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 828 毫秒
1.

Abstract  

Near-UV irradiation of solutions of (Bu4N)AuCl4 in aerated ethanol-stabilized chloroform causes the continuous decomposition of chloroform, as evidenced by the production of many equivalents of HCl and peroxides. At the outset of irradiation, most of the AuCl4 is reduced to AuCl2 , but the reduction stops and is reversed. The same experiments done in ethanol-free chloroform cause chloroform decomposition only until the irreversible reduction of the gold is complete. In deoxygenated ethanol-free chloroform, irreversible reduction to AuCl2 is accompanied by the formation of HCl and CCl4, while the main decomposition products in deoxygenated ethanol-stabilized chloroform are HCl and C2Cl6. It is proposed that, in ethanol-free chloroform, photoreduction of AuCl4 begins with the concerted elimination of HCl from an association complex of CHCl3 with AuCl4 , and that ethanol suppresses { \textCHCl3 ·\textAuCl4 - } \{ {\text{CHCl}}_{3} \cdot {\text{AuCl}}_{4}^{ - } \} complex formation, leaving a slower radical process to carry out the photoreduction of AuCl4 in ethanol-stabilized chloroform. In the presence of oxygen, the radical process causes a build-up of CCl3OOH, which reoxidizes AuCl2 to AuCl4 and allows the photodecomposition of CHCl3 to continue indefinitely.  相似文献   

2.
Broadband (λ > 320 nm) irradiation of chloroform solutions of either [Ru(bpy)2Cl2] or [Ru(bpy)2Cl2]Cl exposed to air led to a photostationary state, in which [Ru(bpy)2Cl2]+ predominated, and to the continuous decomposition of CHCl3, as evidenced by the accumulation of HCl, hydroperoxides (CCl3OOH and CHCl2OOH), and tetra-, penta-, and hexachloroethane. The addition of Cl? increased the rate of photodecomposition, while the replacement of Cl? by F? greatly decreased the rate. The observations are consistent with a photocatalytic cycle in which [Ru(bpy)2Cl2]+ is photochemically reduced to [Ru(bpy)2Cl2], which is thermally reoxidized by CCl3OO or CCl3OOH. In the absence of air a much slower photodecomposition reaction takes place leading to continuously increasing concentrations of chloroethanes. The data are consistent with a catalytic cycle in which [Ru(bpy)2Cl2]+ is photoreduced, as in aerated solutions, while [Ru(bpy)2Cl2] is photooxidized with chloroform as the substrate.  相似文献   

3.
Thermal behaviour of nickel amine complexes containing SO4 2−, NO3 , Cl and Br as counter ions and ammonia and ethylenediamine as ligands have been investigated using simultaneous TG/DTA coupled with mass spectroscopy (TG/DTA–MS). Evolved gas analyses detected various transient intermediates during thermal decomposition. The nickel ammonium sulphate complex produces NH, N, S, O and N2 species. The nickel ammonium nitrate complex generated fragments like N, N2, NO, O2, N2O, NH2 and NH. The halide complexes produce NH2, NH, N2 and H2 species during decomposition. The ligand ethylenediamine is fragmented as N2/C2H4, NH3 and H2. The residue hexaamminenickel(II) sulphate produces NiO with crystallite size 50 nm. Hexaammine and tris(ethylenediamine)nickel(II) nitrate produce NiO in the range 25.5 nm and 23 nm, respectively. The halide complexes produce nano sized metallic nickel (20 nm) as the residue. Among the complexes studied, the nitrate containing complexes undergo simultaneous oxidation and reduction.  相似文献   

4.
For the system liquid anion-exchanger—Cr(III)−NCS, an investigation has been made of the dependence of the percentage extraction of Cr(III) on parameters such as standing time of the Cr(III)−NCS solution, temperature, pH and type of exchanger. Quantitative extraction of e.g. 4·10−4 M Cr(III) by 0.1M Aliquat in CCl4 is easily achieved at room temperature, using 4.75M KNCS−0.05N HCl as aqueous phase. At high Cr(III) concentrations, the complex anion present in the organic phase is Cr(NCS) 6 3− ; when working with dilute metal ion solutions, the species extracted is Cr(NCS)4 (H2O) 2 . Separations of mixtures containing 10−2−10−4 M Co(II), Ni(II) and Cr(III) have successfully been accomplished.  相似文献   

5.
Summary This paper is devoted to the characterization of FeCl3 solutions in isopropanol containing water. For this goal optical absorption and e.p.r. techniques have been used in conjunction with magnetization and M?ssbauer data reported very recently. It is shown that in a 10−2 M solution of FeCl3 containing 0.4 M of water the main iron(III) species present in the solution are [FeCl4] (55%) and [FeCl2(H2O)4]+ (20%) while the remainding 25% is due to dihydroxo dimers, . When the water concentration increases the [FeCl4] anions are progressively destroyed, the main iron(III) species present in the solution being the dihydroxodimers and [FeCl2(H2O)4]+. The variation of the concentration of the three species mentioned with the water content and FeCl3 concentration is presented in this paper.  相似文献   

6.
Raman spectra of aqueous 1 M chloroferrate(III) solutions indicate the presence of FeCl3 and FeCl4- in proportion to the hydrochloric acid concentration, but FeCl4- is only present in solutions of acidity higher than 5 M HCl. The liquid-liquid extraction of the aqueous chloroferrate phases ( 1 M HCl) by trilaurylamine hydrochloride in cyclohexane, shows that iron(III) is extracted only in its FeCl4- form, even though this species does not appear in the aqueous solutions. A quantitative spectroscopic study of the organic phases by means of the v1 line of FeCl4- at 332 cm-1 leads to the following conclusions: (a) the plot of the scattering coefficient of the 332 cm-1 line versus iron(III) concentration shows the presence of two complex species in the organic phase; (b) the distribution curves of the complex species could be calculated. On the basis of the Raman results, an ion-pair type compound is proposed for the two complexes.  相似文献   

7.

Abstract  

This article proposes a simple and fast method of In(III) determination in the presence of Cd(II) and Pb(II). The catalytic activity of N-methylthiourea was used in the In(III) electroreduction, which also had a slight effect on the electroreduction process of Cd(II) and Pb(II). By applying square wave voltammetry it was possible to determine 3 × 10−7 mol dm−3 In(III) in the presence of 5 × 10−5 mol dm−3 Cd(II) and 1 × 10−4 mol dm−3 Pb(II) in 5 mol dm−3 NaClO4 at pH 2. The calibration curve for In(III) was linear from 3 × 10−7 to 5 × 10−4 mol dm−3. The relative standard deviation for In(III) determination was about 3.0%.  相似文献   

8.
Palladium(II) coordination compounds of general formula trans-[PdX2(isn)2], X = Cl (1), N3 (2), SCN (3), NCO (4), isn = isonicotinamide; were synthesized and characterized in solid state by elemental analysis, infrared spectroscopy, and simultaneous TG–DTA. TG experiments reveal that the compounds 14 undergo thermal decomposition in three or four stages, yielding Pd0 as final residue, according to calculus and identification by X-ray powder diffraction.  相似文献   

9.
Reactions between tungsten halides are discussed along the series of compounds WCl6 → WCl4 → W6Cl12 ↔ W6Cl18 → [W6CCl18]n− ← [W3Cl13]u−, focusing on the two closely related tungsten chloride compounds whose structures compromise the well-known octahedro W6Cl18 cluster and the carbon-centered triprismo W6CCl18 cluster. Both clusters can be regarded as being built by merging two trigonal [W3Cl13]u− units in different ways. Syntheses, structural transformation reactions, and concepts regarding electronic structures are reported.  相似文献   

10.
The keto-enol equilibrium of 2 × 10−3 M solutions of (acetoacetyl)ferrocene and 1,1′-bis(acetoacetyl)ferrocene was studied by 1H NMR spectroscopy in a series of aprotic solvents such as DMSO-d 6, (CD3)2CO, CDCl3, CD2Cl2, CCl4, and C6D6 at a temperature of 20°C. It was established that the calculated enolization constants increase with a decrease in the polarity of solvent molecules. The results complement and combine known empirical rules about the effect of the medium on keto-enol equilibria.  相似文献   

11.
The kinetics of the intra-molecular electron transfer of an adduct of l-ascorbic acid and the [Fe3IIIO(CH3COO)6(H2O)3]+ cation in aqueous acetate buffer was studied spectrophotometrically, over the ranges 2.55 ≤ pH ≤ 3.74, 20.0 ≤ θ ≤ 35.0 °C, at an ionic strength of 0.50 and 1.0 mol dm−3 (NaClO4). The reaction of l-ascorbic acid and the complex cation involves the rapid formation of an adduct species followed by a slower reduction in the iron centres through consecutive one-electron transfer processes. The final product of the reaction is aqueous iron(II) in acetate buffer. The proposed mechanism involves the triaqua and diaqua-hydroxo species of the complex cation, both of which form adducts with l-ascorbic acid. At 25 °C, the equilibrium constant for the adduct formation was found to be 86 ± 15 and 5.8 ± 0.2 dm3 mol−1 for the triaqua and diaqua-hydroxo species, respectively. The kinetic parameters derived from the rate expression have been found to be: k 0 = (1.12 ± 0.02) × 10−2 s−1 for the combined spontaneous decomposition and k 1 = (4.47 ± 0.06) × 10−2 s−1H 1 = 51.0 ± 2.3 kJ mol−1, ΔS 1 = −100 ± 8 J K−1 mol−1), k 2 = (4.79 ± 0.38) × 10−1 s−1H 2 = 76.5 ± 0.8 kJ mol−1, ΔS 2 = 6 ± 3 J K−1 mol−1) for the triaqua and diaqua-hydoxo species, respectively.  相似文献   

12.
A series of trinuclear Cu(II) complexes have been prepared by Schiff base condensation of 1,8-[bis(3-formyl-2-hydroxy-5-methyl)benzyl]-l,4,8,11-tetraazacyclotetradecane and 1,8-[bis(3-formyl-2-hydroxy-5-bromo)benzyl]-l,4,8,11-tetraazacyclotetradecane with aromatic and aliphatic diamines, Cu(II) perchlorate and triethylamine. The complexes were characterized by elemental and spectroscopic analysis. Electrochemical studies of the complexes in DMF solution show three irreversible one-electron reduction processes around Epc 1 = −0.73 to −0.98 V, Epc 2 = −0.91 to −1.20 V and Epc 3 = −1.21 to −1.33 V. ESR spectra and magnetic moments of the trinuclear Cu(II) complexes show the presence of antiferromagnetic coupling. The rate constants for hydrolysis of 4-nitrophenylphosphate by the Cu(II) complexes are in the range of 3.33 × 10−2 to 7.58 × 10−2 min−1. The rate constants for the catecholase activity of the complexes fall in the range of 2.67 × 10−2 to 7.56 × 10−2 min−1. All the complexes were screened for antifungal and antibacterial activity.  相似文献   

13.
The photocatalytic transformations of carbon tetrachloride and aliphatic primary alcohols in the presence of iron trichloride and a molar ratio of components FeCl3: CCl4: ROH = 1: 300: 2550 were studied. CCl4 is transformed into chloroform and hexachloroethane after exposure to a mercury lamp (250 W) to the FeCl3–CCl4–ROH system at 20°C, whereas the primary ROH alcohols are selectively oxidized into acetals (1,1-dialkoxyalkanes). The maximum conversion of CCl4 reaches 80%. The kinetics and mechanism of the photocatalytic conversion of the FeCl3–CCl4–ROH system are considered.  相似文献   

14.
Four manganese(II) coordination compounds with bis(1-methylimidazol-2-yl)ketone (BIK) of general formula Mn(BIK)2X2 (X = Cl, Br, NO3, ClO4) were synthesized and characterized by elemental analysis, by UV–vis, and FTIR spectroscopies to be compared with the literature data. Following our previous thermoanalytical studies on imidazole-substituted coordination compounds, the thermal behavior of the synthesized Mn(II) complexes was investigated using TG and DTG techniques: the thermal profile is characterized by three substantial consecutive releasing steps for all the three complexes and the releasing supposed behavior is confirmed by EGA analysis performed by coupling the TG analyzer to an MS spectrometer. In particular, the first step is ascribed to the release of the two anions, followed by the loss of four methyl groups (side chains of the ligand) and two bridge-carbonyl groups. The residual tetra-imidazole manganese compound decomposes in a final step to give MnO as the final residue. Both the initial decomposition temperatures and the kinetic rate constants associated to the first decomposition step indicated a higher stability of the Mn(BIK)2Cl2 complex, the bromide complex being very close to the chloride one (first-step thermal stability: ClO4 <NO3 ≤Br <Cl). Finally, the three-dimensional diffusion reaction model (D3) was selected to describe the first decomposition step for all the four complexes examined.  相似文献   

15.
It was found that in a wide range of pH in the presence of boric acid the oxidation of diethyl sulfide (Et2S) with hydrogen peroxide in an i-PrOH–H2O medium occurs with the participation of H2O2, HOO, monoperoxo-(B(OH)3OOH), and diperoxoborates (B(OH)2(OOH)2). The stability constants of peroxoborates and the rate constants for the reactions of H2O2, HOO, B(OH)3(OOH)2, and B(OH)2(OOH)2 with Et2S under these conditions were determined.  相似文献   

16.
3,3-Dinitroazetidinium (DNAZ) salt of perchloric acid (DNAZ·HClO4) was prepared, it was characterized by the elemental analysis, IR, NMR, and a X-ray diffractometer. The thermal behavior and decomposition reaction kinetics of DNAZ·HClO4 were investigated under a non-isothermal condition by DSC and TG/DTG techniques. The results show that the thermal decomposition process of DNAZ·HClO4 has two mass loss stages. The kinetic model function in differential form, the value of apparent activation energy (E a) and pre-exponential factor (A) of the exothermic decomposition reaction of DNAZ·HClO4 are f(α) = (1 − α)−1/2, 156.47 kJ mol−1, and 1015.12 s−1, respectively. The critical temperature of thermal explosion is 188.5 °C. The values of ΔS , ΔH , and ΔG of this reaction are 42.26 J mol−1 K−1, 154.44 kJ mol−1, and 135.42 kJ mol−1, respectively. The specific heat capacity of DNAZ·HClO4 was determined with a continuous C p mode of microcalorimeter. Using the relationship between C p and T and the thermal decomposition parameters, the time of the thermal decomposition from initiation to thermal explosion (adiabatic time-to-explosion) was evaluated as 14.2 s.  相似文献   

17.
A brown and transparent ionic liquid (IL), [C4mim][FeCl4], was prepared by mixing anhydrous FeCl3 with 1-butyl-3-methylimidazolium chloride ([C4mim][Cl]), with molar ratio 1/1 under stirring in a glove box filled with dry argon. The molar enthalpies of solution, Δs H m, of [C4mim][FeCl4], in water with various molalities were determined by a solution-reaction isoperibol calorimeter at 298.15 K. Considering the hydrolyzation of anion [FeCl4] in dissolution process of the IL, a new method of determining the standard molar enthalpy of solution, Δs H m0, was put forward on the bases of Pitzer solution theory of mixed electrolytes. The values of Δs H m0 and the sum of Pitzer parameters: and were obtained, respectively. In terms of thermodynamic cycle and the lattice energy of IL calculated by Glasser’s lattice energy theory of ILs, the dissociation enthalpy of anion [FeCl4], ΔH dis≈5650 kJ mol−1, for the reaction: [FeCl4](g)→Fe3+(g)+4Cl(g), was estimated. It is shown that large hydration enthalpies of ions have been compensated by large the dissociation enthalpy of [FeCl4] anion, Δd H m, in dissolution process of the IL.  相似文献   

18.
The synthesis, spectroscopic characterization, and thermal analysis of the compounds [Pd(X)2(mtu)(PPh3)] (X = Cl (1), SCN (2); mtu = N-methylthiourea; PPh3 = triphenylphosphine) and [Pd(X)2(phtu)(PPh3)] (X = Cl (3), SCN (4); phtu = N-phenylthiourea) are described. The thermal decomposition of the compounds occurs in two, three, or four stages and the final decomposition products were identified as Pd0 by X-ray powder diffraction. The thermal stability order of the complexes is 4 > 3>2 > 1.  相似文献   

19.
Under similar hydrothermal synthetic conditions, the reactions of Fe(NO3)3/FeCl2, CuCl2, NiCl2, and CdCl2 with phenanthroline (phen) and 3,3′,4,4′-biphenyltetracarboxylic acid (H4BPTC) afforded complexes [Fe(phen)3](H3BPTC)2 (1), [Cu(phen)(BPTC)0.5 · H2O] · H2O (2), [Ni3(phen)3(BPTC)1.5(H2O)5] · 4H2O (3) and [Cd(phen)(BPTC)0.5] · H2O (4). The short Fe–N distance in the monomeric Fe(phen)3(H3 BPTC)2 (1) shows that the Fe(II) is in a low-spin state. H3 BPTC4− acts as a counter-ion in this complex. In [Cu(phen)(BPTC)0.5 · H2O] · H2O (2), the central Cu(II) is five-coordinated in a square-pyramidal geometry. The ligand BPTC4− is centrosymmetric and the four deprotonated carboxylic groups of BPTC4− are coordinated to four different copper ions to form a 1D ladder complex indicating a comparatively strong coordination. In [Ni3(phen)3(BPTC)1.5(H2O)5] · 4H2O (3), all nickel(II) atoms are in an octahedral coordination environment. There are two different BPTC4− ligands; one is centrosymmetric and the other is asymmetric. Metal ions are linked through fully deprotonated BPTC4− ligands to form a 2D metal-organic sheet. [Cd(phen)(BPTC)0.5] · H2O (4) has a 3D metal-organic framework. TG, IR, and fluorescence data for the complexes are presented.  相似文献   

20.
(trienH2)[CoCl4], which contains tetrahedral chlorocobaltate(II) anions, decomposes under argon in two stages via a stepwise deprotonation of the cation. The decomposition starts at 310°C with the liberation of HCl, followed at 400°C by the simultaneous release of a further mole of HCl and triene and/or its cracking products. The second decomposition stage is strongly influenced by the atmosphere. In the lower temperature region (<400°C), increasing oxygen contents of the carrier gas lead to decreasing mass losses. Therefore, the solid residues contain various amounts of C,N-containing products as well as coke. The thermal decomposition of the iron(III) compound, which contains μ-oxalato-bridged FeCl4 units, begins with the dehydratation followed by the decay of the complex anion to produce CO, CO2, and HCl. Instead of a binuclear, monooxobridged chloroferrate(III) complex, a [FeCl4]? — containing compound is proposed as one of the final products. The third decomposition stage, partially overlaying the preceding one, is the degradation of the organic cation as found for the cobalt compound. The results of thein situ-TA-MS measurements are compared with those obtained from usual TA techniques as well as from the residue characterization by X-ray diffraction, Raman spectroscopy, and chemical analysis.  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