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.
Summary The kinetics and mechanism of the reaction between [Fe2L(OH)2]2– and cyanide ion (L = TTHA, triethylenetetraaminehexaacetate) have been studied spectrophotometrically atpH=11.0±0.1,I=0.1 M(NaClO4) and T = 25±0.1 °C. The overall reaction consists of three distinct, observable stages. The first stage involves the dissociation of the binuclear complex into a mononuclear complex [FeL(OH)]4– which then reacts with cyanide to form [Fe(CN)5OH]3–. The species [Fe(CN)5OH]3– reacts further with an excess of cyanide and forms [Fe(CN)6]3– in the second stage of reaction. The last stage involves the reduction of [Fe(CN)6]3– formed in the second stage by the TTHA6– released in the first stage of reaction. The formation of [Fe(CN)5OH]3– in the first stage is firstorder in [Fe2L(OH)2]2– and third-order in cyanide over a large range of cyanide concentrations but becomes zero-order in cyanide at [CN] < 4×10–2M.These observations enable us to suggest the presence of a slow step in which [Fe2L(OH)2]2– dissociates into [FeL(OH)]4– and [FeOH]2+ at low cyanide concentrations and a cyanide assisted rapid dissociation of [Fe2L(OH)2]2– to [FeL(OH)(CN)]5– at higher cyanide concentrations. The species [FeL(OH)(CN)]5– reacts further with an excess of cyanide to produce [Fe(CN)5OH]3– finally.The reverse reaction between [Fe(CN)5OH]3– and TTHA6– follows first-order dependence in each of [Fe(CN)5OH]3– and TTHA6– and inverse first-order dependence on cyanide concentration. A six-step mechanism has been proposed for the first stage of reaction in which the fifth has been identified as the rate-determining step.  相似文献   

2.
Summary The kinetics and mechanism of exchange of HPDTA in [Fe2HPDTA(OH)2] with cyanide ion (HPDTA=2-hydroxytrimethylenediaminetetraacetic acid) was investigated spectrophotometrically by monitoring the peak at 395 nm ( max of [Fe(CN)5OH]3– at pH=11.0±0.02,I=0.25m (NaClO4) at ±0.1°C).Three distinct observable stages were identified; the first is the formation of [Fe(CN)5OH]3–, the second the formation of [Fe(CN)6]3– from it and the third the reduction of [Fe(CN)6]3– to [Fe(CN)6]4– by HPDTA4– released in the first stage.The first stage follows first-order kinetics in [Fe2HPDTA(OH)2] and second-order in [CN] over a wide range of [CN], but becomes zero order at [CN]<5×10–2 m. We suggest a cyanide-independent dissociation of [Fe2HPDTA)(OH)2] into [FeHPDTA(OH)] and [Fe(OH)]2+ at low cyanide concentrations and a cyanide-assisted rapid dissociation of [Fe2HPDTA(OH)2] to [FeHPDTA(OH)(CN)]3– and [Fe(OH)]2+ at higher cyanide concentrations. The excess of cyanide reacts further with [FeHPDTA(OH)(CN)]3– finally to form [Fe(CN)5OH]3–.The reverse reaction between [Fe(CN)5OH]3– and HPDTA4– is first-order in [Fe(CN)5OH]3– and HPDTA4–, and exhibits inverse first-order dependence on cyanide concentration.A six-step mechanism is proposed for the first stage of reaction, with the fifth step as rate determining.  相似文献   

3.
Photosubstitution by OH? ligand was concluded from a photochemical study of the [Cr(CN)6]3? and [Cr(CN)5OH]3? complexes in alkaline medium. Photoaccelerated aquation was found to proceed in the case of aquocyanochromates(III): [Cr(CN)5H2O]2? and [Cr(CN)3(H2O)3].  相似文献   

4.
Cyanide-bridged trinuclear heterometallic Ag(I)-Mn(III) complex {[Mn(TClPP)(H2O)]2[Ag(CN)2]}2 · 2Br · 2C3H6O · 3H2O (I) and ion-pair complex {[Mn(TClPP)(CH3OH)2][Ag(CN)2]} · 0.5H2O (II) have been synthesized with [Mn(TClTPP)(H2O)2]Br (H2TClTPP = meso-tetra(4-chlorophenyl)porphyrin) as assembling segment and K[Ag(CN)2] as building block by using different crystallization method. These two complexes have been characterized by elemental analysis, IR spectroscopy and X-ray structure determination. In the trinuclear complex I, [Ag(CN)2]? as bidentate ligand coordinates with the two central Mn(III) atom of [Mn(TClPP)(H2O)2]+ through its two trans cyanide groups to form the complex cation of [Mn(TClPP)(H2O)]2[Ag(CN)2]+, which further constructs the neutral complexes with the help of one Br? as balanced anion. For the ion-pair complex II composed by free [Mn(TClPP)(CH3OH)2]+ cation and free [Ag(CN)2]? anion, it can be linked into one-dimensional supramolecular structure with the dependence of the intermolecular O-H...N and O-H...O hydrogen bond interactions.  相似文献   

5.
A new three-dimensional diamagnetic metal nitronyl nitroxide radical coordination polymer with an aqua cadmium cyanide framework Cd(NIT4py)(H2O)Cd(CN)4·H2O 1, was synthesized. X-ray crystallography reveals that the structure consists of 3D aqua cadmium cyanide built up by octahedral CdII coordinating to NIT4py and tetrahedral CdII (CN)4 units. Magnetic measurements show that the χmT values are nearly constant at higher temperature. The lower χmT values at lower temperature are related to intermolecular antiferromagnetic interactions of the radicals, which arise due to the hydrogen bonded network (TN = 21 K).  相似文献   

6.
Surleva AR  Neshkova MT 《Talanta》2008,76(4):914-921
A new flow injection approach to total weak acid-dissociable (WAD) metal–cyanide complexes is proposed, which eliminates the need of a separation step (such as gas diffusion or pervaporation) prior to the detection. The cornerstone of the new methodology is based on the highly selective flow-injection potentiometric detection (FIPD) system that makes use of thin-layer electroplated silver chalcogenide ion-selective membranes of non-trivial composition and surface morphology: Ag2 + δSe1 − xTex and Ag2 + δSe. An inherent feature of the FIP-detectors is their specific response to the sum of simple CN + Zn(CN)42− + Cd(CN)42−. For total WAD cyanide determination, ligand exchange (LE) and a newly developed electrochemical pre-treatment procedure for release of the bound cyanide were used. The LE pre-treatment ensures complete recovery only when the sample does not contain Hg(CN)42−. This limitation is overcome by implementing electrochemical pre-treatment which liberates completely the bound WAD cyanide through cathodic reduction of the complexed metal ions. A complete recovery of toxic WAD cyanide is achieved in the concentration range from 156 μg L−1 up to 13 mg L−1. A three-step protocol for individual and group WAD cyanide speciation is proposed for the first time. The speciation protocol comprises three successive measurements: (i) of non-treated, (ii) LE-exchange pre-treated; (iii) electrochemically pre-treated sample. In the presence of all WAD complexes this procedure provides complete recovery of the total bound cyanide along with its quantitative differentiation into the following groups: (1) Hg(CN)42−; (2) CN + Cd(CN)42− + Zn(CN)42−; (3) Cu(CN)43− + Ni(CN)42− + Ag(CN)2. The presence of a 100-fold excess in total of the following ions: CO32−, SCN, NH4+, SO42− and Cl does not interferes. Thus the proposed approach offers a step ahead to meeting the ever increasing demand for cyanide-species-specific methods. The equipment simplicity makes the procedure a good candidate for implementing in portable devices for in-field cyanide monitoring.  相似文献   

7.
Summary The title reaction has been followed spectrophotometrically at 325 nm (max of [Mn(CN)6]3–) under pseudo-first order conditions with cyanide in a large excess at pH=10.0, I=0.1M (NaClO4) and 25°C. The reaction follows first-order kinetics in [MnEDTA(OH)]2– and exhibits variable-order dependence in [CN] one at high cyanide concentration, and two at low cyanide concentration. The product of above reaction has been identified as [Mn(CN)6]3–.The kinetics of the reverse reaction,i.e., the reaction of [Mn(CN)6]3- with EDTA4– have also been followed spectrophotometrically. This reactions is first-order with respect to both [Mn(CN) 6 3– ] and [EDTA4–] and exhibits an inverse first-order dependence on [CN]. A six-step mechanism has been proposed in which the penultimate step is rate-determining. The activation parameters have been obtained and support the postulated mechanism.  相似文献   

8.
Summary The Ru(phen)3(CN)2 · 6 H2O, Ru(bipy)3(CN)2 · 6H2O, Fe(phen)3(CN)2 · H2O and Ru(5-NO2P)3(CN)2 · 2 H2O compounds have been isolated during the reaction of the parent cations with aqueous cyanide solutions. It is evident, that in each case, attack at the ligand has taken placevia the cyanide nucleophile, though the equilibrium constant for the formation of the Reissert-type species are widely different. The implications of the findings with respect to the known reaction kinetics of the parent ions in aqueous cyanide solution are discussed.Part 14: R.D. Gillard and P.A. Williams,Transition Met. Chem., 2, 109(1977)  相似文献   

9.
By using the node‐and‐spacer approach in suitable solvents, four new heterotrimetallic 1D chain‐like compounds (that is, containing 3d–3d′–4f metal ions), {[Ni(L)Ln(NO3)2(H2O)Fe(Tp*)(CN)3] ? 2 CH3CN ? CH3OH}n (H2L=N,N′‐bis(3‐methoxysalicylidene)‐1,3‐diaminopropane, Tp*=hydridotris(3,5‐dimethylpyrazol‐1‐yl)borate; Ln=Gd ( 1 ), Dy ( 2 ), Tb ( 3 ), Nd ( 4 )), have been synthesized and structurally characterized. All of these compounds are made up of a neutral cyanide‐ and phenolate‐bridged heterotrimetallic chain, with a {? Fe? C?N? Ni(? O? Ln)? N?C? }n repeat unit. Within these chains, each [(Tp*)Fe(CN)3]? entity binds to the NiII ion of the [Ni(L)Ln(NO3)2(H2O)]+ motif through two of its three cyanide groups in a cis mode, whereas each [Ni(L)Ln(NO3)2(H2O)]+ unit is linked to two [(Tp*)Fe(CN)3]? ions through the NiII ion in a trans mode. In the [Ni(L)Ln(NO3)2(H2O)]+ unit, the NiII and LnIII ions are bridged to one other through two phenolic oxygen atoms of the ligand (L). Compounds 1 – 4 are rare examples of 1D cyanide‐ and phenolate‐bridged 3d–3d′–4f helical chain compounds. As expected, strong ferromagnetic interactions are observed between neighboring FeIII and NiII ions through a cyanide bridge and between neighboring NiII and LnIII (except for NdIII) ions through two phenolate bridges. Further magnetic studies show that all of these compounds exhibit single‐chain magnetic behavior. Compound 2 exhibits the highest effective energy barrier (58.2 K) for the reversal of magnetization in 3d/4d/5d–4f heterotrimetallic single‐chain magnets.  相似文献   

10.
The structure of aqueous lithium tetraborate solutions was investigated by species distribution calculation and synchrotron X-ray scattering. It shows that the dominant species in supersaturated solution at 298.15 K is B4O5(OH) 4 2? and the minor species are B3O3(OH) 5 2? , B3O3(OH) 4 ? and B(OH)3. The ‘intramolecular’ structural parameters of B4O5(OH) 4 2? , such as bond length and coordination number, were gives out using density function theory calculation. X-ray scattering study shows that the distance Li–O(H2O)I of [Li(H2O)4]+ is about 0.1983 nm with the coordination number(CN) 4 in tetrahedral configuration. The B–O(H2O) distance in hydrated anion B4O5(OH)4(OH2) 8 2? is 0.3662 nm with the CN 12. The Li+–B distance is about 0.3364 nm with a coordination number ~1.0. The temperature effect on solution structure was also discussed.  相似文献   

11.
The sodium cuprous cyanide salt, Na2[Cu(CN)3], has been adsorbed onto alumina, and the i.r. spectra, fast atom bombardment mass spectra and scanning electron micrographs of the reagent over a range of Na2[Cu(CN)3] loadings have been obtained. These show that, at high loadings, Na2[Cu(CN)3] is present on the surface in a crystalline form. At low loadings Na2[Cu(CN)3] is dispersed over the alumina surface, with the i.r. spectrum corresponding to that of aqueous [Cu(CN)3]−2. There appears to be good correlation in the mass spectrum between the percentage intensity of the sodium (m/z 23) peak relative to the sum of the Na+ and Al+ intensities, and the loading of Na2 [Cu(CN)3]. The scanning electron micrographs of the reagent at high loading clearly show crystalline material; at low loadings there is no difference in appearance between the reagent and uncoated alumina.  相似文献   

12.
When copper(II) acetate is treated with the ionic liquid n‐butylmethylimidazolium cyanide (BMIm‐CN), in ethanol solution, two new copper coordination compounds are obtained. (BMIm)2[Cu4(CN)7] comprises a 3D coordination polymer of cyanide bridged copper ions. This anionic coordination polymer contains CuI as well as CuII ions, i.e. it is a mixed‐valent compound. The polymer can be described as honeycomb structure with the BMIm+ cation being located in the cages. The second compound obtained from the chemical reaction is (BMIm)[Cu2(OAc)5][Cu(OAc)2(H2O)]2 · C2H5OH, which can be described as double‐salt. The first unit (BMIm)[Cu2(OAc)5] contains paddle wheel copper(II) acetato moieties, which are bridged by additional acetato ligands and form infinite chains. The second part of the double salt is the neutral, [Cu(OAc)2(H2O)]2 complex. These two parts as well as the co‐crystallized ethanol molecule are connected through a network of hydrogen bridges.  相似文献   

13.
Sulfur cyanide trifluoride, SF3CN, and sulfur dicyanide difluoride, SF2(CN)2, have been prepared by metathesis between sulfur tetrafluoride, SF4, and trimethyl silyl cyanide, (CH3)3SiCN, at – 30°C. Treatment of SF3CN with freshly sublimed selenium dioxide, SeO2, lead to sulfinyl cyanide fluoride, FS(O)CN. IR, Raman, 19F-NMR, uv and mass spectra of the novel compounds are presented as well as some physical and chemical properties.  相似文献   

14.
Substitution reactions take place following the photonic excitation of aqueous K4M(CN)8 (where M = Mo or W) in the presence of 1,10-phenanthroline and 2,2-?bipyridyl. Changes in absorbance with time show that the overall reaction is dependent on photochemical activation of potassium octacyanomolybdate(IV) and -tungstate(IV). The species [K2Mo(CN)4(OH)2(phen)], [K2W(CN)4(OH)2(phen)], [K2Mo(CN)4(OH)2(bipy)] and [K2W(CN)4(OH)2(bipy)] exist in solution. The final photosubstitution products [Mo(OH)3(CN)(phen)2] · 2H2O], [Mo(OH)3(CN)(bipy)2] · 3H2O, [W(OH)3(CN)(phen)2] · 2H2O and [W(OH)3(CN)(bipy)2] · H2O have been isolated in the solid state. Their IR spectra have been discussed. The quantum yield of the photosubstitution reactions has been determined and its variation with change of concentration of the complex as well as the H+ ion concentration has been studied.  相似文献   

15.
The mechanism of propene loss from protonated phenyl n-propyl ether and a series of mono-, di-, and trimethylphenyl n-propyl ethers has been examined by chemical ionization (CI) mass spectrometry in combination with tandem mass spectrometry experiments. The role of initial proton transfer to the oxygen atom and the aromatic ring, respectively, has been probed with the use of deuterated CI reagents, D2O, CD3OD, and CD3CN (given in order of increasing proton affinity), in combination with deuterium labeling of the β position of the n-propyl group or the phenyl ring. The metastable [M + D]+ ions of phenyl n-propyl ether—formed with D2O as the CI reagent—eliminate C3H5D and C3H6 in a ratio of 10:90, which indicates that the added deuteron is incorporated to a minor extent in the expelled neutral species. In the experiments with CD3OD as the CI reagent, the ratio between the losses of C3H5D and C3H6 from the metastable [M + D]+ ions of phenyl n-propyl ether is 18:82, whereas the ratio becomes 27:73 with CD3CN as the reagent. A similar trend in the tendency to expel a propene molecule that contains the added deuteron is observed for the metastable [M + D]+ ions of phenyl n-propyl ether labeled at the β position of the alkyl group. Incorporation of a hydrogen atom that originates from the aromatic ring in the expelled propene molecule is of negligible importance as revealed by the minor loss of C3H5D from the metastable [M + H]+ ions of C6D5OCH2CH2CH3 irrespective of whether H2O, CH3OH, or CH3CN is the CI reagent. The combined results for the [M + D]+ ions of phenyl n-propyl ether and deuterium-labeled analogs are suggested to be in line with a model that assumes that propene loss occurs not only from species formed by deuteron transfer to the oxygen atom, but also from ions generated by deuteron transfer to the ring. This is substantiated by the results for the methyl-substituted ethers, which reveal that the position as well as the number of methyl groups bonded to the ring exert a marked effect on the relative importances of the losses of C3H5D and C3H6 from the metastable [M + D]+ ions of the unlabeled methyl-substituted species.  相似文献   

16.
The hydroxamic acids (RC(O)NHOH, HA) exhibit diverse biological activity, including hypotensive properties associated with formation of nitroxyl (HNO) or nitric oxide (NO). Oxidation of two HAs, benzohydroxamic and acetohydroxamic acids (BHA, AHA) by [Fe(CN)5NH3]2? or [Fe(CN)6]3? was analyzed by spectroscopic, mass spectrometric techniques, and flow EPR measurements. Mixing BHA with both Fe(III) reactants at pH 11 allowed detecting the hydroxamate radical, (C6H5)C(O)NO˙?, as a one-electron oxidation product, as well as N2O as a final product. Successive UV–vis spectra of mixtures containing [Fe(CN)5NH3]2? (though not [Fe(CN)6]3?) at pH 11 and 7 revealed an intermediate acylnitroso-complex, [Fe(CN)5NOC(O)(C6H5)]3? (λmax, 465 nm, very stable at pH 7), formed through ligand interchange in the initially formed reduction product, [Fe(CN)5NH3]3?, and characterized by FTIR spectra through the stretching vibrations ν(CN), ν(CO), and ν(NO). Free acylnitroso derivatives, formed by alternative reaction paths of the hydroxamate radicals, hydrolyze forming RC(O)OH and HNO, postulated as precursor of N2O. Minor quantities of NO are formed only with an excess of oxidant. The intermediacy of HNO was confirmed through its identification as [Fe(CN)5(HNO)]3? (λmax, 445 nm) as a result of hydrolysis of [Fe(CN)5(NOC(O)(C6H5)]3? at pH 11. The results demonstrate that hydroxamic acids behave predominantly as HNO donors.  相似文献   

17.
Summary The kinetics and mechanism of the system [FeHIDA-(OH)2]+5CN[Fe(CN)5OH+HIDA2–+OH (HIDA=N-(2-hydroxyethyl) (iminodiacetate) at pH=9.5±0.02, I=0.1 M and at 25±0.1°C have been studied spectrophotometrically at 395 nm ( max of [Fe(CN)5OH]3–]. The reaction has three distinguishable stages; the first is formation of [Fe(CN)5OH]3–, the second is conversion of [Fe(CN)5OH]3– into [Fe(CN)6]3–, and last is the reduction of [Fe(CN)6]3– to [Fe(CN)6]4– by the HIDA2– released in the first stage. The first stage shows variable-order dependence on cyanide concentration, unity at high cyanide concentration and zero at low cyanide concentration. The second stage exhibits first-order dependence on the concentration of [Fe(CN)5OH]3– as well as on cyanide. The reverse reaction between [Fe(CN)5OH]3– and HIDA2– is first-order in each of these species and inverse first-order in cyanide. On the basis of forward and reverse rate studies, a five-step mechanism has been proposed for the first stage. The first step involves a slow loss of one OH, by a cyanide-independent path.  相似文献   

18.
Positive and negative cluster ions in methanol have been examined using a direct fast atom bombardment (FAB) probe technique. Positive ion (CH3OH)IIH + clusters with n = 1-28 have been observed and their clusters are the dominant ions in the low-mass region. Cluster-ion reaction products (CH3OH)II(H2O)H+ and (CH3OH)II(CH3OCH3)H+ are observed for a wide range of n and the abundances of these ions decrease with increasing n. The negative ion (CH3OH)II(CH3O)? clusters are also readily observed with n = 0-24 and these form the most-abundant negative ion series at low n. The (CH3OH)II(CH2O)?, (CH3OH)II(HIIO)(CH2O)? and (CH3OH)II(H2OXCH3O)? cluster ions are formed and the abundances of these ions approach those of the (CH3OH)II(CH3O)? ion series at high n. Cluster-ion structures and energetics have been examined using semi-empirical molecular orbital methods.  相似文献   

19.
[Co(CN)2(H3Cdta)] · 2H2O (I) and Na2[Co(CN)(Cdta)] · 3H2O (II) (Cdta4− is the 1,2-cyclohexanediaminetetraacetate) were synthesized. In I, the two cyanide ions and the Cdta nitrogen atoms are located in the equatorial plane, the carboxyl groups connected to different N atoms are in axial positions; in addition one of them, like two other uncoordinated carboxyl groups, is protonated. In II, Cdta is a pentadentate ligand coordinated in the cis-equatorial mode. UV excitation of the dicyano complexes at the ligand-to-metal charge transfer band gives organic derivatives of cobalt(III) stable enough for chromatographic isolation. Photolysis of II in neutral medium gives aqua(cyclohexanediaminetriacetato)cobalt (III) complexes as the final products.  相似文献   

20.
A variety of double collision experiments, whereby fast species undergo collisional interactions in two distinct regions of a mass spectrometer, are described. These include two-stage charge reversal of negative ions, two-stage double electron transfer from targets to cations, neutralization-reionization experiments as well as delayed analysis of organic cations formed in a one-step charge reversal of anions. Experiments have been performed on a number of systems of current interest in gas-phase ion chemistry. It is concluded that autoelectron detachment of benzyl anions leads to benzyl radicals, whereas the collisionally induced electron detachment produces a mixture of benzyl and tropyl radicals. By contrast, electron detachment from [H3CNH]? is not a metastable process and occurs only after excitation to produce H3CNH˙ radicals, which do not rearrange into the thermodynamically more stable H2CNH2˙. It is shown that in the double electron transfer reactions H+ + Xe→H˙ + Xe+˙ and H˙ + Xe→H? + Xe+˙, excited states are produced. From double collision experiments on methyl formate ions, it is concluded that the non-decomposing ions have undergone rearrangement on the time-scale of 10 μs into the distonic isomer, \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm H} - \mathop {\rm C}\limits^ + ({\rm OH}){\rm O}\mathop {\rm C}\limits^. {\rm H}_2 $\end{document}. Finally, it is shown that short-lived (<0.2 μs) [H2, C, N]+ ions generated by charge-reversal of [H2CN]? have the [H2CN]+ structure, whereas most of the long-lived (10 μs) ions have the [HCNH]+ structure.  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