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.
Typical terephthalate polyesters such as poly(butylene terephthalate) (PBT) and poly(ethylene terephthalate) (PET) were characterized by pyrolysis-gas chromatography (Py-GC) in the presence of tetramethylammonium hydroxide in terms of the cross-linking structure formed during their material recycling. In the pyrograms of PBT and PET thermally treated at 270 °C for 1 h, which were prepared as model polymers containing cross-linking structures, an additional peak was commonly observed as well as the main reactive pyrolysis products for the original polyesters such as dimethyl terephthalate. Based on the observed spectra obtained by Py-GC/mass spectrometry and Py-GC/Fourier transform infrared spectrometry measurements, this peak was assigned to the product reflecting a biphenyl-type cross-linking structure. Furthermore, in the pyrograms of kneaded PBT and PET samples also at 270 °C for a total of 1 h, which were prepared to simulate material recycling, the same peak for the cross-linking structure was also observed, although its intensity was slightly lower than that in the samples thermally treated in air. This fact verified that the biphenyl-type cross-linking structure would be considerably formed during the recycling of PBT and PET, which might in turn contribute to the deteriorated properties of the recycled materials from waste polyesters. Moreover, difference in the formation of the cross-linking between PBT and PET is discussed on the basis of the observed results.  相似文献   

2.
The Newtonian and non-Newtonian melt viscosities of bisphenol A polycarbonate (PC) at 280°C were treated according to the generalized multivariable power function, where the average molecular weights, polydispersity degree and branching degree are considered as variables. The shear rate was also considered as a variable for non-Newtonian conditions. In the same way, the melt fluidity was treated as a multivariable power function. It has been found that the same melt flow properties of polymer can be obtained by an appropriate combination of Newtonian melt viscosity (being a function of molecular weight) and long chain branching. The experimental data on PC agree with the theoretical approach of Bucche and Graessley.  相似文献   

3.
Miscibility of bisphenol-A polycarbonate (PC) and syndiotactic polymethyle methacrylate (sPMMA) blends was investigated by differential scanning calorimetry (DSC) and small-angle light scattering. Cloud-point measurements indicated the existence of both a lower critical solution temperature and an upper critical solution temperature, forming an immiscibility loop. This immiscibility gap was observed for PC blends with sPMMA of various molecular weights ranging from 8,300 to 55,000. The DSC study on solvent-cast and coprecipitated PC/sPMMA blends from tetrahydrofuran solutions showed a single glass transition, shifting regularly with composition. The annealing of the 50/50 composition within the immiscibility loop exhibited dual glass transitions, but the system reverted to a single phase upon annealing above the loop. Phase dissolution took place during annealing above the loop, followed by thermoxidative branching (cross-linking) reaction. Dry pellets of PC and sPMMA were melt mixed above the loop in a Mini-Max mixer/molder; these molded blends exhibited a single phase. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
铁皮石斛的裂解气相色谱指纹图谱及其系统聚类分析   总被引:4,自引:0,他引:4  
王丽丽  王聪  潘再法  孙法 《色谱》2008,26(5):613-617
采用裂解气相色谱/质谱法(Py-GC/MS)测定了10种不同产地的铁皮石斛并结合系统聚类分析法比较了这些铁皮石斛的指纹图谱,采用释放气体分析法考察了裂解温度对指纹图谱的影响。结果表明,0.4 mg样品在450 ℃下可瞬间裂解,10种样品的指纹图谱具有相似性,且重现性好;采用系统聚类分析能区别不同产地的样品。本法快速、简便、准确,不失为药材质量控制的良好方法。  相似文献   

5.
In the present work, we use the viscoelastic moduli of a large number of industrially available polyethylenes in order to evaluate/test some of the previously proposed correlations between levels of long chain branching and polydispersity with the rheological properties. These correlations together with some new ones can be used to correct for the effects of polydispersity or long chain branching in order to assess the effect of these two molecular features on the rheological properties independently. The effects of short and long chain branching are studied providing a methodology to detect rheologically levels of short and long chain branching.  相似文献   

6.
The branching and cross-linking of poly(ethylene terephthalate) were investigated using two chain extenders: glycidyl methacrylate-styrene copolymer (GS) and poly(butylene terephthalate)-GS (PBT-GS) in order to improve the melt viscosity and melt strength of poly(ethylene terephthalate). An obvious increase in torque evolution associated with chain extending, branching and cross-linking was observed during the process. The properties of modified poly(ethylene terephthalate) were characterized by intrinsic viscosity and insoluble content measurements, rheological and thermal analysis. The intrinsic viscosity and rheological properties of modified PET were improved significantly when using PBT-GS, indicating that PBT-GS should be a better chain extender. Good foaming of poly(ethylene terephthalate) materials were obtained using supercritical CO2 as blowing agent. The average cell diameter and cell density were 61 μm and 1.8 × 108 cells/cm3, respectively.  相似文献   

7.
The carbon-13 spin-lattice relaxation times T1 of the crystalline components of four solid ethylene-octene copolymers have been studied as a function of thermal history, branching number, and branching distribution. Slowly cooled samples (1 deg/min from melt to room temperature) exhibited similar or longer T1s with respect to the same sample quench cooled (from the melt into 20°C water). The greater the degree of branching and the more homogeneous the branching distribution, the shorter were the observed crystal lattice T1s. Differences of up to a factor of 3 in T1 were observed for the same sample undergoing the two thermal treatments. Different degrees of branching homogeneity (for the same total number of branches) resulted in differences approaching a factor of 7 for samples with the same thermal history. These variations were attributed to the differing effects of side-chain disruption of the crystal lattice.  相似文献   

8.
The determination of end groups in polystyrenes (PSts) polymerized anionically with n-butyllithium as the initiator was carried out by pyrolysis-gas chromatography (Py-GC). The relative proportions of the end groups decrease with increasing molecular weight (MW). Consequently, the peaks which reflect the structure of the end groups can be distinguished by comparing pyrograms of samples with different MW. By comparing the intensities of these peaks relative to total intensities of all the peaks in the pyrogram, the numberaverage molecular weight (Mn) of the polymer can be estimated. The observed Mn values of PSts estimated by this method are compared with those determined by size exclusion chromatography (SEC). By this method it was possible to make direct determination of Mn values of PSts with MWs between 1000 and ca. 1 million. Furthermore, the calibration curve produced by the relative intensity of one of the most characteristic peaks (2-phenyl-1-heptene; C4H9? CH2C (Ph) = CH2) for the end group to total intensity of all the peaks in the pyrogram, gave rapid and highly reproducible Mn values. © 1994 John Wiley & Sons, Inc.  相似文献   

9.
Bisphenol A polycarbonate (PC) was irradiated with Co-60 γ-rays at room temperature in presence of air or nitrogen. Changes were followed by IR and NMR spectroscopy as well as GPC, X-ray diffraction, melting temperature, DTA, and isothermal TG. The changes observed include chain scission, chain branching/cross-links, decrease in thermal stability, crystallinity and scission of the product. However, there are some differences in samples irradiated in presence of air as opposed to those irradiated in nitrogen, but many changes are similar. Air accelerates the radiolytic degradation of PC. The radiochemical yield of chain scission, G(s), increased to 0.20 from 0.13; the chain branching/cross-links are less numerous as compared to those after irradiation in nitrogen the thermal stability decreases along with the temperature of that rate of maximum mass loss, crystallinity and the melting temperature. The chemical products in both air and nitrogen irradiations are almost identical. The formation of these products could be explained by multiple pathways-free radical formation and by ring as well as side chain attack. The studies suggest that Fries' rearrangement is not an important pathway during the radiolytic degradation as compared with photodegradation.  相似文献   

10.
Chong CP  Lin TY  Chang CL  Yang YL  Tsai MH  Yu YS  Liu MY 《Electrophoresis》2011,32(10):1241-1251
A simple and fast micellar electrokinetic chromatography (MEKC) method was developed to investigate phospholipids isolated from human high-density lipoproteins (HDL). To optimize the MEKC conditions, several factors including bile salt concentration and organic modifier concentration in the separation buffer as well as temperature have been examined. The optimal separation buffer chosen was a mixture of 50 mM bile salts, 30% v/v 1-propanol and 10 mM sodium phosphate (pH 8.5). The applied voltage and temperature selected were 25 kV and 40°C, respectively. Meanwhile, high-salt stacking has been performed for sample pre-concentration to enhance peak sensitivity. Several factors including organic modifier concentration and salt concentration in the sample matrix as well as sample injection time have been optimized. The optimal sample buffer selected was a mixture of 100 mM NaCl and 20% 1-propanol, and the optimal sample injection time selected was 32 s under a pressure of 0.5 psi. Several phospholipid standards including lysophosphatidyl choline, phosphatidyl choline (PC), sphingomyelin, phosphatidyl ethanolamine, phosphatidyl inositol, phosphatidyl serine and phosphatidic acid have been studied using the optimal MEKC method. The MEKC profile of the mixed phospholipid standards showed good separation and reproducibility. The linear ranges for PC and sphingomyelin were 0.025-1.2 and 0.025-2.0 mg/mL, respectively. The concentration limits of detection of PC and sphingomyelin were 0.0156 and 0.0199 mg/mL, respectively. Using phosphatidic acid as an internal standard, precision and accuracy have been measured for PC and sphingomyelin. The intraday and interday quantitative analysis showed good results. The new MEKC method has been used to characterize native, in vitro oxidized and glycated human HDL phospholipids within 16 min. At absorbance 200 nm, two similar peaks were observed for native and oxidized HDL phospholipids, but three peaks were observed for glycated HDL phospholipids. Interestingly, at absorbance 234 nm, distinctively different MEKC profiles were observed for the three HDL phospholipids.  相似文献   

11.
Corn starch was modified by cross-linking with epichlorohydrin and plasticizer glycerol. X-ray diffraction studies showed that relative crystallinity of the native and cross-linked starch were similar and were not affected by cross-linking. Different films were prepared by blending corn starch, cross-linked starch or glycerol modified starch in LDPE. The mechanical properties of the films were studied for tensile strength, elongation, melt flow index, and burst strength. The properties of the blend films were compared with LDPE films. It was observed that with the blending of 7.5% native starch, there was a decrease in tensile strength, elongation and melt flow index but burst strength increased. The tensile strength, elongation and melt flow index of the films containing cross-linked starch was considerably higher than those containing native starch but the burst strength showed a reverse trend. For native starch and cross-linked starch modified with glycerol, the elongation and melt flow index of the films increased but burst strength decreased. Surface scanning of the blend films were done by scanning electron microscope. Film containing cross-linked starch/glycerol modified starch in the blend was observed to be smoother than the native starch blend films.  相似文献   

12.
Cross-linked polyethylene (XLPE) is notable for its use as power cable insulation. Its longevity is limited by space charge buildup linked to impurities such as the byproducts left behind by the cross-linking agent dicumyl peroxide (DCP). The goal of this work is to determine the impacts of these byproducts on charge trapping and detrapping in XLPE using the thermally stimulated depolarization current technique. XLPE with byproducts has one source of trapped charge, which originates from the byproducts. XLPE that was thermally treated via degassing exhibits two other sources of trapped charge, which are charge injection and dipolar relaxations. Oxidation from degassing was shown to control the trapping from these sources, which is useful knowledge for processing this material prior to its use. Reintroducing acetophenone, one of the major byproducts of DCP, suppresses those two peaks once more, showing that it controls the overall space charge buildup characteristics in XLPE.  相似文献   

13.
(Meth)acrylate/pt‐butylphenol (PTBP)‐coterminated bisphenol A polycarbonates (PCs) were prepared by interfacial processes and subsequently were reacted at high temperatures (≥200 °C) to form new branched polymers. Two interfacial methods were used to prepare the precursor linear PCs, one with (meth)acryloyl chloride [(M)AC] and the other with (meth)acrylic acid [(M)AA]. Both processes involve phosgenation in the presence of catalytic amounts of triethylamine. The process that used (M)AC formed disproportionately large amounts of bisphenol A di(meth)acrylate, whereas the process using (M)AA required about 50% more phosgene to achieve high (M)AA conversions than typical interfacial PC processes. The branching of the acrylate/PTBP PCs occurred with heating at temperatures greater than or equal to 250 °C. The molecular weight and degree of branching depended on the mole ratio of the thermally reactive and nonreactive coterminators, the total amount of coterminators, and the reaction conditions. The functionality of the branch points formed appeared to be dependent on the acrylate concentration. The branching of the methacrylate/PTBP PCs required the presence of a free‐radical initiator and temperatures up to about 200 °C. The methacrylate end group was less effective than the acrylate on a molar basis in increasing the branched polymer molecular weight and degree of branching. The melt rheology of the branched acrylate/PTBP PCs showed the expected increase in low shear viscosity and shear rate sensitivity with increasing weight‐average molecular weight and acrylate‐end‐group concentration. Small changes in the total terminator concentration and, therefore, the linear precursor polymer molecular weight produced large effects in the low shear rate melt viscosity. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2340–2351, 2000  相似文献   

14.
Three kinds of ionic liquids, 1-alkyl-3-methylimidazolium tetrafluoroborate (n=2–4), were prepared and fundamental properties of ionic liquids and those mixed with industrially used organic solvents (PC, GBL and AN) were investigated compared to solid salts, TEMABF4. It was found that degree of ionization of the ionic liquids were almost same as that of TEMABF4 from the conductivity measurement in diluted system of PC. The ionic liquids and the organic solvents intermingle with each other. Some enhancement in conductivity was observed compared to TEMABF4.  相似文献   

15.
Blends of Poly(butylene terephthalate), PBT, with Polycarbonate, PC, were studied for a range of molecular weights and blend compositions. Blends were available in PBT/PC compositions 80/20 and 40/60, and with Mw designated by H (high) or L (low). Samples were prepared by melt crystallization, or by cold crystallization following a rapid quench from the melt. Addition of PC reduces the crystallization kinetics of PBT so that the resulting crystals are more perfect than those which form in the homopolymer. Degree of crystallinity of the blends followed the rank ordering: L/L > L/H > H/L = H/H. The glass transition behavior was investigated using dynamic mechanical analysis (DMA) and modulated differential scanning calorimetry (MDSC). All blends exhibited two glass transitions at intermediate temperatures between the Tgs of the homopolymers, indicating existence of a PBT-rich phase and a PC-rich phase. Blends L/L were most, and H/H the least, miscible. Small-angle X-ray scattering was performed at room temperature on cold crystallized blends, or at elevated temperature during melt crystallization. The long period was consistently larger, and the linear stack crystallinity was consistently smaller, in blends L/L or H/L. These results indicate that in blends containing low Mw PC, there is more PC located within the PBT-rich phase. The long period was consistently smaller in cold crystallized samples, while the linear stack crystallinity was nearly the same, regardless of melt or cold crystallization treatment. Reduction of the average long period in cold crystallized samples could result from crystallization of PBT within the PC-rich phase. This is consistent with thermal analysis results, which indicate that cold crystallized samples have greater overall crystallinity than melt crystallized samples. A hypothetical liquid phase diagram is presented to explain the differences between melt and cold crystallized blends. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
Thermoluminescence (TL) and thermally stimulated conductivity (TSC) glow curves in poly(vinyl chloride), polyethylene, polystyrene, polytetrafluoroethylene, and polyimide have been compared, and many similarities have been observed. Comparison with available NMR, dynamic mechanical loss, and dielectric loss, molecular mobility data shows that most TL and TSC peaks occur at temperatures similar to those assigned to the onset of specific molecular motions, suggesting that the peaks are due to the liberation of electrons from traps formed by the polymer chains themselves, e.g., potential wells or cavities due to chain entanglement in amorphous regions, or main-chain branching points. Peaks for which correlation with molecular motion is not apparent are tentatively assigned to liberation of electrons from traps centered on impurities. The TSC peak temperatures in PVC were not affected in any consistent fashion by the application of high-strength electric fields during the warming process, indicating that the electron traps are electrically neutral when empty and charged when filled; the direction of the TSC currents appears to be determined by temperature gradients existing within the samples. The TL glow-curves are generally in good agreement with the results of other workers. The dark dc conductivity of PVC not exposed to ionizing radiation rises sharply in the temperature region assigned to the β-relaxation process, suggesting that the electron mobility in that polymer is dominated by molecular chain motion, i.e., the interchain charge transport process is probably best described in terms of a hopping process.  相似文献   

17.
Thermally reversible polyesters were obtained by the ester formation reaction of thermoplastic polyesters with hydroxyl end groups and the diacid anhydride of tetra carboxylic acid as a thermally reversible chain extender. Typical example of the thermally reversible polyesters was obtained by the reaction of PBT (polyburylene terephthalate) and PMA (pyromellitic dianhydride). This material having twice as large molecular weight as the original PBT exhibited almost the same melt viscosity as the original. Also that thermally reversible chain extension reaction occurred without unfavorable side reaction such as cross-linking. This material shows both good processability and superior mechanical properties due to its thermally reversible characters.  相似文献   

18.
The mechanical properties of fiber molded samples and monofilaments of thermally treated 73/27 4‐hydroxy benzoic acid/2‐hydroxy‐6‐napthoic acid (HBA/HNA) copolyester have been investigated using both tensile tests and flexural three‐point bending tests. The thermal treatment which involves step annealing at temperatures well below the degradation temperature of the 73/27 system has been shown to produce branching and crosslinking in the crystalline regions of these polymers. The flexural strength of the degraded sample decreased up to 10% of the untreated fiber molded sample. In case of tensile strength of a single fiber, the values for the degraded samples are in line with the untreated fiber in the low draw ratio region while a slight decrease in tensile strength was observed in the high draw ratio region. The decrease in flexural and tensile strength appears to result from a small amount of branching and crosslinking reactions which arise uniquely in the orthorhombic phase of the 73/27 HBA/HNA copolyester. The branching and crosslinking would prevent the molecular orientation along flow direction in the molten state. For the fiber molded samples of degraded 73/27 HBA/HNA the destruction of the chain regularity along fiber axis direction was observed by wide‐angle X‐ray diffraction. The 73/27 HBA/HNA copolyester including 1 wt% of a crosslinked oligomer was used to simulate the branching and crosslinking of the degraded 73/27 HBA/HNA copolyester. Plots of tensile strength versus draw ratio were similar for the degraded 73/27 HBA/HNA and a copolyester which included 1 wt% of a crosslinkable oligomer. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

19.
In this paper, we showed the simplified continuous process to crystallize the polycarbonate (PC) melts using sonocrystallization which is based on ultrasonic cavitation and ‘hot spot’ theory. High power ultrasonic wave was introduced to crystallize the PC prepolymers which is generally used for solid‐state polymerization (SSP) to obtain the ultra‐high molecular weight PCs. The structure and properties of PC treated with ultrasonic wave such as torque changes and crystallinities were studied. The experimental results showed that ultrasonic wave could improve the crystallinities of PCs and sonication time was proportional to their crystallinity. This is attributed to the decrease of the specific volume and the resulting shrinkage is compensated by a flux of a polymer melt toward the crystallization front. Torque data were collected directly from the Haake internal mixer and crystallinities were evaluated from differential scanning calorimeter (DSC). This process is an environmentally benign process so that no additional solvents or treatments are needed and will be very useful in the conventional solid‐state process which is used industrially. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

20.
Yokoyama Y  Sato H  Tsuchiya M 《Talanta》2002,57(4):667-673
The origin of the intense but unknown peaks at m/z 235 observed in liquid ionization (LI) mass spectra of middle ear effusion and serum was investigated by using related standard compounds and the collision induced dissociation techniques. The ions were observed as the base peaks in mass spectra of the aqueous fractions of middle ear effusion and serum after chloroform extraction and in those of authentic human serum albumin (HSA) too. The ions commonly observed in serous fluids could be estimated as tyrosil-valil interchain immonium ions arising from thermal decomposition of HSA during the measurement. Such thermally stable interchain immonium ions, also observed in some oligopeptides having Val-Tyr sequence as their fragment ions, are likely to be characteristic ions for large protein molecules.  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