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 46 毫秒
1.
The new thiostannate Na4Sn2S6 was prepared by directed crystal water removal from the hydrate Na4Sn2S6 ⋅ 5H2O at moderate temperatures. While the structure of the hydrate comprises isolated [Sn2S6]4− anions, that of the anhydrate contains linear chains composed of corner-sharing SnS4 tetrahedra, a structural motif not known in thiostannate chemistry. This structural rearrangement requires bond-breakage in the [Sn2S6]4− anion, movements of the fragments of the opened [Sn2S6]4− anion and Sn−S−Sn bond formation. Simultaneously, the coordination environment of the Na+ cations is significantly altered and the in situ formed NaS5 polyhedra are joined by corner- and edge-sharing to form a six-membered ring. Time-dependent in situ X-ray powder diffraction evidences very fast rehydration into Na4Sn2S6 ⋅ 5H2O during storage in air atmosphere, but recovery of the initial crystallinity requires several days. Impedance spectroscopy demonstrates a mediocre room-temperature Na+ ion conductivity of 0.31 μS cm−1 and an activation energy for ionic transport of Ea=0.75 eV.  相似文献   

2.
Na4P2S7, Na2FeP2S7, and Ag4P2S7 were prepared by elemental synthesis at high temperatures and were characterized by vibration spectra and differential thermal analysis (DTA). A normal coordinate analysis was performed for P2S4−7. Additional vibration frequencies indicate the presence of the decathiotriphosphate anion P3S5−10. The formation of higher thiophosphates of the type PnS(n+2)-3n+1 with n ≥ 4 cannot be excluded.  相似文献   

3.
The title compound, 2C14H13N2+·S2O82−·2H2O, is a protonated amine salt which is formed from two rather uncommon ionic species, namely a peroxodisulfate (pds2−) anion, which lies across a crystallographic inversion centre, and a 2,9‐dimethyl‐1,10‐phenanthrolin‐1‐ium (Hdmph+) cation lying in a general position. Each pds2− anion binds to two water molecules through strong water–peroxo O—H...O interactions, giving rise to an unprecedented planar network of hydrogen‐bonded macrocycles which run parallel to (100). The atoms of the large R88(30) rings are provided by four water molecules bridging in fully extended form (...H—O—H...) and four pds2− anions alternately acting as long (...O—S—O—O—S—O...) and short (...O—S—O...) bridges. The Hdmph+ cations, in turn, bind to these units through hydrogen bonds involving their protonated N atoms. In addition, the crystal structure also contains π–π and aromatic–peroxo C—H...O interactions.  相似文献   

4.
Six new thioantimonates(III) with the [Sb4S7]2− anion were obtained under solvothermal conditions with in‐situ formed transition metal complexes as structure directors. In the two isostructural compounds [Fe(dien)2]Sb4S7·H2O ( 1 ) and [Co(dien)2]Sb4S7·0.5 H2O ( 2 ) (dien = diethylenetriamine; space group: P21/c) the layered [Sb4S7]2− anion is characterized by Sb8S8 rings with a diameter of about 9.6·7.6Å. The cation complexes are located above and below the pores of the rings. Despite the larger size of the cation complex the network topology of the third thioantimonate [Ni(dien)(tren)]Sb4S7 ( 3 ) (tren = tris(2‐aminoethyl)amine; space group: P21/n) is similar to that of the first two compounds. In the isostructural thioantimonates [M(trien)]Sb4S7 (M = Zn ( 4 ); M = Mn ( 5 ); trien = triethylenetetramine; space group: ) the M2+ ions are fivefold coordinated by four N atoms of the amine molecule and by one S atom of the thioantimonate anion forming a MN4S trigonal bipyramid. Sb8S16 building blocks are the central structural motifs of the anion. Two of the terminal S atoms at the periphery of the Sb8S16 units are bound to M2+ ions and the four remaining terminal S atoms connect adjacent Sb8S16 groups into the final [Sb4S7]2− chain. [Ni(tren)]Sb4S7 ( 6 ) (space group: ) contains a one‐dimensional anionic chain. The Ni2+ ion has two bonds to the [Sb4S7]2− anion which is a unique feature in the thioantimonate(III) chemistry. The NiN4S2 octahedron is severly distorted with one very long Ni‐S bond of 2.782(2) Å. In all compounds several short S···H distances indicate hydrogen bonding interactions.  相似文献   

5.
Sulphur deposited on gold by the anodic oxidation of sulphur(−II) species in solution has been studied by X-ray photoelectron spectroscopy. The initial layer behaved as gold sulphide. Multilayers of sulphur had a lower volatility and a smaller electron binding energy than bulk elemental sulphur, indicating that there is interaction with the underlying gold or gold sulphide.The anodic oxidation of sulphur(−II) to sulphur, and the reverse process, has been investigated on gold using the rotating ring disc electrode technique. Polysulphide ions were formed as intermediates in both processes. Polysulphides were also produced by chemical reaction of deposited sulphur with sulphur(−II) species in solution. The polysulphide intermediates were identified as S2−5 at pH 13, a mixture of species with average stoichiometry S2−3.3 at pH 9.2 and S2−2, possibly HS2, at pH 6.8.  相似文献   

6.
Reductive elimination of alkyl−PdII−O is a synthetically useful yet underdeveloped elementary reaction. Here we report that the combination of an H-bonding donor [PyH][BF4] and AgNO3 additive under toluene/H2O biphasic system can enable such elementary step to form alkyl nitrate. This results in the Pd0-catalyzed asymmetric carbonitratations of (Z)-1-iodo-1,6-dienes with (R)-BINAP as the chiral ligand, affording alkyl nitrates up to 96 % ee. Mechanistic studies disclose that the reaction consists of oxidative addition of Pd0 catalyst to vinyl iodide, anion ligand exchange between I and NO3, alkene insertion and SN2-type alkyl−PdII−ONO2 reductive elimination. Evidences suggest that H-bonding interaction of PyH⋅⋅⋅ONO2 can facilitate dissociation of O2NO ligand from the alkyl−PdII−ONO2 species, thus enabling the challenging alkyl−PdII−ONO2 reductive elimination to be feasible.  相似文献   

7.
The reaction of hexachlorophosphazene, P3N3Cl6, with SO3 and the gold halides AuCl3 and AuBr3, respectively, leads to the new cyclic anionic tetramer, [S4N2O10]2−, which is coordinated to Au3+ in the dimeric complexes [Au2X2(S4N2O10)2] (X=Cl, Br). The [S4N2O10]2− anion can be seen as the condensation product of two sulfate anions, [SO4]2−, and two amidosulfate anions, [NH2SO3].  相似文献   

8.
The synthesis and crystal structure (at 100 K) of the title compound, Cs[Fe(C11H13N3O2S2)2]·CH3OH, is reported. The asymmetric unit consists of an octahedral [FeIII(L)2] fragment, where L2− is 3‐ethoxysalicylaldehyde 4‐methylthiosemicarbazonate(2−) {systematic name: [2‐(3‐ethoxy‐2‐oxidobenzylidene)hydrazin‐1‐ylidene](methylamino)methanethiolate}, a caesium cation and a methanol solvent molecule. Each L2− ligand binds through the thiolate S, the imine N and the phenolate O atoms as donors, resulting in an FeIIIS2N2O2 chromophore. The O,N,S‐coordinating ligands are orientated in two perpendicular planes, with the O and S atoms in cis positions and the N atoms in trans positions. The FeIII cation is in the low‐spin state at 100 K.  相似文献   

9.
The sulfur rich difluoropentathiodiphosphate dianion [S5P2F2]2−, from fluoride addition to P4S10, has a somewhat checkered history and proves to be the main product of the reaction in acetonitrile. Its optimized synthesis, and structural characterization, as either a tetraphenylphosphonium or a tetrapropylammonium salt, [NnPr4]2[S5P2F2] allows for the first coordination chemistry for this dianion. Reactions of [S5P2F2]2− with d10 metal ions of zinc(II), and cadmium(II), and d9 copper(II) resulted in a surprising diverse array of binding modes and structural motifs. In addition to the simple bis-chelate coordination of [S5P2F2]2− with zinc, cleavage of the P−S bond resulted in complexes with the unusual [S3PF]2− fluorotrithiophosphate dianion. This was observed in two cluster complexes: a trinuclear cadmium complex with mixed [S5P2F2]2−/[S3PF]2− ligands, [Cd3(S5P2F2)3(S3PF)2]4− as well as an octanuclear copper cluster, [Cu8(S3PF)6]4− which form rapidly at room temperature. These new metal/sulfur/ligand clusters are of relevance to understanding multimetal binding to metallothionines, and to potential capping strategies for the condensed nanoparticulate cadmium chalcogenide semiconductors CdS and CdSe.  相似文献   

10.
The generation of negative ions from SO2 in the gas phase was studied using the thermal surface ionization method. Six anion types were measured: O, S, SO, and SO2 and anions with m/z=96 and m/z=128. The most abundant anion formed was S and the formation routes are discussed for each of the six anions. O, S, and SO are formed via dissociative electron attachment to the molecule, whereas the generation of SO2 and anions with m/z=96 and m/z=128 are probably associated with the formation of H2SO4 in the gas inlet system and the ion source. Using statistical thermodynamics the dissociation temperatures of SO2 and SO in the gas phase are calculated and values of above 1800 °C are obtained for both molecules. We also estimated the optimal filament temperatures for the formation of all anions measured, indicating that for SO2 the optimal temperature is related to the electron affinity of the molecules: the optimal temperature increases with decreasing value of the electron affinity for the molecule corresponding to the respective anion.  相似文献   

11.
A highly unusual solid-state epitaxy-induced phase transformation of Na4SnS4 ⋅ 14H2O ( I ) into Na4Sn2S6 ⋅ 5H2O ( II ) occurs at room temperature. Ab initio molecular dynamics (AIMD) simulations indicate an internal acid-base reaction to form [SnS3SH]3− which condensates to [Sn2S6]4−. The reaction involves a complex sequence of O−H bond cleavage, S2− protonation, Sn−S bond formation and diffusion of various species while preserving the crystal morphology. In situ Raman and IR spectroscopy evidence the formation of [Sn2S6]4−. DFT calculations allowed assignment of all bands appearing during the transformation. X-ray diffraction and in situ 1H NMR demonstrate a transformation within several days and yield a reaction turnover of ≈0.38 %/h. AIMD and experimental ionic conductivity data closely follow a Vogel-Fulcher-Tammann type T dependence with D(Na)=6×10−14 m2 s−1 at T=300 K with values increasing by three orders of magnitude from −20 to +25 °C.  相似文献   

12.
Novel catalytically active monooxomolybdenum(IV) species containing four thiolate ligands obtainable in solution by NaBH4 reduction of [MovO(SC6H5)4], [MovO(Z-cys-Val-OMe)4], (Z=benzyloxycarbonyl), or [MovO(S2C6H4)2] perform the pyridine-N-oxide oxidation of benzoin in N,N-dimethylformamide at 30 °C. The order of catalytic activity is [MovO(Z-cys-Val-OMe)4] > [MovO(S2C6H4)2] > [MovO(SC6H5)4] ([benzoin]/[oxidant]/[catalyst]= 20/20/1), while the oxidation by air under the same catalytic conditions gives a different order, [MovO(Z-cys-Val-OMe)4]> [MovO(SC6H5)4] >[MovO(S2C6H4)2]. During the catalytic cycle in the amine-N-oxide oxidation, two intermediate species, [MoIVOL4]2− and [MoVIO2L4]2−, were detected by 1H NMR, while in the air oxidation an unidentified Mo(VT) species is involved.  相似文献   

13.
The Cd(II)-, Pb(II)-, Ni(II)- and Zn(II)-complexes of small terminally protected peptides containing CXXX, XXXC, XCCX, CXnC (n=1–3) sequences have been studied with potentiometric, UV/Vis and CD spectroscopic techniques. The cysteine thiolate group is the primary binding site for all studied metal ions, but the presence of a histidyl or aspartyl side chain in the molecule contributes to the stability of the complexes. For two-cysteine containing peptides the (S,S) coordinated species are formed in the physiological pH range and the stability increases in the Ni(II)<Zn(II)<Pb(II)<Cd(II) order. As a conclusion, the inserting of −CXXC− sequence into the peptide makes the synthesis of peptides with high selectivity to toxic Cd(II) or Pb(II) ion possible. In addition, the spectroscopic characterization of these complexes can contribute to the discovery of the exact binding site and binding mode of longer peptides mimicking the biologically important proteins.  相似文献   

14.
《Comptes Rendus Chimie》2008,11(8):852-860
IR spectroelectrochemical studies of bis(thiolate) and dithiolate-bridged diiron carbonyl compounds, [Fe2(μ-SR)2(CO)6], show that the primary reduction process results in rapid chemical reaction, leading to two-electron reduced products. When the reaction is conducted under an inert atmosphere, the major product is [Fe2(μ-SR)(μ-CO)(CO)6]1−, where in the case of dithiolate-bridged neutral compounds the product has one bridging and one non-bound sulfur atom. This product is formed in near-quantitative yield for solutions saturated with CO. Reduction of [Fe2(μ-SR)(μ-CO)(CO)6]1− occurs at potentials near −2.0 V vs. SCE to give a range of products including [Fe(CO)4]2−. Reduction of thiolate-bridged diiron compounds at mild potentials in the presence of CH3COOH leads to formation of [Fe2(μ-SR)(μ-CO)(CO)6]1− and this is accompanied by an acid-base reaction with the dissociated thiolate. The reaction is largely reversible with recovery of ca. 90% of the starting diiron compound and CH3COOH. In the presence of acid, reduction of [Fe2(μ-SR)2(CO)6] proceeds without generation of observable concentrations of the structurally related one-electron reduced compound. Electrocatalytic proton reduction is achieved when the potential is stepped sufficiently negative to reduce [Fe2(μ-SR)(μ-CO)(CO)6]1−, an observation in keeping with the cyclic voltammetry of the system. Since the catalytic species involved in the weak-acid reactions is structurally distinct from the starting material, and the diiron subsite of the hydrogenase H-cluster, these experiments are of dubious relevance to the biological system.  相似文献   

15.
Photolysis of S2O8= in strong alkaline solutions (pH>13) in the presence of molecular oxygen yields ozonide radical ions, . These radicals show a complex decay rate sensitive to the peroxodisulfate concentration. A reaction mechanism, which includes the reaction of O•− and S2O8= with a rate constant k=(3−6)×106M−1s−1 and accounts for the experimental results is discussed. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 491–496, 1998  相似文献   

16.
The title compound, [Pb(C4H3N2S)2]n, was prepared by the reaction of [Pb(OAc)2]·3H2O (OAc is acetate) with pyrimidine‐2‐thione in the presence of triethylamine in methanol. In the crystal structure, the PbII atom has an N4S4 coordination environment with four ligands coordinated by N‐ and S‐donor atoms. This compound shows that the pyrimidine‐2‐thiolate anion can lead to a three‐dimensional network when the coordination number of the metal ion can be higher than 6, as is the case with the PbII ion. This compound presents only covalent bonds, showing that despite the possibility of the hemidirected geometries of PbII, the eight‐coordinated ion does not allow the formation of an isolated molecular structure with pyrimidine‐2‐thiolate as the ligand.  相似文献   

17.
The salts 1‐(diaminomethylene)thiouron‐1‐ium hydrogen difluoride, C2H7N4S+·HF2, (I), and bis[1‐(diaminomethylene)thiouron‐1‐ium] hexafluoridosilicate, 2C2H7N4S+·SiF62−, (II), have both been obtained from the reaction of (1‐diaminomethylene)thiourea (HATU) with hydrofluoric acid. Both compounds contain extensive networks of N—H...F hydrogen bonds. The hydrogen difluoride salt contains four independent asymmetric [HF2] anions. In the hexafluoridosilicate salt, the centrosymmetric [SiF6]2− anion is distorted, although this distortion is not clearly correlated with the N—H...F hydrogen‐bonding network.  相似文献   

18.
《Comptes Rendus Chimie》2003,6(4):485-491
Reactivity of 5-thioxylopyranosyl bromide and 1,5-dithioxylopyranoside towards thiolate anions. Reactivity of thiolate anion RS 2′ (C6H5–S) and 3′ (p-CH3–C6H4–S) towards 5-thioxylopyranosyl bromide Xyl–Br yields to the corresponding 1,5-dithioxylopyranoside Xyl–SR 7 R = C6H5– and 8 R = p-CH3–C6H4, respectively. In the presence of 4′ (C6H5–CH2–S) or 5′ (CH3–S), the reaction gives the 5-thioxylopyranose 9. Anions 5′ and 4′ react with 1,5-dithioxylopyranoside 10 Xyl–SR′ (R′ = –C6H4–CO–C6H4–CN) to give sulphide derivative 11 (CH3–S–C6H4–CO–C6H4– CN) and 13 (C6H5–CH2–S–C6H4–CO–C6H4–CN), respectively, and the 1,5–dithioxylose 12. These differences in terms of reactivity could be explained by the nucleophilic behaviour of the formed thiolate anion. To cite this article: D. Brevet et al., C. R. Chimie 6 (2003).  相似文献   

19.
P4S10 reacts with (Me3Si)2NH, to give SP(NHSi-Me3)x(SSiMe3)3−x, where x = 0–3. These species were shown by 31P NMR spectroscopy to subsequently form cyclic and linear condensation products. The molecular structure of a linear trimer (Me3SiNH)P(S)[(μ-NH)P(S)(NHSiMe3)2]2, 12 , was authenticated by an X-ray diffraction study. © 1998 John Wiley & Sons, Inc. Heteroatom Chem 9:115–121, 1998  相似文献   

20.
The sulfur-rich iron carbonyl dimer complexes [Fe(CO)2(S′SiS2)]2 (2), and [Fe(CO)(S′SiS2)]2 (3) have been prepared. The [2Fe-2S] cores of the new complexes are planar. The binding mode of the tridentate sulfur ligand in complex 2 is facial with a S(thiolate)-Fe-S(thiolate) angle of 92°, while in complex 3, the S′SiS2 ligand binds the metal with a S(thiolate)-M-S(thiolate) angle of 120°. The Fe-Fe distance is reduced from 3.45 Å in complex 2 to 2.78 Å in the 32 electron dimer complex 3. Complexes 2 and 3 are at equilibrium in solution and can be readily interconverted by addition or removal of CO.  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