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 937 毫秒
1.
T. Bleha  J. Mlýnek  D. Berek 《Polymer》1980,21(7):798-804
A theoretical model has been developed for the concentration effect in gel chromatography, i.e. the dependence of the elution volume Ve on the concentration of injected polymer c. On the basis of the theoretical relations of Yamakawa and Eizner for coil shrinkage with increasing concentration in the range of dilute polymer solutions, relations predicting the extent of the concentration effect have been derived. A comparison of the calculated and experimental data for polystyrene in tetrahydrofuran and toluene has shown that both theories slightly underestimate the extent of the concentration effect but qualitatively correctly describe its dependence on molecular mass M and on thermodynamic quality of an eluent given by the product A2M, where A2 is the second virial coefficient of the polymer-eluent system. The proposed model explains the recently established correlation between the slope of the concentration dependence of Ve and the thermodynamic quality of the eluent and theoretically accounts for the method for estimating the coefficient A2 from gel chromatographic measurements. The possibility of using the measurements of concentration effects for examining the reliability of the theoretical relations for coil shrinkage with concentration in dilute polymer solution as well as for eventual semiempirical modification of these relations is examined.  相似文献   

2.
A polystyrene (PS) with M?w = 9,7 · 104 was investigated by means of light scattering in the isorefractive polymer/solvens-mixture polymethylmethacrylate (PMMA)/benzene. It was found, that the second osmotic virial coefficient A2 of PS was strongly dependent on the average viscosimetric molecular weight M?v and on the concentration of PMMA, but scarcely on the temperature in the range of 20°C to 60°C. The θ-Point, where A2 is zero, was independent of the temperature within experimental error. By defining the PMMA concentration at the θ-Point as cθ, and by reducing the measured PMMA concentration c to c/cθ, an unequivocal relation was obtained between A2 and c/cθ, which is independent of molecular weight and molecular weight distribution of PMMA. PS shows a high second virial coefficient in dimer and trimer MMA as well as in non-hydrogenated and hydrogenated MMA. The investigated PS constitutes a θ-System in PMMA of a degree of polymerisation of P?w, ~ 17without the use of benzene.  相似文献   

3.
Poly(methacrylic acid) hydrogels were synthesized. The effects of the synthesis parameters: the neutralization degree of methacrylic acid and the concentrations of monomer, crosslinker and initiator on the xerogels structural properties: the xerogel density (ρxg), the number average molar mass between the network crosslinks (Mc), the crosslink degree (ρc), the number of elastically effective chains totally induced in a perfect network per unit volume (Ve), the distance between the macromolecular chains (ξ) and the equilibrium swelling degree (SDeq) and the swelling kinetics was investigated. As the concentrations of crosslinker, monomer and initiator increase, the value of ρxg, ρc and Ve increases and decreases the value of Mc, ξ and SDeq. With the increase in the neutralization degree of methacrylic acid, the values of ρxg, Mc, ξ, SDeq increase, while the ρc and Ve decrease. The xerogels structural properties, SDeq and swelling kinetic parameters are mainly in power law form functional relationships with the synthesis parameters as well as with the xerogels crosslinking degree.  相似文献   

4.
T. Spychaj  D. Berek 《Polymer》1979,20(9):1108-1114
Chromatogram shapes, concentration effects and calibration curves in g.p.c. were studied for dry tetrahydrofuran and its mixtures with water up to 8.9 vol. %. Polystyrene reference materials were separated with silica-based column packings. The results can be generalized fairly well and lead to the conclusion that the gel chromatographic data are considerably influenced by humidity present in the eluent. Ghost-peaks appeared in the domain of high elution volumes (Ve) using mixed eluents. Simultaneously, the solute peaks changed their widths and the slope of the plot of Veversus injected polymer concentration changed with water content. However, most important were the shifts in the polymer elution volumes that were caused by the complex of interactions in the system gel-mixed eluent-solute and depended on the amount of water present in eluent. Thus, the dry eluent is an inevitable condition for obtaining g.p.c. results of high precision and reproducibility. The use of a guard column filled with a sorbent strongly trapping water is proposed in g.p.c. with hygroscopic eluents. Similar precautions are to be considered for any admixture in the eluent, especially if its content cannot be kept constant and if its polarity differs substantially from the polarity of the eluent.  相似文献   

5.
Gel permeation chromatography (GPC) calibration curves log hydrodynamic volume versus elution volumes were obtained from a series of single and mixed eluents for polystyrene on inorganic carriers. The observed calibration shifts were interpreted as a result of adsorption and partition effects on elution volume Ve. Secondary contributions to the separation mechanism can be qualitatively described by means of thermodynamic relations from liquid adsorption (LSC) and liquid partition (LLC) chromatography using the solubility parameters of the eluents. The universal calibration procedure based on the hydrodynamic volume of the coils is directly applicable only for systems in which adsorption and partition is approximately the same.  相似文献   

6.
The elastic behavior of concentrated solution of acrylonitrile copolymer was investigated by the capillary end correction method. The results were as follows. (1) The shear stress is proportional to recoverable shear strain in accordance with Hooke's law below critical concentration; above a critical concentration, however, the shear modulus depends on shear stress. (2) The log–log plots of zero shear modulus against polymer concentration and molecular weight fall on two straight lines with different slopes. The intersection of lines is considered to be the onset of elastically deformable entanglement network. We denote this inflection point as (Cc)e or (Mc)e. (3) The log–log plot of viscosity against polymer concentration does not show a change of slope at the critical concentration (Cc)e. (4) By the application of the kinetic theory of rubberlike elasticity to the pseudo-network structure of concentrated polymer solution, in the range of Cc < C < (Cc)e or Mc < M < (Mc)e, the number of chain entanglement per molecule is kept one; moreover, in the range of C > (Cc)e, or M > (Mc)e, the number of chain entanglement increases to three.  相似文献   

7.
A. Bennett  R. Shanks 《Polymer》2004,45(25):8531-8540
We have used small angle neutron scattering and dynamic light scattering to measure the static and hydrodynamic screening lengths of polystyrene, polymethylmethacrylate and polydimethylsiloxane solutions ranging from marginal to good solvent quality. A universal plot is found for the scaled static screening length when the concentration is scaled using the second virial coefficient in the way suggested by renormalization group theories. The same concentration units do not produce a universal plot for the hydrodynamic screening length at the molecular weights that we have studied (all around 1-2×105 g/mol). However, when the concentration is expressed in terms of kDc, where kD is the virial expansion coefficient for the cooperative diffusion coefficient and c is the concentration, most of the variation between different polymer-solvent combinations is eliminated. The ratio of hydrodynamic screening length to static screening length increases with concentration for all of the polymer solvent pairs studied, and its value differs for different polymer solvent pairs.  相似文献   

8.
Weight average molar masses (Mw) and second virial coefficients (A2) have been measured for linear and cyclic poly(dimethyl siloxane) (PDMS) fractions in toluene at 298K and for linear PDMS in bromocyclohexane at 301K. The values of Mw are compared with those deduced previously using gel permeation chromatography, broadly confirming the values already assigned to the fractions. The values of A2 for linear PDMS in toluene are shown to be consistent with previously published values for oligomeric linear PDMS. The values of A2 for cyclic PDMS approach those of linear PDMS for Mw<~1000 g mol?1 and decrease more rapidly as Mw increases in approximate agreement with theoretical predictions. In addition, the conventional relations between A2 and the expansion factor αs are shown to be inapplicable at low molar mass.  相似文献   

9.
Several narrowly disperse polystyrenes (PS) were mixed with three series of random copolymers of 2,6-dimethyl phenylene oxide (also known as xylenyl ether (XE)) and 3-bromo-2,6-dimethyl phenylene oxide (BrXE), each series having a fixed chain length. The critical BrXE comonomer concentration necessary to induce phase separation, xc′ was characterized for each blend series by a number of techniques. xc was extrapolated to a value of about 0.5 at infinite chain lengths of each polymer of the blend. This finding supports the negative heat of mixing of PS with PXE reported by other workers. Small-angle neutron scattering studies were carried out using small amounts of perdeuteropolystyrene in several matrices for which single phase behaviour had been shown to exist. The radius of gyration R2 and the second virial coefficient A2 were greatest in the pure PXE matrix reflecting the goodness of this material as a ‘solvent’ for PS. In matrices containing 50% PS50% copolymer, A2 extrapolated to zero at xc ≥ 1 which was consistent with the phase behaviour of this series of blends.  相似文献   

10.
W. Mandema  H. Zeldenrust 《Polymer》1977,18(8):835-839
The diffusion of polystyrene in tetrahydrofuran has been studied by line width measurements in a laser light scattering spectrometer. The diffusion coefficient D has been determined as a function of concentration c for the molecular weight range of 2.04 × 104 to 1.8 × 106. At low concentration the relationship between D and c is linear to a very good approximation over a relatively large concentration region, as has also been found for some other systems. The molecular weight dependence of D, extrapolated to zero concentration, D0 can be written as D0 = kTM?bw, where b = 0.564, in good agreement with thermodynamic predictions derived from intrinsic viscosity studies.  相似文献   

11.
The diffusion coefficients of polystyrene (PS) in decahydronaphthalene (DHN) and in solutions of carbon dioxide (CO2) and DHN were measured for dilute PS solutions over a range of temperatures and CO2-DHN ratios using high pressure dynamic light scattering. Infinite dilution diffusion coefficients (D0) of PS and dynamic second virial coefficients (kD) were determined for essentially monodisperse 308 kDa PS. At a system pressure of 20.7 MPa, PS diffusion coefficients increased by a factor of 2.5, and the activation energy of diffusion decreased by approximately 16% when DHN was “expanded” with 44 mol% CO2. However, the hydrodynamic radius of PS at a given temperature was not particularly sensitive to the CO2 concentration. Solvent quality, as measured by kD, decreased at higher CO2 concentrations. The addition of CO2 to polymer solutions may offer a way to “tune” the properties of the solution to facilitate the heterogeneous catalytic hydrogenation of polymers.  相似文献   

12.
Crosslinked 1-octene-isodecyl acrylate copolymers were synthesized and evaluated for oil-absorbency application. The copolymer was crosslinked at different concentrations of ethylene glycol diacrylate (EGDA) and ethylene glycol dimethacrylate (EGDMA) crosslinkers via catalytic initiation or by electron-beam irradiation at dose rate 80 kGy. The concentration of both crosslinkers was varied from 0.5% to 2%. The effect of crosslinking conditions, such as crosslinker concentration, method of polymerization and monomers concentration on conversion and gel fraction was examined through an oil-absorption test using petroleum crude oil. It was found that, the oil absorbency was influenced mainly by the degree of crosslinking and the hydrophobicity of the copolymer units. The final equilibrium oil content, volume fraction of polymer and swelling capacity were determined at 298 K. The effective crosslinking density Ve, the average molecular weight between the crosslinks Mc and the polymer-toluene interaction parameter were determined from stress–strain measurements. The crosslinking efficiencies of EGDA and EGDMA towards copolymers were determined.  相似文献   

13.
Polystyrene and poly(phenylene oxide) are miscible over the entire range of compositions. Thin films of five blends of high molecular weight polystyrene (PS) with high molecular weight poly(phenylene oxide) (PPO), and four blends of low molecular weight PS (whose molecular weight lies below its entanglement molecular weight Me) with the same PPO have been prepared. Following bonding of these films to copper grids, crazes were grown by uniaxial straining in air. Suitable crazes were then observed by transmission electron microscopy. From microdensitometry of the image plates it is possible to measure the extension ratio λcraze within crazes in the nine blends. These measured values are compared with predicted values of λmax, computed from λmax = Ied, where Ie is the chain contour length between entanglements and d is the root mean square end-to-end distance for a chain of molecular weight Me. For the high molecular weight PS blends λmax depends on the entanglement properties of both PS and PPO chains. For the low molecular weight PS blends, the PS chains cannot form part of the entanglement network and the correct value of λmax is obtained from appropriate scaling of the pure PPO value. Comparison of λcraze and λmax for both types of blends shows excellent agreement, demonstrating the importance of the entanglement network in determining craze parameters and hence the toughness of a given polymer.  相似文献   

14.
Tensile retraction measurements of Mc were made on a variety of polyurethanes or polyurethane-ureas. These samples were prepared using a number of prepolymer MW and functionalities. In addition, the stoichiometry of the cure was allowed to range widely. In all cases, excellent linear correlation coefficients were obtained between the Mc and the Amax of the test. The slope and Mr (the value of Mc extrapolated to Amax = 1) determined from this was used to characterize the polymer. These data allowed the determination of χ for each of the different systems such that comparable values of Mc could be measured by swelling. Identical formulations cured at different temperatures were shown to have measurable differences in Mc, slope, and percentage set. These variations appear to relate to the perfection of the network and morphology of the hard domains.  相似文献   

15.
Emulsion polymerization of vinylacetate leads to branched polymers which at high monomer conversions form microgels of the shape and size of the latex particles. Quasielastic light scattering measurements from samples in the pre-gel state give at small q2 a linear angular dependence of Dapp = Гq2 which resembles that of randomly branched chain molecules, where Г is the decay constant of the time correlation function. Extrapolation of Dapp towards zero scattering angle yields the translational diffusion constant Dz. The diffusion constant follows the molecular weight dependence Dz = 9.78 10?5Mw?0.478. The diffusion constant of the microgels, i.e. at molecular weights Mw > 14 106, remains constant because of the finite and constant size of the latex particles. The coefficients kf and kD in the concentration dependence of the frictional and diffusion coefficients are related according to the equation kD = kf ? 2A2Mw ? v? where A2 is the second virial coefficient and v? the partial specific volume of the particle. The coefficient kf is calculated from the experimentally determined quantities kD, A2 and Mw, and the result is compared with the theory by Pyun and Fixman. Accordingly the branched coils in the pre-gel state resemble soft spheres, but the microgels behave more like spheres of some rigidity.  相似文献   

16.
The mass action kinetic model of the irreversible Michaelis-Menten reaction mechanism is mathematically intractable: an explicit analytical solution cannot be obtained. This difficulty is overcome by applying simplifying kinetic assumptions but a full understanding of their dynamic implications and applicability is not readily available. This paper shows how simple modal analysis can provide both a conceptually appealing insight into the reaction dynamics and justification of the commonly used quasi-steady-state and quasi-equilibrium assumptions.The key results are that the quasi-steady-state assumption is applicable when the initial enzyme concentration, e0, is much smaller than the Michaelis constant, Km, or when the initial substrate concentration, s0, is much greater than Km. These results show that the commonly accepted criterion e0 < <s0 is incomplete and should be decomposed into e0 < <Km and Km <<s0. The quasi-equilibrium assumption is valid when e0 ⪢⪢ Km and when the rate of product formation is much slower than reversion to the substrate from the intermediate state, or k2 ⪢⪢ k−1. The important dimensionless parameter ratios characterizing the reaction dynamics are e0/Km, s0/Km and k2/k−1.  相似文献   

17.
The temperature-composition phase diagrams for six pairs of diblock copolymer and homopolymer are presented, putting emphasis on the effects of block copolymer composition and the molecular weight of added homopolymers. For the study, two polystyrene-block-polyisoprene (SI diblock) copolymers having lamellar or spherical microdomains, a polystyrene-block-polybutadiene (SB diblock) copolymer having lamellar microdomains, and a series of polystyrene (PS), polyisoprene (PI), and polybutadiene (PB) were used to prepare SI/PS, SI/PI, SB/PS, and SB/PB binary blends, via solvent casting, over a wide range of compositions. The shape of temperature-composition phase diagram of block copolymer/homopolymer blend is greatly affected by a small change in the ratio of the molecular weight of added homopolymer to the molecular weight of corresponding block (MH,A/MC,A or MH,B/MC,B) when the block copolymer is highly asymmetric in composition but only moderately even for a large change in MH,A/MC,A ratio when the block copolymer is symmetric or nearly symmetric in composition. The boundary between the mesophase (M1) of block copolymer and the homogeneous phase (H) of block copolymer/homopolymer blend was determined using oscillatory shear rheometry, and the boundary between the homogeneous phase (H) and two-phase liquid mixture (L1+L2) with L1 being disordered block copolymer and L2 being macrophase-separated homopolymer was determined using cloud point measurement. It is found that the addition of PI to a lamella-forming SI diblock copolymer or the addition of PB to a lamella-forming SB diblock copolymer gives rise to disordered micelles (DM) having no long-range order, while the addition of PS to a lamella-forming SB diblock copolymer retains lamellar microdomain structure until microdomains disappear completely. Thus, the phase diagram of SI/PI or SB/PB blends looks more complicated than that of SI/PS or SB/PS blends.  相似文献   

18.
Solutions of the random coiling polymers; polystyrene, poly (methylmethacrylate), and sodium polystyrene sulfonate (NaPSS), all at concentrations well below the critical value for entanglement, were subjected to transient, high, elongational strain rates by passage from a cylinder through an orifice into a gently diverging section, driven by a piston at constant, high velocity over a short stroke. It is shown that a critical orifice flow velocity Vc. exists for each polymer species, above which scission of polymer molecules occurs creating new molecules. By gel permeation chromatography, the number of additional polymer molecules created per initial polymer molecule, the scission index, was determined as M n,0/M n – 1 where M n is the number average molecular weight, and M n,0 is the initial value thereof. Vc is found to vary as approximately M . Above Vc the scission index was found to be proportional to M , to the difference: orifice velocity V less Vc, and to the number of passes N of the polymer solution through the orifice. Expansion of NaPSS coils by reducing ionic strength of their aqueous solutions, at constant polymer molecular weight, decreases the scission index, The hypothesis is proposed that intramolecular entanglements are responsible for scission. The random coiling macromolecules in the solution cannot respond to the strain rate (imposed in ca, 100 microseconds) so as to avoid having internal sections, caught by loop entanglement, pulled to nearly full extension and thus broken.  相似文献   

19.
The force constants of the square-well potential for normal fluids were correlated in terms of the critical properties Pc, Tc, and the acentric factor ω, by means of a non-linear regression technique. The second virial coefficients, a total fo 100 points chosen uniformly with respect to Pr and ω from the compilation of Dymond, together with the compressibility factors, taken from the generalized tables of Pitzer. were employed in the correlation. The valid ranges of the correlation are 0.5 ?Tr ? 2·0 and 0 ? ρr ? 1·0. The calculated results for the second virial coefficients compare favourably with the generalized correlation of Pitzer and Curl. The calculated third virial and cross third virial coefficients, and the compressibility factors also agree well with the literature values.  相似文献   

20.
Adsorption equilibrium constants for methyl oleate and methyl linoleate in vapor phase on supported copper and nickel catalysts have been determined using the technique of pulse gas chromatography. The results are discussed in relation to selectivity in fat hydrogenation. Notation: A, column cross-section, m2 ; an,bn, nth Fourier coefficients; c, concentration of adsorbate in bulk flow, mol/m3 ; c* = c/ ∫ 0 cdt, normalized concentration of adsorbate in bulk flow; Ci, concentration of adsorbate in catalyst pores, mol/m3 ; ca, concentration of adsorbate on catalyst surface, mol/kg; cTOT, active area of catalyst as measured by hydrogen adsorption, mol/kg; De, effective diffusion coefficient of adsorbate in catalyst, m2/s; Dea, axial dispersion coefficient based on void cross-section, m2/s; hn, nth coefficient in Hermite polynomial expansion; Hn nth Hermite polynomial; ΔHA, adsorption enthalpy, kJ/mol; ΔHvap , heat of vaporization, kJ/moll; ka, adsorption rate constant, m3/kgs;KA, adsorption equilibrium constant, m3/kg; K0 , preexponential factor defined in Eqn. 8, m3/kg; kf, mass transfer coefficient, m/s; L, bed length, m; q, flow rate, m3/s; R, particle radius, m; Rg, gas constant; t, time, s; T, temperature, K; TF, period of Fourier expansion, s; u = q/A, linear velocity, m/s; z, length coordinate in packed column, m. Greek symbols: δ(t), Dirac delta function; ∈B, void fraction of bed; ∈-p, particle void fraction, ρrp, particle density, kg/m3 ; ξ, radial coordinate in particle, m; μ1, first absolute moment, μ2, second central moment.  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