首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Klaus J. Hüttinger  P. Schleicher 《Fuel》1981,60(11):1005-1012
The catalysis of hydrogasification of carbon by Fe, Co and Ni was studied using a special petroleum coke with extremely low reactivity. The kinetics were studied with impregnated coke in a fixed-bed flow reactor between 1133 and 1235 K and up to 2 MPa, yielding the following rate equation: ?rH2 = k(CH2?CH2e)(1 + Kads · CH2)2Apparent activation energies and heats of adsorption are: Fe, 152 and ?92 kJ mol?1; Co, 201 and ?82 kJ mol?1 Ni, 165 and ?50 kJ mol?1. These studies with impregnated coke as well as further gasification experiments with cokes heat-treated after impregnation with metal salts up to 2273 K confirmed a spillover mechanism and excluded any influence of electronic interactions between carbon and the catalyst metals.  相似文献   

2.
For solutions of polystyrene (M=105–106 g mol?1), intrinsic viscosities [η] have been measured at 34.5°C, which is the θ temperature for the polymer in cyclohexane. The solvents comprised cyclohexane in admixture with a thermodynamically good solvent, 1,2,3,4-tetrahydronaphthalene (tetralin, TET) over the whole range of solvent composition. From an assessment of several extrapolation procedures, a value of 85 × 10?3(±1 × 10?3) cm3g?32mol12 was obtained for Kθ (in the relationship [η] = KθM12α3, where α is the expansion factor), thus yielding 0.681 A? g?12mol12, 2.25 and 10.2 for the unperturbed dimensions, steric factor σ and characteristic ratio C respectively. The value of Kθ was independent of solvent composition despite the finite excess free energy of mixing for the solvent components alone, which has been asserted elsewhere to affect Kθ. The present results, in conjunction with previous ones relating to 98.4°C, indicate a value of ?0.89 × 10?3 deg?1 for the temperature coefficient of the unperturbed dimensions.  相似文献   

3.
J.E.L. Roovers 《Polymer》1975,16(11):827-832
A new method for the synthesis of comb shaped polystyrenes of predetermined structure is described. Silicon-chlorine bonds are introduced into the backbone polystyrene by reaction of SiMe2Cl2 with hydrolysed styrene/vinyl acetate copolymers and coupled with polystyryl-lithium in benzene. From a common backbone polymer a series of comb polymers are prepared that have a constant number of branches but vary in branch length. The MwMn of the whole comb polymers is about 1.3. The comb polymers with high branch density show θ (A2) temperatures lower than that for linear polystyrene. The radius of gyration at θ (A2) [〈S2θ (A2)] is always larger than calculated from random flight statistics. For comb polymers with 20–30 branches 〈S2θ (A2)〈S20,bb increases with λ?0.46 where λ is the fraction of polymer in the backbone. The intrinsic viscosities of the comb polystyrenes at θ (A2) are equal to that of the parent backbone polymer when λ > 0.25 and increase only little when λ becomes equal to 0.1. Similar behaviour is found in toluene. Intrinsic viscosities in cyclohexane at 35°C show a complex pattern because of the θ-temperature variation.  相似文献   

4.
A flow microcalorimeter designed to measure the heat of mixing of dilute polymer solutions is described. The instrument is sensitive to steady state heating rates of ~10 μJ/sec. Measurements of heats of mixing of solutions of differing concentrations of n-hexane and cyclohexane are reported and are compared with recommended data of McGlashan and Stoeckli. Values of:
K1=limV2→ 0
(H?1 ? H?01RTv22 are obtained for four polymer—solvent systems: polyisobutylene—benzene, 0.22; polystyrene (PS)—cyclohexane, 0.33; PS—n-butyl acetate, ?0.06 all at 25°C; and PS—toluene, ?0.05 at 40°C. Various theoretical calculations of second virial coefficients A2 made with use of the calorimetric data are compared with previously measured A2 for the first two mixtures.  相似文献   

5.
J.C. Ravey  P. Mazeron 《Polymer》1975,16(5):329-337
The studies of the changes ΔHH of the HH component of the light scattered by dilute suspensions of particles oriented by an electric field E0 are shown to be convenient and reliable methods for obtaining simultaneous information on the electrical (μ), geometrical (ω) and optical (δ) anisotropies of the particles. Indeed, when E0 is applied in three particular directions, which are related to the exterior and interior bisectrices of the observation angle θ, ΔHH(θ = 90°) generally has the same sign as (?μω), (+μω) or (?μ), according to the direction of E0. Moreover, under the same conditions, the slopes of ΔHH(θ) versus θ are shown to be much more informative than those of ΔHH(θ)HH(θ), having the sign of (?μδ), (+μδ) or (?μδω). Experiments on TMV particles illustrate in a very satisfactory manner these theoretical conclusions.  相似文献   

6.
The apparent diffusion coefficients for Ti, V, Cr, Nb, Mo and Hf as carbides and for elementary Fe, Ni and Cu in electro graphite have been determined by means of an electron-microprobe analyzer. These pseudo diffusion coefficients were found to vary with the heat treatment time. However, after one hour these remain constant and follow the Arrhenius type of relation D = D0exp(?Q/RT). The activation energy Q was nearly constant for the metals investigated. An attempt was made to correlate the frequency factor D0 with the heat of formation ΔH?298 of the corresponding carbides. A plot of log D0vsΔHf yielded two straight lines, one for the negative ΔH?, the other for positive ΔH?. This method was satisfactorily applied to predict the diffusion coefficients of Zr, Sb and Bi.  相似文献   

7.
E. Straube 《Polymer》1985,26(1):105-108
A polymer chain consisting of Nr segments with a repulsive interaction (binary cluster integral βr) and Na ? Nr segments with a stronger, attractive and pairwise saturable interaction (βa), which is at the averaged θ-point N2rβr + N2aβa = 0 deviations from the predictions of the two parameter theory: α2R ? 1 ~ δzr < 0 and A2δzr > 0 with δzr ~ βr(NaNr)12. It is shown that the deviations from the universal behaviour are due to the existence of an intermediate length scale NaNr.  相似文献   

8.
9.
Hydrotreatment of spent oil distillate was carried out on a commercial Ni-Mo-alumina catalyst in the temperature range 260–340 °C, with a liquid hourly space velocity (LHSV) of 0.7–2.0 h?1, pressure of 4.5 MPa and H2oil ratio of 300 NL L?1 (normal litre of H2 per litre of feedstock). U.v. spectra of hydrogenated and original spent oil distillates (measured in normal hexane) gave a band with a maximum at 230 nm. The change in absorbance at three selected wavelengths for original oil distillate and hydrotreated oil at different operating conditions was taken as a guide for the determination of hydrogenation reaction rates (including partial saturation of aromatics and sulphur compound hydrogenolysis). The rate constants of hydrogenation reactions (k) using a second-order equation and a model of two parallel first-order reactions (k1 and k2) were calculated. Finally, the apparent activation energy (Ea), enthalpy of activation (ΔH1) and entropy (ΔS1) were calculated based on the values of k, k1, and k2. The calculated values of Ea based on k, k1 and k2 were 81.479, 71.188 and 62.882 kJ mol?1, respectively. The values of ΔH1 based on the same rate constants were 76.670, 66.564 and 58.433 kJ mol?1, while the values of ΔS1 were ?117.150, ?133.779 and ?150.823 J mol?1 K?1, respectively.  相似文献   

10.
J.C. Radon  L.E. Culver 《Polymer》1975,16(7):539-544
The effects of frequency and temperature on fatigue crack propagation rate in poly(methyl methacrylate) and polycarbonate have been studied using centrally notched plate specimens cycled in tension between constant stress intensity limits. Crack growth was monitored at frequencies between 0.1 Hz and 100 Hz and at temperatures between ?60°C and 40°C. A linear relationship between the cyclic crack growth rate d(2a)dN and appropriate levels of toughness, K, has been proposed: d(2a)dN = A?α, where ? = (λ ? λth)(K21C ? K2max), λ = K2max ? K2min, λth is the threshold limit and A and α are constants. Also, the influence of mean stress intensity was briefly discussed.  相似文献   

11.
Simple techniques for studies of X-ray diffraction from PdSiO2 catalysts during catalytic reactions are reported here. The catalytic activity of PdSiO2 of low percentage metal exposed for methylcyclopropane hydrogenolysis at 0 °C is less for catalyst cooled from 450 °C in H2 than for that cooled in He. This effect appears to result from differing amounts of hydride formation. Ease of formation of hydride decreases with decreasing Pd particle size. Exposing the catalysts with Dh = 13.8 and 29.3% to hydrogen at ~25 °C results in nearly complete conversion of particles of Pd to hydride. Purging with helium (even at 0 °C) reconverts the hydride to palladium, but this reconversion exhibits an induction period; hydride formation is more rapid. Passing hydrogen plus methylcyclopropane results in the conversion into hydride of a substantial fraction of catalyst particles originally present as palladium. The palladium hydride particles of catalyst produced by cooling from 450 °C in H2 remain as PdH0.7 during the hydrogenolysis reaction. The lattice parameter of palladium particles is indistinguishable from that of bulk palladium, at least for particle diameters ≥45 Å.  相似文献   

12.
Shaul M. Aharoni 《Polymer》1980,21(12):1413-1422
When plotted against concentration V02, the viscosity η0 curve of isotropic solutions of lyotropic nematic mesomorphic polymers increases according to:
η00[[η]V02+(π4)[η]2(V02)2+K2(lnx)2[η]3(V02)3+…]
in which η0 is the solvent viscosity, K2 a numerical constant, x? is the average molecular axial ratio and [η] the intrinsic viscosity as defined by:
[η]=2x2451ln 2x?1.84+3ln 2x?0.61 (+1415)
At the concentration v12 an anisotropic phase appears and the viscosity curve shows a decrease in slope followed by a change in direction at a peak viscosity ηp at vp2. Upon further increase in v02 the system undergoes a phase inversion and finally turns fully anisotropic at vA2. In the biphasic interval the viscosity is described by:
η0mat 1+(5λ+2)2(λ+1)Vinc
where λ = ηincηmat, the ratio of the inclusions viscosity to the matrix viscosity, and vinc the volume fraction of the inclusions in the system. It must be emphasized that the molecular weight of the polymer in both phases changes continuously with concentration, resulting in commensurate changes in ηince, ηmat and their ratio λ. In the anisotropic region the viscosity first decreases moderately and then increases precipitously with v02, according to:
η00[2x290S(5ln 2x?1.8+6ln 2x?0.61)+2]V021-V25°VE
in which S is an order parameter and V25, VE are volumes swept by the orbits of the flowing rodlike macromolecules. The equations give results in good qualitative and fair quantitative agreement with experimental data in the literature.  相似文献   

13.
14.
15.
Laser light scattering including angular dependence of total integrated scattered intensity and of the spectral distribution has been used to characterize five samples of poly(1,4-phenylene terephthalamide), PPTA (commercially known as Kevlar), of different molecular weights in 96% sulphuric acid and 0.1 NK2SO4. The data are supplemented by intrinsic viscosity measurements used to detect the possible effects of association, by differential refractometry providing a measure of the refractive index increments in mixed solvents (H2O, H2SO4 and K2SO4) and by spectrophotometry for the extinction coefficient needed in the correction of attenuation in light scattering studies. The results show 〈DZ = 2.11 × 10?5M?W?0.75cm?2s?1 in reasonable agreement with an average of many of the published intrinsic viscosity data obeying [η] = 1.09 × 10?3 Mw1.25 ml g?1 and w expressed in g mol?1.  相似文献   

16.
Francis T.C. Ting  Hoom B. Lo 《Fuel》1978,57(11):717-721
Maximum reflectance, Rmax, can be calculated by the following equation: Rmax = (R1 + R32) + ((R1 ? R2)2 + (R2 ? R3)22)12 in which R1, R2, and R3 are three separate reflectance readings on the same vitrinite grain at 45° angular intervals. The equation is derived from the reflectance distribution function for a central section of a reflectance indicatrix: Rα = Rmaxcos2α + Rminsin2α where Rα is the reflectance measured in the direction α degrees from Rmax, Rmin is an apparent minimum which has the minimum value on this central section, and Rmax ? Rmin ? Rmin. This new technique can be further simplified by rotating the polarizer in 45-degree intervals to obtain three photometer readings and then converting them to R1, R2, and R3, by the use of empirically-established conversion factors. The new techniques do not require the full revolution of the microscope stage and, therefore, are particularly suitable for measuring small grains of vitrinite.  相似文献   

17.
Under COH2O systems at initial pH values s> 12.6, an Illinois No. 6 coal, PSOC-26, was converted to a fully pyridine-soluble product, with benzene and hexane solubilities of 50% and 18%, respectively. The product gases were H2 and CO2. However, the expected H2CO2 ratio of 1.0 based on the water gas shift reaction was not observed, but the deficit in hydrogen was found in the increased hydrogen content of the coal product. 95% coal carbon recovery and good hydrogen balances were obtained, and the coal products were found to be very similar to those from conventional tetralin systems. The results suggest an efficient base-catalysed process, and that COH2O systems are useful for coal studies.  相似文献   

18.
This article reports the synthesis and free radical polymerization of ortho-vinylbenzophenone. The glass transition temperature Tg of the homopolymer is 136°C. The products synthesized appeared to be atactic and amorphous. The Mark-Houwink constants for poly (o-vinylbenzophenone) in tetrahydrofuran are K = 4.2 × 10?2 cm3 g?1 and a = 0.765. The pre-exponential constant under theta conditions, Kθ, is estimated to be 5.93 × 10?2 cm3 g?1. The ratio of unperturbed dimensions of the actual polymer and free rotating analogue chain is 3.93, which is almost double that of polystyrene. The Flory-Huggins interaction parameter for poly (o-vinylbenzophenone)tetrahydrofuran is 0.48 at room temperature. The kpk12t ratio at 60°C is 1.1 × 10?2l12mol?12s?12. In free radical copolymerizations with styrene at 70°C, r1 (o-vinylbenzophenone) = 1.216, r2 = 0.751. This copolymerizations is virtually random.  相似文献   

19.
Thirteen fractions of poly(phenyl acrylate) have been prepared with weight-average molecular weight ranging from 0.158 × 106 to 2.57 × 106 g mol?1. The temperature coefficient of the unperturbed dimensions and the glass transition temperature were found to be ?1.8 × 10?3 deg?1 and 55.6°C respectively. Good accord was obtained among different methods for establishing θ-conditions of 11.5°C in ethyl lactate. From viscometry, osmometry and light scattering under θ-conditions, as well as in a good solvent, the unperturbed dimensions were determined via several procedures yielding a value of [〈r20wM?w]12 = 6.0 (±0.2) × 10?9cm g?12mol12. This corresponds to a steric factor υ = 2.37 (±0.08) and a characteristic ratio C = 11.3 (±0.8). The polymer chain is thus more rigid than poly(methyl acrylate), but less rigid than poly(phenyl methacrylate). With respect to its Tg and flexibility, poly(phenyl acrylate) bears a strong similarity to poly(benzyl methacrylate).  相似文献   

20.
Valery P. Privalko 《Polymer》1978,19(9):1019-1025
Analysis of spherulitic growth rate data for a number of linear polymers has shown that the temperature at maximum growth rate, T1, is related to the glass transition temperature, Tg, through the empirical equation, T1 = 1.26 Tg. The universal master curve for the temperature dependence of the growth rate of crystals from the melt in reduced Gandica—Magill coordinates, ln(GG1) = f(T ? T)(Tm ? T), is possible only on the condition that the following empirical equation holds: 0.26 = TTg ? TTm. Finally, limits of variation of the ‘conformational’ contribution to the excess entropy, and of the free volume fraction at T1 were evaluated for some polymers.  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