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 296 毫秒
1.
The presence of a high temperature (>Tg) relaxation in amorphous polystyrene has been investigated further. In the previous work,1 the techniques of differential thermal analysis (DTA) and torsional braid analysis (TBA) were employed to study polystyrene as a function of “monodisperse” molecular weight. The occurrence of the Tll transition appeared to be associated with the attainment of a critical viscosity level with also corresponded with a free volume level. An entanglement network developed at a critical value of molecular weight, Mc, giving a break in the Tll-versus-M plots. The present work deals with the influence of dispersity on the Tll transition, below and above Mc. A series of binary blends of “monodisperse” anionically polymerized polystyrenes with systematic changes in M?n and heterogeneity index (M?w/M?n) was tested by TBA. The results show that when both components have molecular weights below Mc, single and average values of Tg and Tll are observed which are linearly related to M?n?1, as predicted by free volume arguments. Although a single Tg is observed when one component has a molecular weight above and the other has a molecular weight below Mc, the components appear to undergo the Tll relaxation independently. The results indicate that both the glass transition and the Tll transition are basically governed by the same type of molecular motion but at different length ranges.  相似文献   

2.
Torsional braid analysis (TBA) (~0.3 Hz) and differential thermal analysis (DTA) data are presented for the temperature for the region 0–200°C for two series of atactic polystyrenes with narrow molecular weight distributions: (a) anionic series, M?n = 600–2×106, M?w/M?n ? 1.1; (b) fractionated thermal series, M?n = 2,000–1.1×105, M?w/M?n < 1.25. Preliminary results on bimodal blends are also reported. Heating and cooling cycles were employed with TBA; only the heating mode was used with DTA. In addition to a dynamic mechanical loss peak at Tg, a higher temperature loss peak was also found. Designated the Tll or liquid–liquid transition (relaxation), its temperature is 1.1 to 1.2 Tg (°K) for polymers with molecular weight below the critical molecular weight (Mc) for chain entanglements. Above Mc ? 35,000, it rises steeply, being ?200°C for M?n = 110,000. The common dependence of Tg and Tll on M?n?1 below Mc suggests a common molecular origin. The two facts, (a) that Tll > Tg and (b) that Tll reflects chain entanglements, further suggest that Tll involves a longer chain segment length and possibly the entire molecule. Comparison of Tll versus log M plots with T versus log M isoviscous state plots based on zero-shear melt viscosity data from the literature implies that Tll measured by the TBA technique corresponds to an isoviscous state of 104–105 poises. The employment of narrow molecular weight polymers is presumably responsible for both the linear variation of the Tll transition with M?n?1 (which suggests a free volume basis for the relaxation) and the form of the variation of the Tll transition with log M (which suggests an isoviscous basis for the relaxation). The sharpness of the Tll loss peak by TBA decreases with increasing molecular weight and dispersity. The DTA endothermic event corresponding to Tll is clearly related to the occurrence of flow since the fused films which result from heating granules to 200°C and cooling to R.T. do not reveal a Tll on reheating. If a fused film is crushed, a Tll event is observed on heating. For bimodal blends with M?n < Mc for both components, the Tll transition was averaged; with one component less than and one greater than Mc, the Tll transitions of the components appeared to occur independently at temperatures corresponding to those of the isolated components. In accordance with Ueberreiter and Orthmann, Tg appears to separate a glassy state from a fixed liquid state, whereas Tll separates the fixed liquid from a true liquid state. Possible molecular interpretations for the Tll process are discussed. Systematic bodies of data from the literature which indicate the presence of the Tll process in other polymers are summarized.  相似文献   

3.
Reexamination of published DSC traces for first and second heatings of atactic polystyrenes with M?n from 2050 to 1.99 × 106, M?w/M?n ~ 1.1, is discussed in terms of some newly enunciated principles of calorimetry for the liquid state of atactic polymers. These traces were originally stated to reveal only an endothermic peak above Tg, identified with Tll at M?n < ca. 105. Three types of behavior are now distinguished: (1) First heating, M?n < Mc (the entanglement molecular weight), an endothermic slope change attributed to Tu, followed by an exotherm at Texo ascribed to melt flow and/or wetting of the DSC pan; (2) first heating for M?n above Mc, only an endothermic slope change; (3) second heating, all molecular weights, only an endothermic slope change for Tll and, in some cases, Tlp. Tll thus defined reaches an asymptotic limit when plotted against log M?n, just as Tg does. Texo begins to level off, and then, starting at Mc, increases without limit. Tll is an isofree volume state as is Tg; Texo is an isoviscous state and hence a pure relaxation process. DSC traces on first heating at M?n < 105 provide the first known experimental evidence for both the transitional and the relaxational aspects of Tll in the same experiment. The role of zero shear melt viscosity, η0, in DSC is noted: Tll depends only on M?n; Texo on η0 and hence on M?w. The Tll temperature is independent of physical form of the specimen: powder, pellet, film. DSC studies on bimodal blends of PS show Tll uniquely dependent on M?n, while the Tf – (Texo) exotherm depends on M?w. Such blends can eliminate interference between the Tll endotherm and the Tf exotherm. Exotherms using weighted pan lids and exotherms from sintering are also presented. DSC evidence for Tlp > Tll is indicated.  相似文献   

4.
This paper concerns the synthesis, characterization, and physical properties of novel polyisobutylene (PIB)-based urethane model networks prepared from diphenylmethane diisocyanate (MDI) and three-arm star PIBs capped with ? CH2OH end groups (PIB(CH2OH)3). The PIB(CH2OH)3 starting materials were produced by the inifer method in the M?n = 13,550–27,000 range. The best networks were obtained with NCO : OH = 1. Solvent extractions showed uncatalyzed network formation to be essentially complete and swelling studies indicated the expected architecture, i.e., M?c = 2M?n/3 and virtual absence of dangling chains. Stannous octoate was found to increase the rate of network formation in the absence of side reactions (i.e., allophanate formation), while other catalysts also accelerated undesirable reactions. The effect of molecular weight between crosslinks (M?c) on network physical properties has been studied in the M?c = 900–18,500 range. The tensile strength increases with decreasing M?c up to a limiting value after which it sharply declines. Elongations at break decrease monotonously with decreasing M?c. Low temperature tensile studies show higher tensile data at ?20°C, and, surprisingly, a retention of elongations at break. The Tg's decrease with increasing M?c's until a plateau is reached at ?73°C. Hysteresis is fairly constant and permanent set is constant and low. The PIB-based urethane networks exhibit excellent hydrolytic stability, negligible moisture absorption, and outstanding heat-aging stability, far beyond what is expected for a polyurethane. Conceivably the thermal deblocking of the urethane group may be reversible (? NH? COO? CH2? ? ? NCO + HOCH2? ) because the isocyanate that arises in the highly hydrophobic PIB matrix recombines with the alcohol, and cannot react with moisture as in conventional urethane networks.  相似文献   

5.
A series of epoxy networks were synthesized in which the molecular weight between crosslinks (Mc) and crosslink functionality were controlled independent of the network chain backbone composition. The glass transition temperature (Tg) of these networks was found to increase as Mc decreased. However, the rate at which Tg increased depended on crosslink functionality. The dependency of Mc on Tg is well described by two models, one based on the concept of network free volume while the other model is based on the principle of corresponding states. Initially, neither model could quantitatively predict the effect of crosslink functionality in our networks. However, our tests indicated that both the glass transition and the rubbery moduli of our networks were dependent on Mc and crosslink functionality, while the glassy state moduli were independent of these structural variables. The effect of crosslink functionality on the rubbery modulus of a network has been addressed by the front factor in rubber elasticity theory. Incorporation of this factor into the glass transition temperature models allowed for a quantitative prediction of Tg as a function of Mc and crosslink functionality. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 66: 387–395, 1997  相似文献   

6.
The relation between the structure and the viscoelastic properties of seven kinds of epoxy resins was studied. Seven tetraglycidylethers were synthesized from four-nuclei novolacs in which the positions of methylene linkage or number of kind of substituents were different. These epoxy compounds were cured with diaminodiphenylmethane as a hardener. From the viscoelastic properties of the fully cured resins with the hardener, characteristic properties such as glass transition temperature (Tg), average molecular weight between crosslinking points (M̄c), and front factor (ϕ) were obtained. It was concluded that higher linearity in the main chain of epoxy resins gave a cured resin with a higher Tg, a smaller M̄c, and a larger ϕ.  相似文献   

7.
Polymer blends of polybenzoxazine (PBA‐a) and polycaprolactone (PCL) of different molecular weights (Mn = 10,000, 45,000, and 80,000 Da) were prepared at various PBA‐a/PCL mass ratios and their properties were characterized. The results from dynamic mechanical analyzer (DMA) revealed two glass transition temperatures implying phase separation of the two polymers in the studied range of the PCL contents. Moreover, a synergistic behavior in glass transition temperature (Tg) was evidently observed in these blends with a maximum Tg value of 281°C compared with the Tg value of 169°C of the PBA‐a and about ?50°C of the PCL used. The blends with higher Mn of PCL tended to provide greater Tg value than those with lower Mn of PCL. The modulus and hardness values of PBA‐a were decreased while the elongation at break and area under the stress?strain curve were increased with an increase of the content and Mn of PCL, suggesting an enhancement of toughness of the PBA‐a. Scanning electron micrographs (SEM) of the sample fracture surface are also used to confirm the improvement in toughness of the blends. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41915.  相似文献   

8.
Anionic copolymerizations of styrene (M1) with excess 1-(4-dimethyl-aminophenyl)-1-phenylethylene (M2) were conducted in benzene at 25°C for 24h, using sec-butyllithium as initiator. Narrow molecular weight distribution copolymers with M?;n = 16.1 × 103 g/mol (M?w/M?n = 1.04) and 38.2 × 103g/mol (M?w/M?n = 1.05), and 24 and 38 moles of M2 per macromolecule, respectively, were characterized by size exclusion chromatography, 1H NMR spectroscopy and DSC. The monomer reactivity ratio, r1 = 5.6, was obtained from the copolymer composition at complete consumption of M1, assuming that the rate constant k22 =0,i.e. r2 =0. The polymers exhibited Tg values of 128 and 119°C, respectively, which correspond to an estimated Tg = 217°C for the hypothetical homopolymer of M2.  相似文献   

9.
Three copolymeric perfluoroethers with the structure CF3[(OCF2CF2)p(OCF2)q] OCF3, having different p/q ratios, have been fractionated. The fractions obtained have been characterized by Gel Permeation Chromatography and 19F-NMR. The viscosity η the specific volume v and the glass transition temperature, Tg have been measured by standard techniques for all the above samples as well as for some other perfluorinated polyethers. The temperature dependence of viscosity of the unfractionated samples is described by the W.L.F. equation. The values of fg (fractional free-volume at Tg) and of af (free-volume expansion coefficient) are independent of composition, for p/q ratios from 0.53 to 1.15. The critical molecular weight, Mc, is of the order of 8–9,000. From the molecular weight dependence of specific volume, the contribution to the molar volume of the in-chain CF2 group and the excess molar free volume of the chain ends have been determined. The limiting value of Tg for an infinite molecular weight polymer was found to depend linearly on the compositional ratio O/C and the extrapolated values for polytetrafluoroethylene and for the homopolymer (CF2O)n were found to be respectively 200 K and 120 K.  相似文献   

10.
Investigation of the Tu (>Tg) relaxation in amorphous polymers of styrene by the technique of torsional braid analysis is reviewed. For the most part the relaxation behaves like the glass transition (Tg) in its dependence on molecular weight, on average molecular weight in binary polystyrene blends, and on composition in a polystyrene homogeneously plasticized throughout the range of composition. Diblock and triblock copolymers also display a T > Tg relaxation above the Tg, of the polystyrene phase. Two results in particular suggest that the Tu relaxation is molecularly based. (1) The Tu temperature is determined by the number average molecular weight for binary blends of polystyrene when both components have molecular weights below Mc. (the critical molecular weight for chain entanglements). (2) Homopolymers, and diblock and triblock copolymers of styrene, have a T > Tg relaxation at approximately the same temperature when the molecular weight of the styrene block is equal to that of the homopolymer.  相似文献   

11.
Dielectric properties have been investigated for a bisphenol-A type epoxide oligomer, whose weight average molecular weight (M?w) was 9454. The dielectric α-relaxation of the oligomer was found to be governed by the Havriliak–Negami equation as well as the same series of oligomers with smaller M?ws (388≦M?w ≦ 3903). The dielectric relaxation times (τ)s for the oligomers with different M?ws (1396 ≦ M?w ≦ 9454) can be expressed by the Williams–Landel–Ferry (WLF) equation as a function of the glass transition temperature (Tg) at fixed temperatures from 70 to 100°C. The finding indicates that the Tg of the epoxide oligomer is calculated from the τ through the WLF equation, providing the relation between Tg and τ. The same type of WLF equation was also successfully applied to describe the Tg, dependence of the practical dielectric relaxation time (τp), which was obtained from the peak of the dielectric loss vs. frequency curve. The τp can be calculated more easily than the τ, based on the Havriliak–Negami equation, not only in the measurement of epoxide oligomer, but also in that of the reactive epoxy resin systems during curing. The Tg of an epoxy–aromatic amine system, which was determined from the τp nondestructively detected in the dielectric cure monitoring, was consistent with the Tg experimentally measured by differential scanning calorimetry (DSC).  相似文献   

12.
Molecular weight distributions for polypropylene samples have been determined by a permeation fractionation method (GPC). Porous silica beads were used as a packing material for the columns. The set of columns allows a good separation of the polypropylene macromolecular chains in a range of molecular weights from 5000 to 1.5 × 106, and the thermal and mechanical stabilities of these beads are very good. The calibration has been carried out with fractions of polypropylene of narrow molecular weight distribution prepared by a large-scale column fractionation. The molecular weights M?w and M?n and the ratios M?w/M?n calculated from the GPC curves show, in general, good agreement with the ones calculated from the column fractionation curves. However, the M?w/M?n ratios are always highter in the case of GPC fractionation. This could be due to diffusion phenomena.  相似文献   

13.
Melting and crystallization behavior of virgin polytetrafluoroethylene have been studied using a differential scanning calorimeter. Following quantitative relationship was found between number average molecular weight of polytetrafluoroethylene and the heat of crystallization in the molecular weight range of 5.2 × 105 to 4.5 × 107: M?n = 2.1 × 1010 ΔHc?5.16, where M?n is number average molecular weight and ΔHc is the heat of crystallization in cal/g. The heat of crystallization is independent of cooling rate ranging from 4 to 32°C/min. This relationship provides a simple rapid and reliable method for measuring the molecular weight of polytetrafluoroethylene.  相似文献   

14.
The influence of the reactive endgroup on the synthesis, cure behavior and network properties of thermosetting polyetherimides was investigated. Reactive phenylethynyl, ethynyl and maleimide terminated etherimide oligomers were prepared and characterized. Optimal reaction conditions were established to produce fully endcapped oligomers with imidized structures and controlled molecular weight. The phenylethynyl and ethynyl endcapped systems were synthesized by ester-acid methods. The maleimide endcapped system utilized an amic-acid route. Phenylethynyl endcapped oligomers had good processibility and were thermally cured at high temperatures (350–380°C). The networks exhibited good thermal and hydrolytic stability and good adhesion strength, and are candidates for “primary'' bonding adhesives. In contrast, more reactive ethynyl and maleimide endcapped systems were prepared as “secondary'' bonding materials, which could be cured at temperatures lower than that of the T g of the primary structure. Lap shear test results obtained from NMP-cast/methanol-extracted scrim-cloth-supported precursors confirmed that good adhesion to titanium at both room temperature and at 177°C was achieved when cured at 250°C-280°C. High glass transition temperatures and good thermal stability were achieved as determined by thermal analysis (DSC, TGA and DMA). Solvent extraction measurements confirmed that very high gel fractions were obtained, which is consistent with good chemical resistance.

The influence of molecular weight between crosslinks (Mc) on thermal and mechanical behavior was also investigated for 2,3,5,7 and 10k initial M n values. Lower molecular weight oligomers exhibited lower T g and cure temperatures, but higher cured network crosslink densities afforded higher T g and higher gel fractions, but with reduced toughness.  相似文献   

15.
Various low-density polyethylenes ranging in initial weight-average molecular weight (M?w) from 7600 to 589000 having a ratio of M?w to number-average molecular weight (M?n) of about 5 were irradiated by γ-rays in vacuo at 30°C. Gel fractions were determined and analyzed by using the equation derived by Charlesby and Pinner. The following relationships were obtained when M?w was used as the molecular weight: where rg represents the gel point dosage (Mrad), [C?C]0 is the sum of the initial contents of terminal vinyl and vinylidene unsaturations (mole/g polyethylene), and q0 and p0 are the probabilities of crosslinking and main-chain scission per monomer unit for a unit radiation dose in Mrad, respectively. Similar relationships to the equations described above were also obtained when M?n was used. From the results, it was concluded that terminal vinyl and vinylidene unsaturations play an important role for the gel formation in the γ-radiation-induced crosslinking of polyethylene in vacuo at room temperature.  相似文献   

16.
The temperature dependence of 1H spin-lattice, T1(H), and spin-spin, T2(H), relaxation times of isotactic polypropylene with relatively low molecular weight, M?w = 1.95 × 105, and ultra-high molecular weight, M?w = 1.78 × 106, crystallized at high temperature (155° C) for a long time (200 h), was measured with a broad line pulse spectrometer. T1(H) for the ultra-high molecular weight sample is shorter than that for the low molecular weight sample, while T2(H) for the ultra-high molecular weight sample is longer than that for the low molecular weight sample. From the analysis of free induction decay (FID), three components of T2(H), i.e. T2c(H), T2m(H) and T2a(H), were obtained. These relaxation times are associated with the crystalline, intermediate and amorphous regions, respectively. Here the intermediate regions are the regions in which chain molecules have intermediate mobility between that of crystalline and amorphous regions. A decrease in signal intensity of methylene, methine and methyl carbons was measured as a function of delay time following a 90° pulse before cross-polarization. From the slope of a log (intensity) versus t2 plot a relaxation time associated with rigid regions of proton nuclei, T2r(H), was obtained, where the rigid regions are the regions in which cross-polarization occurs easily. The value of T2r(H) lies between that of T2c(H) and T2m(H), indicating that the intermediate regions act as the rigid regions so far as cross-polarization is concerned.  相似文献   

17.
A series of polyurethane (PU) films was prepared from chain-extended hydroxypropyl lignins (CEHPL). In appearance, these films ranged from brittle and dark brown to rubbery and bronze. The thermal, mechanical, and network properties of these PUs were investigated by DMTA and DSC analysis. All films exhibited single Tg's which varied between ?53° and 101°C, depending on lignin content. From swelling experiments, molecular weight between crosslinks (M c) was determined and found to vary over 2.5 orders of magnitude. The M c's were related to the change in Tg that accompanied network formation. Stress–strain experiments showed a variation in Young's modulus between 7 and 1300 MPa. Most of the variation in material properties was related to lignin content and to a lesser extent to diisocyanate type, hexamethylene diisocyanate, or toluene diisocyanate. The source of the CEHPL had no effect on the observed properties. From these results it was concluded that the properties of PUs can be controlled and engineered for a wide variety of practical uses.  相似文献   

18.
New polyesters have been prepared by solution and interfacial polycondensation of acid chlorides obtained from bis‐α,ω‐polymethylene‐[(2‐methoxy)phenoxy]acetic acids, acid‐ether siderophores AE n (n being the number of methylene units), and diols including bisphenol A, ethylene glycol and propanediol. For interfacial polycondensation, average molecular weights range from 1000 to 4500 g mol?1; at higher temperature the molecular weights increase but the yield is smaller. Differential scanning calorimetry shows Tg ranging from ?5 to + 110 °C. Polyesters are crystalline with n = 6 and amorphous with n = 2 and 4; in addition, increasing n in AE n increases molecular weights and Tg values. In solution polycondensation, the average molecular weights are around 3000 g mol?1 and Tg is low. © 2001 Society of Chemical Industry  相似文献   

19.
Dihydroxy-terminated polyacetals had been synthesized from aldehydes and glycols and used as soft segments to obtain segmented polyurethane block copolymers. For soft segment ≥ 1700 M n, the Tg ranges from–48 to ?58°C and is insensitive to the structures of diisocyanate and chain extender. The Tg of PacPU with 1350 M n polyacetals is raised to ?38°C, and none was observed for shorter polyacetal chains. The copolymers can be synthesized to have a broad range of mechanical properties, such as modulus from 0.5 to 130 MPa, stress at break from 0.7 to 21 MPa, and elongation at break from 66 to 1300% through the variation of the constituents and composition. The rheologic properties are only slightly dependent on temperature for symmetrical diisocyanates but quite temperature sensitive with asymmetric diisocyanate copolymers. The polyacetals are selected to build in acid-catalyzed thermal decomposition of the thermoplastic elastomers. The extreme acid sensitivity of the polyacetal block is buffered in the coplymers.  相似文献   

20.
An extrusion-grade of high density polyethylene (HOPE) (3 ethyl groups per 1000 carbons) has been divided into 16 fractions by preparative GPC and selective p-xylene extraction. The fractions, with molecular weights ranging from 900 to 1,000,000, have been studied by IR spectros-copy, DSC, WAXS, polarized microscopy, and small-angle light scattering (SALS), The average degree of chain branching (percent C2H5) is 0.5 percent for the part of the sample having a molecular weight lower than 10,000 and it decreases monotonically with increasing molecular weight, finally approaching 0.1 percent C2H5. A crystallinity depression with respect to linear PE equivalent to 20 percent/(percent C2H5) is recorded for all samples except for the very low molecular weight samples for which the crystallinity depression is much larger (30 to 35 percent/ (percent C2H5)). The unit cell volume increases with increasing percent C2H5, presumably due to the inclusion of ethyl groups in the crystals as interstitlals at 2gl kinks. The concentration of ethyl groups in the crystals (?c) unanimously follows the relationship: ?c(percent) = 0.32 + 0.25 log(percent C2H5) except for the low molecular weight fractions which have significantly lower values for ?c. Our admittedly speculative explanation for this major discrepancy between high and low molecular weight samples is based on the idea that segments with ethyl groups close to chain ends have a greater difficulty in crystallizing than segments containing ethyl groups located at positions far from the chain ends. The fractions obtained from the extrusion-grade HDPE show a solidification temperature depression with respect to linear PE which can only be explained by the presence of chain branches in these samples. The depression is particularly pronounced for the low molecular weight samples as is expected from the data on molecular structure. Well-developed non-banded spherulites are observed in rapidly cooled (crystallized at about 35 K supercooling), low molecular weight samples (6,000 < Mw < 8,000)from the extrusion-grade HDPE in contrast to the axialites observed in linear PE of the same molecular weight and thermal treatment. This discrepancy in morphology has been related to the presence of ethyl groups in the extrusion grade HDPE fractions. Higher molecular weight samples (20,000 < Mw < 1,000,000)from the extrusion-grade HDPE and linear PE both display well-developed banded spherulites of similar nature as is expected due to the similarity in molecular structure of the two sets of sample.  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