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 703 毫秒
1.
The electrochemical behaviour of [Ir(bipy)2Cl2]+ and [Ir(phen)2Cl2]+ (bipy = 2,2′-bipyridine; phen = 1,10-phenanthroline) has been investigated in N,N-dimethylformamide (DMF). In potential sweep voltammetry [Ir(bipy)2Cl2]+ exhibits four reduction peaks. The first two processes involve one electron and are reversible in our conditions. The third reduction step is irreversible and has been attributed to the addition of three electrons to [Ir(bipy)2Cl2]+ followed by fast liberation of ligands. The data obtained for the fourth peak are consistent with a one-electron reversible process. The behaviour of [Ir(phen)2Cl2]+ is more complicated than that found for the bipy complex. In this case in fact, in addition to the four peaks observed in the case of the bipy complex, two other peaks appear. The latter have been attributed to the reduction of phen molecules liberated by the reduction of the complex. A qualitative MO discussion of the nature of the molecular levels involved in the reduction processes is also reported.  相似文献   

2.
On the Preparation of Dimercapto(methyl)Sulfonium Salts [CH3S(SH)2]+ AsF6? and [CH3S(SH)2]+SbCl6? and the Bis(chlorothio)methylsulfonium Salts [CH3S(SCI)2]+ AsF6? and [CH3S(SCI)2]+ SbCl6? The preparation of the dimercapto(methyl)sulfonium salts [CH3S(SH)2]+ AsF6? and [CH3S(SH)2]+SbCl6? from [CH3SCl2]+ salts and H2S at 195 K is reported. The salts are stable below 210 K. They are characterized by additional Raman spektroscopic measurements of the isotopic labelled cations [CH3S(SD)2]+, [CH3S(34SH)2]+ and [CH3S(34SD)2]+. The dimercapto(methyl)sulfonium salts are transfered into bis(chlorthio)methylsulfonium salts by reaction with Cl2 at 195 K.  相似文献   

3.
The title compound, Cs3[Cr(C2O4)3]·2H2O, has been synthesized for the first time and the spatial arrangement of the cations and anions is compared with those of the other members of the alkali metal series. The structure is built up of alternating layers of either the d or l enantiomers of [Cr(oxalate)3]3−. Of note is that the distribution of the [Cr(oxalate)3]3− enantiomers in the Li+, K+ and Rb+ tris(oxalato)chromates differs from those of the Na+ and Cs+ tris(oxalato)chromates, and also differs within the corresponding BEDT‐TTF [bis(ethylenedithio)tetrathiafulvalene] conducting salts. The use of tris(oxalato)chromate anions in the crystal engineering of BEDT‐TTF salts is discussed, wherein the salts can be paramagnetic superconductors, semiconductors or metallic proton conductors, depending on whether the counter‐cation is NH4+, H3O+, Li+, Na+, K+, Rb+ or Cs+. These materials can also be superconducting or semiconducting, depending on the spatial distribution of the d and l enantiomers of [Cr(oxalate)3]3−.  相似文献   

4.
A new complex compound, bis(2,2,2-cryptand potassium) tetrakis(isocyanato)cuprate(II), 2[K(Crypt-222)]+ [Cu(NCS)4]2? was prepared and its crystal structure was studied by X-ray structural analysis. The structure includes one symmetrically independent complex cation [K(Crypt-222)]+ of a guest-host type and independent one half of [Cu(NCS)4]2? anion. Through the center of the anion passes crystallographic symmetry axis 2, the approximate point symmetry of the anion is D 2, while the approximate point symmetry of the complex cation is D 3. The coordination polyhedron of the [Cu(NCS)4]2? anion (four N atoms) conjugated with Cu2+ cation is a nonplanar square considerably screwed into a flattened tetrahedron. The K+ cation (coordination number 8) of the complex cation [K(Crypt-222)]+ is coordinated by all eight heteroatoms (6O + 2N) of the 2,2,2-cryptand ligand, and its coordination polyhedron can be described as bis-basecentered trigonal prism slightly screwed into an anti-prism.  相似文献   

5.
The kinetics of oxidation of Fe2+ by [Co(C3H2O4)3]3? in acidic solutions at 605 nm showed a simple first-order dependence in each reactant concentration. The second-order rate constant dependence on [H+] is in accordance with eqn (i) k2 = k′2 + k3[H+] (i) where k′2 and k3 have values of 73.4 ± 14.0 M ?1 s?1 and 353 ± 41 M?2 s?1, respectively, at 1.0 M ionic strength (NaClO4) and 25°C. At 310 nm the formation and decomposition of an intermediate, believed to be [FeC3H2O4]+, was observed. The increase in the rate of oxidation with increasing [H+] was interpreted in terms of a “one-ended” dissociation mechanism which facilitates chelation of Fe2+ by the carbonyl oxygens of malonate in the transition state.  相似文献   

6.
CuCl or pre‐generated CuCF3 reacts with CF3SiMe3/KF in DMF in air to give [Cu(CF3)4]? quantitatively. [PPN]+, [Me4N]+, [Bu4N]+, [PhCH2NEt3]+, and [Ph4P]+ salts of [Cu(CF3)4]? were prepared and isolated spectroscopically and analytically pure in 82–99 % yield. X‐ray structures of the [PPN]+, [Me4N]+, [Bu4N]+, and [Ph4P]+ salts were determined. A new synthetic strategy with [Cu(CF3)4]? was demonstrated, involving the removal of one CF3? from the Cu atom in the presence of an incoming ligand. A novel CuIII complex [(bpy)Cu(CF3)3] was thus prepared and fully characterized, including by single‐crystal X‐ray diffraction. The bpy complex is highly fluxional in solution, the barrier to degenerate isomerization being only 2.3 kcal mol?1. An NPA study reveals a huge difference in the charge on the Cu atom in [Cu(CR3)4]? for R=F (+0.19) and R=H (+0.46), suggesting a higher electron density on Cu in the fluorinated complex.  相似文献   

7.
A spectroelectrochemical sensor was developed for [Re(dmpe)3]+ as a nonradioactive analog for [Tc(dmpe)3]+. The sensor consists of an optically transparent electrode (OTE) coated with a thin film of sulfonated polystyrene‐block‐poly(ethylene‐ran‐butylene)‐block‐polystyrene (SSEBS). Colorless [Re(dmpe)3]+ was reversibly oxidized to [Re(dmpe)3]2+ (λmax=530 nm). [Re(dmpe)3]+ preconcentrated by ion‐exchange into the SSEBS film, resulting in a 20‐fold increase in peak current compared to a bare OTE after 1 h of exposure to aqueous [Re(dmpe)3]+ solution. Detection of [Re(dmpe)3]+ at concentrations down to 2×10?6 M was accomplished by electrochemical modulation of the complex and monitoring absorbance by attenuated total reflectance (ATR).  相似文献   

8.
The systematic study of the reaction of M[PF6] salts and Me3SiCN led to a synthetic method for the synthesis and isolation of a series of salts containing the unprecedented [PF2(CN)4]? ion in good yields. The reaction temperature, pressure, and stoichiometry were optimized. The crystal structures of M[PF2(CN)4] (M=[nBu4N]+, Ag+, K+, Li+, H5O2+) were determined. X‐ray crystallography showed the exclusive formation of the cis isomer in accord with 31P and 19F solution NMR spectroscopy data. Starting with the K[PF2(CN)4] the room temperature ionic liquid EMIm[PF2(CN)4] was prepared exhibiting a rather low viscosity.  相似文献   

9.
Crystals of the novel title arsenic(III) phthalocyanine complex, [As(C32H16N8)]2[As4I14] or [(AsPc)+]2·[As4I14]2−, where Pc is phthalocyaninate(2−), have been obtained by the reaction of pure powdered As with phthalo­nitrile under a stream of iodine vapour at 493 K. The crystals are built up of separate but interacting [AsPc]+ cations and [As4I12]2− anions. The As atom of the [AsPc]+ unit is bonded to the four iso­indole N atoms of the Pc macrocycle and lies 0.743 (2) Å out of the plane defined by these four N atoms. The anionic part of the complex consists of AsI3 and [AsI4] units joined together into an [As4I14]2− anion. The arrangement of the oppositely charged moieties, [AsPc]+ and [As4I14]2−, in the crystal is determined mainly by ionic attraction and by donor–acceptor interactions between the [AsPc]+ and [As4I14]2− ions.  相似文献   

10.
Preparation of Trimercaptosulfonium Salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? The preparation of the trimercaptosulfonium salts [S(SH)3]+AsF6? and [S(SH)3]+SbCl6? from SCl3+ salts with excessive H2S at 193 K is reported. The [S(SH)3]+SbCl6? is transferred into [S(SCl)3]+SbCl6? by reaction with Cl2 at low temperatures. The new [S(SH)3]+ cation is isoelectronic to P(PH2)3. In addition, its existence is supported by an ab-initio calculation. The results show a potential well for C3v configuration with SH bonds bended towards the top of the pyramid for the isolated ion. Also the results of a force-field calculation are reported.  相似文献   

11.
The kinetics of Ruthenium(III) chloride mediated oxidation of acetone, 2-butanone, 4-methyl-2-pentanone, 2-pentanone, cyclopentanone, and cyclohexanone by sodium periodate in aqueous HClO4 media was zero-order in [IO4] and first-order in [ketone]. The reaction was independent of added [Ru(III)] and showed first-order dependence on [H+] for all the ketones studied, except acetone. In the case of acetone at [H+] < 0.05 M, the rate was independent of [H+], the order in [Ru(III)] being unity; but at [H+] > 0.05 M the reaction showed unit dependence on [H+] and the order in [Ru(III)] was zero. Ruthenium(VIII) generated in situ is postulated as the hydride abstracting species. A mechanism involving enolization as the rate determining step is proposed. Acetone at lower acidity of the medium is shown to react directly with Ru(VIII). In the absence of ruthenium(III) chloride, the kinetics were first-order in [IO4], [ketone], and [H+]. Structure-reactivity relationship is discussed and thermodynamic parameters are reported. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
In the title compound, [PtI(C15H11N3)][AuI2], the [PtI(terpy)]+ cations (terpy is 2,2′:6′,2′′‐terpyridine) stack in pairs about inversion centers through Pt...Pt interactions of 3.5279 (5) Å. The [AuI2] anions also exhibit pairwise stacking, with Au...I distances of 3.7713 (5) Å. The [PtI(terpy)]+ cations and [AuI2] anions aggregate forming infinite arrays of stacked ...({[PtI(terpy)]+...[PtI(terpy)]+}...{[AuI2]...[AuI2]})... units.  相似文献   

13.
Mass spectroscopic investigations of 2H- and 13C-labelled derivatives of e.g. 1-, 2-chloro-methyl-naphthalene and 1-chloro-phenyl(5)-pentene-(2)-in-(4) show that the [C11H9]+ ion, which gives [C9H7]+ by acetylene elimination, has the structure of a benztropylium cation. Model considerations show that the formation of this cation [C11H9]+ through a common transition state with a three-membered ring is very probable.  相似文献   

14.
The bis(ferrocenyl)phosphenium ion, [Fc2P]+, reported by Cowley et al. (J. Am. Chem. Soc. 1981 , 103, 714–715), was the only claimed donor‐free divalent phosphenium ion. Our examination of the molecular and electronic structure reveals that [Fc2P]+ possesses significant intramolecular Fe???P contacts, which are predominantly electrostatic and moderate the Lewis acidity. Nonetheless, [Fc2P]+ undergoes complex formation with the Lewis bases PPh3 and IPr to give the donor–acceptor complexes [Fc2P(PPh3)]+ and [Fc2P(IPr)]+ (IPr=1,3‐bis(2,6‐diisopropylphenyl)imidazole‐2‐ylidene).  相似文献   

15.
The complexation of manganese(II), cobalt(II) and nikel(II) with bromide ions has been studied in N,N-dimethylacetamide(DMA) by calorimetry and spectrophotometry. The formation of [MBr]+, [MBr2] and [MBr3] (M=Mn, Co, Ni) was revealed in all the metal systems. Interestingly, the complexation is significantly enhanced in DMA over N,N-dimethylformamide (DMF). This is unusual because physicochemical properties of DMA and DMF as solvent are similar. Furthermore, extracted electronic spectra of individual complexes of NiII suggested the presence of a geometry equilibrium, [NiBr(DMA)5]+=[NiBr(DMA)4]++ DMA, in DMA. A similar geometry equilibrium is also suggested, [NiBr2(DMA)3]=[NiBr2(DMA)2]+DMA. Such geometry equilibria were not observed in DMF. With regard to cobalt(II), electronic spectra show the presence of the four-coordinated [CoBr(DMA)3]+ complex in DMA, unlike the six-coordinated [CoBr(DMF)5]+ one in DMF. These facts suggest that a specific strong steric interaction operates between coordinating solvent molecules, which plays a key role in the complexation behavior of the divalent transition metal ions in DMA.  相似文献   

16.
The title compounds are salts of the general form (Q+)2[Zn(dmit)2]2?, where dmit corresponds to the ligand (C3S5)? present in both and Q+ to the counter‐cations (nBu4N)+ [or C16H36N+] and (Ph4As)+ [or C24H20As+], respectively. In the first case, Zn is in the 4e special positions of space group C2/c and hence the [Zn(dmit)2]2? dianion possesses twofold axial crystallographic symmetry. Including these, there are now 11 known examples of [Zn(dmit)2]2? or its analogues, with O replacing the exocyclic thione S, and [Zn(dmio)2]2? dianions in nine structures with various Q. Comparison of these reveals a remarkable variation in details of the conformation which the dianion may adopt in terms of Zn coordination, equivalence of the Zn—S bond lengths, displacement of Zn from the plane of the ligand and overall dianion shape.  相似文献   

17.
The kinetics of oxidation of PdII by CeIV have been studied spectrophotometrically in HClO4 media at 40 °C. The reaction is first order each in [CeIV] and [PdII] at constant [H+]. Increasing [H+] accelerates the reaction rate with fractional order in [H+]. The initially added products, palladium(IV) and cerium(III) do not have any significant effect on the reaction rate. At constant acidity, increasing the added chloride concentration enhances the rate of reaction. H3Ce(SO4)4 and PdCl42− are the active species of oxidant and reductant respectively. The possible mechanisms are proposed and the reaction constants involved have been determined.  相似文献   

18.
The mass spectra of a series of β-ketosilanes, p-Y? C6H4Me2SiCH2C(O)Me and their isomeric silyl enol ethers, p-Y? C6H4Me2SiOC(CH3)?CH2, where Y = H, Me, MeO, Cl, F and CF3, have been recorded. The fragmentation patterns for the β-ketosilanes are very similar to those of their silyl enol ether counterparts. The seven major primary fragment ions are [M? Me·]+, [M? C6H4Y·]+, [M? Me2SiO]+˙, [M? C3H4]+˙, [M? HC?CCF3]+˙, [Me2SiOH]+˙ and [C3H6O]+˙ Apparently, upon electron bombardment the β-ketosilanes must undergo rearrangement to an ion structure very similar to that of the ionized silyl enol ethers followed by unimolecular ion decompositions. Substitutions on the benzene ring show a significant effect on the formation of the ions [M? Me2SiO]+˙ and [Me2SiOH]+˙, electron donating groups favoring the former and electron withdrawing groups favoring the latter. The mass spectral fragmentation pathways were identified by observing metastable peaks, metastable ion mass spectra and ion kinetic energy spectra.  相似文献   

19.
The tetraoxido ruthenium(VIII) radical cation, [RuO4]+, should be a strong oxidizing agent, but has been difficult to produce and investigate so far. In our X-ray absorption spectroscopy study, in combination with quantum-chemical calculations, we show that [RuO4]+, produced via oxidation of ruthenium cations by ozone in the gas phase, forms the oxygen-centered radical ground state. The oxygen-centered radical character of [RuO4]+ is identified by the chemical shift at the ruthenium M3 edge, indicative of ruthenium(VIII), and by the presence of a characteristic low-energy transition at the oxygen K edge, involving an oxygen-centered singly-occupied molecular orbital, which is suppressed when the oxygen-centered radical is quenched by hydrogenation of [RuO4]+ to the closed-shell [RuO4H]+ ion. Hydrogen-atom abstraction from methane is calculated to be only slightly less exothermic for [RuO4]+ than for [OsO4]+.  相似文献   

20.
The products of the reaction between the electrophilic alkenylxenonium cation [1-Xe+–C6F9] and the halide anions I?, Br?, Cl? and F? depend on the hardness of the halide anion. With the soft halides I? and Br? Xe(II) is formally displaced by halogen as well in basic MeCN as in superacidic (AHF1), whereas with hard fluoride and chloride no reaction takes place in AHF. In MeCN F? initiates the formation of alkenyl radicals, which abstract hydrogen from the solvent, whereas Cl? exhibits borderline character: RH and RCl formation. Possible reaction paths are discussed. The reactivity of the arylxenonium cation [C6F5Xe]+ in AHF toward halide ions is reported and the relative electrophilicity of the cations [C6F5Xe]+ and [1-Xe+–C6F9] is determined by the competitive reaction with Cl?. In addition the synthesis of cyclohexene 1-CF3–C6F9 from C6F5CF3 and XeF2 is performed and its electrophilicity is compared with that of the aromatic compound C6F5CF3.  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