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.
Yozo Chatani  Tamotsu Irie 《Polymer》1988,29(12):2126-2129
Linear poly(ethylenimine) was found to form a crystalline complex with hydrogen chloride when poly(ethylenimine) in the state of its anhydrate or hydrates was immersed in concentrated hydrochloric acid or exposed to HCl vapour. The crystal structure of the HCl complex was determined by X-ray diffraction. The crystals are metrically tetragonal with cell constants a = B = 5.06 Å and c (chain axis) = 7.57 Å and the unit cell includes one planar zigzag polymer chain (two monomeric units) and two HCl molecules (the molar ratio is 1:1). The space group of the crystals is, however, not of the tetragonal system: the structure determined can be expressed in terms of any of the following crystal systems and space groups: orthorhombic P2221, orthorhombic Pcm21, orthorhombic Pc2m, and monoclinic P21/m (c unique). The Cl···N distance of 3.05 Å strongly indicates that every NH and HCl hydrogen atom participates in hydrogen bondings between nitrogen and chlorine atoms. The HCl complex has a fairly high melting temperature of 265°C, which is about 200°C higher than those of the anhydrate and hydrates.  相似文献   

2.
Mass transfer coefficients at cylindrical, H2 evolving electrodes, were measured by determining the reduction rate of K3Fe(CN)6. The variables studied were: gas discharge rate V, diameter of the cylinder D, height and position of the cylinder. The diameters ranged from 0·2–2·5 cm, the cd from 25-380 mA/cm2. For horizontal cylinders, the following correlation was found: log K = a + 2·17 log(V0·11/D0·08). The application of gas evolving cylindrical electrodes in industrial electrolysis is discussed in comparison with rotating electrodes.  相似文献   

3.
Light scattering by monodisperse solutions of rigid rod-like anisotropic macromolecules, with linear dimensions l of the order of incident wavelength λ, oriented in an external d.c. electric field, →E has been analysed. The relative variations δVEv, δHEv, δVEh, δHEh of the scattered light components are discussed for the three values [l/λ] = 1, 0·5, 2, and various reorientation parameters, p = [μE/kT] of the permanent dipole moment μ and q = [(31)E2/2kT] of the moment induced by the principal polarizabilities 1 = 23. The saturation orientation field strength has been calculated for certain macromolecules with the aim of determining their optical anisotropy numerically.  相似文献   

4.
新型除磷填料的制备及除磷吸附床运行参数的优化   总被引:1,自引:0,他引:1       下载免费PDF全文
为解决现有除磷吸附剂粒径小造成的材料易流失和系统压降过大等问题,以实现吸附除磷工艺在实际工程中的应用,以聚氨酯填料为载体,水溶性聚氨酯为介质,将水化硅酸钙负载到聚氨酯填料上制成负载型除磷填料。研究了制备条件对除磷填料除磷效果的影响,采用扫描电子显微镜(SEM)和傅里叶变换红外光谱仪(FTIR)观察分析了负载前后水化硅酸钙微观结构及化学基团的变化;利用除磷填料作为除磷吸附床的滤料,研究了运行条件对吸附床除磷效果的影响。在此基础上,利用响应曲面法研究了除磷吸附床磷酸盐去除率和各变量之间的关系,并对工艺参数进行了优化。结果表明,水性聚氨酯溶液的浓度和用量分别为100 g/L和50 ml,水化硅酸钙的质量为12 g的条件下所制备的除磷填料除磷效果最好;SEM和FTIR分析结果显示,水化硅酸钙负载前后其孔隙结构和化学基团没有明显的变化;预测模型的方差分析结果表明,HRT(X 1)、进水ρ(PO4 3--P)(X 2)、温度(X 3)、初始pH(X 4)以及X 1 X 2,X 1 X 4,X 2 X 3,X 2 X 4的交互作用均对磷酸盐的去除具有显著影响 (P<0.05),但X 1 X 3的交互作用对磷酸盐的去除影响不显著。通过预测模型获得的最佳运行条件为:HRT为79.77 min,进水ρ(PO4 3--P)为1.70 mg/L, 温度为34.04℃,pH为9.68。在该条件下,反应器对磷酸盐的去除率可以达到93.46%。  相似文献   

5.
The oxidative coupling of methane (OCM) over a La2O3/CaO catalyst was studied in a poly tropic fixed-bed reactor (I.D. = 15 mm, W/F= 0.15 g · s/ml). Reaction conditions for stable operation were determined. (1) A minimum inlet temperature of 580°C was necessary to initiate the reaction. (2) The maximum hot-spot temperature of 1000°C limited the highest oxygen inlet concentration to 20%. The temperature gradients in the bed amounted to 250 K. The influence of the reaction conditions on the C2+ selectivity was investigated by testing the effects of temperature (Tinlet = 580–860°C), oxygen concentration (CO2 = 5–20%) and particle diameter (dp = 250–350 μm, and pellets of hp = 4 mm and dp = 4 mm). The C2+ selectivity ran through a maximum with increasing temperature and decreased with rising inlet oxygen concentration. Mass-transfer limitations, which occurred when applying pellets, resulted in a drop of C2+ selectivity. Highest C2+ yields amounted to 15.5% (XCH4 = 31%, S 2+ = 51%). Distributed feed of oxygen was tested as a means to cope with the high temperature gradients and to increase C2+. selectivity. Upon applying this mode of operation, oxygen concentrations up to 30% could be converted. However, no improvement of C2+ selectivity and yield compared to cofeed operation was achieved.  相似文献   

6.
Conductance measurements are reported for s-alkylisothiouronium bromide, iodide and picrate salts in the dipolar aprotic solvent DMSO at 25°C. The data were analysed by Fuoss's equation (1975–1980) for 1:1 electrolytes, from which the values of Λ0, the Gurney's cosphere diameter R and KA are obtained. The results are discussed in the light of the recorded values of the constants KR and KS and the free energy term Gs.  相似文献   

7.
A series of poly(ethylene oxide)-substituted triphenylphosphines, Ph3−mP[C6H4-p-(OCH2CH2)nOH]m (PEO-TPPs; 1a m=1, 1b m=2, 1c m=3; N=m×n=8–25), have been prepared by the ethoxylation of mono-, di-, and tri-p-hydroxytriphenylphosphines. PEO-TPPs demonstrate an inverse temperature-dependent solubility in water, and possess distinct cloud points range from 26°C to 90°C.

Based on the clouding property of PEO-TPPs, a new line of aqueous/organic two-phase catalysis termed the thermoregulated phase-transfer catalysis (TRPTC) has been described. That is, the catalyst transfers into the organic phase to catalyze a reaction at a higher temperature, and returns to the aqueous phase to be separated from the products at a lower temperature. Application of this novel strategy to the rhodium-catalyzed two-phase hydroformylation of higher olefins gave desirable results with an average turnover frequency of 180 h−1 for 1-dodecene. The TRPTC is suitable for carrying out a reaction with extremely water-immiscible substrate in the aqueous/organic two-phase system. Thus, the application scope of the classical two-phase catalysis has been widened.  相似文献   


8.
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   

9.
The synthesis of a novel 3D aluminophosphate is described. The thermal properties of the material were investigated, and the existence of three high-temperature variants was revealed. The crystal structures of the as-synthesized material (UiO-26-as) and the material existing around 250°C (UiO-26-250) were solved from powder X-ray diffraction data. UiO-26-as with the composition [Al4O(PO4)4(H2O)]2−[NH3(CH2)3NH3]2+ crystallizes in the monoclinic space group P21/c (no. 14) with a=19.1912(5), b=9.3470(2), c=9.6375(2) Å and β=92.709(2)°. It exhibits a 3D open framework consisting of connections by PO4 tetrahedra with AlO4 tetrahedra, AlO5 trigonal bipyramids and AlO5(H2O) octahedra forming two types of layers stacked along [1 0 0] and connected by Al–O–P bondings. The structure possesses a 1D 10-ring channel system running along [0 0 1], in which doubly protonated 1,3-diaminopropane molecules are located. UiO-26-250 with the composition [Al4O(PO4)4]2−[NH3(CH2)3NH3]2+ crystallizes in the monoclinic space group P21/c with a=19.2491(4), b=9.27497(20), c=9.70189(20) Å and β=93.7929(17)°. The transformation to UiO-26-250 involves removal of the water molecule which originally is coordinated to aluminum. The rest of the structure remains virtually unchanged. The crystal structures of the two other variants existing around 400 (UiO-26-400) and 600°C (UiO-26-600) remain unknown.  相似文献   

10.
The electrical conductivity of KI solutions in anhydrous acetonitrile has been determined at 0, 25 and 35°C in the concentration range 0·9– 600 × 10−4 mole/l. The values of Λ0, K and a calculated from the results are, respectively: 145·9 mho/cm, 0·95 × 10 −2 and 1·72 Å at 0°C; 186·2 mho/cm, 8·98 × 10−2 and 4·6 Å at 25°C; and 204·8 mho/cm, 5·17 × 10−2 and 3·5 Å at 35°C. The phoreograms at all the three temperatures are catabatic at lower concentration, but become anabatic at 0·017, 0·022 and 0·024 respectively, at 0, 25 and 35°C.  相似文献   

11.
The catalytic effect of a heteropolyacid, H4SiW12O40, on nitrobenzene (20 and 30 μM) oxidation in supercritical water was investigated. A capillary flow-through reactor was operated at varying temperatures (T=400–500 °C; P=30.7 MPa) and H4SiW12O40 concentrations (3.5–34.8 μM) in an attempt to establish global power-law rate expressions for homogenous H4SiW12O40-catalyzed and uncatalyzed supercritical water oxidation. Oxidation pathways and reaction mechanisms were further examined via primary oxidation product identification and the addition of various hydroxyl radical scavengers (2-propanol, acetone, acetone-d6, bromide and iodide) to the reaction medium. Under our experimental conditions, nitrobenzene degradation rates were significantly enhanced in the presence of H4SiW12O40. The major differences in temperature dependence observed between catalyzed and uncatalyzed nitrobenzene oxidation kinetics strongly suggest that the reaction path of H4SiW12O40-catalyzed supercritical water oxidation (average activation Ea=218 kJ/mol; k=0.015–0.806 s−1 energy for T=440–500 °C; Ea=134 kJ/mol for the temperature range T=470–490 °C) apparently differs from that of uncatalyzed supercritical water oxidation (Ea=212 kJ/mol; k=0.37–6.6 μM s−1). Similar primary oxidation products (i.e. phenol and 2-, 3-, and 4-nitrophenol) were identified for both treatment systems. H4SiW12O40-catalyzed homogenous nitrobenzene oxidation kinetics was not sensitive to the presence of OH√ scavengers.  相似文献   

12.
Three metal-organic framework compounds [HZn3(OH)(BTC)2(2H2O) (DMF)] · H2O (MOF-CJ3), [Co6(BTC)2(HCOO)6(DMF)6] (MOF-CJ4), and [Co18(HCOO)36] · 3H2O (MOF-CJ5) have been solvothermally synthesized in mixed solvents of DMF and HAc, respectively. These MOFs are characterized by single-crystal X-ray diffraction, X-ray powder diffraction, ICP, TG analyses, IR, and photoluminescence spectroscopy analyses. MOF-CJ3 crystallizes in tetragonal, space group I4cm (No. 108) with a = 20.588(3) Å, b = 20.588(3) Å, c = 17.832(4) Å. Its framework can be described as a 3D decorated (3, 6)-connected net based on the assembly of trigonal prismatic SBUs and triangular links. MOF-CJ4 crystallizes in hexagonal, space group P-3 (No. 147) with a = 13.975(2) Å, b = 13.975(2) Å, c = 8.1650(16) Å. The 2D network of MOF-CJ4 is constructed from [Co6(R(CO2)3)2(HCO2)6(DMF)6] (R = C9H3–) clusters and 1,3,5-benzene-tricarboxylates linkers. MOF-CJ5 crystallizes in triclinic, space group (No. 2) with a = 15.205(3) Å, b = 18.005(4) Å, c = 21.500(4) Å,  = 71.21(3)°, β = 84.47(3)°, γ = 67.15(3)°. MOF-CJ5 has a diamond framework with Co-centered CoCo4 tetrahedra as nodes. It is noteworthy that the formic ligands in MOF-CJ4 and MOF-CJ5 are generated by the decomposition of DMF under acid conditions and incorporated into these two compounds.  相似文献   

13.
Preparations and physico-chemical characterizations of NASICON-type compounds in the system Li1+xAlxA2−xIV(PO4)3 (AIV=Ti or Ge) are described. Ceramics have been fabricated by sol-gel and co-grinding processes for use as ionosensitive membrane for Li+ selective electrodes. The structural and electrical characteristics of the pellets have been examined. Solid solutions are obtained with Al/Ti and Al/Ge substitutions in the range 0≤x≤0·6. A minimum of the rhombohedral c parameter appears for x about 0·1 for both solutions. The grain ionic conductivity has been characterized only in the case of Ge-based compounds. It is related to the carrier concentration and the structural properties of the NASICON covalent skeleton. The results confirm that the Ti-based framework is more calibrated to Li+ migration than the Ge-based one. A grain conductivity of 10−3 S cm−1 is obtained at 25°C in the case of Li1·3Al0·3Ti1·7(PO4)3. A total conductivity of about 6×10−5 S cm−1 is measured on sintered pellets because of grain boundary effects. The use of such ceramics in ISE devices has shown that the most confined unit cell (i.e. in Ge-based materials) is more appropriate for selectivity effect, although it is less conductive.©  相似文献   

14.
The synthesis and structure of (CH3CH[NH3]CH2NH3)1/2·ZnPO4, an organically templated zincophosphate (ZnPO) analogue of aluminosilicate zeolite thomsonite (THO), are described. The ZnPO framework is built up from an alternating, vertex-sharing, network of ZnO4 and PO4 groups (dav(Zn–O)=1.944 (8) Å, dav(P–O)=1.535 (9) Å, θav(Zn–O–P)=130.5°) involving distinctive 4=1 secondary building units. The 1,2-diammonium propane cations are highly disordered in the [0 0 1] 8-ring channels. Crystal data: (CH3CH[NH3]CH2NH3)1/2·ZnPO4, Mr=198.42, orthorhombic, space group Pncn (no. 52), a=14.119 (6) Å, b=14.136 (5) Å, c=12.985 (5) Å, V=2591 (3) Å3, Z=10, R(F)=0.057, Rw(F)=0.061 (for a twinned crystal).  相似文献   

15.
An attempt has been made to evaluate several experimental methods by which the potentials of zero charge (pzc) of metals could be concordantly measured. The methods chosen involved measurements of the adsorption in the double layer of a neutral organic compound as a function of the concentration of the supporting electrolyte; double-layer capacitance as a function of potential; and static friction between surfaces in solution.

The pzc for Pt containing a negligible amount of H is Eq=0 = 0·56 (±0·07) − 0·067 pH (3 methods). For Au, Eq=0 = 0·15 ± 0·02 (2 methods). For Ag, Eq=0 = −0·44 ± 0·02 (2 methods). For Ni, Eq=0 = −0·26 ± 0·06 (3 methods). Introduction of H into Pt shifts the pzc in the negative direction. Anodic and cathodic pulsing of Au and Ag electrodes produced surfaces which gave pzc values 0·2 and 0·4 V respectively more positive than those for unpulsed electrodes.

Results reported earlier for Pt are for Pt containing dissolved H. The dissolved proton concentration, abnormally high near the surface, contributes significantly to the surface potential. The results obtained oil pulsing of silver and gold are for dermasorbed surfaces. The most satisfactory methods for the determination of the pzc are the capacitance, static friction, immersion, and tension vibration methods.  相似文献   


16.
In the present study, we have examined sulfation of cerium oxide via impregnation of (NH4)2SO4, followed by heating in the temperature range of 220–720°C, using Raman Spectroscopy. Based on the SO and SO stretching frequencies in the range of 900–1400 cm−1, a wide range of surface oxysulfur species and bulk cerium-oxy-sulfur species are identified. At 220°C, a mixture of (NH4)2SO4 crystals, SO2−4(aq) and HSO1−r(aq) is found to have formed on ceria's surface, whereas complete conversion of (NH4)2SO4 to SO2−4(aq) and HSO1−4(aq) occurs at 280°C. At 350°C, formation of a mixture of surface pyrosulfate S2O2−7(surf.0, consisting of two SO oscillators and a bulk type compound identified as Ce(IV)(SO4)x(SO3)2−x (0 < x < 2) have been observed. Upon introduction of moisture, the former transforms to HSO1−4(surf.), whereas the latter remains unchanged. At 400°C, only the bulk type compound can be observed. At 450°C, only Ce2(SO4)3 is generated and remains stable until 650°C. Further increase in the temperature to 720°C results in the formation of CeOSO4. The present study not only provides a more thorough understanding of the sulfation of cerium oxide at a molecular level, but also demonstrates that Raman spectroscopy is a highly effective technique to characterize sulfation of metal oxides.  相似文献   

17.
We report that ultrastable faujasite-based ruthenium zeolites are highly active catalysts for N2O decomposition at low temperature (120–200°C). The faujasite-based ruthenium catalysts showed activity for the decomposition of N2O per Ru3+ cation equivalent to the ZSM-5 based ruthenium catalysts at much lower temperatures (TOF at 0.05 vol.-% N2O: 5.132 × 10−4 s−1 Ru−1 of Ru-HNaUSY at 200°C versus 5.609 × 10−4 s−1 Ru−1 of Ru-NaZSM-5 at 300°C). The kinetics of decomposition of N2O over a Ru-NaZSM-5 (Ru: 0.99 wt.-%), a Ru-HNaUSY (Ru: 1.45 wt.-%) and a Ru-free, Na-ZSM-5 catalyst were studied over the temperature range from 40 to 700°C using a temperature-programmed micro-reactor system. With partial pressures of N2O and O2 up to 0.5 vol.-% and 5 vol.-%, respectively, the decomposition rate data are represented by: −dN2O/dt=itk(PN2O) (PO2)−0.5 for Ru-HNaUSY, −dN2O/dt=k(PN2O) (PO2)−0.1 for Ru-NaZSM-5, and −dN2O/dt=k(PN2O)−0.2 (PO2)−0.1 for Na-ZSM-5. Oxygen had a stronger inhibition effect on the Ru-HNaUSY catalyst than on Ru-NaZSM-5. The oxygen inhibition effect was more pronounced at low temperature than at high temperature. We propose that the negative effect of oxygen on the rate of N2O decomposition over Ru-HNaUSY is stronger than Ru-NaZSM-5 because at the lower temperatures (<200°C) the desorption of oxygen is a rate-limiting step over the faujasite-based catalyst. The apparent activation energy for N2O decomposition in the absence of oxygen is much lower on Ru-HNaUSY (Ea: 46 kJ mol−1) than on Ru-NaZSM-5 (Ea: 220 kJ mol−1).  相似文献   

18.
The phase transformation and subsequent droplet growth of the mixed salt aerosols NaCl—KCl and (NH4)2SO4—H2SO4 were investigated in a continuous-flow apparatus at 25 and 30°C as a function of relative humidity. Monodisperse salt aerosols (d = ≈ 0.5 μm, OG = 1.07–1.13) were prepared and mixed with N2 carrier gas at controlled humidities. The particle-size distribution of the aerosol before and after growth by water vapor condensation was continuously monitored with an optical particle counter. It was found that mixed salt aerosols were characterized by stage-wise growth when the relative humidity in the atmosphere was increased. The onset of growth took place at a specific deliquescence humidity determined by the water activity at the eutonic composition. Thus, mixed NaCl—KCl aerosols deliquesce at 73.8 ± 0.5% r.h. regardless of initial compositions. For sulfate aerosols containing 0.75 to 0.95 mole fraction (NH4)2SO4 (the balance being H2SO4), the onset of growth occurs at 69.0 ± 0.5% r.h.. In the composition range of 0.5 to 0.75, a deliquescence humidity of 39.0 ± 0.5% is noted. Below 0.5 mole fraction, however, the mixed-sulfate aerosols are expected to exhibit hygroscopic properties on the basis of thermodynamic considerations.  相似文献   

19.
The standard molal potentials E°m of the Hg/Hg2(OPr)2, OPr electrode at 15°, 20°, 25°, 30° and 35° C have been determined. The E°m values obtained are 0.5114, 0.5072, 0.5031, 0.4988 and 0.4942 V respectively, which can be fitted to the equation Edgm/V = 0.5031 −8.56 × 10−4 (itt/°C − 25)−3.0588 × 10−6 (t/ °C -25)2. The changes in standard free energy, entropy and enthalpy for the cell reaction have been calculated.  相似文献   

20.
The paper is concerned with the dispersion of a solute in a Bingham plastic fluid flowing in a pipe or a parallel plate channel. For pipe flow, the dispersion coefficient K2 first increases with ξ0 (the dimensionless radius of the plug flow region), reaches a maximum and then decreases. But in a channel flow, K2 decreases monotonically with increasing ξ0. Further K2 for channel flow is found to be larger than that for pipe flow for all values of ξ0 except 0.8≤ξ0≤1.  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