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 0 毫秒
1.
John D. Hoffman 《Polymer》1982,23(5):656-670
The theory of polymer crystallization with chain folding is extended to include the effect of reptation in the melt on the rates of crystallization GI and GII in régimes I and II. The result is that the pre-exponential factors for GI and GII contain a factor 1n, Where n is the number of monomer units in the pendant chain being reeled onto the substrate by the force of crystallization; n is proportional to the molecular weight. The predicted fall in growth rate with increasing molecular weight is found experimentally in nine polyethylene fractions Mz=2.65 × 104 to Mz=2.04 × 105, corresponding to nz=1.90 × 103 to 1.45 × 104. The data on these fractions are analysed to find the reptation or ‘reeling’ rate r and the substrate completion rate g. The values gnuc~0.5/nz cm s?1 and rnuc~21/nz cm s?1 at 400K are obtained from the data in conjunction with nucleation theory adapted to account for reptation assuming a substantial degree of regular folding. These results are consistent with a melting point in the range of ~142° to ~145°C. (The analysis using T°m(∞)=145°C gives values of such quantities as σ σe and α that are quite similar to those deduced in earlier studies.) An estimate of g (denoted gexpt) that is independent of the molecular details of nucleation theory gives gexpt~0.4/nz cm s?1 and r~17/nz cm s?1 at 400K. Calculations of the reptation rate from r1,2 = (force of crystallization ÷ friction coefficient for reptation in melt), where the friction coefficient is determined from diffusion data on polyethylene melts, leads to r1,2~17/nz to 34/nz cm s?1 at at 400K, or g1,2~0.4/nz to 0.8/nz cm s?1. The conclusion is that the reptation rate characteristic of the melt is fast enough to allow a significant degree of adjacent re-entry or ‘regular’ folding during substrate completion at the temperature cited, and that the substrate completion process is governed jointly by the activation energy for reptation Q1D and the work of chain folding q. The nucleation theory and the friction coefficient theory approaches are compared, and the formulations found to be essentially equivalent; the ‘reeling’ rate r is found to be proportional to (1n)A0(Δf)v0exp[?(Q1D+q)RT], where v0 is a frequency factor, and A0(Δf) is the force of crystallization on the pendant chain. The data analysis on the fractions confirms the detailed applicability of régime theory. The growth rate theory presented allows the possibility that the growth front may be microfaceted in régime I.  相似文献   

2.
W.D. Reents  M.L. Kaplan 《Polymer》1982,23(2):310-313
Low molecular weight and high molecular weight samples of poly(p-phenylene sulphide) were extracted exhaustively with methylene chloride. The extracts were examined by h.p.l.c. and by solid probe mass spectroscopy. Four families of compounds have been identified in these extracts: (1) cyclic (p-phenylene sulphide) oligomers; (2) chlorine terminated linear (p-phenylene sulphide); (3) single chlorine terminated linear (p-phenylene sulphide); and (4) linear (p-phenylene sulphides). The fully extracted polymer did not exhibit any mass spectrum. In addition to the compounds specified above, dibenzothiophene was found in all samples while phenyl biphenyl ether was found only in the extract of the high molecular weight material.  相似文献   

3.
The equilibrium thermodynamic properties and crystallization behavior of poly(p-phenylene succinate) (PPSc) were examined using differential scanning calorimetry (DSC) and a wide-angle X-ray diffractometer. When PPSc quenched into liquid nitrogen, the PPSc still showed 60 J/g of the heat of fusion, indicating that PPSc crystallizes too fast to be quenched to a glassy state. The equilibrium heat of fusion (ΔH°) of PPSc and its equilibrium entropy of fusion (ΔS°) were estimated to be 28 and 48 J/mol K, respectively. The effect of the molecular weight and undercooling on PPSc crystallization was analyzed by the Avrami expression. The crystallization rate of PPSc was highest at about 276°C, regardless of molecular weight when the sample was isothermally crystallized. This temperature was higher than that of poly(ethylene terephthalate) (PET) by 100°C. The crystallization rate was highest when the number average molecular weight of PPSc was about 20,000 g/mol. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 801–806, 1999  相似文献   

4.
The spherulitic morphology of pure poly(L ‐lactide) (PLLA) was investigated with polarized optical microscopy as a function of the crystallization temperature and molecular weight. After being melted at 210°C for 3 min, samples were cooled quickly to designated temperatures for isothermal crystallization. It was shown for the first time that a clear banding‐to‐nonbanding morphological transition took place at a critical temperature for PLLA with a number‐average molecular weight of 86,000. With the increasing molecular weight of the material, the spherulite growth rates decreased notably, and the band spacing decreased significantly. On the basis of the main‐chain chirality in PLLA and the observation of a nonbanded spherulitic morphology in a certain temperature region, it was suggested that the crystallization temperature might have an effect on the relationship between the sense of lamellar twisting and the main‐chain chiral structure in PLLA. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007  相似文献   

5.
The spherulitic growth rates of a series poly (?‐caprolactone) homopolymers and poly(?‐caprolactone)‐b‐ poly(ethylene glycol) (PCL‐b‐PEG) block copolymers with different molecular weights but narrow polydispersity were studied. The results show that for both PCL homopolymers and PCL‐b‐PEG block copolymers, the spherulitic growth rate first increases with molecular weight and reaches a maximum, then decreases as molecular weight increases. Crystallization temperature has greater influence on the spherulitic growth rate of polymers with higher molecular weight. Hoffman–Lauritzen theory was used to analyze spherulitic growth kinetics and the free energy of the folding surface (σe) was derived. It is found that the values of σe decrease with molecular weight at low molecular weight level and become constant for high molecular weight polymers. The chemically linked PEG block does not change the values of σe significantly. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

6.
Copolymers of ethylene glycol with 4,4′‐bibenzoic acid and terephthalic acid are known to crystallize rapidly to surprisingly high levels of crystallinity. To understand this unusual behavior, the isothermal crystallization of poly(ethylene bibenzoate‐co‐terephthalate) in the molar ratio 55:45 (PETBB55) was studied. Poly(ethylene naphthalate) (PEN) was included in the study for comparison. The kinetics of isothermal crystallization from the melt and from the amorphous glass was determined using differential thermal analysis. The results were correlated with the crystalline morphology as observed with atomic force microscopy (AFM). Crystallization of PEN exhibited similar kinetics and spherulitic morphology regardless of whether it was cooled from the melt or heated from the glass to the crystallization temperature. The Avrami coefficient was close to 3 for heterogeneous nucleation with 3‐dimensional crystal growth. The copolymer PETBB55 crystallized much faster than did PEN and demonstrated different crystallization habits from the melt and from the glass. From the melt, PETBB55 crystallized in the “normal” way with spherulitic growth and an Avrami coefficient of 3. However, crystallization from the glass produced a granular crystalline morphology with an Avrami coefficient of 2. A quasi‐ordered melt state, close to liquid crystalline but lacking the order of a recognizable mesophase, was proposed to explain the unusual crystallization characteristics of PETBB55. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 98–115, 2002  相似文献   

7.
The crystal morphology and nonisothermal crystallization kinetics of short carbon fiber/poly(trimethylene terephthalate) (SCF/PTT) composites were investigated by polarized optical microscopy (POM) and differential scanning calorimetry (DSC). The optical micrographs suggest that the more content of SCF in composites, the smaller size of the spherulites is. Moreover, the addition of SCF can lead to forming banded spherulites in composites. The Avrami equation modified by Jeziorny, Ozawa theory and the method developed by Mo were used, respectively, to fit the primary stage of nonisothermal crystallization of various composites. The results suggest that the SCF served as nucleation agent, accelerates the crystallization rate of the composites, and the more content of SCF, the faster crystallization rate is. Effective activation energy calculated by the differential iso‐conversional method developed by Friedman also concludes that the composite with more SCF component has higher crystallization ability than that with less SCF content. The kinetic parameters U* and Kg are determined, respectively, by the Hoffman–Lauritzen theory. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

8.
He Huang  Lixia Gu 《Polymer》2006,47(11):3935-3945
A study has been made of the non-isothermal crystallization behavior and thermal transitions of a biodegradable, partially hydrolyzed poly(vinyl alcohol) with 80% degree of saponification (PVA80). Possible sample degradation was first investigated, but no significant degradation or dehydration was detected using FTIR and DSC under the experimental condition. The non-isothermal crystallization of PVA80 was analyzed with Ozawa equation, and the Mo method of combining Ozawa and Avrami equations. Ozawa equation was only applicable in a narrow temperature range from 80 to 100 °C. The deviation from the Ozawa equation is not due to the secondary crystallization or the quasi-isothermal nature of the treatment. It is only a result of the large relative difference of the relative crystallinity values under different cooling rates. The Mo method demonstrated a success in the full temperature range investigated. The isoconversional method developed by Friedman failed to estimate the activation energy for this non-isothermal crystallization. Thermal transitions of PVA80 are associated with its complex hydrogen-bonding interactions. The melt-crystallized PVA80 sample, as that from film casting, followed by annealing at 60 and 80 °C, has a broad melting temperature range measured by DSC and FTIR. It was found that the melting behavior of a semicrystalline polymer can be probed via a non-crystalline hydrogen-bonded CO band using FTIR. The glass transition temperature Tg of PVA80 was raised about 20 °C, after the sample was melt-crystallized. The intensity of the hydrogen-bonded CO band increases when temperature was increased from 110 to 180 °C, due to the promoted hydrogen-bonding interactions between the CO groups in the amorphous phase and the hydroxyl groups from the crystalline phase, which is also the main reason for the increased Tg transition.  相似文献   

9.
M. Kojima  J.H. Magill   《Polymer》1985,26(13):1971-1978
Spherulite formation in poly(bis(trifluoroethoxy)phosphazene) has been investigated from the melt and from solution over a relatively wide crystallization and annealing range. Three polymorphic forms and one mesoform is found in this polymer. Several of these crystal modifications coexist in amounts which depend upon the crystallization conditions. Negatively birefringent spherulites increase in birefringence slightly upon being heated through the thermotropic T(1) transition. This change involves the formation of a chain extended morphology from a chain folded one. Microbeam X-ray analysis made within the spherulite shows that the unit cell [a] direction is along the spherulite radius while the [c] chain direction lies transverse to the spherulite radius. Moreover, below T(1) it has been established that the X-ray long period is invariant with annealing time and temperature, and above T(1) the periodicity disappears, or cannot be recorded. Whenever heating or cooling occurs through T(1), the spherulite birefringence appears to be invariant after initially heating through T(1). However, a substantial volume change (by dilatomery) of 6% occurs through T(1) and this is consistent with a change in crystal structure from a 3D orthorhombic structure below, to a 2D hexagonal form above T(1). From the molten state the transformation occurs rapidly from isotropic to the 2D form of the mesostate, not far below Tm. The 2D hexagonal form reverts to 3D chain extended orthorhombic on cooling below T(1). This stable chain extended morphology also arises whenever PBFP is melted and then cooled below T(1). All transformations from the isotropic melt or the folded chain conformation must pass through the mesophase. All specimens are friable upon cooling below T(1), whereas the solution case spherulitic polymer film is ductile here.  相似文献   

10.
The crystallization kinetics of binary blends of poly(ethylene oxide) and poly(methyl methacrylate) were investigated. The isothermal spherulitic growth rates were measured by means of a polarized light microscope. The temperature and composition dependence on the growth rates have been analysed. The temperature range studied was from 44° to 58°C. The introduction of poly(methyl methacrylate) into poly(ethylene oxide) resulted in a reduction of the spherulitic growth rate as the proportion of poly(methyl methacrylate) was increased from zero to 40% by weight. Results have been analysed using the theoretical equations of Boon and Azcue for the growth rate of polymer-diluent mixtures. The experimental results are in good agreement with this equation. The temperature coefficient is negative as is the case in the crystallization of bulk homopolymers.  相似文献   

11.
A polyacetal (POM)/poly(ε‐caprolactone) (PCL) reactive blend prepared via a chain‐transfer reaction was investigated with respect to its morphology and nonisothermal crystallization, and the results were compared with those of a simple POM/PCL blend. The reactive blend had a microscopically phase‐separated morphology in which the diameter of the PCL microphase was below 100 nm, and it clearly yielded ring‐banded spherulites, whereas between the two blends, there were no significant differences in the diameters and polygonal edges of the spherulites and in the long period of the POM phases. The PCL part of the reactive blend crystallized within the confined microspace with about 10% lower crystallinity than that of the corresponding simple blend. A lower Avrami exponent and crystallization rate parameter of the PCL part were observed in the primary crystallization process of the reactive blend. In contrast, the crystallinity of the POM component and the nonisothermal crystallization kinetic parameters of the POM part showed no noticeable differences between the two blends at any given cooling rate. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

12.
The nonisothermal crystallization kinetics of poly(ethylene terephthalate) (PET) and poly(methyl methacrylate) (PMMA) blends were studied. Four compositions of the blends [PET 25/PMMA 75, PET 50/PMMA 50, PET 75/PMMA 25, and PET 90/PMMA 10 (w/w)] were melt‐blended for 1 h in a batch reactor at 275°C. Crystallization peaks of virgin PET and the four blends were obtained at cooling rates of 1°C, 2.5°C, 5°C, 10°C, 20°C, and 30°C/min, using a differential scanning calorimeter (DSC). A modified Avrami equation was used to analyze the nonisothermal data obtained. The Avrami parameters n, which denotes the nature of the crystal growth, and Zt, which represents the rate of crystallization, were evaluated for the four blends. The crystallization half‐life (t½) and maximum crystallization (tmax) times also were evaluated. The four blends and virgin polymers were characterized using a thermogravimetric analyzer (TGA), a wide‐angle X‐ray diffraction unit (WAXD), and a scanning electron microscope (SEM). © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3565–3571, 2006  相似文献   

13.
Poly(ethylene terephthalate) was submitted to five reprocessing cycles by extrusion. The materials were analyzed with oligomer and after oligomer extraction. The nonisothermal crystallization of the five samples was investigated by differential scanning calorimetry. Samples with oligomer content and carboxylic end group concentrations between 44 and 98 eqw × 106 g presented a nonlinear correlation with the crystallization temperature. After the oligomer extraction of the polymer, this correlation is linear. The nonisothermal crystallization results were analyzed using the Ozawa model. The polymers containing oligomers obey the Ozawa model for the first reprocessing cycle. After oligomer extraction, the polymers obey the Ozawa model from the first to the third reprocessing cycle. In both cases, the exponential n values are close to 2.0. For the other cycles, deviations from this model occur. The activation energy was calculated using the Kissinger and Varma models. The values obtained for the five reprocessed samples were inversely proportional to the molar mass when analyzed by both models. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 91: 525–531, 2004  相似文献   

14.
The melting behavior and crystallization kinetics of poly(trimethylene terephthalate) (PTT)/stainless steel fiber (SSF) composites were investigated with differential scanning calorimetry. The morphology was studied with scanning electron microscopy and polarized optical microscopy. Differential scanning calorimetry analysis revealed that the crystallization temperature increased by 27°C with the addition of 1 vol % SSF to the matrix. The Avrami exponents, analyzed in isothermal crystallization kinetics, were determined to be 2–3 for both neat PTT and PTT/SSF composites. SSF, as a nucleating agent in the composites, greatly increased the crystallization rate. The activation energies of the composites were obviously lower than that of pure PTT, and this indicated much easier crystallization of the composites. All these samples exhibited banded spherulites, and the spherulite size gradually decreased with the SSF loading increasing. Subsequent melting behaviors revealed that all of these samples, especially of the composites, exhibited triple melting peaks at all crystallization temperatures studied. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

15.
The miscibility and crystallization kinetics of the blends of random poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) [P(HB‐co‐HV)] copolymer and poly(methyl methacrylate) (PMMA) were investigated by differential scanning calorimetry (DSC) and polarized optical microscopy (POM). It was found that P(HB‐co‐HV)/PMMA blends were miscible in the melt. Thus the single glass‐transition temperature (Tg) of the blends within the whole composition range suggests that P(HB‐co‐HV) and PMMA were totally miscible for the miscible blends. The equilibrium melting point (T°m) of P(HB‐co‐HV) in the P(HB‐co‐HV)/PMMA blends decreased with increasing PMMA. The T°m depression supports the miscibility of the blends. With respect to the results of crystallization kinetics, it was found that both the spherulitic growth rate and the overall crystallization rate decreased with the addition of PMMA. The kinetics retardation was attributed to the decrease in P(HB‐co‐HV) molecular mobility and dilution of P(HB‐co‐HV) concentration resulting from the addition of PMMA, which has a higher Tg. According to secondary nucleation theory, the kinetics of spherulitic crystallization of P(HB‐co‐HV) in the blends was analyzed in the studied temperature range. The crystallizations of P(HB‐co‐HV) in P(HB‐co‐HV)/PMMA blends were assigned to n = 4, regime III growth process. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 91: 3595–3603, 2004  相似文献   

16.
The thermal behavior and spherulitic morphologies of poly(L ‐lactic acid) (PLLA)/poly(D ‐lactic acid) (PDLA) 1/1 blend with weight‐molecular‐weight of 105 order, together with those of pure PLLA and PDLA, were investigated using differential scanning calorimetry and polarized optical microscopy. It was found that in the blend, stereocomplex crystallites could be formed exclusively or coexisted with homocrystallites depending on thermal history. Banded to nonbanded spherulitic morphological transition occurred for melt‐crystallized PLLA and PDLA, while the blend presented exclusively nonbanded spherulitic morphologies in the temperature range investigated. The spherulite growth of the blend occurred within a wider temperature range (≤180°C) compared with that of homopolymers (≤150°C), while the spherulite growth rates were comparable for both the blend and homopolymers. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

17.
用非等温结晶动力学研究了聚乳酸(PLA)的结晶行为,利用Hoffman-Weeks外推法以及Baur等提出的方程反推出PLA中的右旋组分摩尔分数为7.2%.较高的右旋组分摩尔分数是PLA结晶过程中成核速率低的主要因素.添加质量分数为2%滑石粉等成核剂后,PLA结晶速率没有明显提高;含10%滑石粉的PLA结晶速率略上升,...  相似文献   

18.
19.
聚乙二醇增塑聚乳酸的非等温结晶动力学研究   总被引:2,自引:0,他引:2  
田怡  钱欣 《聚酯工业》2007,20(2):15-18
采用DSC方法对聚乙二醇(PEG)增塑聚乳酸的非等温结晶动力学进行了研究。结果表明,PEG的加入明显提高了聚乳酸的结晶速度。对所得数据分别用Ozawa方程和莫志深方法进行了处理,发现在给定温度范围里非等温结晶时,PLA/PEG主要是以均相成核的三维生长方式结晶;PLA的结晶速度随着PEG分子质量的增加而升高。  相似文献   

20.
Nonisothermal melt crystallization kinetics of biodegradable PBSU/PVPh blend was investigated with differential scanning calorimetry (DSC) from the viewpoint of practical application. PBSU/PVPh blends were cooled from the melt at various cooling rates ranging from 2.5 to 40°C/min. The crystallization peak temperature decreased with increasing the cooling rate for both neat and blended PBSU. Furthermore, the crystallization peak temperature of PBSU in the blend was lower than that of neat PBSU at a given cooling rate. Two methods, namely the Avrami equation and the Tobin method, were used to describe the nonisothermal crystallization of PBSU/PVPh blend. It was found that the Avrami equation was more suitable to predict the nonisothermal crystallization of PBSU/PVPh blend than the Tobin method. The effects of cooling rate and blend composition on the crystallization behavior of PBSU were studied in detail. It was found that the crystallization rate decreased with decreasing the cooling rate for both neat and blended PBSU. However, the crystallization of PBSU blended with PVPh was retarded compared with that of neat PBSU at the same cooling rate. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 104: 972–978, 2007  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