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.
We introduce a method for assessing the glass transition temperature (Tg) of thin polymer films. The technique may be applied to any polymer film that can effect liquid crystal alignment, and is demonstrated here for a commercial polyamide-imide. The method leverages the ability of the polymer film to align nematic liquid crystals on its surface, when that surface has been prepared by mechanical brushing. Relaxation of the alignment layer, brought about by thermal cycling through Tg, is seen to affect liquid crystal alignment, and thus serve as a Tg indicator. The technique reveals a three-order-of-magnitude change in the measured property. The method allows the assessment of that portion of the film responsible for aligning liquid crystals, and provides an indication of the efficacy of alignment. Our results imply that the relaxation of the surface in Probimide 32 occurs over a range of temperatures, and relaxation is not complete until the film is heated to a temperature above the glass transition of the bulk polymer.  相似文献   

2.
Influence of styrene‐acrylate latexes with varied glass transition temperature (Tg) on cement hydration was studied and the mechanism was analyzed. Results show that polymer latexes with varied Tg retard cement hydration to different extents. Specifically, low Tg polymer shows stronger retardation effect than the high Tg polymer. Despite similar surface charges, colloidal particles with lower Tg exhibit higher affinity to surface of cement grains than the high Tg polymer, indicated by the higher adsorption amount and denser covering layer. The low Tg polymer experiences particle packing, deforming, and film forming processes along with the consumption of water during cement hydration, which eventually produces a covering layer of polymer surrounding cement grains. However, for the high Tg polymer, film forming process is absent. Consequently, the higher adsorption amount and the film‐formation process along with cement hydration are the two reasons for the stronger retardation effect of the low Tg polymer. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 45264.  相似文献   

3.
The surface crystallization behavior of poly(ethylene terephthalate) (PET) and poly(ethylene 2,6‐naphthalate) (PEN) spin‐coated thin films was compared by means of atomic force microscopy (AFM) with an in situ heating stage. As the films were heated up stepwise, characteristic surface crystals appeared at a crystallization temperature (Tc) in the near‐surface region which is about 15 °C under the bulk Tc, and were replaced by bulk crystals when the temperature was increased to the bulk Tc. In the case of films whose thickness is less than 70 nm (PET) and 60 nm (PEN), significant increases in the bulk Tc were observed. Scanning force microscopy (SFM) force‐distance curve measurements showed that the glass transition temperature (Tg) of the near‐surface region of PET and PEN were 22.0 and 26.6 °C below their bulk Tg (obtained by DSC). After the onset of surface crystallization, edge‐on and flat‐on crystals appeared at the free surface of PET and PEN thin films, whose morphologies are very different to those of the bulk crystals. Although the same general behavior was observed for both polyesters, there are significant differences both the influence of the surface and substrate on the transition temperatures, and in morphology of the surface crystals. These phenomena are discussed in terms of the differences in the mobility of polymer chains near the surface. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 44269.  相似文献   

4.
The glass transition temperatures (Tg) of poly(acrylic acid), poly(methyl acrylate), and poly(ethyl acrylate) filled with submicron particulate silicas and silicates have been measured by dynamic mechanical spectroscopy and differential scanning calorimetry. The peak temperatures of the damping factor (tan δ) and the dynamic shear loss modulus (G″) were shifted by an amount which depended upon the quantity and type of filler added to each polymer. The temperatures corresponding to the step discontinuity of specific heat also shifted, but to a lesser extent than those measured mechanically. The degree of Tg shift per unit of volumetric filler addition increased with polymer pendant group polarity for both measurement methods. Utilizing a Tg-crosslinking analogue, a model was developed that related positive Tg shifts to polymeric segmental adsorption onto filler surfaces. This model also incorporated negative contributions to the Tg shift from energy storage mechanisms arising from particle–particle interactions, as well as corrections due to effective surface area available for polymer adsorption.  相似文献   

5.
Because of its slowly crystallizing nature, poly(ethylene terephthalate) (PET) can be supercooled into an amorphous glass by rapid quenching. Upon reheating between Tg and Tm, the amorphous PET are subjected to two competing processes: rubber softening and crystallization. Fusion bonding of two such crystallizable amorphous polymer sheets in this processing temperature window is thus a complex process, different from fusion of purely amorphous polymer above Tg or semicrystalline polymer above Tm. In this study, the interfacial morphological development during fusion bonding of supercooled PET in the temperature window between Tg and Tm was studied. A unique double‐zone interfacial morphology was observed at the bond. Transcrystals were found to nucleate at the interface and grow inward toward the bulk and appeared to induce nucleation in the bulk to form a second interfacial region. The size and morphology of the two zones were found to be significantly affected by the fusion bonding conditions, particularly the fusion temperature. The fusion bonding strength determined by the peeling test was found to be significantly affected by the state of crystallization and the morphological development at the bonding interface. Based on the interfacial morphology observed and the bonding strength measured, a fusion bonding mechanism of crystallizable amorphous polymer was proposed. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

6.
A mathematical model was developed for plasticating single-screw extrusion of amorphous polymers. We considered a standard metering screw design. By introducing a ‘critical flow temperature’ (Tcf), below which an amorphous polymer may be regarded as a ‘rubber-like’ solid, we modified the Lee-Han melting model, which had been developed earlier for the extrusion of crystalline polymers, to model the flow of an amorphous polymer in the screw channel. Tcf is de facto a temperature equivalent to the melting point of a crystalline polymer. The introduction of Tcf was necessary for defining the interface between the solid bed and the melt pool, and between the solid bed and thin melt films surrounding the solid bed. We found from numerical simulations that (1) when the Tcf was assumed to be close to its glass transition temperature (Tg), the viscosity of the polymer became so high that no numerical solutions of the system of equations could be obtained, and (2) when the value of Tcf was assumed to be much higher than Tg, the extrusion pressure did not develop inside the screw channel. Thus, an optimum modeling value of Tcf appears to exist, enabling us to predict pressure profiles along the extruder axis. We found that for both polystyrene and polycarbonate, Tcf lies about 55°C above their respective Tgs. In carrying out the numerical simulation we employed (1) the WLF equation to describe the temperature dependence of the shear modulus of the bulk solid bed at temperatures between Tg and Tcf, (2) the WLF equation to describe the temperature dependence of the viscosity of molten polymer at temperatures between Tcf and Tg + 100°C, (3) the Arrhenius relationship to describe the temperature dependence of the viscosity of molten polymer at temperatures above Tg + 100°C, and (4) the truncated power-law model to describe the shear-rate dependence of the viscosity of molten polymer. We have shown that the Tg of an amorphous polymer cannot be regarded as being equal to the Tm of a crystalline polymer, because the viscosities of an amorphous polymer at or near its Tg are too large to flow like a crystalline polymer above its Tm. Also conducted was an experimental study for polystyrene and polycarbonate, using both a standard metering screw and a barrier screw design having a length-to-diameter ratio of 24. For the study, nine pressure transducers were mounted on the barrel along the extruder axis, and the pressure signal patterns and axial pressure profiles were measured at various screw speeds, throughputs, and head pressures. In addition to significantly higher rates, we found that the barrier screw design gives rise to much more stable pressure signals, thus minimizing surging, than the metering screw design. The experimentally measured axial pressure profiles were compared with prediction.  相似文献   

7.
Summary The influence of thermal history of the stationary phase in inverse gas chromatography has been studied, using polystyrene and solvents of very different solubility behaviour. While no differences in retention volumes are observable well above Tg,strongly deviating retention values are determined around and below Tg. Changes in retention behaviour can be correlated to a change in surface to volume ratio of the polymer and its influence on combined surface and bulk retention mechanism. Examplified by the system polystyrene/nitromethane the results give evidence for a different solution mechanism above and below Tg of the polymer.  相似文献   

8.
Relaxation of surface defects shows that the surface of a crosslinked epoxy system, under normal laboratory conditions, apparently has a softening transition temperature that is approximately 20°C lower than calorimetric measures of T g in the bulk. Physical aging data here confirms that this transition does resemble a glass transition. This observation is significant for properties that are determined by the surface of a coating. When ambient humidity is such that an epoxy may absorb significant quantities of moisture, it is plasticized, reducing its T g. When the relaxation of nanoindentations was measured in a humid environment, the value of this softening surface transition temperature was further reduced by 10°C. Thus, the surface properties of a polymer coating, such as wear, durability, friction, will depend not only on the ambient temperature, but also on whether the environment is humid. This is very important when such a coating responds to an environment that changes, e.g., natural weather.  相似文献   

9.
Temperature dependencies of viscoelastic functions of the three-component models of composite materials in the transition state temperature range of polymer binder have been studied. On the basis of theoretical calculations for the models, a conclusion has been made about the conditions for shift of the relaxation maxima along the temperature axis. Also conditions for their resolubility on tan δ curves were determined for materials such as filled polymers as well as anisotropic laminated and reinforced plastics with deformation of the components in series. These effects are due to the change in properties of the boundary layer of the polymer. They are entirely dependent on the concentration ratio between the boundary layer and the bulk of the binder polymer and on the difference in their glass temperatures Tg. Concentration of the high-modulus filler affects Tg of the composition. This is due to the change in the ratio of concentrations of the polymer in the boundary layer and in the bulk. With parallel deformation of the components of the three-component model, resolubility and shift of the relaxation maxima depend not only on the above factors, but also on the reinforcing filler concentration.  相似文献   

10.
A model is presented for the calculation of the time to vitrify vs. temperature for isothermal polymerization by the chain growth mechanism. The model is based on the glass transition temperature (Tg) rising from its initial value to the reaction temperature. The relationships between Tg and the volume fraction of polymer and monomer, the volume fraction of polymer and the extent of reaction, and the extent of reaction and time are also required. In a plot of temperature vs. time the vitrification curve is generally S-shaped; the time passes through a maximum just above the glass transition temperature. The model applies to linear polymerization in which monomer and high molecular weight polymer are the dominant species, i.e., to chain reactions. In this communication the model is applied to the bulk polymerization of styrene by the free radical mechanism.  相似文献   

11.
Two component topologically-interpenetrating polymer networks were made of the SIN type (simultaneous interpenetrating network) composed of two polyurethanes (a polyether-based and a polyester-based) in combination with an epoxy resin, a polyacrylate and two unsaturated polyesters. The linear polymers and/or prepolymers were combined in solution and in bulk together with the necessary crosslinking agents and catalysts. Films were cast and chains extended and crosslinked in situ. All of the IPN's exhibited one glass transition (Tg) intermediate in temperature to the Tg's of the component networks, and as sharp as the Tg's of the components. This suggests that phase separation may not occur and thus some chain entanglement (interpenetration) of the two networks is involved. The observed Tg's are always several degrees lower than the arithmetic means of the component Tg's. A theory based on interpenetration is developed to account for this.  相似文献   

12.
The coalescence of a poly(vinylidene chloride/methyl acrylate/acrylic acid) latex was found to occur only above the Tg. The various theories of film formation from polymer dispersions attribute latex coalescence to three types of forces: the pressure due to the polymer–air interface, the pressure due to a polymer–water interface, and the pressure due to capillary action. Analysis of the forces indicates that none are sufficient to cause coalescence below Tg while any, or all, may cause coalescence above Tg.  相似文献   

13.
The thermal properties of poly(ethylene terephthalate) (PET) conventional fibers and microfibers are measured and compared to bulk samples. It is shown that the glass transition temperature (Tg) of the fibers can be monitored with modulated differential scanning calorimetry (MDSC). The Tg region is about 30°C wide and shifted to approximately 110°C for conventional as well as for micro‐PET fibers. The Tg of these fibers is compared to the Tg of cold‐crystallized bulk samples. Upon crystallization, a shift and even a split up of Tg is observed. The second Tg is much broader and is situated around 90°C. This Tg is related to the appearance of a rigid amorphous phase. In comparison, the mobility of the amorphous phase in fibers is even more restricted. The whole multiple melting profile observed on the fibers is the result of a continuous melting and recrystallization process, in contrast to bulk PET. The heat‐set temperature is shown to trigger the start of melting and recrystallization. It is seen in the MDSC as an exotherm in the nonreversing signal and an excess contribution in the heat‐capacity signal. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 3840–3849, 2003  相似文献   

14.
The diffusion of nonionic penetrant, m-nitroaniline, into polyacrylonitrile was studied in detail on a range of temperature from 50.6°C to 95.0°C. The penetrant distribution in polymer is Fickian, which is different from that of cationic dye, Malachite Green reported earlier. The diffusion coefficient D increases with the rise of temperature. The sharp inflection point (72°C) of the Arrhenius plot, log D versus 1/T, corresponds to Tg of polyacrylonitrile in the presence of water, which is lower than that measured in the dry state by a dynamic mechanical testing method. The activation energy is constant below Tg (ca. 10 kcal/mole), suddenly reaches a maximum at Tg and then gradually decreases with increasing temperature. General trends of Arrhenius plot for different polymer–penetrant systems are discussed. The temperature dependence of penetrant diffusion above Tg can be described by a general form of the WLF equation, log aT = log (DTg/DT) = ? C1g(T ? Tg)/(C2g + T ? Tg), where the values of C1g and C2g were calculated to be 4.03 and 24.54, respectively. A comparison was made between m-nitroaniline and Malachite Green. The difference in the respective Tg and the constants C1g and C2g of the WLF equation in polyacrylonitrile is attributed to the size of the penetrants and their ionic character. The surface concentration increases below Tg and decreases above Tg with rise in temperature.  相似文献   

15.
Blends of an amorphous and a semi‐crystalline polymer—polystyrene and polypropylene, respectively—were prepared by melt processing in an extruder at 220°C. These polymers are known to be immiscible and the composite morphologies were characterized by electron microscopy and thermal analysis. Fine micron‐scale morphologies, ranging from 0.5 to 20 microns were observed. Thermal analysis and dynamic mechanical analysis showed changes in both the polystyrene and polypropylene glass transition temperatures (Tg) over the composition range. The major effect was a sharp increase in polystyrene Tg with increasing polypropylene content in the blend. A Tg elevation of 5.5°C was observed at 85% polypropylene. The polypropylene Tg also increases with increasing polypropylene content, starting at a depressed value in discrete polypropylene domain environments and approaching the bulk polypropylene value after the phase inversion is crossed. Qualitative structural models are proposed based on spatial and mechanical interactions between the components. POLYM. ENG. SCI., 45:1187–1193, 2005. © 2005 Society of Plastics Engineers  相似文献   

16.
Reversible cross-linking reactions of alkoxyamine-appended polymers with low glass transition temperature (Tg) were successfully carried out under bulk conditions. The low-Tg polymers with alkoxyamine units in the side chains were synthesised by radical copolymerisation of 2-ethylhexyl acrylate and two kinds of alkoxyamine-containing acrylate monomers. By heating the low-Tg polymers under bulk conditions at 100 °C, cross-linked polymers were formed by radical exchange reactions between alkoxyamine units, and a transition from a liquid-like flowable polymer state to a rubber-like polymer state was confirmed. A de-cross-linking reaction was also accomplished by radical exchange reactions between the cross-linked polymers and an added alkoxyamine-containing small molecule or stable nitroxyl radical, which resulted in transition to the flowable state again. The structural transition between low-Tg linear polymers and cross-linked polymers were characterised by 1H and 13C NMR spectroscopy, Fourier transform infrared spectroscopy, rheology measurement, swelling experiment, and gel permeation chromatography measurement.  相似文献   

17.
Differential thermal analysis has been used to examine the process of dry blending of plasticizer and PVC. The rate of transformation of the glass transition from that of the polymer initially in the cold mix to the glass transition of the blend (blend Tg) has been examined at various temperatures from room temperature to above the polymer Tg. The dependence on temperature of this rate of transformation of the observed Tg is similar to the temperature dependence of the diffusion of plasticizer into PVC. It is concluded that diffusion of plasticizer into polymer particles is the rate-determining step in the dry blending of PVC. It also appears that a single mechanism of diffusion is involved both below and above the glass transition of the polymer.  相似文献   

18.
An equation, based on thermodynamic considerations to relate the glass transition temperature, Tg, to compositional variation of a polymer system, is adapted in this article for modeling the Tg vs. fractional conversion (x) relationship of reactive thermosetting systems. Agreement between the adapted equation and experimental Tg vs. x data is found for several thermosetting crosslinking systems (i.e., epoxies and cyanate ester/polycyanurate) as well as for reactive thermosetting linear polymer systems (i.e., polyamic acid and esters to polyimides). The equation models the experimentally obtained Tg vs. x behavior of thermosetting systems which include competing reactions. Agreement for widely varying molecular structures demonstrates the generality of the equation. The entire Tg vs. x relationship can be predicted for a thermosetting material by using the Tg vs. x equation and the values of the initial glass transition temperature, Tg0, the fully reacted system glass transition temperature, Tg∞, and the ratio of the change in specific heat from the liquid or rubbery state to the glassy state (Δcp) at Tg0 and Tg∞, Δcp∞cp0. The values of Tg0, Tg, and Δcp∞cp0 can be measured generally from two differential scanning calorimetric experiments. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 64: 3–14, 1997  相似文献   

19.
The physical aging behavior of a high-Tg amine/epoxy thermosetting system has been investigated vs. change of chemical structure induced by cure and vs. aging temperature (Ta) using the Torsional Braid Analysis (TBA) technique. The chemical structure was changed systematically from monomer to highly crosslinked polymer by curing in the equilibrilium state (T > Tg). The aging temperatures ranged from just below the glass transition temperature to deep in the glassy state (Ta > Tg). In the absence of chemical reaction, the physical aging rate at a given temperature, Ta, passes through a minimum with increasing chemical conversion (i.e., change of chemical structure). Analysis of this behavior is simplified by using Tg as an index of measurement of extent of cure. There is a superposition principle for normalizing the physical aging behavior of the thermosetting glasses, which involves a shift of TgTa and a shift of C(Ta) (a function of aging temperature), regardless of chemical structure. Analysis reveals that: (1) this behavior is the consequence of the Tg and Tβ transitions, (2) the segmental mobility (1/τ) is a function of the deviation from equilibrium (as measured by TgTg and the aging time), (3) the segmental mobility, which is involved in the physical aging process in the glassy state, is insensitive to the extreme changes of chemical structure (from monomer, to sol/gel polymer, and to highly crosslinked polymer), and (4) physical aging deep in the glassy state affects both segmental mobility and cohesive energy density. © 1993 John Wiley & Sons, Inc.  相似文献   

20.
Chemical‐modified pulps were synthesized from four types of waste pulps (Pulp1–4) and succinic anhydride (SAn) or maleic anhydride (MAn). The solubility of the modified pulps was evaluated in common organic solvents, and their thermal properties were investigated by DSC measurement. The solubility of the modified pulps increased with an increasing degree of substitution (DS). However, no Tg or Tm of these modified pulps was confirmed. Pulps and modified pulps were graft‐polymerized with ε‐caprolactone (CL) in bulk and in DMAc/LiCl. Although the solubility of the graft copolymers was similar to modified pulps, some graft copolymers showed a Tg by the introduction of CL units. In the bulk, graft copolymers obtained from modified pulps and nonmodified pulps showed a Tg of about 75°C and no Tg, respectively. In DMAc/LiCl, the obtained graft copolymers from both modified and nonmodified pulps exhibited a Tg of 95–110°C. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 90: 2059–2065, 2003  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