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 390 毫秒
1.
The variations of size, intensity, and size distribution of free volumes generated in the network of molecular chains of gelatin at the sol-gel transition were studied by means of the positron annihilation lifetime technique. Although variation in average free-volume radius was not recognized, a variation of free-volume content was observed at the sol-gel transition point of gelatin with an addition of saccharose.  相似文献   

2.
Positron annihilation spectroscopy has been used to study free volume in an arnine-cured epoxy as a function of external pressure at temperatures above and below the glass transition temperature. The observed ortho-positronium lifetime τ3 and formation probability I3 decreased with increasing pressure. The decrease in τ3 is interpreted in terms of a corresponding decrease in average free-volume hole size over the range from 0.135 to 0.045 nm3. The fractional free-volume and the free-volume compressibility in the epoxy are calculated as functions of pressure at 100°C.  相似文献   

3.
Positronium annihilation spectroscopy (PAS) has been used to study the microstructural properties of amine-cured epoxy polymers. We have determined the free-volume “hole” sizes in these polymers by comparing the observed ortho-positronium lifetimes with the known lifetime–free volume correlation for low-molecular-weight systems. The free volumes for four epoxies with different crosslink densities are found to vary significantly over the temperature range between ?78° and 250°C. The free-volume holes for these polymers are found to range from 0.025 to 0.220 nm3. Two important transition temperatures were found: one corresponds to the glass transition temperature Tg determined by differential scanning calorimetry (DSC), and the other occurs about 80–130°C below Tg. The sub-Tg transition temperature is interpreted tentatively as being where hole size reaches dimensions adequate for positronium trapping or else the onset temperature for local mode or side-chain motions. These two transition temperatures plus two additional onset temperatures are found to be correlated with crosslink densities calculated from stoichiometry.  相似文献   

4.
Positron lifetime studies were performed on well-characterized annealed and quenched samples of isotactic polypropylene. The positron experiments were conducted from ?20 to 110°C as a function of both heating and cooling. Of the three decaying exponential components resolved from the lifetime spectra, only the long-lifetime ortho-positronium (o-Ps) pickoff component was affected by the changes in temperature. The behavior of both the lifetime and intensity of the o-Ps component was interpreted with the aid of x-ray diffraction, densitometry, and optical microscopy examinations and results from previously reported investigations of the thermal transition behavior of polypropylene. The present experiments demonstrate that o-Ps lifetimes were similar for both the annealed and quenched samples, independent of thermal cycling, while the o-Ps component intensity was significantly larger for the quenched material during heating, with both sample types exhibiting a significant hysteresis upon cooling. These results suggest that the mean free-volume cavity size is independent of prior thermal treatment, while the density of free-volume sites is a sensitive function of structure and prior thermal history. The variations of lifetime and of intensity with temperature have provided insight into polypropylene's glass transition phenomena.  相似文献   

5.
The free-volume, of size ranging from 0.2 to 0.4 nm in radius, in an ethylene-vinyl alcohol copolymer was estimated using positronium lifetime measurement to elucidate the dependence of oxygen permeability on the free-volume size and fraction, on the ethylene content and on the crystallinity. The permeability and the free-volume fraction with varying the ethylene content were well related and the relation was interpreted based on the free-volume theory near below and above the glass transition temperature. On the other hand, the crystallinity significantly influenced the fraction of the amorphous region, where the free-volume hole exists, along with a slight change of the free-volume size. The variation of the permeability with the crystalline degree cannot be explained from the averaged free-volume fraction estimated by the whole volume of the polymer, but the permeability correlated with the free-volume size apparently.  相似文献   

6.
荧光光谱跟踪结冷胶水溶液的溶液-凝胶转变   总被引:2,自引:0,他引:2  
将异硫氰酸荧光黄(FITC)标记在结冷胶分子链,用荧光光谱跟踪了结冷胶水溶液凝胶化过程中FITC荧光强度和各向异性比的变化.结果表明在结冷胶的凝胶化转变中,FITC的荧光相对强度和各向异性随温度降低而增大,在某一温度荧光相对强度和各向异性比对温度的曲线出现了明显的转折点,这个转折点的温度低于流变温度扫描曲线中G′=G″的温度.利用荧光的方法确定物理交联体系的关于重均聚合度和凝胶分数的相关临界指数γ和β.γ和β不符合Flory-Stockmayer和逾渗模型的预测.  相似文献   

7.
Positron annihilation lifetime measurement was applied to the study of free-volume properties in three kinds of polypropylene as a function of temperature in the range of 25–180°C at thermal equilibrium. Positron lifetime data for polypropylenes were analyzed with a Laplace inversion technique in order to obtain continuous positron annihilation lifetime (PAL) distributions. At each temperature, four distinct PAL distributions were recognized. The distribution of the longest lived component was associated with a pick-off annihilation of ortho-positronium (o-Ps) trapped in free-volume of amorphous region, which grew bigger as the temperature increased. The hole radius distributions of free-volumes were estimated from the results of o-Ps lifetime distributions. A detailed analysis showed a mean radius of free volumes was 0.34 nm at room temperature and that was 0.42 nm near the melting point for each specimen. The distributions of hole radii of free volumes were found to be broader after thermal treatments. The relaxation of free volumes was attributed to the thermal equilibrium and the evacuation of included molecules in free volumes. © 1995 John Wiley & Sons, Inc.  相似文献   

8.
The positron lifetime spectroscopy (PLS), a non-destructive characterization method, utilizes positronium (Ps; an electron–positron bound state) as a probe and measures its lifetime in polymer free volumes. For the first time the free volumes have been estimated by PLS in polyaniline (PANI) complexes with various inorganic and organic acids. It was found that the o-Ps lifetime increases and the intensity decreases with increasing ionic radius of the counter-ions in PANI complexes. Obviously, larger counter anions result in enhanced mean size of the voids corresponding to the free volume in the bulk polymer.Electrical conductivity has been measured by conventional four-probe technique. The glass transition temperature and temperature of removal of the absorbed water have been determined by using differential scanning calorimetry. It was established fairly well correlation of the mentioned polymer parameters with the o-Ps lifetime and the free volume of PANI complexes, respectively. The greater free-volume results in a decrease of conductivity, glass transition temperature and temperature of removal of the absorbed water.  相似文献   

9.
Unique information about the properties of free-volume sites in polymers is gained from Positron Annihilation Lifetime (PAL) measurements. After calibration with data from other techniques the method may be used to determine free-volume fractions. From pressure–volume–temperature (PVT) and PAL (ortho-Ps lifetime τ3) data, measured on identical amorphous poly(methyl methacrylate) (PMMA) samples with controlled thermal histories, we find a linear relationship between free-volume fractions, derived from PVT measurements, the Simha–Somcynsky equation-of-state theory and the mean subnanometer free-volume size both below and above the glass transition temperature.  相似文献   

10.
The size distribution of free-volume (<~0.1 nm3) of ethylene-vinyl alcohol copolymer with various crystalline degrees was estimated by positronium lifetime measurement. With increasing degree of crystallinity, the size distribution significantly narrowed and the intensity of positronium decreased. This indicates that the inhomogeneity reduces with the increase of the degree of crystallinity. It is found that the free-volume fraction estimated is reflected by the fraction of the amorphous region.  相似文献   

11.
A photopolymerizable/cross-linkable dry-film system is generally a compatible mixture of three main ingredients: a preformed polymeric binder, a multifunctional polymerizable monomer, and a photoinitiating system. The photographic speed of systems sensitized with certain free-radical sensitizers which undergo intermolecular hydrogen-abstraction (e.g., benzophenone/Michler's ketone) is proportional to the excess fractional free volume (Vf ? Vfg), of the dry coating. Free-volume dependence appears to be determined by the molecular size or configuration of the actual initiating species. We used as a working model the copolymer system poly[4,4′-isopropylidenediphenylene-1,1,3-trimethyl-3-phenylindan-5,4′-dicarboxylate:azelate] and tri(6-acryloyloxyhexyl)-1,3,5-benzenetricarboxylate as the polymerizable/cross-linkable monomer. A mathematical model based on experimental data and the free-volume theory was developed. The photographic speed of a photopolymerizable/cross-linkable dry-film system suffering from free-volume dependence can be correlated with the glass transition temperatures of the polymeric binder and the monomer, the volume fraction of the monomer, and the temperature of exposure.  相似文献   

12.
Position annihilation spectroscopy (PAS) was used to measure the relative free-volume fraction of protective epoxy coatings before and after exposure to liquid water at room temperature. The relative free volume fraction determined before water exposure was compared to the equilibrium water uptake of each coating and a correlation was found. The relative free-volume fraction of the epoxy coatings decreased slightly after water exposure. This decrease is contrary to the free volume theory of plasticization, but is consistent with the antiplasticization process. Larger decreases in the relative free volume fraction were sought by repeating the water uptake experiments with nitrobenzene which in the bulk, liquid form quenches ortho-positronium (o-Ps). Since the o-Ps lifetime remained approximately constant and the o-Ps intensity decreased after nitrobenzene absorption, nitrcbenzene is believed to be inhibiting the formation of o-Ps in the epoxy free volume cavities. Larger decreases in the relative free volume fractions were found after nitrobenzene exposure than after water exposure. These larger decreases are due to the fact that nitrobenzene is a better inhibitor of o-Ps formation than water in the epoxy free volume cavities. Larger volume fractions of nitrobenzene were absorbed by the coatings than water and were interpreted to be due to interactions between nitrobenzene and the epoxies. © 1993 John Wiley & Sons, Inc.  相似文献   

13.
采用开环聚合法和自由基聚合法合成了生物可降解嵌段共聚物OSM1-PCLA-PEG-PCLA-OSM1, 并对其进行了结构表征. 采用荧光分光光度计和激光粒度仪对共聚物溶液临界胶束浓度(CMC)和粒径大小及分布进行了考察, 研究了温度和pH对共聚物胶束形成的影响. 相转变过程研究结果表明, 共聚物溶液具有pH和温度双重敏感性. 共聚物溶液在一定温度和pH条件下可发生溶液-凝胶相转变.  相似文献   

14.
杨海洋 《高分子科学》2013,31(2):263-274
The sol-gel transition of methylcellulose (MC) solutions in the presence of ortho-methoxycinnamic acid (OMCA) or cetyltrimethylammonium bromide (CTAB) and in the coexistence of OMCA and CTAB was determined by the rheological measurement. It has been found that the sol-gel transition temperature of MC solutions increases linearly with the concentration of either OMCA or CTAB in solution, respectively. However, in the coexistence of OMCA and CTAB, the sol-gel transition temperature of MC solutions remains invariable, independent of the concentration of CTAB in solution. The experimental results show that OMCA has priority to adsorb on the methyl group of MC chains to form polymer-bound aggregates. In particular, these aggregates inhibit the hydrophobic interaction between CTAB and the methyl group of MC chains completely. Taking into account the fact that OMCA is almost insoluble in MC-free solutions but dissolves very well in aqueous MC solutions, we propose the formation of the core-shell architecture prompted by OMCA and the methyl group of MC chains, with the methyl group of MC chains serving as the core and the self-assembly of OMCA molecules serving as the shell. Obviously, the formation of the core-shell structure increases the solubility of OMCA, improves the stability of methyl groups of MC chains at high temperatures and inhibits the hydrophobic interaction between CTAB and the methyl group of MC chains in solution. The abnormal behavior relating to the sol-gel transition of MC solutions in the presence of OMCA or in the coexistence of OMCA and CTAB is therefore explained. Upon UV irradiation, the sol-gel transition temperature of MC solutions in the presence of OMCA, or in the coexistence of OMCA and CTAB, decreases notably. However, the dependence of the sol-gel transition temperature of MC solutions as a function of OMCA concentration, or CTAB concentration in the presence of OMCA, does not change after UV irradiation.  相似文献   

15.
The size and shape of free-volume holes available in membrane materials control the rate of gas diffusion and its permeability. Based on this principle, two segmented thermo-sensitive polyurethane (TSPU) membranes with functional gates, i.e. the ability to sense and respond to external thermo-stimuli, were synthesized and used for water vapor controllable permeation. Differential scanning calorimetry (DSC), positron annihilation lifetimes (PAL), water swelling and water vapor permeability (WVP) were used to evaluate how the structure of the polyurethane (PU) and the temperature influence the free-volume holes size and the water vapor permeability (WVP) of the PU membranes. DSC study reveals that TSPU with a glass transition or a crystalline transition reversible phase shows an obvious phase-separated structure and a phase transition temperature (defined as switch temperature, Ts). PAL study indicates that the free-volume holes size of TSPU is closely related to the Ts. When the temperature is higher than the Ts, the ortho-positronium (o-Ps) lifetime (τ3) and the average radius (R) of free-volume holes of TSPU membrane increase dramatically. As a result, the WVP of TSPU membrane shows a dramatic increase. Additionally, the water swelling and the WVP of TSPU membrane are found to depend on the inner structure of the polymer, and they also give different responses to temperature variation. When the temperature is higher than the Ts, there is a significant increase of WVP from 3.80 kg/m2 day to 7.63 kg/m2 day for TSPU(a) and from 4.30 kg/m2 day to 8.58 kg/m2 day for TSPU(b), respectively. Phase transition accompanying significant changes in free-volume holes size and WVP can be used to develop “smart membranes” with functional gates and controllable gas permeation.  相似文献   

16.
Viscoelastic experiments were performed to study the influence of nonsolvent and temperature on critical viscoelastic behaviors of ternary polyacrylonitrile (PAN) solutions around the sol-gel threshold. The dynamic critical parameters around the sol-gel threshold were determined using dynamic rheometer. The sol-gel transition takes place at a critical gel temperature at which the scaling law of G′(ω) ∼ G″(ω) ∝ ωn holds, allowing an accurate determination of the critical gel temperature by means of the frequency independence of the loss tangent. Although the gel points of PAN solutions increase with increasing H2O content, the results show that the scaling exponent n at the gel point is found to be universal for all ternary PAN solutions, which is independent of temperature and H2O content, indicating the similarity of the fractal structure in the critical PAN gels. The gelation of ternary PAN solutions induced by adding a nonsolvent and by decreasing the temperature is demonstrated to be a thermoreversible process, which implies that the PAN gels are physical gels. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 46: 2637–2643, 2008  相似文献   

17.
Free volumes in thermotropic side-chain liquid-crystalline polymers were probed by positron annihilation technique. Lifetime spectra of positrons were measured in the temperature range between 130 and −60°C in cooling. For a nematic liquid-crystalline polymer (polyacrylate), the lifetime of ortho-positronium (τ3) was decreased with decreasing temperature above the glass transition temperature (Tg, 21°C) with larger temperature coefficient than that below Tg. The intensity of ortho-positronium (I3) was constant above Tg. These facts mean that the size of the free-volume holes decreased with the decreasing the temperature but the concentration was almost constant in nematic phase. For a smectic liquid-crystalline polymer (poly(p-methylstyrene) derivative), a discontinuous decrease in the value of τ3 and that of I3 were observed at 107°C, which was the transition temperature from smectic to crystalline phase. Such discontinuous changes were not observed for the polyacrylate specimen. This difference was considered to be attributed to the higher-ordered structure of the smectic phase. © 1996 John Wiley & Sons, Inc.  相似文献   

18.
Positron annihilation measurements as a function of temperature and time have been carried out on a poly(butadiene). The measurements were performed at several temperature points from 14 to 225 K. The measurement time was several hours to four days. The analysis of data shows the following features:
(i) the value of τ3 does not depend on the rate of cooling or time,
(ii) the value of I3 depends on the rate of cooling and the history of thermal treatment,
(iii) the dependence of I3 on time can be described by Debye function. But the rise in I3 is observed at very low temperatures,
(iv) the I3 decays to value of I3 observed during very slow cooling.

Article Outline

1. Introduction
2. Experiments
3. Results
4. Discussion
5. Conclusions
6. Uncited Reference
Acknowledgements
References

1. Introduction

If a glass is formed by rapid cooling of a super-cooled liquid to a temperature below the glass–liquid transition temperature, Tg, its properties will not be static, but will relax toward values characteristic of the corresponding “equilibrium” supercooled liquid as extrapolated from above to below Tg. This process named as structural relaxation or “physical aging” is of great practical importance because of its relevance to the designing and engineering of amorphous materials with desired properties. The relaxation property and transport phenomena of disordered polymers can be explained within the free-volume concept (Ferry, 1980). However, an unsettled problem is a way of quantifying the free-volume properties, such as the free-volume fraction, the average and the distribution of the free-volume size. In the last decade, the positron annihilation lifetime spectroscopy (PALS) technique has been recognised as a useful method to detect atomic scale free-volume holes of polymers ( Schrader and Jean, 1988). This technique involves using a positron source, mostly 22Na, to emit positrons into the sample. But these positrons and the accompanying gamma–quanta have sufficient energy (average positron energy 200 keV, gamma 890 keV) to induce radiation effects, and the positron probe can thus affect the sample being investigated during PALS experiments.The basic assumption of positron annihilation lifetime spectroscopy (PALS) data interpretation in terms of the free-volume concept is the proportionality of the intensity of long-lived ortho-positronium (o-Ps) component, I3, to the concentration of free-volume holes (Kobayashi et al., 1989). However, there are different findings regarding the influence of external factors on the “true” intrinsic value of I3. Its variation with the measurement time is regarded as a manifestation of the relaxation of free-volume fraction. On the other hand, the decrease in I3 with PALS measurement time is related to the activity of the positron source and the chemical processes in the positron spur, e.g., formation of free radicals. There are PALS measurements on semi-crystalline samples (Suzuki et al., 1996), observing the I3 increase with elapsed time when the temperature of the sample is below Tg.All these reports indicate that the o-Ps formation in polymers is more complicated and the basic assumption of PALS interpretation may be questionable.In this work, PALS results will be presented on the amorphous cistrans-1,4-poly(butadiene), cistrans-1,4-PBD, in a wide temperature range from 14 to 350 K. The aim of this paper is the study of the influence of temperature, time and sample history on the intensity I3, life time of o-Ps, τ3, as well as the S-parameter from Doppler broadening measurements.

2. Experiments

The PALS experiments were conducted using a conventional fast–fast coincidence system having a time resolution of ca. 320 ps (FWHM). Cistrans-1,4-PBD has a molecular weight of Mw = 2 × 104, the glass transition temperature Tg = 178 K (Zorn et al., 1995). The isomer composition was 41% cis, 52% trans and 7% vinyl form. This isomer composition was chosen to avoid a crystallisation process on the PBD sample (Zorn et al., 1995).The positron source, consisting of 2 MBq 22N a sealed between two 3.5 μm Ni foils, was sandwiched between polymer discs, each of about 3 mm thick and with a diameter of 10 mm. At a chosen temperature, each spectrum was accumulated for 1 h, resulting in a total number of counts of about 1.14 mil. At least, two such spectra were recorded at each temperature point.The 22Na source–sample assembly was mounted on a closed cycle helium gas refrigerator. The assembly was kept in a rotary pump vacuum of about 4 Pa. Automatic temperature regulation was used during all the measurements and the temperature was controlled within ±1 K. Several different temperature scans on the specimens were performed. The first sequence (heating) was the following: I3, τ3 were first evaluated at room temperature of 300 K immediately after the source installation. Then, fast cooling to the temperature of 40 K at a rate 4 K/min was performed and the temperature increased in steps of 10 K. The second sequence (cooling) started at 300 K, then the temperature decreased to 14 K in steps of 10 K.For the PALS measurement as a function of time, the PBD was annealed in the chamber at 300 K for several hours, then cooled to the measurement temperature and the measurement began immediately.The positron life-time spectra were measured as a function of the elapsed time at 14 different temperature points below and above Tg.The PALS data were also accumulated during heating of the samples to 300 K and cooling of PBD to chosen temperature below 300 K. The total irradiation time of 1080 h was divided between PALS and calibration (Bi) measurements. To clearly describe the thermal history of the experiment, the time dependence of I3 and τ3 is shown in Fig. 1 and Fig. 2, respectively. The values of I3 and τ3 at room temperature were the same despite the long irradiation time and complicated thermal history. This indicates that a possible irradiation damage does not influence the annihilation observables.  相似文献   

19.
用正电子湮没寿命谱(PALS)研究了温度和PEG含量对以聚己二酸丁二醇酯(PBA)为软段的聚酯聚氨酯的自由体积特性和透气性的影响.实验结果表明,自由体积空洞的体积随着温度的升高而增加,分布变宽,导致透气性增大.不同PEG含量的聚酯聚氨酯PBA-10,PBA-15和PBA-20的水汽渗透系数(WVP)基本相同.结果表明,在这类聚氨酯中,影响透气性的主要因素不仅是自由体积,而且与材料的亲水性有关.  相似文献   

20.
A series of polyurethane films based on hard segments consisting of toluene diisocyanate and 1,4-butanediol and different soft segments consisting of hydroxyl terminated polybutadiene, hydroxyl terminated polybutadiene/styrene and hydroxyl terminated polybutadiene/acrylonitrile were synthesized by solution polymerization separately. Positron annihilation lifetimes were measured at room temperature for all samples studied. We found that both the free volume size and fractional free-volume decreased with the increase of hard segment content. On the other hand, direct relationship between the gas permeability and the free-volume has been established based on the free-volume parameters and gas diffusivity measured. Experimental results revealed that the free-volume plays an important role in determining the gas permeability.  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