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 473 毫秒
1.
Fundamental correlations between the molar mass and the solution viscosity (dimethylformamide (DMF)/1% di-n-butylamine (DBA)) are acquired and discussed on selected polyesterurethanes (PUR-Es) and polyetherurethanes (PUR-Et) with NCO/OH ratios ranging from 955 to 1 030. Static light scattering measurements reveal mass average molar masses in the range 20 000–95 000 g mol–1; membrane osmometry reveals number average molar masses in the range 15 000–50 000 g mol–1. Comparative tests with the analytical ultracentrifuge reveal the agreement with respect to the molar mass and the A2-values; additionally this measurements confirm unimodal Schulz-Zimm molar mass distributions with an average sedimentation constant of S = 1.1 sved (1 sved = 10–13 s). The results of light scattering und ultracentrifuge measurements show that the solvent mixture DMF/1 wt.-% DBA is well suitable for molar mass measurements and solution viscosity measurements. The linear light scattering Zimm diagram and the sedimentation run exclude microgels and higher molecular associates (for example allophanate and biuret crosslinkage). Because of the good correlation between the determined molar masses and the intrinsic viscosities [η] and the viscosities of the concentrated solutions it was possible to establish a calibration function [η] = K Ma. The exponent a = 0.6–0.9 of the Mark-Houwink-relation indicates that the polymer chains, in diluted solution, are isolated coiled molecules. In concentrated solutions the chains are partly crosslinked, depending on the molar mass. The correlation between the molar mass and the viscosity (Fox-relation) shows an unsteadiness at Mw = 55 000 g mol–1 (critical molar mass). Above the critical molar mass the solution viscosity increases with the molar mass with an exponent of ϵ = 3.4. Below the critical molar mass of Mw = 55 000 g mol–1 the solution viscosity increases straight proportional with the molar mass (ϵ = 1).  相似文献   

2.
This study presents the development of an improved technique for viscosity measurements under high pressure. The apparatus is based on the principle of the falling ball viscometer, implemented in a high-pressure autoclave fitted with visualisation windows. The originality here is that the balls fall through a tube open at both ends with a diameter slightly greater than that of the balls, allowing a simplified modelling and numerical simulation. A numerical approach has been used for viscosity determination. Calculations have been made with COMSOL Multiphysics® with the laminar Navier-Stokes model for Newtonian mixtures. It includes the specific hydrodynamic effects without the need for a calibration fluid. However, validation experiments were carried out at atmospheric pressure with dimethylsulfoxide (DMSO) at 298, 308 and 318 K and with cocoa butter at 313 and 353 K, with values of viscosity in the range from 1.4 to 45.4 mPa s. Comparative measurements with literature data have been conducted with cocoa butter saturated with carbon dioxide at 313 and 353 K and for pressures ranging from 0.1 to 25 MPa. At 313 K, viscosity varies from 45.4 mPa s to 3.1 mPa s while at 353 K it varies from 12.4 to 1.9 mPa s. For both isotherms tested, within the range 0-15 MPa, the higher the CO2 dissolution in the cocoa butter, the lower the viscosity. However, this decrease in viscosity is more pronounced at the lowest temperature. Above 15 MPa the CO2 dissolution effect on viscosity becomes insignificant, i.e. within the experimental error, due to a counter effect linked with the high hydrostatic pressure. Furthermore, the limits of use of this method have been determined. This technique is revealed as reliable and can therefore be used with other binary systems.  相似文献   

3.
Solid polymer electrolytes comprising blends of poly(vinyl chloride) (PVC) and poly(methyl methacrylate) (PMMA) as host polymers and lithium bis(trifluoromethanesulfonyl) imide (LiTFSI) as dopant salt were prepared by solution‐casting technique. The ionic conductivity and dielectric behavior were investigated by using AC‐impedance spectroscopy in the temperature range of 298–353 K. The highest ionic conductivity of (1.11 ± 0.09)×10?6 S cm?1 is obtained at room temperature. The temperature dependence of ionic conductivity plots showed that these polymer blend electrolytes obey Arrhenius behavior. Conductivity–frequency dependence, dielectric relaxation, and dielectric moduli formalism were also further discussed. Apart from that, the structural characteristic of the polymer blend electrolytes was characterized by means of horizontal attenuated total reflectance–Fourier transform infrared (HATR–FTIR) spectroscopy. HATR–FTIR spectra divulged the interaction between PMMA, PVC, and LiTFSI. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

4.
In this study, the changes in the rheological curves of polyvinyl chloride (PVC) plastisols with increasing storage time and the factors affecting these changes were studied. The results show that with increasing storage time, all the “viscosity–temperature” and “viscosity–time” rheological curves of PVC plastisols exhibit nonnormal distribution change trends, that is, the viscosity first decreases, and then changes from slow increasing to rapid increasing, forming a shoulder peak, reaches to the maximum value and gradually decreases. With increasing storage time, the complex viscosities of PVC plastisols increased generally in the first, the second, and the fourth stages, and the gelation process shortened in the third stage. The first and second stages of the viscosity changes reflect the “time–temperature” equivalence principle of PVC plastisol in suspension stage. However, the maximum viscosity of PVC plastisol corresponding to temperature max does not change with increasing storage time.  相似文献   

5.
PVC melts are predicted to be homogeneous with single molecules as the stable flow units (monomolecular melts) at corresponding values of high temperatures and/or high shear stresses. Under these conditions, it is found that the zero shear viscosity in simple shearing flow of rigid compounds depends on the average molecular weight by weight to the 3.5 power for molecular weights between 24,000 and 100,000. All data measured under conditions where monomolecular melts are predicted fall on a master curve of reduced viscosity versus reduced shear rate when a relaxation time proportional to η0/c2T is used. It is, therefore, concluded that monomolecular melts of PVC compounds follow the same structure–viscosity relations as found for other linear melts in viscometric flow.  相似文献   

6.
A rheological investigation of the saturated fatty acids as lubricants for PVC was made in a Sieglaff-McKelvey rheometer over a wide temperature range. Results reveal that with an increase in carbon chain, fatty acids became more effective in reducing the melt viscosity of PVC. It was also discovered that the maximum viscosity variation due to these materials was observed at 350°F. At test temperatures of 400°F and above, no major viscosity differences were observed. Furthermore, the changes in the melt viscosity were more pronounced at shear rates less than 200 sec?1. The usefulness of the measured viscosity variations at the critical test temperature of 350°F was shown by extrusion. An excellent correlation between the apparent melt viscosity and extrusion melt temperature was found. The criticalness of the test temperature in assessing the effect of lubricants on PVC flow is discussed.  相似文献   

7.
引 言松节油的主要组分是α 和 β 蒎烯 ,含量约91%左右[1] .蒎烯在温度 35 3~ 4 33K、压力 2 0~7 0MPa下 ,经催化加氢制得蒎烷 .蒎烷是一种具有温和松针样气息的透明状液体 ,有重要的工业用途[2 ] .蒎烷经氧化、还原及裂解等反应可制备芳樟醇 ,芳樟醇是重要的香料 ,同时也  相似文献   

8.
A poly(vinyl chloride) (PVC) matrix of final particles is investigated by means of quasi-elastic light scattering (QELS) and high resolution porosimetry (HRP). It is demonstrated that the QELS method is a powerful tool for the detection of various higher organized structures in particle interiors, which in all likelihood will be a difficult-to-process fraction of the resin. Measurements of common PVC grades revealed that the QELS records are usually composed of three distinctly separated size distributions, which were assigned to the size distribution of single macromolecules dissolved (maximum at ca. 10 nm), associates of macromolecules (primary particles, the maximum at ca. 100–200 nm), and microglobules (the maximum at ca. 0.8–1 (μm)). The properties of primary particles are presented, as well as the correlation between QELS and HRP records. Based on this correlation, it can be suggested that the pore size distribution is in fact a morphological fingerprint of individual PVC resins. Some consequences of the results obtained for the evaluation of the processsing properties of PVC are discussed.  相似文献   

9.
Ten commercial linear low-density polyethylenes (LLDPE) were characterized by solution viscosity, size exclusion chromatography, SEC, and 13C nuclear magnetic resonance. The resins were copolymers of ethylene with butene, hexene, or octene. They were prepared in gas phase (with narrow or very broad molecular weight distribution), or in solution. The macromolecules were found to be linear. For all but the very broad molecular weight distribution resins the average comonomer sequence length was found to be 1; in the other case diad formation was observed. The weight average molecular weights calculated from SEC, and intrinsic viscosities agreed quite well. Mechanical degradation of LLDPE was observed during the solution viscosity measurements.  相似文献   

10.
The influence of relatively nontoxic thermal stabilizers including different types of organic calcium complex (Ca/Zn system of liquid stabilizers) and organotin on photodegradation of poly(vinyl chloride) (PVC) was investigated by color difference measurement, viscosity‐average molecular weight determination, UV–vis spectroscopy, Fourier transform infrared (FTIR), and thermogravimetric (TG) analysis. PVC films containing relatively nontoxic thermal stabilizers were prepared by solution casting and then exposed to xenon‐arc light source with the irradiance of 0.51 W/(m2·nm) at 65°C. Two major chain processes, photodehydrochlorination and photo‐oxidation, occur simultaneously during photodegradation of PVC. It has been confirmed by both color difference and UV–vis spectra that during the former 300 h of irradiation, organic calcium complex stabilizers retard photodehydrochlorination as well as initial color development of PVC films while organotin stabilizers remarkably accelerate photodehydrochlorination after 100 h. Relative carbonyl index (RCI) is first introduced to the analysis of FTIR results, which implies that organotin has a better ability to inhibit photo‐oxidation than organic calcium complex and ensures longer stabilization time. The antioxidation of mercaptan organotin has been observed because it is an effective decomposer of peroxides and hydroperoxides. TG analysis reveals that some unstable structures generated due to the irradiation of ultraviolet can easily split away off from PVC macromolecular backbones under relatively low temperature. POLYM. ENG. SCI., 2010. © 2009 Society of Plastics Engineers  相似文献   

11.
The influence of UV absorber (Chimassorb81) on the photodegradation mechanism of different average degrees of polymerization (DP ) of poly (vinyl chloride) (PVC) with UV‐irradiation time was investigated by viscosity‐average molecular weight determination, UV‐vis spectroscopy, FTIR, contact angle measurement, and scanning electron microscopy (SEM). The PVC films with different DP (1000 and 3000), which contained 0.3 or 0.5 phr Chimassorb81, were prepared by solution casting. It was carried out by exposing specimens to xenon‐arc light source with a spectral irradiance of 0.68 W/(m2 nm) at 63°C. It is found that the Chimassorb81 is efficient photostabilizer for PVC with different DP . Although the Chimassorb81 delays the photodegradation of PVC, it does not influence the photodegradation mechanisms of PVC with different DP . The main photodegradation reaction for the lower DP of PVC is dehydrochlorination in the initial stage of UV‐irradiation, and then the crosslinking and chain scission reactions occurred after long irradiation. However, the main reaction of the higher DP of PVC is not dehydrochlorination but crosslinking and chain scission in the initial stage of UV‐irradiation. The results of carbonyl index, C? Cl index, contact angle measurement, and SEM also show that the photostability of Chimassorb81 is more effective for the higher DP of PVC, especially in the presence of higher concentration of Chimassorb81. POLYM. ENG. SCI., 47:1480–1490, 2007. © 2007 Society of Plastics Engineers  相似文献   

12.
A series of hollow‐fiber membranes was produced by the dry–wet spinning method from PEEKWC, a modified poly(ether ether ketone) with good mechanical, thermal, and chemical resistance. The fibers were prepared under different spinning conditions, varying the following spinning parameters: polymer concentration in the spinning solution, height of the air gap, and bore fluid composition. The effect of these parameters on the water permeability, the rejection of macromolecules (using dextrane with an average molecular weight of 68,800 g/mol), and the morphology of the membranes was studied. The results were also correlated to the viscosity of the spinning solution and to the ternary polymer/solvent/nonsolvent phase diagram. The morphology of the cross section and internal and external surfaces of the hollow fibers were analyzed using scanning electron microscopy (SEM). All membranes were shown to have a fingerlike void structure and a skin layer, depending on the spinning conditions, varying from (apparently) dense to porous. Pore size measurements by the bubble‐point method showed maximum pore sizes ranging from 0.3 to 2 μm. Permeability varied from 300 to 1000 L/(h?1 m?2 bar) and rejection to the dextrane from 10 to 78%. The viscosity of polymer solutions was in the range of 0.2 to 3.5 Pa s. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 91: 841–853, 2004  相似文献   

13.
This paper deals with the main concepts of the modern theory of poly(vinyl chloride) (PVC) degradation, which embrace the key problems concerning the chemical structure and the content of anomalous groups in PVC, their influence on the thermal stability of polymer products, the kinetics of HCL elimination. In contrast to the universally recognized ~-chloroallyl activation of the process of PVC degradation, a conception of the prime role of oxygen-containing chloroallyl groups of type ~C(O)? CH?CH? CHCl~ (CAG) has been developed. It has been shown that PVC real macromolecules contain approximately 10?4 mole/PVC mole oxygen-containing ~C(O)? CH?CH? CHCl~ groups determining PVC low thermal stability. It has been found that the monomer purity, the presence of oxygen in the reaction area, the temperature of vinyl chloride polymerization, etc., considerably affect CAG content in PVC macromolecules.  相似文献   

14.
Thermal stability of poly(vinyl chloride) (PVC) samples polymerized under a temperature trajectory was studied from the point of view of morphological and microstructural characteristics. The results are compared with those of the PVC samples obtained by polymerization at constant temperature having the same K value. The Brabender® plastograph data indicated that the final PVC synthesized with the temperature trajectory showed lower fusion time and higher thermal stability time. The nonisothermal condition also increased the degree of fusion of the final PVC resin, reflecting lower temperature/time required to process it. It was found that the thermal stability of nonisothermally produced PVC as characterized by dehydrochlorination rate decreased (improved) with the increasing monomer conversion until a minimum value was reached that corresponded to the conversion at the pressure drop. However, the dehydrochlorination rate remains almost constant with conversion for an isothermal grade PVC resin. Although the evolution of the number of internal double bonds as well as extent of discoloration of PVC with conversion shows a decreasing trend, the labile chlorine concentration exhibits a maximum at early conversion. The reason for the former can be explained by the temperature dependence of reactions forming defect structures, which are kinetically controlled and thus favored at higher temperatures. The latter, however, can be explained because of the increasing importance of transfer reactions to polymer with increasing polymer concentration. Finally, the results from differential thermogravimetry verify an improvement in thermal stability of the final PVC prepared by using a temperature trajectory during vinyl chloride monomer suspension polymerization. J. VINYL ADDIT. TECHNOL., 23:259–266, 2017. © 2015 Society of Plastics Engineers  相似文献   

15.
The recent studies deal with a diblock copolymer, polystyrene–poly(ethylene oxide). Infrared spectroscopy, proton resonance spectroscopy (1H‐NMR), and laser light scattering techniques have been used to characterize the polymer. It has been concluded that the sample investigated is diblock copolymer polystyrene–poly(ethylene oxide) having molecular mass 1.656 × 104 g/mol and blocks ratio 1 : 2. The micellization behavior is explored through 1H‐NMR, laser light scattering, light absorption, surface tension, and conductance and viscosity measurements. The results conclude that the critical micelles concentration of copolymer is 0.0951 g/dL at 25°C. It has been observed that the surface tension of solution decreases with the temperature and its impact is maxima in dilute concentration region. In addition, new methodologies have been introduced to get accurate critical micelles concentration and critical micelles temperature. © 2010 Wiley Periodicals, Inc., J Appl Polym Sci, 2010  相似文献   

16.
Linear and nonlinear rheological properties of poly(vinyl chloride) (PVC)‐poly(n‐butyl acrylate)‐PVC triblocks of different compositions, obtained by single electron transfer‐degenerative chain transfer living radical polymerization, are investigated, focusing on the effect of crystallites. Dynamic mechanical thermal analysis results show the existence of two glass transition temperatures, denoting microphase segregation. However, rather than phase separation, it is the presence of two types of crystals that melt at Tm1 = 127 ± 0.8°C and Tm2 = 185 ± 2°C, respectively, the factor that determines the rheological response of the copolymers. To the difference with PVC homopolymers, extrusion flow measurements at very low temperatures (T = 100°C) are possible with the copolymers. A change in the viscosity‐temperature dependence is observed below and above the lowest melting temperature. Notwithstanding the microphase separation and the presence of crystallites, experiments carried out in conditions similar to industrial processing reveal a remarkable viscosity reduction for our copolymers with respect to PVC obtained by single electron transfer‐degenerative chain transfer living radical polymerization, conventional PVC, and PVC/[diethyl‐(2‐ethylhexyl) phthalate] compounds. Extrudates free of surface instabilities are obtained at low extrusion temperatures, such as 90–100°C. J. VINYL ADDIT. TECHNOL., 21:24–32, 2015. © 2014 Society of Plastics Engineers  相似文献   

17.
通过BrookField DV-Ⅲ Ultra可编程旋转流变仪实时测定了以偶氮二异丁腈为引发剂,二甲基亚砜为溶剂,丙烯腈-衣康酸二元溶液共聚合体系的零剪切黏度变化。同时设计相同条件的封管实验测定转化率,并利用Arrhenius-Frenkel-Eyring方程导出了适合描述丙烯腈溶液共聚合体系的零剪切黏度随固含量、温度及黏均分子量变化的经验关联式。结果表明,各工艺条件对体系黏度都有显著影响,原料配比细微的差别将导致最终原液黏度相差数百帕秒以上,调控黏度时需综合考虑。值得注意的是,所有实验条件下反应2~3 h后,体系黏度激增1000倍以上,即从约2 mPa·s的初始黏度迅速增加至2~3 Pa·s;而当总单体含量达到28%时体系产生凝胶效应,反应一定时间后黏度急剧增长。另外发现,当体系达到某一临界固含量时,溶液中大分子发生链缠结,固含量微小的增量就会导致体系黏度剧增。  相似文献   

18.
Summary Grafted copolymer of poly(vinyl chloride) (PVC) with N-isopropylacrylamide (NIPAAm) was prepared by radiation-grafting method using γ-ray source. NIPAAm was graft polymerized from its aqueous solution onto PVC tubes by preirradiation method, all samples were exposed in the presence of air at room temperature to 60Co. Conditions for achieving maximum grafting yield were observed between 0.5 and 1 moldm-3 of monomer concentration, pre-irradiation dose of PVC from 5 to 110 kGy, and reaction temperature of 323 and 333 K. Characterization of the grafted copolymer was conducted by various methods: FTIR-ATR, TGA, and SEM. The temperature-responsive behavior of grafted copolymer was studied by swelling at various temperatures and pH 6.8.  相似文献   

19.
The processing and structural effects occurring in a poly(vinyl chloride) (PVC) compound processed in a plastograph, at the time at which the maximum value of the torque (point X on the torque curve) was reached, were studied. The unplasticized PVC compound was processed at various temperatures (150–200°C) with a kneader operating at rotors speeds of 5–40 min?1. The changes in the temperature of the compound due to kneading and its influence on the progress of PVC gelation were analyzed. The gelation progress was studied with differential scanning calorimetry, rheological measurements, and scanning electron microscopy. The maximum value of the torque occurred within the temperature range (170–180°C) of the PVC compound, and it was independent of the adjusted temperature of the mixing chamber. The processing of the PVC compound in the kneader, with high mechanical charges and a comparatively low adjusted temperature (150–170°C), spontaneously generated an effect of self‐heating, resulting in an increasing temperature of the compound. Despite the surprisingly high degree of gelation (80–98%) of the PVC compound processed under these conditions (as determined by differential scanning calorimetry), the scanning electron microscopy observations indicated significant nonhomogeneity of the fracture surface. Consequently, the kneaded PVC compound at point X of the torque curve could be treated as a two‐phase system composed of a liquid, amorphous phase and elements of a grain structure. The transformation of both phases (particularly the quantitative and qualitative changes), which was significantly dependent on the PVC temperature and the shear rate applied by the rotors, determined the state of the melted compound at this point of the torque curve. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

20.
Poly(vinyl chloride) (PVC) is one of the most common polymers used in the water treatment industry due to outstanding hydrophobicity and mechanical strength. Generating eco‐friendly membranes derived from natural polymers has gained attention, particularly for water purification and producing potable water. In this study, nonwoven mats were prepared by electrospinning polymer solutions. Mats with a tailorable hydrophilicity were prepared by electrospinning solution mixtures containing PVC and an eco‐friendly, hydrophilic natural polymer: soy protein. As the viscosity of the solution decreased, the average fiber diameter, and average pore surface area reduced. However, when the PVC concentration remained constant and the soy protein concentration increased, the viscosity decreased and average fiber diameter became reduced, while the average pore diameter remained relatively constant. The mats with volumetric ratios of PVC:soy protein of 85:15 and 80:20 displayed optimal characteristics suitable for mat fabrication based on the fiber diameter and average pore surface area. © 2018 American Institute of Chemical Engineers AIChE J, 64: 2737–2744, 2018  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