首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The measurements of ac conductivity [σm(ω)], dielectric constant [?′(ω)] and loss [?″(ω)] have been performed on solution grown (thickness ∼85 μm) and uniaxially stretched (thickness ∼25, 45 and 80 μm) films of poly(vinylidene fluoride) (PVDF) in the frequency range 0.1 kHz-10 MHz and in the temperature range 77-400 K. The σm(ω) can be described by the relation σ(ω) = s, where s is close to unity and decreases with increase in temperature. Three relaxations, observed in the present investigation, have been designated as the αc-, the αa- and the β-relaxations appearing from high temperature side to the low temperature side. The αc-relaxation could not be observed in the case of uniaxially stretched poly(vinylidene fluoride) films. The αc- and αa-relaxations are associated with the molecular motions in the crystalline regions and micro-Brownian motion in the amorphous regions of the main polymer chain, respectively, whereas the β-relaxation is attributed to the rotation of side group dipoles or to the local oscillations of the frozen main polymer chain.  相似文献   

2.
S. Saeki  F. Wang  Y. Tanaka 《Polymer》2006,47(21):7455-7459
An equation of state for zero internal pressure in rare gas solids and semi-crystalline polymers has been determined based on the empirical functions of thermal pressure coefficient γV with respect to volume at constant pressure. The experimental data of PVT over wide range of temperature and pressure published by Anderson and Swenson and Syassen and Holzapfel for rare gas solids and Olabisi and Simha and Zoller for semi-crystalline polymers are used to evaluate γV. The function of γV with respect to volume determined at constant pressure is given by where V0 is the volume at 0 K, A, ? and c are constants. The function of internal pressure Pi = γVT − P with respect to temperature at constant pressure is determined by converting the function of γV(V) to a function of temperature γV(T). An empirical equation of state for zero internal pressure determined by pressure P, volume V and temperature T at which Pi = 0 is expressed by PV/RT=CDV for rare gas and semi-crystalline polymer where C and D are constants. The practical meaning of the equation of state for Pi = 0 in the semi-crystalline polymers has been discussed.  相似文献   

3.
Long-chain aliphatic α,ω-diols containing up to 32 consecutive methylene groups were synthesized by several methods and characterized. 1,22-Docosanediol HO-(CH2)22-OH and 1,32-dotriacontanediol HO-(CH2)32-OH both exhibited a solid-solid phase transition before melting. The α,ω-diols HO-(CH2)m-OH, where m=12, 22, or 32, were reacted in the melt with much shorter aliphatic α,ω-diisocyanates OCN-(CH2)n-NCO, where n=4, 6, 8, or 12, producing a series of linear, aliphatic, and increasingly polyethylene-like m,n-polyurethanes. Characterization (by DSC, TGA, and SAXS) of the m,n-polyurethane series showed that when the aliphatic segments were increased, and the hydrogen-bonding densities thus decreased, the polymers displayed physical and thermal properties (for example, solubility and melting temperature) typical of polyethylene.  相似文献   

4.
Ryo Konishi 《Polymer》2004,45(15):5191-5198
High-resolution solid-state 15N and 13C NMR and dynamic mechanical measurements were carried out for solution grown crystals of the α- and γ-forms of nylon-6 to understand the relation between segmental mobility including interchain interactions and ductility of polyamides. The ductility, in the temperature range of 100-180 °C, was lower for the α-crystals than for γ-crystals. 15N chemical shift revealed that the hydrogen bonding was stronger for the γ-crystals than for α-crystals. In addition, lower mobility of NH group in the γ-crystals than in α-crystals was shown by results. The results suggest that the ductility of nylon-6 could not be simply related to the strength of hydrogen bonding. Highly crystalline films of aggregates of solution grown α-crystals and γ-crystals showed no crystalline relaxation in the temperature dependence of dynamic mechanical loss factor, suggesting an existence of strong intermolecular interactions in the crystalline regions although data indicated that the mobility of methylene units was higher in the γ-crystals than in α-crystals. Information on the large-scale chain dynamics in the crystalline regions might be necessary to understand the low ductility of polyamides.  相似文献   

5.
K.L. Ngai  T.R. Gopalakrishnan 《Polymer》2006,47(20):7222-7230
The voluminous amount of data in the literature on the structural α- and the Johari-Goldstein β-relaxations of the poly(n-alkyl methacrylate)s allows a systematic study of the interrelation between the two important relaxation processes. The data bring out the systematic changes in the interrelation between the structural α- and the Johari-Goldstein β-relaxations with changes in molecular structure, molecular weight, tacticity and size (by nanoconfinement), and modifications by copolymerization, and crosslinking. The results can all be interpreted as primarily due to changes in intermolecular coupling, which have significant effects on the many-molecule dynamics constituting the structural α-relaxation, but not on the precursory Johari-Goldstein β-relaxation. Theoretically, the Coupling Model predicts a relation of intermolecular coupling (or degree of cooperativity of the α-relaxation) to the ratio of the α- and the β-relaxation times, and a correlation of intermolecular coupling to the steepness or “fragility” index. The predicted relation and correlation are compared with experimental data of the poly(alkyl methacrylate)s.  相似文献   

6.
The dielectric and dynamic mechanical behaviour of bi-stretched non-treated and annealed semicrystalline poly(ethylene 2,6-naphthalene dicarboxylate) (PEN) films are studied as a function of different morphologies obtained by thermal treatments at temperatures close to the melting temperature of a semicrystalline film. Differential scanning calorimetry (DSC) shows that the glass transition temperatures do not change significantly with the thermal treatment for bi-stretched films. However, the melting temperatures and the degree of crystallinity increase with the value of annealing temperature. Both dielectric relaxation spectroscopy (DRS) and dynamic mechanical analysis (DMA) display three relaxation processes. In order of decreasing temperature, can be observed: the α-relaxation due to the glass transition, the β-process assigned to cooperative molecular motions of the naphthalene groups which aggregate and the β-relaxation due to local fluctuations of the carbonyl groups. The α-relaxation process shifts to higher temperatures for the 250 and 260 °C treated bi-stretched semicrystalline samples compared to the sample thermally treated at 240 °C according to DRS data but shifts to lower temperatures according to the DMA measurements for the three annealed samples. This discrepency results from the different sensitivity of each methods with regards to the release of orientation. At a fixed frequency the temperature associated to β-relaxation is lower for the non-treated bi-stretched semicrystalline samples than for the treated ones using DMA but no difference can be seen in DRS. The associated apparent activation energies are rather high which suggest cooperative motions. It is assumed that the orientation of the samples prevents coupling between the naphthalene groups due to the stretched chain configuration in the amorphous phase. The activation energy for the β-process given by DRS is independent of the thermal treatment and the value agrees with those found for poly(ethylene terephthalate) (PET) and amorphous PEN. Evidence of the decrease of orientation in the sample with thermal treatment can be seen via the onset of mobility, both by DRS and DMA. Thus, the orientation induces a greater change of properties compared to the crystalline samples obtained from the thermal treatment of an amorphous sample. Finally, a three phase model is proposed since there is evidence of a rigid amorphous phase present in PEN biaxially stretched samples which was favoured by the dependence of dielectric relaxation strengths on the degree of crystallinity for the β- and α-relaxation.  相似文献   

7.
Single crystals of the β-Ca3B2SiO8 new monoclinic modification have been obtained by cooling the melt of a stoichiometric composition. The crystal structure has been determined from the single crystal X-ray diffraction data and refined with R = 0.059 (wR = 0.069) in the monoclinic space group P21/m. The thermal behavior of the synthetic borosilicate has been studied. At 472 ± 5°С, a reversible phase transition of the first order occurs, leading to the formation of the orthorhombic α-Ca3B2SiO8 modification. The thermal expansion of α- and β-modifications of Ca3B2SiO8 is anisotropic: (α11 = 15, α22 = 16, α33 =–1, α V = 30 × 10–6°С–1) and α11 = 9, α22 = 28, α33 = 1, α V = 38 × 10–6°C–1, respectively.  相似文献   

8.
9.
Young Gyu Jeong  Won Ho Jo 《Polymer》2008,49(6):1693-1700
It was revealed that poly(octamethylene 2,6-naphthalate) (PON) existed in two different crystal structures, α- and β-form, depending on crystallization process: The α-form crystal was dominantly developed from the cold-crystallization, whereas the β-form was from the melt-crystallization. The apparent melting temperatures of α- and β-form crystals were characterized to be 175 and 183 °C, respectively. On the basis of X-ray diffraction and molecular modeling studies, the crystal structure of β-form, developed dominantly from the melt-crystallization, was identified to be triclinic with dimensions of a = 0.601 nm, b = 1.069 nm, c = 2.068 nm, α = 155.68°, β = 123.25°, γ = 52.85°, and with the space group of . The calculated crystal density was 1.243 g/cm3, supporting that one repeating unit of PON exists in a unit cell. The octamethylene units in the PON backbone take largely all-trans conformation in the β-form unit cell.  相似文献   

10.
M. Soccio  L. Finelli  A. Munari 《Polymer》2007,48(11):3125-3136
Polyesters based on 1,3-propanediol glycol and aliphatic dicarboxylic acids with different chain length were synthesized by melt polycondensation, obtaining samples characterized by high and comparable Mn. The polymers were subjected to molecular and thermal characterization. All polymers showed a good thermal stability, even though depending on the chain length. At room temperature all the polymers appeared as semicrystalline materials; the effect of the chain length was a lowering in the Tg value, an odd-even fluctuation for Tm and an increase of the crystallization rate. A comparison of the X-ray data revealed that the polymers with odd carbon number per repeat unit, show similar patterns, different from those of samples with even carbon atoms number. Multiple endotherms were evidenced in melt isothermally crystallized samples, due to melting and recrystallization processes. By applying the Hoffman-Weeks' method, the of the samples was derived. Lastly, the presence of an interphase was not evidenced.  相似文献   

11.
The dielectric behaviour of solution-grown thin films of poly(methylmethacrylate) containing p-nitroaniline (p-NA) as a dopant was investigated within the temperature range 60–90°C and a 20–105 Hz frequency band. It is shown that these mixtures exhibit only one relaxation process similar to that of pure PMMA. However, the addition of p-NA increases both the height and the relaxation strength of the peaks and also shifts log fm to higher frequencies. The additive causes narrowing of the loss curves and increases the activation energy for relaxation. These results are interpreted in terms of the hydrogen bonding effects of p-NA on localized motions of carboxymethyl dipoles in PMMA. An insight into the possible origins of the β- and α-relaxations in PMMA is also presented.  相似文献   

12.
Dynamic mechanical properties have been investigated over the temperature range of?150–360°C for the quinazolone–imide copolymers, prepared by condensation of the amine-terminated quinazolone prepolymer with a stoichiometric quantity of pyromellitic dianhydride or 3,3′,4,4′-benzophenone-tetracarboxilic dianhydride. Both copolymers have, respectively, the low-temperature β-relaxation and the β*-relaxation, as well as the case of poly(4,4′-oxydiphenylene–pyromellite-imide) (Kapton) examined for comparison. These relaxations seem to contribute to toughness of the copolymers. The α-relaxations for both copolymers occurred at much the same temperature of 320°C, which can be assigned to a large scale segmental motion of the quinazolone chain sequence. The α-peak temperatures shifted into higher temperatures by heat aging. This can be explained in terms of crosslinking in the copolymers, supported by swelling test in hot m-cresol and IR spectroscopy.  相似文献   

13.
Curdlan is a high-molecular-weight linear β-1,3-glucan synthesized by microorganisms. A series of curdlan branched esters with a degree of substitution of three were synthesized and their physical properties and structures were compared with those of curdlan linear esters. Thermal degradation temperatures of all the curdlan branched esters were ca. 360 °C; almost the same as those of curdlan linear esters. The curdlan branched esters had melting temperatures (T m ) higher than those of the corresponding curdlan linear ester with the same side-chain carbon number. In particular, comparing T m of curdlan propionate, curdlan isobutyrate, and curdlan pivalate, the latter two had high T m of over 335 °C, suggesting that the degree of branching of the side chain affects the stability of molecular chains with helix structure in their crystals. Highly transparent films were prepared from the curdlan branched esters. These films exhibited higher Young’s modulus and tensile strength compared with those of films composed of the linear equivalents with the same side-chain carbon numbers. These results indicate that curdlan branched esters are promising thermoplastics with favorable thermal and mechanical properties because of the closer packing structure of their molecular chains than that of the corresponding curdlan linear esters.  相似文献   

14.
A phenylacetylene bearing an n-butylsulfoxide group, i.e., p-(n-butylsulfoxide)phenylacetylene (1) was prepared in high yields using the [Rh(norbornadiene)Cl]2-NEt3 catalyst in the presence of various solvents under mild conditions. The resulting polymer, poly[p-(n-butylsulfoxide)phenylacetylene] (poly(1)), was characterized in detail by 1H NMR, ESR, laser Raman, and diffuse reflective UV-vis methods. The data clearly showed that cis-to-trans isomerization of the polymer can be induced when pressure is imposed to the polymer at room temperature, rotationally breaking the cis CC bonds to generate the cis and trans radicals. Further, the spin density in the cis radical was migrated from the main-chain to the sulfoxide moiety as the side chain of the phenyl ring to magnetically interact with the first two methylene protons in the n-butyl group giving a triplet line ESR spectrum with an extremely large g value, g = 2.0081.  相似文献   

15.
Poly(butylene fumarate) (PBF) bearing reactive double bonds on the polymer main chains has been designed and synthesized by coupling with hexamethylene diisocyanate (HDI) under very mild condition. The chemical structure, conformational structure, crystal structure and molecular weight of PBF were systematically characterized by ATR-FTIR, 1H NMR, 13C NMR, GPC and WARD. The thermal properties, mechanical properties and biodegradability of PBF were carefully studied by DSC, mechanical testing and enzymatic degradation. The results of 1H NMR and 13C NMR spectra indicate that no isomerization or Ordelt saturation reaction of trans CC took place during the bulk polymerization and the reaction just proceeded in the way we designed. Linear PBF with high-molecular-weight has been successfully synthesized. This new type of uncrosslinked polyester is shown to have many merits such as relatively high melting point (Tm), satisfactory processability and good mechanical properties. The impact strength of PBF is higher than 200 J/m; tensile and flexural strength can reach to 41.0 and 26.7 MPa, respectively.  相似文献   

16.
From a consideration of the work required for expansion of a liquid, the following relationship between viscosity η, pressure P and temperature T is put forward. For unassociated liquids with molecules which are not too large, V* is taken as the parachor, log10 (η* in Ns/m2) is ?3.88, P* is 8.58 × 106 N/m2, R is the gas constant, and T* is a constant characteristic of each liquid. The equation can be applied to polymeric liquids if V* and η* are taken as disposable constants. For example, for polystyrene V* is found to be 3 × 10?3 m3 mol?1 and log10 (η* in Ns/m2) to be 3.4 log10 M?w ?10.2 where M?w is the weight-average molecular weight (kg/mol) from 5 kg/mol upwards. In the equation, the same constants serve for the variation of viscosity with pressure and with temperature. The viscosity under a high pressure can therefore be estimated from viscosities all measured at normal pressures but at different temperatures. The viscosities of a number of polymers have been measured over a range of temperature and pressure and the results support the equation. Support is found for the view that segments are involved in the flow of polymeric liquids and V* gives a measure of the volume of the segment. The size of the segment seems to increase as the flexibility of the polymer chain decreases. The lowest values for V* are found for polysiloxanes in which the segment seems to be only four atoms long. Larger values of V* are found for polymers with units of the type –CH2–CHR-. Larger values still of V* are given by polymers with units of the type –CH2-CR1R2- and even larger V* values are found for those polymers with benzene rings constituting a major part of the main chain. As V* rises the viscosity of the polymeric liquid becomes much more dependent upon pressure and temperature. Thus whilst the polysiloxanes have viscosities which are relatively insensitive to pressure and temperature, the aromatic polysulphones and poly(2,6-dimethylphenylene oxide) have viscosities which are very sensitive to pressure and temperature.  相似文献   

17.
The synthesis of monodisperse hard segments containg 2,4-TDI and 1,4-butanediol by a simple technique is described. High purity material is obtained in good yield. The structures of these hard segments are confirmed by proton n.m.r., 13Cn.m.r., and i.r. Glass transition temperatures and melting temperatures are reported. The Tg′s of the hard segments are inversely proportional to the reciprocal of their molecular weights. The Tm′s show an odd-even effect relative to the number of 2,4-TDI units in the hard segment.  相似文献   

18.
Multiple melting behavior of poly(butylene-2,6-naphthalate) (PBN) was studied with X-ray analysis and differential scanning calorimetry (DSC). Double endothermic peaks L and H attributed to the α-form crystal modification, a small peak attributed to the β-form crystal modification, and a new shoulder peak S at a lower temperature of peak H appeared in the DSC melting curves. Wide-angle X-ray diffraction patterns of the samples isothermally crystallized at 200 and 220 °C were obtained at a heating rate of 1 K min−1, successively. In this heating process, change of crystal structure and increase of quantity of the β-form crystallites could not be observed up to the final melting. With increasing temperature, the diffraction intensity decreased gradually and then increased distinctly before a steep decrease due to the final melting. The X-ray analysis clearly proved the melt-recrystallization during heating. The β-form crystal modification was formed during slow heating process in the high temperature region.  相似文献   

19.
Y.S. Hu  A.R. Kamdar  S.P. Chum  E. Baer 《Polymer》2006,47(18):6387-6397
The crystallization behavior and morphological patterns of a miscible blend of two propylene/ethylene (P/E) copolymers that differed in ethylene content were studied. Metallocene-catalyzed P/E copolymers containing 3.1 and 11.0 mol% ethylene were chosen for blending. The difference in ethylene content was small enough to ensure miscibility of the pair in the melt, and the ethylene content was low enough to ensure that both were crystallizable. The blends were characterized by differential scanning calorimetry (DSC), wide angle X-ray diffraction (WAXD), optical microscopy (OM) and atomic force microscopy (AFM). The complex melting endotherm of the blends consisted of a broad low temperature peak at Tm1, a high temperature peak at Tm2, and an intermediate peak at which was not characteristic of either constituent and depended on blend composition. The multiple melting peaks arose from distinct crystal populations. All the blends exhibited a mixed morphological texture of α-radial lamellae with short, densely packed γ-overgrowths, interspersed with areas of α-crosshatch. The high temperature peak at Tm2 was assigned to the melting of the α-radial lamellae which formed from chains of the lower comonomer constituent. The broad low temperature peak at Tm1 was attributed to the melting of γ-crystal overgrowths on the radial lamellae. The new peak at was thought to arise from the melting of the α-crosshatch lamellae. The lamellar thickness, and hence , correlated with the crystallization temperature, which decreased as the blend was made richer in the higher comonomer constituent.  相似文献   

20.
A series of processable semi-aromatic polyamides containing thioether and methylene units were synthesized through the reaction of 4,4-thiobis(methylene)dibenzoyl chloride and aliphatic diamine by the method of interfacial polycondensation. These polyamides had excellent thermal properties with glass transition temperatures (T g) of 104.3–130.6 °C, melting temperatures (T m) of 300.3–303.8 °C, and initial degradation temperatures (T d) of 405.2–410.3 °C. They had wider processing windows than traditional semi-aromatic polyamides (such as PA6T can not be processed by melting) and can be processed by melting method. They had better tensile strengths of 57.6–64.1 MPa, low-temperature mechanical properties, low water absorption of 0.19–0.27 %, low dielectric constants of 3.11–3.95 at 100 kHz, and better melt flowability properties of 232–60.7, 301.9–78.8, and 423.1–83.6 Pa s under a shear rate ranging from 20 to 1,170 s?1, respectively. In addition, these polyamides showed good corrosion resistance, they did not dissolve in solvents such as NMP, DMSO, hydrochloric acid (6 mol/l), and solution of NaOH (1 mol/l) and so on.  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