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 kinetics and mechanism of reduction of the surfactant-cobalt(III) complex ions, cis-[Co(bpy)2(C12H25NH2)2]3+ and cis-[Co(phen)2(C12H25NH2)2]3+ (bpy = bipyridyl, phen = 1,10-phenan-throline, C12H25NH2 = dodecylamine) by Fe(CN6)4− in self-micelles were studied at different temperatures. Experimentally the reaction was found to be second order and the electron transfer postulated as outersphere. The rate constant for the electron transfer reaction for both the complexes was found to increase with increase in the initial concentration of the surfactant-cobalt(III) complex. This peculiar behaviour of dependence of second-order rate constant on the initial concentration of one of the reactants has been attributed to the presence of various concentration of micelles under different initial concentration of the surfactantcobalt(III) complexes in the reaction medium. The effect of inclusion of the long aliphatic chain of the surfactant complex ions into β-cyclodextrin on these reactions has also been studied.  相似文献   

2.
The outer sphere electron transfer reaction of surfactant cobalt(III) complexes, Cis-[Co(en)2(4CNP)(C12H25NH2)](ClO4)3 1, Cis-[Co(trien)(4CNP)(C12H25NH2)](ClO4)3 2 and Cis-[Co(trien)(4AMP)(C12H25NH2)](ClO4)3 3 (en: ethylenediamine, trien: triethylenetetramine, 4CNP: 4-cyanopyridine, 4AMP: 4-aminopyridine, C12H25NH2: dodecylamine) have been investigated by Fe2+ ion in liposome vesicles (DPPC) and ionic liquids medium at different temperatures under pseudo first order conditions using an excess of the reductant. In the presence of ionic liquid medium the second order rate constant for this electron transfer reaction was found to increase with increasing concentration of ionic liquids. Below the phase transition temperature of DPPC, the rate decreased with increasing concentration of DPPC, while above the phase transition temperature the rate increased with increasing concentration of DPPC for the same complexes has also been studied. Experimentally the reactions were found to be second order and the electron transfer postulated as outer sphere. The results have been discussed in terms of increased hydrophobic effect, self aggregation and the presence of pyridine ligand containing 4-amino and 4-cyano substituent.  相似文献   

3.
The effect of unilamellar vesicles of dipalmitoylphosphotidylcholine (DPPC), both below and above the phase transfer region, on the second-order rate constants for outer-sphere electron transfer between Fe2+ and the surfactant?Ccobalt(III) complexes, cis-[Co(en)2(C12H25NH2)2]3+ and cis-[Co(trien)(C12H25NH2)2]3+ (en?=?ethylenediamine, trien?=?triethylenetetramine, C12H25NH2?=?dodecylamine) was studied by UV?CVis absorption spectroscopy. Below the phase transition temperature of DPPC, the rate decreased with increasing concentration of DPPC, while above the phase transition temperature the rate increased with increasing concentration of DPPC. It is concluded that below the phase transition temperature, there is an accumulation of surfactant?Ccobalt(III) complexes at the interior of the vesicle membrane through hydrophobic effects, and above the phase transition temperature the surfactant?Ccobalt(III) complex is released from the interior to the exterior surface of the vesicle. Through isokinetic plots, we have established that the mechanism of the reaction does not alter during the phase transition of DPPC.  相似文献   

4.

Abstract  

The surfactant complex ion cis-[Co(tmd)2(C12H25NH2)2]3+ (tmd = 1,3-propanediamine, C12H25NH2 = dodecylamine) has been synthesized and characterized by elemental analysis and spectral data. In addition we have determined the critical micelle concentration of the surfactant–cobalt(III) complex and studied the kinetics and mechanism of the complex with ferrocyanide anion. The reaction is found to be second order, and the second-order rate constant increases with increasing initial concentration of the surfactant–cobalt(III) complex due to the presence of self-micelles formed by the complex itself. The thermodynamic parameters were determined. The results have been analyzed.  相似文献   

5.
The electron‐transfer reaction of some surfactant cobalt(III) complexes, cis‐[Co(ip)2(C12H25NH2)2]3+ 1 , cis‐[Co(dpq)2(C12H25NH2)2]3+ 2 , and cis‐[Co(dpqc)2(C12H25NH2)2]3+ 3 (ip = imidazo[4,5‐f][1,10]phenanthroline, dpq = dipyrido[3,2‐d:2′‐3′‐f]quinoxaline, dpqc = dipyrido[3,2‐a:2′,4′‐c](6,7,8,9‐tetrahydro)phenazine, C12H25NH2 = dodecylamine) with the Fe(CN)64? ion has been investigated in microheterogeneous media (micelles, β‐cyclodextrin) at different temperatures by the spectrophotometric method under pseudo‐first‐order conditions using an excess of the reductant. Experimentally, the reaction was found to be second order and the electron transfer postulated as an outer sphere. The rate constant for the electron‐transfer reaction in micelles was found to increase with an increase in the initial concentration of the surfactant–cobalt(III) complex. This peculiar behavior of dependence of the second‐order rate constant on the initial concentration of one of the reactants has been attributed to the presence of various concentrations of micelles under different initial concentrations of the surfactant–cobalt(III) complex in the reaction medium. Inclusion of the long aliphatic chain of the surfactant complex ion into β‐cyclodextrin leads to decrease in the rate constant. Kinetic data and activation parameters are interpreted in terms of an outer‐sphere electron‐transfer mechanism. All these results have been interpreted in terms of the hydrophobic effect and the reactants with the opposite charge.  相似文献   

6.
The reactivity of the cobalt(III) complexes dichlorido[tris(2‐aminoethyl)amine]cobalt(III) chloride, [CoCl2(tren)]Cl, and dichlorido(triethylenetetramine)cobalt(III) chloride, [CoCl2(trien)]Cl, towards different amino acids (l ‐proline, l ‐asparagine, l ‐histidine and l ‐aspartic acid) was explored in detail. This study presents the crystal structures of three amino acidate cobalt(III) complexes, namely, (l ‐prolinato‐κ2N,O)[tris(2‐aminoethyl)amine‐κ4N,N′,N′′,N′′′]cobalt(III) diiodide monohydrate, [Co(C5H8NO2)(C6H18N4)]I2·H2O, I , (l ‐asparaginato‐κ2N,O)[tris(2‐aminoethyl)amine‐κ4N,N′,N′′,N′′′]cobalt(III) chloride perchlorate, [Co(C4H7N2O3)(C6H18N4)](Cl)(ClO4), II , and (l ‐prolinato‐κ2N,O)(triethylenetetramine‐κ4N,N′,N′′,N′′′)cobalt(III) chloride perchlorate, [Co(C4H7N2O3)(C6H18N4)](Cl)(ClO4), V . The syntheses of the complexes were followed by characterization using UV–Vis spectroscopy of the reaction mixtures and the initial rates of reaction were obtained by calculating the slopes of absorbance versus time plots. The initial rates suggest a stronger reactivity and hence greater affinity of the cobalt(III) complexes towards basic amino acids. The biocompatibility of the complexes was also assessed by evaluating the cytotoxicity of the complexes on cultured normal human fibroblast cells (WS1) in vitro. The compounds were found to be nontoxic after 24 h of incubation at concentrations up to 25 mM.  相似文献   

7.
The surfactant–cobalt(III) complex, cis-[Co(trien)(4AMP)(DA)](ClO4)3, trien = triethylenetetramine, 4AMP = 4-aminopyridine, DA = dodecylamine was synthesized and characterized by various spectroscopic and physico-chemical techniques. The critical micelle concentration (CMC) value of this surfactant–cobalt(III) complex in aqueous solution was found out from conductance measurements. The conductivity data (at 303, 308, 313, 318 and 323 K) were used for the evaluation of the temperature-dependent CMC and the thermodynamics of micellization (ΔG m ° , ΔHm and ΔS m ° ). Also the kinetics of reduction of this surfactant–cobalt(III) complex by hexacyanoferrate(II) ion in micelles, β-cyclodextrin, ionic liquids (ILs) and in liposome vesicles (DPPC) media were studied at different temperature. The rate constant for the electron transfer reaction in micelles was found to increase with increase in the initial concentration of the surfactant–cobalt(III) complex. This peculiar behaviour of dependence of second-order rate constant on the initial concentration of one of the reactants has been attributed to the presence of various concentration of micelles under different initial concentration of the surfactant–cobalt(III) complex in the reaction medium. Inclusion of the long aliphatic chain of the surfactant complex ion into β-cyclodextrin leads to decrease in the rate constant. Below the phase transition temperature of DPPC, the rate decreased with increasing concentration of DPPC, while above the phase transition temperature the rate increased with increasing concentration of DPPC. It is concluded that below the phase transition temperature, there is an accumulation of surfactant–cobalt(III) complex at the interior of the vesicle membrane through hydrophobic effects, and above the phase transition temperature the surfactant–cobalt(III) complex is released from the interior to the exterior surface of the vesicle. In the presence of ionic liquid medium the second order rate constant for this electron transfer reaction for the same complex was found to increase with increasing concentration of ILs has also been studied. An outer-sphere mechanism is proposed for all these reactions and the results have been explained based on the hydrophobicity of the ligand and the reactants with opposite charges.  相似文献   

8.
Abstract The critical micelle concentrations (CMC) of three metallosurfactant Schiff base cobalt(III) complexes of the type [Co(trien)(C19H30NO)]Cl2,[Co(trien) (C23H38NO)]Cl2 and [Co(trien)(C25H42NO)]Cl2, where trien = triethylenetetramine, have been studied in n-alcohol and in formamide at different temperatures by the electrical conductivity method. CMCs have also been measured as a function of percentage concentration of alcohol in the mixed solvents with formamide. Specific conductivity data (at 303–323 K) served for the evaluation of the temperature-dependent CMC and thermodynamic parameters such as the standard Gibbs energy changes (DG°mic)\Delta G^{\circ}_{\mathrm{mic}}), enthalpy changes (DH°mic)\Delta H^{\circ}_{\mathrm{mic}}), and entropy changes (DS°mic)\Delta S^{\circ}_{\mathrm{mic}}) of micelle formation. It is suggested that addition of an alcohol leads to increased penetration of formamide into the micellar interface, the extent depending on the alcohol’s chain length. The results have been discussed in terms of the solvophobic interaction, dielectric constant of the medium, the chain length of the alcohol, and the surfactant in the solvent mixture.  相似文献   

9.
The kinetics and mechanism of reduction of the surfactant complex ions, cis-chloro/bromo(dodecylamine)(triethylenetetramine)cobalt(III) by iron(II) in aqueous solution were studied at 303, 308 and 313 K by spectrophotometry under pseudo-first-order conditions using an excess of the reductant. The second-order rate constant increases with cobalt(III) concentration and the presence of aggregation of the complex itself alters the reaction rate. The reductions are acid-independent in the range [H+] = 0.05–0.25 mol dm−3. Variation of ionic strength (μ) influences the reaction rate. Activation and thermodynamic parameters have been computed. It is suggested that the reaction of Fe2+(aq) with the cobalt(III) complex proceeds by an inner-sphere mechanism. The critical micelle concentration (CMC) values of these surfactant metal complexes in aqueous solution were obtained from conductance measurements. Specific conductivity data (at 303, 308 and 313 K) served for the evaluation of the temperature-dependent CMC and the standard Gibbs energy of micellization (ΔGm0).  相似文献   

10.
Surface active micelle formable surfactant-Cr(III) complexes of the type cis-α-[Cr(trien)(C16H33NH2)X]2+ (where trien = triethylenetetramine; X = F, Cl, Br) have been studied in n-alcohol and in formamide at different temperatures by conductance measurements. Standard Gibbs energy changes (ΔG o mic), enthalpies (ΔH o mic) and entropies (ΔS o mic) of micelle formation have been determined by studying the variation of the Critical Micelle Concentration (CMC) with temperature. Critical micelle concentrations have also been measured as a function of percentage concentration of alcohol added. It is suggested that alcohol addition leads to an increase in formamide penetration into the micellar interface that depends on the alcohol chain length. The results are discussed in terms of an increased hydrophobic effect, dielectric constant of the medium, the chain length of the alcohols and the surfactant in the solvent mixture.  相似文献   

11.
The surfactantCo(III) complexes of the type cis-[Co(en)2AX]2+ (A?=?Tetradecylamine, X?=?Cl?,?Br?) were synthesised from corresponding dihalogeno complexes by the ligand substitution method. The critical micelle concentration (CMC) values of these surfactant complexes in aqueous solution were obtained from conductance measurements. The kinetics and mechanism of iron(II) reduction of surfactantCo(III) complexes, cis-[Co(en)2(C14H29NH2)Cl](ClO4)2 and cis-[Co(en)2(C14H29NH2)Br] (ClO4)2 ions were studied spectrophotometrically in an aqueous acid medium by following the disappearance of Co(III) using an excess of the reductant under pseudo-first-order conditions: [Fe(II)]?=?0.25?mol?dm?3, [H+]?=?0.1?mol?dm?3, [μ]?=?1.0?mol?dm?3 ionic strength in a nitrogen atmosphere at 303, 308 and 313?K. The reaction was found to be of second order and showed acid independence in the range [H+]?=?0.05–0.25?mol?dm?3. The second-order rate constant increased with surfactant–Co(III) concentration and the presence of aggregation of the complex itself altered the reaction rate. The effects of [Fe(II)], [H+] and [μ] on the rate were determined. Activation and thermodynamic parameters were computed. It is suggested that the reaction of [Fe(II)] with Co(III) complex proceeds by an inner-sphere mechanism.  相似文献   

12.
31P N.M.R. Spectroscopic Investigations of Phosphato Complexes of Cobalt(III) and Rhodium(III) 31P n.m.r. spectra of phosphato complexes of cobalt and rhodium(III) are recorded. The coordination shift of monodentate phosphate is 8–9 ppm, of bidentate phosphate 18 ppm, of phosphite 10–11 ppm, of fluorophosphate 6–7 ppm. Monophosphato and pentaamminaquacobalt complexes condense with the elimination of water to a m?-complex. The ability of the phosphato complexes of cobalt(III) to condense phosphate to diphosphate was investigated. After heating [CoPO4en2] · 2 H2O with an excess of dihydrogenphosphate only small amounts of the expected diphosphate complex could be detected. The analogous reaction with fluorophosphate results in an appreciably higher yield of the diphosphate complex.  相似文献   

13.
The novel (E,E)-dioxime 5,6:13,14-dibenzo-9,10-benzo(15-crown-5)-2,3-bis(hydroxyimino)-7,12-dioxo-1,4,8,11-tetraazacyclotetradecane (H2L) has been synthesized by the reaction of 4′,5′-diaminobenzo(15-crown-5) with N,N′-bis(2-carbomethoxyphenyl)diaminoglyoxime (1). Only mononuclear CoIII and RuII complexes with a metal/ligand ratio of 1:2 have been isolated. The cobalt(III) complex bridged with BF2+ is achieved with H-bonded cobalt(III) complex and borontrifluoride ethyl ether complex. The reaction of BF2 bridged cobalt(III) complex with bis(benzonitril)palladium(II) chloride gives a trinuclear complex. The structures of dioxime and its complexes are proposed according to elemental analyses, 1H and 13C-NMR, IR and mass spectral data.  相似文献   

14.
Twelve surfactant Schiff base ligands were synthesized from salicylaldehyde and its chloro-, bromo- and methoxy- derivatives by condensation with long-chain aliphatic primary amines, and a number of mixed ligand cobalt(III) surfactant Schiff base coordination complexes of the type [Co(trien)A]2+ were synthesized from the corresponding dihalogeno complexes by ligand substitution. The Schiff bases and their complexes were characterized by physico-chemical and spectroscopic methods. The complexes form foams in aqueous solution upon shaking. The critical micelle concentration (CMC) values of the complexes in aqueous solution were obtained from conductance measurements. Specific conductivity data (at 303–323 K) served for the evaluation of the thermodynamics of micellization ( \Updelta G\textm0 \Updelta G_{\text{m}}^{0} , \Updelta H\textm0 \Updelta H_{\text{m}}^{0} , \Updelta S\textm0 \Updelta S_{\text{m}}^{0} ). The complexes were tested for its antimicrobial activity.  相似文献   

15.
New Organobis(thiophenolate)bismuth(III) derivatives C6H5Bi(SC6H4Cl-p)2 (I), Ch3Bi(SC6H4NH2-p)2 (II) and CH3AS(SC6H4NH2-p)2 (III) can be prepared by the reaction of phenyldibromobismuth(III), methyldibromobismuth(III) and -arsenic(III), respectively, with stoichiometric amounts of lithium-p-chlorothiophenolate (for I), and lithium-p-aminothiophenolate. II reacts with an excess of iodomethane to form the bis(mercaptoanilinium) complex [CH3Bi(SC6H4NH2CH3-p2]2+ (I)2 (IV). Treating IV with adequate amounts of AgNO3 or TINO3 yields the analogous nitrate derivative [CH3Bi(SC6H4NH2CH3-p)2]2+ (NO3)2 (V). The new compounds were characterized by elemental analysis, 1H NMR, IR and mass spectroscopy and some of the compounds were also characterized by vapour pressure molecular weight osmometry too.The microbiological effect of the organometallic arsenic(III)- and bismuth(III)-bis(thiolates) I-IV was studied, the bismuth compounds generally inhibited bacterial growth more than the arsenic derivative. The aminothiophenolate complexes II and IV in particular were not inferior to the well-known organothiolatomercury(II) complexes with regard to their inhibition of bacterial growth. The hydrophilic mercaptoanilinium derivative IV has potential as a novel bactericide.  相似文献   

16.
Tetra-azidodiamminecobaltates(III): cis-[Co(N3)4(NH3)2]? and [Co(N3)4en]? The preparation and the properties of complexes containing the anions cis-[Co(N3)4(NH3)2]? and [Co(N3)4en]? are described. The compounds [Co(NH3)6][Co(N3)4(NH3)2 · H2O], [Co(N3)2(NH3)4][Co(N3)4(NH3)2], [As(C6H5)4][Co(N3)4en], cis- and trans-[Co(N3)2en2][Co(N3)4en] have been isolated.  相似文献   

17.
2-Pyridinamine Adducts of Transition Metal Bis(acetylacetonates) and their Reactions. Hydrogencarbonate as a Chelating Ligand in cis-(Ampy)2Co(acac)(HCO3) The reaction of cobalt(II) salts, acetylacetone (acacH), 2-pyridinamine (Ampy), and the carbon dioxide of the air in methanol affords a mixture of (Ampy)2Co(acac)2( II ) and (Ampy)2Co(CO3)(H2O)2. On heating in toluene, appropriately in the presence of carbon dioxide, these complexes are converted into cis-(Ampy)2Co(acac)(HCO3) ( III ). Characteristic of compound III is a four-membered ring with the hydrogencarbonate as a bidentate ligand. The two Co? O distances are distinctly different (215.9 and 224.4 pm). In the complexes II and III 2-pyridinamine is a bidentate ligand coordinating by the endo-nitrogen. The Co-n-N bond lengths vary between 210.9 and 225.3 pm. Reasons are both the different trans-influence of the hydrogencarbonate or the acetylacetonato donor atoms and the π-interaction between cobalt(II) and the pyridine ring. This interaction is more significant in the cis-complex III . II and III are stabilized by a system of N? H …? O- and O? H …?O-bridges. With nickel(II) complexes analogous to II and III were obtained, while only the type II was characterized for manganese( II ).  相似文献   

18.
Summary The ammoniation ofcis-[Rh(en)2Cl2] · (ClO4) in liquid NH3 was studied at constant ionic medium of 0.20 m perchlorate in the 0 to 35° range. The complex reacts in two distinct steps to givecis-[Rh(en)2(NH3)2] · (ClO4)3, with the intermediate formation ofcis-[Rh(en)2(NH3)Cl] · (ClO4)2. Both steps follow a conjugate-base mechanism. Activation parameters were obtained for the acid-base preequilibrium and the rate-determining step. The entropies of activation for the rate-determining step are 0 and –42 JK–1mol–1 for the first and second ammoniations respectively. These values are considerably lower than those found for the cobalt(III) analogues. The entropy changes for the acid-base equilibria are –84 and –36 JK–1mol–1 respectively, which is less negative than those values found for the cobalt(III) analogues. Trans-[Rh(en)2I2] · (ClO4) ammoniates totrans-[Rh(en)2(NH3)I] · (ClO4)2. The contribution of spontaneous ammoniation to the overall reaction oftrans-[Rh(en)2I2] · (ClO4) is negligible, so the uniqueness oftrans-[Co(en)2Cl2] · (ClO4) among cobalt(III) complexes in this respect is not reproduced for thetrans-dihalotetraamine structure in rhodium(III) complexes. A comparison of cobalt(III) and rhodium(III) amines with respect to activation parameters and the influence of formal charge of the metal complex on reactivity indicates a more associative type of activation for rhodium(III).  相似文献   

19.
Complexes of the type [Pt R2 (dppma-PP′)] (R─Me, Et, Ph, CH2Ph, C6H4 Me-p, C6H4OMe-2, CH2CMe3, 1-naphthyl, C6H4Me-o, dppma = Ph2PNMe PPh2) have been prepared from [PtCl2, (dppma-PP′)] and the corresponding alkyl-lithium or Grignard reagents. Equilibrium constants, k, for the conversion of [PtR2 (dppma-PP′)] into cis-[PtR2(dppma-P)2] with dppma were studied using 31P NMR spectroscopy at room temperature. Equilibrium is rapidly established for R─C6H4-Me-o, at 20°C. Complex of the type cis-[PtR2 (dppma-P)2] was isolated R─C6H4 Me-o. The complexes [PtMe2(dppma-P)2] and [Pt(o-methoxyphenyl)2(dppma-P)2] were prepared, but unfortunately decomposed once isolated, the only evidence for its formation being from 31P-{1H} NMZR spectroscopy. The o-tolyl or 1-naphthyl complexes exist as syn-anti mixtures in solution, due to restricted rotation around the platinum aryl bonds. Treatment of several complexes of the type [PtR2(dppma-PP′)] with MeI gives [PtR2Me(I)(dppma-PP′)] with trans addition of MeI. Treatment of [PtR2(dppma-PP′)] with HCl gives [Pt Cl (R) (dppma-PP′)] for R─C6H2Me3-2,4,6, C6H4-CH3-2, C6H4-Me-4, Me, 1-naphthyl. The 1H, 31P NMR parameters for these complexes are discussed. Attempted preparation of complexes of the type [PtR2 (dppma-P)2M] (R─C6H4-Me-2, Me CN-C6H4-Me-4); M─Pd, Pt, Au,) are reported.  相似文献   

20.
The kinetics of formation of dinuclear di-μ-hydroxo-diaquo-bis-cobalt(III) complexes from the corresponding tri-μ-hydroxo complexes: has been investigated with three different compounds to start with: Ammonia-Triol [N3 = (NH3)3], Dien-Triol [N3 = H2N—CH2—CH2—NH—CH2—CH2—NH2] and Tach-Triol [N3 = C6H9(NH2)3 = cis-cis-1,3,5-triaminocyclohexane]. With respect to the otherwise very inert cobalt(III)-complexes, reaction (1) is unusually rapid and takes place in two steps, the first being about 100 times faster than the second. The process begins with an exceptionally slow proton transfer to one of the bridging OH with half life of 0,2, 0,03 and 0,025 sec. respectively (perchlorate medium μ 1 M , 20°, pH = 0). The rate of the back reaction could also be determined, yielding ratios of the two rate constants corresponding to pK-values between 0 and 1,5. Whereas Tach-Triol is protonated to the di-μ-hydroxo-μ-aquo complex at about pH 1, it is deprotonated to the di-μ-hydroxo-μ-oxo complex at about pH 14 (Siroky [16]). The second step of (1), the aquation of the di-μ-hydroxo-μ-aquo species to the di-μ-hydroxo-diaquo complex (Diol) takes place with half lives between 4 and 9 sec. The final cleavage to the mononuclear triaquo complex needs many hours to go half way and is again initiated by protonation of one of the remaining two OH-bridges. Neither the μ-aquo complex produced thereby, nor the Mono-ol [mono-μ-hydroxo-tetraaquo complex] formed as an intermediate cause any observable changes of the spectrum. The Mono-ol aquates at a rate which is inverse in [H+] because of the labilisation brought about by deprotonation of its aquo ligands.  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