首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 296 毫秒
1.
Jun-Ting Xu  Jian Ji 《Polymer》2003,44(20):6379-6385
Crystallization and solid state structure of a poly(styrene)-graft-poly(ethylene oxide) (PS-g-PEO) graft copolymer with crystallizable side chains were studied using simultaneous small angle X-ray scattering/wide angle X-ray scattering/differential scanning calorimetry (SAXS/WAXS/DSC). It is found that the glass transition temperature (Tg) of PS main chain is remarkably higher than that of PS homopolymer. The start cooling temperature (To) has a great influence on crystallization of the PEO side-chain. When the graft copolymer is cooled from the temperature above Tg, phase separation is suppressed due to the low mobility of the PS main chain and the homogeneous melt is vitrified. The unfavorable conformation of the rigid main chain results in a single crystallization peak and lower crystallinity. When PS-g-PEO is only heated to a temperature lower than the Tg and then cooled, phase separation is retained. Both the PEO side chains with high and low crystallizability can crystallize in the phase-separated state, leading to double crystallization peaks and higher crystallinity. The effect of solvent on crystallization of the graft copolymer was also examined. It is observed that addition of toluene reduces the Tg of the PS main chain and leads to the disappearance of the vitrification effect.  相似文献   

2.
The crystallization behavior of poly(ε-caprolactone) (PCL) blocks starting from a solid lamellar morphology formed in advance by the crystallization of polyethylene (PE) blocks (PE lamellar morphology) in a PCL-b-PE diblock copolymer was investigated by differential scanning calorimetry (DSC), small-angle X-ray scattering with synchrotron radiation (SR-SAXS), and polarized optical microscope (POM). The crystallization behavior was quantitatively compared with that of a PCL-block-polybutadiene copolymer, where the crystallization of PCL blocks started from a rubbery lamellar microdomain. DSC and SR-SAXS results revealed that the crystallization rate of PCL blocks in PCL-b-PE increased drastically with decreasing crystallization temperature Tc and the Avrami exponent depended significantly on Tc. SR-SAXS curves during the crystallization of PCL blocks at high Tc showed a bimodal scattering character, that is, the peak position moved discontinuously with crystallization time. At low Tc, on the other hand, no shift of the SAXS peak position was observed. The macroscopic change in morphology was detected only at high Tc by POM observations. These experimental results for the crystallization behavior of PCL blocks in PCL-b-PE all support our previous conclusions obtained by static measurements; the crystallization mechanism at low Tc is completely different from that at high Tc, that is, the PCL blocks crystallize within the PE lamellar morphology at low Tc while the crystallization of PCL blocks at high Tc yields a morphological transition from the PE lamellar morphology into a new solid morphology.  相似文献   

3.
Sixun Zheng  Yongli Mi 《Polymer》2003,44(4):1067-1074
The blends of poly(hydroxyether of bisphenol A) (phenoxy) with poly(4-vinyl pyridine) (P4VPy) were investigated by differential scanning calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR) and high-resolution solid-state nuclear magnetic resonance (NMR) spectroscopy. The single, composition-dependent glass transition temperature (Tg) was observed for each blend, indicating that the system is completely miscible. The sigmoid Tg-composition relationship is characteristic of the presence of the strong intermolecular specific interactions in the blend system. FTIR studies revealed that there was intermolecular hydrogen bonding in the blends and the intermolecular hydrogen bonding between the pendant hydroxyl groups of phenoxy and nitrogen atoms of pyridine ring is much stronger than that of self-association in phenoxy. To examine the miscibility of the system at the molecular level, the high resolution 13C cross-polarization (CP)/magic angle spinning (MAS) together with the high-power dipolar decoupling (DD) NMR technique was employed. Upon adding P4VPy to the system, the chemical shift of the hydroxyl-substituted methylene carbon resonance of phenoxy was observed to shift downfield in the 13C CP/MAS spectra. The proton spin-lattice relaxation time T1(H) and the proton spin-lattice relaxation time in the rotating frame T(H) were measured as a function of the blend composition. In light of the proton spin-lattice relaxation parameters, it is concluded that the phenoxy and P4VPy chains are intimately mixed on the scale of 20-30 Å.  相似文献   

4.
Summary In this work are examined changes of the relaxation spectrum of PMA during the structural relaxation process which follows a quenching from a temperature To above Tg to another T1 below Tg, which is mantained constant. It is found that the relaxation times spectrum is shifted towards longer times as the ageing time increases. At the same time, its form becomes more and more wide, approaching the form of the relaxation times spectrum calculated at temperatures above Tg from alternative dielectric measurements. A parameter measuring the shift of the spectrum, a(ta), is defined and its dependence on ageing time and annealig temperature is studied. The effective relaxation time of the structural relaxation process is calculated from a(ta) and it is found dependent on the temperature of annealing as well as on the value of a(ta)itself.  相似文献   

5.
The miscibility and crystallization behavior of binary crystalline blends of poly(butylene terephthalate) [PBT] and polyarylate based on Bisphenol A and a 27/73 mole ratio of isophthalic and terephthalic acids [PAr(I27-T73)] have been investigated by differential scanning calorimetry (DSC). This blend system exhibits a single composition-dependent glass transition temperature over the entire composition range. The equilibrium melting point depression of PBT was observed, and Flory interaction parameter χ12 = −0.96 was obtained. These indicate that the blends are thermodynamically miscible in the melt. The crystallization rate of PBT decreased as the amount of PAr(I27-T73) increased, and a contrary trend was found when PAr(I27-T73) crystallized with the increase of the amount of PBT. The addition of high-Tg PAr(I27-T73) would suppress the segmental mobility of PBT, while low-Tg PBT would have promotional effect on PAr(I27-T73). The crystallization rate and melting point of PBT were significantly influenced when the PAr(I27-T73) crystallites are previously formed. It is because not only does the amorphous phase composition shift to a richer PBT content after the crystallization of PAr(I27-T73), but also the PAr(I27-T73) crystal phase would constrain the crystallization of PBT. Thus, effects of the glass transition temperature, interaction between components, and previously formed crystallites of one component on the crystallization behavior of the other component were discussed and compared with blends of PBT and PAr(I-100) based on Bisphenol A and isophthalic acid.  相似文献   

6.
Poly(lactide) (PLA) is rapidly gaining interest as a biodegradable thermoplastic for general usage in degradable disposables. To improve mechanical properties, a PLA with low stereoregularity was blended with polyethylene glycol (PEG). Blends with up to 30 wt% PEG were miscible at ambient temperature. Blending with PEG significantly decreased the Tg, decreased the modulus and increased the fracture strain of PLA. However, the PLA/PEG 70/30 blend became increasingly rigid over time at ambient conditions. The mechanism of aging primarily under ambient conditions of temperature and humidity was studied. Changes in mechanical properties, thermal transitions and solid state morphology were examined over time. Aging was caused by slow crystallization of PEG. Crystallization of PEG depleted the amorphous phase of PEG and gradually increased the Tg. As Tg approached the aging temperature, reduced molecular diffusivity slowed the crystallization rate dramatically. Aging essentially ceased when Tg of the amorphous phase reached the aging temperature. The increase in matrix Tg and the reinforcing effect of the crystals produced a change in mechanical properties from elastomer-like to thermoplastic-like.  相似文献   

7.
The thin films of a symmetric crystalline-coil diblock copolymer of poly(l-lactic acid) and polystyrene (PLLA-b-PS) formed lamellae parallel to the substrate surface in melt. When annealed at temperatures well above the glass transition temperature of PLLA block (TgPLLA), the PLLA chains started to crystallize, leading to reorientation of lamellae. Such reorientation behavior exhibited dependence on the correlation between the crystallization temperature (Tc), the glass transition temperature of PS (TgPS), the peak melting point of PLLA crystals (TmPLLA), and the end melting point of PLLA crystals (Tm,endPLLA). When annealed at (Tc=) 80 °C (Tc < TgPS < TODT, order-disorder transition temperature), 123 °C (TgPS < Tc < TmPLLA < TODT), 165 °C (TgPS < TmPLLA < Tc < Tm,endPLLA < TODT), the parallel lamellae became perpendicular to the substrate surface, exclusively starting at the edge of surface relief patterns. Meanwhile, the corresponding lamellar spacing was significantly enhanced. The PLLA crystallization between PS layers was hypothesized to account for the lamella reorientation during annealing. The crystallization, chain conformation, and possible chain folding mechanisms were discussed, based on detailed analysis of the lamellar structure before and after crystallization.  相似文献   

8.
Structure and properties of a bioabsorbable poly(glycolide-co-lactide) (PGA-co-PLA) fiber during several processing stages and the final in vitro degradation stage were investigated by means of wide-angle X-ray diffraction, dynamic mechanical analysis and mechanical property tests. In the orientation stage, an increase in the temperature of the first encountered orientation roll resulted in a lower level of crystallinity and larger crystallites. The temperature of the second encountered pre-annealing roll (PR) imposed a smaller effect on the structure. In the hot-stretching stage after fibers were braided, the maximum crystallinity was achieved at around 126 °C. Higher hot-stretching temperatures increased the crystal size, glass transition temperature (Tg) and tensile strength, but decreased the elongation at break and the heat shrinkage near Tg. In the post-annealing stage, it was found that crystallinity, Tg and tensile strength all increased significantly while the heat shrinkage near Tg sharply decreased after annealing. This suggests that the internal stress accumulated in the orientation and hot-stretching stages can be effectively reduced by post-annealing. During in vitro degradation, crystallinity was found to increase with time while the heat shrinkage near Tg and in the supercooling region (Tg<T<Tm) was greatly reduced. These results support the process of cleavage-induced crystallization.  相似文献   

9.
The influence of the solvent evaporation rate on the crystallization of the poly(vinylcyclohexane)-b-poly(ethylene)-b-poly(vinylcyclohexane) (PVCH-PE-PVCH) triblock copolymer with the high Tg of PVCH segment in chloroform was investigated. The competition between the crystallization of PE block and the vitrification of PVCH in the triblock copolymer was controlled through changing the solvent evaporation rate in the solution system at different temperatures (Te). It was found that the melting temperature (Tm) of PVCH-PE-PVCH samples increased with increasing the Te when the Te was lower than the solvent boiling point (bp), depending on the crystalline temperatures. However, when Te was just a little above the solvent bp, two melting peaks, which corresponding to the fusion of the confined and unconfined crystals, respectively, were observed on the DSC curves of the samples. As Te increased to be higher temperature, only one lower melting peak, which corresponding to the fusion of the confined crystals, existed for each samples. It was also found that the crystallinity (Xc) of the samples decreased gradually when Te was lower than the solvent bp, and then decreased suddenly when Te was just a little above the solvent bp, finally reached a plateau (about 13.5%) at higher Te. The changes in Xc of the samples depend on the evaporating time in the solution. The results should be related to the competition between the crystallization of PE block in solvent and the vitrification of PVCH block with the solvent evaporation. Furthermore, the competition was controlled through changing the solvent evaporation rate. The confined and unconfined crystallization of the samples could be freely adjusted.  相似文献   

10.
Phthalates pose adverse health effects due to their propensity to leach and the most common, di(2-ethylhexyl) phthalate (DEHP) and di-n-octyl phthalate (DOP), are petroleum-based. Conversely, di-esters, succinates are biobased (produced from fermentation of biomass), biodegradable, and therefore potential sustainable replacements for phthalates. A series of succinates, di-octyl succinate (DOS), di-hexyl succinate (DHS), di-butyl succinate (DBS), and di-ethyl succinate (DES), were mixed with poly(vinyl chloride) (PVC). The interaction of the plasticizer ester carbonyl with PVC shows an average −5 cm−1 shift of the carbonyl absorbance peak energy. The glass transition temperatures (T g), were monitored by differential scanning calorimetry and dynamic mechanical analyses. The T gs of DOS and DHS plasticized PVC were significantly lower than DOP plasticized PVC at a lower percent mass. On the other hand, PVC plasticized with either DBS or DES exhibited a similar trend in lowering the T g as that of DOP plasticized PVC.  相似文献   

11.
Mohammad K. Hassan 《Polymer》2007,48(7):2022-2029
Broadband dielectric spectroscopy was used to examine carboxylic acid-terminated poly(d,l-lactide) samples that were hydrolytically degraded in 7.4 pH phosphate buffer solutions at 37 °C. The dielectric spectral signatures of degraded samples were considerably more distinct than those of undegraded samples and a Tg-related relaxation associated with long range chain segmental mobility was seen. For both degraded and undegraded samples, a relaxation peak just beneath a DSC-based Tg was observed, which shifts to higher frequency with increasing temperature. Thus, this feature is assigned as the glass transition as viewed from the dielectric relaxation perspective. Linear segments on log-log plots of loss permittivity vs. frequency, in the low frequency regime, are attributed to d.c. conductivity. An upward shift in relaxation peak maximum, fmax, observed especially after 145 d of immersion in buffer, implies a decrease in the time scale of long range segmental motions with increased degradation time.Permittivity data for degraded and undegraded materials were fitted to the Havriliak-Negami equation with subtraction of the d.c. conductivity contribution to uncover pure relaxation peaks. Parameters extracted from these fits were used to construct Vogel-Fulcher-Tammann-Hesse (VFTH) curves and distribution of relaxation time, G(τ), curves for all samples. It was seen that the relaxation times for the α-transition in both degraded and undegraded samples showed VFTH temperature behavior. G(τ) curves showed a general broadening and shift to lower τ with degradation, which can be explained in terms of a broadening of molecular weight within degraded samples and faster chain motions.  相似文献   

12.
A star polymer with a γ-CD core and PS arms (CD-star) is used to partially compatibilize blends of the immiscible polymers polystyrene (PS) and poly(dimethylsiloxane) (PDMS). The mechanism of compatibilization is threading of the CD core by PDMS and subsequent solubilization in the PS matrix facilitated by the star arms. Films cast from clear solutions in chloroform exhibit large wispy PDMS domains, indicating that some dethreading of CD-star and agglomeration of PDMS takes place during the slow process of solvent evaporation. However, DSC and DMA data show that partial compatibilization takes place, as evidenced by a shift in the PS and PDMS Tgs toward each other. The shift in PS Tg is greater when CD-star is present compared to samples without CD-star. PDMS also tends to leach out of the solution-cast films during solvent evaporation and post-processing of the films. The amount of retained PDMS is significantly increased when CD-star is present. The DMA data also show that PDMS has a lower molecular mobility when CD-star is present.  相似文献   

13.
syndiotactic Polystyrene (sPS) glass crystallizes into the α form when it is heated above the glass transition temperature (Tg, about 100 °C). sPS can be crystallized also into the δ form in the solvent atmosphere at room temperature. In order to trace the structural evolution process, the time-resolved infrared spectral measurements have been performed in the isothermal crystallization from the glass to α form and in the solvent-induced crystallization from the glass to δ form at the various temperatures. Absorbance of crystallization-sensitive infrared bands was plotted against time, from which the crystallization kinetics were analyzed on the basis of Avrami equation: X(t)=1−exp[−(kt)n] where X is a normalized crystallinity, n is an index, k is a rate constant, and t is a time. The isothermal crystallization was investigated also by carrying out the temperature jump experiment of DSC thermograms, giving almost the same results as the infrared spectral measurements. The Avrami index n was 2-5 depending on the crystallization temperature (Tc). The k was also dependent on the Tc, about 10−1-10−4 s−1 and could be fitted reasonably by the equation of crystallization kinetics. An extrapolation of the k vs Tc plot to the negligibly small k value allowed us to predict the temperature at which no crystallization should occur, ca. 100 °C, in good agreement with the observed Tg value. On the other hand, the solvent-induced crystallization was investigated for the first time at the various temperatures from 50 to 9 °C by the time-resolved measurement of infrared spectra. Compared with the experiment at room temperature, the crystallization was highly accelerated at 40-50 °C, while the crystallization rate was reduced remarkably at such a low temperature as 9 °C. The time dependence of infrared absorbance was analyzed for the crystallization-sensitive bands on the basis of Avrami equation as the first approximation, although the crystallization mechanism was more complicated than the isothermal crystallization case. The logarithm of the k value was found to change almost linearly with temperature and an extrapolation to infinitesimally small k value gave a Tg of about −15 °C. That is to say, the glass transition temperature was estimated to shift remarkably from 100 to −15 °C by absorbing solvent molecules or by a plasticizing effect.  相似文献   

14.
Several papers have demonstrated that structural changes in the amorphous regions of semicrystalline polymers can be produced by heat treatment below the glass transition temperature (Tg). In this paper, we report structural change in the amorphous phase of poly(ethylene terephthalate), heat-treated below and above Tg. The density, the Tg, the endothermic peak at Tg and the relative spectral intensity in the 973 cm?1 band (due to the CO stretching vibration), all increased with heat treatment below Tg, but the specific heat decreased. The stability of the amorphous state was examined by further heat treatment at temperatures above Tg and sufficiently high for crystallization, and it was verified that structural changes in the amorphous regions do not result in acceleration of the rate of crystallization. We therefore suggest that the amorphous region is one phase, rather than two phases consisting of random and regular regions.  相似文献   

15.
A unique rapid scanning rate differential scanning calorimeter is used to examine the differences in melt and cold crystallized poly (l-Lactide) (PLLA), a biodegradable semi-crystalline polymer. After isothermal melt and cold crystallization at various temperatures, both melt and cold crystallized PLLA are characterized by similar melting temperatures (Tm) and exhibit multiple melting behavior on heating at 500 °C/min. However, cold crystallization results in a higher degree of crystallinity (wc) compared to melt crystallization. While the overall amorphous fraction is higher for melt crystallization, the mobile amorphous fraction (wa) is found to be higher for cold crystallization. The rigid amorphous fraction (wraf) in PLLA is determined to be higher for melt crystallization than for cold crystallization at almost all temperatures. The higher values of wraf also appear to result in higher values of the glass transition temperature (Tg) for melt crystallized samples due to a reduction in mobility of amorphous phase. These dramatic differences depending on whether the material is brought to the crystallization temperature from the melt or the glassy state, could have profound implications for processing and optimizing the properties of PLLA.  相似文献   

16.
A series of poly(di-n-alkyl itaconic acid esters) with side chain lengths from 7 to 20 carbon atoms have been prepared. For derivatives with chain lengths of 7 to 11 carbons, two glass transition temperatures have been detected. The transition occurring at the lower temperature T2g originates in the side groups and is a result of the independent cooperative relaxation of the alkyl side chains, while that observed at higher temperatures T1g reflects the glass-rubber transition of the main chain backbone and the cooperative motion of the total molecule. Derivatives with chain lengths ?12 carbon atoms display melting temperatures Tm, and measurement of the enthalpy of fusion ΔHf indicates that side chain crystallization takes place with part of each side chain entering into a regularly packed hexagonal lattice.  相似文献   

17.
Longitudinal relaxation times in the laboratory (T1) and rotating (T1ρ) frames have been measured for poly(N-amyl maleimide) and poly(N-dodecyl maleimide). The results are compared with information previously published using dielectric relaxation techniques. A total of three relaxation processes have been detected corresponding to methyl rotation, motion within the substituent alkyl chain and out-of-plane deformation of the maleimide ring. An attempt to fit the T1 data in the dodecyl derivative to theoretical equations is described.  相似文献   

18.
Blending poly(ethylene glycol) (PEG) with poly(lactide) (PLA) decreases the Tg and improves the mechanical properties. The blends have lower modulus and increased fracture strain compared to PLA. However, the blends become increasingly rigid over time at ambient conditions. Previously, it was demonstrated that a PLA of lower stereoregularity was miscible with up to 30 wt% PEG. Aging was due to slow crystallization of PEG from the homogeneous amorphous blend. Crystallization of PEG depleted the amorphous phase of PEG and gradually increased the Tg until aging essentially ceased when Tg of the amorphous phase reached the aging temperature. In the present study, this aging mechanism was tested with a crystallizable PLA of higher stereoregularity. Changes in thermal transitions, solid state structure, and mechanical properties were examined over time. Blends with up to 20 wt% PEG were miscible. Blends with 30 wt% PEG could be quenched from the melt to the homogenous amorphous glass. However, this composition phase separated at ambient temperature with little or no crystallization. Changes in mechanical properties during phase separation reflected increasing rigidity of the continuous PLA-rich phase as it became richer in PLA. Construction of a phase diagram for blends of higher stereoregular PLA with PEG was attempted.  相似文献   

19.
BACKGROUND: The thermal behaviour of poly(lactic acid) (PLA) in contact with compressed CO2 was studied using high‐pressure differential scanning calorimetry. In particular, the effect of annealing below and above the glass transition temperature (Tg) on the glass transition, cold crystallization and melting temperatures was studied systematically as a function of annealing time and pressure. RESULTS: The effect of compressed CO2 on the thermal properties of PLA is time dependent. Annealing below Tg decreases the temperature and enthalpy of cold crystallization. Similar, but more evident, behaviours are observed when annealing above Tg. Crystallization temperature and enthalpy during cooling decrease with increasing pressure, and the peak is narrower. CONCLUSION: Annealing PLA in the presence of compressed CO2 accelerates cold crystallization, but retards crystallization during cooling. Copyright © 2009 Society of Chemical Industry  相似文献   

20.
Amorphous poly(ethylene terephthalate) film was uniaxially drawn over a wide range of temperatures from below to above the Tg at a constant strain rate. The geometry of the deformation in macroscopic dimensions of the sample demonstrates that homogeneous deformation can be obtained when the drawing temperature (Tdef) is not lower than 69°C. The change of the cold crystallization peak temperature (Tcc) and crystallinity determined by differential scanning calorimetry and density measurement, respectively, were studied in terms of the Tdef and the draw ratio (λ). The orientation, relaxation, and crystallization during drawing were investigated as a function of Tdef as well as of λ. The results suggest that 69°C is the critical temperature at which the sample with the highest orientation and the least slippage of the molecular chain and without obvious crystallization can be obtained. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 77: 2044–2048, 2000  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-23

京公网安备 11010802026262号