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.
In the present redetermination of the complex cis‐tetra­carbonyl­bis­(tri­cyclo­hexyl­phosphine)molybdenum(0), (I), [Mo(C18H33P)2(CO)4] or cis‐{η1‐[P(C6H11)3]2}Mo(CO)4, the Mo atom has a distorted octahedral geometry with a large P—Mo—P angle of 104.8 (1)°. A strong trans influence on the carbonyls in (I) is seen in a shortening of the Mo—C and a lengthening of the C—O distances opposite the phosphines compared with those that are cis. This influence is greatly diminished in the complex penta­carbonyl­(tri­cyclo­hexyl­phosphine)­molyb­denum(0), (II), [Mo(C18H33P)(CO)5] or {η1‐[P(C6H11)3]}­Mo(CO)5, the core of which has a slightly distorted C4v geometry.  相似文献   

2.
DFT calculations at the BP86/TZ2P level were carried out to analyze quantitatively the metal–ligand bonding in transition‐metal complexes that contain imidazole (IMID), imidazol‐2‐ylidene (nNHC), or imidazol‐4‐ylidene (aNHC). The calculated complexes are [Cl4TM(L)] (TM=Ti, Zr, Hf), [(CO)5TM(L)] (TM=Cr, Mo, W), [(CO)4TM(L)] (TM=Fe, Ru, Os), and [ClTM(L)] (TM=Cu, Ag, Au). The relative energies of the free ligands increase in the order IMID<nNHC<aNHC. The energy levels of the carbon σ lone‐pair orbitals suggest the trend aNHC>nNHC>IMID for the donor strength, which is in agreement with the progression of the metal–ligand bond‐dissociation energy (BDE) for the three ligands for all metals of Groups 4, 6, 8, and 10. The electrostatic attraction can also be decisive in determining trends in ligand–metal bond strength. The comparison of the results of energy decomposition analysis for the Group 6 complexes [(CO)5TM(L)] (L=nNHC, aNHC, IMID) with phosphine complexes (L=PMe3 and PCl3) shows that the phosphine ligands are weaker σ donors and better π acceptors than the NHC tautomers nNHC, aNHC, and IMID.  相似文献   

3.
Group 6 metal carbonyls [cis-M(CO)4(amine)(EPh3)] (M = Mo, W; amine = piperidine (pip), pyridine (py); E = As, Sb) have been prepared and characterized. These complexes react thermally in chlorobenzene solutions with phosphine or phosphite ligands (= L) to give cis- and trans-M(CO)4(L)(EPh3) products. Kinetics of amine substitution by L in these complexes, under pseudo-first-order conditions, indicate that these reactions proceed by a rate law that is first-order in concentration of the metal complex. Rate constants and activation parameters for these reactions have been determined and are discussed. Competition studies for the [M(CO)4(EPh3)] intermediates show that these intermediates are highly reactive and react almost indiscriminately with various incoming nucleophiles with slight preference for more basic ones.  相似文献   

4.
Equimolar reactions of molybdenum hexacarbonyl with tris(piperidino)phosphine (L1), bis(morpholino)(phenyl)phosphine (L2), and (Di-isopropylamino)(morpholino)(phenyl) phosphine (L3), which are examples of symmetrically and unsymmetrically substituted tertiary(amino)phosphines, afford the corresponding mono derivatives, Mo[P(NC5H10)3)](CO)5 (1), Mo[P(Ph)(NC4H8O)2](CO)5 (2), and Mo[P(Ph){N(i-C3H7)2}(NC4H8O)](CO)5 (3) in moderate yields as air stable crystalline solids. In the case of L1, some amount of trans-bis derivative, Mo[P(NC5H10)3)]2(CO)4 (4), was also isolated. X-ray structures of 1, 2, and 3 have been determined. Compounds 1 and 2 crystallize in the monoclinic system with space group P21/c, while 3 crystallizes in the triclinic system with space group Pī. While the Mo-P and Mo-Cax bond distances in these complexes are comparable with those of other P-C bonded phosphines, the presence of chiral phosphine seems to induce a relatively weaker interaction with the molybdenum center. Intermolecular hydrogen bonding is seen in compound 1.  相似文献   

5.
Pseudohalogenogermylenes [(iBu)2ATI]GeY (Y=NCO 4 , NCS 5 ) show different coordination behavior towards group 6 metal carbonyls in comparison to the corresponding halogenogermylenes [(iBu)2ATI]GeX (X=F 1 , Cl 2 , Br 3 ) (ATI=aminotroponiminate). The reactions of compounds 4 – 5 and 1 – 3 with cis‐[M(CO)4(COD)] (M=Mo, W, COD=cyclooctadiene) gave trans‐germylene metal complexes {[(iBu)2ATI]GeY}2M(CO)4 (Y=NCO, M=Mo 6 , W 11 ; Y=NCS, M=Mo 7 ) and cis‐germylene metal complexes {[(iBu)2ATI]GeX}2M(CO)4 (M=Mo, X=F 8 , Cl 9 , Br 10 ; M=W, X=Cl 12 ), respectively. Theoretical studies on compounds 7 and 9 reveal that donor–acceptor interactions from Mo to Ge atoms are better stabilized in the observed trans and cis geometries than in the hypothetical cis and trans structures, respectively.  相似文献   

6.
Summary A new method of evaluating the Tolman cone angle from X-ray structural data available from the Cambridge Crystallographic Data Base has been developed and a statistical analysis of the cone angles of the phosphines PPh3, PMe2Ph, PMePh2, PMe3, PEt3 and PCy3 (Cy = cyclohexyl) in transition metal complexes has been completed.  相似文献   

7.
Activation parameters have been obtained for the chelation of Mo(CO)5dpe (dpe = Ph2PCH2CH2PPh2) and of Mo(CO)5dmpe (dmpe = Me2PCH2CH2PMe2) to give cis-Mo(CO)4dpe and cis-Mo(CO)4dmpe respectively. The results are compared with those for the analogous chromium complexes and show that the enthalpy contribution determines the more rapid chelation in the molybdenum complexes. The preparation and properties of the chelate-bridged hetero-metallic complex (CO)5ModmpeMn(CO)4Br are reported. The reaction between Et4N[Mn(CO)4X2] (X = Cl, Br) and bidentate ligands dpe, dmpe and ape (ape = Ph2PCH2CH2AsPh2) in the presence of either silver(I) tetrafluoroborate or Et3OBF4 produces cis-Mn(CO)4X(bidentate) which is identified by infrared and mass spectrometry. At room temperature the Mn(CO)4X(bidentate) complex is rapidly converted to the chelated fac-Mn(CO)3X(bidentate) complex. The chelation process is approximately 104 times more rapid than in the isoelectronic chromium(O) complexes. The preparation and characterisation of fac-Mn(CO)3Br(dmpe), cis-Mn(CO)4Br(PMe3) and fac-Mn(CO)3Br(PMe3)2 are reported.  相似文献   

8.
A rational approach to the synthesis of heterobi‐ or ‐trimetallic complexes based upon self‐assembly of a flexible ditopic catechol‐phosphine ligand with [(cod)PdCl2] and simple metal halides such as GaCl3, BiCl3, SnCl4, or ZrCl4 is described. All products were characterized by spectroscopic and analytical data and single‐crystal X‐ray diffraction studies. The molecular structures can be described in terms of cis‐configured palladium complexes with supramolecular bisphosphine ligands that are formed by the assembly of two phosphine catecholate fragments on a main group/transition metal template. Of particular interest are the distinct decreases in P‐Pd‐P bite angles and P???P distances between the ligating atoms with increasing covalent radii of the templates. The range of these variations is of a magnitude similar to that of the geometrical changes in known families of complexes containing molecular bidentate ligands. Solution NMR studies give further evidence that in several cases the μ2‐bridging coordination of two of the catechol oxygen atoms in the template complexes is broken under the influence of donor solvents, thus allowing the supramolecular ligand to be switched between tetradentate ‐O2P2 and bidentate ‐P2 coordination modes.  相似文献   

9.
Substitution reactions of phosphine ligands, triphenylphosphine (PPh3), tri(m-chlorophenyl)phosphine (m-ClPPh3), tri(p-methoxyphenyl)phosphine (p-MeOPPh3) and tri(benzyl)phosphine (PBz3) with [M(CO)4(PCA)] (M?=?Cr, Mo and W, PCA?=?pyrazinecarboxamide) were found to be dependent on the type of metal and phosphine ligand. The complexes were characterized by elemental analysis, mass spectrometry, and IR and 1H NMR spectroscopy. UV–vis spectra of the complexes in different solvents showed bands due to metal-to-ligand charge transfer.  相似文献   

10.
唐典勇  胡常伟 《化学学报》2008,66(6):647-651
采用密度泛函理论B3LYP方法研究了配体和配位数对乙烯插入杂双核(CO)4Cr(m-PH2)2RhH(Ln) (L=CO或PH3, n=1或2)配合物中Rh—H键反应的影响. 计算结果表明, 六配位乙烯复合物中乙烯与铑之间轨道相互作用主要为乙烯到铑中心的s供体相互作用; 而五配位乙烯复合物中乙烯与铑中心间相互作用涉及乙烯到铑中心的s供体相互作用和铑到乙烯的p反馈作用. PH3配体在热力学上不利于该反应. 处于氢配体对位的膦配体能加速乙烯插入反应. 乙烯插入的五配位反应途径占优势. Cr(CO)4部分的引入降低了乙烯插入反应的活化能.  相似文献   

11.
The new triphosphine (Me2P)2N? N(Me)(PMe2) ( 1 ), has been synthesized in pure form by the reaction of methylhydrazine with dimethylchlorophosphine in the presence of triethylamine. This triphosphine represents a bridge between phosphinoamine (>P? N(R)? P<) and phosphanyl hydrazide (>P? N(R)? N(R)? P<) backbones. Reaction of 1 with cis-[W(CO)4(NHC5H10)2] proceeds via two coordination modes to give four-membered M? P? N? P and five-membered M? P? N? N? P metallacyclic frameworks. The tungsten complex cis-[W(CO)4{(Me2P)2NN(Me)(PMe2)}] ( 2 ) possessing an uncoordinated phosphine moiety to prepare multimetallic organometallic compounds. For example, reactions of 2 with PdCl2(PhCN)2 and PtCl2(COD) produced the trimetallic complexes consisting of W(0)? Pt(II) and W(0)? Pd(II) centers respectively in good yields. The different isomers of these trinuclear complexes have been clearly identified by 31P NMR spectroscopy.  相似文献   

12.
Addition of phosphines to the coordinatively unsaturated species (Cp)2M2(CO)4 (M = Mo, W) and H2Os3(CO)10 yield a series of products incorporating terminal phosphine along with bridging and capping phosphido ligands formed via insertion into PH bonds. The complexes were characterised by 1H and 31P NMR spectroscopy and mass spectrometry.  相似文献   

13.
Preparation and spectroscopical Investigations of M(CO)4L2 and M(CO)3L3 Complexes (M = Cr, Mo, W; L = Me3SiOCH2PMe2, Me2(CH2?CH)SiOCH2PMe2 The coordinating properties of the ligands L1 (?Me3SiOCH2PMe2) and L2 (?Me2ViSiOCH2PMe2)1) have been studied by synthesis and spectroscopic investigations (IR, NMR, MS) of their complexes M(CO)4L2 and M(CO)3L3(M = Cr, Mo, W). The complexes are obtained by replacement of norbornadiene (NBD) in M(CO)4NBD or cycloheptatriene CHT in M(CO)3CHT. Spectroscopic data (v(CO), δ δ) support the σ-donor/-π-acceptor model of the MP bonds.  相似文献   

14.
Summary Upon u.v. irradiation of [Fe(CO)4(PR 3 )] with HSiR3 (HSiR3 = HSiMePh2, PR3 = PPh3; HSiR3 = HSiMe2Cl, PR3 = PPh3 or PMe2Ph; HSiR3 = HSiMeCl2, PR3 = PPh3, PMePh2, PMe2Ph or PMe3; HSiR3 = HSiCl3, PR3 = PPh3, PMePh2, PMe2Ph, PMe3 or PBu 3 n ) the corresponding hydridosilyl complexes [Fe(CO)3H(PR3)SiR3] are formed. The complexes have themer configuration with acis disposition of the hydride and the silyl ligands. Prolonged irradiation with an excess of silane results in the formation of bis-silyl complexes [Fe(CO)3(PR3)(SiR3)2], if electron density at the metal is not too high. Thus, [Fe(CO)3H(PPh3)SiMePh2] and [Fe(CO)3-H(PMe2Ph)SiMe2Cl] can be obtained but not the corresponding bis-silyl complexes. Most bis-silyl complexes are obtained asmer-isomers with acis-arrangement of the silyl ligands. Only for [Fe(CO)3(PR3)(SiCl3)2] with small phosphine ligands (PR3 = PMe3 or PMe2Ph) is thefac-isomer formed.Part VII of this series, ref. (1).  相似文献   

15.
The title compounds have been prepared in low yields starting with cis-(CH3CN)3-Mo(CO)3 or Mo(CO)6 (photochemical route). The new complexes are generally less stable than the corresponding hexaalkylborazine-chromium compounds and react very smoothly with Lewis bases by cleavage of the borazine-metal bond. The IR.-spectra indicate the similarity of type of bonding in hexaalkylborazine-molybdenum and -chromium complexes.  相似文献   

16.
Carbonyl–iridium half-sandwich compounds, Cp*Ir(CO)(EPh)2 (E=S, Se), were prepared by the photo-induced reaction of Cp*Ir(CO)2 with the diphenyl dichalcogenides, E2Ph2, and used as neutral chelating ligands in carbonylmetal complexes such as Cp*Ir(CO)(μ-EPh)2[Cr(CO)4], Cp*Ir(CO)(μ-EPh)2[Mo(CO)4] and Cp*Ir(CO)(μ-EPh)2[Fe(CO)3], respectively. A trimethylphosphane–iridium analogue, Cp*Ir(PMe3)(μ-SeMe)2[Cr(CO)4], was also obtained. The new heterodimetallic complexes were characterized by IR and NMR spectroscopy, and the molecular geometry of Cp*Ir(CO)(μ-SePh)2[Mo(CO)4] has been determined by a single crystal X-ray structure analysis. According to the long Ir…Mo distance (395.3(1) Å), direct metal–metal interactions appear to be absent.  相似文献   

17.
The reduction of [WCl4(PMe3)3] with dispersed sodium, under dinitrogen, gives cis-[W(N2)2(PMe3)4], while under ethylene trans-[W(C2H4)2(PMe3)4] is obtained. The ethylene complex can also be prepared by displacement of the dinitrogen molecules in cis-[W(N2)2(PMe3)4] by ethylene at room temperature and pressure. Interaction of cis-[M(N2)2(PMe3)4] complexes (M = Mo, W), with PMe3, under helium or argon, yields [M(N2)(PMe3)5]. The molybdenum complex crystallizes in the orthorhombic space group Pnma, with a 22.063(6), b 12.106(4), c 9.745(4) Å. The Mo—P distance trans to the dinitrogen ligand (2.483(7) Å) is slightly longer than the average of the other four Mo—P bonds (2.460(5) Å).  相似文献   

18.
Irradiation of the title compound 1 through quartz in toluene solution at − 20°C produces cis-Mo(CO)2(PMe3)4 and OPMe3 as the major products along with lesser amounts of mer- and fac-Mo(CO)3(PMe3)3 and Mo(CO)PMe3) 5. Irradiation of 1 through Pyrex produces in addition substantial amounts of an unstable species formulated as trans-Mo (CO)2(PMe3)4.  相似文献   

19.
Irradiation of bis(phosphine) tetracarbonyl complexes L2M(CO)4 (M = Cr, Mo, W) in the presence of donor ligands (amine, nitrile, halide ion) leads, via loss of one phosphine ligand, to neutral (LL′M(CO)4) or ionic ([LM(CO)4X]?) metal carbonyl compounds. The use of this reaction as the first step in a general synthesis of unsymmetrically disubstituted derivatives of Group VIA hexacarbonyls is discussed.  相似文献   

20.
A general approach to the first compounds that contain rhenium–germanium triple and double bonds is reported. Heating [ReCl(PMe3)5] ( 1 ) with the arylgermanium(II) chloride GeCl(C6H3‐2,6‐Trip2) ( 2 ; Trip=2,4,6‐triisopropylphenyl) results in the germylidyne complex mer‐[Cl2(PMe3)3Re?Ge? C6H3‐2,6‐Trip2] ( 4 ) upon PMe3 elimination. An equilibrium that is dependent on the PMe3 concentration exists between complexes 1 and 4 . Removal of the volatile PMe3 shifts the equilibrium towards complex 4 , whereas treatment of 4 with an excess of PMe3 gives a 1:1 mixture of 1 and the PMe3 adduct of 2 , GeCl(C6H3‐2,6‐Trip2)(PMe3) ( 2 ‐PMe3). Adduct 2 ‐PMe3 can be selectively obtained by addition of PMe3 to chlorogermylidene 2 . The NMR spectroscopic data for 2 ‐PMe3 indicate an equilibrium between 2 ‐PMe3 and its dissociation products, 2 and PMe3, which is shifted far towards the adduct site at ambient temperature. NMR spectroscopic monitoring of the reaction of complex 1 with 2 and the reaction of complex 4 with PMe3 revealed the formation of two key intermediates, which were identified to be the chlorogermylidene complexes cis/trans‐[Cl(PMe3)4Re?Ge(Cl)C6H3‐2,6‐Trip2] (cis/trans‐ 3 ) by using NMR spectroscopy. Labile chlorogermylidene complexes cis/trans‐ 3 can be also generated from trans‐[Cl(PMe3)4Re?Ge? C6H3‐2,6‐Trip2]BPh4 ( 9 ) and (nBu4N)Cl at low temperature, and decompose at ambient temperature to give a mixture of complexes 1 and 4 . Complex 4 reacts with LiI to give the diiodido derivative mer‐[I2(PMe3)3Re?Ge? C6H3‐2,6‐Trip2] ( 5 ), which undergoes a metathetical iodide/hydride exchange with Na(BEt3H) to give the dihydrido germylidyne complex mer‐[H2(PMe3)3Re?Ge? C6H3‐2,6‐Trip2] ( 6 ). Carbonylation of 4 induces a chloride migration from rhenium to the germanium atom to afford the chlorogermylidene complex mer‐[Cl(CO)(PMe3)3Re?Ge(Cl)C6H3‐2,6‐Trip2] ( 7 ). Similarly, MeNC converts complex 4 into the methylisocyanide analogue mer‐[Cl(MeNC)(PMe3)3Re?Ge(Cl)C6H3‐2,6‐Trip2] ( 8 ). Chloride abstraction from 4 by NaBPh4 in the presence of PMe3 gives the cationic germylidyne complex trans‐[Cl(PMe3)4Re?Ge? C6H3‐2,6‐Trip2]BPh4 ( 9 ). Heating complex 4 with cis‐[Mo(PMe3)4(N2)2] induces a germylidyne ligand transfer from rhenium to molybdenum to afford the germylidyne complex trans‐[Cl(PMe3)4Mo?Ge? C6H3‐2,6‐Trip2] ( 10 ). All new compounds were fully characterized and their molecular structures studied by X‐ray crystallography, which led to the first experimentally determined Re? Ge triple‐ and double‐bond lengths.  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