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.
Phenylhydrazine (R) quantitatively reduces [Fe2(μ-O)(phen)4(H2O)2]4+ (1) (phen?=?1,10-phenanthroline) and its conjugate base [Fe2(μ-O)(phen)4(H2O)(OH)]3+ (2) to [Fe(phen)3]2+ in presence of excess 1,10-phenanthroline in the pH range 4.12–5.55. Oxidation products of phenylhydrazine are dinitrogen and phenol. The reaction proceeds through two parallel paths: 1?+?R?→?products (k 1), 2?+?R?→?products (k 2); neither RH+ nor the doubly deprotonated conjugate base of the oxidant, [Fe2(μ-O)(phen)4(OH)2]2+ (3) is kinetically reactive though both are present in the reaction media. At 25.0°C, I?=?1.0?M (NaNO3), the rate constants are k 1?=?425?±?10?M?1?s?1 and k 2?=?103?±?5?M?1?s?1. An inner-sphere, one-electron, rate-limiting step is proposed.  相似文献   

2.
The synthesis, characterization, and theoretical studies of titanium-μ-N2 complexes with di-anionic guanidinate ligands were reported as the first example of its kind. Thus, with(Me3Si)2N-guanidinate ligands, the mono-anionic guanidinate-supported titanium-μ-N2complex 1 was obtained. Then, reduction of 1 with potassium afforded the di-anionic guanidinate-supported titanium-μ-N2complex 2 via cleavage of one N–Si bond of the(Me3Si)<...  相似文献   

3.
This Perspective article discusses three interesting chemical transformations promoted by the octahedral [Re(6)(μ(3)-Se)(8)](2+) cluster core. These include (1) nucleophilic addition of alcohols to cluster-bound nitrile(s) to produce imino ester complexes; (2) synthesis of cluster-imine complexes with geometric specificity by reacting cluster nitrile solvates with organic azides; and (3) preparation and reactivity studies of carbonyl complexes of the cluster. We found that cluster-bound nitrile ligands were activated toward nucleophilic attack by methanol or ethanol, affording predominantly the Z-configured cluster-imino ester complexes, for which a mechanism evoking bifurcated hydrogen bonding interactions involving both the alcohol OH group and two nearest Se atoms of the cluster core was proposed. When reacted with organic azides, cluster-bound nitrile ligands were displaced and cluster-imine complexes were obtained, presumably through the formation of the corresponding cluster-azide complexes, followed by their photodecomposition. Lastly, cluster complexes featuring all-terminal carbonyl ligands were synthesized. Back-bonding interactions were verified, both experimentally and by computational studies. Their thermal and photo-stabilities were also evaluated, so was their reactivity toward methyl lithium for the eventual making of cluster-carbene catalysts. These findings, together with those by others, portend an exciting, new direction in the chemistry of solid-state type transition metal clusters.  相似文献   

4.
The electronic and geometric structures of a series of iron(II) complexes supported by tetradentate tris(pyrid-2-ylmethyl)amine-type ligands with different numbers of 4-nitropyridine groups, [(PyCH(2))(3-n)(4-NO(2)PyCH(2))(n)N] (n = 0-3), were examined by X-ray absorption fine-structure and variable-temperature (1)H NMR spectroscopies and theoretical calculations to reveal how the low-spin state is stabilized through π back-bonding interactions between iron(II) and 4-nitropyridine donor group(s).  相似文献   

5.
Benzyldimethyltetradecylammonium (BA14(+)) salts with anionic ligands (X(-)), such as bis(2-ethylhexyl)sulfosuccinate, bis(2-ethylhexyl)phosphate (BEHP(-)), and benzotriazole (BTA(-)) anions, were prepared. These salts were soluble in various organic solvents. The luminescence emission spectra of organic solutions of a red luminescent, tris(1,1,1,2,2,3,3-heptafluoro-7,7-dimethyl-4,6-octadionato)europium(III) complex in the presence of the BA14X's were recorded. The emission intensity of the Eu(III) complex was increased remarkably by the addition of BA14X (X(-) = BEHP(-) and BTA(-)). This effect can be attributed to the formation of 1:1 X(-)-adducts of the Eu(III) complex, in which the asymmetry of the ligand field is increased so as to enhance the emission efficiency of the (5)D(0)→(5)F(2) transition. The enhancement effect by BA14X was higher than that of charge-neutral ligands, such as tri-n-octylphosphine oxide and 1,10-phenanthroline, which have been used as second ligands to enhance the emission efficiency of tris(β-diketonato)europium(III) complexes.  相似文献   

6.
Summary The kinetics of the oxidation of ascorbic acid by diaquatetrakis (2,2-bipyridine)--oxo diruthenium(III) in aqueous HClO4 were investigated. The dependence of the second order rate constantk 2 on [H+] is given by k 2=a+b[H+], indicating that both the undissociated form and the monoanion of ascorbic acid are reactive. Marcus theory was used to estimate the redox potential for the RuIII-O-RuIII/RuIII-O-RuII couple and a feasible mechanism has been proposed to explain the results.  相似文献   

7.
By the DFT (U)PBE0 method the structural parameters of molecules, cations, dications, and anions of gallium(III) (μ-oxo)bis[phthalocyaninate], gallium(III) (μ-oxo)bis[perfluorophthalocyaninate], and heteroleptic bis-phthalocyaninate FPcGaOGaPc are determined. The ∠GaOGa bond angle and the Ga?Ga internuclear distance depend non-monotonically on the charge. The ionization potential of the (PcGa)2O molecule of 5.71 eV, the second electron detachment energy of 7.94 eV, and the electron affinity of 2.14 eV increase to 6.14 eV, 8.37 eV, and 2.72 eV after the perfluorination of one Pc moiety and to 6.60 eV, 8.70 eV, and 3.13 eV respectively after complete fluorination.  相似文献   

8.
The kinetics of oxidation of the neutralized -hydroxy acids: lactic, -hydroxyisobutyric, mandelic, benzilic and atrolactic acids by tris(pyridine-2-carboxylato)manganese(III) have been studied. The reactions were carried out in a Na(pic)-picH [Na(pic) = sodium salt of pyridine-2-carboxylic acid and picH = pyridine-2-carboxylic acid] buffer medium in the 4.89–6.10pH range. The oxidation rate was found to be independent of pH, and rate follows the order: benzilate > mandelate >atrolactate>lactate > -hydroxy isobutyrate. The oxidation products are MeCHO, Me2CO, PhCHO, Ph2CO and PhCOMe for the respective reactions. A mechanism is proposed involving intermediate formation of hepta-coordinated MnIII complexes in a fast step. The complexes then decompose to give free radicals and MnII in the rate determining step. The free radicals subsequently react with another molecule of the MnIII species to give the respective carbonyl compounds in a fast step.  相似文献   

9.
Rate enhancements for the reduction of dioxygen by a Mn(II) complex were observed in the presence of redox-inactive group 2 metal ions. The rate changes were correlated with an increase in the Lewis acidity of the group 2 metal ions. These studies led to the isolation of heterobimetallic complexes containing Mn(III)-(μ-OH)-M(II) cores (M(II) = Ca(II), Ba(II)) in which the hydroxo oxygen atom is derived from O(2). This type of core structure has relevance to the oxygen-evolving complex within photosystem II.  相似文献   

10.
To obtain structural and spectroscopic models for the diiron(II,III) centers in the active sites of diiron enzymes, the (μ-alkoxo)(μ-carboxylato)diiron(II,III) complexes [Fe(II)Fe(III)(N-Et-HPTB)(O(2)CPh)(NCCH(3))(2)](ClO(4))(3) (1) and [Fe(II)Fe(III)(N-Et-HPTB)(O(2)CPh)(Cl)(HOCH(3))](ClO(4))(2) (2) (N-Et-HPTB = N,N,N',N'-tetrakis(2-(1-ethyl-benzimidazolylmethyl))-2-hydroxy-1,3-diaminopropane) have been prepared and characterized by X-ray crystallography, UV-visible absorption, EPR, and M?ssbauer spectroscopies. Fe1-Fe2 separations are 3.60 and 3.63 ?, and Fe1-O1-Fe2 bond angles are 128.0° and 129.4° for 1 and 2, respectively. M?ssbauer and EPR studies of 1 show that the Fe(III) (S(A) = 5/2) and Fe(II) (S(B) = 2) sites are antiferromagnetically coupled to yield a ground state with S = 1/2 (g= 1.75, 1.88, 1.96); M?ssbauer analysis of solid 1 yields J = 22.5 ± 2 cm(-1) for the exchange coupling constant (H = JS(A)·S(B) convention). In addition to the S = 1/2 ground-state spectrum of 1, the EPR signal for the S = 3/2 excited state of the spin ladder can also be observed, the first time such a signal has been detected for an antiferromagnetically coupled diiron(II,III) complex. The anisotropy of the (57)Fe magnetic hyperfine interactions at the Fe(III) site is larger than normally observed in mononuclear complexes and arises from admixing S > 1/2 excited states into the S = 1/2 ground state by zero-field splittings at the two Fe sites. Analysis of the "D/J" mixing has allowed us to extract the zero-field splitting parameters, local g values, and magnetic hyperfine structural parameters for the individual Fe sites. The methodology developed and followed in this analysis is presented in detail. The spin Hamiltonian parameters of 1 are related to the molecular structure with the help of DFT calculations. Contrary to what was assumed in previous studies, our analysis demonstrates that the deviations of the g values from the free electron value (g = 2) for the antiferromagnetically coupled diiron(II,III) core in complex 1 are predominantly determined by the anisotropy of the effective g values of the ferrous ion and only to a lesser extent by the admixture of excited states into ground-state ZFS terms (D/J mixing). The results for 1 are discussed in the context of the data available for diiron(II,III) clusters in proteins and synthetic diiron(II,III) complexes.  相似文献   

11.
The synthesis and characterization of the oxo-centered carboxylato-bridged trinuclear iron(III) complex, triaquahexakis(2-betaine)(3-oxo)triiron(III) perchlorate heptahydrate are described. X-ray crystallography shows that the FeIII atom in the complex has a slightly distorted octahedral geometry, coordinated by four oxygen atoms from different betaine ligands [Fe—;O = 2.009(3) 2.034(3) Å], one aqua ligand [Fe—O = 2.028(4) and 2.031(3) Å] and the central 3-oxo atom [Fe—O = 1.917(2) and 1.917(3) Å]. The central oxygen is ideally coplanar with the plane of the three metal atoms. Magnetic susceptibility data (4–320 K) show the presence of an antiferromagnetic exchange interaction with a coupling constant of J = –20.2 cm–1.  相似文献   

12.
The synthesis and characterization of mononuclear tris(κ2-amidate) aluminium complexes supported by the tripodal ligands, [N(o-PhNC(O)R)3]3− (R = iPr and tBu), are described. The molecular structures of [Al(N(o-PhNC(O)iPr)3)] and [Al(N(o-PhNC(O)tBu)3)] have been determined by X-ray diffraction studies. Both neutral six-coordinate aluminium complexes display coordination geometries that are intermediate between octahedral and trigonal prismatic. Solution-state NMR studies (1H, 13C and 27Al) indicate that these structures are non-fluxional in solution. Detailed analysis of the solid-state structures shows that slight changes in the relative size of the amidate acyl substituents do not significantly impact the solid-state structures. However, large substituents may be required to prevent the formation of multinuclear species.  相似文献   

13.
The heterolytic dissociation of water with the previously unknown four-coordinated syn-linear (μ-oxo)divanadium core to form the five-coordinated bis(μ-hydroxo)divanadium core was confirmed by a crystal-to-crystal transformation.  相似文献   

14.
Base mismatches arise naturally in the life cycleof a cell as a result of either polymerase error or DNAdamage. Under most circumstances the cell correctsthese mispairings using a complex repair system toprevent mutations in the genetic code. Experimental…  相似文献   

15.
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   

16.
The reduction of the octahedral cobalt(III) complex CoIII(HL)·9H2O, H4L = 1,8-bis(2-hydroxybenzamido)-3,6-diazaoctane by glutathione (GSH) has been studied by conventional spectrophotometry at 25.0 ≤ t/°C ≤ 45.0, 0.02 ≤ [H+]/mol dm?3 ≤ 0.20 and I = 0.3 mol dm?3 (NaClO4). The reaction is biphasic. The fast initial phase is attributed to the H+-induced formation of the mixed ligand complex, [CoIII(H2L)GSH]+, for which the rate-limiting step is the chelate ring opening via CoIII–NH (amide–N) bond cleavage of the protonated species, [CoIII(H2L)]+. Outer-sphere association equilibria between GSH/GSH2 + and [CoIII(H2L)]+ substantially retard the ring opening process and consequently the mixed ligand complex formation. This is then followed by a slow phase involving reduction of [CoIII(H2L)GSH]+ by both GSH and GSH2 +. The final products are the corresponding Co(II) complex and the oxidized form of GSH, GS–SG. The kinetic data and activation parameters for the redox process are interpreted in terms of an outer-sphere electron transfer mechanism.  相似文献   

17.
Heterobimetallic compounds [L(1)OV(V)═O→Fe(metsalophen)(H(2)O)] (1) and [L(2)OV(V)═O→Fe(metsalophen)(H(2)O)]CH(3)CN (2), where H(2)L(1) and H(2)L(2) are tridentate dithiocarbazate-based Schiff base ligands, containing a discrete V(V)-μ-O-Fe(III) angular core have been synthesized for the first time through a targeted synthesis route: confirmation in favor of such a heterobimetallic core structure has come from single-crystal X-ray diffraction analysis and electrospray ionization mass spectrometry.  相似文献   

18.
《Tetrahedron: Asymmetry》2001,12(16):2289-2293
The chiral [5-(4-hydroxybutyl)-5′-methyl-2,2′-bipyridine]-bis(2,2′-bipyridine)-ruthenium(II)-bis(hexafluoroantimonate) complex 3 was prepared and characterized by different NMR techniques and successfully separated into enantiomers by electrokinetic chromatography using anionic carboxymethyl-β-cyclodextrin as chiral mobile phase additive (CMPA). The optimum separation conditions were obtained with 40 mM borate buffer at pH 9.5 and 7.5 mg/mL of the chiral selector at 20°C.  相似文献   

19.
Bis(μ3-2-oxybenzaldoximato-O,O′,N)-(μ2-oxo)-bis(5-bromo-2-methoxyphenyl)antimony, which crystallizes from toluene in the form of a solvate, has been synthesized by the reaction between tris(5-bromo-2-methoxyphenyl)antimony and 2-oxybenzaldoxime in the presence of hydrogen peroxide. According to X-ray diffraction data, the structurally equivalent antimony atoms in a binuclear complex molecule are linked via two tridentate bridging ligands and an oxygen atom and have a distorted octahedral coordination to the C2O3N surrounding The trans-angles of the octahedron are CSbO (168.7(1)°), CSbN (164.1(1)°), and OSbO (159.3(1)°), and the bond lengths are Sb-C (2.127(4), 2.159(3) Å) and Sb-N (2.238(3) Å). A molecule contains the three types of Sb-O distances with a bridging oxygen atom (1.950(1) Å), the oxygen atoms of oxy groups (2.002(2) Å), and the oxygen atoms of oxime groups (2.091(2) Å). Sb?OCH3 intramolecular contacts (3.097, 3.290 Å) also exist.  相似文献   

20.
Eight new dicobalt-iron clusters have been synthesised and structurally characterized. Treatment of (μ3-S)FeCo2(CO)9 (A) with monophosphane ligands tris(4-fluorophenyl)phosphane, tris(4-methoxyphenyl)phosphane, or tris(2-furyl)phosphane in the presence of Me3NO?2H2O afforded monosubstituted complexes (μ3-S)FeCo2(CO)8L [L = P(4-C6H4F)3, 1; P(4-C6H4OMe)3, 3; P(2-C4H3O)3, 5] and disubstituted complexes (μ3-S)FeCo2(CO)7L2 [L = P(4-C6H4F)3, 2; P(4-C6H4OMe)3, 4; P(2-C4H3O)3, 6]. Reaction of complex A with Ph2PN[CH(CH3)2]PPh2 in refluxing toluene gave complex (μ3-S)FeCo2(CO)7{Ph2PN[CH(CH3)2]PPh2} (7) with an intramolecular bridging diphosphane ligand. Reaction of complex A with trans-1,2-bis(diphenylphosphino)ethylene (trans-dppv) and Me3NO?2H2O yielded complex [(μ3-S)FeCo2(CO)8]2(trans-Ph2PCH = CHPPh2) (8) with an intermolecular bridging diphosphane ligand. The new complexes 1–8 were characterized by elemental analysis, IR, 1H NMR, 31P{1H} NMR, and 13C{1H} NMR spectroscopy, particularly for 1, 3, and 68 by X-ray crystallography.  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