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.
ABSTRACT

Small specimens of Pinus radiata have been tested to determine the creep strain that occurs during the kiln drying of boards. The samples have been tested over a range of temperatures from 20°C to 140°C. The samples, measuring 150 × 50 × 5 mm, were conditioned at various relative humidities in a pilot-plant kiln, in which the experiments at constant moisture content (MC) in the range of 5-20% MC were undertaken to eliminate mechano-sorptive strains. To determine the creep strain, the samples were brought to their equilibrium moisture content (EMC), then mechanically loaded under tension in the direction perpendicular to the grain. The strain was measured using small linear position sensors (LPS) which detect any elongation or shrinkage in the sample. The instantaneous compliance was measured within 60 sec of the application of the load (stress). The subsequent creep was monitored by the continued logging of strain data from the LPS units.

The results of these experiments are consistent with previous studies of Wu and Milota (1995) on Douglas-fir ( Pseudotsuga menziesii ). An increase in temperature or moisture content causes a rise in the creep straw while the sample is under tension. Values for the instantaneous compliance range from 1.7 × 10?3 to 1.28 × 10?7 MPa?1 at temperatures between 20°C and 140°C and moisture content in the range of 5-20%. The rates of change of the creep strains are in the Order of magnitude 10?7to10?8s?1 for these temperatures and moisture contents. The experimental data have been fitted to the constitutive equations of Wu and Miloia (1996) for Douglas-fir to give material parameters for the instantaneous and Creep strain components for Pinus radiata.  相似文献   

2.
The high-temperature compression creep of additive-free β/α silicon carbide ceramics fabricated by rapid hot pressing (RHP) was investigated. The creep tests were accomplished in vacuum at temperature range 1500 °C–1750 °C and compressive loads of 200 MPa to 400 MPa. Under investigated condition the RHP ceramics possessed the lowest creep rate reported in the literature. The observed strain rates changed from 2.5 × 10?9 s?1 at 1500 °C and a lowest load of 275 MPa to 1.05 × 10?7 s?1 at 1750 °C and a highest load of 400 MPa. The average creep activation energy and the stress exponent remain essentially constant along the whole range of investigated parameters and were 315 ± 20 kJ?mol?1, and 2.22 ± 0.17, respectively. The suggested creep mechanism involves GB sliding accommodated by GB diffusion and β?α SiC phase transformation.  相似文献   

3.
A simple, versatile biconical rehemoter has been developed. This device provides shear creep and creep recovery data for polymer melts over a temperature range of 200–500°F. and a range of applied shear stresses from 2 × 103 to 9 × 105 dynes/cm2. Extensive reheological data have been obtained for two samples each of polyisobutylene and high-density polyethylene. These illustrate the value of the device in obtaining data useful for predicting and understanding the processing properties of polymer melts.  相似文献   

4.
The creep behaviour of SiCN materials derived from polyvinylsilazane (PVS) and polyhydridomethylsilazane (PHMS) precursors was investigated in the temperature range between 1200 and 1550°C at compressive stresses between 30 and 250 MPa in air. Both materials show very similar creep behaviour. Decreasing strain rates with time were observed. Even after 4×106 s creep deformation, stationary creep was not observed. Temperature dependence of the creep behaviour of such materials is very low. Dense passivating oxide layers were found on the surface of creep specimens tested in the temperature range up to 1500°C. At 1550°C active oxidation was observed.  相似文献   

5.
6.
The radiation-induced emulsion polymerization of tetrafluoroethylene was carried out with the use of ammonium perfluorooctanoate as an emulsifier at an initial pressure of ca. 30–35 Kg/cm2. The polymerization rate was shown to be proportinal to about the 0.8 power of the dose rate in the range of 2 × 104 to 105 R/hr and to be almost independent of emulsifier concentration. The molecular weight of the polymer lies in the range of 104 to 105, increases with reaction time at the initial stage, and decreases with emulsifier concentration, but is independent of the dose rate from 2 × 104 to 6 × 104 R/hr. If the emulsifier is not used, a polymer with a molecular weight as high as 1.8 × 106 to 2 × 107 is obtained. Apparently, the emulsifier and its radiolysis products act as chain transfer agents. Postirradiation polymerization was found to take place with the formation of products with increased molecular weight.  相似文献   

7.
The creep behaviour of (Ta-Hf-Zr-Nb)C high entropy ceramic (HEC) was investigated at temperatures between 1400 and 1600 °C in vacuum under compressive stresses from 150 to 300 MPa. The measured steady-state creep rates ranged from 2 × 10−9/s to 8 × 10-8/s, which are approximately 10 times lower than the published creep rates of the corresponding monocarbides. The stress exponent n is in the range of 2.34 ∼ 2.89 and the average activation energy is 212 kJ/mol. The creep mechanisms involve dislocation glide/climb and the formation of voids and cracks. The voids formed at the grain boundaries parallel to the loading direction, which often connected to form cracks at the highest load/temperature The active dislocation slip system during creep was &lt,110&gt,{111}. The reason why (Ta-Hf-Zr-Nb)C has enhanced creep resistance compared to the monocarbides can be explained by lattice distortion and the higher thermodynamic stability of HEC ceramic at high temperatures.  相似文献   

8.
Cyclic voltammetry (CV) and UV–Visible spectroscopy (UV–Vis) techniques were used to calculate binding parameters of 4-methyl-5-((phenylimino)methyl)-3H-1,2-dithiole-3-thione (MPDT) and 5-(4-fluorophenyl)-3H-1,2-dithiole-3-thione (FPDT) with DNA. The results obtained from both techniques were confirmed by computational molecular docking using AutoDock molecular docking software. The anodic peak potential shift in CV indicated an intercalative mode of binding. The binding constants (M?1) of the adducts MPDT-DNA and FPDT-DNA obtained from voltammetric measurements were found to be 8.0?×?104 and 2.4?×?104, respectively, with binding free energy being ?27.99 and ?25.01?KJ?mol?1, respectively. These results are in good agreement with those obtained from UV–Visible spectroscopic studies. The diffusion coefficients of MPDT and FPDT (2.06?×?10?10 and 2.42?×?10?9, respectively) were found to be higher than those of DNA-bound compounds (1.27?×?10?10 and 1.65?×?10?9?cm2/s, respectively). The binding free energy of MPDT and FPDT to DNA was also calculated by molecular docking study. The docking study gave excellent approximation with experimental results, shedding light on the sites of binding.  相似文献   

9.
The high pressure crystallized specimens described in the preceding paper (1) were characterized by X-ray diffraction, density, and differential scanning calorimetry, Nitric acid etching, followed by gel permeation chromatography and DSC on the residues provided further characterization of the morphology. The crystallinity decreased from 100 to 80 percent with increasing molecular weight, over the range Mw = 4.9 × 104 to 4.6 × 106. The crystallites were predominately in extended chain morphology, but the percent in extended chain form decreased from 98 to 85 percent over the indicated molecular weight range. In all cases, there was a wide, bimodal or multi-modal, distribution of lamellar thickness. The physical characterization results were consistent with and complementary to the results from electron microscopy. The large differences in crystallinity and gross morphology readily account for observed molecular weight dependent differences in mechanical properties.  相似文献   

10.
The pressure dependence of the upper critical solution temperature (dTdp)c in the polystyrene-cyclohexane system has been measured over the pressure range of 1 to 50 atm. The value of (dTdp)c determined over the molecular weight (Mw) range of 3.7 × 104 to ~145 × 104 greatly depends on the molecular weight of polystyrene. The value of (dTdp)c for a polystyrene solution of low molecular weight (Mw = 3.7 × 104) is positive (3.14 × 10?3 degree atm?1), while the values are negative (?0.52 × 10?3~?5.64 × 10?3 degree atm?) for solutions of polystyrene over the high molecular weight range of 11 × 104 to ~145 × 104. The Patterson-Delmas theory of the corresponding state and the newer Flory theory have been used to explain this behaviour.  相似文献   

11.
In the determination of molecular weight distributions by GPC, the traditional column arrangement is such that the fractionation process proceeds from a high-permeability limit to a low-permeability limit column. We report computer comparisons of data obtained from columns in their normal ordering (high- to low-permeability limit), reverse ordering, and random ordering. The columns had a permeability limit range from 1 × 106 Å down to 1 × 103 Å, and the polymers had a molecular weight range of 1.8 × 106 down to 2.1 × 103. The concentrations used varied from 0.05% up to 0.5%. The data show significantly different results, with the random arrangement the preferred ordering. A qualitative model for the separation mechanism is presented to account for the improvement in resolution. Additional data are presented which show that serious errors (as high as 45%, depending on concentration) will be encountered in GPC studies, unless the calibration curve is obtained at the same relative concentration as the samples, with definite overloading effects occurring at the higher concentrations. A new method of curve fitting was used in the higher molecular weight region to give meaningful calibration curves.  相似文献   

12.
The supermolecular structure of crystallised molecular weight fractions of poly(hexa-methylene oxide), covering the molecular weight range (M) 4.5. × 103 to 8. 5 × 104, was studied by polarised light microscopy and small angle light scattering. Different forms were observed as a function of molecular weight and crystallisation temperatures. Perfect spherulites are formed after rapid crystallisation, and these forms deteriorate as both the molecular weight and crystallisation temperature increase. The morphology in the isothermal crystallisation region corresponds to an intermediate state which represents a transition from spherulites to hedrites.  相似文献   

13.
Specimens cut from 9-in, brick of nine brands of firebrick, including two high-alumina, four fire-clay, two siliceous fire-clay, and one silica, were subjected to tensile and compressive creep tests at eleven temperatures from 25° to 950°C., inclusive. The duration of each test was approximately 240 days. Small length changes, independent of stress direction (that is, compressive or tensile), occurred at the lower temperatures. The lowest temperatures at which creep was significant were (a) high-alumina brick, 700° to 850°C.; (b) fire-clay brick, 600° to 700°C.; and (c) siliceous and silica brick, 950°C. Creep results under compressive stress could not be correlated with results under tensile stress. Specimens of different brands, at 950° C. showed greatly different capacities to carry load. Repeated heatings caused growth of silica brick of approximately 0.27%. Moduli of elasticity at room temperature were determined before and after the various heat-treatments and resultant changes were recorded. The changes in moduli were 15% or greater for silica and siliceous brick and 4% or less for the fire-clay brick. The moduli of elasticity at room temperature were approximately 2.7–4.3 × 106 for high-alumina brick, 0.6–1.9 × 106 for fire-clay brick, 0.3–1.7 × 106 for siliceous fire-clay brick, and 0.4 × 106 for silica brick.  相似文献   

14.
Determination of the sedimentation–diffusion equilibrium is one of the methods furnishing data on the thermodynamic parameters of polymer solutions. By means of this procedure measurements can be performed over a considerable range of concentrations. The equilibria in the ultracentrifuge have been determined for five polystyrene samples of narrow distribution covering the molecular weight range 2 × 104 to 2 × 106, at concentrations up to 40 wt-% and at temperatures of 30, 45 and 65 °C. The results are expressed in the data for the chemical potential of the solvent, the number-average chemical potential of the polymer and the interaction parameter ξ. The results are compared with those obtained from osmotic pressure, light scattering and critical miscibility measurements. At very low concentrations the molecular weight dependence of the interaction parameter is in agreement with literature values derived from osmotic pressure and light scattering figures. At higher concentrations this molecular-weight dependence decreases sharply but remains noticeable up to 40% concentration.  相似文献   

15.
S. Saeki  D.C. Bonner 《Polymer》1978,19(3):319-324
The activity of benzene in poly(ethylene oxide), PEO, has been determined over the concentration range 0.3 to ~35 wt% benzene using the piezoelectric sorption method. The temperature range was 64° to 97°C and the molecular weights of polymer samples were 1.0 × 105, 6.0 × 105, and 5.0 × 106 g/g mol. The Flory-Huggins interaction parameters, χ, determined in this work agree within experimental error with χ values determined by gas chromatography and by vapour pressure measurement. The values of χ extrapolated to zero solvent concentration, χ, have a minimum and negative value in the vicinity of 85°C for the sample of molecular weight 1.0 × 105, 84°C for the sample of molecular weight 6.0 × 105, and 77°C for the sample of molecular weight 5.0 × 106. In the vicinity of 65°C, χ is negative for each molecular weight of polymer and increases to positive values with an increase in temperature. For all samples studied, the χ parameters reach constant values (0.22 to 0.30) at temperatures higher than 90°C.  相似文献   

16.
Sodium bisulfite-initiated polymerization of methyl methacrylate (MMA) in water medium was carried out in the absence and in the presence of cupric oxide and manganese dioxide using various initiator concentrations at various temperatures ranging from 30° to 60°C. It seems that the metal oxide–water interface plays an important role, as it has been found that both oxides accelerate the rate of polymerization. Cupric oxide was found to be more effective than manganese dioxide. The cupric oxide was found to have nearly the same catalytic effect as the cuprous oxide, and manganese dioxide was found to be somewhat more effective than titanium dioxide. The initial rate of polymerization increased from 2.3 × 10?5 mole/(l.sec) to 3.4 × 10?4 mole/(l.sec) and to 6.6 × 10?5 mole/(l.sec) when the metal oxide concentration increased from 0 to 3 g/l. in case of cupric oxide and manganese dioxide, respectively. The initial rate of polymerization increased from 3.7 × 10?4 mole/(l.sec) to 4.2 × 10?4 mole/(l.sec) and from 7.2 × 10?5 to 2.2 × 10?4 mole/(l.sec) when the temperature was raised from 30° to 60°C in the presence of cupric oxide and manganese dioxide, (9 g/l.), respectively. Both the rate of polymerization and the number-average molecular weights were found to increase with increase the monomer concentration; the rate values were higher while the number-average molecular weights were lower in case of cupric oxide than in case of manganese dioxide. For example, the rate of polymerization increased from 2 × 10?5 mole/(l.sec) to 8.1 × 10?5 mole/(l.sec) and from 1.9 × 10?5 mole/(l.sec) to 6.9 × 10?5 mole/(l.sec); and the number-average molecular weight increased from 0.7 × 105 to 2.2 × 105 and from 1.5 × 105 to 4.9 × 105 in the presence of cupric oxide and manganese dioxide (10 g/l.), respectively, when the monomer concentration was increased from 23.5 g to 94 g/1. water. The apparent energy of activation for the polymerization of methyl methacrylate in water medium between 40° and 50°C was found to be 0.8 and 4.3 kcal/mole when using cupric oxide and manganese dioxide (9 g/l.), respectively.  相似文献   

17.
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   

18.
By means of x-ray diffraction, the lattice spacing of the (100) plane for molded polytetrafluoroethylene was measured at different temperatures from 25 to 190°C. In the crystalline region, the linear expansion coefficient, in the direction perpendicular to the molecular chain axis, was obtained as 1.1 × 10?4°C.?1 below 60°C., as 1.2 × 10?4°C.?1 above 90°C., and as a minimum value of some 0.2 × 10?4°C.?1 at about 80°C. As the linear expansion coefficient of the crystalline region in bulk was observed as some 0.6 × 10?4°C.?1, the expansion coefficient in the direction of molecular chain axis must be negative except in the transition region near 80°C. The variation of molecular chain axis separation with temperature showed an irregularity at about 80°C. but none near 130°C. in the crystalline region. Infrared absorbance of film samples of PTFE was measured at different temperatures of 25 to 150°C. range for 518, 627, and 639 cm.?1 bands. On absorbance–temperature curves for those b?ands, irregularities were observed near 30, 50, 90, and 130°C. Particularly with 518 cm.?1 band, a more crystalline sample gave more distinct irregularities near 50 and 90°C. than a less crystalline sample. The change at about 90°C. in infrared spectra may correspond to that obtained by x-ray measurements near 80°C., which was thought to occur in the crystalline region. The results obtained by x-ray and infrared measurements support the previous results by thermal, rheological, and dielectric methods: there exist first-order transitions in the crystalline region at about 90°C. and second-order transitions in the amorphous region at about 130°C.  相似文献   

19.
The effects of concentration and aspect ratio of carbon fiber and temperature on volume resistivity of polychloroprene-carbon fiber composites have been studied. The critical concentration of fiber wherein a sharp transition of electrical performance occurs from insulative range to conductive range is in the range of 5 to 10 phr for fibers of higher aspect ratio (~ 100) and in the range of 20 to 25 phr for fibers with lower aspect ratio (~ 25). Experimental values of electrical conductivity agreed reasonably well with the calculated values from a theory based on probability of formation of conductive network. Heating and cooling curves in the variation of volume resistivity with temperature do not follow the same path and results in a “hysteresis loop,” up to carbon fiber loadings of 30 phr. Such loops disappear at 40 phr loading. The activation energy of conduction for carbon fiber-filled polychloroprene varies from 1.10 × 10?20 to 0.96 × 10?20 J. Linearity in current-voltage relation was observed at room temperature. Increasing interfiber distance at higher temperatures disrupts the linear relation. The type of carrier in carbon fiber-filled polychloroprene thermovulcanizate was found to be n-type. The carrier concentration and drift mobility determined by studying the Hall effect were found in the range of 2.5 × 1022 to 2.90 × 1025 m?3 and 1.66 × 10?4 to 14.25 × 10?4 m2 V?1 s?1, respectively.  相似文献   

20.
The molecular weights of low density lipoprotein (LDL) subfractions were determined precisely by meniscus depletion sedimentation equilibrium. Equilibrium speeds ranged from 9743 to 5896 rpm. The average molecular weights of various LDL subfractions of Sf o values 9.49, 7.94, 6.42, 5.17, and 3.71 determined by sedimentation equilibrium were 2.97×106; 3.13×106; 2.89×106; 2.45 ×106; and 2.61×106 daltons, respectively; and their respective densities were 1.0267, 1.0306, 1.0358, 1.0422, and 1.0492 g/ml. Minimal hydrated molecular weights for these fractions determined by flotation velocity at 37,020 rpm were 2.57×106; 2.37×106; 2.09×106; 1.94×106; and 1.81 ×106 daltons; whereas similar molecular weights determined at 52,640 rpm were 2.53×106; 2.27 ×106; 1.99×106; 1.86×106; and 1.74×106 daltons for the respective LDL subfractions. Higher molecular weights of fractions 2 and 5 compared to their adjacent fractions 1 and 4 by sedimentation equilibrium are of great interest. The calculated frictional ratio f/fo from sedimentation equilibrium and flotation velocity data ranges from 1.10 to 1.31, suggesting complexity and asymmetry of LDL subfraction molecules. There is also evidence that compressibility of LDL molecules may be different than that for the salt solution under high g-force. Assuming that redistributed LDL molecules at equilibrium under low g-force are spherical, it is possible that the shape of LDL molecules undergoing flotation velocity determinations may be distorted in high g-force conditions. Such distortion may be consistent with the high f/fo values obtained and may also be a basis for structural rearrangement and/or lipoprotein degradation with prolonged preparative ultracentrifugation at high g-force and pressure.  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