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 46 毫秒
1.
The dependency of the mechanical properties (Young's modulus, maximum load, breaking strain, and breaking energy) of polystyrene (PS) and high-impact polystyrene (HIPS) on the tensile deformation speeds was examined without changing the mode of deformation or the shape of the test specimen. It was found that HIPS has an excellent mechanical balance compared with PS for both low (1.7 × 10?4 to 2.9 × 10?2 m/sec) and high (1.3–16m/sec) speeds. This is due to the following two mechanisms ( which have different time responses) originating from the dispersed rubber particles: (1) at low speeds, the generation of large numbers of microcrazes, and (2) at high speeds, tensile-orientation hardening of the rubber and cold-drawing of the PS matrix resulting from the rise in temperature accompanied by the abrupt eleongation of the rubber phases.  相似文献   

2.
We report on the process–structure–property relationships for Poly(lactic acid) (PLA) filaments produced through the spunbond process. The influence of spinning speed, polymer throughput, and draw ratio on crystallinity and birefringence of fibers were evaluated. We established that increasing spinning speed increases crystallinity and birefringence of fibers. We also investigate the role of fiber structures on fiber tensile properties—breaking tensile strength, strain at break, initial modulus, and natural draw ratio. An increase in spinning speed leads to a higher breaking tensile strength, higher initial modulus and lower strain at break. We have shown an almost linear relationship between breaking tensile strength of PLA fibers and birefringence. This indicates that improved tensile properties at high spinning speeds can be attributed to enhanced molecular orientation. The dependency of fiber breaking tensile strength and strain at break on spun orientation were explained with natural draw ratio, as a measure of spun orientation. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 44225.  相似文献   

3.
In this paper, we study the relaxation behavior of initially amorphous poly(ethylene terephthalate) (PET) films drawn, at 80°C using a draw rate of 2 cm/min, to a draw ratio (λ) from 1 to 5 and then quenched to room temperature. These films were then heated at different temperatures from 68 to 80°C for different times and their orientation determined. The orientation measurements were performed by transmission infrared spectroscopy and the bands used for the determination of orientation were those at 1340 and 970 cm?1 for the trans conformers, normalized using the 1410 cm?1 benzene ring vibration. The crystallinity was determined by thermal analysis. It is shown that when PET is drawn to λ values up to 2–2.5 (before stress-induced crystallization), the orientation relaxes rapidly at temperatures close to the glass transition temperature of PET. For λ values of 3 or higher, the orientational relaxation of the amorphous regions is hindered. This effect is ascribed to the development of strain-induced crystallites, which are believed to act as pseudo-crosslinks.  相似文献   

4.
《国际聚合物材料杂志》2012,61(1-4):219-229
Abstract

A number of poly(vinyl alcohol) fibers with different draw ratios was characterized by measuring the birefringence, crystalline orientational order, crystallinity, tensile strength, and modulus. The birefringence, tensile strength and modulus increased with increasing draw ratio whereas the crystallinity and crystalline order parameters remained constant within narrow limits. The increase in birefringence has to be attributed solely to an increase in chain orientation in the amorphous phase of the semicrystalline fiber. The tensile strength and modulus are therefore directly related to the chain orientation in the amorphous phase. With the aid of a simple two-phase model it was found that the modulus of the amorphous phase in its disordered conformation was 4.8 GPa. The intrinsic birefringence of the amorphous phase was found to be 79 × 10?3, i.e. much higher than the value obtained for the crystalline phase (52 × 10?3). When this value was used in calculations, it was found that the order parameter of the amorphous phase increased from around 0.1 for a draw ratio of 1 to approximately 0.6 for a draw ratio of 5, whereas the order parameter of the crystalline phase was close to 1 for all draw ratios.  相似文献   

5.
Nylon 66 microfibers were obtained by a carbon dioxide (CO2) laser‐thinning method. A laser‐thinning apparatus used to continuously prepare microfibers consisted of spools supplying and winding the fibers, a continuous‐wave CO2‐laser emitter, a system supplying the fibers, and a traverse. The diameter of the microfibers decreased as the winding speed increased, and the birefringence increased as the winding speed increased. When microfibers, obtained through the laser irradiation (at a power density of 8.0 W cm?2) of the original fiber supplied at 0.23 m min?1, were wound at 2000 m min?1, they had a diameter of 2.8 μm and a birefringence of 46 × 10?3. The draw ratio calculated from the supplying and winding speeds was 8696×. Scanning electron microscopy showed that the microfibers obtained with the laser‐thinning apparatus had smooth surfaces not roughened by laser ablation that were uniform in diameter. To study the conformational transition with winding speed, the changes in trans band at 936 cm?1 and gauche band at 1136 cm?1 were measured with a Fourier transform infrared microscope. The trans band increased as the winding speed increased, and the gauche band decreased. Young's modulus and tensile strength increased with increasing winding speed. The microfiber, which was obtained at a winding speed of 2000 m min?1, had a Young's modulus of 2.5 GPa and tensile strength of 0.6 GPa. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 99: 802–807, 2006  相似文献   

6.
In this work we use thermal shrinkage for estimating the internal stresses accumulated in a sample during a technological process. This can be a precious instrument available for every manufacturer interested in the thermal treatment efficiency. Based on the Kelvin‐Voight rheological model, we have established a correlation between the internal stresses appearing during the technological formation processes of yarns and their shrinkage at different temperatures. In this work (PET) fibers were used for study, with different birefringences ranging between 4.2×10–3–2.2×10–3. Thermal treatment was performed with warm air at temperatures between 333–453 K. The applied unitary stress changed within the range 5.09×1010–18.84×1013 N·m–2. From the examination of the curves obtained for free contraction Δε = (T) for different birefringences Δn for the preoriented PET fibers, contraction increases with increasing temperature up to a maximum situated at ≈ 353 K and then decreases to a constant value. From the examination of the curves, the dimensional change vs. contraction σ = f (Δε), to a thermal treatment under stress for different birefringences Δn, when an external stress σ is applied the thermal stability of the fibers already having on orientation, made evident by the birefringence magnitude, that in the case of a slight orientation, the retraction decreases with increasing temperature. Singular points registered for fibers with Δn = 22×10–3 for a temperature exceeding 393 K indicate that for the partially orientated fibers the sample is subjected to a stretching process during the thermal stability under stress. We established experimentally a correlation between the strain σ and Δn, Δn = ct. σ/T, i. e., a decrease of the birefringence with increasing temperature.  相似文献   

7.
An isotactic polypropylene (i‐PP) microfiber was continuously produced by using a carbon dioxide (CO2) laser‐thinning apparatus developed in our laboratory. The CO2 laser‐thinning apparatus could wind up the obtained microfiber in the range of 100 m min?1 to 2500 m min?1. The diameter of the microfiber decreased and its birefringence increased with increasing winding speed. When the microfiber obtained by irradiating the CO2 laser operated at a power density of 31.8 W cm?2 to the original fiber supplied at 0.30 m min?1 was wound at 1,387 m min?1, the obtained microfiber had a diameter of 3.5 μm and a birefringence of 25 × 10?3. The draw ratio calculated from the supplying and the winding speeds was 4,623‐fold. The SEM photographs showed that the obtained microfibers had a smooth surface without a surface roughened by a laser‐ablation and were uniform in diameter. The wide‐angle X‐ray diffraction photographs of the microfibers wound at 848 and 1,387 m min?1 showed the existence of the oriented crystallites. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 99: 27–31, 2006  相似文献   

8.
Films of pure and high-impact atactic polystyrene were prepared by the recently developed technique of solid-state coextrusion. The films were produced at extrusion rates ≥4 cm/min at 126°C with a maximum extrusion draw ratio (EDR) of 11.6. These ultradrawn films are fibrous, have a high birefringence of ?2.24 × 10?2, and exhibit a 72% elastic recovery. The material has a tensile modulus of ~4–5 GPa and a tensile strength to break of 85 MPa. Thermal analysis suggests a constant Tg.  相似文献   

9.
Poly(ethylene terephthalate) (PET) filament yarns were photostabilized by addition of 0.5–4.0 mole % dimethyl 4,4′-biphenyldicarboxylate (4,4′-BPDC) to the polymerization feed. The mechanism of photostabilization is proposed to be a triplet–triplet energy transfer from excited terephthalate units to ground-state biphenyldicarboxylate units. The mechanism of transfer is reported to be via an electron exchange mechanism, with the “quenching sphere” calculated to be 14.9 Å. Kinetic studies show the “pseudo” zero-order rate constant of initial photodegradation to decrease from 3.4 × 10?19 for the PET homopolymer to 2.0 × 10?19% breaking strength loss/quantum exposure/cm2 for the copolyester containing 4.0 mole % of the 4,4′-biphenyldicarboxyl moieties. The photophysical processes available to the dimethyl 4,4′-biphenyldicarboxylate monomer were characterized by absorption and luminescence studies. In solution, dimethyl 4,4′-biphenyldicarboxylate was shown to emit an intense fluorescence from a 2(π,π*) state and a weaker (~10?2×) phosphorescence from a 1(π,π*) state derived from the 1A→supn1Lb absorption. The copolymer yarns were shown to exhibit both fluorescence and phosphorescence from the biphenyldicarboxylate units: the fluorescence from direct excitation, the phosphorescence by sensitized transfer.  相似文献   

10.
Triplet–triplet energy transfer has been shown to occur from poly(ethylene terephthalate) (PET) units to the 2,6-naphthalenedicarboxylate (2,6-ND) monomer units in a series of poly(ethylene terephthalate–co–2,6-naphthalenedicarboxylate) (PET–2,6-ND) copolymers, as filament yarns, by an exchange mechanism at 77°K. The radius of the “quenching sphere” has been calculated to be 19.7 Å, indicating the presence of triplet energy migration. Photostabilization was observed in the copolymer yarns with the concentration of the monomer dimethyl 2,6-naphthalenedicarboxylate (2,6-DMN) at or above 2 mol %; the rate of phototendering in an air atmosphere was shown to decrease from 2.0 × 10?19% breaking strength loss/quantum absorbed/cm2 in the homopolymer PET to 0.7 × 10?19% breaking strength loss/quantum absorbed/cm2 in the copolymer yarns. The photophysical processes in the monomers, dimethyl terephthalate and 2,6-DMN, were examined by absorption and luminescence studies. The lowest excited singlet and triplet in both monomers were identified to be the 1(π, π*) and 3(π, π*) states, respectively. The phosphorescence of PET was shown to originate from a 3(π, π*) state, while the complex fluorescence spectrum may arise from some oriented aggregates in the polymer matrix. In copolymer yarns, only the fluorescence emission from the 2,6-ND monomer units at 380 nm was observed. The phosphorescence spectra of the copolymer yarns showed phosphorescence emissions from the PET and 2,6-ND monomer units; in addition, delayed fluorescence from the 2,6-ND monomer was also observed.  相似文献   

11.
Murta (Ugni molinae Turcz) berries were dried under convective and combined convective-infrared conditions at 40, 50 and 60°C and 400–800 W in order to determine the drying characteristics and to compare the dried product's quality. To model the drying kinetics, seven mathematical equations were fitted to experimental data. According to statistical tests performed, the Midilli-Kuçuk model best fitted experimental data and was closely followed by the logarithmic model. Effective moisture diffusivity also showed dependency on drying conditions and varied between 7.59 × 10?10 to 44.18 × 10?10 m2/s and 11.34 × 10?10 to 85.41 × 10?10 m2/s for air-convective drying and combined infrared-convective drying. As to quality attributes of the berries, total surface color difference (ΔE) and total phenolic content (TPC) were determined. It was found that chromaticity coefficients a* and b* changed significantly, showing ΔE to be dependent on the mode of heat supply. TPC under all drying conditions decreased and was significantly different from the initial value in fresh samples. However, at a constant drying temperature, an increase in infrared power enhanced retention of TPC in samples. In particular, working at 40°C/800 W resulted in dried samples with the highest TPC.  相似文献   

12.
The electrochemical impedance spectroscopy (EIS) technique was used to evaluate the water transport (diffusion and equilibrium water uptake) and the dielectric properties of free-standing poly(ethylene terephthalate) (PET) membranes at 40°C. Permeability and diffusion coefficients were also obtained using the Payne cup method and the MacBain quartz spring balance to assess the reliability of the EIS method when compared to other techniques. In addition, an electromigration (dc) technique was used to estimate the NaCl diffusion coefficient across PET films. Results obtained indicate that PET is highly permeable to water and much less permeable to salt. The water diffusion coefficient, D, varies from 2.11× 10?9 to 9.97× 10?9 cm2s?1 for thicknesses between 22 and 205μm, whereas the equilibrium water uptake, W, varies from 0.54 to 0.95 wt % for the same given range of thicknesses. The average calculated dielectric constant of the free-standing PET films is 3.6. An estimate of the NaCl diffusion coefficient, Ds, is 9.34× 10?14 cm2s?1. Transport properties results obtained via the electrochemical technique are in reasonable agreement with those obtained with the classical gravimetric method. © 1993 John Wiley & Sons, Inc.  相似文献   

13.
We determined that a poly(ethylene terephthalate) microfiber was easily obtained by irradiating a carbon dioxide laser to an annealed fiber. The annealed fiber was prepared by zone drawing and zone annealing. First, an original fiber was zone drawn at a drawing temperature of 90°C under an applied tension of 4.9 MPa, and the zone‐drawn fiber was subsequently zone annealed at 150°C under 50.9 MPa. The zone‐annealed fiber had a degree of crystallinity of 48%, a birefringence of 218.9 × 10?3, tensile modulus of 18.8 GPa, and tensile strength of 0.88 GPa. The microfiber prepared by laser heating the zone‐annealed fiber had a diameter of 1.5 μm, birefringence of 172.8 × 10?3, tensile modulus of 17.6 GPa, and tensile strength of 1.01 GPa. The draw ratio estimated from the diameter was 9165 times; such a high draw ratio has thus far not been achievable by any conventional drawing method. Microfibers may be made more easily by laser heating than by conventional technologies such as conjugate spinning. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 90: 1955–1958, 2003  相似文献   

14.
The graft copolymerization of methyl methacrylate onto polyester fibers (PET) was investigated using tetravalent cerium as the initiator. The rate of grafting was found to increase progressively with the initiator and monomer concentrations up to 2.5 × 10?2M and 70.41 × 10?2M, respectively. The reaction was found to be catalysed by acid up to 15.0 × 10?2M. The graft yield increased by increasing temperature. The effect of addition of some solvents and thiourea on the rate of grafting was also investigated. A suitable kinetic scheme has been pictured, and rate equations have been derived.  相似文献   

15.
Optical adhesives combine the traditional function of structural attachment with a more advanced function of providing an optical path between optical interconnects. This article aims to characterize refractive index and birefringence of such adhesives under environmental exposure to different temperature conditions. Optical time domain reflectometery (OTDR) and prism coupling methods were employed to measure optical properties of an optical adhesive. Thermo‐optic coefficient (dn/dT) of the adhesive was observed to decrease noticeably from ?2 × 10?4°C?1 to ?4 × 10?4°C?1 around the glass transition temperature (Tg ~ 78°C). It is observed that refractive indices for both TE and TM modes increase with increasing annealing temperature, but the birefringence (TE ? TM) is decreasing. This suggests that the material has become more isotropic due to the annealing. The environmental changes in optical properties of the adhesive are discussed in the light of Lorentz–Lorenz equations. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 98: 950–956, 2005  相似文献   

16.
Measurements of shrinkage stress and birefringence have been undertaken on drawn samples of a copolyterephthalate based on ethylene glycol and 1,4 cyclohexane-dimethanol. The temperature location and the peak value of the maximum shrinkage stress have been related to the drawing conditions (temperature, strain rate and draw ratio), comparing the results with those reported for poly (ethylene terephthalate). Moreover, a value of the stress optical coefficient of the copolymer equal to 14.5 × 10?9 m2 N?l has been obtained.  相似文献   

17.
Two sets of soluble high performance polyimides synthesized from 2,2′-bis(3,4-dicarboxyphenyl)hexafluoropropane dianhydride (6FDA) and 2,2′-(trifluoromethyl)-4,4′-diaminobiphenyl diamine (PFMB), and from 2,2′-bis(trifluoromethyl)-4,4′,5,5′-biphenyl-tetracarboxylic dianhydride (HFBPDA) and 2,2′-(trifluoromethyl)-4,4′-diaminobiphenyl diamine (PFMB) have been investigated by static and dynamic laser light scattering (LLS) in tetrahydrofuran (THF) at 30°C. The calibrations, for 6FDA-PFMB: <Rg> (nm) = 3.87 × 10?2 <Mw> 0.568, <Rh> (nm) = 2.38 × 10?2 <Mw>0.560 and <D> (cm2/s) = 2.13 × 10?4 <Mw>?0.560; for HFBPDA-PFMB: <Rg> (nm) = 2.24 × 10?2 <Mw>0.626, <Rh> (nm) = 1.27 × 10?2 <Mw>0.621 and <D> (cm2/s) = 3.99 × 10?4 <Mw>?0.621, have been established, where <M2>, <Rg>, <Rh> and <D> are the weight-average molar mass, the root mean square z-average radius of gyration, the z-average hydrodynamic radius and the z-average translational diffusion coefficient, respectively. A combination of <Mw> and the translational diffusion coefficient distribution G(D) leads to the calibrations of D (cm2/s) = 2.41 × 10?4M?0.564 and D (cm2/s) = 6.16 × 10?4M?0.656 for 6FDA-PFMB and HFBPDA-PFMB, respectively, where D and M correspond to monodisperse species. With these calibrations, we can convert a translational diffusion coefficient distribution G(D) into a corresponding molar mass distribution fw(M). On the basis of the Kratky-Porod wormlike chain model, the persistence lengths (q) were found to be ? 3.3 nm and ? 4.5 nm, respectively, for 6FDA-PFMB and HFBPDA-PFMB, which indicates that both polyimide chains have an extended conformation. In addition, <Rg> / <Rh> ? (1.7-1.9) shows that they are in coil conformation. Therefore, we conclude that both polyimides have an extended coil conformation.  相似文献   

18.
Melt viscosity and flow birefringence of bisphenol A-type polycarbonate were measured and analyzed by the application of rubber-like photoelastic theory. The melt viscosity in the Newtonian flow region increased with the molecular weight to the power of 3.4. In polycarbonate, the shear stress of the Newtonian flow region was to 106 dyn/cm2, whereas in PMMA it was at most 3 = 105 dyn/cm2. The flow birefringence δn has a linear relation with shear stress S, that is δn = 5.7 × 10?10 S. The principal polarization difference of flow unit α1 – α2 was 1.62 × 10?23 cm3, which was obtained by the application of the rubber-like elastic theory. In PMMA, it was 3.9 = 10?25 cm3; about 1/40 of that was polycarbonate. The anisotropy of polarizability of the flow unit of polycarbonate was also about 40 times larger than that of PMMA. So the anisotropy reflected the large flow birefringence of the polycarbonate.  相似文献   

19.
The nonlinear optical response of the liquid crystalline multiarm star‐shaped azodendrimer was investigated in picosecond pulse and CW regimes at 532 nm. The polymer exhibited large nonlinear refractive coefficient in two regimes (n2 = −2.88 × 10−13 cm2/W and −1.1 × 10−10 cm2/W under picoseconds pulse excitation, whereas n2 = −1.4 × 10−6 cm2/W and n2 = −8.8 × 10−5 cm2/W under CW laser excitation in solution and film, respectively). The mechanism accounting for the process of nonlinear refraction was discussed. The value of photoinduced birefringence in the polymer film was also measured (Δn ∼ 10−3) under CW laser excitation at 532 nm. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

20.
In preliminary experiments to optimize the condition of a laser heating, zone drawing for poly(ethylene terephthalate) (PET) fiber, a microfiber was prepared by a continuous‐wave carbon dioxide (CW CO2) laser heating. CW CO2 laser heating was carried out at an extremely low applied tension (σa) at a higher laser power density (PD) as compared to the optimum condition for the laser heating, zone drawing of PET fiber reported previously. The microfibers were obtained by CO2 laser heating carried out at a PD of 15.8 W cm?2 and under a σa of 0.66 MPa or lower. The diameter of the fiber decreased with a decreasing σa and increasing PD. The smaller the diameter, the higher was its birefringence. The smallest diameter fiber obtained at σa = 0.17 MPa at PD = 21 W cm?2 had a diameter of 4.5 μm and a birefringence of 0.112, and its draw ratio estimated from the diameter reached 3086 fold. Such a high draw ratio was not previously attained by any drawing method. In a wide‐angle X‐ray diffraction photograph of the smallest diameter fiber, indistinct reflections due to oriented crystallites were observed. An SEM micrograph of the smallest diameter fiber showed a smooth surface without any crack and was uniform in diameter. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 3297–3283, 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