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.
《Polyhedron》1999,18(20):2605-2608
The interaction of iron carbonyls, Fe(CO)5, Fe2(CO)9 and Fe3(CO)12 with Me3NO occurs according to a one-electron redox-disproportionation scheme giving rise to iron carbonyl radical anions: Fe2(CO)8·− (1), Fe3(CO)12·− (2), Fe3(CO)11·− (3) and Fe4(CO)13·− (4). The role of Me3NO, inducing CO-substitution, consists of the generation of reactive 17-electron species with a labile coordination sphere in which the substitution for other ligands occurs, resulting from fast ligand and electron exchange in the confines of the ETC-reaction.  相似文献   

2.
The reactions of dimethylthiocarbamoyl chloride with a number of neutral and ionic iron carbonyl complexes in tetrahydrofuran are described. A variety of unusual products were obtained, viz. Fe(CO)2(S2CNMe2)2 from Fe(CO)5; Fe(CO)2(S2CNMe2)(CSNMe2) from Fe2)CO)9, Fe3(CO)12, and Fe(CO)42?; [Fe-(CO)2(S2CNMe2)(CNMe2)(CNMe2)2S]+ from Fe(CO)42?, and Fe4(CO)12S(CSNMe2)-(CNMe2) from Fe2(CO)82?, as well as Fe2(CO)6(CSSEt)2 from Fe2(CO)9 and ClCSSEt. The structures and behavior and some reactions of these complexes are described.  相似文献   

3.
Quaternary ammonium borohydrides, used directly or generated in phase transfer reactions, are highly effective reagents for preparing metal carbonyl anions from metal carbonyls [Mo(CO)6, Mn2(CO)10, Re2(CO)10, CO2(CO)8, Fe3(CO)12, Ru3(CO)12 and (η5-C5H5)2Mo2(CO)6] and from some metal carbonyl halides [BrMn(CO)5 and η5-C5H5Mo(CO)3Cl]. Where strongly basic anions would be formed from a halide [BrMn(CO)4PPh3 and η5-C5H5Ru(CO)2Br], the reactions provide efficient syntheses of the corresponding hydrides instead. The anion η5-C5H5Fe(CO)2? is not accessible by these techniques; reaction of η5-C5H5Fe(CO)2Br yields the iron dimer (via the highly nucleophilic anion) and the dimer is unreactive toward Q+BH4?. Reductions of Re2(CO)10 conducted in CH2Cl2 provide Re2(CO)9Cl? in high yield.  相似文献   

4.
《Polyhedron》1987,6(11):1959-1970
The interaction of Fen(CO)m, (n and m equal 1 and 5, 2 and 9, 3 and 12, respectively) with 2-methyl-2-nitrosopropane and sodium salts of nitromethane and nitrocyclohexane was studied. The initial stages of the process, following the activating complex-formation, involves redox disproportionation to give rise to the radical Fe(I) carbonyl complexes and radical anions Fe2(CO)8. (I) Fe3(CO)11. (II), Fe4(CO)13. (III) and Fe3(CO)12. (IV). Also, radical anions IIV are formed in the interaction of salts of carbonyl ferrate anions Na2Fe(CO)4·1.5 diox and PPN2[Fen(CO)m−1] (where PPN = (PPh3)2N+), with nitro- and nitroso-tert-butane.Radical anions IIII act as catalytically active species in the coordination sphere of which the nitro compounds undergo a successive deoxygenation to nitrene radical complexes with their subsequent carbonylation to isocyanates. A scheme of the reductive carbonylation is proposed.  相似文献   

5.
The equilibrium structures and vibrational frequencies of the iron complexes [Fe0(CN)n(CO)5?n]n? and [FeII(CN)n(CO)5?n]2?n (n = 0–5) have been calculated at the BP86 level of theory. The Fe0 complexes adopt trigonal bipyramidal structures with the cyano ligands occupying the axial positions, whereas corresponding Fe2+ complexes adopt square pyramidal structures with the cyano ligands in the equatorial positions. The calculated geometries and vibrational frequencies of the mixed iron Fe0 carbonyl cyanide complexes are in a very good agreement with the available experimental data. The nature of the Fe? CN and Fe? CO bonds has been analyzed with both charge decomposition and energy partitioning analysis. The results of energy partitioning analysis of the Fe? CO bonds shows that the binding interactions in Fe0 complexes have 50–55% electrostatic and 45–50% covalent character, whereas in Fe2+ 45–50% electrostatic and 50–55% covalent character. There is a significant contribution of the π‐ orbital interaction to the Fe? CO covalent bonding which increases as the number of the cyano groups increases, and the complexes become more negatively charged. This contribution decreases in going from Fe0 to Fe2+ complexes. Also, this contribution correlates very well with the C? O stretching frequencies. The Fe? CN bonds have much less π‐character (12–30%) than the Fe? CO bonds. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

6.
The reaction between Fe(CO)5, and group V donor ligands L, (L  PPh3, AsPh3, SbPh3, PMePh2, PMe2Ph, Asme2Ph, P(C6H11)3, P(n-Bu)3, P(i-Bu)3, P(OPh)3, P(OEt)3, P(OMe)3) in the presence of [(η5-C5Me5Fe(CO)2]2 (R  H, Me) or [(η5-C5Me5)Fe(CO)2]2 as catalyst in refluxing toluene, rapidly gives the complexes Fe(CO)4L in yields > 85%. The reaction rate is essentially independent of the nature of L for [(η5-C5Me5)Fe(CO)2]2 as catalyst. For the other catalysts, the rate is influenced predominantly by the steric properties of L. These results are interpreted in terms of the interaction between the catalyst and the ligand L to give derivatives of the type (η5-C5H4R)2Fe2,(CO)3,(L). These derivatives were also found to catalyse the reaction between Fe(CO)5, and L. The complexes [(η-C5H4R)Fe(CO)2]2 (R  H, Me) and [(η5-C5Me5)Fe(CO)2]2 also catalyse the reaction between Mn2(CO)10 and PPh3 to give Mn2(CO)8- PPh3)2 in > 80% yield.  相似文献   

7.
The photolysis of iron carbonyl (Fe(CO)5) adsorbed on titanium dioxide (TiO2, anatase) was studied by FT-IR spectroscopy. When adsorbed Fe(CO)5 is illuminated by visible and near-UV light, the IR spectrum of its photolysis products is hardly observed, indicating that most of the Fe(CO)5 is photodecomposed to iron(0) or iron oxides on TiO2. The carbon monoxide (CO) evolution rate upon illumination depends on the wavelength of light; 433 nm light is more effective for CO evolution than 366 nm light. This result implies that the band-gap excitation of TiO2 has little effect on the photolysis of adsorbed Fe(CO)5, since the absorption edge of TiO2 (anatase) lies at around 400 nm. The effects of substrates on the photolysis of adsorbed Fe(CO)5 are discussed with reference to previous results obtained for aluminium oxide (Al2O3) and silicon dioxide (SiO2), on which the photolysis leads to the formation of Fe2(CO)9 or Fe3(CO)12.  相似文献   

8.
Kinetic data for the exchange of 1,3-cyclohexadiene with (η4-benzylideneacetone)Fe(CO)2L complexes (L = CO, PPh3-xMex (x = 0-2) or P(OPh)3) to give (η4-1,3-cyclohexadiene)Fe(CO)2L derivatives indicate a mechanism involving stepwise competing D and Id opening of the ketonic M-CO π-bond. Rates increase in the order CO ? PPh3 ≈ P(OPh)3 > PPh2Me ? PPhMe2, and both steric and electronic factors appear to be important. (η4-1,3-cyclohexadiene)Fe(CO)2L complexes of potential use in enantioselective synthesis (L=(+)-Ph2P(menthyl) or (+)-Ph2PCH2CH(Me)Et) may be prepared via their (η4-benzylideneacetone)Fe(CO)2L complexes.  相似文献   

9.
Die Struktur von Tetracarbonyl(1,1,2,3,3-pentaphenyltriphosphan-P2)-eisen(0) ( I ) wurde röntgenographisch bestimmt. I kristallisiert in der Raumgruppe P1 . Es weist trigonal bipyramidale Koordination am Fe-Atom auf. Der Phosphanligand befindet sich in äquatorialer Position mit einem Fe? P? ;Abstand von 226,2(1) pm. Der mittlere Abstand Fe? C für die axial gebundenen CO-Gruppen beträgt 178,8(3) pm und unterscheidet sich nicht signifikant von dem für die äquatorial gebundenen CO-Gruppen. Die P? P? Abstände im Triphosphanliganden unterscheiden sich um 1,9 pm [225,6(2) und 223,7(1) pm]. Mit PtCl2 bildet I einen Komplex ( III ) der Zusammensetzung (CO)4FePPh(PPh2)2PtCl2, in dem die terminalen Phosphoratome an das Platin koordiniert sind und einen viergliedrigen Ring bilden. Beim Erhitzen lagert sich I unter Wanderung der Fe(CO)4-Einheit von der medialen in die terminale Position der Triphosphankette in den isomeren Komplex II um. Die Bestrahlung von I mit UV-Licht liefert als Hauptprodukte Ph2P? PPh2 und (CO)4Fe? PPhH? PPh2. Reactions of Coordinated Ligands. IX. Reactions and Molecular Structure of Tetracarbonyl(1,1,2,3,3-pentaphenyltriphosphane-P2)iron(0) – a Monosubstitution Product of Pentacarbonyliron with an Equatorial Phosphane Ligand The structure of tetracarbonyl(1,1,2,3,3-pentaphenyltriphosphane-P2)iron(0) (I) was determined by X-ray analysis. The molecule displays a trigonal-bipyramidal coordination geometry at the iron atom with the phosphane ligand in an equatorial position and an Fe? P distance of 226.2(1) pm. The average Fe? C axial distance of 178.8(3) pm is not significantly different from that of 178.0(5) pm for the average Fe? C equatorial distance. A difference of 1.9 pm is observed between the two P? P distances [225.6(2) and 223.7(1) pm]. With PtCl2 I yields a complex of composition (CO)4FePPh(Ph2P)2PtCl2 (III) in which the terminal phosphorus atoms of I are coordinated to platinum forming a four membered ring system. On heating, complex II , an isomer of I , is formed by migration of the Fe(CO)4-unit from the medial to the terminal position within the triphosphane chain. Irradiation of I with UV light affords Ph2P? PPh2 and Fe(CO)4? PPhH? PPh2 as main products.  相似文献   

10.
Substituted carbonylmanganese cations [Mn(CO)5L]+, where L=py, PPh3 and PPh2Me, readily react with various organoborate anions (tetramethylborate, methyltriphenylborate and tetraphenylborate) in THF solution to afford a mixture of dimanganese carbonyls, hydridomanganese carbonyls and alkylmanganese carbonyls. The formation of the dimanganese carbonyl dimers as well as the hydridomanganese carbonyls suggests the involvement of 19-electron carbonylmanganese radicals that stem from an initial electron transfer. On the other hand, the acetonitrile-substituted analogue [Mn(CO)5(CH3CN)]+ reacts with the same borate anions to afford the alkylated RMn(CO)5, where R=CH3 and C6H5, as the sole carbonylmanganese product. As such, this alkylative annihilation is best formulated as a direct attack on the carbonyl carbon by the borate nucleophile. The two different pathways can be understood in terms of the balance between the electrophilicity of the carbonyl ligand and the electron affinity of the carbonylmanganese cation.  相似文献   

11.
Reactions of ruthenium(II) carbonyl complexes of the type [RuHCl(CO)(PPh3)2(B)] [B?=?PPh3, pyridine (py), piperidine (pip) or morpholine (mor)] with bidentate Schiff base ligands derived from the condensation of 2-hydroxy-1-naphthaldehyde with aniline, o-, m- or p-toluidine in a 1?:?1 mol ratio in benzene resulted in the formation of complexes formulated as [RuCl(CO)(L)(PPh3)(B)] [L?=?bidentate Schiff base anion, B?=?PPh3, py, pip, mor]. The complexes were characterized by analyses, IR, electronic and 1H NMR spectroscopy, and cyclic voltammetric studies. In all cases, the Schiff bases replace one molecule of phosphine and a hydride ion from the starting complexes, indicating that Ru–N bonds in the complexes containing heterocyclic nitrogenous bases are stronger than the Ru–P bond to PPh3. Octahedral geometry is proposed for the complexes.  相似文献   

12.
Photolysis of (η5-C5H5Fe(CO)(CNMe)2]PF6 in the presence of excess nucleophiles resulted in efficient substitution of the carbonyl ligand, generating the new isocyanide complexes (η5-C5H5Fe(CNMe)2)(L)]PF6 (L = PPh3, AsPh3, SbPh3, pyridine, acetonitrile, and ethylene). Similar reactions of (η5-C5H5Fe(CO)2)(CNMe)PF6 led to sequential replacement of both carbony groups with the exception of L  ethylene. No evidence of photochemical isocyanide substitution was found. The same carbonyl complexes failed to reach with L thermally. In the absence of light, ethylene, pyridine, and acetonitrile complexes were found to disporportionate in the manner [η5-C5H5Fe(CNMe)(L)2]PF6→ [η5C5H5Fe(CNMe)2(L)]PF6 → [η5-C5H5Fe(CNMe)3]PF6 with the first rearrangement occurring much faster than the second. The new isocyanide complexes are characterized by their infrared and NMR (1H, 13C) spectra.  相似文献   

13.
Synthesis of S2Fe2(CO)6 and S2Fe3(CO)9 by direct substitution of the carbonyl ligands in iron carbonyls with sulfur is described. The relative reactivities were determined for both iron carbonyls and organosulfur compounds in reactions of R2S, R2S2 and RSH (R = alkyl, aryl) with Fe(CO)5, Fe2(CO)9 and Fe3(CO)12. The properties of the [RSFe(CO)3]2 obtained are studied. New methods are found for their regeneration in the form of sulfides, disulfides and thiols.  相似文献   

14.
[Fe2sb‐CO)(CO)3(NO)(μ‐PtBu2)(μ‐Ph2PCH2PPh2)]: Synthesis, X‐ray Crystal Structure and Isomerization Na[Fe2(μ‐CO)(CO)6(μ‐PtBu2)] ( 1 ) reacts with [NO][BF4] at —60 °C in THF to the nitrosyl complex [Fe2(CO)6(NO)(μ‐PtBu2)] ( 2 ). The subsequent reaction of 2 with phosphanes (L) under mild conditions affords the complexes [Fe2(CO)5(NO)L(μ‐PtBu2)], L = PPh3, ( 3a ); η‐dppm (dppm = Ph2PCH2PPh2), ( 3b ). In this case the phosphane substitutes one carbonyl ligand at the iron tetracarbonyl fragment in 2 , which was confirmed by the X‐ray crystal structure analysis of 3a . In solution 3b loses one CO ligand very easily to give dppm as bridging ligand on the Fe‐Fe bond. The thus formed compound [Fe2(CO)4(NO)(μ‐PtBu2)(μ‐dppm)] ( 4 ) occurs in solution in different solvents and over a wide temperature range as a mixture of the two isomers [Fe2sb‐CO)(CO)3(NO)(μ‐PtBu2)(μ‐dppm)] ( 4a ) and [Fe2(CO)4(μ‐NO)(μ‐PtBu2)(μ‐dppm)] ( 4b ). 4a was unambiguously characterized by single‐crystal X‐ray structure analysis while 4b was confirmed both by NMR investigations in solution as well as by means of DFT calculations. Furthermore, the spontaneous reaction of [Fe2(CO)4(μ‐H)(μ‐PtBu2)(μ‐dppm)] ( 5 ) with NO at —60 °C in toluene yields a complicated mixture of products containing [Fe2(μ‐CO)(CO)4(μ‐H)(μ‐PtBu2)(μ‐dppm)] ( 6 ) as main product beside the isomers 4a and 4b occuring in very low yields.  相似文献   

15.
The 57Fe Mössbauer spectra of Fe3(CO)12-related clusters [Fe3(CO)11]2−, [Fe2Ru(CO)12], [FeRu2(CO)12], [Fe3(CO)11PPh3], [Fe3(CO)11PPh2Me], [Fe3(CO)11PPhMe2], [Fe3(CO)9(PPh2Me)3], [Fe2Ru(CO)11P(OMe)3], [FeRu2(CO)11PPh3] and [FeRu2(CO)10(PPh3)2] have been recorded at 78 K. The data are compared with published data for other M3 clusters.Generally, the isomer shifts (δ) fall within a narrow range, for example with compounds containing Fe or Fe and Ru and four or five CO ligands per metal, all δ values lie between 0.29 and 0.36 mm s−1 even though the ligands may be terminal, doubly bridging or triply bridging. Values of quadrupole splitting (Δ) are much more susceptible to changes in the Fe environment, for example the Fe(CO)4 sites have Δ values from about zero {Fe(CO)4t in [Fe3(CO)12)]} to 1.52 {Fe(CO)3tCOtbr in [Fe3(CO)11]2−}. The quadrupole splitting of the Fe site in [FeRu2(CO)12] (0.77 mm s−1) clearly indicates that the structure of this cluster is not exactly similar to that of [Ru3(CO)12]. Substitution of CO by phosphine in general leads to small changes in Δ and Δ if the geometry of the Fe site is unaltered. However, Δ especially can be affected if phosphine substitution cause changes in geometry or if there is multiple substitution.  相似文献   

16.
During the past 10 years iron‐catalyzed reactions have become established in the field of organic synthesis. For example, the complex anion [Fe(CO)3(NO)]?, which was originally described by Hogsed and Hieber, shows catalytic activity in various organic reactions. This anion is commonly regarded as being isoelectronic with [Fe(CO)4]2?, which, however, shows poor catalytic activity. The spectroscopic and quantum chemical investigations presented herein reveal that the complex ferrate [Fe(CO)3(NO)]? cannot be regarded as a Fe?II species, but rather is predominantly a Fe0 species, in which the metal is covalently bonded to NO? by two π‐bonds. A metal–N σ‐bond is not observed.  相似文献   

17.
The interaction of iron carbonyls Fe n (CO) m (wheren = 1,m = 5;n = 2,m = 9;n = 3,m = 12) with anionic Lewis bases (H, F, Cl, Br , I, CN, SCN, N3 , MeSO3 , MeCO2 , CF3CO2 , S2 , CO3 2–, and SO4 2–) passes through two-stage redox-disproportionation. The first stage is the formation of an iron carbonyl-base complex, [Fe n (CO) m–1C(O)L], and the second is a single-electron reduction of this complex by another molecule of the initial iron carbonyl, giving rise to Fe(l) and Fe(–l) derivatives.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 248–249, January, 1996.  相似文献   

18.
Formation of Organosilicon Compounds. 80. Si-Metalation of 1,3,5-Trisilacyclohexanes by Means of Trisition Metal Complexes Several Si-transition metal-substituted 1,3,5-trisilacyclohexanes are reported. l-Bromo-1,3,5-trisilacyclohexane reacts with the metal carbonyl anions W(CO)5cp?, Mo(CO)3cp-, Cr(CO)3cp?, Mn(CO)3?, Fe(CO)2cp?, or Co(CO)4minus;, resp., yielding monosubstituted derivatives as 6, e. g.(cp = π-cyclopentadienyl). 1,3-Dibromo-1,3,5-trisilacyclohexane forms disubstituted compounds aa 7, e. g., with 2 moles of the metal carbonyl anions Fe(CO)2cp?, Mn(CO)5? or Co(CO)4?. Starting from (H2c? SiHBr)3 compound 13 is accessible by reaction with KCo(CO)4. In the soluted compounds the metal carbonyl groups occupy the equatorial positions in the chair form of the six membered ring. The reaction of 13 with Co2(CO)8 yields 17 , whereas 6 preferrably forms 18 . Starting from (H2C? SiH2)3 the reaction with Co2(CO)2 preferrably yields 19. The reported compounds are crystalline, air – and moisture – sensitive. The reported formulae are assured by analysis, IR, and NMR investigations.  相似文献   

19.
Reaction of carbene‐stabilized disilicon ( 1 ) with Fe(CO)5 gives the 1:1 adduct L:Si?Si[Fe(CO)4]:L (L:=C{N(2,6‐Pri2C6H3)CH}2) ( 2 ) at room temperature. At raised temperature, however, 2 may react with another equivalent of Fe(CO)5 to give L:Si[μ‐Fe2(CO)6](μ‐CO)Si:L ( 3 ) through insertion of both CO and Fe2(CO)6 into the Si2 core, which represents the first experimental realization of transition metal‐carbonyl‐mediated cleavage of a Si?Si double bond. The structures and bonding of both 2 and 3 have been investigated by spectroscopic, crystallographic, and computational methods.  相似文献   

20.
Synthesis of diastereoisomeric manganese complexes is described, starting from LMn(CO)3, carbonyl substitutions gave LMn(CO)(PPh3)[P(OMe)3] (L = h5-C5H3, 1-CO2Me, 3-Me).  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