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.
The reactivity of [Pt2(μ-S)2(PPh3)4] towards a range of nickel(II) complexes has been probed using electrospray ionisation mass spectrometry coupled with synthesis and characterisation in selected systems. Reaction of [Pt2(μ-S)2(PPh3)4] with [Ni(NCS)2(PPh3)2] gives [Pt2(μ-S)2(PPh3)4Ni(NCS)(PPh3)]+, isolated as its BPh4 − salt; the same product is obtained in the reaction of [Pt2(μ-S)2(PPh3)4] with [NiBr2(PPh3)2] and KNCS. An X-ray structure determination reveals the expected sulfide-bridged structure, with an N-bonded thiocyanate ligand and a square-planar coordination geometry about nickel. A range of nickel(II) complexes NiL2, containing β-diketonate, 8-hydroxyquinolinate, or salicylaldehyde oximate ligands react similarly, giving [Pt2(μ-S)2(PPh3)4NiL]+ cations.  相似文献   

2.
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   

3.
Electrospray ionisation mass spectrometry (ESI-MS) has been used as an analytical tool in a wide-ranging scoping study of the alkylation and arylation reactions of [Pt2(μ-S)2(PPh3)4]. From these experiments, the factors that influence the formation of different product species - formed by mono- or di-alkylation - are determined. If the alkylating agent is an alkyl chloride or sulfate, monoalkylation followed by dialkylation of the two sulfido groups occurs, dependent on the alkylating power of the reagent used. For example, n-butyl chloride gives solely [Pt2(μ-S)(μ-SBu)(PPh3)4]+ while dimethyl sulfate gives [Pt2(μ-SMe)2(PPh3)4]2+. This species, previously unisolated is stable in the absence of good nucleophiles, but the addition of potassium iodide results in rapid conversion to [Pt2(μ-SMe)2(PPh3)3I]+. This iodo complex is also observed from the reaction of [Pt2(μ-S)2(PPh3)4] with excess MeI, after the initial formation of mono- and di-methylated species. In these reactions, the iodide presumably displaces a phosphine ligand, which is then quaternised by excess alkylating agent. Changing the alkylating agent to a longer chain alkyl iodide or methyl bromide decreases the rate of alkylation of the sulfide in the initially formed [Pt2(μ-S)(μ-SR)(PPh3)4]+. Mixed-thiolate species of the type [Pt2(μ-SMe)(μ-SR)(PPh3)4]2+ are easily generated by reaction of [Pt2(μ-S)(μ-SR)(PPh3)4]+ with excess Me2SO4 and is also dependent on the avoidance of nucleophiles. Reactions towards α,ω-dialkylating agents are surveyed; the chain length is found to have a dramatic effect on the rate of the second intramolecular cyclisation process, illustrated by a competitive reactivity study involving a mixture of Br(CH2)4Br and Br(CH2)5Br; on completion of the reaction the former gives [Pt2{μ-S(CH2)4S}(PPh3)4]2+ while the latter predominantly gives monoalkylated[Pt2(μ-S){μ-S(CH2)5Br}(PPh3)4]+. The reactivity of o- and p-dihaloxylenes has been explored, with the reaction with p-BrCH2C6H4CH2Br giving the bridged species [(PPh3)4Pt2(μ-S)(μ-SCH2C6H4CH2S)(μ-S)Pt2(PPh3)4]2+. Arylation reactions of [Pt2(μ-S)2(PPh3)4] with halobenzenes and 2-bromoheterocyclic compounds (pyridine, thiophene) are also described.  相似文献   

4.
The reaction of the Tc(II) nitrosyl complex (Bu4N)[Tc(NO)Cl4] with di-(2-picolyl)(NEt)amine in methanol yields the neutral complex [Tc(NO)Cl(py-N(Et)-py)]. The reaction of the Tc(I) nitrosyl complex [Tc(NO)Cl2(HOMe)(PPh3)2] with this tridentate ligand yields cationic [Tc(NO)Cl(py-N(Et)-py)(PPh3)]Cl. These two complexes have been structurally characterized. The reaction of [Tc(NO)Cl2(HOMe)(PPh3)2] with the tetradentate ligand 1,4-bis-(2-pyridylmethyl)-1,4-diazobutane yields a mixture of products including cationic [Tc(NO)Cl(py-NH-NH-py)]Cl and cationic [Tc(NO)Cl(PPh3)(py-NH-NH∼py)]Cl, with a pyridyl terminus left dangling.  相似文献   

5.
Reaction of HSi(OEt)3 with IrCl(CO)(PPh3)2 (5:1 molar ratio) at room temperature for 1 h gives IrCl(H){Si(OEt)3}(CO)(PPh3)2 (1), which is observed by the 1H and 31P{1H} NMR spectra of the reaction mixture. The same reaction, but in 20:1 molar ratio at 50 °C for 24 h produces IrCl(H)2(CO)(PPh3)2 (2) rather than the expected product Ir(H)2{Si(OEt)3}(CO)(PPh3)2 (3) that was previously reported to be formed by this reaction. Accompanying formation of Si(OEt)4, (EtO)3SiOSi(OEt)3, and (EtO)2HSiOSi(OEt)3 is observed. On the other hand, trialkylhydrosilane HSiEt3 reacts with IrCl(CO)(PPh3)2 (10:1 molar ratio) at 80 °C for 84 h to give Ir(H)2(SiEt3)(CO)(PPh3)2 (4) in a high yield, accompanying with a release of ClSiEt3.  相似文献   

6.
Thermolysis of [CpRuCl(PPh3)2] and NaS2CNPr2 or NaS2CNMeBu in methanol affords the ruthenium(II) dithiocarbamate complexes, [CpRu(PPh3)(S2CNPr2)] and [CpRu(PPh3)(S2CNMeBu)], which have been crystallographically characterized. A similar treatment of two equivalents of [CpRuCl(PPh3)2] with the bis(dithiocarbamate) ligand derived from 1,3-homopiperazine affords [{CpRu(PPh3)}2(μ-S2CNC5H10NCS2)].  相似文献   

7.
Reaction of the precursor Ir complex [Ir(H)2(PPh3)2(Me2CO)2]PF6 with 3,6-bis(2-pyridyl)tetrazine (bptz) in CH2Cl2 gave a novel dinuclear Ir hydrido complex [Ir2(H)4(PPh3)4(bptz)](PF6)2 · 4CH2Cl2. Crystallographic study described an interesting coordination environment having a π-π interaction and 1H NMR study showed unique upfield shifts of pyridyl rings that are likely induced by the ring current effect of neighboring PPh3 ligands.  相似文献   

8.
The organometallic tin(IV) complexes [SnPh2(SRF)2] SRF = SC6F4-4-H (1), SC6F5 (2), were synthesized and their reactivity with [MCl2(PPh3)2] M = Ni, Pd and Pt explored. Thus, transmetallation products were obtained affording polymeric [Ni(SRF)(μ-SRF)]n, monomeric cis-[Pt(PPh3)2(SC6F4-4-H)2] (3) and cis-[Pt(PPh3)2(SC6F5)2] (4) and dimeric species [Pd(PPh3)(SC6F4-4-H)(μ-SC6F4-4-H)]2 (5) and [Pd(PPh3)(SC6F5)(μ-SC6F5)]2 (6) for Ni, Pt and Pd, respectively. The crystal structures of complexes 1, 2, 3, 4 and 6 were determined.  相似文献   

9.
The crystal structures of [Cr(NO)(NH3)5](PF6)2 (red) and [Cr(NO)(NH3)5]Cl(PF6) (brown) have been determined. The [Cr(NO)(NH3)5]2+(A) complex cations in these compounds have a slightly distorted octahedral geometry with a strictly linear Cr-N-O arrangement (from symmetry). The short interatomic distances (2.399 Å × 4) between the O (nitrosyl) and H (ammonia in adjacent complex cations) atoms in A(PF6)2 indicate the existence of hydrogen bonds, while the interatomic distances (3.258 Å × 8) between those in ACl(PF6) are much longer, and the hydrogen bonds should be weak in spite of the presence of the smaller counter anion of chloride ion in ACl(PF6). Comparisons of the five crystal structures of A(PF6)2, ACl2, ACl(ClO4), ACl(PF6), and A(ClO4)2 have led to the conclusion that the existence of the strong hydrogen bonds gives red crystals of A(PF6)2, while the absence of hydrogen bonds results in the formation of green crystals of A(ClO4)2 (O ? H, 3.595 Å × 2). The color change of the crystals (from red to green) with the change of outer sphere anions is attributed to the change of the strength of the hydrogen bonding between the complex cations.  相似文献   

10.
The crystal structures of the four-coordinate trans-[Rh(Cl)(CO)(SbPh3)2] (1) and the five-coordinate trans-[Rh(Cl)(CO)(SbPh3)3] (2) are reported, as well as the unexpected oxidative addition product, trans-[Rh(I)2(CH3)(CO)(SbPh3)2] (3), obtained from the reaction of 2 with CH3I. The formation constants of the five-coordinate complex were determined in dichloromethane, benzene, diethyl ether, acetone and ethyl acetate as 163±8, 363±10, 744±34, 1043±95 and 1261±96 M−1, respectively. While coordinating solvents facilitate the formation of the five-coordinate complex, the four-coordinate complex could be obtained from diethyl ether due to the favorable low crystallization energy. The tendency of stibine ligands to form five-coordinate rhodium(I) complexes is attributed mainly to electron deficient metal centers in these systems, with smaller contributions by the steric effects. The average effective cone angle for the SbPh3 ligand in the three crystallographic studies was determined as 139° with individual values ranging from 133 to 145°.  相似文献   

11.
Reaction between the carbonyl, nitrosyl complex, OsCl(CO)(NO)(PPh3)2 (1) and dioxygen results in combination of CO and O2, forming a chelating peroxycarbonyl ligand in the yellow complex, Cl(NO)(PPh3)2 (2). Confirmation of the unique peroxycarbonyl ligand arrangement in 2 is provided by crystal structure determination. When 2 is heated, as a suspension in heptane under reflux, there is a rearrangement to the regular chelating carbonate ligand in the orange complex, Cl(NO)(PPh3)2 (3). The structure of 3 has also been determined by X-ray crystallography. Compound 2 also undergoes the following reactions: with water, releasing CO2 and forming Os(OH)2Cl(NO)(PPh3)2 (4); with HCl releasing CO2 and forming Os(OH)Cl2(NO)(PPh3)2 (5); and with excess triphenylphosphine releasing CO2 and triphenylphosphine oxide forming OsCl(NO)(PPh3)3 (6).  相似文献   

12.
Reaction of [Ru2(O2CMe)4]Cl with K3[Cr(CN)5NO] in water forms Hx[RuII/III2(O2CMe)4]3−x-[Cr(CN)5NO]·zH2O (x = 0.2) that magnetically orders at 4.0 K and possesses an interpenetrating body centered cubic [a = 13.2509(2) Å] structure with random locations of the bridging nitrosyl ligands, and x/3 vacant cation sites. Similarly, the aqueous reaction of [Ru2(O2CMe)4]Cl with Na2[Fe(CN)5NO] forms paramagnetic [Ru2(O2CMe)4]2[Fe(CN)5NO]·H2O, which has a similar tetragonal interpenetrating structure [a = 13.0186(1) Å, c = 13.0699(2) Å] where the NO ligands are presumably nonbridging and 1/3 of the expected cation sites are unoccupied. The presence of uncoordinated NO sites in addition to missing neighboring [Ru2(O2CMe)4]+ units, results in significant vacancies (or holes) in the lattice.  相似文献   

13.
The rhodium dimer [Rh2H(PPh2)2(PPh3)3] was prepared from RhCl(PPh3)3 and K4Sn9 in the presence of 2,2,2-cryptand in ethylenediamine/toluene solvent mixtures. The [K(2,2,2-crypt)]+ salt was isolated and characterized via NMR and X-ray diffraction studies. The solid state structure reveals a binuclear, diphenylphosphido-bridged, 32 electron Rh(I)-Rh(I) complex with edge-shared tetrahedral and square planar Rh centers with overall Cs point symmetry. 1-D and 2-D 1H, 31P, and 31P{1H} NMR experiments were used to characterize the complex.  相似文献   

14.
The new aryl phosphinites PPh2OR (R = 2,4,6-Me3C6H2, 1; R = 2,6-Ph2C6H3, 2) have been prepared from chlorodiphenylphosphine and the corresponding phenols. In these ligands, the ortho-positions of the aromatic phosphite function are blocked by methyl and phenyl substituents, which allows coordination to metal centres without ortho-metallation. Thus, reaction with [PdCl2(cod)] leads to the complexes trans-[PdCl2(PPh2OR)2] (R = 2,4,6-Me3C6H2, 3; R = 2,6-Ph2C6H3, 4), while the reaction with [Rh2(CO)4Cl2] gives trans-[Rh(CO)Cl(PPh2OR)2] (R = 2,4,6-Me3C6H2, 5; R = 2,6-Ph2C6H3, 6). The single-crystal X-ray structure analyses of 3 and 5 confirm the trans-coordination of the new ligands in these square-planar complexes.  相似文献   

15.
Mo(CO)4(LL) complexes, where LL = polypyridyl ligands such as 2,2′-bipyridine and 1,10-phenanthroline, undergo quasi-reversible, one-electron oxidations in methylene chloride yielding the corresponding radical cations, [Mo(CO)4(LL)]+. These electrogenerated species undergo rapid ligand substitution in the presence of acetonitrile, yielding [Mo(CO)3(LL)(CH3CN)]+; rate constants for these substitutions were measured using chronocoulometry and were found to be influenced by the steric and electronic properties of the polypyridyl ligands. [Mo(CO)3(LL)(CH3CN)]+ radical cations, which could also be generated by reversible oxidation of Mo(CO)3(LL)(CH3CN) in acetonitrile, can be irreversibly oxidized yielding [Mo(CO)3(LL)(CH3CN)2]2+ after coordination by an additional acetonitrile. Infrared spectroelectrochemical experiments indicate the radical cations undergo ligand-induced net disproportionations that follow first-order kinetics in acetonitrile, ultimately yielding the corresponding Mo(CO)4(LL) and [Mo(CO)2(LL)(CH3CN)3]2+ species. Rate constants for the net disproportionation of [Mo(CO)3(LL)(CH3CN)]+ and the carbonyl substitution reaction of [Mo(CO)3(LL)(CH3CN)2]2+ were measured. Thin-layer bulk oxidation studies also provided infrared characterization data of [Mo(CO)4(ncp)]+ (ncp = neocuproine), [Mo(CO)3(LL)(CH3CN)]+, [Mo(CO)3(LL)(CH3CN)2]2+ and [Mo(CO)2(LL)(CH3CN)3]2+ complexes.  相似文献   

16.
Abnormal and normal carbene complexes are formed in reactions of 2-pyridylmethylimidazolium salts with [Ir(H)2(PPh3)2(OCMe2)2]BF4 at room temperature in tetrahydrofuran (THF) or dichloromethane (CH2Cl2). Reactions in THF lead to the formation of abnormal carbene (C-5 bound), while reactions in CH2Cl2 lead to formation of normal carbene (C-2 bound).  相似文献   

17.
Reaction of [Rh(CO)2I]2 (1) with MeI in nitrile solvents gives the neutral acetyl complexes, [Rh(CO)(NCR)(COMe)I2]2 (R=Me, 3a; tBu, 3b; vinyl, 3c; allyl, 3d). Dimeric, iodide-bridged structures have been confirmed by X-ray crystallography for 3a and 3b. The complexes are centrosymmetric with approximate octahedral geometry about each Rh centre. The iodide bridges are asymmetric, with Rh-(μ-I) trans to acetyl longer than Rh-(μ-I) trans to terminal iodide. In coordinating solvents, 3a forms mononuclear complexes, [Rh(CO)(sol)2(COMe)I2] (sol=MeCN, MeOH). Complex 3a reacts with pyridine to give [Rh(CO)(py)(COMe)I2]2 and [Rh(CO)(py)2(COMe)I2] and with chelating diphosphines to give [Rh(Ph2P(CH2)nPPh2)(COMe)I2] (n=2, 3, 4). Addition of MeI to [Ir(CO)2(NCMe)I] is two orders of magnitude slower than to [Ir(CO)2I2]. A mechanism for the reaction of 1 with MeI in MeCN is proposed, involving initial bridge cleavage by solvent to give [Rh(CO)2(NCMe)I] and participation of the anion [Rh(CO)2I2] as a reactive intermediate. The possible role of neutral Rh(III) species in the mechanism of Rh-catalysed methanol carbonylation is discussed.  相似文献   

18.
A new tri-cyanometalate building block for heterometallic complexes, [PPh4]2[FeII(Tpms)(CN)3] (2) (PPh4 = tetraphenylphosphonium; Tpms = tris(pyrazolyl) methanesulfonate), has been prepared. Using it as a building block, a one-dimensional chain compound, {[FeII(Tpms)(CN)3][MnII(H2O)2( DMF)2]} · DMF (3), has been synthesized and structurally characterized. The magnetic properties of 3 correspond to a ferromagnetic chain with weak long-range superexchanged magnetic interaction between the high-spin manganese(II) ions.  相似文献   

19.
The reaction of cis-[Os(CO)4Me2] with Me3NO in the THF or MeCN yields the complexes fac-[Os(CO)3(L)Me2] (where L = THF or MeCN). Whereas the THF complex is unstable and only characterised spectroscopically, fac-[Os(CO)3(MeCN)Me2] has been isolated as a white solid and fully characterized by both analytical and spectroscopic methods. These complexes fac-[Os(CO)3(L)Me2] are shown to be useful intermediates. Thus, reaction with PPh3 gives fac-[Os(CO)3(PPh3)Me2] in good yield.Reactions of fac-[Os(CO)3(L)Me2] (L = CO or MeCN) with CPh3PF6 or B(C6F5)3 have been investigated. Whereas cis-[Os(CO)4Me2] showed no reaction with either CPh3PF6 or B(C6F5)3, the reaction of fac-[Os(CO)3(MeCN)Me2] with CPh3PF6 in CH2Cl2 occurred over 16 h at room temperature to give an unstable cationic product and CPh3Me. The reaction was monitored by both IR and NMR spectroscopies. When this reaction of fac-[Os(CO)3(MeCN)Me2] was carried out in the presence of a trapping ligand such as MeCN, the stable cationic product [Os(CO)3(MeCN)2Me]+ could be isolated and identified spectroscopically.  相似文献   

20.
Complex [PdCl(bdtp)](BF4), in presence of AgBF4 or NaBF4, reacts with pyridine (py), triphenylphosphine (PPh3), cyanide (CN), thiocyanate (SCN) or azide (N3) ligands, leading to the formation of the following complexes: [Pd(bdtp)(py)](BF4)2 [1](BF4)2, [Pd(bdtp)(PPh3)](BF4)2 [2](BF4)2, [Pd(CN)(bdtp)](BF4) [3](BF4), [Pd(SCN)(bdtp)](BF4) [4](BF4) and [Pd(N3)(bdtp)](BF4) [5](BF4). These complexes were characterised by elemental analyses, mass spectrometry, conductivity measurements, infrared and NMR spectroscopies. The crystal structure of [2](BF4)2 was determined by single-crystal X-ray diffraction methods. The metal atom is coordinated by two azine nitrogen atoms, and one sulfur atom of the thioether-pyrazole ligand and one triphenylphosphine in a distorted square-planar geometry.  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