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 15 毫秒
1.
The formation of palladium(II) complexes with aliphatic amines and their oxidation by chloramine‐T in perchloric acid medium has been studied. The spectrophotometric studies showed the formation of 1:1 and 1:2 complexes between palladium(II) and amine in absence of HClO4. An increase in [HClO4] in reaction mixture suppresses the complex formation and in presence of [HClO4] ~10?3 mol dm?3 only a 1:1 complex between palladium(II) and amine has been observed. The effect of Cl? on the complex formation has also been studied. Palladium(II)‐catalyzed oxidation of these amines by chloramine‐T showed a first‐order dependence of rate with respect to each—oxidant, substrate, catalyst, and H+. The mechanism consistent with kinetic data for the oxidation process has been proposed in absence as well as in presence of initial [Cl?]. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 603–612, 2002  相似文献   

2.
The formation kinetics of ferroin is studied under varied acid conditions at 25°C and fixed ionic strength (0.48 mol dm?3) under pseudo‐first‐order conditions with respect to Fe2+ by using the stopped‐flow technique. The reaction followed is first and third order with respect to Fe2+ and 1,10‐phenanthroline (phen)T, respectively. Increasing the acid concentration retarded the reaction, and the reaction rate showed a positive salt effect. The rate‐limiting step involved the complexation of the phen or protonated phen with [Fe(phen)2]2+ complex ion, leading to formation of [Fe(phen)3]2+ ion. The observed retardation of the reaction rate with increasing [H+]0 is due to the increased [phenH+]eq and low reactivity of phenH+ with [Fe(phen)2]2+ complex ion. Simulated curves for the acid variation experiments agreed well with the corresponding experimental curves and the estimated rate coefficients supporting the proposed mechanism. Relatively low energy of activation (26 kJ mol?1) and high negative entropy of activation (?159.8 J K?1 mol?1) agree with the proposed mechanism and the formation of compact octahedral complex ion. © 2008 Wiley Periodicals, Inc. 40: 515–523, 2008  相似文献   

3.
Densities and ultrasonic velocities were measured at 25°C for aqueous solutions of bipyridine and phenanthroline complexes [M(bpy)3]Cl2 and [M(phen)3]Cl2 (M=Fe, Co, Ni, and Cu, bpy=2,2-bipyridine, and phen=1,10-phenanthroline), and chlorides of these metals. The partial molar volumes V 2 o and partial molar adiabatic compressibilities K s o were calculated. For the complex ions, [M(bpy)3]2+ and [M(phen)3]2+, electrostatic interactions with the solvent are not nearly as important as effects due to the hydrophobic ligands bpy and phen. The relationship between V 2 o and K s o of the complex ions and common metal ions are examined.  相似文献   

4.
The reaction of [SnMe2Cl2] with the bidentate ligand 4,7‐phenanthroline (4,7‐phen) resulted in the formation of [SnMe2Cl2 (4,7‐phen)]n ( 1a ) which is probably a polymeric chain in solution. On the other hand, the reaction of [SnEt2Cl2] with 4,7‐phen afforded the complex [Sn2Et4Cl41‐N‐4,7‐phen)2(μ‐κ2‐N,N‐4,7‐phen)] ( 1b ) which dissociates in dimethylsulfoxide solution. The reaction of [SnR2Cl2] (R = Me, Et) with 2,2′‐biquinoline (biq) yielded the complexes [SnMe2Cl22‐N,N‐biq)] ( 2a ) and [SnEt2Cl21‐N‐biq)2] ( 2b ) in the solid state. Moreover, the reaction of [SnR2Cl2] (R = Me, Et) with the tridentate ligand 4′‐(2‐furyl)‐2,2′:6′,2″‐terpyridine (ftpy) resulted in the formation of ionic penta‐ and hexa‐coordinated tin complexes [SnMe2Cl (ftpy)][SnMe2Cl3] ( 3a ) and [SnEt2Cl (ftpy)]Cl ( 3b ). The reaction of [SnMe2 (NCS)2] with ftpy afforded the hepta‐coordinated complex [SnMe2 (NCS)2(ftpy)] ( 4a ). The products were fully characterized using elemental analysis, and infrared, UV–visible, multinuclear (1H, 13C, 119Sn) NMR, DEPT‐135°, HH‐COSY and HSQC NMR spectroscopies. The crystal structure of complex 3a reveals that it contains the simultaneous presence of penta‐ and hexa‐coordinated tin (IV) atoms. Notably, the crystal structure of complex 4a shows that tin (IV) is hepta‐coordinated in a pentagonal bipyramidal geometry SnC2N5 by three nitrogen atoms of ftpy, two nitrogen atoms of NCS? and two Me groups with trans‐[SnMe2] configuration. These data indicate the influence of halide or pseudo‐halide group on the coordination number and geometry of tin. Hirshfeld surface analysis and two‐dimensional fingerprint plots were calculated for 3a and 4a which show the π–π interaction between molecules in the solid is relatively weak.  相似文献   

5.
The kinetics of the oxidation of tris(2,2′-bipyridyl)iron(II) and tris(1,10-phenanthroline)iron(II) complexes ([Fe(LL)3]2+, LL = bipy, phen) by nitropentacyanocobaltate(III) complex [Co(CN)5NO2]3? was investigated in acidic aqueous solutions at ionic strength of I = 0.1 mol dm?3 (HCl/NaCl). The reactions were carried out at fixed acid concentration ([H+] = 0.01 mol dm?3) and the temperature maintained at 35.0 ± 0.1 °C. Spectroscopic evidence is presented for the protonated oxidant. Protonation constants of 360.43 and 563.82 dm3 mol?1 were obtained for the monoprotonated and diprotonated Co(III) complexes respectively. Electron transfer rates were generally faster for [Fe(bipy)3]2+ than [Fe(phen)3]2+. The redox complexes formed ion-pairs with the oxidant with increasing concentration of the oxidant over that of the reductant. Ion-pair constants for these reaction were 160.31 and 131.9 dm3 mol?1 for [Fe(bipy)3]2+ and [Fe(phen)3]2+, respectively. The activation parameters measured for these systems have values as follows: ?H (kJ K?1 mol?1) = +113.4 ± 0.4 and +119 ± 0.3; ?S (J K?1) = +107.6 ± 1.3 and 125.0 ± 1.6; ?G (kJ K?1) = +81 ± 0.4 and +82.4 ± 0.4; and E a (kJ mol?1) = 115.9 ± 0.5 and 122.3 ± 0.6 for LL = bipy and phen, respectively. Effect of added anions (Cl?, $ {\text{SO}}_{4}^{2 - } $ and $ {\text{ClO}}_{4}^{ - } $ ) on the systems showed decrease in the electron transfer rate constant. An outer-sphere mechanism is proposed for the reaction.  相似文献   

6.
The kinetics of the reduction of octacyanomolybdate(V) and octacyanotungstate(V) by sulphite ions has been studied over a wide pH range. The reaction is catalysed by alkali metal ions. The rate law is found to be of the form:
The third order rate constants at [OH?] = 0.05 mol dm?3 for the reduction of Mo(CN)83? and W(CN)3?8 were determined as 6.2 x 103dm6mol?2 s?1 and 22.3 dm6mol?2s?1 respectively at 298 K for A+ = Na+ while Ka for the hydrogen sulphite ion was determined as 2.4 x 10?8 mol dm?3. It was established that the reaction proceeds via an outer-sphere mechanism. An explanation for the alkali metal ion catalysis is proposed.  相似文献   

7.
Abstract

Ruthenium (II) complexes of the type RuL(CO)2Cl2, [RuL(CO)2L? 2]2+ and [RuL(CO)2Cl L′]+ [L = bipyridine (bpy), phenanthroline (phen), biquinoline (biq) and L′ = pyridine (py), 4-chloropyridine (Cl-py), 4-methoxypyridine (MeO-py)] were synthesized from [Ru(CO)2Cl2]n and L, to produce the intermediate RuL(CO)2Cl2 followed by hydrolysis and reaction with L′. The catalytic activity of these complexes in epoxidation of olefins with iodosylbenzene under ambient conditions was investigated. A possible mechanism of these reactions, explaining the effects of the ligands on the reaction was explored. At least one carbonyl ligand remained bound to the metal through the reaction. The formation of an oxo intermediate was inferred from spectroscopic detection of bridged oxygen Ru—O—Ru and Ru=O species.  相似文献   

8.
At bromide concentrations higher than 0.1 M, a second term must be added to the classical rate law of the bromate–bromide reaction that becomes ?d[BrO3?]/dt = [BrO3?][H+]2(k1[Br?] + k2[Br?]2). In perchloric solutions at 25°C, k1 = 2.18 dm3 mol?3 s?1 and k2 = 0.65 dm4 mol?4 s?1 at 1 M ionic strength and k1 = 2.60 dm3 mol3 s?1and k2 = 1.05 dm4 mol?4 s?1 at 2 M ionic strength. A mechanism explaining this rate law, with Br2O2 as key intermediate species, is proposed. Errors that may occur when using the Guggenheim method are discussed. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 39: 17–21, 2007  相似文献   

9.
于杰辉  施展等 《中国化学》2002,20(6):560-563
The title compound Cu2Cl2phen (phen = 1,10‐phenanthroline, C12H8N2) 1 was synthesized from CuCl2·2H2O, CuCl and phen by hydrothermal method and its structure was determined by single crystal X‐ray analysis. With phen, CuG forms one‐dimensional chains, which comprise two zigzag chains based on fused Cu‐X units and connected via covalent bonds. The compound contains two crystallographically unique monovalent copper ions, Cu(1) and Cu(2). The Cu(1) atom in the tetrahedral site, is coordinated to two bridging Cl? and two N atoms in phen. The Cu(2) atom with a slightly distorted triangular planar geometry, is coordinated to three Cl?. The compound 1 was crystallized in monoclinic, space group P21/n with a = 0.37338(4), b = 1.9510(2), c = 1.68008(19) nm, β = 95.605 (3)°, R = 0.0458, and was characterized by elemental analysis, IR spectrum and TGA analysis.  相似文献   

10.
The crystal structures, magnetic properties, and catalase-like activities of assymmetric dinuclear manganese(III, III) complexes, [Mn2III, III(spa)2(μ-Me3CCO2)(Me3CCO2)(CH3OH)] ( 1 ) and [Mn2III, III(vpa)2(μ-Me3CCO2)(Me3CCO2)(CH3OH)] ( 2 ), (H2spa = 3-salicyclideneamino-1-propanol, H2vpa = O-vanillin), were reported. The crystal structures of complexes 1 and 2 consist of the same discrete asymmetric coordination environment of dinuclear clusters, where the two manganese atoms are bridged by two alkoxo oxygens of the spa or vpa ligands and one bidentate carboxylate ion, whereas an additional oxygen atom of monodentate carboxylate coordinated to the first metal ion, and the second metal ion was coordinated by one oxygen atom of the solvent CH3OH. Magnetic investigations (2–300 K) reveal an intramolecular antiferromagnetic spin exchange interaction with axial-field splittings: J = ?12.3 cm?1 (D = ?0.10 cm?1) and J = ?13.3 cm?1 (D = ?0.15 cm?1) for complexes 1 and 2 , respectively. The complexes should show catalase-like activity for H2O2 disproportionation in CH3OH solvent at 25° with rate constants of k = 6.35 dm3moI?1s?1 and 6.20 dm3mol?1s?1 for complexes 1 and 2 , respectively.  相似文献   

11.
The spectrocoulometric technique reported earlier is applied to verify the mechanism and to evaluate the contributions kBi of the individual bases to the total rate constant k of the hydrolysis of the tris (1,10-phenanthroline) iron(III) complex, Fe (phen)3+3. Both normal and “open-circuit” spectrocoulometric experiments are used. Partial rate constants for four bases in the acetate-buffered solutions are kH2O=(3.4±1.2) × 10?4s?1 (kH2O includes the H2O concentration), kOH=(1.20±0.06)×107 mol?1dm3s?1, kphen=(1.4±0.2) mol?1dm3s?1, kAc=(3.8±0.3)×10?2 mol?1dm3s?1, at 25°C and ionic strength 0.5 mol dm?3. The Fe(phen)3+3 hydrolysis, with (phen)2 (H2O) Fe-O-Fe (H2O) (phen)4+2 formation, is first order with respect to Fe (phen)3+3 and the bases present in the solution. The rate-determining step in the hydrolysis is the entry of a base to the coordinating sphere of the complex, as in the hydrolysis of the analogous 2,2′-bipyridyl complex.  相似文献   

12.
The title compound, [Cu{N(CN)2}(C12H8N2)2]BF4, was prepared as part of our study of the shape of coordination polyhedra in five‐coordinated copper(II) complexes. Single‐crystal X‐ray analysis reveals that the structure consists of [Cu{N(CN)2}(phen)2]+ cations (phen is 1,10‐phenanthroline) and BF4 anions. The Cu centre is five‐coordinated in a distorted trigonal bipyramidal manner by four N atoms of two phen ligands and one N atom of a dicyanamide anion, which is coordinated in the equatorial plane at a distance of 1.996 (2) Å. The two axial Cu—Nphen distances have similar values [average 1.994 (6) Å] and are shorter than the two equatorial Cu—Nphen bonds [average 2.09 (6) Å]. This work demonstrates the effect of ligand rigidity on the shape of coordination polyhedra in five‐coordinated copper(II) complexes.  相似文献   

13.
In the cationic complex present in the title compound, chloro­[2‐(4‐imidazolyl‐κN1)­ethyl­amine‐κN](1,10‐phenanthroline‐κ2N,N′)copper(II) chloride monohydrate, [CuCl(C5H9­N3)­(C12H8N2)]Cl·H2O, the metal centre adopts a five‐coordinate geometry, ligated by the two phenanthroline N atoms, two amine N atoms of the hist­amine ligand (one aliphatic and one from the imidazole ring) and a chloro ligand. The geometry around the Cu atom is a distorted compressed trigonal bipyramid, with one phenanthroline N and one imidazole N atom in the axial positions, and the other phenanthroline N atom, the histamine amine N atom and the chloro ligand in the equatorial positions. The structure includes an uncoordinated water mol­ecule, and a Cl ion to complete the charge. The water mol­ecule is hydrogen bonded to both Cl ions (coordinated and uncoordinated), and exhibits a close Cu⋯H contact in the equatorial plane of the bipyramid.  相似文献   

14.
The title compound, [Cu(C2N3)(C12H8N2)2]ClO4, represents a relatively rare class of compounds with dicyan­amide coordinated in a monodentate manner. The structure is formed by the [Cu{N(CN)2}(phen)2]+ complex cation (phen is 1,10‐phenanthroline) and an uncoordinated ClO4 anion. The Cu atom is five‐coordinate, with a slightly distorted trigonal–bipyramidal environment. The dicyan­amide ligand is coordinated through one nitrile N atom in the equatorial plane, at a distance of 2.033 (6) Å from the metal. The two axial Cu—N distances are similar [mean 1.999 (4) Å] and are substantially shorter than the remaining two equatorial Cu—N bonds [mean 2.087 (1) Å].  相似文献   

15.
Diorganotin (IV) complexes SnR2X2 (R = Me, Ph; X = Cl, NCS) form a series of versatile complexes when react with bidentate substituted pyridyl ligands. The reaction of dimethyltin dichloride with 5,5′‐dimethyl‐2,2′‐bipyridine (5,5′‐Me2bpy) resulted in the formation of [SnMe2Cl2(5,5′‐Me2bpy)] ( 1 ). Moreover, the reaction of SnMe2(NSC)2 with 4,4′‐di‐tert‐butyl‐2,2′‐bipyridine (bu2bpy), 1,10‐phenanthroline (phen) and 4,7‐diphenyl‐1,10‐phenanthroline (bphen) affords the hexa‐coordinated complexes [SnMe2(NCS)2(bu2bpy)] ( 2 ), [SnMe2(NCS)2(phen)] ( 3 ) and [SnMe2(NCS)2(bphen)] ( 4 ), respectively. The resulting complexes have been characterized using elemental analysis, IR, multinuclear NMR (1H, 13C, 119Sn) and DEPT‐135° NMR spectroscopy. On the other hand, the reaction of diphenyltin dichloride with 2,2′‐biquinoline (biq) and 4,7‐phenantroline (4,7‐phen) led to the formation of polymeric complexes of [SnPh2Cl2(4,7‐phen)]n ( 5 ) and [SnPh2Cl2(biq)]n ( 6 ). The NMR spectra, however, reveal the ligand lability in solution and suggest a coordination number of 5 . The X‐ray crystal structures of complexes [SnMe2Cl2(5,5′‐Me2bpy)] ( 1 ), [SnMe2(NCS)2(bu2bpy)] ( 2 ) and [SnMe2(NCS)2(bphen)] ( 4 ) have been determined which reveal that the geometry around the tin atom is distorted octahedral with trans‐[SnMe2] configuration. Interestingly, the crystal structure of (H2biq)2[SnPh2Cl4]?2CHCl3 ( 7 ) was characterized by X‐ray crystallography from a chloroform solution of [SnPh2Cl2(biq)]n ( 6 ) indicating the formation of doubly protonated [H2biq]+ and [Ph2SnCl4]2? which are stabilized by a network of hydrogen bonds with a feature of trans‐[SnPh2]. The 3D Hirshfeld surface analysis and 2D fingerprint maps were used for quantitative mapping out of the intermolecular interactions for 1 , 2 , 4 and 7 which show the presence of π‐π and hydrogen bonding interactions which are associated between donor and acceptor atoms (N, S, Cl) in the solid state.  相似文献   

16.
两种镍的配合物[Ni(NH2CH2CH2CH2NH2)3]Cl2 (1)和[Ni(C6H4N2H4)2Cl2] (2)已经被合成并且通过红外和单晶X射线衍射分析对其进行了表征。在配合物1中,镍原子处于手性假八面体[NiN6]的几何构型中,它与三个1,3-丙二胺分子形成了三个六元环。在配合物2中,镍原子除了与两个o-苯二胺分子通过四个Ni-N键形成两个五元环外,它还与两个Cl原子配位形成了反式Ni-Cl2,这不同于以往报道过的镍的二胺配合物。这两个镍的配合物被MAO, MMAO或Et2AlCl活化后,对乙烯的二聚合或三聚合显示了很好的催化活性[对于配合物2,催化活性达到3.59×106 g mol-1 (Ni) h-1]。  相似文献   

17.
We report the development of a series of rhenium(I) polypyridine complexes appended with an electron‐rich diaminoaromatic moiety as phosphorogenic sensors for nitric oxide (NO). The diamine complexes [Re(N^N)(CO)3(py‐DA)][PF6] (py‐DA=3‐(N‐(2‐amino‐5‐methoxyphenyl)aminomethyl)pyridine; N^N=1,10‐phenanthroline (phen) ( 1 a ), 3,4,7,8‐tetramethyl‐1,10‐phenanthroline (Me4‐phen) ( 2 a ), 4,7‐diphenyl‐1,10‐phenanthroline (Ph2‐phen) ( 3 a )) have been synthesized and characterized. In contrast to common rhenium(I) diimines, these diamine complexes were very weakly emissive due to quenching of the triplet metal‐to‐ligand charge‐transfer (3MLCT) emission by the diaminoaromatic moiety through photoinduced electron transfer (PET). Upon treatment with NO, the complexes were converted into the triazole derivatives [Re(N^N)(CO)3(py‐triazole)][PF6] (py‐triazole=3‐((6‐methoxybenzotriazol‐1‐yl)methyl)pyridine; N^N=phen ( 1 b ), Me4‐phen ( 2 b ), Ph2‐phen ( 3 b )), resulting in significant emission enhancement (I/I0≈60). The diamine complexes exhibited high reaction selectivity to NO, and their emission intensity was found to be independent on pH. Also, these complexes were effectively internalized by HeLa cells and RAW264.7 macrophages with negligible cytotoxicity. Additionally, the use of complex 3 a as an intracellular phosphorogenic sensor for NO has been demonstrated.  相似文献   

18.
The kinetics and mechanism of the reaction of complexation of iron(III) with 2,4-octanedione and 2,4-nonanedione have been investigated spectrophotometrically in aqueous solution at 10°C and ionic strength 0.5 mol dm?3 NaClO4. The equilibrium constants of the mono-complexes have been determined. The mechanism proposed to account for the kinetic data involves a double reversible pathway where both Fe3+ and Fe(OH)2+ react with the enol tautomer of the ligand. 2,4-Octanedione reacts with Fe3+ and Fe(OH)2+ with rate constants of 0.65 dm3 mol?1 s?1, and 14.07 dm3 mol?1 s?1, respectively. For 2,4-nonanedione complexation the rate constants determined are 0.49 dm3 mol?1 s?1, and 11.39 dm3 mol?1 s?1, respectively. Some discussions are made on the basis of Eigen-Wilkins theory considering the effect of solvent exchange on the complex formation. © 1993 John Wiley & Sons, Inc.  相似文献   

19.
The reaction between Au(I), generated by reaction of thallium(I) with Au(III), and peroxydisulphate was studied in 5 mol dm?3 hydrochloric acid. The reaction proceeds with the formation of an ion‐pair between peroxydisulphate and chloride ion as the Michealis–Menten plot was linear with intercept. The ion‐pair thus formed oxidizes AuCl2? in a slow two‐electron transfer step without any formation of free radicals. The ion‐pair formation constant and the rate constant for the slow step were determined as 113 ± 20 dm?3 mol?1 and 5.0 ± 1.0 × 10?2 dm3 mol?1 s?1, respectively. The reaction was retarded by hydrogen ion, and formation of unreactive protonated form of the reductant, HAuCl2, causes the rate inhibition. From the hydrogen ion dependence of the reaction rate, the protonation constant was calculated to be as 0.6 ± 0.1 dm3 mol?1. The activation parameters were determined and the values support the proposed mechanism. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 589–594, 2002  相似文献   

20.
The kinetics of the reactions between Fe(phen) 3 2+ [phen = tris–(1,10) phenanthroline] and Co(CN)5X3− (X = Cl, Br or I) have been investigated in aqueous acidic solutions at I = 0.1 mol dm−3 (NaCl/HCl). The reactions were carried out at a fixed acid concentration ([H+] = 0.01 mol dm−3) and the second-order rate constants for the reactions at 25 °C were within the range of (0.151–1.117) dm3 mol−1 s−1. Ion-pair constants K ip for these reactions, taking into consideration the protonation of the cobalt complexes, were 5.19 × 104, 3.00 × 102 and 4.02 × 104 mol−1 dm−3 for X = Cl, Br and I, respectively. Activation parameters measured for these systems were as follows: ΔH* (kJ K−1 mol−1) = 94.3 ± 0.6, 97.3 ± 1.0 and 109.1 ± 0.4; ΔS* (J K−1) = 69.1 ± 1.9, 74.9 ± 3.2 and 112.3 ± 1.3; ΔG* (kJ) = 73.7 ± 0.6, 75.0 ± 1.0 and 75.7 ± 0.4; E a (kJ) = 96.9 ± 0.3, 99.8 ± 0.4, and 122.9 ± 0.3; A (dm3 mol−1 s−1) = (7.079 ± 0.035) × 1016, (1.413 ± 0.011) × 1017, and (9.772 ± 0.027) × 1020 for X = Cl, Br, and I respectively. An outer – sphere mechanism is proposed for all the reactions.  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