共查询到20条相似文献,搜索用时 15 毫秒
1.
Akira Kaito Yuko Iwakura Yongjin Li Kazuo Nakayama Hiroshi Shimizu 《Macromolecular chemistry and physics.》2007,208(5):504-513
Unique orientation textures have been induced by the confined crystal growth of PVDF in drawn films of PVDF/PA6 blends. Oriented films of PVDF/PA6 blends were prepared by uniaxially drawing melt‐mixed blends. The drawn films with fixed lengths were heat‐treated at 180 °C for 3 minutes to melt the PVDF component, followed by non‐isothermal crystallization of PVDF at a cooling rate of 0.5 °C · min?1. The crystal orientation was studied by WAXD. When PVDF was melted and recrystallized in the drawn films of the PVDF/PA6 = 50/50 blend at a slow cooling rate, the crystal b‐axis of the α‐crystalline form of PVDF was oriented in the drawing direction, forming orthogonal orientation textures. SEM showed that stretched domains of PVDF with diameters of 0.2–0.5 µm were dispersed in the PA6 phase in the drawn films of the PVDF/PA6 = 50/50 blend. Spatial confinement of the crystal growth resulted in the alignment of the crystal b‐axis along the long axis of the domains, because PVDF is crystallized in thin cylindrical domains. The orientation behavior is different from the oriented crystallization of PVDF/PA11 (Y. Li, A. Kaito, Macromol. Rapid Commun. 2003 , 24, 255), in which transcrystallization from the interface causes the a‐axis orientation to be in the drawing direction. It is thought that the domain size influenced the mechanism of oriented crystallization and the resultant crystal orientation.
2.
Forrest A. Landis Stephen R. March Jr. Dakshinamoorthy Deivasagayam Robert T. Mathers 《Macromolecular chemistry and physics.》2014,215(2):153-162
A series of poly(methyl methacrylate‐co‐methacrylic acid) (PMMA‐co‐MAA) random copolymers ranging in MAA content from 0–15 mol% is synthesized and blended with poly(vinylidene fluoride) (PVDF). Using infrared spectroscopy, it is observed that the absorption bands attributed to hydrogen‐bonded carbonyl groups increase in intensity as the amount of MAA in the copolymer increases. In DSC analysis, the crystallization temperature of the PVDF in the blend initially decreases by ca. 12 °C with MAA contents ranging from 0 to 5.5 mol%; however, a PVDF blend with a 15 mol% MAA copolymer has a crystallization temperature that is only ca. 3 °C below that of pure PVDF. Similarly, spherulitic growth rate analysis initially shows a decrease in radial growth rate for PVDF in blends with PMMA‐co‐MAA copolymers containing less than 5.5 mol% MAA. At higher MAA copolymer contents, the spherulitic growth rate approaches that of pure PVDF. It is concluded that the presence of the MAA comonomer in the PMMA‐co‐MAA copolymer initially (<5.5 mol% MAA) increases the intermolecular interactions between the copolymer and the PVDF. However, as the MAA content of the copolymer rises above 5.5 mol%, intramolecular hydrogen bonding interactions within the PMMA‐co‐MAA copolymer cause the copolymer to be less compatible with PVDF.
3.
Meral Akkoyun Christian Carrot Benoît Blottière 《Macromolecular chemistry and physics.》2012,213(5):587-593
Experiments on the thermoreversible gelation of a ternary system composed of poly(vinylidene fluoride) (PVDF) and poly[(vinylidene fluoride)‐co‐hexafluoropropene] (P(VDF‐co‐HFP)) in propylene carbonate (PC) or DMSO solutions are carried out and the resulting ternary diagrams are plotted for various systems. The results show that the gelation occurs at relatively low concentrations. The miscibility area of the ternary diagram is larger with DMSO than with PC. In addition, the concentration at the gelation point increases with increasing temperature. The gelation line is almost independent of the molecular structure of PVDF in PC but varies in DMSO. This variation of the miscibility is essentially connected to the carboxylic functions on the PVDF chain. 相似文献
4.
Kyung‐Moo Shin Tungalag Dong Yong He Yoshio Inoue 《Macromolecular chemistry and physics.》2006,207(8):755-762
Summary: The effects of methylated CDs, α‐ and β‐DMCD and α‐ and β‐TMCD on the crystallization behavior of P(3HB) have been investigated by employing DSC and polarized optical microscopy (POM). The crystallization and nucleation of P(3HB) was improved by an addition of appropriate amount of each methylated CD, while negative effect of miscible blend was observed when the amount of addition exceeds the optimum value for the maximum nucleating effect. β‐TMCD was shown to be the most effective for the enhancement of the P(3HB) crystallization, thought to be due to the possibility of IC formation with P(3HB).
5.
Samy A. Madbouly Nahed Y. Abdou Ashraf A. Mansour 《Macromolecular chemistry and physics.》2006,207(11):978-986
Summary: Phase behavior and isothermal crystallization kinetics of poly(ε‐caprolactone) (PCL) blends with tetramethyl polycarbonate (TMPC) and poly(styrene‐co‐acrylonitrile) with 27.5 wt.‐% acrylonitrile content have been investigated using broadband dielectric spectroscopy and differential scanning calorimeter. An LCST‐type phase diagram has been observed for PCL/SAN blend while all the different blend compositions of PCL/TMPC were optically clear without any phase separation structure even at high temperatures up to 300 °C. The composition dependence of Tgs for both blends has been well described by the Gordon‐Taylor equation. The phase diagram of PCL/SAN was theoretically calculated using the Flory‐Huggins equation considering that the interaction parameter is temperature and composition dependent. The equilibrium melting point of PCL depressed in the blend and the magnitude of the depression was found to be composition dependent. The interaction parameters of PCL with TMPC and SAN could not be calculated from the melting point depression based on Nishi‐Wang approach. The isothermal crystallization kinetics of PCL and in different blends was also investigated as a function of crystallization temperature using broadband dielectric spectroscopy. For pure PCL the rate of crystallization was found to be crystallization temperature (Tc) dependent, i.e., the higher the Tc, the lower the crystallization rate. The crystallization kinetics of PCL/TMPC blend was much slower than that of PCL/SAN at a constant crystallization temperature. This behavior was attributed to the fact that PCL is highly interacted with TMPC than SAN and consequently the stronger the interaction the higher the depression in the crystallization kinetics. It was also attributed to the different values of Tg of TMPC (191 °C) and SAN (100 °C); therefore, the tendency for crystallization decreases upon increasing the Tg of the amorphous component in the blend. The analysis of the isothermal crystallization kinetics was carried out using the theoretical approach of Avrami. The value of Avrami exponent was almost constant in the pure state and in the blends indicating that blending simply retarded the crystallization rate without affecting the crystallization mechanism.
6.
M. Habibur Rahman Sudip Malik Arun K. Nandi 《Macromolecular chemistry and physics.》2003,204(14):1765-1770
Blends of poly(vinylidene fluoride) (PVF2) and dodecyl benzene sulfonic acid (DBSA) were found to phase separate in the melt state from hot‐stage polarizing optical microscopy and differential scanning calorimetry (DSC). Both these studies also indicated the formation of a reversible mesomorphic order in the DBSA‐rich phase of the blends. Scanning electron microscopy (SEM) showed a fibrillar network morphology of the DBSA‐rich phase. Wide‐angle X‐ray diffraction (WAXD) revealed the formation of a lamellar structure in the blends corresponding to a long period of 26 Å. The mesomorphic nature of the blends is believed to arise from the supramolecular organization of the surfactant molecules within the lamella. Thermal study indicated that the supramolecularly organized complex is the VF2 monomer unit:DBSA ≈ 4:1, and the enthalpy of complexation is ~723 J · mol?1.
7.
Chin Han Chan Claudia Kummerlwe Hans‐Werner Kammer 《Macromolecular chemistry and physics.》2004,205(5):664-675
Summary: Blends of high molecular weight poly(R‐3‐hydroxybutyrate) (PHB) ( = 352 000 g · mol?1), comprising of either low molecular weight poly(R‐3‐hydroxybutyrate) (D‐PHB) ( = 3 900 g · mol?1) or poly[(R‐3‐hydroxybutyrate)‐co‐(R‐3‐hydroxyvalerate)] (PHBV) ( = 238 000 g · mol?1) with 12 mol‐% hydroxyvalerate (HV) content as a second constituent, were investigated along with the thermal properties and morphologies. After isothermal crystallization, a lowering of the melting temperature of PHB can be observed with increasing content of the second component in the blends. This behavior points towards miscibility of the constituents both in the liquid and the solid state. Crystallization kinetics was studied under isothermal and non‐isothermal conditions. The overall kinetics of isothermal crystallization was analyzed in terms of the Avrami equation. Only one crystallization peak is observed in all cases for the PHB/D‐PHB and PHB/PHBV blends under the conditions studied. This demonstrates co‐crystallization of the constituents. The addition of D‐PHB or PHBV to PHB reduces the rate of crystallization of the blends compared to that of neat PHB. The corresponding activation energies of crystallization also decrease with an increasing concentration of the second constituent. Non‐isothermal crystallization, carried out with different cooling rates held constant, is discussed in terms of a quasi‐isothermal approach. The corresponding rate constants as functions of reciprocal undercooling show Arrhenius‐like behavior in a certain range of temperatures. At sufficiently high undercooling, the rate constants of crystallization for the isothermal process exceed those reflecting non‐isothermal conditions, whereas in the limit of low undercoolings, the rate constants become similar. Ring‐banded morphologies are observed when PHB is in excess. When the respective second component is the major component, fibrous textures of the spherulites develop.
8.
Summary: The isothermal crystallization process of poly(ε‐caprolactone) (PCL) and its blend with styrene‐acrylonitrile random copolymer containing 27.5 wt.‐% acrylonitrile has been investigated as a function of crystallization and melting temperatures for different melting times in the vicinity of the lowest critical solution temperature (LCST)‐type phase diagram using DSC. The obtained results showed that the isothermal crystallization kinetics from the melt of PCL in the blend was greatly affected by the presence of SAN. For a given crystallization temperature, the crystallization half time (t0.5) in the case of PCL/SAN = 80/20 blend was much longer than the corresponding value for pure PCL. This depression in the crystallization kinetics of PCL in the blend was mainly attributed to the favorable interaction between the two components and the reduction in chain mobility as a result of increasing the Tg by adding the amorphous component (SAN). In addition, the value of t0.5 was found to be independent of the melting temperature (Tm) and melting time for pure PCL. The isothermal crystallization kinetics for samples that were crystallized after decomposition into different stages of phase separation were also investigated. A major acceleration in the crystallization kinetics was observed isothermally for the samples that had undergone phase separation in the early and middle stages. The isothermal crystallization kinetics was also analyzed based on the Avrami approach. Although the crystallization kinetics was accelerated to a great extent by the liquid‐liquid phase separation, no change in the crystallization mechanism could be predicted. Furthermore, the phase separation process of PCL/SAN blend has been studied indirectly by following the variation in t0.5 at different melting temperatures and melting times above the LCST phase diagram.
9.
Teri A. Walker Yuri B. Melnichenko George D. Wignall J. S. Lin Richard J. Spontak 《Macromolecular chemistry and physics.》2003,204(17):2064-2077
Previous efforts have demonstrated that high‐pressure CO2 can markedly influence the phase behavior of amorphous polymer blends. In this work, we examine the effect of high‐pressure CO2 on the miscibility of blends composed of glassy poly(methyl methacrylate) (PMMA) and semicrystalline poly(vinylidene fluoride) (PVDF). Blends of this type are known to exhibit lower critical solution temperature (LCST) behavior with partial miscibility up to ≈50–60 wt.‐% PVDF at ambient conditions. Two miscible PMMA/PVDF blends have been systematically exposed to high‐pressure CO2 at 35 °C and pressures below and above the critical pressure. Small‐angle X‐ray scattering reveals that the scattering intensity at high scattering angles shows little dependence on pressure at low CO2 pressures, but increases substantially at relatively high CO2 pressures. Transmission electron microscopy and differential scanning calorimetry analyses confirm that the blends are initially quasi‐homogeneous with diffuse PVDF‐rich dispersions and a single glass transition temperature. After exposure to relatively high CO2 pressures, however, the PVDF is found to crystallize within the PMMA‐rich matrix. Thermal recycling of these blends promotes homogenization, indicating that such CO2‐altered phase behavior is reversible.
10.
Poly(ethylene terephthalate) (PET) and poly(vinylidene fluoride) (PVF2) are miscible in the melt‐state for the whole composition range. The glass transition temperature (Tg) of the solvent cast film decreases with the decrease in WPET (weight fraction of PET) in the blend, however, the Tg for the repeated melt quenched blends remains invariant with WPET. The melting point (Tm) and crystallization temperature (Tc) of PET decrease significantly with decrease in WPET in the blend, but the Tm and Tc of PVF2 decrease slightly with increase in WPET. The crystallinity of both PET and PVF2 decreases with increasing concentration of the other component in the blend, however, the decrease is larger for the former. The equilibrium melting points (Tm0's) of PET in the blends are determined by the extrapolation procedures using (i) Tm–Tc method for 5% crystallinity and (ii) Tm–Ta method, where Tm, Tc and Ta are melting, crystallization and annealing temperatures, respectively. The data of both the methods indicate a large depression of Tm0 of PET with increase in PVF2 concentration. The χ12 values determined from both the data are essentially the same, –0.14. This negative value of χ12 indicates that the two polymers are miscible in the melt‐state, however, they are not miscible in the crystalline state. The onset of degradation of PET increases with increase in PVF2 concentration in the blend. 相似文献
11.
Hideto Tsuji Kohji Tashiro Leevameng Bouapao Makoto Hanesaka 《Macromolecular chemistry and physics.》2012,213(20):2099-2112
Poly(L ‐lactide) (PLLA) and poly(L ‐lactide‐co‐glycolide) (78/22)0 [P(LLA‐GA)] were phase‐separated in PLLA/P(LLA‐GA) blends, forming independent domains and thence crystallizing separately. The crystallization of PLLA was not disturbed or delayed by the presence of P(LLA‐GA) and vice versa. PLLA and poly(L‐lactide‐co‐D ‐lactide) (77/23) [P(LLA‐DLA)] were miscible in the PLLA/P(LLA‐DLA) blends, overall crystallization was delayed, and the growth of crystallites was disturbed in the presence of P(LLA‐DLA). In isothermal crystallization, the originally noncrystallizable P(LLA‐DLA) became crystallizable in the presence of PLLA, with which it cocrystallized. The disturbance effect on periodical lamella twisting of PLLA was larger for P(LLA‐GA) than for P(LLA‐DLA). 相似文献
12.
Summary: Small amplitude oscillatory shear rheology is employed to investigate the linear viscoelastic behavior of the LCST‐type PMMA/α‐MSAN polymer blends as a function of annealing temperature and time. The results reveal that when temperature approaches the separation temperature, the blends exhibit some characteristic of complex thermorheological behavior, such as a shoulder in the dynamic storage modulus G′ or the linear relaxation modulus G(t), the appearance of a loss tangent (tan δ) peak and the additional relaxation in the relaxation spectrum H(τ). All of these can be attributed to the enhanced concentration fluctuations near the phase boundary. The anomalous pretransitional behavior can be quantified to yield the binodal temperature from the inflexion of variation and the spinodal temperature on the basis of the mean field theory. Furthermore, the linear viscoelastic properties of the phase‐separated PMMA/α‐MSAN (80/20) blends can be described well by either Palierne's or Bousmina's emulsion model in virtue of the calculated value in the whole frequency region.
13.
Kyung‐moo Shin Tungalag Dong Yong He Yoshio Inoue 《Macromolecular chemistry and physics.》2006,207(3):310-317
Summary: The effects of β‐CD, partially methylated and per‐methylated β‐CDs on the crystallization behavior of P(3HB) have been investigated. All measurements, including FT‐IR spectra, WAXD, and DSC, showed that the details of interaction between P(3HB) and respective CDs were varied with the degree of methyl substitution of CDs. The crystallization of P(3HB) was enhanced by an addition of immiscible β‐CD, while it was restricted by the introduction of methylated CDs via the formation of miscible blend with di‐O‐methyl β‐CD and via the formation of an IC with tri‐O‐methyl β‐CD, respectively.
14.
Molecular modeling is used to explain how the resistance of poly[(L ‐lactide)‐co‐(D ‐lactide)] to hydrolysis is affected by the percentages of L ‐ and D ‐lactide and their arrangements in blocks or random arrangements in the polymer. Previous studies on improving the hydrolysis resistance of PLA have involved forming either poly(L ‐lactide)/poly(D ‐lactide) (PLLA/PDLA) polyblends or copolymers of L ‐ and D ‐lactide. In this study, molecular modeling was used to study the hydrolysis resistance of PLA containing various arrangements of L ‐ and D ‐lactide in the polymers. PLA copolymers are found to have less resistance to hydrolysis than a PLLA/PDLA polyblend having the same percentages of L ‐ and D ‐lactide because a polyblend can form more stereocomplexes, which is the most stable structure PLA can form.
15.
Summary: The objective of this study was to synthesize thermoplastic elastomers by the direct copolyesterification of reactive oligomers of poly[(ethylene terephthalate)‐co‐(ε‐caprolactone)] (PET) and poly(propylene oxide) (PPO). The synthesis of hard segment oligomers was achieved in two steps. The first step consisted of the glycolysis of PET leading to α,ω‐hydroxyl functionalized oligomers. The second step corresponded to the ring opening polymerization of ε‐caprolactone onto the hydroxyl end groups of the PET oligomers. Commercially available hydroxytelechelic poly(propylene oxide) was modified to obtain carboxytelechelic poly(propylene oxide). The chemical structure of the product was investigated by 1H NMR and size exclusion chromatography (SEC). Multiblock poly(ester‐ether) was then synthesized by polyesterification of hydroxytelechelic poly[(ethylene terephthalate)‐co‐(ε‐caprolactone)] with carboxytelechelic poly(propylene oxide) oligomers, using different catalysts and reaction conditions. The best stoichiometric ratio for the reaction was determined in order to obtain the highest possible . The chemical structure of the synthesized poly(ester‐ether) was investigated by size exclusion chromatography and 1H NMR. The thermal and thermomechanical behavior of the synthesized poly(ester‐ether) was investigated by differential scanning calorimetry and dynamic mechanical analysis, which showed that the poly(ester‐ether) behaved as a thermoplastic elastomer. This product could also be an interesting way for chemical recycling of PET waste.
16.
Maximilian Maier Christian Dollendorf Helmut Ritter Joachim E. Klee 《Macromolecular chemistry and physics.》2012,213(24):2598-2605
Acrylic acid copolymers bearing radically polymerizable methacrylic and itaconic moieties are synthesized. Starting from copolymers of tBA and APVE, prepared at different ratios, methacrylic or itaconic amides are obtained via polymer‐analogous reaction. After acidic elimination of isobutylene, polymers with free carbon acids and staggered reactivity can be prepared. Varying amounts of attached acrylamide moieties allow the adjustment of network densities. Rheological properties in HEMA/water matrices during curing are evaluated. 相似文献
17.
Hongjun Yang Manqing Yan Stergios Pispas Guangzhao Zhang 《Macromolecular chemistry and physics.》2011,212(23):2589-2593
Poly[(ethylene carbonate)‐co‐(ethylene oxide)] is synthesized by ROP of 1,3‐dioxolan‐2‐one using 1‐tert‐butyl‐4,4,4‐tris(dimethylamino)‐2,2‐bis[tris(dimethylamino)phosphoranylidenamino]‐2Λ5,Λ5‐catenadiphosphazene as the catalyst. NMR spectroscopy shows that the copolymer chain consists of EO‐EC‐EO‐EO sequences. GPC measurements indicate a PDI of 1.43–1.66 and show that the molecular weight increases with monomer conversion. Thermal analysis demonstrates that the copolymer is amorphous with a glass transition temperature of about –42 °C and an onset decomposition temperature of about 200 °C. 相似文献
18.
Nilhan Kayaman‐Apohan Bahattin M. Baysal 《Macromolecular chemistry and physics.》2001,202(7):1182-1188
Semi‐interpenetrating hydrogel networks (IPNs) composed of poly(2‐hydroxyethyl methacrylate), PHEMA, and poly[(D ,L ‐lactic acid)‐co‐(ε‐caprolactam)], copoly(D,L ‐LA/ε‐CLM), were synthesized. Linear copoly(D,L ‐LA/ε‐CLM) chains were interpenetrated into the crosslinked three‐dimensional networks of PHEMA. Physical properties of IPN samples such as stress‐strain behavior, swelling, extraction and glass transition (Tg) were examined. Glass transition temperature of the PHEMA was shifted to lower temperatures, indicating interpenetration of PHEMA and copoly(D,L ‐LA/ε‐CLM) chains. The extraction results showed that the entrapping level between two components of PHEMA/copoly(D,L ‐LA/ε‐CLM) in simultaneous semi‐IPN is very high. These results confirm that the decrease in swelling ratios with increase in copoly(D,L ‐LA/ε‐CLM) contents is probably due to the entanglement effect. Semi‐IPNs showed higher ultimate strength and modulus but lower elongation with respect to pure PHEMA hydrogel. The semi‐IPN samples exhibited brittle fracture morphology which is consistent with tensile properties. 相似文献
19.
《Macromolecular chemistry and physics.》2017,218(17)
Poly[(R )‐3‐hydroxybutyrate] (P3HB) is a potential candidate for biomaterials due to its biocompatibility and biodegradability. However, P3HB needs to have tunable hydrophobicity, modification through chemical functionalization and the right hydrolytic stability to increase their potential for water‐based biomedical applications such as using them as in situ forming gels for drug delivery. This work focuses on using a copolymer, poly[(R )‐3‐hydroxybutyrate‐co‐4‐hydroxybutyrate] (P3HB4HB) in a thermogelling multiblock system with polypropylene glycol and polyethylene glycol to study the effect of the hydrophobic P3HB4HB on gelation properties, degradation, and drug release rates with reference to P3HB. Thermogels containing P3HB4HB segments show lower critical micellization concentration values in a range from 3 × 10−4 to 1.08 × 10−3 g mL−1 and lower critical gelation concentration values ranging from 2 to 6 wt% than that of gels containing P3HB. Furthermore, gels containing P3HB4HB degrade at a slower rate than the gels containing P3HB. Drug release studies of 5 µg mL−1 of doxorubicin show that gels containing P3HB4HB exhibit sustained release although the release rates are faster than gels containing P3HB. However, this can be modified by varying the concentration of the gels used. Process optimization of purifying the starting material is one important factor before the synthesis of these biomaterials.
20.
Two approaches to obtain fast‐degrading polymer films based on poly(sebacic anhydride) (PSA) are presented, both of which target polymer films with a lower degree of crystallinity than pure PSA homopolymer: first, thin films are prepared from poly(adipic anhydride)/poly(sebacic anhydride) blends at different ratios, and second, films are made from the copolymer poly(salicylic acid‐co‐sebacic acid). These films are intended as sacrificial layers for self‐regenerating functional coatings, for example, to regenerate antimicrobial surface activity. The degradation kinetics of these films are analyzed by surface plasmon resonance spectroscopy. The results of the blends approach indicate that the blend degradation rate is accelerated only in the initial degradation phase (compared to PSA). The degradation kinetics study of the poly(salicylic‐acid‐co‐sebacic acid) film shows that this copolymer degrades faster than poly(sebacic anhydride) initially, releasing antimicrobial salicylic acid in the process. However, its degradation rate slows down at a mass loss >60% and approaches the PSA degradation curve at longer degradation times. When tested as sacrificial layer in self‐regenerating antimicrobial polymer stacks, it is found that the degradation rate is too low for successful layer shedding. 相似文献