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 15 毫秒
1.
The miscibility of poly(methyl methacrylate) (PMMA) and styrene-acrylonitrile random copolymers (SAN) blends was investigated on the basis of the Flory—Orwoll—Vrij equation of state theory. To obtain the equation of state parameters (P*, V*sp, T*: characteristic parameters), the pressure—volume—temperature (PVT) behaviour was measured for PMMA and a series of SANs with various acrylonitrile contents. The exchange energy parameter Xij was also calculated by fitting the theory to some phase diagrams of PMMA/SAN blends. The Flory—Huggins interaction parameter χ was separated into two contributions based on the equation of state theory for mixtures: the exchange energy term χinter and the free volume term χfree. Both the temperature and copolymer composition dependences of χinter and χfree were estimated by calculations using the equation of state parameters. There exists a region in which χinter is negative, leading to a miscibility window in PMMA/SAN blends. However, the immiscibility at high temperatures in the blends cannot be explained only by χinter; it is caused by the free volume contribution, χfree. The miscibility window behaviour in PMMA/SAN blends may be explained within the framework of the equation of state theory.  相似文献   

2.
O.S. Tyagi  S.M. Sajjad  Sajid Husain 《Polymer》1987,28(13):2329-2334
Inverse gas chromatography was employed to determine the apparent polymer-polymer interaction parameter χ′23,app for the following blends using them as binary stationary phases: poly(vinyl acetate)-poly(n-butyl methacrylate) at 100 and 120°C, poly(vinyl acetate)-atactic poly(vinyl isobutyl ether) at 70°C, and poly(n-butyl methacrylate)-atactic poly(vinyl isobutyl ether) at 70°C. The interaction parameter χ′23,app depended significantly on the chemical nature of the solvent (probe) used and the composition of the stationary phase. The lowest values of χ′23,app were obtained when aromatic and chlorinated-alkane probes were eluted on stationary phases having weight fractional compositions of the component polymers in the range 0.4–0.6. The results predict a better compatibility for poly(n-butyl methacrylate)-atactic poly(vinyl isobutyl ether) than for the other blends.  相似文献   

3.
J.P. Montfort 《Polymer》1976,17(12):1054-1058
Narrow molecular weight distribution polystyrenes and their blends, with molecular weights exceeding Mc have been measured in the molten state by means of a Contraves-Kepes balance rheometer. The vertical and the horizontal shift factors are the same for the fractions and their blends. In the terminal zone, the effect of blending on the complex viscosity η* and its imaginary η″ results in the appearance of two fields representing the respective contributions of both components to the blend. The importance of the coupling between both fields as a function of the weight ratio of both components is discussed. The zero-shear viscosity η0 is equal to that of a narrow molecular weight distribution polystyrene having the same weight-average molecular weight Mw measured at the same temperature.  相似文献   

4.
Characteristic electrochemical transport parameters for an experimental poly(ethylene)terephtalate (PET) track-etched membrane with well-defined structure and low porosity (Θ=0.13%) were determined with the membrane in contact with KCl solutions at different concentrations. Membrane potential, Δφm, measurements were performed to investigate the effective fixed charge concentration, Xf, and transport number of the ions, ti, in the membrane using two different procedures: keeping the concentration ratio constant, or keeping one concentration constant and changing the other one. Results show the membrane presents a weak cation-exchanger character, since the following values were obtained: Xf=−(2.5±0.2)×10−2 M, tK =(0.56±0.06), tCl=(0.44±0.05); taking into account these values, concentration dependence of membrane potential was predicted. Membrane electrical resistance, Rm, was obtained from Impedance Spectroscopy (IS) measurements using equivalent circuits as models, and the membrane porosity Θ=(0.11±0.02)% was also obtained from resistance values, which agrees very well with the value determined from geometrical parameters. From Rm, Δφm and Θ values, the diffusion coefficient of the ions in the membrane pores can be calculated, and the following average values were obtained: DK+=(1.9±0.4)×10−9 m2/s and DCl=(0.8±0.2)×10−9 m2/s, but for an average concentration higher than 0.06 M, their values do not differ practically from solution in agreement with the small negative charge previously indicated.  相似文献   

5.
Conductance measurements are reported for several salts in binary aqueous mixtures containing up to 60 mole % sulfolane, 20 mole % acetonitrile and 20 mole % dimethylsulfoxide. The variations of R = (λ±0η0)s/(λ±0η0)w with solvent composition have been compared with those observed in other water-rich mixtures. Alkali cations show R values greater than one with maxima in all the solvent mixtures. This behaviour has been discussed in terms of “sorting”, “averaging” and “steric” effects. Contrary to what happens to alkali cations, halide ions show R values greater or lesser than one according to whether the organic solvent respectively increases or decreases water structure. On these bases we suggest that conductometric behaviour of the halide ions may be indicative of the effect of the cosolvent on the water structure in water-rich mixtures and that DMSO is a water structure breaker.  相似文献   

6.
Forward recoil spectrometry (FRES) was used to measure the tracer diffusion coefficients D*PS and D*PXE of deuterated polystyrene (d-PS) and deuterated poly(xylenyl ether) (d-PXE) chains in high molecular weight protonated blends of these polymers. The D*s were shown to be independent of matrix molecular weights and to decrease as M−2, where M is the tracer molecular weight, suggesting that the tracer diffusion of both species occurs by reptation. These D*s were used to determine the monomeric friction coefficients ζ0,PS and ζ0,PXE of the individual PS and PXE macromolecules as a function of ф, the volume fraction of PS in the PS:PXE blend. Since ζ0,PSζ0,PXE at each ф, the rate at which a PS molecule reptates is much greater than that of a PXE molecule, even though both chains are diffusing in identical surroundings. Part of this difference may be due to the difficulty of backbone bond rotation of the PXE molecule. However, even when measured at a constant temperature increment above the glass transition temperature, ζ0,PS and ζ0,PXE were observed to be markedly composition dependent. In addition the ratio ζ0,PS0,PXE varied from a maximum of 4 × 10−2 near ф=0.85 to a minimum of 5 × 10−5 for ф=0.0. These results show that intramolecular barriers do not solely determine the ζ0s of the components in this blend. Clearly, the interactions between the diffusing chains and the matrix chains also influence ζ0.  相似文献   

7.
Model networks of ,ω-dihydroxy-poly(dimethylsiloxane) (PDMS) were prepared by tetrafunctional crosslinking agent tetraethyl orthosilicate (TEOS) and the catalyst stannous 2-ethylhexanoate. Hydroxylterminated chains of PDMS having molecular weights 15 × 103 and 75 × 103 g mol−1 were used in the crosslinking reaction. Bimodal networks were obtained from a 50% (w/w) mixture of PDMS chains with Mn = 15 × 103 and 75 × 103 g mol−1. A sequential interpenetrating network of these PDMS chains was also prepared. Physical properties of the elastomers were determined by stress-strain tests, swelling and extraction experiments, and differential scanning calorimetry measurements.  相似文献   

8.
The theoretical basis and the results of a computer simulation are presented which describe the operational limits of size and concentration for aerosol sizing by laser Doppler spectroscopy LDS,. This analysis suggests that a state of the art LDS system has the capability of sizing 0·03 μm diameter particles when the number concentration is 108 cm−3 or greater and 0·2 μm diameter for coocentrations as low as 100 particles cm−3. An evaluation of the effect on the laser Doppler spectroscopy measurements of a polydisperse aerosol having a log normal size distribution is presented and methods for combining these measurements with other averaged measurements to determine both count median diameter (CMD) and geometric standard deviation (δg) are proposed. For aerosols having log normal distributions with 0·3 < CMD < 3 μm and 1·0 < δg < 2·0, laser Doppler spectroscopy is able to measure the surface area median diameter within ± 15 per cent, independent of polydispersity. Applications of LDS to aerosol sizing are evaluated and its advantages and disadvantages relative to other sizing methods are discussed.  相似文献   

9.
The reaction rate of the alite reaction, C + C2S → C2S, was measured by means of a sandwich technique, for experimental conditions pertinent to burnability studies. The growth of the alite layer followed the parabolic rate law, X2 = k·t, where X is width of layer and t reaction time. k (cm2 sec−1) was found to increase with increasing melt content in the structure, with increasing temperature, and with increasing substitution of Fe2O3 for Al2O3 in the melt. Substitution of MgO for Al2O3 up to MgO-saturation had no significant effect on k.  相似文献   

10.
The effect of SO2 addition on the oxidation of ethyl acetate, ethanol, propane and propene, over Pt/γ-Al2O3 and Pt/SiO2 has been investigated. The reactants (300–800 vol. ppm) were mixed with air and led through the catalyst bed. The conversions below and above light-off were recorded both in the absence and in the presence of 1–100 vol. ppm SO2. For the alumina-supported catalyst, the conversion of ethyl acetate, ethanol and propane was promoted by the addition of SO2, while the conversion of propene was inhibited. The effect of SO2 was reversible, i.e. the conversion of the reactants returned towards the initial values when SO2 was turned off. However, this recovery was quite slow. The oxidation of propane was inhibited by water, both in absence and presence of SO2. For the silica-supported catalyst no significant effect of SO2 could be observed on the conversion of ethyl acetate, ethanol or propane, whereas the conversion of propene was inhibited by the presence of SO2. In situ FTIR measurements revealed the presence of surface sulphates on the Pt/γ-Al2O3 catalyst with and after SO2 addition. It is proposed that these sulphate groups enhance the oxidation of propane, ethyl acetate and ethanol by creating additional reaction pathways to Pt on the surface of the Pt/γ-Al2O3 catalyst.  相似文献   

11.
Ahmed Harrach  Andr  M  trot 《Electrochimica acta》1989,34(12):1877-1881
The electrochemical intercalation of H2SO4 into graphite leads to the well known graphite salts C+n HSO4. χH2SO4 (with n → 20), χ ≈ 2.5). The potential—charge relations observed during intercalation agree well with a model involving the extended graphene layer-intercalate interfacial capacitance. The components of this capacitance are studied, including the effect of Fermi level lowering. A computed electrocapillary curve is compared to the separation work of graphene layers.  相似文献   

12.
Edmund A. Di Marzio 《Polymer》1990,31(12):2294-2298
The entropy theory of glasses is used to derive the glass temperature, Tg, of a binary polymer blend in terms of the glass temperatures of the two substituents. The formula is Tg = B1Tg1 + B2Tg2, where Bi is the fraction of flexible bonds of substituent i. A bond is flexible if rotation about it changes the shape of the molecule. Bonds in side groups as well as in the backbone are to be counted. This formula assumes that the free volume, taken here to be the volume fraction of empty lattice sites, is the same for each of the three materials. It has no parameters. The above equation expressed in weight fractions, Wi, is (TgTg1)W111) + (TgTg2)W222) = 0, where ωi is the weight of a monomer unit and ggi is the number of flexible bonds per monomer unit. A more general treatment is given. One variation of the more general treatment which expresses the properties of the blend in purely additive terms gives Tg = B1Tg1 + B2Tg2 + KB1B2(Tg1Tg2)(V01V02), where V0i are the free volume fractions of the homopolymers at their glass temperatures and K is a constant. The added term is usually small. The most general form of the equation requires the energy of interaction between the two unlike molecules, which can be estimated by volume measurements on the blend.  相似文献   

13.
Stoichiometric polycrystalline samples of MnxMg1−xFe2O4 (0·5 ≤ x ≤ 0·66) have been synthesized by following a novel route using stabilized MnO and Fe2O3 at high temperatures. This route precludes the formation of large amounts of Mn3+ and Fe2+ and precipitation of MgO and -Fe2O3 which are generally observed during the usual route of preparation by conventional ceramic techniques. These samples have been characterized for their structural and magnetic properties using X-ray diffraction, Fe57 Mössbauer spectroscopy and bulk magnetic properties such as initial permeability, loss factor, ferromagnetic transition temperature, remanance and coercivity. For X = 0·62, these ferrites exhibit the highest remanance ratio 0·96, suitable for square loop applications.  相似文献   

14.
The swelling features of gelatine gels in water (good solvent) were studied as a function of thermodynamic conditions of sol—gel transition and ripening. It is shown that the degree of equilibrium swelling Qe varies with the volume fraction of the polymer in a casting solution φo in accordance with the prediction of the classic theory: Qe φo−0.4. Qc, as a function of the gelation temperature Tg, the ripening time tr and φo, can be rescaled and described by the single empirical equation: Qe Tgx tryφo−0.4, where x = 0.1, y = 0.15 for wet gels and X = −0.5, y = 0.04 for dried gels. The kinetics of macroscopic swelling is described by the equation of Peters and Candau, with values of collective diffusion coefficients being in good agreement with values obtained by other workers via photon correlation spectroscopy.  相似文献   

15.
J. T. van Oosterhout  M. Gilbert   《Polymer》2003,44(26):8081-8094
Solid–gel transition temperatures, tm, were measured for PVC in binary and ternary mixtures of plasticizers. Data were used to investigate interactions between plasticizers, and their effect on tm and polymer–plasticizer interaction (χ) values. The UNIFAC-FV method was used to predict Gibbs free energy of mixing, and χ for a range of plasticizers, and interactions for one quaternary mixture. Synergy was observed when one or more branched phthalates were mixed with ODPP, but none when three phthalates were mixed. Mixtures of DOP with aliphatic ester plasticizers again produced synergy. Equations were produced to enable interaction coefficients of different plasticizer blends to be predicted. It was found that interaction coefficients increase as the difference in molar mass between the two plasticizers increases. Calculated χ values were generally similar to experimental values. The lower limit of Gibbs free energy of mixing required for plasticization of PVC using monomeric plasticizers was calculated to be about −0.9 J/g, and that for polymeric plasticizers −0.15 J/g. Equations were produced enabling Gibbs free energy of mixing to be predicted from the amount of C atoms in the plasticizer.  相似文献   

16.
To measure size distributions of submicrometer aerosols with an electrical differential mobility analyzer (DMA) requires an inversion procedure. The Knutson (1976) and the Hoppel (1978) inversion procedures were numerically investigated for the case of log-normal aerosol size distributions. It was found that the Hoppel procedure converges to the same result as that given by the Knutson procedure. The computational range for geometric mean diameter ( g) was 0.025-0.25 μm, and for geometric standard deviation (σg) was 1.1–2.4. The inversion error was found to be greater than 10% in certain “forbidden zones” of g and σg values. For the case of an ideal DMA having no lower mobility limit, only one forbidden zone exists, this consisting of small σg values. The boundary of this forbidden zone intercepts the computational range boundaries at σg = 1.25, and σg = 1.62, . These results also apply to an actual DMA when the size distribution of particles larger than the DMA singly charged mobility limit is available a priori. If such information is not available, the concentration of these larger particles is assumed to be zero in performing the inversion. This assumption adds a second forbidden zone, consisting of large σg values and having the intercepts σg = 2.44, and σg = 1.50, . The first forbidden zone remains nearly the same.  相似文献   

17.
Blends of poly(ethylene oxide) (PEO) and poly(4-vinylphenol-co-2-hydroxyethyl methacrylate) (PVPh-HEM) were studied by means of synchrotron small and wide angle X-ray scattering (SAXS and WAXS, respectively) and by differential scanning calorimetry (DSC). The DSC measurements were used in the determination of the Flory–Huggins interaction parameter and also to study the isothermal and non-isothermal crystallisation kinetics of the PEO/PVPh-HEM blend. The interaction parameter, χ12, was found to be negative (between −0.5 and −2.5, approximately) and presented a significant dependence on the blend composition, which is expected for a system with specific interactions such as hydrogen bonding. From the kinetic studies with Kissinger, Friedman and Avrami models, it was shown that crystallisation of PEO chains is slower in the blend than in the pure polymer, despite the decrease in the energy barrier to the crystallisation with the increase in PVPh-HEM concentration.

From the SAXS and WAXS profiles, the nanostructure of the blend was elucidated, exhibiting the formation of PEO lamellae even in the blends containing high concentrations of PVPh-HEM, which are non-crystalline (as observed by the WAXS profiles). The thickness of the PEO lamellae (Rc, approximately 8 nm) remains almost unchanged with the blend composition, while the crystalline peaks, observed at 19.78 and 23.98°, vanish, and the WAXS profile exhibits only a non-crystalline halo. For the non-crystalline blends with high concentrations of PVPh-HEM, PEO chains keep their crystalline structural memory.  相似文献   


18.
NiMo/(X)SiO2–Al2O3 catalysts were synthesized with various SiO2 contents (X = 0, 10, 25 and 50 wt%) using the pH-swing method. In order to find the optimum SiO2 content, the catalysts were evaluated in the hydrodesulfurization of 4,6-DMDBT, hydrogenation of naphthalene and hydrodenitrogenation of carbazole. Kinetic parameters of Langmuir–Hinshelwood type equations for all the reaction systems were estimated. FTIR analysis of CO adsorption for the sulfided catalysts shows that the amount of coordinatively unsaturated Mo sites promoted by nickel (CUS-NiMoS) follows the order NiMo/10ASA > NiMo/25ASA > NiMo/0ASA. This tendency agrees with the results obtained in catalytic activity.  相似文献   

19.
A new approach of analyzing the depth profile of double bond conversion as a function of film depth has been studied. By using a combination of statistical calculation and traditional FTIR, a new approach to analyze the depth profile of conversion “layer by layer” in the characterization of photopolymerization was explored. Utilizing a formula (X1 + X2 +  + Xn)/n = average conversion, n = 1, 2, 3, n is a number of layers (μm), an average conversion of any 5 μm depth could be calculated from the prior 5 μm conversion and the total average conversion. More detailed information of photopolymerization, such as the depth profile of conversion and a difference in conversion between the top 5 μm and the bottom 5 μm in a 25 μm film as a function of film depth, was obtained. This investigation was accomplished using a variation of film depth, non-photo bleaching photo initiator [PhI] as well as the concentration of PhIs in the presence of air and in the absence of air. Results of analyzing double bond conversion between traditional FTIR and the new approach (statistical calculation/FTIR) were compared.  相似文献   

20.
Ammonium polyacrylate (NH4PA) was introduced into powdered mixtures consisting of anatase-structured TiO2 nanoparticles and silicon alkoxide precursors at the sol level, and the rheological behavior of the mixtures was examined under various solid loadings (φ=0.05–0.13 in volumetric ratios), shear rates (  s−1) and NH4PA concentrations. The alkoxide precursors were mixtures of tetraethyl orthosilicate (TEOS, Si(OC2H5)4), ethyl alcohol (C2H5OH), H2O and HCl in a constant [H2O]/[TEOS] ratio of 11. The nanoparticle–sol mixtures generally exhibited a pseudoplastic flow behavior over the shear-rate regime examined. The NH4PA appeared to serve as an effective surfactant which facilitates the suspension flow by reducing the flow resistance at low NH4PA concentrations. At φ=0.10, a viscosity reduction ca. 85% was found at  s−1 when the NH4PA concentration was held at 2.5 wt.% of the solids. As the NH4PA exceeded a critical level, e.g., [NH4PA]≥3.0 wt.%, the NH4PA acted as a catalyst which quickly turned the TiO2–silica sol mixtures (φ=0.10) into a gelled structure, resulted in a pronounced increase of mixture viscosity. The maximum solids concentration (φm) of the mixtures was experimentally determined from a derivative of relative viscosity, i.e., (1−ηr−1/2)–φ dependence. The estimated φm increased from 0.127 to 0.165 when NH4PA of 0.5 wt.% was introduced into the TiO2–silica sol mixtures.  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