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 859 毫秒
1.
The synthesis of a new series of 2,6-bis(imino)pyrazinyl ligands, [ArNCPyzCNAr] where the aryl groups Ar = naphtyl, 2,6-dimethylphenyl, 2,6-diisopropylphenyl, 2,4,6-trimethylphenyl, and their iron(II) complexes is described starting from monoacetylpyrazine.  相似文献   

2.
Amberlite XAD-16 resin has been functionalized using nitrosonaphthol as a ligand and characterized employing elemental, thermogravimetric analysis and FT-IR spectroscopy. The sorption of Ni(II) and Cu(II) ions onto this functionalized resin is investigated and optimized with respect to the sorptive medium (pH), shaking speed and equilibration time between liquid and solid phases. The monitoring of the influence of diverse ions on the sorption of metal ions has revealed that phosphate, bicarbonate and citrate reduce the sorption up to 10–14%. The sorption data followed Langmuir, Freundlich, and Dubinin–Radushkevich (D–R) isotherms. The Freundlich parameters computed are 1/n = 0.56 ± 0.03 and 0.49 ± 0.05, A = 9.54 ± 1.5 and 6.0 ± 0.5 mmol g−1 for Ni(II) and Cu(II) ions, respectively. D–R isotherm yields the values of Xm = 0.87 ± 0.07 and 0.35 ± 0.05 mmol g−1 and of E = 9.5 ± 0.23 and 12.3 ± 0.6 kJ mol−1 for Ni(II) and Cu(II) ions, respectively. Langmuir characteristic constants estimated are Q = 0.082 ± 0.005 and 0.063 ± 0.003 mmol g−1, b = (4.7 ± 0.2) × 104 and (7.31 ± 0.11) × 104 l mol−1 for Ni(II) and Cu(II) ions, respectively. The variation of sorption with temperature gives thermodynamic quantities of ΔH = −58.9 ± 0.12 and −40.38 ± 0.11 kJ mol−1, ΔS = −183 ± 10 and −130 ± 8 J mol−1 K−1 and ΔG = −4.4 ± 0.09 and −2.06 ± 0.08 kJ mol−1 at 298 K for Ni(II) and Cu(II) ions, respectively. Using kinetic equations, values of intraparticle transport and of first order rate constant have been computed for both the metal ions. The sorption procedure is utilized to preconcentrate these ions prior to their determination in tea, vegetable oil, hydrogenated oil (ghee) and palm oil by atomic absorption spectrometry using direct and standard addition methods.  相似文献   

3.
The polymerization of styrene with novel catalytic systems of anilido-imino nickel complexes (Ar1N = CHC6H4NAr2) NiBr (Ar1 = Ar2 = 2,6-dimethylphenyl, 1; Ar1 = 2,6-dimethylphenyl, Ar2 = 2,6-diisopropylphenyl, 2; Ar1 = Ar2 = 2,6-diisopropylphenyl, 3; Ar1 = 2,6-diisopropylphenyl, Ar2 = 2,6-dimethylphenyl, 4) activated by methylaluminoxane was investigated. The influence of reaction parameters (temperature, Al/Ni mole ratio, and polymerization time) on styrene polymerization was evaluated. The influence of the bulkiness of the substituents on polymerization activity and polymer characteristics was also ascertained. The obtained polystyrene was an iso-rich atactic polymer and its weight-average molecular weight reached 70 500. NMR analysis of the end groups further confirmed that styrene polymerization catalyzed by anilido-imino nickel complexes/MAO systems proceeded through a coordination mechanism, and the chain was initiated through styrene secondary insertion into the NiH and terminated mainly through β-H elimination of styrene producing the chain-end group (CHCHPh).  相似文献   

4.
A series of Cu(II) complexes of disubstituted 2,2′-bipyridine bearing ammonium groups [Cu(L1−4)2Br]5+ (1–4, L1 = [5,5′-(Me2NHCH2)2-bpy]2+, L2 = [5,5′-(Me3NCH2)2-bpy]2+, L3 = [4,4′-(Me2NHCH2)2-bpy]2+, L4 = [4,4′-(Me3NCH2)2-bpy]2+ and bpy = 2,2′-bipyridyl) were synthesized, of which complexes 1 and 4 were structurally characterized. Both coordination configurations of Cu(II) ions can be described as distorted trigonal bipyramid. The interaction between all complexes and CT-DNA was evaluated by thermal-denaturation experiments and CD spectroscopy. Results show that the complexes interact with CT-DNA via outside electrostatic interactions and their binding ability follows the order: 1 > 2 > 3 > 4. In the absence of any reducing agents, the cleavage of plasmid pBR322 DNA by these complexes was investigated and the hydrolysis kinetics of DNA was studied in Tris buffer (pH 7.5) at 37 °C. Obtained pseudo-Michaelis–Menten kinetic parameters: 15.0, 13.6, 2.01 and 1.69 h−1 for 1, 2, 3 and 4, respectively, indicate that complexes 1 and 2 exhibit very high DNA cleavage activities. According to their crystal data, the high nuclease activity may be attributed to the strong interaction of the metal moiety and two ammonium groups with phosphate groups of DNA.  相似文献   

5.
The crystal and magnetic structure of Sr2ErRuO6 has been studied by means of neutron powder diffraction as well as magnetization and susceptibility measurements. Neutron diffraction profile measured at 50 K shows that the Ru5+ and Er3+ are ordered in the B-sites of the perovskite-type structure, while the Sr atoms occupy the A-site. This compound crystallizes with a monoclinic unit cell, space group P21/n and lattice parameters are approximately √2ap × √2ap × 2ap. Magnetic susceptibility measurements reveal the existence of antiferromagnetic interactions in which Ru5+ and Er3+ sublattices are involved. The field dependence of the magnetization indicates the presence of a weak ferromagnetic component at the transition temperature, arising from the spin canting of the antiferromagnetically ordered Ru5+ and Er3+ moments. Thermal evolution of the neutron diffraction patterns indicate that the Nèel temperature is 36 K and the magnetic reflections can be indexed on the basis of a propagation vector k = [0, 0, 0]. The spin arrangement is described by the AxAz magnetic modes where the Ru5+ and Er3+ moments are mainly aligned along the c-axis of the structure, forming an angle of 6° with the c-axis in the case of the Er3+ sublattice and 15° for the Ru5+ moment.  相似文献   

6.
A tetra-nuclear copper(II) complex [Cu4(C54H46N4O14)(OH)2] · 10H2O (1) has been synthesized starting from l-tyrosine, NaOH, 2,6-diformyl-4-methylphenol (dfp) and CuSO4 · 5H2O. Compound 1 crystallizes from an ethanol–water mixture in triclinic space group. In the crystal of 1, two binuclear copper units, related by a center of symmetry, are bridged by two hydroxo bridges and results in the formation of a tetra-nuclear {Cu4} structure. Five lattice water molecules, located in the asymmetric unit, interact among themselves and form an unusual form of a water nonamer. In the crystal, the water nonamer is again hydrogen bonded to the next nonamer forming a chainlike polymer. Each {Cu4} complex unit attaches four such water nonamer chains. Variable temperature magnetic data fit to the Bleaney–Bower’s equation with a Curie type of impurity of S = 0.5. The best fit of the magnetic data to this equation yielded 2J = −217, g = 2.019 and a TIP value of 60 × 10−6 cm3 mol−1.  相似文献   

7.
Ruthenium(III) acetylacetonate was employed for the first time as homogeneous catalyst in the hydrolysis of sodium borohydride. Ruthenium(III) acetylacetonate was not reduced by sodium borohydride under the experimental conditions and remains unchanged after the catalysis. Poisoning experiments with mercury and trimethylphosphite provide compelling evidence for the fact that ruthenium(III) acetylacetonate is indeed a homogenous catalyst in the hydrolysis of sodium borohydride. Kinetics of the ruthenium(III) acetylacetonate catalyzed hydrolysis of sodium borohydride was studied depending on the catalyst concentration, substrate concentration, and temperature. The hydrogen generation was found to be first order with respect to both the substrate concentration and catalyst concentration. The activation parameters of this reaction were also determined from the evaluation of the kinetic data: activation energy; Ea = 58.2 ± 2.6 kJ mol−1, the enthalpy of activation; ΔH# = 55.7 ± 2.5 kJ mol−1 and the entropy of activation ΔS# = 118 ± 5 J mol−1 K−1. Ruthenium(III) acetylacetonate was found to be highly active catalyst providing 1200 turnovers over 180 min in hydrogen generation from the hydrolysis of sodium borohydride before deactivation.  相似文献   

8.
2-氨基吡啶镍配合物/MAO高活性催化β-蒎烯聚合研究   总被引:2,自引:0,他引:2  
合成了一系列2-氨基吡啶镍配合物(2-PyCH2NAr)NiBr,Ar=2,6-二甲基苯基(a),2,6-二异丙基苯基(b),2,6-二氟苯基(c).在助催剂甲基铝氧烷(MAO)存在下,该系列配合物能高活性催化β-蒎烯聚合,得到的聚β-蒎烯分子量明显比传统正离子聚合所得到的聚合物高.对配合物配体结构以及聚合条件对该聚合的催化活性以及聚合物分子量的影响进行了研究.所得聚合物经1H-NMR和13C-NMR分析表明,β-蒎烯聚合是通过正离子方式进行的,聚合中产生开环异构化,得到由环己烯和异丁烷结构单元交替组成的聚β-蒎烯.  相似文献   

9.
A series of hydroxyl-conducting anion-exchange membranes were prepared by blending chloroacetylated poly(2,6-dimethyl-1,4-phenylene oxide) (CPPO) with bromomethylated poly(2,6-dimethyl-1,4-phenylene oxide) (BPPO), and their fuel cell-related performances were evaluated. The resulting membranes exhibited high hydroxyl conductivities (0.022–0.032 S cm−1 at 25 °C) and low methanol permeability (1.35 × 10−7 to 1.46 × 10−7 cm2 s−1). All the blend membranes proved to be miscible or partially miscible under the investigations of scanning electron microscopy (SEM) and differential scanning calorimeters (DSC). By condition optimization, the blend membranes with 30–40 wt% CPPO are recommended for application in direct methanol alkaline fuel cells because they showed low methanol permeability, excellent mechanical properties and comparatively high hydroxyl conductivity.  相似文献   

10.
The enthalpy and entropy of sublimation of N-ethylthiourea were obtained from the temperature dependence of its vapour pressure measured by both the torsion–effusion and the Knudsen effusion method in the temperature range 360–380 K. The compound undergoes no solid-to-solid phase transition or decomposition below 380 K. The pressure against reciprocal temperature resulted in lg(p, kPa) = (13.40 ± 0.27) − (6067 ± 102) /T(K). The molar sublimation enthalpy and entropy at the mid interval temperature were ΔsubHm(370 K) = (116.1 ± 2.0) kJ mol−1 and ΔsubSm(370 K) = (218.0 ± 5.2) J mol−1 K−1, respectively. The same quantities derived at 298.15 K were (118.8 ± 2.1) kJ mol−1 and (226.1 ± 5.5) J mol−1 K−1, respectively.  相似文献   

11.
Rate coefficients for the reactions of cyclohexadienyl (c-C6H7) radicals with O2 and NO were measured at 296 ± 2 K. The c-C6H7 radicals were detected selectively by laser-induced fluorescence. The rate coefficient for the reaction of c-C6H7 with O2, (4.4 ± 0.5) × 10−14 cm3 molecule−1 s−1, was independent of the bath-gas (He) pressure (13–80 Torr). In the reaction of c-C6H7 with NO, thermal equilibrium among c-C6H7, NO, and C6H7NO was observed. The forward and reverse reactions were in the falloff region, and the equilibrium constant was (1.5 ± 0.6) × 10−15 cm3 molecule−1.  相似文献   

12.
A variety of neutral palladium(II) complexes [Pd(L–L)Cl2] containing 1,3-di(2-pyridyl)propane (1), 1,3-bis(2-pyridyl)-2-pentylpropane (2), 1,3-bis(2-pyridyl)-2-phenylpropane (3a), 1,3-bis(2-pyridyl)-2-tolylpropane (4), and 1,3-bis(2-pyridyl)-2-ferrocenylpropane (5) as chelate ligands (L–L) have been synthesized. The crystal structures of 1,3-diphenyl-2,4-di-pyridin-2-yl-butan-1-ol (3b), 5, [(2)PdCl2], [(4)PdCl2], and [(5)PdCl2] have been determined and show a square planar geometry at palladium(II). The neutral complexes were tested in the polymerization of norbornene and copolymerization of norbornene with norbornene derivatives. The complex bearing the pentyl group exhibited high reactivity to give up to 5.9×105 in molecular weight for the homopolymerization. When [(4)PdCl2] or [(5)PdCl2] was used as a catalyst, homopolymers insoluble at 150 °C in trichlorobenzene were obtained. However, copolymerization of norbornene with norbornene derivatives 8a–d catalyzed by [(4)PdCl2] gave soluble copolymers with molecular weights up to 5.1×105.  相似文献   

13.
Trace amounts of nickel(II) can function as a trigger (=reaction initiator) in an autocatalytic reaction with the sodium sulfite/hydrogen peroxide system. Based on this finding, sub-μg L−1 levels of nickel(II) were determined by a time measurement using the autocatalytic reaction. The detection range using the above method was 10−9–10−5 M, the detection limit (3σ) was 8.1 × 10−10 M (0.047 μg L−1), and the relative standard deviation was 2.66% at nickel(II) concentration of 10−7 M (n = 7). This method was applied to length detection-flow injection analysis. The detection range for the flow injection analysis was 2 × 10−9–2 × 10−3 M. The detection limit (3σ) was 1.4 × 10−9 M (0.082 μg L−1), and the relative standard deviation was 1.86 at initial nickel(II) concentration of 10−6 M (n = 7).  相似文献   

14.
Neutral salicylaldiminato Ni(II) complexes bearing a single N-heterocyclic carbene (NHC) ligand [3,5-tBu2-2-(O)C6H2CHNAr]Ni(C{RNCHCHNiPr})Ph [Ar = 2,6-iPr2C6H3, R = Bn (1); Ar = 2,6-iPr2C6H3, R = iPr (2)], have been synthesized via a one-pot procedure in high yield. The X-ray structure analysis reveals that both of 1 and 2 adopt distorted square-planar coordination geometry and NHC carbon (Ccarbene) is trans to the ketimine nitrogen. Preliminary study indicates that complex 1 is inert toward the insertion of ethylene, however, it can catalyze the dimerization of ethylene in the presence of modified methylaluminoxane (MMAO) with a moderate activity of 3.05 × 104 g(mol Ni)−1 h −1 atm−1 in a highly selective fashion.  相似文献   

15.
Activated carbons are produced from wastes of Algerian date stones by pyrolysis and physical activation in the presence of water vapor into a heated fixed-bed reactor. The effect of pyrolysis temperature and activation hold time on textural and chemical surface properties of raw date stones and carbon materials produced are studied. As expected, the percentage yield decreases with increase of activation temperature and hold time. The characterization of carbon materials is performed by scanning electron microscopy (SEM). X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR) and nitrogen adsorption (BET). Results show the presence of cellulose and hemicellulose in the raw material, and the predominance of carbon and graphite after pyrolysis. Different oxygen-containing functional groups are found in the raw material while aromatic structures are developed after pyrolysis and activation. The best specific surface area (635 m2 g−1) and microporous volume (0.716 cm3 g−1) are obtained when the date stones are grinded, pyrolysed at 700 °C under a 100 cm3 min−1 nitrogen flow and then activated under water vapor at 700 °C for 6 h.  相似文献   

16.
Using zinc hexamethylenedithiocarbamate (Zn(HMDC)2) and flame atomic absorption spectrometry (FAAS) and/or flow injection hydride generation atomic absorption spectrometry (FI-HGAAS), solvent extraction of As(III) from HCl and H2SO4 media into 2,6-dimethyl-4-heptanone (diisobutyl ketone, DIBK) was examined. Arsenic(III) was quantitatively extracted with 2.41×10−3 mol l−1 Zn(HMDC)2 from about 0.004 (pH 2.4) to 4 mol l−1 HCl and H2SO4 aqueous solutions. The logarithmic conditional extraction constant of As(HMDC)3 in the HCl–DIBK system was determined to be 8.3±0.7, by the measurement of the distribution ratios of Zn(II) and As(III). The effectiveness of the proposed extraction method was ascertained in the determination of As in geochemical standard reference materials supplied by the Geological Survey of Japan. Furthermore, the analysis of arsenic in procedural blanks was 0.083±0.003 μg l−1.  相似文献   

17.
The ring-opening metathesis polymerization (ROMP) of norbornene catalyzed by bis(acetonitrile) molybdenum and tungsten complexes, [M(η3-C3H5)Cl(CO)2(NCMe)2] (1-Mo: M = Mo, 1-W: M = W), which have two labile acetonitrile ligands, has been investigated. These complexes catalyzed the ROMP of norbornene as a single-component initiator. The highly cis-selective polymerization proceeded in a THF solution (95% for 1-Mo and 96% for 1-W), whereas polymerization in CH2Cl2 or toluene resulted in lower cis selectivity. The polymerization of terminal acetylenes using these complexes was also examined. The tungsten complex 1-W showed a high catalytic activity for the polymerization of terminal acetylenes, such as phenyl- and tert-butylacetylene. A highly active catalytic system for the ROMP of norbornene was achieved by the activation of the tungsten complex, 1-W, with one equivalent of phenylacetylene, giving poly(norbornene) with a high molecular weight (Mn = 391 × 104) and a high cis selectivity (cis  89%).  相似文献   

18.
The heat capacities and enthalpy increments of strontium bismuth niobate SrBi2Nb2O9 (SBN) and strontium bismuth tantalate SrBi2Ta2O9 (SBT) were measured by the relaxation method (2–150 K), Calvet-type heat-conduction calorimetry (305–570 K) and drop calorimetry (773–1373 K). The temperature dependences of non-transition heat capacities in the form Cpm = 324.47 + 0.06371T − 5.0755 × 106/T2 J K−1 mol−1 (298–1400 K) and Cpm = 320.22 + 0.06451T − 4.7001 × 106/T2 J K−1 mol−1 (298–1400 K) were derived for SBN and SBT, respectively, by the least-squares method from the experimental data. Furthermore, the standard molar entropies at 298.15 K Sm°(SBN)=327.15±0.80 and Sm°(SBT)=339.23±0.72 J K−1 mol−1 were evaluated from the low-temperature heat capacity measurements.  相似文献   

19.
With the aim of understanding the influence of donor solvents on the reactivity of the amine complexes [RuCl2(PPh3)2(piperidine)] (1) and [RuCl2(PPh3)2(imidazole)2] (2) in the presence of ethyldiazoacetate, and on the properties of the resulting polymer, a ring opening metathesis polymerization of norbornene was carried out in the presence of small amounts of common solvents such as additives (isopropanol, THF, N,N-dimethylformamide, 2,6-lutidine, isopropanethiol, acetonitrile, dimethyl sulfoxide, NEt3, NH2Me and pyridine). From observations, typical coordinating solvents like DMSO, NEt3, NH2Me and pyridine, hardly affected the yields when either complex was employed. With other additives, the major advantage was the decrease in the polydispersity indices. On using complex 1 with 2,6-lutidine, observed values of Mw/Mn were as low as 1.3, while the yield decreased from 99% to about 20–30% at RT for 1 min in pure solution. In the case of complex 2, which is almost inactive to ROMP (19% at 50 °C for 5 min with Mw/Mn = 6.30), the yield was three-fold (60% at 50 °C for 5 min with Mw/Mn = 1.95) compared to that of without THF. Further, the Mw/Mn was observed to decrease to 1.34 with 200 eq. of THF.  相似文献   

20.
The hydrogen permeation and stability of tubular palladium alloy (Pd–23%Ag) composite membranes have been investigated at elevated temperatures and pressures. In our analysis we differentiate between dilution of hydrogen by other gas components, hydrogen depletion along the membrane length, concentration polarization adjacent to the membrane surface, and effects due to surface adsorption, on the hydrogen flux. A maximum H2 flux of 1223 mL cm−2 min−1 or 8.4 mol m−2 s−1 was obtained at 400 °C and 26 bar hydrogen feed pressure, corresponding to a permeance of 6.4 × 10−3 mol m−2 s−1 Pa−0.5. A good linear relationship was found between hydrogen flux and pressure as predicted for rate controlling bulk diffusion. In a mixture of 50% H2 + 50% N2 a maximum H2 flux of 230 mL cm−2 min−1 and separation factor of 1400 were achieved at 26 bar. The large reduction in hydrogen flux is mainly caused by the build-up of a hydrogen-depleted concentration polarization layer adjacent to the membrane due to insufficient mass transport in the gas phase. Substituting N2 with CO2 results in further reduction of flux, but not as large as for CO where adsorption prevail as the dominating flow controlling factor. In WGS conditions (57.5% H2, 18.7% CO2, 3.8% CO, 1.2% CH4 and 18.7% steam), a H2 permeance of 1.1 × 10−3 mol m−2 s−1 Pa−0.5 was found at 400 °C and 26 bar feed pressure. Operating the membrane for 500 h under various conditions (WGS and H2 + N2 mixtures) at 26 bars indicated no membrane failure, but a small decrease in flux. A peculiar flux inhibiting effect of long term exposure to high concentration of N2 was observed. The membrane surface was deformed and expanded after operation, mainly following the topography of the macroporous support.  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