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 750 毫秒
1.
In this paper, we summarise our recent research interest in the hydrothermal synthesis and structural characterisation of multi-dimensional coordination polymers. The use of N-(phosphonomethyl)iminodiacetic acid (also referred to as H4pmida) in the literature as a versatile chelating organic ligand is briefly reviewed. This molecule plays an important role in the formation of centrosymmetric dimeric [V2O2(pmida)2]4− anionic units, which were first used by us as building blocks to construct novel coordination polymers. Starting with [V2O2(pmida)2]4− in solution, we have isolated [M2V2O2(pmida)2(H2O)10] species (where M2+ = Mn2+, Co2+ or Cd2+) via the hydrothermal synthetic approach, which were then employed for the construction of [CdVO(pmida)(4,4′-bpy)(H2O)2]·(4,4′-bpy)0.5·(H2O), [CoVO(pmida)(4,4′-bpy)(H2O)2]·(4,4′-bpy)0.5, [Co(H2O)6][CoV2O2(pmida)2(pyr)(H2O)2]·2(H2O) and [Cd2V2O2(pmida)2(pyr)2(H2O)4]·4(H2O) by the inclusion of bridging organic ligands in the reactive mixtures, such as pyrazine (pyr) and 4,4′-bipyridine (4,4′-bpy). These materials can contain channel systems, and exhibit magnetic behaviour, not only due to the V4+ centres but also to the transition metal centres which establish the links between neighbouring dimeric [V2O2(pmida)2]4− anionic units. A closely related anionic moiety, [Ge2(pmida)2(OH)2]2−, was engineered to allow the study of such crystalline hybrid materials using one- and two-dimensional high-resolution solid-state NMR.  相似文献   

2.
The effects of doping of Co3O4with MgO (0.4–6 mol%) and V2O5 (0.20–0.75 mol%) on its surface and catalytic properties were investigated using nitrogen adsorption at −196°C and decomposition of H2O2 at 30–50°C. Pure and doped samples were prepared by thermal decomposition in air at 500–900°C, of pure basic cobalt carbonate and basic carbonate treated with different proportions of magnesium nitrate and ammonium vanadate. The results revealed that, V2O5 doping followed by precalcination at 500–900°C did not much modify the specific surface area of the treated Co3O4 solid. Treatment of Co3O4 with MgO at 500–900°C resulted in a significant increase in the specific surface area of cobaltic oxide. The catalytic activity in H2O2 decomposition, of Co3O4 was found to suffer a considerable increase by treatment with MgO. The maximum increase in the catalytic reaction rate constant (k) measured at 40°C on Co3O4 due to doping with 3 mol% MgO attained 218, 590 and 275% for the catalysts precalcined at 500, 700 and 900°C, respectively. V2O5-doping of Co3O4 brought about a significant progressive decrease in its catalytic activity. The maximum decrease in the reaction rate constant measured at 40°C over the 0.75 mol% V2O5-doped Co3O4 solid attained 68 and 93% for the catalyst samples precalcined at 500 and 900°C, respectively. The doping process did not modify the activation energy of the catalyzed reaction but much modified the concentration of catalytically active constituents without changing their energetic nature. MgO-doping increased the concentration of CO3+–CO2+ ion pairs and created Mg2+–CO3+ ion pairs increasing thus the number of active constituents involved in the catalytic decomposition of H2O2. V2O5-doping exerted an opposite effect via decreasing the number of CO3+–CO2+ ion pairs besides the possible formation of cobalt vanadate.  相似文献   

3.
The possibility to change the Seebeck coefficient sign has been evidenced in the LaCoO3 perovskites. A small hole doping (Co3+/Co4+) will result in a large positive Seebeck coefficient, while a small electron doping (Co2+/Co3+) will give a large negative Seebeck coefficient at room temperature. This mechanism is shown to be efficient as well in 1D Ca3Co2O6 deriving from hexagonal perovskites. By doping Ca3Co2O6 with Ti4+, a mixed valency Co2+/Co3+ is introduced and the thermopower turns negative.

At high temperature, the Seebeck coefficients of LaCoO3 and related compounds decrease to small values due to the spin state transition. The values converge towards a positive value, close to +20 μV/K at 800 K. This suggests that at high T, the Seebeck coefficients in the case of localized charges do not depend on the doping, but only on the spin and orbital degeneracies. On the other hand, in the case of metallic-like samples as electron-doped manganites, the properties can be described up to high T in terms of a single-band metal. Due to the linear variation of S as a function of T and the almost constant value of ρ, the ratio S2/ρ which is crucial for high temperature applications increases.  相似文献   


4.
The activity of enzyme I (EI), the first protein in the bacterial PEP:sugar phosphotransferase system, is regulated by a monomer–dimer equilibrium where a Mg2+-dependent autophosphorylation by PEP requires the homodimer. Using inactive EI(H189A), in which alanine is substituted for the active-site His189, substrate binding effects can be separated from those of phosphorylation. Whereas 1 mM PEP (with 2 mM Mg2+) strongly promotes dimerization of EI(H189A) at pH 7.5 and 20 °C, 5 mM pyruvate (with 2 mM Mg2+) has the opposite effect. A correlation between the coupling of N- and C-terminal domain unfolding, measured by differential scanning calorimetry, and the dimerization constant for EI, determined by sedimentation equilibrium, is observed. That is, when the coupling between N- and C-terminal domain unfolding produced by 0.2 or 1.0 mM PEP and 2 mM Mg2+ is inhibited by 5 mM pyruvate, the dimerization constant for EI(H189A) decreases from >108 to <5 × 105 or 3 × 107 M−1, respectively. With 2 mM Mg2+ at 15–25 °C and pH 7.5, PEP has been found to bind to one site/monomer of EI(H189A) with KA′106 M−1G′=−33.7±0.2 kJ mol−1 and ΔH=+16.3 kJ mol−1 at 20 °C with ΔCp=−1.4 kJ K−1 mol−1). The binding of PEP to EI(H189A) is synergistic with that of Mg2+. Thus, physiological concentrations of PEP and Mg2+ increase, whereas pyruvate and Mg2+ decrease the amount of dimeric, active, dephospho-enzyme I.  相似文献   

5.
The influence of Li+, Na+, K+, Cs+, Mg2+, Ca2+, Sr2+, Ba2+, Co2+, Ni2+, Cu2+, Zn2+, Pb2+ and Al3+ ions on the spectroscopic properties of the dansyl group covalently linked to crown ether or diazacrown ethers was investigated by means of absorption and emission spectrophotometry. Interaction of the alkali metal ions with all fluoroionophores studied is weak, while alkaline earth metal ions interact strongly causing about 50% quenching of dansyl fluorescence of A21C5-Dns and A218C6-Dns. The Cu2+, Pb2+ and Al3+ cations interact very strongly with dansyl chromophore regardless the crown ether type, causing a major change in absorption spectrum of the chromophore and forming non-fluorescent complexes. The Co2+, Ni2+, Zn2, Mg2+ and Ag+ cations interact moderately with all fluoroionophores studied causing about 20% of fluorescence quenching of dansyl, except for a strong dansyl fluorescence quenching of 15C5-Dns by Co2+ ion. The quenching efficiency of didansylated fluoroionophores by the alkali metal ions and alkaline earth metal ions is weaker than monodansylated ones.  相似文献   

6.
A series of γ-Al2O3 samples modified with various contents of sulfate (0–15 wt.%) and calcined at different temperatures (350–750 °C) were prepared by an impregnation method and physically admixed with CuO–ZnO–Al2O3 methanol synthesis catalyst to form hybrid catalysts. The direct synthesis of dimethyl ether (DME) from syngas was carried out over the prepared hybrid catalysts under pressurized fixed-bed continuous flow conditions. The results revealed that the catalytic activity of SO42−/γ-Al2O3 for methanol dehydration increased significantly when the content of sulfate increased to 10 wt.%, resulting in the increase in both DME selectivity and CO conversion. However, when the content of sulfate of SO42−/γ-Al2O3 was further increased to 15 wt.%, the activity for methanol dehydration was increased, and the selectivity for DME decreased slightly as reflected in the increased formation of byproducts like hydrocarbons and CO2. On the other hand, when the calcination temperature of SO42−/γ-Al2O3 increased from 350 °C to 550 °C, both the CO conversion and the DME selectivity increased gradually, accompanied with the decreased formation of CO2. Nevertheless, a further increase in calcination temperature to 750 °C remarkably decreased the catalytic activity of SO42−/γ-Al2O3 for methanol dehydration, resulting in the significant decline in both DME selectivity and CO conversion. The hybrid catalyst containing the SO42−/γ-Al2O3 with 10 wt.% sulfate and calcined at 550 °C exhibited the highest selectivity and yield for the synthesis of DME.  相似文献   

7.
The hydrothermal reactions of vanadium oxide starting materials with divalent transition metal cations in the presence of nitrogen donor chelating ligands yield the bimetallic cluster complexes with the formulae [{Cd(phen2)2V4O12]·5H2O (1) and [Ni(phen)3]2[V4O12]·17.5H2O (2). Crystal data: C48H52Cd2N8O22V4 (1), triclinic. a=10.3366(10), b=11.320(3), c=13.268(3) Å, =103.888(17)°, β=92.256(15)°, γ=107.444(14)°, Z=1; C72H131N12Ni2O29.5V4 (2), triclinic. a=12.305(3), b=13.172(6), c=15.133(4), =79.05(3)°, β=76.09(2)°, γ=74.66(3)°, Z=1. Data were collected on a Siemens P4 four-circle diffractometer at 293 K in the range 1.59° <θ<26.02° and 2.01°<θ<25.01° using the ω-scan technique, respectively. The structure of 1 consists of a [V4O12]4− cluster covalently attached to two {Cd(phen)2}2+ fragments, in which the [V4O12]4− cluster adopts a chair-like configuration. In the structure of 2, the [V4O12]4− cluster is isolated. And the complex formed a layer structure via hydrogen bonds between the [V4O12]4− unit and crystallization water molecules.  相似文献   

8.
This work presents chemical modeling of solubilities of metal sulfates in aqueous solutions of sulfuric acid at high temperatures. Calculations were compared with experimental solubility measurements of hematite (Fe2O3) in aqueous ternary and quaternary systems of H2SO4, MgSO4 and Al2(SO4)3 at high temperatures. A hybrid model of ion-association and electrolyte non-random two liquid (ENRTL) theory was employed to fit solubility data in three ternary systems H2SO4–MgSO4–H2O, H2SO4–Al2(SO4)3–H2O at 235–270 °C and H2SO4–Fe2(SO4)3–H2O at 150–270 °C. Employing the Aspen Plus™ property program, the electrolyte NRTL local composition model was used for calculating activity coefficients of the ions Al3+, Mg2+ Fe3+ and SO42−, HSO4, OH, H3O+, respectively, as well as molecular species. The solid phases were hydronium alunite (H3O)Al3(SO4)2(OH)6, hematite Fe2O3 and magnesium sulfate monohydrate (MgSO4)·H2O which were employed as constraint precipitation solids in calculating the metal sulfate solubilities. A correlation for the equilibrium constants of the association reactions of complex species versus temperature was implemented. Based on the maximum-likelihood principle, the binary interaction energy parameters for the ionic species as well as the coefficients for equilibrium constants of the reactions were obtained simultaneously using the solubility data of the ternary systems. Following that, the solubilities of metal sulfates in the quaternary systems H2SO4–Fe2(SO4)3–MgSO4–H2O, H2SO4–Fe2(SO4)3–Al2(SO4)3–H2O at 250 °C and H2SO4–Al2(SO4)3–MgSO4–H2O at 230–270 °C were predicted. The calculated results were in excellent agreement with the experimental data.  相似文献   

9.
Single-phase Co-doped TiO2(CoxTi1-xO2) nanoparticles(NPs) synthesized via a simple sol-hydrothermal method were used as surface-enhanced Raman scattering(SERS) substrates. Interestingly, it was found that SERS signals were enhanced greatly compared to those of pure TiO2 nanoparticles when an amount of Co2+ ions were doped into the TiO2 lattice. Detailed results clearly show that Co element as Co2+ was incorporated into the TiO2 lattice and the defects were created due to the substitution of Co2+ ions for the Ti4+ ions. The Co2+ doping increases the defect concentration of CoxTi1-xO2 NPs. An amount of defects is beneficial to the charge-transfer so as to increase the SERS activities. A possible mechanism of charge-transfer from CoxTi1-xO2 NPs to molecules was then briefly discussed.  相似文献   

10.
The thermal decomposition of CaOsO3 by differential thermal analyses, thermogravimetry and X-ray powder diffraction has been studied. In nitrogen CaOsO3 decomposes at 880 ± 10°C into CaO, osmium metal and oxygen due to the reaction CaOsO3 → CaO + Os + O2. In static air the decomposition occurs in three stages: 2CaOsO3 + 1/2O2 → Ca2Os2O7 (in region 775–808°C), Ca2Os2O7 → Ca2Os2O6,5 + 1/4O2 (at a temperature interval of 850–1000°C) and in the third stage Ca2Os2O6,5 → 2CaO + OsO4 ÷ 1/4 O2 (at 1005 ± 5°C). The first intermediate Ca2Os2O7 is isostructural with orthorhombic Ca2Nb2O7 and its cell parameters are: a0 = 3.745 Å, b0 = 25.1 Å, c0 = 5.492 Å, Z = 4, space group Cmcm or Cmc2. Ca2Os2O7 exhibits metallic conductivity and its electrical resistivity is 4.6 × 10−2 ohm-cm at 296K.  相似文献   

11.
山西典型无烟煤灰流动性的调控   总被引:1,自引:0,他引:1  
为满足气化炉液态排渣的要求,考察和比较了CaO、MgO和Fe2O3三种助熔剂对山西典型无烟煤煤灰流动性(熔融性和黏温特性)的影响.研究发现,MgO对硅铝比在1.2~2.0的高硅铝煤灰的流动温度降低最有效,其次为CaO和Fe2O3,这是由于使用各种助熔剂时生成不同的高温稳定矿物组分造成的.针对三种助熔剂建立了流动温度和完全液相温度的关系式,并得到了CaO和Fe2O3含量与流动温度的关系:FT = 1 593-9.573 × wCaO (R2=0.9429) 和FT =1 576-8.330 6 × wFe2O3 (R2=0.955 9),可以用于指导助熔剂的添加.CaO无论从降低黏度数值或降低临界黏度温度都具有最好的效果.Ca2+、Mg2+、Fe2+的电负性差异和高温下的产物不同是三种助剂对黏度数值影响不同的根本原因; Mg2+、Fe2+具有较小的离子半径以及单质铁在高温下析出是导致临界黏度温度较高的原因.  相似文献   

12.
Bismuth as BiCl4 and BH4 ware successively retained in a column (150 mm × 4 mm, length × i.d.) packed with Amberlite IRA-410 (strong anion-exchange resin). This was followed by passage of an injected slug of hydrochloric acid resulting in bismuthine generation (BiH3). BiH3 was stripped from the eluent solution by the addition of a nitrogen flow and the bulk phases were separated in a gas–liquid separator. Finally, bismutine was atomized in a quartz tube for the subsequent detection of bismuth by atomic absorption spectrometry. Different halide complexes of bismuth (namely, BiBr4, BiI4 and BiCl4) were tested for its pre-concentration, being the chloride complexes which produced the best results. Therefore, a concentration of 0.3 mol l−1 of HCl was added to the samples and calibration solutions. A linear response was obtained between the detection limit (3σ) of 0.225 and 80 μg l−1. The R.S.D.% (n = 10) for a solution containing 50 μg l−1 of Bi was 0.85%. The tolerance of the system to interferences was evaluated by investigating the effect of the following ions: Cu2+, Co2+, Ni2+, Fe3+, Cd2+, Pb2+, Hg2+, Zn2+, and Mg2+. The most severe depression was caused by Hg2+, which at 60 mg l−1 caused a 5% depression on the signal. For the other cations, concentrations between 1000 and 10,000 mg l−1 could be tolerated. The system was applied to the determination of Bi in urine of patients under therapy with bismuth subcitrate. The recovery of spikes of 5 and 50 μg l−1 of Bi added to the samples prior to digestion with HNO3 and H2O2 was in satisfactory ranges from 95.0 to 101.0%. The concentrations of bismuth found in six selected samples using this procedure were in good agreement with those obtained by an alternative technique (ETAAS). Finally, the concentration of Bi determined in urine before and after 3 days of treatment were 1.94 ± 1.26 and 9.02 ± 5.82 μg l−1, respectively.  相似文献   

13.
A new family of heteropolytungstate complexes (NH4)21[Ln(H2O)5{Ni(H2O)}2As4W40O140xH2O(Ln=Y, Ce, Pr, Nd, Sm, Eu, Gd) were prepared by the reaction of Na27[NaAs4W40O140]·60H2O with NiCl2·6H2O and Ln(NO3)3·xH2O at pH≈4.5. The crystal structures of (NH4)21[Gd(H2O)5{Ni(H2O)}2As4W40O140]·51H2O was determined by X-ray diffraction analysis and element analysis. The compound crystallizes in the monoclinic space group P21/n with a=19.754(3), b=24.298(4), c=39.350(6) Å, β=100.612(3)°, V=18564(5) Å3, Z=2, R1(wR2)=0.0544(0.0691). The central site S1 and two opposite sites S2 of the big cyclic ligand [As4W40O140]28− are occupied by one Ln3+and two Ni2+, respectively, each site supply four Od coordinating to metal ion, another one water molecule and other five water molecules coordinate, respectively, to Ni2+and Ln3+. Polyanion [Ln(H2O)5{Ni(H2O)}2As4W40O140]21− has C2v symmetry. IR and UV–vis spectra of [NaAs4W40O140]27− of the title compounds are discussed.  相似文献   

14.
Solid acids – NiSO4/Al2O3, Fe2(SO4)3/Al2O3 and TiO2/SO42− – appeared to be effective catalysts for the acid catalyzed synthesis of methyl ester of trifluoropyruvic acid. They are active at 150–180 °C.  相似文献   

15.
The paper describes the determination of the molybdenum content in white wines based on its catalytical action on the kalium iodide oxidation by hydrogen peroxide in acid medium.

The optimum reaction conditions (the catalyst, KI and H2O2 concentrations, the pH value, the order of the reagent additions, the temperature) have been found by studying the effect of the reaction variables.

The influence of some metallic ions (Ca2+, Mg2+, Zn2+, Cd2+, Fe2+ and Fe3+) and complexing anions (F, C2O2−4, EDTA4−) on the catalyzed reaction rate was elucidated.

The molybdenum concentration was estimated by the tangent, fixed-time and fixed-absorbance method. The obtained average values for molybdenum content in white wines are within the 1.77×10−7–1.83×10−7 mol l−1 range.  相似文献   


16.
The temperature dependence of the rate constants, for the reactions of hydrated electrons with H atoms, OH radicals and H2O2 has been determined. The reaction with H atoms, studied in the temperature range 20–250°C gives k(20°C) = 2.4 × 1010M-1s1 and the activation energy EA = 14.0 kJ mol-1 (3.3 kcal mol-1). For reaction with OH radicals the corresponding values are, k(20°C) = 3.1 × 1010M-1s-1 and EA = 14.7 kJ mol-1 (3.5 kcal mol-1) determined in the temperature range 5–175°C. For reaction with H2O2 the values are, k(20°C) = 1.2 × 1010M-1s-1 and EA = 15.6 kJ mol-1 (3.7 kcal mol-1) measured from 5–150°C. Thus, the activation energy for all three fast reactions is close to that expected for diffusion controlled reactions. As phosphates were used as buffer system, the rate constant and activation energy for the reaction of hydrated electron with H2PO4- was determined to k(20°C) = 1.5 × 107M-1s-1 and EA = 7.4 kJ mol-1 (1.8 kcal mol-1) in the temperature range 20–200°C.  相似文献   

17.
A 1D supramolecular compound [dmbbbi](1) and a 2D cobalt coordination polymer [Co(dmbbbi)(ox)]2(2) [dmbbbi=1,1-(1,4-butanediyl)bis(5,6-dimethylbenzimidazole), ox=oxalate], C48H52Co2N8O8, were obtained under hydrothermal conditions by tuning the molar ratio of the reactants. The crystal structure analysis reveals that in compound 1, the adjacent dmbbbi molecules connect with each other via hydrogen bonds to form a 1D supramolecular chain. In compound 2, two crystallographically independent Co2+ ions show the same six-coordination mode. Each Co2+ ion is coordinated by four oxygen atoms from two ox anions and two nitrogen atoms from two cis-dmbbbi ligands. The adjacent Co2+ ions are bridged by ox anions to generate an infinite 1D zigzag chain, which is extended by pairs of dmbbbi ligands to form a 2D honeycomb-like (6,3) network. Moreover, the thermal stability and the electrochemical property of compound 2 were studied.  相似文献   

18.
Shamsipur M  Esmaeili A  Amini MK 《Talanta》1989,36(12):1300-1302
The complexation reactions between murexide and Co2+, Ni2+ and Cu2+ in C2H5OH-H2O mixtures have been investigated spectrophotometrically. The formation constants of the 1:1 complexes formed increase in the order Co2+ < Ni2+ < Cu2+ for all solvent mixtures studied, and log Kf is a linear function of the mole fraction of ethanol. The heat of complexation was determined calorimetrically for the nickel and copper complexes. The values of ΔH° and ΔS° are solvent-dependent, and all three complexes have negative ΔH° and positive ΔS° values.  相似文献   

19.
As a part of the European EUROCORE and GRIP (Greenland Ice Core Project) operations aimed at recovering deep ice cores at Summit (Central Greenland), we have for the first time successfully performed ion chromatography measurements in the field and investigated in detail the soluble impurities, including Na+, NH+4, K+, Mg2+, Ca2+, F, CH3COO, CH2 OHCOO, HCOO, CH3SO3, Cl, NO2, SO42− and C2O42−, trapped in ice deposited over some 200 000 years in Greenland.  相似文献   

20.
Zuberbühler AD  Kaden TA 《Talanta》1979,26(12):1111-1118
A fully automatic system for combined spectrophotometric and pH titrations was described in Part I. Its performance in the titration of weak acids and metal complexes is discussed, along with a computer program for numerical treatment of the data, based on Marquardt's modification of the Newton—Gauss non-linear least-squares method. The deprotonation of p-nitrophenol at concentrations of 4 × 10−5 and 4 × 10−6M was studied in order to test the sensitivity. Results identical within the reproducibility of the pH-meter were obtained: pKH = 7.00 ± 0.01 and 7.02 ± 0.01, respectively. Three complexation reactions were studied: (1) the interaction of SCN with the Co2+ complex of 1,4,8,11-tetramethyl-1,4,8,11-tetra-azacyclotetradecane (TMC); five independent experiments gave pK [CoTMC (SCN)+ CoTMC2+ + SCN] = 3.099 ± 0.003: (2) the deprotonation of the Cu2+ complex of 3,7-diazanonanediamide (DANA); five experiments gave pK (CuDANA2+ CuDANAH+−1 + H+) = 7.14 ± 0.01 and pK (CuDANAH+−1 CuDANAH−2 + H+) = 8.38 ± 0.01: (3) for the reaction of Cu2+ with 1,3,7-triazacyclodecane (L), data from different ligand: metal ratios had to be combined to obtain pK (CuL2+ Cu2+ + L) = 16.19 ± 0.01, pK (CuL2+2 CuL2+ + L) = 10.30 ± 0.01, and pK (Cu2L2 (OH)2+2 2 CuL2+ + 2 OH) = 14.58 ± 0.03. Titration curves with a total change in absorbance of as little as 0.03 units could be analysed satisfactorily, extending considerably the useful range of concentrations for spectrophotometric titrations. In combined spectrophotometric/pH titrations the accuracy of the glass electrode is normally the limiting factor. Other equilibrium constants can easily be reproduced with standard errors of less than 0.01 log unit.  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