) with equimolar l-lactyl and d-lactyl units and different number-average molecular weights (Mn) of 3.9 × 103, 9.3 × 103, and 1.1 × 104 g mol−1, which are abbreviated as PLLA-b-PDLA copolymers, contained only stereocomplex crystallites as crystalline species, causing higher melting temperatures of the PLLA-b-PDLA copolymers compared to those of PLLA homopolymers. In the case of non-isothermal crystallization, the cold crystallization temperatures of the PLLA-b-PDLA copolymers during heating and cooling were respectively lower and higher than those of PLLA homopolymers, indicating accelerated crystallization of PLLA-b-PDLA copolymers. In the case of isothermal crystallization, in the crystallizable temperature range, the crystallinity (Xc) values of the PLLA-b-PDLA copolymers were lower than those of the PLLA homopolymers, and were susceptible to the effect of crystallization temperature in contrast to that of homopolymers. The radial growth rate of the spherulites (G) of the PLLA-b-PDLA copolymers was the highest at the middle Mn of 9.3 × 103 g mol−1. This trend is different from that of the PLLA homopolymers where the G values increased monotonically with a decrease in Mn, but seems to be caused by the upper critical Mn values of PLLA and PDLA chains as in the case of PLLA/PDLA blends (in other papers), above which homo-crystallites are formed in addition to stereocomplex crystallites. The disturbed crystallization of PLLA-b-PDLA copolymers compared to that of the PLLA/PDLA blend is attributable to the segmental connection between the PLLA and PDLA chains, which interrupted the free movement of those chains of the PLLA-b-PDLA copolymers during crystallization. The crystallite growth mechanism of the PLLA-b-PDLA copolymers was different from that of the PLLA/PDLA blend.  相似文献   

3.
The spherulite growth behavior and mechanism of l-lactide copolymers, poly(l-lactide-co-d-lactide) [P(LLA-DLA)], poly(l-lactide-co-glycolide) [P(LLA-GA)], and poly(l-lactide-co-ε-caprolactone) [P(LLA-CL)] have been studied using polarization optical microscopy in comparison with poly(l-lactide) (PLLA) having different molecular weights to elucidate the effects of incorporated comonomer units. The incorporation of comonomer units reduced the radius growth rate of spherulites (G) and increased the induction period of spherulite formation (ti), irrespective of the kind of comonomer unit. Such effects became remarkable with the content of comonomers. At a crystallization temperature (Tc) of 130 °C, the disturbance effects of comonomers on the spherulite growth decreased in the following order: d-lactide>glycolide>ε-caprolactone, when compared at the same comonomer unit or reciprocal of averaged l-lactyl unit sequence length (ll). The ti estimation indicated that the glycolide units have the lowest disturbance effects on the formation of spherulite (crystallite) nuclei. The PLLA having the number-average molecular weight (Mn) exceeding 3.1×104 g mol−1 showed the transition from regime II to regime III at Tc=120 °C, whereas PLLA with the lowest Mn of 9.2×103 g mol−1 crystallized solely in regime III kinetics and the copolymers excluding P(LLA-DLA) with 3% of d-lactide units crystallized solely according to regime II kinetics. The nucleation and front constant for regime II and III [Kg(II), Kg(III), G0(II), and G0(III), respectively] estimated with each (not with a fixed for high-molecular-weight PLLA) decreased with increasing the amount of defects per unit mass of the polymer for crystallization, i.e. with increasing the comonomer content and the density of terminal group through decreasing the molecular weight.  相似文献   

4.
Yun Hu  Jianming Zhang  Isao Noda 《Polymer》2008,49(19):4204-4210
The miscibility, crystallization and subsequent melting behavior in binary biodegradable polymer blends of poly(l-lactic acid) (PLLA) and low molecular weight poly(3-hydroxybutyrate) (PHB) have been investigated by differential scanning calorimetry (DSC), Fourier-transform infrared (FTIR) spectroscopy, and wide-angle X-ray diffraction (WAXD). DSC analysis results indicted that PLLA showed no miscibility with high molecular weight PHB (Mw = 650,000 g mol−1) in the 80/20, 60/40, 40/60, 20/80 composition range of the PHB/PLLA blends. On the other hand, it showed some limited miscibility with low molecular weight PHB (Mw = 5000 g mol−1) when the PHB content was below 25%, as evidenced by small changes in the glass transition temperature of PLLA. The partial miscibility was further supported by changes of cold-crystallization behavior of PLLA in the blends. During the nonisothermal crystallization, it was found that the addition of a small amount of PHB up to 30% made the cold-crystallization of PLLA occur in the lower temperature. Meanwhile, the crystallization of PHB and PLLA was observed in the heating process by monitoring characteristic IR bands of each component for the low molecular weight PHB/PLLA 20/80 and 30/70 blends. The temperature-dependent IR and WAXD results also revealed that for PLLA component crystallization, the disorder (α′) phase of PLLA was produced, and that the α′ phase changed to the order (α) phase just prior to the melting point.  相似文献   

5.
Kazunori Se  Masaaki Suzuki 《Polymer》2010,51(21):4760-4768
The anionic living-equilibrium polymerization (ALEP) of (p-isopropenylphenethyl)poly(α-methylstyrene) macromonomer (PMStM) was studied: PMStM was simultaneously purified and initiated above a ceiling temperature (Tc) and then propagated below Tc. Three PMStMs (Mn = 1.77 × 103, 3.65 × 103, and 5.36 × 103) were prepared by the coupling of poly(α-methylstyryllithium) with p-isopropenylphenethyl chloride, and were polymerized to produce polymacromonomers, (PMSt)m (m = 9.5, 4.1, and 3.5) by n-BuLi in d8-THF at temperatures below −10 °C under high vacuum via ALEP. From kinetic studies using a 1H NMR technique, the propagation rate constant (0.240, 0.110, and 0.079 l mol−1 s−1) at −78 °C was found to be proportional to a reciprocal of Mn of PMStM, and its Mn dependence was discussed based on a simple diffusion-controlled theory. The thermodynamic characteristics of ΔHMSt (−7.48, −6.73, and −3.82 kJ mol−1) and ΔSMSt (−26.8, −24.1, and −13.7 J mol−1 K−1) were determined from a kinetic analysis, and their Mn dependence was also discussed. The TC values for the three PMStM, including α-methylstyrene, were found to be the same (6.0 °C).  相似文献   

6.
Star-shaped 4-armed stereo diblock poly(lactide)s with the core/shell types of poly(d-lactide) (PDLA)/poly(l-lactide) (PLLA) and PLLA/PDLA (abbreviated as 4-DL and 4-LD, respectively) and the number-average molecular weights of about 1 × 104 g mol−1 were synthesized and the crystallization behavior of neat 4-DL, 4-LD, and their 50/50 blend (abbreviated as 4-DL/4-LD blend) was investigated. Solely stereocomplex (SC) crystallites as crystalline species were formed in the neat 4-DL, 4-LD, and 4-DL/4-LD blend, irrespective of crystallization temperature (100–160 °C). The overall SC crystallization of 4-DL/4-LD blend was highly accelerated compared with that of neat 4-DL and 4-LD, due to the largely elevated spherulite nuclei number per unit mass in the blend. Such high density of nuclei formation in 4-DL/4-LD blend is attributable to the facile intermolecular interaction and subsequent SC nucleation between the PLLA shell of 4-DL and the PDLA shell of 4-LD. The blending method reported in the present study is applicable for various core/shell types of star-shaped stereo diblock stereocomplexationable polymers to accelerate overall SC crystallization and can counterbalance the lowered crystallization rate caused by the star-shaped architecture. Despite the highly accelerated overall SC crystallization of 4-DL/4-LD blend by blending 4-DL and 4-LD, the spherulite growth rate, induction period for spherulite growth, final crystallinity, crystalline species, growth morphology, and crystallization mechanism were not altered by blending 4-DL and 4-LD.  相似文献   

7.
The nitrone mediated polymerization of methyl methacrylate (MMA) via the enhanced (termination) spin capturing polymerization (ESCP) process is made possible via the addition of small amounts of styrene (between 5 and 10 vol.%) to the reaction mixture. Efficient control over the molecular weight between 7000 and 57,000 g mol−1 (at 60 °C) yields macromolecules that feature a mid-chain alkoxyamine functionality and are rich in methyl methacrylate. The collated kinetic and molecular weight data allow for a deduction of the spin capturing constant, CSC, in the range between 0.15 and 0.30. During the ESCP process, the number average molecular weight, Mn, of the formed mid-chain functional polymer is constant up to high monomer to polymer conversions (i.e. 80%). The high degree of alkoxyamine mid-chain functionality present in the generated polymeric material is evidenced via a subsequent nitroxide-mediated polymerization process employing the formed ESCP polymer, indicating a chain extension from 37,700 to 118,000 g mol−1 with a concomitant reduction in polydispersity (from 2.3 to 1.5).  相似文献   

8.
Sunil Dhamaniya 《Polymer》2010,51(23):5392-5399
A series of aliphatic polyesters based on tartaric acid and its derivatives were synthesized starting from naturally occurring L-tartaric acid. The hydroxyl groups of the tartaric acid derivatives were first protected and the polyesters were synthesized by bulk and solution polycondensation methods. Two classes of polyesters were synthesized and characterized, the first by polycondensation of dimethyl 2,3-O-isopropylidene-l-tartrate with various alkanediols, and the second by reaction of 2,3-O-isopropylidene-l-threitol with various diacid chlorides. Acid catalyzed deprotection of isopropylidene groups gave well-defined polyesters having pendant hydroxyl functional groups regularly distributed along the polymer chain. The number average molecular weights (Mn) of the polymers were found to vary in the range of 2.3-15.7 × 103 g mol−1. Differential scanning calorimetry (DSC) analysis showed the glass transition temperatures (Tg) of the polyesters varied from −36.1 °C to 17.9 °C on varying the chain length.  相似文献   

9.
The effects of polymerization time and temperature on the molecular weight and molecular weight distribution of polyethylene, produced over homogeneous catalyst bis[N-(3-tert-butyl salicylidene)anilinato]zirconium(IV) dichloride tBu-L2ZrCl2/MAO have been studied. The data on the number of active centers (CP) and propagation rate constants (kP) at different polymerization time have been obtained as well. It was found that at a short polymerization time two types of active centers, producing low molecular weight PE (Mw = (4-10) × 103 g mol−1) are formed. The number of these centers was estimated to be 11% of total zirconium complex and their reactivity is very high (the kP value was found to be 54 × 103 L mol−1 s−1 at 35 °C). High initial activity of the catalyst fell with the increase in polymerization time, whereas the polydispersity values of the resulting PE increase due to formation of new centers, producing high molecular weight PE (Mw = (30-1300) × 103 g mol−1). It was found that the decrease in activity is caused by reducing the initial active centers number and lower reactivity of the new-formed centers (kP = 17 × 103 L mol−1 s−1).  相似文献   

10.
Visual recognition of the formation of supramolecular graft polymers in aqueous solution driven by host-guest assembly is presented. The interaction between β-cyclodextrin (β-CD) and phenolphthalein moieties can be followed via a distinctive change of color in basic media by optical inspection. The phenolphthalein-containing lateral polymer chain was obtained via reversible addition-fragmentation chain transfer (RAFT) copolymerization of N,N-dimethylacrylamide and N-(2-hydroxy-5-(1-(4-hydroxyphenyl)-3-oxo-1,3-dihydroisobenzofuran-1-yl)benzyl)acrylamide. The RAFT polymerizations afforded statistical copolymers with molecular masses between 6900 and 8800 g mol−1 and molar mass dispersities (?) between 1.1 and 1.2. β-CD-functionalized building blocks were synthesized using a propargyl-functionalized chain transfer agent for the polymerization of N,N-diethylacrylamide followed by a copper(I)-catalyzed cycloaddition with β-CD-azide (Mn = 8200 g mol−1; ? = 1.2). Self-assembly of both polymers to form supramolecular graft polymers was evidenced by 2D NOESY NMR, UV–vis spectroscopy and dynamic light scattering.  相似文献   

11.
Qiang Liu  Yu Zhang  Peng-Fei Yan  Ri-Wei Xu 《Polymer》2010,51(25):5960-5969
The initiating system consisting of AlCl3 with dialkyl ether such as di-n-butyl ether or diisopropyl ether has been successfully developed for providing a cost-effective process of synthesis of highly reactive polyisobutylenes (HRPIBs) with large proportion of exo-olefin end groups up to 93 mol% at temperatures ranging from −20 to +20 °C. The above dialkyl ethers played very important roles in promoting the directly rapid β-proton elimination from -CH3 of the growing chain ends to create exo-olefin end groups and decreasing or even suppressing the carbenium ion rearrangements to form the double bond isomers. Very importantly, the highly reactive PIBs with 80-92 mol% of exo-olefin end groups, having low Mns of 1300-2300 g mol−1 and monomodal molecular weight distribution (Mw/Mn = 1.7-2.0) could be achieved at 0-20 °C. These results are comparable to those of commercial HRPIBs produced industrially by the best BF3-based initiating system at far below 0 °C.  相似文献   

12.
Hideto Tsuji  Ippei Fukui 《Polymer》2003,44(10):2891-2896
Poly(l-lactide) (i.e. poly(l-lactic acid) (PLLA)) and poly(d-lactide) (i.e. poly(d-lactic acid) (PDLA)) and their equimolar enantiomeric blend (PLLA/PDLA) films were prepared and the effects of enantiomeric polymer blending on the thermal stability and degradation of the films were investigated isothermally and non-isothermally under nitrogen gas using thermogravimetry (TG). The enantiomeric polymer blending was found to successfully enhance the thermal stability of the PLLA/PDLA film compared with those of the pure PLLA and PDLA films. The activation energies for thermal degradation (ΔEtd) were evaluated at different weight loss values from TG data using the procedure recommended by MacCallum et al. The ΔEtd values of the PLLA/PDLA, PLLA, and PDLA films were in the range of 205-297, 77-132, and 155-242 kJ mol−1 when they were evaluated at weight loss values of 25-90% and the ΔEtd value of the PLLA/PDLA film was higher by 82-110 kJ mol−1 than the averaged ΔEtd value of the PLLA and PDLA films. The mechanism for the enhanced thermal stability of the PLLA/PDLA film is discussed.  相似文献   

13.
Resin immobilized lipase B from Candida antarctica (CALB) was used to catalyze the condensation polymerization of two difuctional siloxane and poly(ethylene glycol) systems. In the first system, 1,3-bis(3-carboxypropyl)tetramethyldisiloxane was reacted with poly(ethylene glycol) (PEG having a number-average molecular weight, Mn = 400, 1000 and 3400 g mol−1, respectively). In the second system, α,ω-(dihydroxy alkyl) terminated poly(dimethylsiloxane) (HAT-PDMS, Mn = 2500 g mol−1) was reacted with α,ω-(diacid) terminated poly(ethylene glycol) (PEG, Mn = 600 g mol−1). All the reactions were carried out in the bulk (without use of solvent) at 80 °C and under reduced pressure (500 mmHg vacuum gauge). The progress of the polyesterification reactions was monitored by analyzing the samples collected at various time intervals using FTIR and GPC. The thermal properties of the copolymers were characterized by DSC and TGA. In particular, the effect of the chain length of the PEG block on the molar mass build up and on the thermal stability of the copolymers was also studied. The thermal stability of the enzymatically synthesized copolymers was found to increase with increased dimethylsiloxane content in the copolymers.  相似文献   

14.
Double hydrophilic block copolymers (DHBC) were prepared by end-to-end coupling of two biocompatible water-soluble homopolymers: the polysaccharide dextran (Mw 8300 or 14,700 g mol−1) and ω-amino poly(ethylene glycol) (PEG-NH2, Mw 3000 or 7000 g mol−1). The synthesis involved, first, specific oxidation of the dextran terminal aldehyde group and, second, covalent linkage of PEG-NH2 via a lactone aminolysis reaction. The diblock copolymers dextran-PEG (DEX-PEG) were converted in high yield into the corresponding carboxymethyldextran-PEG (CMD-PEG) derivatives with control over the degree of substitution, from 30 to 85 mol% CH2COOH groups per glucopyranosyl unit. Further modifications of a CMD-PEG block copolymer led to N-(2-aminoethyl)carbamidomethyldextran-PEG yielding a pair of oppositely-charged DHBC of identical charge density, chain length, and neutral block/charged block content. The properties of CMD-PEG in aqueous solutions were studied by static and dynamic light scattering as a function of solution pH, providing evidence of the pH-sensitive assembly of the copolymers driven by inter- and intra-chain hydrogen-bond formation.  相似文献   

15.
Nitroxide-mediated radical polymerization of a tailor-made acrylate carrying a 1,2,3-triazole group with an undecanoyl spacer affords a well-defined (Mn = 7860 g mol−1 and D = 1.39) neutral polyacrylate precursor. A series of 1,2,3-triazolium-based poly(ionic liquid)s (TPILs) is then obtained by straightforward quaternization of the 1,2,3-triazole groups with methyl iodide and subsequent anion metathesis reactions. Among the prepared materials, TPIL with bis(trifluoromethane)sulfonimide anion exhibits low glass transition temperature (Tg = −40 °C), high thermal stability (Td10 = 325 °C) and anhydrous ionic conductivity of 4 × 10−6 S cm−1 at 30 °C, as measured by differential scanning calorimetry, thermogravimetric analysis and broadband dielectric spectroscopy, respectively.  相似文献   

16.
The development of the first star-structured poly(l-lactides) (s-PLLAs) with dual functionality as both nucleating agent and plasticizer in polylactide (PLA) blends is described. Mechanisms controlling this functionality are deduced. Blends of PLA containing s-PLLAs show significant improvements in thermal and mechanical properties. The s-PLLAs are made by using appropriate saccharides, e.g., methyl-α-d-glucopyranoside and β-cyclodextrin, as star-shaped core macroinitiators for l-lactide polymerization. Varying the l-lactide mole ratio gives control of the degree of polymerization (DPn) of PLLA branches from 5 to 30. Blending ≈1 wt% of the new s-PLLAs with PLA resins produces a drastic decrease in the glass transition temperature (Tg) (from 57 °C to 22 °C), and an increase in the elongation at break (30–40%) confirms significant increases in chain mobility and plasticity. A concomitant reduction of crystallization temperature (from 127 °C to 81 °C) as well as a higher crystallization rate (a factor of four increase) implies rapid nucleation of crystalline domains. With only 1 wt% of s-PLLA in the blend of PLA film, the elongation at break of PLA increases eightfold. In addition to the dual functionality, the PLLA branches also provide miscibility with PLA in the blend (as confirmed by a single Tg, combined with estimates from the Fox equation and solubility parameters).  相似文献   

17.
Kelly S. Anderson 《Polymer》2006,47(6):2030-2035
A melt blending procedure was developed for the preparation of poly(l-lactide) (PLLA)/poly(d-lactide) (PDLA) stereocomplex crystallites dispersed in a PLLA matrix. All PLLA/PDLA blends were prepared in a batch melt mixer with ≥95% PLLA. Three PDLA homopolymers with a range of molecular weights were used as the minority (≤5%) component. The presence of the stereocomplex in the PLLA matrix was verified by differential scanning calorimetry (DSC) and optical microscopy. The effectiveness of the in situ formed stereocomplex crystallites for nucleating PLLA crystallization was evaluated using self-nucleation and non-isothermal DSC methods. With only 3 wt% of the 14 kg mol−1 PDLA, nucleation efficiencies near 100% could be obtained. In addition, fast crystallization kinetics were observed in isothermal crystallization experiments at 140 °C. The stereocomplex crystallites were much more effective at enhancing the crystallization rate of PLLA compared to talc, a common nucleating agent.  相似文献   

18.
Shaoyong Huang  Hongfei Li  Xuesi Chen 《Polymer》2011,52(15):3478-3487
Shear effect on crystal structure, morphology and melting behavior of poly(l-lactide) (PLLA) were investigated by wide-angle X-ray diffraction (WAXD), differential scanning calorimetry (DSC) and in-situ polarized optical microscope (POM). Influences of steady shear on PLLA melt was investigated during annealing. The impacts of crystallization temperature (Tc) and shear rate (SR) on crystal structure, crystal growth rate, nucleation, and melting behavior were studied. It could be clarified into the low (<20 s−1) and high shear rate ranges in terms of the crystal structure, growth rate, nucleation and melting behavior. It was found that phase transition and recrystallization during heating completely changed the stack structure. A new stable structure was established after a temperature jump in the melting range. Cylindric morphology arranged with spherulite was also observed in the polymer films at high Tc influenced by shear effect.  相似文献   

19.
The number average molar mass Mn of poly(octadecene-alt-maleic anhydride) (PODMA) copolymers calculated from data obtained by size exclusion chromatography (SEC) using a polystyrene (PS) calibration was found to be inaccurate. The use of SEC combined with dilute solution viscometry enabled a method to be developed using an iterative approach, which does not require knowledge of the Mark-Houwink constants for PODMA samples. A new calibration curve was constructed as a plot of molar mass Mu for PODMA. True number-average molar masses Mn (true) calculated using the new calibration are approximately twice the apparent molar mass Mn (app) based on a PS calibration for higher molar mass samples (>10?000 g mol−1).  相似文献   

20.
The present work shows that the enzyme mediated ring-opening polymerization of l-lactide at 65 °C can be achieved using supercritical carbon dioxide. It is reported a biphasic media system where the supercritical phase coexists with a liquid organic phase, which is mainly composed of melted monomer, wherein the growing poly-l-lactide chains are soluble. The immobilized lipase B from Candida antartica was used as the biocatalyst. The results indicated that semi-crystalline polymers with a molecular weight (Mw) up to 12,900 g mol−1 can be attained and that the monomer conversion is related to the biocatalyst concentration and its initial water activity (awi). Experiments carried out with denatured enzyme gave no monomer conversion which confirms that the enzymatic mechanism is only involved in our system.  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