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.
When HCrO4 ? is reduced by formate in solutions buffered by 2-ethyl-2-hydroxybutanoic acid and its anion, chelated complexes of both Cr(IV) and Cr(V), both of them stabilized in the medium used, are formed. It appears that Cr(V) is not generated directly from the Cr(VI)-formate reaction but arises instead from oxidation of Cr(IV) by Cr(VI). When the Cr(VI)-formate reaction is allowed to go to completion in the presence of [Cl(NH3)5Co]2+, a scavenger for Cr(II), 84–86% of the Cr(VI) taken is found to be converted to Cr(II), indicating that nearly all of the reacting system proceeds through Cr(IV) and bypasses the more usual state Cr(III). Initial rates for formation of Cr(IV) lead to a rate law pointing to a transition state containing the two redox partners, two ligating carboxyl groups, and two units of H+. Substitution of DCO2 ? for HCO2 ? retards formation of Cr(IV) by a factor of 3.3, whereas the solvent isotope affect, rateD 2O/rateH 2O, favors the deuterated system by a factor of 1.4. Our observations are in accord with a sequence initiated by the ligation of HCrO4 ? to a chelate derived from the buffering carboxylate anion. Conversions of Cr(VI) to Cr(IV), and Cr(IV) to Cr(II) appear to entail hydride shifts from formate to the Cr(=O) function.  相似文献   

2.
The bis(chelated) complex of CrV(0) derived from the dianion (L2 ) of 2-ethyl-2-hydroxybutanoic acid is readily reduced to a bis(chelate of CrIII, featuring the monoanion (LH) [Cr V(0)(L2−)2]+4H++H2O+2e→[CrIII(OH2)2(LH 2]+ of this acid. Potentials estimated by Ghosh in 1993 for this 2e change, E0 (pH 0) 1.32 V, Eeff (pH 3.3) 0.93 V, are in accord with the nearly irreversible reductions of the Cr(V) species (in 1∶1 ligand buffer) by Fe2+, V02+, IrCl6 3 and I, whereas lower values reported by Bose in 1996, E0 (pH 0) 0.84 V, Eeff (pH 3.3) 0.45 V, are potentiometrically inconsistent with these conversions. A similar discrepancy is noted for potentials for Cr(V,IV) estimated in 1996, E0 (pH 0) 0.84 V, Eeff (pH 3.3) 0.46 V, which, wholly contrary to observation, predict that the reductions of excess Cr(V) to CR(IV) by Fe2+, V02+, and I are thermodynamically disfavored.  相似文献   

3.
The kinetics of oxidation of diaquadichloro(1,10-phenanthroline)chromium(III) complex, [CrIII(phen)(H2O)2Cl2]+, by N-bromosuccinimide (NBS) is biphasic. The first faster step involves the oxidation of Cr(III) to Cr(IV). The second slower step is due to the oxidation of Cr(IV) to Cr(V). The reaction product is isolated and characterized by electron spin resonance (ESR), IR, and elemental analysis. The chromium(V) product is consistent with the formula [CrV(phen)Cl2(O)]Br. The rate constants kf and ks, for the faster and the slower steps respectively, were obtained using an Origin 9.0 software program. Values of both kf and ks, varied linearly with [NBS] at constant reaction conditions. The effect of pH on the reaction rate is investigated over the pH (4.11–6.01) range at 25.0°C. The rate constants kf and ks increased with increasing pH. This is consistent with hydroxo forms of the chromium species being more reactive than the aqua forms. Chromium(III) complexes, more often than not, are inert. The oxidation of the Cr(III) complex to Cr(IV), most likely, proceeds by an outer sphere mechanism. Since chromium(IV) is labile the mechanism of its oxidation to chromium(V) is not certain.  相似文献   

4.
The kinetic characteristics of the concentrated Ⅴ(Ⅳ)/Ⅴ(Ⅴ) couple have been studied at a glassy carbon electrode in sulfuric acid using rotating-disc electrode and cyclic voltammetry. The kinetics of the Ⅴ(Ⅳ)/Ⅴ(Ⅴ) redox couple reaction was found to be electrochemically quasi-reversible with the slower kinetics for the Ⅴ(Ⅴ) reduction than that for the Ⅴ(Ⅳ) oxidation. And, dependence of diffusion coefficients and kinetic parameters of Ⅴ(Ⅳ) species on the Ⅴ(Ⅳ) and H2SO4 concentration was investigated. It is shown that the concentration of active species Ⅴ(Ⅳ) should be over 1 mol·L^-1 for the redox flow battery application. Further, with increasing the Ⅴ(Ⅳ) and H2SO4 concentration, the diffusion coefficients of Ⅴ(Ⅳ) were gradually reduced whereas its kinetics was improved considerably, especially in the case of Ⅴ(Ⅳ) and H2SO4 up to 2 and 4 mol·L^-1.  相似文献   

5.
The asymmetric molybdenum(VI) dioxo complexes of the bis(phenolate) ligands 1,4‐bis(2‐hydroxybenzyl)‐1,4‐diazepane, 1,4‐bis(2‐hydroxy‐4‐methylbenzyl)‐1,4‐diazepane, 1,4‐bis(2‐hydroxy‐3,5‐dimethylbenzyl)‐1,4‐diazepane, 1,4‐bis(2‐hydroxy‐3,5‐di‐tert‐butylbenzyl)‐1,4‐diazepane, 1,4‐bis(2‐hydroxy‐4‐flurobenzyl)‐1,4‐diazepane, and 1,4‐bis(2‐hydroxy‐4‐chlorobenzyl)‐1,4‐diazepane (H2(L1)–H2(L6), respectively) have been isolated and studied as functional models for molybdenum oxotransferase enzymes. These complexes have been characterized as asymmetric complexes of type [MoO2(L)] 1–6 by using NMR spectroscopy, mass spectrometry, elemental analysis, and electrochemical methods. The molecular structures of [MoO2(L)] 1–4 have been successfully determined by single‐crystal X‐ray diffraction analyses, which show them to exhibit a distorted octahedral coordination geometry around molybdenum(VI) in an asymmetrical cis‐β configuration. The Mo? Ooxo bond lengths differ only by ≈0.01 Å. Complexes 1 , 2 , 5 , and 6 exhibit two successive MoVI/MoV (E1/2, ?1.141 to ?1.848 V) and MoV/MoIV (E1/2, ?1.531 to ?2.114 V) redox processes. However, only the MoVI/MoV redox couple was observed for 3 and 4 , suggesting that the subsequent reduction of the molybdenum(V) species is difficult. Complexes 1 , 2 , 5 , and 6 elicit efficient catalytic oxygen‐atom transfer (OAT) from dimethylsulfoxide (DMSO) to PMe3 at 65 °C at a significantly faster rate than the symmetric molybdenum(VI) complexes of the analogous linear bis(phenolate) ligands known so far to exhibit OAT reactions at a higher temperature (130 °C). However, complexes 3 and 4 fail to perform the OAT reaction from DMSO to PMe3 at 65 °C. DFT/B3LYP calculations on the OAT mechanism reveal a strong trans effect.  相似文献   

6.
Oxophthalocyaninato(2–)molybdenum(IV), activated by bromine oxidation prior to use, reacts with fused triphenylphosphine in the presence of bis(triphenylphosphine)iminium bromide to yield linear-bis(triphenylphosphine)iminium trans-dibromophthalocyaninato(2–)molybdate(III), l(PNP)trans[Mo(Br)2pc2?]. It crystallizes triclinic with crystal data: a = 10.506(1) Å, b = 12.436(2) Å, c = 12.918(2) Å, α = 76.186(1)°, β = 67.890(1)°, γ = 68.689(1)°; space group P1 (No. 2); Z = 1. MoIII is in a pseudo-octahedral coordination geometry with the bromo ligands in trans-arrangement. The Mo? Np and Mo? Br distance is 2.043(10) and 2.588(1) Å, respectively. The PNP cation adopts a linear conformation. In the IR spectrum vas(Mo? Br) is observed at 218 cm?1 and vas(P? N) of the linear (P? N? P) core at 1406 cm?1. Cyclic and differential-pulse voltammetry show two quasi-reversible cathodic processes at ?1.15 and ?0.53 V vs. Ag/AgCl. The first is assigned to a phthalocyaninate directed reduction (pc2?/pc3?), while the latter arises from a Mo directed reduction (MoIII/MoII). Spectral monitoring confirms the reversible MoIII/MoII reduction. Two quasi-reversible anodic processes at 0.60 and 1.27 V are assigned to the successive Mo directed oxidation with redox couples MoIII/MoIV and MoIV/MoV. For the first time, three very intense spin-allowed trip-quartet transitions are observed in the electronic absorption spectra at 7140 (TQI), 16890 (TQ2) and 18700 cm?1 (TQ3) together with a sing-quartet transition at 15850 cm?1 and characteristic ?Q”? region with maximum at 28500 cm?1 and ?N”? region at 37400 cm?1. All electronic excitations are of comparable intensity. A prominent low temperature emission at 6690 cm?1 is assigned to a spin-forbidden trip-sextet.  相似文献   

7.
Na superionic conductor (NASICON) structured cathode materials with robust structural stability and large Na+ diffusion channels have aroused great interest in sodium-ion batteries (SIBs). However, most of NASICON-type cathode materials exhibit redox reaction of no more than three electrons per formula, which strictly limits capacity and energy density. Herein, a series of NASICON-type Na3+xMnTi1−xVx(PO4)3 cathode materials are designed, which demonstrate not only a multi-electron reaction but also high voltage platform. With five redox couples from V5+/4+ (≈4.1 V), Mn4+/3+ (≈4.0 V), Mn3+/2+ (≈3.6 V), V4+/3+ (≈3.4 V), and Ti4+/3+ (≈2.1 V), the optimized material, Na3.2MnTi0.8V0.2(PO4)3, realizes a reversible 3.2-electron redox reaction, enabling a high discharge capacity (172.5 mAh g−1) and an ultrahigh energy density (527.2 Wh kg−1). This work sheds light on the rational construction of NASICON-type cathode materials with multi-electron redox reaction for high-energy SIBs.  相似文献   

8.
The redox ability of isolated Mo5+ cations in model MoH-beta and MoH-ZSM-5 systems is studied byin situ ESR. The oxidation of reduced samples by nitrogen monoxide at 20°C proceeds much faster than oxidation by oxygen. The interaction between Mo5+ ions and propene at 20°C results in the formation of a complex in which the oxidizability of Mo(V) by NO is substantially enhanced. UV irradiation increases the oxidation rate for Mo5+ ions in HZSM-5 by nitrogen monoxide at 20°C, indicating the possibility for the photochemical activation of the process. The step of active site reduction during the interaction of the samples with the H2 + NO + He gas mixtures of various compositions at 500°C is fast, and the dynamic equilibrium of the redox Mo(V) ai Mo(VI) reaction is shifted to the left Deceased.  相似文献   

9.
Summary A binuclear oxo MoV hypophosphite of composition [Mo2O4(H2PO2)2(H2O)2], is prepared by direct reduction of MoVI oxide hydrate (MoO3·H2O) with hypophosphorus acid in an argon atmosphere, and characterised by i.r., and electronic spectra, magnetic susceptibility and cyclic voltammetry measurements.1H and31Pn.m.r., x-ray diffraction and thermal analysis data contribute to its molecular structure elucidation, and a dioxobridged dioxo MoV with bidentate hypophosphite ion and water molecule completing the octahedral coordination around each Mo atom is proposed.  相似文献   

10.
The oxidation of cis‐diaquabis(1,10‐phenanthroline)chromium(III) [cis‐CrIII(phen)2(H2O)2]3+ by ‐bromosuccinimide (NBS) to yield cis‐dioxobis(1,10‐phenanthroline)chromium(V) has been studied spectrophotometrically in the pH 1.57–3.56 and 5.68–6.68 ranges at 25.0°C. The reaction displayed biphasic kinetics at pH < 4.0 and a simple first order at the pH > 5.0. In the low pH range, the reaction proceeds by two successive steps; the first faster step corresponds to the oxidation of Cr(III) to Cr(IV), and the second slower one corresponds to the oxidation of Cr(IV) to Cr(V), the final product of the reaction. The formation of both Cr(IV) and Cr(V) has been detected by electron spin resonance (ESR). The ESR clearly showed the formation and decay of Cr(IV) as well as the formation of Cr(V). Each oxidation process exhibited a first‐order dependence on the initial [Cr(III)]. The pseudo–first‐order rate constants k34 and k45, for the faster and slower steps, respectively, were obtained by a computer program using Origin7.0. Both rate constants showed first‐order dependence on [NBS] and increased with increasing pH.  相似文献   

11.
The solvothermal reaction of [ReOCl3(PPh3)2] with 2,2′-seleno-bis(4,6-di-tert-butylphenol) (H2L1) in ethanol at 95 °C resulted in an oxorhenium(V) complex of formulation [ReO(L1)(L2)] due to the in situ formation of 2-selenochloromethyl-4,6-di-tert-butylphenolate (L2) through the cleavage of one C–Se bond of H2L1. The mononuclear oxorhenium(V) complex was characterized by physicochemical and spectroscopic methods along with detailed structural analysis by single crystal X-ray crystallography. Electrochemical studies revealed a one-electron equivalent quasi-reversible voltammogram for the ReO(V)/ReO(VI) redox couple at 1.28 V (versus SCE) with no cathodic response for ReO(V) → ReO(IV) reduction.  相似文献   

12.
Two new reduced molybdenum pyrophosphates, Na28[Na2{(Mo2O4)10(P2O7)10(HCOO)10}]·108H2O ( 1 ) and Na22(H3O)2[Na4{(Mo2O4)10(P2O7)10(CH3COO)8(H2O)4}]·91H2O ( 2 ) have been synthesized and characterized by single‐crystal X‐ray diffraction. Red crystals of 1 are triclinic, space group , with a = 17.946(4) Å, b = 18.118(4) Å, c = 21.579(4) Å, α = 114.47(3)°, β = 93.54(3)°, γ = 114.39(3)° and V = 5581.8(19) Å3, and orange crystals of 2 are monoclinic, space group P21/n, with a = 21.467(4) Å, b = 23.146(5) Å, c = 24.069(5) Å, β = 101.76(3)° and V = 11708(4) Å3. They are both constructed by MoV dimers ({Mo2O4(OP)4(HCOO)} in 1 , {Mo2O4(OP)4(CH3COO)} and {Mo2O4(OP)4(H2O)2} in 2 ) and pyrophosphoric groups. Their structures can be described as two interconnected nonequivalent wheels which are approximately perpendicular, delimiting a large cavity. The larger wheel contains six MoV dimers, while the smaller one has four dimers.  相似文献   

13.
In the title family the tridentate ONO donor ligands are the fully deprotonated forms of acetylhydrazones of 2-hydroxybenzaldehyde (H2L1) and 2-hydroxyacetophenone (H2L2) (general abbreviation H2L), while bidentate mononegative OO donor ligands are the deprotonated salicylaldehyde (Hsal), vanillin (Hvan) and monodeprotonated 1,2-ethanediol (H2ed) (general abbreviation HB). The reaction of VIVO(acac)2 with H2L and Hsal or Hvan in equimolar ratio in MeOH afforded the complexes of the type [VVO(L)(B)], (1)–(4). The reaction of VIVO(acac)2 with H2L1 (in an equimolar ratio) and an excess of H2ed in MeOH yielded the complex [VVO(L1)(Hed)], (5) but the similar reaction with H2L2 ligand failed to produce such a type of complex. Complexes have been characterized by elemental analyses and by i.r., n.m.r. and u.v.-vis. spectroscopies. All the complexes are diamagnetic and display only LMCT bands. 1H-n.m.r. spectral data indicate that complexes (1)–(4) exist in two isomeric forms [(1A), (1B); (2A), (2B); (3A), (3B) and (4A), (4B)] in different ratios in CDCI3 solution. Complexes (1)–(4) display a quasi-reversible one electron reduction peak in the −0.06 to +0.05 V versuss.c.e. region in CH2CI2 solution and (5) displays an irreversible reduction peak at −0.46 V versuss.c.e. in DMF solution. The trend in the redox potential values has been correlated with the basicity of both the primary and auxiliary ligands.  相似文献   

14.
The title centrosymmetric cluster octakis(4-iso­propyl­pyridine-N)-di-μ4-oxo-hexa-μ3-oxo-octa-μ2-oxo-deca­oxo­octa­molyb­denum(V)­dimolybdenum(VI), [Mo10O26(C8H11N)8], consists of ten Mo atoms connected together by bridging oxo groups. Pentavalent Mo atoms are linked into four Mo2V pairs by metal–metal single bonds with lengths of 2.5637 (6) and 2.6132 (6) Å.  相似文献   

15.
Oxidation of quadruply bonded metal-metal dimers in the presence of good π-accepting ligands results in the formation of MoV---MoV compounds of the type [MO2(μ-X)2(Y)(Y′)]2+ (X = O or S; Y,Y′ = O,O; S,S; O,S). Reaction of MO2(O2CCH3)4 with oxygen in the presence of Na2mnt (mnt = 1,2-dicyanoethylene-2,2-dithiolate) gives [MO2(μ-S)2(O)(S)(mnt)2]2− (1). The compound crystallizes in the monoclinic space group P21/c, with cell dimensions a = 19.547(4), b = 15.210(4), c = 18.754(6) Å, β = 101.69(2)°, V= 5460(2) Å3, and Z = 4. Similarly, oxidation of o-dichlorobenzene solutions of Mo2Cl4(CH3CN)4 and 4,4′-dimethyl-2,2′-dipyridyl (dmpby) or, more directly, the reaction of Mo2Cl4(dmbpy)2 with oxygen leads to the formation of a red solid, which was characterized by X-ray crystallography to be Mo2(μ-O)2(O)2(Cl)2(dmbpy)2 (2). Red diamond crystals, prepared by slow evaporation of CH3CN solutions of 2, are trigonal and in the space group P3121 with cell dimensions a = 16.135(4), b = 16.135(4), c = 10.709(3) Å, V = 2414.4(13) Å3 and Z = 3. In both structures, the geometry about each of the molybdenum atoms is a distorted square pyramid with terminal oxygen or sulphur atoms at the apices and in a syn conformation. The molybdenum-molybdenum bond distances of 2.858(1) Å and 2.562(2) Å in structures of 1 and 2, respectively, are typical of other MoV---MoV dimers and indicative of a single Mo---Mo bond.  相似文献   

16.
Molybdenum(V) and (VI) and Niobium (V) μ-Oxalato Complexes The complexation of the « oxalato » ligand as bis-(bidentate) bridge in the solid compounds {(C2H5)4N}2 {M2VO2X6(C2O4)} or {(C2H5)4N}2 {MVIO4X4(C2O4)} (with MV = Mo or Nb; MVI = Mo and X = F, Cl, Br) is characterized on the basis of their infrared spectra and analytical data. The behaviour, of the MoV dimer compounds in near corresponding mononuclear species at room temperature, because of the low interaction between metallic ions.  相似文献   

17.
The solid solutions (V1–xWx)OPO4 with β‐VOPO4 structure type (0.0 ≤ x ≤ 0.01) and αII‐VOPO4 structure type (0.04 ≤ x ≤ 0.26) were obtained from mixtures of VVOPO4 and WVOPO4 by conventional solid state reactions and by solution combustion synthesis. Single crystals of up to 3 mm edge length were obtained by chemical vapor transport (CVT) (800 → 700 °C, Cl2 as a transporting agent). Single crystal structure refinements of crystals at x = 0.10 [a = 6.0503(2) Å, c = 4.3618(4) Å, R1 = 0.021, wR2 = 0.058, 21 parameters, 344 independent reflections] and x = 0.26 [a = 6.0979(2) Å, c = 4.2995(1) Å, R1 = 0.030, wR2 = 0.081, 21 parameters, 346 independent reflections] confirm the αII‐VOPO4 structure type (P4/n, Z = 2) with mixed occupancy V/W for the metal site. Due to the specific redox behavior of W5+ and V5+, solid solutions (V1–xWx)OPO4 should be formulated as (VIVxVV1–2xWVIx)OPO4. The valence states of vanadium and tungsten are confirmed by XPS measurements. V4+ with d1 configuration was identified by EPR spectroscopy and magnetic measurements. Electronic spectra of the solid solutions show the IVCT(V4+ → V5+) and the LMCT(O2– → V5+). (V0.74W0.26)OPO4 powders exhibit semi‐conducting behavior (Eg = 0.7 eV).  相似文献   

18.
Reduction of carboxylato bound chromium(V) by several reducing agents such as N2H4, NH2OH, Ti(III), Fe(II), VO2+, U(IV) formed part of earlier studies in which electron transfer to Cr(V) seems to occur through nitrogen or by inner or outer sphere path. In the chromium(V) oxidation of lactic acid, saturation kinetics has been observed and the association constant,K, evaluated from a double reciprocal plot, indicates weak interaction between reactants. In contrast, the reduction of Cr(V) by thiolactic acid exhibits total second order kinetics-first order with respect to each reactant. Probably the Crv-S bond is weaker than the Crv-O bond and electron transfer mediated through the sulphur system is facile. Under identical conditions Cr(VI) seems to be a better oxidant for lactic than for thiolactic acid.  相似文献   

19.
Summary Complex reactions between MoV.VI ando-hydroxybenzylamine-N,N,O-triacetic acid (HBATA) have been investigated in the 1–3 and 2.8–6.5 pH range by potentiometric titration at 30° C in 0.5 mol dm–3 NaCl. The equilibrium data were analyzed with the SCOGS2 and MINIQUAD programs, taking into account side reactions of MoV.VI and HBATA with hydrogen ion. The favorable reaction model comprises two complexes, (1,1,1)+ and (1,2,2), with formation constants log 111 = 14.85 ± 0.11 and log 122 = 28.51 ± 0.08 for the MoV-HBATA system and the two complexes (1,1,2)3– and (1,1,3)2– with formation constants log 112 = 17.36 ± 0.01 and log 113 = 20.60 ± 0.01 for the MoVI-HBATA system. The numbers in brackets refer to the chemical stoichiometric coefficients of molybdenum, HBATA and hydrogen ion in the complexes. The structure and coordinating behaviours of MoV and MoVI complexes are discussed. The equilibria studied for the polymerization of MoV indicates that dimeric, trimeric and tetrameric species are present at pH 1–3.  相似文献   

20.
《Electroanalysis》2017,29(4):1056-1061
Functionalized high purity carbon nanotubes (CNTs) with various amounts of oxygen containing surface groups were investigated towards the relevant redox reactions of the all‐vanadium redox flow battery. The quinone/hydroquinone redox peaks between 0.0 and 0.7 V vs. Ag|AgCl|KClsat. were used to quantifying the degree of functionalization and correlated to XPS results. Cyclic voltammetry in vanadyl sulfate‐containing 3 M H2SO4 as a common supporting electrolyte showed no influence of the amount of surface groups on the V(IV)/V(V) redox system. In contrast, the reactions occurring at the negative electrode (V(II)/V(III) and V(III)/V(IV)) are strongly affected by oxygen surface groups. However, under modified experimental conditions, SECM experiments detecting the consumption of VO2+ molecules by CNT thin films in pH=2 solution show improved onset potentials with increased surface oxygen content up to ∼ 3 at%. Further increase in surface oxygen up to 8 at% led to minor improvement. These dissimilar results under different experimental conditions are rationalized by suggesting that oxygen functional groups do not form the active site for the V(IV)/V(V) reaction but wetting of the catalyst layer is of high importance.  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