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.
The reactions of H2Os3(CO)10, Ia and H2Os3(CO)9PMe2Ph, Ib with CF3CN have been investigated. Both la and Ib react with CF3CN to give the products HOs3[μ-η2-(CF3)CNH](CO)9Land HOs3[μ-η1-NC(H)CF3](CO)9L, IIa, IIIa, L = CO; IIb and IIIb, L = PMe2Ph. IIb and IIIb have been characterized crystallographically. In each, one nitrile molecule was added to the cluster and one hydride ligand was transferred to the nitrile ligand, but in IIb the hydride was transferred to the nitrogen atom to form a CF3CNH ligand which bridges an edge of the cluster while in IIIb the hydride was transferred to the carbon atom to form a CF3(H)CN ligand which also bridges an edge of the cluster. On the basis of spectroscopic measurements IIa and IIIa are believed to have analogous structures. An isotope scrambling experiment established that the formation of Ilia occurs by an intramolecular process. IIa was decarbonylated to yield the compound HOs332-(CF3)CNH](CO)9, which is believed to contain a triply-bridging iminyl ligand. Ilia reacts with PMe2Ph to give two mono-substitution products, one of which is IIIb.  相似文献   

2.
When Cl atoms react with CHClCHCl in the presence of O2 at 31°C, a long-chain oxidation occurs. The products are the geometrical isomer of the starting olefin and CHClO, HCl, CO, and CCl2O. The quantum yields of the oxygen-containing products are the same with both isomers and are Φ{CHClO} = 30, Φ{CO} = 11.7, and Φ{CCl2O} = 1.29. The chlorine atom adds to the olefin and is followed by O2 addition. The reaction then proceeds with k6a/k6b = 19 and k7a/k7 ~ 0.5, where k7k7a + k7b. The CCl2H radical oxidizes to regenerate the chain carrier. O(3P) reacts with CHClCHCl at 25°C with a rate coefficient of 6.6 × 108 M?1 sec?1 for trans-CHClCHCl and 2.8 × 108 M?1 sec?1 for cis-CHClCHCl. The reaction channels are with k1a/k1 = 0.23 and 0.28, respectively, for the cis and trans compounds. Reaction (1b) occurs < 4% of the time. Reaction (1c) leads to polymer production and presumably, via redissociation, to the geometrical isomer of the starting olefin. In the presence of O2 the same long-chain oxidation is observed as in the chlorine-atom initiated oxidation. The chain-initiating step is   相似文献   

3.
High pressure IR and UV spectroscopic experiments confirm the Heck and Breslow mechanism of the hydroformylation of 1-octene and cyclohexene with Co2(CO)8 as the starting catalyst. The major repeating unit is HCo(CO)4, which is formed via the reaction of acylcobalt tetracarbonyl with H2. The rates are 6.7 × 10?4 mol l?1 min?1 and 8.8 × 10?5 mol l?1 min?1 for 1-octene and cyclohexene, respectively at 80°C and 95 bar CO/H2 = 1 in methylcyclohexane. The alternative reaction of RCOCo(CO)4 with HCo(CO)4 is only a minor pathway, with rates of 1.8 × 10?5 mol l?1 min?1 and 1.1 × 10?5 mol l?1 min?1 for 1-octene and cyclohexene, respectively. It represents an exit from the catalytic cycle. The activation of the catalyst precursor Co2(CO)8 is the slowest step of the reaction.  相似文献   

4.
The compound hydrido cyclopentadienyliron dicarbonyl has been shown by infrared and proton NMR to be present in substantial quantities during hydroformylation of propene and 1-pentene in the presence of [(η5-C5H5)Fe(CO)2]2. This and related observations strongly suggest that the hydride is an important link in the catalytic cycle as is HCo(CO)4 in Co2(CO)8-catalyzed olefin hydroformylation.  相似文献   

5.
Relative rate measurements of the reactions of the HO-radical with CO [HO + CO → H + CO2 (1)] and with isobutane [HO + iso-C4H10 → H2O + t-(or iso-)C4H9 (3)] have been made through the photolysis of dilute mixtures of HONO with CO, iso-C4H10, NO2, and NO in simulated air at 700 and 100 torr pressure and 25 ± 2°C. In situ, long path, FT-IR analysis of the reactants and products provided essentially continuous monitoring of the reactions during the course of the experiments. The kinetic analysis of the data coupled with Greiner's estimate of k3 give new estimates of k1 = 439 ± 24 ppm?1 min?1 in air at 700 torr and k1 = 203 ± 29 ppm?1 in air at 100 torr. The results confirm the recent conclusions of Cox and Sie and their co-workers that the rate constant for reaction (1) is pressure dependent. Modeliers of the chemical changes which occur in the troposphere should adopt a new value for the rate constant k1 which is about a factor of two larger than that in current use by most groups.  相似文献   

6.
The reaction of (HMe2Si)2NH with Co2(CO)8 gives the complex [Co2(CO)7(SiMe2)2NH2]+[Co(CO)4]. Its thermal decomposition starts with dissociation into the “acid” HCo(CO)4 and the “base” Co2(CO)7(SiMe2)2NH. After that, the base and the initial complex decompose further under the action of HCo(CO)4. The final products of this reaction are CO, NH3, Co, volatile dimethylcyclosilazane, and a solid residue consisting of cobalt particles encapsulated into a polymethylsiloxane matrix and possessing properties of mixed para- and ferromagnetics with an ultimate specific magnetization of 64–74 G g−1. Tetramethyldisilazane reacts with iron pentacarbonyl under UV irradiation to give relatively stable 1,3-bis(tetracarbonylthydrideiron)-1,1,3,3-tetramethyldisilazane. This product contains Fe−H…N hydrogen bonds, which stabilize it against dehydrogenation and cyclization to diironcyclodisilazane. Thermal decomposition of this product was investigated. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2537–2544, December, 1998.  相似文献   

7.
The reaction of Ph3COH with 2 mol of HCo(C0)4 gives Ph3CH in quantitative yield. The reaction is cleanly second order (κ2  2.50 X 10-4 1 mol-1 s-1, in CH2Cl2 at 20°C), first order with respect to each reactant. The rate increases markedly with increase in solvent polarity, suggesting Ph3C+ as an intermediate. The rate of the reaction of HCo(C0)4 with
is more than 103 as fast as with Ph3COH. No evidence for the functioning of HCo(C0)4 as a hydride donor could be secured.  相似文献   

8.
Decacarbonyldirhenium reacts with LiSi(C6H5)3 to yield, on subsequent alkylation with FCH3SO3 or (C2H5)3OBF4, the equatorial nonacarbonyl[triphenylsily(alkoxy)carbene] dirhenium complexeseq-(CO)9Re2C(OR)Si(C6H5)3(Ia, R - CH3; Ib, R - C4H8OCH3; Ic. R - C2H5). Reactions of these compounds with Al2Cl6 or Al2Br6 produce novel binuclear, cationic silycarbyne complexes, ax-[(CO)9Re2CSi(C6H5)3]+ AlX4- (IIa, X - Cl; IIb, X - Br). Treatment of these complexes with alcohols results in formation of the axial nonacarbonyl(carbene)dirhenium complexesax-(CO)9Re2C(OR)Si(C6H5)3 (IIIa, R - CH3; IIIb, R - C2H5). The isomeric carbene complexes Ia and IIIa react with dialkylamine affording the isomeric aminocarbene complexeseq-(CO)9Re2C(CH3)2]-Si(C6H5)3 (V) andax-(CO)9Re2Cl(NR2)Si(C6H5)3 (IVa, R - CH3; IVb, R - C2H5). Reaction conditions, properties and spectroscopic data of the new compounds are reported.  相似文献   

9.
The cis-trans interconversion of olefins in the system W(CO)6 + CCl4 + 2-butene is studied, both with initial irradiation of a solution of W(CO)6 in CCl4 (photoinduction), and with continuous irradiation of the system, for cis- and trans-butene concentrations between 0.09 and 6.0 M. Analysis of the results of the photoinduction experiments (rate of conversion and kinetic law as a function of the initial concentration of the olefin) allowed us to propose a simple kinetic scheme comprising a cis-trans interconversion of 2-butene and olefin-catalyzed destruction of the catalytic entity (k2 = (0.62 ± 0.06)x10?4 M s?1). In the continuous irradiation experiments the final distribution of the olefin was independent of the initial butene concentration (cis-2-butene/trans-2-butene 3.0) and the reaction kinetics are of first-order (kobs = (3±1) x10?4 s?1. Comparison of the two experiments suggests continuous photochemical regeneration of the catalytically active entity. The first-order reaction kinetics are in agreement with a carbene-metal carbonyl structure of the intermediate  相似文献   

10.
The dinuclear complex [(h5-1-CH3-3-C6H5C5H3)Fe(CO)2]2 was synthesized by reaction of Fe2(CO)9 with 1-methyl-3-phenylcyclopentadiene; it was converted to (h5-1-CH3-3-C6H5C5H3)Fe(CO)2CH3 by reduction with sodium amalgam and addition of CH3l, and thence to (h5-1-CH3-3-C6H5C5H3)Fe(CO)[P(C6H5)3] (COCH3) (I) by reaction with P(C6H5)3. The acetyl I was separated into two diastereomerically related pairs of enantiomers. Ia and Ib, by a combination of column chromatography on alumina and crystallization from benzene/pentane. The photochemical decarbonylation of Ia and Ib in benzene or THF solution was examined by 1H NMR spectroscopy. This reaction proceeds with high stereospecificity (>84% retention or inversion) at the iron center to yield (h5-1-CH3-3-C6H8C5H3)Fe(CO)[P(C6H5)3]CH3(II), enriched in the diastereomerically related pairs of enantiomers, IIa and IIb, respectively. Since IIa and IIb epimerize under the photolytic conditions of decarbonylation, the actual stereospecificity of the conversion of I to II is higher than 84%, and likely 100%. This is supported by the data from kinetic studies of the decarbonylation of I and the epimerization of II, carried out under identical photolytic conditions. The implications of the foregoing results to the mechanism of the decarbonylation are considered. Also described herein is the synthesis of other complexes with two asymmetric centers of the general formula (h5-cyclopentadienyl)Fe(CO)(L)(COR) and (h5-cyclopentadienyl)Fe(CO)(L)R that contain either an unsymmetrically substituted h5-cyclopentadienyl ring or a chiral tertiary phosphine.  相似文献   

11.
With a newly developed analytical technique, i.e. high temperature/pressure IR cell coupled to the reactor, it was possible to study the mechanism of hydroformylation at reaction conditions. It has been conclusively found that the hydrogenolysis of the acyl cobalt complex is performed by HCo(CO)4 and not by molecular H2, as proposed byHeck andBreslow.Therefore the formation of HCo(CO)4 from Co2(CO)8 is an intermediate step in the sequence of hydroformylation reaction steps. The rate of hydroformylation of any of the olefins is smaller than the rate of formation of HCo(CO)4 from Co2(CO)8. The IR spectra reveal that always more than 30% of the cobalt is in the form of HCo(CO)4 under the reaction conditions.It is found that the formation of HCo(CO)4 from Co2(CO)8 is the slowest and most temperature-dependent step of the hydroformylation reaction. Also the reaction between olefin and HCo(CO)4 is slower than the hydrogenolysis of the acyl complex.The experiments were carried out under industrial oxo conditions. The diffusional effects were eliminated.With 6 FiguresPart of the Ph.D. dissertation 1974. N. H. Alemdarolu, J. M. L. Penninger, andE. Oltay, Mechanism of Hydroformylation, Part II. Mh. Chem.107, 1043 (1976).  相似文献   

12.
Ionising radiations, employed in a broad range of dose-rate, together with a complex non-linear computation of reaction mechanisms, allow the determination of boundary values of rate constants concerning sorbitylfurfural (SF) reactivity towards a wide series of oxidant and/or virtually harmful radicals. SF reacts with some radicals (H, SO4 , CO3 , Br2 , CH3 ·), produced with both pulse and stationary radiolysis in neutral aqueous solution, having electrophilic and/or oxidative behaviour. The rate constants range from diffusional (k = (7–9 ) × 109 M-1 s-1) to relatively low values (k = 2 × 105 M-1 s-1). The possibility to observe these reactions, by means of radiolytical techniques, is heavily influenced by dose-rate. A relation between the radical E NHE 0 and their reactivity with SF is hinted.  相似文献   

13.
Unusually stable [(tC4H9O)3Ti-Co(CO)4] has been prepared by treating the appropriate carbonylmetallate anion with chloro titanium t-butoxide as well as by protolysis of CH3Ti(OtC4H9)3 with HCo(CO)4. Spectroscopic data indicate that the alkoxide and carbonyl ligands are nonbridging, establishing C3v-symmetry at the cobalt atom.  相似文献   

14.
Dicyclopentadienyllutetium monochloride interacts with LiAlH4in benzene (or toluene) and ether in the presence of a Lewis base to form coordination and electron saturated (18-electron configuration of the Lu atom) dimeric complexes (Cp2LuAlH4·L)2), where L = Et2O (I), NEt3 (IIa), C4H8O (III). Complexes IIa and III crystallize in monoclinic lattices with parameters: a = 11.35, b = 13.34, c = 14.20 Å, γ = 102°, space group P22/b for IIa; a = 8.73, b = 11.06, c = 16.42 Å, γ = 95.6°, space group P21/b for III. When a single crystal of IIa is exposed to hard X-rays (Mo-Kα, λ = 0.7106 nm) dissociation of the dimer takes places and a single crystal of monomeric Cp2Lu(μ2-H)AlH3·NEt3 (IIb) with a monodentate AlH4 group and 14-electron configuration of the Lu atom is formed. IIb crystallizes in monoclinic lattice with parameters: a = 13.278(4), b = 9.697(3), c = 14.099(4) Å, γ = 94.22°, space group P21/a, Z = 4, dcalc = 1.60 g/cm3 (R = 0.046, Rw = 0.047).  相似文献   

15.
Summary The chemistry of cobalt carbonyls in the presence of dienes and high pressure of synthesis gas was studied by online i.r. spectroscopy. Dicobalt octacarbonyl reacts with butadiene under 95 bar CO/H2 and 80°C to give [3-C4H7Co(CO)3] (1) and [4-C4H6)2Co2(CO)4] (2). Hydrogenation or hydroformylation are observed only with [HCo(CO)4] as the starting catalyst, and only at the beginning of the reaction. The results are explained by formation of an alkenyl complex, [-C4H7Co(CO)4], which either reacts with [HCo(CO)4] to give butene and [Co2(CO)8], or loses CO to give (1), depending on the [HCo(CO)4] concentration. The butene is hydroformylated. At temperatures >100°C (1) is transformed into a CO-free species, which catalyzes the oligomerisation of butadiene. Addition of tributylphosphine (L) leads to the formation of [3-C4H7Co(CO)2L] (5) and [Co2(CO)6L2] (6). In (5) the -allyl moiety is more labile than in (1) and a slow hydrogenation and hydroformylation of the butadiene is observed. In methanol solution the reaction of the cobalt carbonyls to give (1) is incomplete and the remaining H+ and [Co(CO)4] catalyze the hydroformylation of butadiene. Isoprene is less reactive than butadiene but otherwise behaves similarly.  相似文献   

16.
NO2 was photolyzed at 366 nm and 296 K in the presence of CH2O and O2 and in some runs with added NO or N2. The measured products were CO, CO2 and HCOOH. H2 and N2O were not produced. Both the CO and the CO2 were produced in a linear fashion with irradiation time, but the HCOOH grew after a marked induction period. From the CO2 quantum yields at high [O2]/[NO2] ratios an upper limiting value of 0.16 ± 0.02 was found for k3b/k3 where reactions (3a) and (3b) are
This is lower than the value of approximately 0.30 reported for k3b/k3 by Chang and Barker. From the CO and CO2 yields the competition for HCO between O2 and NO2 could be measured.
The ratio k9/k7a was found to be 0.21 ± 0.07 and was independent of pressure. The analysis required knowledge of some other rate coefficients. If those rate coefficients are completely in error, k9/k7a could be as high as 0.63. With the literature value of 5.6 × 10−12 cm3 s−1 for k9, the best value for k7a is (2.7 ± 0.9) × 10−11 cm3 s−1 with a lower limiting value of (8.9 ± 3.0) × 10−12 cm3 s−1.Information was also obtained on the reaction of HO2 with CH2O which produces HCOOH. An approximate value of (1.4 – 3.2) × 10−13 cm3 s−1 was found for the rate coefficient for this reaction which is about 14 – 32 times greater than the estimate of Su et al.  相似文献   

17.
Pulse radiolysis studies were carried out to determine the rate constants for reactions of ClO radicals in aqueous solution. These radicals were produced by the reaction of OH with hypochlorite ions in N2O saturated solutions. The rate constants for their reactions with several compounds were determined by following the build up of the product radical absorption and in several cases by competition kinetics. ClO was found to be a powerful oxidant which reacts very rapidly with phenoxide ions to form phenoxyl radicals and with dimethoxybenzenes to form the cation radicals (k = 7 × 108 −2 × 109 M-1 s-1). ClO also oxidizes ClO-2 and N-3 ions rapidly (9.4 × 108 and 2.5 × 108 M-1 s-1, respectively), but its reactions with formate and benzoate ions were too slow to measure. ClO does not oxidize carbonate but the CO-3 radical reacts with ClO- slowly (k = 5.1 × 105 M-1 s-1).  相似文献   

18.
Fe3(CO)92-H)(μ3-S-t-Bu) reacts with amines in aprotic solvents to give salts [Fe3(CO)93-S-t-Bu)][AminH]+ under deprotonation. The association of cluster and amine under formation of a solvated ion pair follows a second order rate law. The isotope effects kH/kD as well as the rate constants are strongly correlated with the steric demand of the individual bases used: The largest rate constants and the largest isotope effects (up to kH/kD = 13) are observed for bases with the least steric hindrance.  相似文献   

19.
NO2 was photolyzed with 2288 Å radiation at 300° and 423°K in the presence of H2O, CO, and in some cases excess He. The photolysis produces O(1D) atoms which react with H2O to give HO radicals or are deactivated by CO to O(3P) atoms The ratio k5/k3 is temperature dependent, being 0.33 at 300°K and 0.60 at 423°K. From these two points, the Arrhenius expression is estimated to be k5/k3 = 2.6 exp(?1200/RT) where R is in cal/mole – °K. The OH radical is either removed by NO2 or reacts with CO The ratio k2/kα is 0.019 at 300°K and 0.027 at 423°K, and the ratio k2/k0 is 1.65 × 10?5M at 300°K and 2.84 × 10?5M at 423°K, with H2O as the chaperone gas, where kα = k1 in the high-pressure limit and k0[M] = k1 in the low-pressure limit. When combined with the value of k2 = 4.2 × 108 exp(?1100/RT) M?1sec?1, kα = 6.3 × 109 exp (?340/RT)M?1sec?1 and k0 = 4.0 × 1012M?2sec?1, independent of temperature for H2O as the chaperone gas. He is about 1/8 as efficient as H2O.  相似文献   

20.
A simple method is described for studying the reactions of metastable Ar(3P0) and Ar(3P2) atoms separately in a discharge-flow system. CO and Kr quench these states with rate constants in the ratio k0 (CO)/k2 (CO) = 8±1 and k2 (Kr)/k0(Kr) = 18±2.  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