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 960 毫秒
1.
Thermally stimulated depolarization (TSD) experiments were carried out on several polystyrene samples. They included normal head-to-tail polystyrene (atactic) obtained by anionic polymerization of styrene, amorphous and substantially crystalline isotactic polystyrene, and the newly available head-to-head polystyrene. By TSD, six maxima of current intensity occurred at specific temperatures. Their features are compared for the various samples. Only three peaks could be identified with transitions which had been found by other techniques. Peak 5, located near Tg, is the primary relaxation. Maximum 6 could be the transition found above the Tg by torsional braid analysis and called T11 for polystyrene samples. Maximum 1 seems to correspond to what is sometimes referred to as the γ transition.  相似文献   

2.
Molecular dynamics (MD) simulation of the local motion of a polystyrene (PS) chain with anthryl group at the chain end surrounded by benzene molecules was performed and the results were compared with those obtained experimentally by the fluorescence depolarization method. The molecular weight dependence of the relaxation time of the probe obtained by the MD simulation was qualitatively in agreement with the results obtained by the fluorescence depolarization method. We also estimated the molecular weight dependence of the relaxation time for the end-to-end vector. Below the degree of polymerization (DP)≤3, the mean relaxation time Tm for the end-to-end vector was similar to that for the vector corresponding to the transition moment of the probe. With the increase of DP, the Tm for the probe tended to reach an asymptotic value unlike that for the end-to-end vector, which monotonically increased with DP. This indicates that the entire motion of a polymer coil contributes to the local motion to a lesser extent as the molecular weight increases. The MD simulations using artificial restraints showed that the rotational relaxation of the probe at the chain end for a dynamically stiff PS chain is realized by the cooperative rotation of the main chain bonds. The internal modes which takes place below 5 monomer units mainly led to the rotational relaxation of the probe at the PS chain end. Finally, the change of Tm with the position along the PS main chain was examined.  相似文献   

3.
Wholly aromatic, thermotropic homopolyesters, derived from 4,4′-biphenol, substituted biphenols, or 1,1′-binaphthyl-4,4′-diol and 3,4′-benzophenone dicarboxylic acid, and two copolyesters, each of which contained 30 mol % of 6-hydroxy-2-naphthoic acid, were prepared by acidolysis polycondensation reactions and characterized for their liquid crystalline properties. The solubility behavior of these polymers has also been investigated. The two homopolymers of phenyl-substituted biphenols with 3,4′-benzophenone dicarboxylic acid were soluble in many common organic solvents. All of the homopolymers had lower Tm/Tf values than those with terephthalic acid, which was attributed to the incorporation of the asymmetric 3,4′-benzophenone dicarboxylate units in a head-to-head and head-to-tail fashion along the polyester chain. Two copolymers had lower Tm values than those of the respective homopolymers, as expected. They formed nematic phases which persisted up to 400°C, except those of phenyl-substituted biphenols with 3,4′-benzophenone dicarboxylic acid. Each of these two polymers also exhibited an accessible Ti transition, and had a broad range of LC phase. They had glass transition temperatures, Tg, in the range of 139-209°C and high thermal stabilities in the temperature range of 465-511°C. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
The 13C relaxation times (T1 and T2) and isotropic contact shifts (Δω) of a one molar aqueous solution of l-proline at pH = 11 (or pD = 11.4) containing ca 10?4 M copper(II) perchlorate are measured at 62.86 MHz over a temperature range of 26–70°C. The purely dipolar longitudinal relaxation of carbon-13 nuclei contrasting with purely scalar transverse relaxation allowed us to extract carbon-to-metal distances (through T1 measurements) and hyperfine coupling constants and dynamic parameters (from T2 and Δω measurements). The structure of the complex in solution is found closely similar to that in the solid state. Curve-fitting procedures allowed us to derive the hyperfine electron—carbon coupling constants Ac = ?1.95, + 0.42, + 1.90 and ?1.70 MHz for carbons α, β, γ, δ, of the pyrrolidinic ring, the reorientation correlation time of the complex, τR (25°C) = 1.15 × 10?10 sec, the l-proline exchange rate, kM (25°C) = 4.0 × 105 sec?1 (and the corresponding activation parameters ΔH = 9.0 kcal mol?1 and ΔS = ?0.7 e.u.), and the electronic relaxation time, T1e = 1.13 × 10?8 sec (at 25°C). The latter value was found in agreement with the one computed from ESR data and the above τR value, showing the predominant contributions of spin—rotation interaction and, to a lesser extent, of the effect of g-tensor anisotropy to the electronic relaxation rate.  相似文献   

5.
《Thermochimica Acta》1986,109(1):193-206
Thermogravimetric (TG) runs were made with a Ni sample over a wide range of heating and cooling rates using an electromagnet at three different field strengths. The extrapolated onsets and endpoints of the weight change induced by the magnetic field gradient were obtained from both TG and the differential curves (DTG). The onset upon heating and endpoint on cooling are very clearly dependent upon magnetic field strength and hence unsuitable for temperature calibration using this method. The endpoint upon heating and onset upon cooling, however, showed only a slight dependence upon the cube root of the magnetic field strength. Observed values of Tc showed a linear dependence upon heating or cooling rates which extrapolated to a common point at zero rate. The value of Tc derived from extrapolation to zero rate and then zero field is 339.15°C, which is 16°C below the literature value.Simultaneous measurements of the melting point (Tm) of Pb and the Tm of Ni using a different furnace/thermocouple/magnet combination indicated a value of Tm = 311.8°C and Tc = 343.1°C. The value of Tm is 15.7°C below the value from the International Temperature Scale. The value of Tc is uncorrected to zero magnetic field strength, which would probably lower it several degrees. It would appear that both methods of temperature calibration are suitable for TG and the choice is best determined by the convenience with the particular apparatus involved and the temperature range desired.  相似文献   

6.
The thermally stimulated depolarization current (TSDC) technique has been used to study the slow molecular mobility of polysulfone in the glassy state and in the glass transformation region, i.e., in the temperature ranging from ?155 to 183 °C. Since the polysulfone is a rigid polymer without polar side-groups, a broad and low-intensity secondary relaxation was detected in the temperature region from ?120 °C up to the glass transition; the activation energy of the motional modes of this secondary relaxation is in the range between 35 and 100 kJ mol?1. The glass transition temperature of polysulfone provided by the TSDC technique is T M = T g = 176 °C (at 4 °C min?1). The relaxation time at this temperature is τ(T g) = 33 s and the fragility index was found to be m = 91. Our results are compared with literature values obtained by dynamic mechanical analysis and by dielectric relaxation spectroscopy. The amorphous polysulfone was also characterized by DSC; a glass transition signal with an onset at T on = 185.5 ± 0.3 °C (heating rate 10 °C min?1) was detected, with ΔC p = 0.21 ± 0.01 J g?1 °C?1.  相似文献   

7.
Phase equilibria in Li,K∥F,NO3 and Li,K∥Cl,NO3 three-component reciprocal systems were studied by differential scanning calorimetry (DSC). Eutectic compositions (mol %) in the Li,K∥F,NO3 system were determined to be as follows: 5.0 LiF, 10.0 KF, and 85.0 KNO3 with T m = 281°C and 48.5 KNO3, 44.0 LiNO3, and 7.5 LiF with T m = 105°C. Eutectic compositions (mol %) in the Li,K∥Cl,NO3 system were determined to be as follows: 10.0 LiCl, 32.1 KCl, and 57.9 LiNO3 with T m = 147°C and 44.5 KNO3, 45.0 LiNO3, and 10.5 KCl with T m = 97°C.  相似文献   

8.
Self‐metathesis of erucic acid by [(PCy3)(η‐C‐C3H4N2Mes2)Cl2Ru = CHPh] (Grubbs second‐ generation catalyst) followed by catalytic hydrogenation and purification via the ester yields 1,26‐hexacosanedioate (>99% purity). Polyesterification with 1,26‐hexacosanediol, generated from the diester, affords polyester‐26,26, which features a Tm of 114 °C (Tc = 92 °C, ΔHm = 160 J g−1). Ultralong‐chain model polyesters‐38,23 (Tm = 109 °C) and −44,23 (Tm = 111 °C), generated via multistep procedures including acyclic diene metathesis polymerization, underline that melting points of such aliphatic polyesters do not gradually increase with methylene sequence chain length. Available data suggest that to mimic linear polyethylenes thermal properties, even longer sequences, amounting to at least four times a fatty acid chain, fully incorporated in a linear fashion are required.  相似文献   

9.
Low-angle electron diffraction (LAED) was used to study the microstructure of crazes produced at different temperatures T and strain rates in thin films of monodisperse polystyrene (PS). At a slow strain rate of 4.1 × 10?6 s?1 both the fibril diameter D and the fibril spacing D0 of crazes in 1800k molecular weight PS remained constant with temperature up to T ≈ 70°C and then sharply increased as T approaches Tg. At a higher strain rate of ~ 10?2 s?1, both D and D0 increase only slightly with T. The values of D and D0 over a range of temperature are in very good agreement with those values obtained in bulk samples using small-angle x-ray scattering. The crazing stress was measured as a function of temperature in the thin films of the 1800k molecular weight PS strained at the same slow strain rate used for the LAED measurements. These measurements were analyzed using a simple model of craze growth to reveal the temperature and strain rate dependence of the craze surface energy Γ. At room temperature Γ ≈ 0.076 J/m2 (versus Γ ≈ 0.087 J/m2 predicted) and was observed to remain constant up to T ≈ 70°C and then decrease by approximately a factor of two at T = 90°C. This decrease in Γ is believed to result from chain disentanglement to form fibril surfaces at sufficiently high temperatures and occurs in the same temperature range in which the craze fibril extension ratio λ was observed to increase.  相似文献   

10.
To study the effect of processing history, molecular weight/molecular weight distribution, and thermal history on solid state properties (in particular fracture properties and orientation), carefully characterized polydisperse and monodisperse polystyrene samples were drawn above Tg and the orientation frozen in. The objective was to simulate the incidental orientation of polymer chains after processing, molding, and so forth (e.g., injection or compression, blow molding) as a result of melt flow. A series of polystyrene samples was produced by hot drawing at temperatures of 113 and 148 °C, followed by a relaxation period, and then a quench to below Tg. The level of segmental orientation imposed in the samples was determined by birefringence measurements. The tear energy of the sheets was measured at 20 °C by tearing along the draw direction, ultimately giving a value for the fracture energy, G3C. Samples of high draw ratio and low segmental orientation were unexpectedly found to have highly anisotropic fracture properties despite the low level of optical anisotropy. The fracture properties also depended significantly on whether the samples were drawn with or without lateral constraint. The results are compared with measurements of isotropic samples and the findings of a previous investigation utilizing SANS and birefringence. Modeling the drawing conditions at the chain level using a recent nonlinear tube theory explains how birefringence alone is an inadequate measure of molecular orientation. © 2007 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 45: 377–394, 2007  相似文献   

11.
A phase diagram of the tetraethylammonium hydroxide-water binary system was studied by differential thermal analysis. Seven crystal phases of tetraethylammonium hydroxide were found. Two of them are stable: tetrahydrate (T m = 47.0°C), two phases of a compound with the hydration number 7.5 (T m = −18.1 and −23.2°C); five are metastable phases: two polymorphs of pentahydrate (T m = 39.4 and 0.9°C), hexahydrate (T m = −70.1°C), and a compound with hydrate number 10–11 (T m = −112°C).  相似文献   

12.
Magnetic susceptibility measurements on the rhombic form of manganocene, Mn(C5H5)2, are reported between 0.94 and 297.5 K. A pronouced minimum in the 1/xcorrm versus T curve is located at Tmin = 141 ± 1 K. The results are well reproduced by the solution of Fisher to the quasi one-dimensional Heisenberg linear chain modified such as to include interchain interaction for J/k = ?14 K, g = 2x = ?5 K. A less accurate approximation to the experimental results has been achieved in terms of the pair model (J/k = ?22.5 K, g = 2.0). The results are in excellent agreement with the zigzag chain structure of C5H5Mn units reported for the compound.  相似文献   

13.
A new two‐stage draw technique was successfully applied to the superdrawing of polytetrafluoroethylene (PTFE) virgin powder. A film, compression‐molded from powder below the melting temperature (Tm = 335 °C), was initially solid‐state coextruded to an extrusion draw ratio (EDR) of 6–20 at 325 °C, about 10 °C below the Tm. These extrudates from the first‐stage draw were further drawn by a second‐stage pin draw in the temperature (Td) range of 300–370 °C that covers the static Tm. The maximum achievable total draw ratio was ~60 at a Td = 300 °C and increased rapidly with increasing Td, reaching a maximum of 100–160 at a temperature window between 340 and 360 °C, depending on the initial EDRs. At yet higher Td's, the ductility was lost as a result of melting. The high ductility of the PTFE extrudates at such high temperatures was ascribed to the improvement of interfacial adhesion and bonding between the deformed powder particles upon the first‐stage extrusion combined with the rapid heating of only a portion of the extrudate followed by the elongation at a high rate. The highly drawn fibers were highly crystalline (χc ≤ 87%) and showed high chain orientation (fc ≤ 0.997) and a large crystallite size along the chain axis (D0015 ≤ 160 nm). The molecular draw ratio, estimated from the entropic shrinkage above the Tm, was close to the macroscopic deformation ratio independently of the initial EDRs. These results indicate that the draw was highly efficient in terms of chain extension, orientation, and crystallization. Thus, the maximum tensile modulus and strength achieved in this work were 102 ± 5 and 1.4 ± 0.2 GPa, respectively, at 24 °C. These tensile properties are among the highest ever reported on oriented PTFE. © 2001 John Wiley & Sons, Inc. J Polym Sci Part B: Polym Phys 39: 1995–2004, 2001  相似文献   

14.
The molecular-topological structure of polytetrafluoroethylene (PTFE) has been studied in the range of ?100 to +450°C by thermomechanical spectrometry. Revealed in this temperature range is a fourblock topological structure composed of one amorphous (T g = 16°C) and three crystalline (low-melting (T m = 315°C), intermediate (T m 1 = 355°C), and high-melting (T m 2 = 388°C)) polymorphs. At a dose of 1 kGy, the long-range orientation of chains in the intermediate and high-melting crystalline blocks of PTFE is replaced by short-range orientation of the cluster association structure. At doses of 100?C500 kGy, the latter structure transitions to the amorphous state and the irradiated samples acquire a semicrystalline structure of the two-block type. The molecular-mass distribution function of interjunction chains of the pseudo-network of the amorphous block is bimodal in character and its maxima are noticeable shifted toward lower masses with an increase in the radiation dose. As the dose increases, the crystallinity decreases and the molecular mobility of amorphized chains is enhanced. As a result, both the glass transition and the molecular flow onset temperatures of the polymer are reduced.  相似文献   

15.
The heat of immersion in water was measured at 25°C for three iron(III) oxides using a twin-type microcalorimeter. One of the samples was commercial α-Fe2O3 (sample C) and the other two (samples M and F) were prepared by calcining magnetite and iron(III) hydroxide in air at various temperatures, Tp, from 300 to 700°C. The samples were evacuated at outgassing temperature, To, between room temperature and 500°C at a pressure of 1 × 10?2?2.7 × 10?2N m?2 for 6 h. The heat of immersion, hi(J m?2), of samples C and M increased with an increase in To and showed the maximum hi at To =400°C, while sample F did not show the maximum up to To =500°C. The systematic correlation was not observed between hi and Tp of sample F. The heat of reproduction of the surface hydroxyl group on sample F was approximately estimated as 6.6 × 104 J mole?1 H2O.  相似文献   

16.
The temperature dependence of T1 for 3He gas in the range 0–4°K is calculated for a Lennard-Jones (12,6) potential. The relaxation of the nuclear spins is assumed to be due to a dipolar interaction between the nuclei. A minimum value in the relaxation time, T1,min, is found to occur at a temperature denoted by Tmin. By repeating the calculation for different pairs of values of the potential parameters ? and σ, we have found that for a density of 10?2 g/cm32Tmin = 13.0?1.12 × 1032, T1,min2(Tmin)12 = 17.4?6.56 × 1022, with ?, σ, Tmin and T1,min in eV, Å, °K and minutes, respectively. From measurements of Tmin and T1,min, ? and σ can be determined.  相似文献   

17.
The tensile stress relaxation master curve for polyvinylcyclohexane (completely hydrogenated polystyrene) has been measured. Direct relaxation experiments were carried out at several temperatures above the glass transition temperature over the rather long time range of four orders of magnitude. This long time span was realized by calculating the modulus during the period when a constant small strain rate was applied to the sample as well as during the usual constant strain interval. A computer solution to the Boltzmann superposition equation allowed data from these two regions to be joined into a smooth curve representing E(t), a parameter indicative of an instantaneous strain experiment. The measured Ti was found to be 143°C; Tg is expected to fall within several degrees of this temperature. This result is apparently at odds with a previously reported Tg value of 120°C. More importantly, the maximum value of the negative slope of the stress relaxation master curve of polyvinylcyclohexane in the primary transition region was only slightly different from that for polystyrene. This observation clearly indicates that the molecular factors which result in the highly coupled nature of the primary transition in polystyrene are not strongly dependent upon any side-chain π–π interactions which might be present in polystyrene.  相似文献   

18.
EuSb2 crystallizes in the monoclinic CaSb2 type of structure, space group P21m, with a = 4.768(2), b = 4.299(2), c = 8.9703(3) Å, β = 103.01(3)°. The structural distortions of the ZrSi2 type provide the Sb chains required for a Mooser-Pearson phase. EuSb2 is antiferromagnetic below TN = 26.2°K. From high-field magnetization measurements a weak anisotropy (Hanis. ≈ 6 kOe at 1.5°K) is deduced. The spins are aligned perpendicular to the (001) planes. Susceptibility measurements between 30 and 1100°K gave no indication of a Eu2+ → Eu3+ transition.  相似文献   

19.
A round robin test was performed to determine the reliability of values for the glass transition temperatureT g as determined by DTA on polymers. Ten different instruments were involved. The test material was high molecular weight polystyrene. Values forT g (midpoint) were reported in the range 107°C±2 K. The respective heat flow curves differed considerably in shape. In the literature aT g of 100°C is often given for polystyrene. The discrepancy between this value and the value of 107°C found in the round robin test is due to three differences: the thermal history of the sample, the evaluation of the heat flow curves, and the effect of finite sample size.  相似文献   

20.
The changes in crystalline phase of poly(butylene naphthalate) (PBN), poly(diethylene naphthalate) (PDEN) and poly(thiodiethylene naphthalate) (PTDEN) upon thermal treatments were evaluated by X-ray diffraction technique. The melting behavior and the crystallization kinetics of the polymers under investigation were investigated by means of differential scanning calorimetry. Multiple endotherms were evidenced in PBN and PTDEN, due to melting and recrystallization processes. By applying the Hoffman?CWeeks?? method, the T m° of PBN, PDEN, and PTDEN was derived: the introduction of ethero-atoms along PBN polymer chain causes a decrement of T m° value. The isothermal crystallization kinetics was analyzed according to the Avrami??s treatment: the presence of ether-oxygen or sulphur atoms in the chain deeply reduces the PBN ability of crystallizing. Finally, no interphase was evidenced both in PDEN and PTDEN.  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