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 343 毫秒
1.
Thermally stimulated currents (TSC) were examined for poly(bis(p-fluorophenoxy)phosphazene) (PBFPP) film. TSC showed peaks at the glass transition temperature (Tg=–4 °C) and atT(1) (160 °C – 170 °C), where-form crystal phase transformed to mesophase of-form structure. Another peak was found atT cc betweenTg andT(1). Linear relationship between polarization field and peak current ofT cc -peak was found, which shows thatT cc -peak was caused by motion of dipolar groups in crystalline phase. When heating (up to 200 °C) and cooling (down to 20 °C) thermal process was repeated,T cc -peak shifted to higher temperature region approachingT(1) and simultaneously, the peak current ofT(1)-peak became smaller. Activation energy, time constant of dipolar relaxation and charge mobility were evaluated forT cc -peak. From these results, it was concluded that-form and more ordered- form crystalline phases coexisted in PBFPP once heated aboveT(1) and the content of- form phase increased by repeated thermal hysteresis.  相似文献   

2.
Measurements are reported of the interfacial tension in two-phase ternary systems of water(1) + hydrocarbon(2) + 2-butoxyethanol(3) atT=25°C along a linear path to the critical point (plait point). The two systems studied (with heptane and octane as the hydrocarbon component) exhibit a three-phase region and differ in respect of the distance of the upper critical end pointT u of this three-phase region from the experimental temperature (T>T u). Along the chosen linear path, a stepwise decrease of the tension with increasing concentration of the amphiphile, 3, is found: in region (i) decreases from its initial value (50 mN m–1) to about 2 mN m–1 due to the adsorption of the amphiphile; in region (ii) which extends over a wide composition range of the upper phase, exhibits an inflection point and is only a weak function ofw 3; on approaching the critical point there is a second inflection point in vs.w 3 and finally, in a rather narrow range of concentration (w 3 c -w 3), a follows the asymptotic critical-point scaling (region iii). In region (ii), the interfacial tension of the octane system (for whichT-T u is only about 1 K) falls below that of the heptane system and remains nearly constant, as expected for systems near three-phase coexistence.  相似文献   

3.
Electrical and dielectrical properties of poly(bis(-phenoxyethoxy)phosphazene) (I) and its complexes with various content ratios of AgSO3CF3 to monomeric unit (0.25/1 (II) and 0.5/1 (III) in molar ratio) were investigated.Dc conductivity of respective samples at 18 °C were 6.1×10–12, 4.4×10–9, and 7.1×10–8 S/m.Dc conduction was considered to be due to ion hopping. Charge mobility ranged from 3×10–12 to 6× 10–11 m2/Vs depending on the applied field in sample II. In sample I, a tan peak was found which can be ascribed to molecular relaxation of main chains. The peak vanished upon introducing AgSO3 CF3. Temperature dependence of total conductivity ( T ) measured byac method in the temperature range between –150 °C and 50 °C showed several peaks at the temperatures corresponding to the peak temperatures of tan. Total conductivities of respective samples at 100 kHz were 4.9×10–7 (69 °C), 1.7×10–4 (45 °C), and 1.5×10–4(40°C)S/m.  相似文献   

4.
The study was extended to analysis of mass, size and conformation of micelles formed in aqueous solutions of ethoxylated nonyl phenols. The results obtained by ultracentrifugal technique between 293 and 323 K have proved that the slightly ethoxylated nonyl phenols form micelles with high molecular mass and larger size at constant temperature, while the increasing length of the ethylene oxide chain favours formation of micelles of smaller molecular mass and size. The transformation of conformation from oblate to spherical shapes ensues with increasing temperature at constant ethoxy number or with ethoxylation at constant temperature. The second virial coefficient decreases with increasing temperature and decreasing ethoxy number. In accordance with the earlier conclucions, the change of the second virial coefficient relates to enhanced variation of monomer solubility, stabilization of micelle structure and increased deviation from ideal behaviour of a given micellar system.Symbols a major axis of micelle, Å - a m attractivity factor, cm3 erg molecule2 - b minor axis of micelle, Å - c concentration, g dm–3 - c b equilibrium concentration at the bottom of the cell, g dm–3 - c m equilibrium concentration at the meniscus of the cell, g dm–3 - c o initial concentration in the cell, g dm–3 - c M critical micellization concentration, mol dm–3 - e eccentricity - f IS Isihara-constant - f/f o frictional ratio of micelle - amount of water in micelle per ethoxy group, mol H2O/mol EO - n aggregation number, monomer micelle–1 - n EO number of ethoxy groups - r distance of Schlieren peak from the axis, cm - r b distance of cell bottom from the axis, cm - r m distance of cell meniscus from the axis, cm - R h equivalent hydrodynamic radius of micelle, Å - s t sedimentation coefficient, s - reduced sedimentation coefficient, s - reduced limiting sedimentation coefficient, s - ¯v t volume of micelle, cm3 micelle–1 - partial specific volume of solute, cm3g–1 - partial specific volume of solute reduced to 293 K, cm3 g–1 - B a, Be constants, cm3 mol g–2 - B 2 second virial coefficient, cm3 mol g–2 - M m a mass average apparent molecular mass of micelle, g mol–1 - M m mass average molecular mass of micelle corrected withB 2, g mol–1 - M m cM mass average molecular mass of micelle belonging toc M, g mol–1 - M 1 mass average molecular mass of monomer, gmol–1 - N A the Avogadro's number, molecule mol–1 - R universal gas constant, erg mol–1 K–1 - T temperature, K - t o dynamic viscosity of solvent atT temperature, g cm–1 s–1 - dynamic viscosity of solvent at 293 K, g cm–1 s–1 - t density of solution atT temperature, g cm–3 - t o density of solvent atT temperature, g cm–3 - density of solvent at 293 K, g cm–3 - angular velocity, rad s–1 - time, s  相似文献   

5.
A study has been undertaken of stress relaxation in ovalbumin thermotropic gels with a concentration of 8–20%, depending on time and temperature of heating (respectively, 20–60 min, 70°–110°C), at pH 2.5–10.0. In all instances, the dependence of the initial gel elasticity modulus on heating has a single maximum. Gelation conditions corresponding to this maximum are considered optimal. Optimal gelation time is 30 min, regardless of pH. On the other hand, the optimal heating temperature depends on pH. To the right and left of the isoelectric point of protein (2.5pH<4.0 and 5.5G) of gels on heating conditions, pH and protein concentration (X 1,X 2,X 3,X 4), as well as on time of relaxation (t) may be generally described asG(X 1,X 1,X 1,X 1,t)=G e(X 1,X 2,X 3,X 4)f(t), whereG e is the equilibrium value of the elasticity modulus, and f(t) the relaxation function. Thus, a change in the parameters only affects the value of the equilibrium elasticity modulus, and exerts no effect on the relaxation time spectrum. For this reason, all the relaxation curves obtained may be transformed into two normalized relaxation functions:f(t)=f(t)/f(1)=G(X 1,X 2,X 3,X 4,t)/G(X 1,X 2,X 3,X 4, 1)Each of these normalized functions corresponds to one of the homologous groups. Rheological similarity of gels in each homologous group evidently points to their structural similarity. Invariance of the gel relaxationproperties with regard to protein concentration, leads to a concentration dependence of the equilibrium modulus at various pH values. These dependences are curvilinear on a double logarithmic scale. The slope of the curve exceeds 2 in the entire concentration interval studied. In other words, the dependences obtained cannot be described by the usual law of squares. On the other hand, they adequately match Hermans theoretical relation for a network formed by random association of identical polyfunctional particles without cyclization. This simple model evidently gives a true picture of the major regularities of thermotropic gelation for ovalbumin. An agreement between this theory and experiment was achieved for a protein concentration ofC *=6.0±1.0% at the gel point regardless of pH. Invariance of gelpoint position with regards to pH demands further confirmation.List of symbols T h,t h heating temperature and time - T h * ,t h * optimal heating temperature and time - C protein concentration - C * protein concentration in gel-point - G relaxation modulus - G e equilibrium modulus - f(t) relaxation function - t time of relaxation - f(t) normalized relaxation function - fT A (t), f B (t) normalized relaxation functions of groups A and B - G 1 T h,t h-reduced modulus - G 2 T h,t h, pH-reduced modulus - G 3 C-reduced modulus - b 1 T h, th reduction parameter of modulus - b 2 pH reduction parameter of modulus - b 3 C reduction parameter of modulus - Wg gel-fraction  相似文献   

6.
Phase transitions of poly(N-isopropylacrylamide-co-acrylic acid) (PiPA-AA) and poly(N,N- diethylacrylamide-co-acrylic acid) (PdEA-AA) in water have been investigated by means of turbidimetry, Fourier transform infrared (FTIR) spectroscopy and differential scanning calorimetry (DSC). The phase transition temperatures (Tp) of these copolymers increase with the degree of ionization () of the acrylic acid (AA) units, which in turn is dependent on the pH of the solutions. Apparent values of pKa for the AA units, determined from the pH dependencies of Tp, are 4.7 and 5.4 for PiPA-AA and PdEA-AA, respectively. Differences between Tp for PiPA-AA and Tp for PiPA homopolymer (Tp) are +1.5 and –0.2 °C/mol% of AA at =1 and 0, respectively. The values of Tp for PdEA-AA are +2.6 (ionic) and –0.5 (nonionic)°C/mol%, indicating that the incorporated AA units have a larger effect on PdEA than on PiPA. DSC measurements performed with each of these copolymers at different pH values show a linear relationship between Tp and the enthalpy of transition (H). IR measurements of PiPA-AA show that the profiles of IR bands from both iPA and AA units exhibit critical changes at Tp of the copolymer. Heating the solution above Tp leads to shifts of the amide II, C–H stretch, and C–H bend bands from the iPA units toward lower wavenumbers, as well as a shift of the amide I band from the iPA units toward higher wavenumbers. A decrease in the intensity of the symmetric C=O stretch IR band from carboxylate anions (1560 cm–1), and an increase in the intensity of the C=O stretch band from COOH groups (1705 cm–1) suggest that a partial protonation of the carboxylate groups (COO+H+COOH) takes place upon the phase transition.  相似文献   

7.
Electron spin resonance (E.S.R.) studies of the bulk phase and film balance measurements of the monolayers at the air-water interface of some members of homologous rac-1-acyl-2-hexadecyl-glycerophosphocholines (1-(2C m -16:0)-2-H-PC with m = length of the side chain and PC = phospholipid) are reported.

The results of the E.S.R. measurements using fatty acid spin labelled near the terminal methyl group suggest that in the lamellar gel phase of the short-chain branched phospholipids (m < 9) the hydrocarbon chains are interdigitated. In the monobranched phospholipid with m = 14 a gel phase with non-interdigitated chains is assumed.

The F/A isotherms (F = film pressure and A = molecular area) of four branched-chain lecithins were measured over a wide range of temperatures. A comparison of the isotherms at similar reduced temperatures indicates that both condensed and liquid-expanded states occupy increasing molecular areas with increasing m. The phospholipid with m = 8 does not form condensed films.

The temperature interval of the transition region between the liquid-expanded and the condensed films which is confined by T o (lowest temperature at which the liquid-expanded film occurs) and the critical temperature T o decreases with increasing m and passes through a minimum at m ≈ 9. Although the T o-values are very similar, the T c -values are strongly dependent on m.

The values of T c for the monolayer transition and T m for the main transition in bulk are identical for m = O (unbranched) and for m = 14, whereas for the phospholipids with m = 3, 4 and 8 the values of T c are very much lower than values of T m . This is discussed in terms of the interdigitation of the lecithins in the gel phase.  相似文献   

8.
Three series of segmented polyurethanes based on MDI, variable chain extender, and polypropylene oxide of MW=1000, 2000, and 3000 were synthesized and their dielectric behavior examined.Dielectric relaxations in the segmented polyurethanes were investigated between –150°C and +150°C in the 100 Hz to 10 kHz range. In general, three transitions, designated as, , and were observed, and ascribed in accordance with calorimetric relaxations to glass transitions of the hard and soft segments, and Shatzki-type motions, respectively. The effect of structure variables such as soft segment size, type of chain extender (ethylene glycol, butane diol, and hexane diol) and soft segment concentration, as well as the effect of interaction of the phases on dielectric properties was discussed. It was found that a certain degree of phase mixing exists in all series, detected by the variation of theT g of the soft segment with soft segment concentration, contrary to DSC results, which was ascribed to thermal treatment prior to the dielectric measurements. It appears that interfacial polarization becomes important only above the transition temperature.  相似文献   

9.
A 0.5 dm3 aqueous solution of 0.1 M FeCl3 dissolving 1 wt% poly(vinyl alcohol) (PVA) was treated hydrothermally in a stainless steel autoclave at various temperatures (T h=110–200 °C). Highly ordered red corpuscle-like hematite particles around 2 m in diameter were produced after aging the solution at T h=110 °C for 7 days, though large numbers of spherical PVA microgels around 2–4 m in diameter were produced together with the red corpuscle-like particles at T h120 °C. The number of red corpuscle-like hematite particles decreased but that of spherical PVA microgels increased with increasing T h, leading to the proposal that the method carried out in the present study will become a new synthetic method of polymer microgels. The ferric ions acted as a cross-linking agent to make PVA insoluble in water. The red corpuscle-like hematite particles produced at T h=110 °C had high specific surface areas and showed high mesoporosity. The mesoporosity appeared to be more pronounced after evacuating the particles above 300 °C. The diameter of the mesopores after evacuation above 300 °C ranged from 2 to 20 nm, with a maximum at around 5–6 nm. The H2O and N2 adsorption experiments revealed that there are no ultramicropores in the particles. The H2O and CCl4 adsorption experiments further disclosed that the surface hydrophobicity of the particles is low even though PVA molecules remain after evacuation of the particles at 100–400 °C. Furthermore, the micropores produced after evacuation of the particles at 400 °C exhibited a high size restriction effect, i.e., the micropores produced were accessible to H2O (diameter 0.253 nm) and N2 (diameter 0.318 nm) molecules but not to CCl4 (diameter 0.514 nm).  相似文献   

10.
The groundstate-stable dimers in the non-crystalline regions of uniaxially drawn poly(ethylene terephthalate) (PET) films were used as chain-intrinsic fluorescent labels for studying the orientation distribution in the non-crystalline regions. As far as indicated by the spectra and the fluorescence decay law, the fluorescent group remains unchanged when the sample is uniaxially oriented by drawing above the glass transition temperature. In contrary to the behaviour of physically incorporated probe molecules, the orientation coefficient f 2 F of the dimers is proportional to the amorphous orientation coefficientf 2 A ; concerning the fluorescence signal, amorphous includes all the material outside the perfect crystal.During deformation, the orientation coefficientf 2 F follows approximately a superposition curve of crystallite-like orientation, separable in the initial range of stretching ratio <2.5, and of true-amorphous orientation of the affine network type that becomes noticably at >3.At temperatures closely aboveT g, and within the selected range of stretching parameters, the fluorescence intensity of PET remains nearly constant with increasing stretching ratio; at >2.5, where the maximum crystallite orientation is achieved and the increase of amorphous orientation becomes noticably, a slight augmentation of the dimer concentration is observed.  相似文献   

11.
An isotactic polypropylene film was stretched at 120 °C in poly(ethylene glycol) and thermally shrunk at various temperatures. Proton spin-lattice,T 1, and spin-spin,T 2, relaxation times were measured using a broad line pulse spectrometer operating at 19.8 MHz in the temperature range 40 °C–100 °C. The temperature ofT 1 minimum shifts to higher temperatures and the value ofT 1 minimum increases in magnitude as the stretching ratio is increased. In contrast the temperature ofT 1 minimum shifts to lower temperatures as shrinkage is increased, whereas the value ofT 1 minimum increases in magnitude because of the increase in crystallimty during shrinkage. T2a, the longestT 1 associated with the mobile amorphous regions, increases during shrinkage, indicating that chain mobility in the amorphous regions increases substantially during shrinkage. It was found that an orientation function of the amorphous regions,f a, correlates well withT 2a .Presented in part at the 52nd Annual Meeting of the Japan Chemical Society, Kyoto, April 1986.  相似文献   

12.
Structural changes at annealing temperatures (Tan) of 500–1,100°C were investigated for thin Ta films which were sputter-deposited onto pure Si substrates and onto thermally oxidized Si. In the as-deposited state, the Ta layers predominantly consist of metastable tetragonal -Ta, whereby the [001] texture is independent of the substrate material. At lower annealing temperatures, the microstructural evolution is essentially the same for both Ta films. Incorporation of O atoms causes an increase of the intrinsic compressive stress, and diffusion of C atoms into the Ta layer leads to the formation of Ta2C. Additionally, a partial transformation of the original -Ta phase into a second phase with tetragonal unit cell (denoted as -Ta) occurs. For the Ta/Si system, the formation of a Ta–Si intermixing layer is initiated at Tan=550°C, and nucleation of crystalline TaSi2 occurs at Tan=620°C. The formation of a second Ta silicide was not detected up to Tan=900°C. In the case of the Ta film deposited onto the SiO2 substrate, the metastable -Ta and the -Ta transform completely into the thermodynamically stable cubic -Ta at Tan=750°C. A marked reaction with the substrate indicated by the formation of Ta2O5 and Ta5Si3 occurs at Tan=1,000°C.  相似文献   

13.
Poly(parabanic acid)-CuS composite film (wt-% of CuS=20–50) prepared by using organosol of CuS (=1500 Å) showed electrical conductivity of 0.1–70 S cm–1, high thermal stability up to 250°C, high mechanical strength (breaking stress=7.0–12 × 107 Pa), and good shielding effect of electromagnetic wave.  相似文献   

14.
The variation of amorphous orientation and crystalline regularity of hard elastic polypropylene (HEPP) films during cyclic deformation and stress relaxation processes were studied using a FTIR spectrometer. The result proves entropic elasticity and shows the orientational hysteresis in the amorphous region or within the microfibrils, and also shows that the amorphous orientation increases, but that the crystalline regularity decreases with the increase of extension rate.Three spin-spin relaxation timesT 2f,T 2m, andT 2s and associated mass fractionsF f,F m, andF s of HEPP fibers were measured with a solid echo of NMR method at different elongations and after relaxation or recovery for a long time A new possible interpretation was proposed that, while the microfibrils are formed in HEPP, the medium decay component should be ascribed to inner molecules of the microfibrils, and the slow decay component to the surface molecules of the microfibrils. According to this interpretation, the results implied that subfibrillation is the main process when HEPP is stretched up to 15% strain, and that at above 15% strain thinning and lengthening of the microfibrils become the main process. Thickening of the microfibrils was found in the recovery and relaxation processes.  相似文献   

15.
Organosols of ZnS and HgS in N,N-dimethylformamide (DMF) and dimethyl sulfoxide (DMSO) were prepared by reaction of Zn and Hg salts with H2S. ZnS in the DMF organosol has a particle size of about 300 Å, determined by an argon laser scattering technique. The form of HgS in the organosols varies with temperature, concentration of HgS, and type of solvent;-HgS in a DMF organosol is converted into-HgS at –30 °C or above. Poly(acrylonitrile)-ZnS or -HgS composite films prepared from organosols are semiconductive, and the electrical conductivity of the film increases by a factor of 102–103 on exposure to moisture or alcohol vapour. Poly(acrylonitrile)--HgS composite films, prepared from heated-HgS, show electrical conductivity of a range of 10–1–10–2 S cm–1 and a large electron mobility.  相似文献   

16.
Viscosity measurements were made in the temperature range of 10 °–40 °C. The equation= o exp(B/(T-T o )) was used with the parameterT 0 as structure indicator, which is called the limiting temperature. For instance, hydrocarbons, as liquids with quasifree molecules, haveT 0=O; water as a highly structured liquid hasT 0= 140–150 K.The polymer investigated was ovalbumin in aqueous solution in a concentration comparable to that of blood. Acetylsalicylic acid produces a protein conformation which breaks the water structure in solution at a pH of within the in vivo region.The question of whether only the acidity determines the water structure breaking properties of the protein is investigated by acidifying albumin-water solutions with hydrochloric acid, lysine chloride and ascorbic acid. All these acids exhibit similar effects. A stronger influence is observed for ammonium chloride. Its interaction with ovalbumin produces a strong structure-breaking effect. The most powerful water structure breaker in albumin-water solutions is dextrane. In a concentration of 10 % it changes the polymer conformation so that the water structure is broken to such an extent that the solution behaves as an almost quasifree liquid withT 0=O.  相似文献   

17.
The pressure in the Reverse Wilson Chamber (RWC) was directly measured at different rates of compression of the gas mixture. It was shown that at compression time in the range from 0.06 to 0.3 s an intermediate, between adiabatic and isothermal, process took place in the chamber. To obtain the relative pressure increase P m /P at from the values of the relative gas compression V/V, a calibration of the experimental set-up was carried out. The calibration showed that the values of critical supersaturationSc for water condensation on hexadecane, estimated for intermediate regime of the gas compression, were reduced with respect to the values calculated when the adiabatic regime was assumed. This fact confirmed the conclusions made earlier [1–3] that the classical theory was not applicable in this case of heterogeneous phase formation and that the line tension < 0 should be taken into account. Moreover, in an atmosphere of very pure argon (instead of room air [1–3]) the critical supersaturation turned out to be independent of the initial state of undersaturationS o . The more accurate values obtained for condensation of water on hexadecane were: lnS c =0.204 (instead of the maximum value obtained earlier: lnS c =0.26) and=–1.9×10–5 dyne (instead of=–1.5×10–5 dyne).  相似文献   

18.
The diffusion of six azo and five anthraquinone derivatives through nylon 6, poly(ethylene terephthalate) and secondary cellulose acetate films were studied under high hydrostatic pressures of up to 3000 bar and at temperatures 80–130 °C, by analyzing the diffusion profiles yielded in a stacked multiple film, placed in the solution of the diffusant. It was found that the diffusion coefficient,D, of the diffusant decreased with increasing pressure, giving a linear relationship between InD and the pressure, the slope of which gave the activation volume for the diffusion,V . It was revealedV increased linearly with increasing intrinsic molecular volume of the diffusant,V w , the slopes being different between the azo and the anthraquinone derivatives. The ratio ofV toV w (V /V w ) ranged from 0.13 to 0.93, depending in a sensitive manner on the degree of swelling of the polymer matrix which in turn was varied by the solvent. The overall results could be explained in accordance with the formulation,V f, local +V =V w , whereV f, local represents the free volume contribution. It was proposed thatV w is increased by solvation when the solvent is good for the diffusant.  相似文献   

19.
Following the earlier study of the- and-casein micelle structure, we will now report results from the s1-casein. Static and dynamic light scattering measurements were performed in a concentration range from 0.5 to 6.0 mg/ml atT=35 °C. A constant apparent molecular weight of 3.4×106 daltons was found over the whole range. The apparent radii of gyration and the diffusion coefficients also show no detectable concentration dependence. The ratio of the two radiiR g /R H =2.78+0.21 is characteristic of extended rigid structures.R g is the radius of gyration andR H the hydrodynamic radius defined via the Stokes-Einstein relationship from the translational diffusion coefficient. This is in agreement with the analysis of the pronounced angular dependence of the scattered light, which leads to the conclusion that s1-casein forms very long worm-like micelles. The contour length of one cylinder was found to beL1600 nm and the chains appear to be composed of about 12 Kuhn segments. At higher concentrations, lateral aggregation proportional to the concentration is observed. Beyond the overlap concentrationc * the asymptotic scattering curve changes its shape, which is interpreted as the beginning of a reversible gelation.  相似文献   

20.
The structure-property relationships derived here permit the prediction of both the zero-shear viscosity 0, as well as the shear rate dependent viscosity . Using this molecular modeling it is now possible to predict over the whole concentration range, independently of the molecular weight, polymer concentration and imposed shear rate. However, the widely accepted concept: dilute — concentrated, is insufficient. Moreover it is necessary to take five distinct states of solution into account if the viscous behavior of polymeric liquids is to be described satisfactorily. For non-homogeneous, semi-dilute (moderately concentrated) solutions the slope in the linear region of the flow curve (= must be standardized against the overlap parameterc · []. As with the 0-M-c-relationship, a-M -c- relationship can now be formulated for the complete range of concentration and molecular weight. Furthermore, it is possible to predict the onset of shear induced degradation of polymeric liquids subjected to a laminar velocity field on the basis of molecular modeling. These theoretically obtained results lead to the previously made ad hoc conclusion (Kulicke, Porter [32]) that, experimentally, it is not possible to detect the second Newtonian region.Roman and Italian symbols a exponent of the Mark-Houwink relationship - b exponent of the third term of the 0-M -c relationship - c concentration /g · cm–3 - E number of entanglements per molecule - F(r) connector tension - f function - G i A shear modulus; A indicates that it /Pa has been evaluated by a transient shear flow experiment; i is the shear rate to whichG A refers to - G storage modulus /Pa - G p plateau modulus /Pa - H() relaxation spectrum /Pa - h shift factor (0/r) - K H Huggins constant - K b third constant of the 0-M -c relationship - K constant of the Mark-Houwink relationship - M molecular weight /g · mol–1 - M e molecular weight between two /g · mol–1 entanglement couplings - N number of segments per molecule - n slope in the power-law region of the flow curve - p p-th mode of the relaxation time spectrum - R gas constant /8.314 J·K–1·mol–1 - r direction vector - T temperature /K Greek symbols ß reduced shear rate - shear rate /s–1 - shear viscosity /Pa·s - s solvent viscosity /Pa·s - sp specific viscosity - 0 zero-shear viscosity /Pa·s - apparent viscosity at shear rate - reduced viscosity - viscosity of polymeric liquid in /Pa·s the second Newtonian region - [] intrinsic viscosity/cm3·g–1 - screening length/m - /g·cm –3 density - relaxation time/s - 0 experimentally derived relaxation time/s - angular frequency of oscillation Indices conc concentrated - corr slope corrected - cr critical - deg degradation - e entanglement - exp experimental - mod moderately concentrated/semi-dilute - n number average - p polymer - R Rouse - rep reptation - s solvent - sp specific - theo theoretical - weight average - relaxation time - o experimental or steady state - * critical - ** transition moderately conc. — conc. - + transition dilute — moderately cone. Paper presented at the 2nd bilateral U.S.-West German Polymer Symposium, Yountville, the 7th–11th September 1987.  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