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 31 毫秒
1.
A novel dinuclear cis-dioxomolybdenum(VI) complex [{MoO2(Bz2endtc)}2] coordinated with a quadradentate dithiocarbamate (Bz2endtc2−: ((2-(dithiocarboxybenzylamino)ethyl)benzylamino)-methanedithioate(2−)) has been synthesised. The structural features of [{MoO2(Bz2endtc)}2] have been elucidated by X-ray crystal analysis, elemental analysis and 13C NMR, IR and FAB+ mass spectroscopy: two almost identical cis-dioxomolybdenum(VI) centres are bridged by the two Bz2endtc2− ligands and each molybdenum(VI) centre has a distorted octahedral geometry with four sulphur atoms and two terminal oxo ligands lying in a cis position to each other. There is unlikely to be electronic interaction between the two cis-dioxomolybdenum(VI) centres in [{MoO2(Bz2endtc)}2] because the MoMo distance is long (=7.337 Å). In the [{MoO2(Bz2endtc)}2]/PPh3 system, the oxygen atom transfer reaction (Eq. (A)) occurs to give a tetranuclear oxomolybdenum(VI,V) complex formulated as [MoO2(Bz2endtc)2Mo2O3(Bz2endtc)2MoO2] which has one μ-oxomolybdenum(V) moiety.
(A)  相似文献   

2.
Reaction of cis-[Ru(acac)22-C8H14)2] (1) (acac = acetylacetonato) with two equivalents of PiPr3 in THF at −25 °C gives trans-[Ru(acac)2(PiPr3)2], trans-3, which rapidly isomerizes to cis-3 at room temperature. The poorly soluble complex [Ru(acac)2(PCy3)2] (4), which is isolated similarly from cis-[Ru(acac)22-C2H4)2] (2) and PCy3, appears to exist in the cis-configuration in solution according to NMR data, although an X-ray diffraction study of a single crystal shows the presence of trans-4. In benzene or toluene 2 reacts with PiPr3 or PCy3 to give exclusively cis-[Ru(acac)22-C2H4)(L)] [L = PiPr3 (5), PCy3 (6)], whereas in THF species believed to be either square pyramidal [Ru(acac)2L], with apical L, or the corresponding THF adducts, can be detected by 31P NMR spectroscopy. Complexes 3-6 react with CO (1 bar) giving trans-[Ru(acac)2(CO)(L)] [L = PiPr3 (trans-8), PCy3 (trans-9)], which are converted irreversibly into the cis-isomers in refluxing benzene. Complex 5 scavenges traces of dinitrogen from industrial grade dihydrogen giving a bridging dinitrogen complex, cis-[{Ru(acac)2(PiPr3)} 2(μ-N2)] (10). The structures of cis-3, trans-4, 5, 6 and 10 · C6H14 have been determined by single-crystal X-ray diffraction. Complexes trans- and cis-3, 5, 6, cis-8, and trans- and cis-9 each show fully reversible one-electron oxidation by cyclic voltammetry in CH2Cl2 at −50 °C with E1/2(Ru3+/2+) values spanning −0.14 to +0.92 V (versus Ag/AgCl), whereas for the vinylidene complexes [Ru(acac)2 (CCHR)(PiPr3)] [R = SiMe3 (11), Ph (12)] the process is irreversible at potentials of +0.75 and +0.62 V, respectively. The trend in potentials reflects the order of expected π-acceptor ability of the ligands: PiPr3, PCy3 <C 2H4 < CCHR < CO. The UV-Vis spectrum of the thermally unstable, electrogenerated RuIII-ethene cation 6+ has been observed at −50 °C. Cyclic voltammetry of the μ-dinitrogen complex 10 shows two, fully reversible processes in CH2Cl2 at −50 °C at +0.30 and +0.90 V (versus Ag/AgCl) corresponding to the formation of 10+ (RuII,III) and 102+ (RuIII,III). The former, generated electrochemically at −50 °C, shows a band in the near IR at ca. 8900 cm−1 (w1/2 ca. 3700 cm−1) consistent with the presence of a valence delocalized system. The comproportionation constant for the equilibrium 10 + 102+ ? 2 10+ at 223 K is estimated as 1013.6.  相似文献   

3.
The present paper describes a new tripodal ligand containing imidazole and pyridine arms and its first cis-[RuIII(L)(Cl)2]ClO4 complex (1). The crystal structure of 1 shows RuIII in a distorted octahedral geometry, in which two chloride ions, cis-positioned to each other, are coordinated besides the four nitrogen atoms from the tetradentate ligand L. The cyclic voltammogram of 1 exhibits three redox processes at −67, +73 and +200 mV versus SCE, which are attributed to the RuIII/RuII couple in the cis-[RuIII(L)(Cl)2]+, cis-[RuII(L)(H2O)(Cl)]+ and cis-[RuII(L)(H2O)2]2+, respectively. After chemical reduction (Zn(Hg) or EuII) only the cis-[RuII(L)(H2O)2]2+ species is observed in the cyclic voltammetry. Complex 1 absorbs at 470 nm (ε=1.4×103 mol−1 L cm−1), 335 nm (ε=7.9×103 mol−1 L cm−1), 301 nm (ε=6.7×103 mol−1 L cm−1) and 264 nm (ε=9.9×103 mol−1 L cm−1), in water solution (CF3COOH, 0.01 mol L−1, μ=0.1 mol L−1 with CF3COONa). Spectroelectrochemical experiments show a decrease of the bands at 335 and 301 nm, which are attributed to LMCT transitions from the chloride to the RuIII center and the appearance of a broad band at 402 nm ascribed to MLCT transition from the RuII center to the pyridine ligand. The lability of the water ligands in the cis-[RuII(L)(H2O)2]2+ species has been investigated using the auxiliary ligand pyrazine. Reactions in the presence of stoichiometric and excess of pyrazine yield the same species, cis-[RuII(L)(H2O)(pz)]2+, which exhibits a reversible redox process at 493 mV versus SCE and absorbs at 438 nm (ε=5.1×103 mol−1 L cm−1) and 394 nm (ε=4.2×103 mol−1 L cm−1). Experiments performed with a large excess of pyrazine gave a specific rate constant k1=(2.8±0.5)×10−2 M−1 s−1, at 25 °C, in CF3COOH, 0.01 mol L−1, μ=0.1 mol L−1 (with CF3COONa).  相似文献   

4.
A new class of asymmetric N-capped (dianionic/trianionic) tripodal proligands [Hx(Ln)] (x = 2, n = 1-6; x = 3, n = 7, 8) which possess pendant arms with N2OS, N2S2 or NOS2 donor groups and with different chelate ring sizes {5,5,5} or {5,6,5} has been prepared. Treatment of these ligands with [WO2Cl2(dme)] (dme = 1,2-dimethoxyethane) in the presence of base (triethylamine or KOH) leads to the formation of cis-dioxotungsten(VI) complexes of the types [WO2(Ln)] (n = 1-6) and K[WO2(Ln)] (n = 7, 8). Reaction of these tetradentate ligands with [MoO2(acac)2] (acac = acetylacetonate) gives the corresponding Mo(VI) analogues [MoO2(Ln)] (n = 1-6) and K[MoO2(Ln)] (n = 7, 8). Moreover, a new five coordinate dioxomolybdenum(VI) complex with an NS2 tridentate ligand [MoO2(L9)] has been synthesised using similar procedure. All these compounds have been spectroscopically characterised and the molecular structures of [MoO2(Ln)] (n = 2, 6) and [WO2(L6)] have been established by X-ray diffraction analysis. The electrochemistry and the catalytic activity for oxidation of allylic and benzylic alcohols of these dioxo complexes have also been investigated.  相似文献   

5.
《Inorganica chimica acta》1987,128(2):231-237
Ni(II) dithiocarbamates (Ni(dtc)2) with various substituents on dtc were allowed to react with triphenylphosphine (PPh3). Mixed ligand complexes of the general formulae Ni(dtc)Cl(PPh3) and [Ni(dtc)(PPh3)2]ClO4 were prepared. The complexes were analysed by high resolution IR spectra. Comparison of the ν(C–N) frequencies of different complexes viz., Ni(dtc)2, Ni(dtc)Cl(PPh3) and [Ni(dtc)(PPh3)2]ClO4, showed the following order of decreasing v(C–N) values: [Ni(dtc)(PPh3)2]+> Ni(dtc)Cl(PPh3)> Ni(dtc)2. The observation showed the extent of contribution of the thiouride form in describing the structure of the complexes. The higher the contribution, larger is the value of ν(C–N). Cyclic voltammetric studies on the complexes showed the one electron reduction potentials to decrease in the following order: Ni(dtc)Cl(PPh3)>Ni(dtc)2> [Ni(dtc)(PPh3)2]+. The observations are explained with the nature of the substituents on the dtc moiety and other ligands present around Ni(II). Crystal structure of [Ni(dedtc) (PPh3)2]ClO4 (dedtc = diethyldithiocarbamate) was determined to study the effect of the introduction of PPh3 in place of Cl in the Ni(dtc)Cl(PPh3) complex. The complex is planar with NiS2P2 chromophore. The NiS distances are 2.190(2) and 2.239(2) Å and the NiP distances are 2.230(2) and 2.200(2) Å. The asymmetry in the NiS and NiP distances is ascribed to the steric effect due to bulky PPh3. The structural aspects are compared with those of the Ni(dtc)Cl(PPh3) complex.  相似文献   

6.
The reaction of K2[ReX6] (X = Cl, Br) with oxalic acid and triethylamine in dimethylformamide solution yields the substituted complexes [ReX4(ox)]2− and cis-[ReX2(ox)2]2−, which can be obtained separately depending on the amount of added amine. The crystal structures of (PPh4)2[ReBr4(ox)], cis-(PPh4)2[ReBr2(ox)2] and cis-(AsPh4)2[ReCl2(ox)2] have been determined by single-crystal X-ray diffraction. The anionic complexes are octahedral with only slight distortions. The direct isolation of the pure complexes as well as the formation of only the cis isomers - without the presence of trans isomers and/or [Re(ox)3]2− - is probably due to the kinetic inertness of Re(IV)-X bonds, which increases with the number of oxalato ligands bound to the metal ion.  相似文献   

7.
A reaction of the octahedral bidentate metalloligand, trans(N)-[Co(d-pen)2] (d-pen=d-penicillaminate) with Cd(NO3)2 or Cd(ClO4)2 gave a novel S-bridged trinuclear complex, [Cd(H2O){Co(d-pen)2}2] (1). In this complex molecule, the central Cd atom is surrounded by four S atoms from two [Co(d-pen)2] units and one O atom of a H2O molecule to form a distorted five-coordinated geometry. Each of two terminal [Co(d-pen)2] units takes an approximately octahedral geometry and has a similar trans(N) geometry to that of the starting material. On the other hand, the reaction of trans(N)-[Co(d-pen)2] with CdCl2 in the molar ratio of 1:1 gave an S-bridged dinuclear complex, [CdCl{Co(d-pen)2}(H2O)mnH2O (m+n=4) (2). The reactivity of trans(N)-[Co(d-pen)2] toward CdCl2 is significantly influenced by the ratio of two components, and the formation of a similar trinuclear species to 1 is also suggested under the condition with excess amount of trans(N)-[Co(d-pen)2]. Some spectrochemical properties of these complexes are also discussed in relation to their structures.  相似文献   

8.
In order to examine the effects of coordinated hydroxide ion and free hydroxide ion in configurational conversion of a tetraamine macrocyclic ligand complex, the kinetic of the cis-to-planar interconversion of cis-[Ni(isocyclam)(H2O)2]2+ (isocyclam = 1,4,7,11-tetraazacyclotetradecane) has been examined spectrophotometrically. All kinetic data have been satisfactorily fitted by the rate law, R = (k1KOH[OH]2 + k2[OH])(1 + KOH[OH])−1(cis-[Ni(isocyclam)(H2O)2]2+ + [Ni(isocyclam)(OH)]+), where k2 = (3.40 ± 0.12) × 103 dm3 mol−1 s−1 is almost equal to kOH determined in buffer solution (lowly basic media), KOH = 22.7 ± 1.4 dm3 mol−1 at I (ionic strength) = 0.10 mol dm−3 (NaClO4 + NaOH) and 25.0 °C. Rate constants, k2 and KOH, are functions of ionic strength, giving a good evidence for an intermolecular pathway. The reaction follows a free-base-catalyzed mechanism where nitrogen inversion, solvation and ring conformational changes are occurred.  相似文献   

9.
In order to examine the effects of coordinated hydroxide ion and free hydroxide ion in configurational conversion of a tetraamine macrocyclic ligand complex, the kinetics of the cis-to-planar interconversion of cis-[Ni(isocyclam)(H2O)2]2+ (isocyclam, 1,4,7,11-tetraazacyclotetradecane) has been studied spectrophotometrically in basic aqueous solution. The interconversion requires the inversion of one sec-NH center of the folded cis-complex to have the planar species. Kinetic data are satisfactorily fitted by the rate law, R = kOH[OH][cis-[Ni(isocyclam)(H2O)2]2+], where kOH = 3.84 × 103 dm3 mol−1 s−1 at 25.0 ± 0.1 °C with I = 0.10 mol dm−3 (NaClO4). The large ΔH, 61.7 ± 3.2 kJ mol−1, and the large positive ΔS, 30.2 ± 10.8 J K−1 mol−1, strongly support a free-base-catalyzed mechanism for the reaction.  相似文献   

10.
Two 15N-labelled cis-Pt(II) diamine complexes with dimethylamine (15N-dma) and isopropylamine (15N-ipa) ligands have been prepared and characterised. [1H,15N] HSQC NMR spectroscopy is used to obtain the rate and equilibrium constants for the aquation of cis-[PtCl2(15N-dma)2] at 298 K in 0.1 M NaClO4 and to determine the pKa values of cis-[PtCl(H2O)(15N-dma)2]+ (6.37) and cis-[Pt(H2O)2(15N-dma)2]2+ (pKa1 = 5.17, pKa2 = 6.47). The rate constants for the first and second aquation steps (k1 = (2.12 ± 0.01) × 10−5 s−1, k2 = (8.7 ± 0.7) × 10−6 s−1) and anation steps (k−1 = (6.7 ± 0.8) × 10−3 M−1 s−1, k−2 = 0.043 ± 0.004 M−1 s−1) are very similar to those reported for cisplatin under similar conditions, and a minor difference is that slow formation of the hydroxo-bridged dimer is observed. Aquation studies of cis-[PtCl2(15N-ipa)2] were precluded by the close proximity of the NH proton signal to the 1H2O resonance.  相似文献   

11.
The salt elimination reaction between FeCl2(THF)1.5 and the sterically hindered thallium tert-butyl-tris(3-isopropylpyrazolyl)borate, Tl[t-BuTpi-Pr], affords the new air and moisture-sensitive, high-spin, four-coordinate complex Fe[t-BuTpi-Pr]Cl (1) in 52% yield. Compound 1 has been characterized by elemental analysis, paramagnetic 1H NMR, and X-ray crystal structure determination. Fe[t-BuTpi-Pr]Cl is monomeric, and the homoscorpionate ligand is κ3-coordinated. The Fe(II) ion lies in a trigonally distorted tetrahedral environment with average N-Fe-N and N-Fe-Cl angles of 89.4(2)° and 125.7(2)°, respectively. Upon reactions with methyllithium followed by carbonylation, the six-coordinate acetyl-dicarbonyl derivative Fe[t-BuTpi-Pr](CO)2{C(O)Me} was identified by IR spectroscopy. The reaction of 1 with ethylmagnesiumbromide furnished a mixture of mononuclear Grignard complexes Mg[t-BuTpi-Pr]X (X = Et, Cl, Br), resulting formally from a facile and quantitative replacement of the [Fe-Cl]+ fragment by [Mg-X]+.  相似文献   

12.
Crystal structure of [ReO2(4-MeOpy)4][PF6] (4-MeOpy = 4-methoxypyridine) complex has been examined by the single crystal X-ray analytical method. This complex shows a trans-dioxo geometry (average Re-O bond length = 1.766(2) Å) and its equatorial plane is occupied by four 4-MeOpy molecules (average Re-N bond length = 2.156(4) Å). Electrochemical reaction of [ReO2(4-MeOpy)4]+ in CH3CN solution containing tetra-n-butylammonium perchlorate as a supporting electrolyte has been studied using cyclic voltammetry at 24 °C. Cyclic voltammograms show one redox couple around 0.65 V (Epa) and 0.58 V (Epc) [versus ferrocene/ferrocenium ion redox couple, (Fc/Fc+)]. Potential differences between two peaks (ΔEp) at scan rates in the range from 0.01 to 0.10 V s−1 are 65 mV, which is almost consistent with the theoretical ΔEp value (59 mV) for the reversible one electron transfer reaction at 24 °C. The ratio of anodic peak currents to cathodic ones is 1.04 ± 0.03 and the (Epa + Epc)/2 value is constant, 0.613 ± 0.001 V versus Fc/Fc+, regardless of the scan rate. Spectroelectrochemical experiments have also been carried out by applying potentials from 0.40 to 0.77 V versus Fc/Fc+ with an optically transparent thin layer electrode. It was found that the UV-visible absorption spectra show clear isosbestic points at 228, 276, and 384 nm, and that the electron stoichiometry is evaluated as 1.03 from the Nernstian plot. These results indicate that the [ReO2(4-MeOpy)4]+ complex is oxidized reversibly to the [ReO2(4-MeOpy)4]2+ complex. Furthermore, it was clarified that the [ReO2(4-MeOpy)4]2+ in CH3CN has the characteristic absorption bands at 236, 278, 330, 478, and 543 nm and their molar absorption coefficients are 4.3 × 104, 4.5 × 103, 1.0 × 104, and 6.1 × 103 M−1 cm−1 (M = mol dm−3), respectively.  相似文献   

13.
The reaction of trans(N)-[Co(d-pen)2] (pen = penicillaminate) with HgCl2 or HgBr2 in the molar ratios of 1:1 gave the sulfur-bridged heterodinuclear complex, [HgX(OH2){Co(d-pen)2}] (X = Cl (1a) or Br (1b)). A similar reaction in the ratio of 2:1 produced the trinuclear complex, [Hg{Co(d-pen)2}2] (1c). The enantiomers of 1a and 1c, [HgCl(OH2){Co(l-pen)2}] (1a′) and [Hg{Co(l-pen)2}2] (1c′), were also obtained by using trans(N)-[Co(l-pen)2] instead of trans(N)-[Co(d-pen)2]. Further, the reaction of cis · cis · cis-[Co(d-pen)(l-pen)] with HgCl2 in the molar ratio of 1:1 resulted in the formation of [HgCl(OH2){Co(d-pen)(l-pen)}] (2a). During the formations of the above six complexes, 1a, 1b, 1c, 1a′, 1c′, and 2a, the octahedral Co(III) units retain their configurations. On the other hand, the reaction of cis · cis · cis-[Co(d-pen)(l-pen)] with HgCl2 in the molar ratio of 2:1 gave not [Hg{Co(d-pen)(l-pen}2] but [Hg{Co(d-pen)2}{Co(l-pen)2}] (2c), accompanied by the ligand-exchange on the terminal Co(III) units. The X-ray crystal structural analyses show that the central Hg(II) atom in 1c takes a considerably distorted tetrahedral geometry, whereas that in 2c is of an ideal tetrahedron. The interconversion between the complexes is also examined. The electronic absorption, CD, and NMR spectral behavior of the complexes is discussed in relation to the crystal structures of 1c and 2c.  相似文献   

14.
The reaction between [PtII(Ox)2]2− and an appropriate oxidant resulted in the formation of the dimeric unbridged platinum complex [{PtIII(Ox)2 }2]2− where (Ox) is oxalate. This complex was moderately stable under ambient conditions and was studied via a variety of NMR and spectrophotometric techniques. Reaction of the [{PtIII(Ox)2}2]2− complex with [PtII(Ox)2]2− in the presence of H+ lead to the formation of a series of longer platinum oligomers with non-integral oxidation states, culminating in the formation of partially oxidized platinum polymers of general formula [{Pt(Ox)2}n]n. The concentration of H+ was an important factor leading to higher oligomers and the approximate number of protons associated with each oligomer was determined. The analogous [{PtIII(Mal)2}2]2− complex, where (Mal) is the malonate anion, was also synthesized and studied but was shown to be significantly less stable.  相似文献   

15.
Reduction of the model platinum(IV) complexes cis-[PtCl4(NH3)2] (1), trans-[PtCl4(NH3)2] (2), trans-[PtCl2(en)2]2+ (3), trans-[PtBr2(NH3)4]2+ (4), [PtCl6]2− (5), and [PtBr6]2− (6) with l-ascorbic acid (H2Asc) in 1.0 M aqueous medium at 25 °C in the region 1.75≤pH≤7.20 has been investigated using stopped-flow spectrophotometry. The redox reactions follow the rate law: −d[Pt(IV]/dt=k[H2Asc]tot[Pt(IV)] where k is a pH-dependent second-order rate constant and [H2Asc]tot, the total concentration of ascorbic acid. The pH-dependence of k is attributed to parallel reduction of Pt(IV) by the protolytic species HAsc and Asc2−. Analysis of the kinetics data reveals that the ascorbate anion Asc2− is up to seven orders of magnitude more reactive than HAsc while H2Asc is unreactive. Electron transfer from HAsc/Asc2− to the Pt(IV) compounds is suggested to take place by a mechanism involving a reductive attack on any one of the mutually trans-halide ligands by Asc2− and/or HAsc forming a halide-bridged activated complex. The rapid reduction of these complexes supports the assumption that ascorbate Asc2− might be an important reductant at physiological conditions for anticancer active Pt(IV) pro-drugs capable of undergoing reductive trans elimination. The parameters ΔH and ΔS for reduction of Pt(IV) with Asc2− have been determined from the study of the temperature dependence of k.  相似文献   

16.
Pt(II) complexes of the types K[Pt(R2SO)X3], NR4[Pt(R2SO)X3] and Pt(R2SO)2Cl2 (where X = Cl or Br) were characterized by multinuclear magnetic resonance spectroscopy (195Pt, 1H and 13C). In 195Pt NMR, the chloro ionic compounds have shown signals between −2979 and −3106 ppm, while the cis disubstituted complexes were observed at higher fields, between −3450 and −3546 ppm. The signal of the compound trans-Pt(DPrSO)2Cl2 was found at higher field (−3666 ppm) than its cis analogue (−3517 ppm), since π-back-donation is considerably less effective in the trans geometry. In 1H NMR, a single signal was observed for the sulfoxide in [Pt(DMSO)Cl3], but for the other more sterically hindered ligands, two series of resonances were observed for the protons in α and β positions. The coupling constant 3J(195Pt-1H) are between 15 and 33 Hz. The 13C NMR results were interpreted in relation to the concept of inversed polarization of the π sulfoxide bond. The 2J(195Pt-13C) values vary between 35 and 66 Hz, while a few 3J(195Pt-13C) couplings were observed (13-26 Hz). The crystal structures of five monosubstituted ionic compounds N(n-Bu)4[Pt(TMSO)Cl3], N(Me)4[Pt(DPrSO)Cl3], K[Pt(EMSO)Cl3], K[Pt(TMSO)Br3] · H2O and N(Et)4[Pt(DPrSO)Br3] and one disubstituted complex cis-Pt(DBuSO)2Cl2 were determined. The trans influence of the different ligands is discussed.  相似文献   

17.
The reactivity of the metalloligand [Pt2(μ-S)2(PPh3)4] towards a variety of indium(III) substrates has been explored. Reaction with excess In(NO3)3 and halide (KBr or NaI) gave the four-coordinate adducts [Pt2(μ-S)2(PPh3)4InX2]+[InX4] (X = Br, I). An X-ray structure determination on the iodo complex revealed a slightly distorted tetrahedral coordination geometry at indium. In contrast, reaction of [Pt2(μ-S)2(PPh3)4] with indium(III) chloride was more complex; the ion [Pt2(μ-S)2(PPh3)4InCl2]+ was initially observed in solution (using ESI mass spectrometry), and isolated as its BPh4 salt. Analysis of [Pt2(μ-S)2(PPh3)4InCl2]+[BPh4] by ESI MS showed the parent cation when analysed in MeCN solution. However in solutions containing methanol, partial solvolysis occurred to give the di-indium species [{Pt2(μ-S)2(PPh3)4InCl(OMe)}2]2+ (proposed to contain an In2(μ-OMe)2 unit with five-coordinate indium) and its fragment ion [Pt2(μ-S)2(PPh3)4InCl(OMe)]+. Reaction of [Pt2(μ-S)2(PPh3)4] with InCl3·3H2O, 8-hydroxyquinoline (HQ) and trimethylamine in methanol gave the adduct [Pt2(μ-S)2(PPh3)4InQ2]+, isolated as its PF6 salt. The same cationic complex is formed when [Pt2(μ-S)2(PPh3)4] is reacted with InQ3 in methanol, but in this case the product is contaminated with the mononuclear complex [(Ph3P)2PtQ]+ formed by disintegration of the trinuclear complex [Pt2(μ-S)2(PPh3)4InQ2]+ with byproduct Q. [(Ph3P)2PtQ]+BPh4 was independently prepared from cis-[PtCl2(PPh3)2] and HQ/Me3N, and is the first example of a platinum 8-hydroxyquinolinate complex containing phosphine ligands.  相似文献   

18.
Reaction of Ni(OAc)2 with the symmetric `end-off' compartmental proligand 2,6-[N,N-bis(2-hydroxy-phenylmethyl)-N,N-bis(2-pyridylmethyl)aminomethyl]-4-methylphenol (H3L) in the presence of NaPF6 has been found to generate a homotetranuclear nickel(II) complex [(Ni4HL)(L)(OAc)2(H2O)2(HOAc)2]PF6. The crystal structure of the complex reveals that the complex is donor asymmetric and that the extended supra-ligand periphery is maintained by a tight hydrogen-bond between two pendant phenol/phenoxy groups of adjacent ligands and by further tight hydrogen-bonds between coordinated acetic acid molecules and the remaining pendant phenols of the ligand, generating a double acid salt of the type [CH3COO?H?LH?L?H?OOCCH3]5−. Reaction of H3L with Ni(OAc)2 and NaClO4 in methanol gave the complex [Ni2(HL)(OAc)2(OH2)2][ClO4]. The structure was determined by X-ray diffraction and showed that the complex exists as a dimer promoted by intermolecular hydrogen-bonding.  相似文献   

19.
The kinetics of the formation of the purple complex [FeIII(EDTA)O2]3−, between FeIII-EDTA and hydrogen peroxide was studied as a function of pH (8.22-11.44) and temperature (10-40 °C) in aqueous solutions using a stopped-flow method. The reaction was first-order with respect to both reactants. The observed second-order rate constants decrease with an increase in pH and appear to be related to deprotonation of FeIII-EDTA ([Fe(EDTA)H2O] ⇔ Fe(EDTA)OH]2− + H+). The rate law for the formation of the complex was found to be d[FeIIIEDTAO2]3−/dt=[(k4[H+]/([H+] + K1)][FeIII-EDTA][H2O2], where k4=8.15±0.05×104 M−1 s−1 and pK1=7.3. The steps involved in the formation of [Fe(EDTA)O2]3− are briefly discussed.  相似文献   

20.
The synthesis and characterisation of cis- and trans-[Co(tmen)2(NCCH3)2](ClO4)3 are described. Solvolysis rates have been measured by both 1H NMR spectroscopy and UV-Vis spectrophotometry in dimethyl sulfoxide at 298.2 K. The cis isomer undergoes solvolysis by consecutive first-order reactions, k1=5.61 × 10−4 and k2=5.35 × 10−4 s−1, each with steric retention. The measured solvolysis rate (single step reaction) for the trans isomer is k=1.54 × 10−5 s−1. The solvent exchange rates have been measured by 1H NMR spectroscopy in CD3CN at 298.2 K: kex(cis)=kct + kcc=2.0 × 10−5 and kex(trans)=ktc + ktt=4.56 × 10−6 s−1. From these data, the measured cis-trans isomerisation rate (1.71 × 10−6 s−1) and equilibrium position in CH3CN (17% trans), the steric course for substitution in the exchange processes has been determined: trans reactant - 69% trans product; cis reactant - 99% cis product. Aquation rates for cis- and trans-[Co(tmen)2(NCCH3)2](ClO4)3 have also been determined spectrophotometrically and by NMR; kcis=1.3 × 10−4 and ktrans=2.7 × 10−5 s−1. In both cases the steric course for the primary aquation step is indeterminate because the subsequent steps are faster. Where data are available, the [Co(tmen)2X2]n+ complexes are found to be consistently much more reactive than their [Co(en)2X2]n+ analogues.  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