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 109 毫秒
1.
Negative ion fast atom bombardment mass spectra (FAB-MS) recorded for ZnCl2-1-ethyl-3-methylimidazolium chloride (ZnCl2-EMIC) ionic liquids with various compositions indicate that various Lewis acidic chlorozincate clusters (ZnCl3, Zn2Cl5 and Zn3Cl7) are present in ZnCl2-EMIC ionic liquids depending on the percentage of ZnCl2 used in preparing the ionic liquids; higher ZnCl2 percentage favors the larger clusters. Cyclic voltammetry reveals that the potential limits for a basic 1:3 ZnCl2-EMIC melt correspond to the cathodic reduction of EMI+ and anodic oxidation of Cl, giving an electrochemical window of approximately 3.0 V which is the same as that observed for basic AlCl3-EMIC ionic liquids. For acidic ionic liquids that have a ZnCl2/EMIC molar ratio higher than 0.5:1, the negative potential limit is due to the deposition of metallic zinc, and the positive potential limit is due to the oxidation of the chlorozincate complexes. All the acidic ionic liquids exhibit an electrochemical window of approximately 2 V, although the potential limits shifted in the positive direction with increasing ZnCl2 mole ratio. Underpotential deposition of zinc was observed on Pt and Ni electrodes in the acidic ionic liquids. At proper temperatures and potentials, crystalline zinc electrodeposits were obtained from the acidic ionic liquids.  相似文献   

2.
The crystal structures of trifluoroacetato-[meso-tetra(p-chlorophenyl)porphyrinato]thallium(III), Tl[(p-Cl)4tpp](O2CCF3) (1), and pentafluoropropionato-[meso-tetra(p-chlorophenyl)porphyrinato]thallium(III) Tl[(p-Cl)4tpp](O2CCF2CF3) (2), were determined. The coordination sphere around the Tl3+ ion is described as five-coordinate regular square-based pyramid (RSBP) in which the apical site is occupied by an unidentate CF3CO2 ligand for 1 whereas the unidentate CF3CF2CO2 ligand occupies the axial site for 2. The plane of the four pyrrole nitrogen atoms [i.e., N(1)–N(4)] strongly bonded to Tl3+ is adopted as a reference plane 4N. The Tl3+ is moderately out of the 4N plane; its displacement of 0.58 Å (or 0.59 Å) for 1 (or 2) is in the same direction as that of the trifluoroacetate oxygen for 1 (or pentafluoropropionate oxygen for 2). The free energy of activation at the coalescence temperature Tc for the intermolecular trifluoroacetate exchange for 1 in CD2Cl2 is found to be ΔG178=36.6 kJ/mol whereas the intermolecular pentafluoropropionate exchange for 2 in CD2Cl2 is determined to be ΔG213=41.5 kJ/mol through 19F and 13C NMR temperature-dependent measurements.  相似文献   

3.
《Reactive Polymers》1990,12(1):59-73
Eleven novel metal ion-chelating resins based upon porous poly(glycidyl methacrylate-co-ethylene glycol dimethacrylate) have been prepared by direct reaction of the polymer with ligands containing an amino group. Those species involving a 2-pyridyl-2-imidazole and a 2-aminomethylpyridine proved to be the most interesting and, for comparison, polystyrene resin analogues of these were also produced. As expected from the literature, the above pyridine-containing ligands remained effective for the extraction of Cu2+ down to pH 2. Despite having superficially similar overall physical characteristics (particle size, surface area, porosity) the glycidyl methacrylate-based resins showed remarkably enhanced kinetic behaviour in the batch extraction of base metal ions such as Cu2+, with an extraction half life as low as ∼ 8 minutes, four times faster than a polystyrene-based analogue.In addition the glycidyl methacrylate-based resins displayed superior selectivity in the extraction of particular metal ions from feed liquors containing a mixture of metal ions. The most remarkable separation discovered was that of Cu2+ from Zn2+ where the glycidyl methacrylate-based species with the above ligands achieved essentially quantitative separation of Cu2+, even with Zn2+ present in 250 times excess. Extracted metal ions were rapidly and quantitatively eluted from these resins with 2 M H2SO4, and in the case of Cu2+ an acid strength of only 0.25 M was adequate.  相似文献   

4.
Macroreticular cation exchange resins containing phosphoric acid groups (RGP) were prepared by the reaction of glycidyl methacrylate–divinylbenzene copolymer [or poly(glycidyl methacrylate)]beads (RG) with phosphoric acid or phosphorous oxychloride, and the adsorption behavior of metal ions on the RGP was investigated. The phosphorylation of the polymer beads could be effectively carried out by treatment of the polymer beads with 85% phosphoric acid at 80°C for 3 h. The RGP obtained from glycidyl methacrylate–divinylbenzene (2 mol %) copolymer beads showed high cation exchange capacity, salt splitting capacity, and adsorption capacity for Cu2+, Zn2+, Cd2+, Ca2+, and Ag+. On the other hand, the RGP obtained from poly(glycidyl methacrylate)beads had high adsorption capacity for Al3+, Fe3+, and UO22+. The RGP prepared by treating the RG with phosphoric acid had a higher selective adsorption for Li+ than for Na+.  相似文献   

5.
The catalytic activity in the polymerization of styrene has been examined using commercially available simple rare earth metal compounds such as Sm(OiPr)3, Sm(acac)3, Sm(OCOMe)3, SmI2(THF)2 or SmCl3 coupled with Et3Al or methylaluminoxane (MAO). Among these compounds, the Sm(OiPr)3/AlEt3 system shows the highest catalytic activity, especially in the presence of a minor amount of toluene at 60 °C. The random copolymerization of styrene with methyl methacrylate suggests that the present polymerization proceeds with a radical polymerization mechanism. (C5Me5)SmCl3Na(THF) and (C5Me5)SmCl3Li(THF) systems exhibit relatively low catalytic activity, even in the presence of AlEt3. © 2001 Society of Chemical Industry  相似文献   

6.
Little is known about the hydrogenation and cracking of fused aromatic nuclei during the liquefaction of coal under the influence of Lewis acid catalysts. This study was conducted to establish the effects of catalyst acidity on the activity and selectivity of Lewis acid catalysts, the sources of hydrogen involved in hydrogenation and cracking, and the relationships between reactant structure and reactivity. Three-ring aromatic and hydroaromatic compounds were used to simulate some of the structural units present in coal. The catalysts examined were ZnCl2 and AlCl3. It has been established that the rates of both processes are strongly influenced by the Brönsted acidity of the active catalyst, e.g. H+ (MXnY)?, and the Brönsted basicity of the aromatic portions of the reactant. The source of the hydrogen used for hydrogenation depends on the choice of catalyst. In the presence of AlCl3, Scholl condensation of aromatic nuclei serves as the principal source of hydrogen. Molecular hydrogen is used exclusively, however, when hydrogenation is catalysed by ZnCl2. The formation of reaction products and the trends in reactant reactivity are discussed on the basis of cationic mechanisms. The results of this study contribute to an understanding of the processes which occur during the liquefaction of coal using ZnCl2 or AlCl3.  相似文献   

7.
Crystallization behavior and melt structure of two typical mold fluxes A (CaO–SiO2-based) and B (CaO–Al2O3-based) for casting high-aluminum steel were investigated using double hot thermocouple technology (DHTT), X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy. The results suggest that the crystallization temperature of Flux B is higher, and its crystallization incubation time is shorter compared with Flux A. The precipitated phase in Flux A is CaSiO3, whereas BaAl2O4 and Ca2Al2SiO7 form in Flux B. The structure analyses suggest that the degree of polymerization of Flux A is larger than that of Flux B. In addition, the major structural units of Flux A are Si–O–Si, Q0Si, Q1Si, Q2Si and Q3Si, but those of Flux B are mainly aluminate (Al–O–Al, Al–O-), aluminosilicate (Al–O–Si) and silicate units (Q0Si, Q1Si, Q2Si and Q3Si). These different melt structures are the main reasons why the precipitated phases in these two mold fluxes are different, and the crystallization ability of Flux A is weaker than Flux B.  相似文献   

8.
Clay (kaolin, mont-K10 and mont-KSF) supported InCl3, GaCl3, FeCl3 and ZnCl2 catalysts (metal chloride loading=1.1 mmol g−1) show high selectivity (⩾98%) at high conversion in the esterification of tert-butanol by acetic anhydride to tert-butyl acetate (t-BA) and very low activity for the dehydration of tert-butanol at ⩽50°C. For all the catalysts, mont-K10 is the best support and the order of their esterification activity (at 26°C) is: InCl3/mont-K10 (TOF=0.025 s−1) > GaCl3/mont-K10 (0.023 s−1) > FeCl3/mont-K10 (0.02 s−1)  ZnCl2/mont-K10 (0.019 s−1). InCl3/mont-K10 is highly active, selective and reusable catalyst for the esterification.  相似文献   

9.
The in situ preparation of the unique complex [CoClL1][ZnI3L2] (1) (L1 = tris(1-(3,5-dimethylpyrazolyl)-methyl)amine, L2 = hexamethylenetetramine) via redox and condensation processes of zerovalent cobalt, ZnCl2 and 1-hydroxymethyl-3,5-dimethylpirazole has been reported. The crystal structure of 1, determined by single-crystal X-ray diffraction exhibited a distorted trigonal bipyramidal coordination sphere for [CoCl(L1)]+ ion and a distorted tetrahedral arrangement of the [ZnI3(L2)]? anion, both of the C1 point group symmetry. The structure of 1 was confirmed by IR, FTIR, UV–Vis, magnetic and thermal investigations.  相似文献   

10.
Radical copolymerization of acrylonitrile (AN) with styrene (Sty), using x,x′-azobisisobutyronitrile as initiator, was carried out in the presence of zinc chloride (ZnCl2) dilatometrically at 65/pm 0.1 C for 120min. The rate of polymerization was a direct function of the concentrations of ZnCl2, AN and Sty, and polymerization temperature. The viscosity-average molecular weight of the copolymer increased with ZnCl2 concentration. The energy of activation in the presence and absence of the complex was evaluated as 82.5 kJ mol?1 and 115.5 kJ mol?1, respectively. The copolymerization of AN with Sty proceeded via the radical-complex mechanism.  相似文献   

11.
In this work, an approach towards ionic liquid crystals is strategically designed. A pyridine group has been attached to an 1,3-diketone containing an alkyloxyphenyl R substituent (R = C6H4OCnH2n+1; n = 10–18) and protonated with hydrochloric acid towards the formation of the chloride salts of 2-[3-(4-n-alkyloxyphenyl)propane-1,3-dion-1-yl]pyridinium cations [OOR(n)py]Cl (I), which have been proved to be non-mesomorphic. Reaction of these organic–inorganic compounds (I) with ZnCl2 yields to the ionic liquid crystals of the type bis{2-[3-(4-n-alkyloxyphenyl)propane-1,3-dion-1-yl]pyridinium}tetrachlorozincate [OOR(n)py]2[ZnCl4] (II).The crystalline structure of compounds [OOR(12)py]Cl and [OOR(10)py]2[ZnCl4] as representative examples of both kinds of salts have been solved and discussed. Both compounds exhibit layered structures containing cationic and anionic sublayers. In addition for the tetrachlorozincate salt [OOR(16)py]2[ZnCl4] structural relationships between the mesophases and the crystalline phase are proposed on the basis of the variable-temperature small-angle X-ray diffraction studies.  相似文献   

12.
A new ligand, 4-(p-chlorophenyl)-1,2,4-triazole (pCltrz), and its complexes, [Zn(pCltrz)6]2(ClO4)2 · H2O (1) and [Zn(pCltrz)2Cl2] (2), were synthesized. The structure of 1 exhibits the first case that zinc(II) ion coordinated with six triazole ligands to form octahedral environment. Both 1 and 2 display very strong blue photoluminescence as the result of the fluorescence from the intraligand emission excited state.  相似文献   

13.
Hydrogenation of Big Ben coal (Australian bituminous coal) with 9.8 MPa H2 for 3h at 400 °C has been carried out using a batch autoclave system in the presence of a molten salt catalyst such as ZnCl2, SnCl2, ZnCl2-KCl-NaCl (3:1:1 mol ratio) and SnCl2-KCl (3:2). The hexane-soluble (HS) yield decreases in the order: SnCl2-KCl, SnCl2, ZnCl2-KCl-NaCl, ZnCl2. The use of SnCl2-containing melt is characterized by higher yields of both HS and benzene-soluble (BS) fractions and the suppression of gas yield compared with the use of ZnCl2. The average aromatic unit and the molecular weight of HS increase in the order: ZnCl2, ZnCl2-KCl-NaCl, SnCl2, SnCl2-KCl. Chromatographic separation of HS fractions indicates that the saturate content is lowest and polar-material content is highest with the SnCl2-KCl melt: this fact coupled with the structural parameters suggests that HS material with SnCl2-KCl melt has, on average, a reasonably-high-molecular-weight skeleton which has more alkyl chains and heteroatoms. Micrographs of resulting benzene-insoluble (Bl) materials clearly indicated that the Bl particles were larger when ZnCl2 was used than when the other melts were used, the Bl particles in these cases being of similar size. Lewis acidity and the viscosity of the molten salt appear to be related, in part, to the size of the Bl particles.  相似文献   

14.
The well established activated carbon manufacturing process has been investigated as a novel treatment for contaminated soil from gaswork sites by converting it into a porous carbonaceous solid with adsorbent properties. Several activation methodologies were evaluated: CO2, air, ZnCl2, H2SO4, H3PO4, FeSO4 and HNO3. Thermal analysis of the soil provided information regarding appropriate carbonisation and activation conditions. Bulk samples were prepared using contaminated soil samples, with ZnCl2 being found to be the most effective agent for the process, producing an adsorbent which possessed a BET surface area of 131m2g−1. The aqueous adsorption ability of the soil carbons was studied using phenol and 4‐nitrophenol as representative micropollutant organic molecules. The Langmuir monolayer capacity of the ZnCl2‐activated soil was found to be 0.12 mmg−1 for phenol and 0.23 mmg−1 for 4‐nitrophenol. © 2000 Society of Chemical Industry  相似文献   

15.
Diphenyl carbonate (DPC) was synthesized from phenol and dense phase CO2 in the presence of CCl4 and K2CO3 using different catalysts of ZnCl2, ZnBr2, Lewis acid ionic liquids including 1-butyl-3-methylimidazolium chloride (BMIMCl) and bromide (BMIMBr). It was found that K2CO3 was not required, ZnCl2 and ZnBr2 were similar in the catalytic performance, and the use of BMIMCl and BMIMBr was not effective for the production of DPC. For the reactions with ZnCl2 in CCl4, the effects of such reaction variables as temperature, CO2 pressure, the amount of ZnCl2, and the volume of CCl4 were studied in detail. It was shown that the pressure was less influential while a larger amount of ZnCl2, a smaller volume of CCl4, and a low temperature of around 100°C were beneficial for the synthesis of DPC. On the basis of the results obtained, possible reaction mechanisms were discussed.  相似文献   

16.
A hydrogel composite that has been prepared by using waste linear low‐density polyethylene, acrylic acid, and organo‐montmorillonite (LLDPE‐g‐PAA/OMMT) is used as a hydrogel electrolyte. An absorbency test was used to determine the percentage of ZnCl2 solution absorbed by the hydrogel composite. The swelling behavior of the hydrogel composite in the ZnCl2 solution was then studied. The highest absorbency was recorded when the concentration of ZnCl2 solution was 3 M. The conductivity of ZnCl2‐hydrogel composite electrolytes is dependent on the solution's concentration. A mixture of ZnCl2 solution with hydrogel composite yields a good hydrogel composite electrolyte with a conductivity of 0.039 S cm?1 at 3 M ZnCl2. The hydrogel composite electrolyte was used to produce zinc‐carbon cells. The fabricated cell gives capacity of 7.8 mAh, has an internal resistance of 9.9 Ω, a maximum power density of 15.78 mWcm?2, and a short‐circuit current density of 43.75 mAcm?2 for ZnCl2‐hydrogel composite electrolytes. J. VINYL ADDIT. TECHNOL., 22:279–284, 2016. © 2014 Society of Plastics Engineers  相似文献   

17.
Polymerization of 1-butyne and isopropylacetylene was studied using MoCl5, WCl6, and two Ziegler catalysts [Fe(acac)3-Et3Al, Ti(On-Bu)4-Et3Al]. 1-Butyne was polymerized in high yields with WCl6 and Fe(acac)3-Et3Al to give a yellow, air-sensitive polymer. The cis content of poly(1-butyne), evaluated by 13C n.m.r., was about 80% irrespective of polymerization conditions. Isopropylacetylene was polymerized well by any of MoCl5, WCl6, and Fe(acac)3-Et3Al; the polymer formed was a light yellow air-sensitive powder. The cis content of poly(isopropylacetylene) varied from 65% to 90% according to polymerization conditions. Substituent effects on polymerization and polymer structure are discussed.  相似文献   

18.
Yuan Kong  Xiuli Dou  Qigu Huang  Kejing Gao 《Polymer》2010,51(17):3859-3179
Comparison with the conventional Ziegler-Natta catalyst TiCl4/MgCl2 (I), the modified supported Ziegler-Natta catalysts (iso-PentylO)TiCl3/MgCl2 (II) and (BzO)TiCl3/MgCl2 (III) were prepared as efficient catalysts for copolymerization of ethylene with 1-octene. The complexes (II) and (III) were desirable for the production of random ethylene/1-octene copolymers coupled with higher molecular weight, higher comonomer incorporation within copolymer chain and good yield even at high temperature 80 °C and fairly low Al/Ti molar ratio of 100. The effects of catalysts ligands, Al/Ti molar ratio, polymerization temperature, as well as concentration of 1-octene on the catalytic activity, molecular weight and microstructure of the copolymer were investigated in detail. The structure and properties of the copolymers were characterized with 13C NMR, GPC, DSC and WAXD. The kinetic results also indicate that these catalysts (II) and (III) show higher catalytic activity and the produced polymers feature higher molecular weight, because of lower ratio of Ktrm/Kp and Ktra/Kp, and higher ratio of Ktra/Ktrm which indicates that chain transfer to cocatalyst is predominant.  相似文献   

19.
The formation constants of the species ZnC+, ZnCl2, ZnBr+ and ZnBr2 were determined from emf measurements in suitable concentration cells. Mixtures of NH4NO3Ca(NO3)2 with a molar ratio 0.844/0.156 containing variable amounts of water (from 0 to 1.2 moles per mol of salt) were used as a solvent. The bromide complexes were found to be more stable than the chloride complexes in the anhydrous melt, while the opposite was found in the concentrated aqueous solutions. The results are discussed on the basis of models for chemical equilibria in molten salts and aqueous melts.  相似文献   

20.
The production of spiramycin using a synthetic basal medium, to which amino acids, vitamins and metallic ions were added, was investigated. L -Tryptophan was found to be the amino acid which best increased the antibiotic yield. Addition of vitamins to the fermentation medium did not increase the antibiotic output, while the addition of different concentrations of trace elements such as Co(NO3)2 ZnCl2 (NH4)2Mo7O24 and MnCl2 improved the antibiotic yield. High concentrations of these trace elements decreased the antibiotic yield and directed the microbial activities towards biosynthesis of spiramycin II and III—these are less toxic than spiramycin I. Optimum concentrations of Co(NO3)2, ZnCl2, (NH4)2Mo24O7, MnCl2 which supported the antibiotic yield were 5 × 10?1, 5 × 10?3, 5 × 10?3 and 5 × 10?3 g dm?3, respectively.  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