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.
Using the relative kinetic method, rate coefficients have been determined for the gas‐phase reactions of chlorine atoms with propane, n‐butane, and isobutane at total pressure of 100 Torr and the temperature range of 295–469 K. The Cl2 photolysis (λ = 420 nm) was used to generate Cl atoms in the presence of ethane as the reference compound. The experiments have been carried out using GC product analysis and the following rate constant expressions (in cm3 molecule?1 s?1) have been derived: (7.4 ± 0.2) × 10?11 exp [‐(70 ± 11)/ T], Cl + C3H8 → HCl + CH3CH2CH2; (5.1 ± 0.5) × 10?11 exp [(104 ± 32)/ T], Cl + C3H8 → HCl + CH3CHCH3; (7.3 ± 0.2) × 10?11 exp[?(68 ± 10)/ T], Cl + n‐C4H10 → HCl + CH3 CH2CH2CH2; (9.9 ± 2.2) × 10?11 exp[(106 ± 75)/ T], Cl + n‐C4H10 → HCl + CH3CH2CHCH3; (13.0 ± 1.8) × 10?11 exp[?(104 ± 50)/ T], Cl + i‐C4H10 → HCl + CH3CHCH3CH2; (2.9 ± 0.5) × 10?11 exp[(155 ± 58)/ T], Cl + i‐C4H10 → HCl + CH3CCH3CH3 (all error bars are ± 2σ precision). These studies provide a set of reaction rate constants allowing to determine the contribution of competing hydrogen abstractions from primary, secondary, or tertiary carbon atom in alkane molecule. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 651–658, 2002  相似文献   

2.
The photooxidations of n‐butyraldehyde initiated by Cl atom were carried out at room temperature (298 ± 2K) and 1 atm pressure. The rate coefficient for the reactions of Cl atom with n‐butyraldehyde was determined as k = (2.04 ± 0.36) × 10?10 cm3 molecule?1 s?1 by using relative rate techniques. The photooxidation products of n‐butyraldehyde reaction with Cl atom were also studied by using both gas chromatography‐mass spectrometry (GC‐MS) and gas chromatography techniques. C2H5CHO, CH3CHO, CO and CO2 were the major products observed. In the absence of NO, the observed yields of C2H5CHO, CH3CHO, and CO were 60%, 3%, and 9%, respectively. However, when NO was introduced into the reaction chamber and the initial ratios of [NO]0/[n‐butyraldehyde]0 were between 1 and 8, the yield of C2H5CHO decreased to 33%, whereas that of CH3CHO and CO rose up to 21% and 25%, respectively. On the basis of mechanism data deduced in this study and the fraction molar yields, the approximate branching ratios for Cl atom attack at ? C(O)H, α‐, β‐, and γ‐positions in n‐butyraldehyde could be derived as ?42%, <25%, 21%, and ?12%, respectively. © 2007 Wiley Periodicals, Inc. 39: 168–174, 2007  相似文献   

3.
Irradiation of solutions of n5-C5H5W(CO)3R (R  CH3n1-CH2C6H5) in cyclohexane at ca. 310490 nm leads to the formation of [n5-C5H5W(CO)3]2 and methane and of n5-C5H5W5(CO)2(n3-CH2C6H5) and some [n5-C5H5W(CO)3]2, respectively. When the irradiation is carried out in the presence of excess P(C6H5)3, the photoproducts are n5-C5H5W(CO)2[P(C6H5)3]CH3 (R  CH3) and n5-C5H5W(CO)2(n3-CH2C6H5) and trace [n5-C5H5W(CO)3]2 (R  n1-CH2C6H5). Photolysis of the n5-C5H5W(CO)3R in the presence of benzyl chloride affords n5-C5H5W(CO)3Cl (R  CH3) and both n5-C5H5W(CO)2(n3-CH2C2H5) and n5-C5H5W(CO)3Cl (R  n1-CH2C6H5), the relative amounts of the latter products depending on the quantity of added C6H5CH2Cl. Irradiation of n5-C5H5W(CO)3-CH3 in the presence of both P(C6h5)3 and C6H5CH2Cl affords n5-C5H5W(CO)2-[P(C6H5)3]CH3, but no n5-C5H5W(CO)3Cl. It is proposed that the primary photo-reaction in these transformations is dissociation of a CO group from n5-C5H5W-(CO)3R to generate n5-C5H5W(CO)2R, which can either combine with L to form a stable 18 electron complex, n5-C5H5W(CO)2(L)R (L  CO, P(C5H5)3; LR  n3-CH2C6H5), or lose the group R in a competing, apparently slower step. This proposal receives support from the observation that, light intensifies being equal, n5-C5H5W(CO)3CH3 undergoes a considerably faster photoconversion to [n5-C5H5W(CO)3]2 under argon than under carbon monoxide.  相似文献   

4.
The highly insoluble organic-inorganic hybrid ionic compounds N,N??-methylenedipyridinium tetrachloroplatinate(II) [(C5H5N)2CH2] · [PtCl4] and N,N??-methylenedipyridinium hexachloroplatinate(IV) [(C5H5N)2CH2] · [PtCl6] were obtained by the treatment of N,N??-methylenedipyridinium dichloride monohydrate [(C5H5N)2CH2]Cl2 · H2O with K2[PtCl4] or (NH4)2[PtCl6], respectively, in an aqueous solution. Both complexes were isolated, purified, characterised by elemental analysis, and their molecular structures were confirmed by powder X-ray diffraction. The crystal structure of both compounds consists of separated discrete dications [(C5H5N)2CH2]2+ and anions [PtCl n ]2? (n = 4 or 6). As anticipated, the dications formed a butterfly shape consisting of two pyridine rings bound to the methylene group via their N atoms, while the Pt centre had a square planar geometry in [(C5H5N)2CH2] · [PtCl4] and an octahedral coordination in [(C5H5N)2CH2] · [PtCl6]. Interestingly, both crystal structures are stabilised by intermolecular C-H??Cl non-standard hydrogen bonds, ??-?? ring interactions between two pyridine rings of adjacent dications, and also by Cl-?? interactions.  相似文献   

5.
Coordination Compounds of tert-Butyliminovanadium(V) Trichloride with O-Donor-Ligands The reaction of tert-butyliminovanadium(V)trichloride ( 1 ) with cyclic and acyclic ethers, ethylene carbonate and thietane has been studied. The 1:1-complexes have a different stability; reversible and irreversible cleavage of ether in the coordination sphere of the vanadium atom rearranging in ω-chloroalkanolato ligands are observed. The reaction of 1 with 2-chloroethanol, 3-chloropropanol and 5-chloropentanol yields the complexes tC4H9N = V(OR)Cl2 (R = CH2CH2CH2CH2CH2Cl) and [tC4H9N = V(OR)Cl2 · ROH]; in the presence of triethylamine the disubstituted compounds tC4H9N = V(OR)2Cl are formed. The 51V NMR spectra are discussed. The crystal structure of [tC4H9N = VCl3 · DME] ( 12 ) and [tC4H9N = V(OCH2CH2Cl)Cl2 · HOCH2CH2Cl] ( 13 ) has been determined. The vanadium atoms in 13 have a distorted octahedral coordination and are linked by the oxygen atoms of the 2-chloroethanolato ligands forming a binuclear complex. In solution molecular weight measurement and 51V NMR data indicate the equilibrium between a mononuclear complex 13 and its isomer [tC4H9N = V(OCH2CH2Cl)2Cl · HCl].  相似文献   

6.
The reaction of dicarbonyl- and carbonyl(trimethylphosphine)(cyclopentadienyl)-carbyne complexes of molybdenum and tungsten η5-C5H5(CO)2−n(PMe3)nMCR (n = 0, 1; M = Mo, W; R = CH3, C6H5, C6H4CH3, C3H5) with protic nucleophiles HX (X = Cl, CF3COO, CCl3COO) leads, through a combined protonation/carbon-carbon coupling reaction, to η2-acyl complexes η5-C5H5(CO)1−nX2(PMe3)n-M(η2-COCH2R). The reaction conditions, the results of the spectroscopic measurements and the X-ray structure of η5-C5H5(CO)(Cl2)W(η2-COCH2CH3) are reported.  相似文献   

7.
Hydrates of 3-phenylpropenal thiosemicarbazone (HL·H2O) and semicarbazone (HL′·H2O) react in methanol with cobalt, nickel, copper, and zinc chlorides, nitrates, and acetates to form coordination compounds MX2·2HL·nSolv [M = Co, Ni, Cu, Zn; X = Cl, NO3; HL = C6H5CH=CH-CH=N-NHC(O)NH2; n = 0–3; Solv = H2O, CH3OH], CuX2·HL·nH2O [M = Ni, Cu; n = 0, 1], ML2·nH2O and ML′·nH2O [M = Co, Ni, Zn; HL′ = C6H5CH=CH-CH=N-NHC(O)NH2; n = 0–3]. In the presence of amines (A = C5H5N, 2-CH3C5H4N, 3-CH3C5H4N, and 4-CH3C5H4N) these reactions yield the complexes Cu(A)LCl·CH3OH and M(A)LX·nH2O [M = Cu, Ni; X = Cl, NO3; n = 0–2]. The copper complexes with the amine ligands are of polynuclear structure, and other complexes are monomeric. Carbazones (HL and HL′) are included in the complexes as bidentate N,S-and N,O-ligands. The thermolysis of the complexes involves the stages of removing solvent crystallization molecules (70–90°C), deaquation (150–170°C), and full thermal decomposition (500–580°C).  相似文献   

8.
Electron density distribution in n-alkyl radicals (from ethyl to n-octyl) was studied by the B3LYP/6-311++G(3df,3pd) DFT method. The theory of atoms in molecules was used to show that the inductive effect of a free valence extends to two neighboring CH2 groups. The electronegativities χ(C?H2) > χ(CH3) > χ(CH2) of groups and χ(C?) > χ(H) > χ(C) atoms were qualitatively determined. The group method for calculating the enthalpies of formation of n-alkyl radicals Δf H°(n-C n H2n+ 1, n > 5) was substantiated.  相似文献   

9.
The (liquid + liquid) solubility curves have been determined by a synthetic method for six binary mixtures of [acetonitrile + {heptyl methyl ether CH3OnC7H15, or ethyl hexyl ether C2H5OnC6H13, or pentyl propyl ether nC3H7OnC5H11, or isopentyl propyl ether nC3H7Oi C5H11, or dibutyl ether nC4H9OnC4H9, or butyl isobutyl ether nC4H9OiC4H9}]. The possibility of the COSMO-SAC model to account for the thermodynamic differences between these systems has been tested and the discussion on the influence of screening charge of ethers on the system properties was undertaken.  相似文献   

10.
The differential thermal analyses of n-long chain alcohols CH3(CH3)n-1-OH, designated as Cn-OH (n = 16, 18, 20 and 22), and corresponding alkoxy ethanols CH3(CH2)n-1-OCH2CH2OH, designated as Cn-OC2H4OH, have been carried out at a heating rate of 1°C min?1. The differential e.m.f (ΔV) has been plotted against the temperature of the reference material (°C) and the onset of the peak has been taken as the appearance of the polymorphic phase transition or melting of the compound. Heats of transition (ΔH1) and melting or fusion (ΔHf) were computed from the areas under the respective peaks.  相似文献   

11.
Cyclisation of New Trimethylsilylalkylaminohalosilanes Compounds of the composition RSiCl2NR′SiMe3 (R = Cl, CH3, C2H5, C6H5; R′ = CH3, C2H5, C(CH3)3) are obtained by the reaction of silicon halides with the lithium salts of silylamines. Under suitable experimental conditions the reaction leads to the formation of the corresponding Si? N four- and six-membered ring systems (RSiHalNR′)n (Hal = F, Cl; n = 2 or 3) under elimination of trimethylhalosilane. The i.r., mass, 35Cl-n.q.r., 1H and 19F-n.m.r. spectra of these compounds are reported.  相似文献   

12.
Radiation yields of gases from n-paraffins of n-C20H42 to n-C24H50 and squalane (C30H62), as polymer model compounds, in the liquid and solid phases were analyzed by gas chromatography. G(H2) in the liquid phase was 3.2–3.3 for all samples and found to be almost independent of the chemical structure and molecular weight; G(H2) in the crystalline state of n-paraffins was 2.2–2.5 at -77 to 25°C. G(CH4) was about 1% of G(H2) for n-paraffins and increased with the methyl content of the branched chain for squalane. G(C2H6) in the liquid phase was about 0.05 for n-paraffins, but G(C2H6) in the crystalline state was found to depend on the crystal structure; that is, nearly zero for triclinic of an even number of carbons and about 0.02 for orthorhombic of an odd number. C3H8 and C2H4(C3H6 in squalane) were observed only in the liquid phase of n-paraffins and in glass and the liquid phase of squalane; G(C3H8) = 0.03–0.05 and G(C2H4 or C3H6) = 0.01–0.03. But the C4-compounds were not detected in any phase of any of the samples.Chain scission by radiation is supposed to proceed mainly at chain end carbons until the third carbon in the liquid phase of n-paraffins, only at the chain end carbon of the crystalline surface in triclinic crystals and at chain end carbons until the second carbon in orthorhombic crystals. These chain scission phenomena in the liquid phase and crystalline state of n-paraffins and in the liquid phase of squalane would be analogous to those in the amorphous and crystalline states of polyethylene, and in amorphous ethylene-propylene copolymer, respectively.  相似文献   

13.
Rate constants for H + Cl2, H + CH3CHO, H + C3H4, O + C3H6, O + CH3CHO, and Cl + CH4 have been measured at room temperature by the discharge flow—resonance fluorescence technique. The results are (1.6 ± 0.1) × 10?11, (9.8 ± 0.8) × 10 ?14, (6.3 ± 0.4) × 10?13) (2.00 torr He), (3.95 ± 0.41) × 10?12, (4.9 ± 0.5) × 10|su?13 and (1.08 ± 0.07) × 10?13, respectively, all in units of cm3 molecule?1 s?1. Also N atom reactions with C2H2, C2H4, C3H4, and C3H6 were studied but in no case was there an appreciable rate constant. These results are compared to previous studies.  相似文献   

14.
The dipole moments of the (C2H5)nPX3?n (X = Cl, Br, I; n = 0, 1, 2, 3) and P{EIVB(CH3)3}3 compounds (EIVB = C, Si, Sn) have been determined by dielectric measurements in benzene at 25°C by Higasi's method. The results are interpreted in terms of an additive vector model from bond moment calculations and a maximum phosphorus lone pair contribution of 1.75 D as calculated for P(C2H5)3. The high dipole moment of the mixed PA2B-type compounds in comparison with PA3 moments is mainly ascribed to a loss of C3v symmetry, characterized by a quite asymmetric orientation of the phosphorus lone pair with respect to the phosphorus bonding orbitals.  相似文献   

15.
A new method for the synthesis of the B3H 8 ? anion based on the reactions of alkyl and aryl halides (C2H4Cl2, C6H5CH2Cl, C4H9Br, (C6H5)3CCl, C10H7CH2Cl, CH2I2, C2HCl3) with sodium tetrahydroborate in diglyme was proposed. The method is characterized by high (up to quantitative) yields and easy isolation of the target products, (n-C4H9)4N)[B3H8] and Cs[B3H8], and also easy preparation of a salt of higher hydroborate anion, B12H 12 2? , without multistage purification from impurities of other anions of this series.  相似文献   

16.
Intensively coloured stibinidene complexes (LnM)2SbR (LnM = (CO)5Cr; R = tBu, Cl, I, EtS. LnM = C5H5(CO)2Mn; R = Cl. LnM = CH3C5H4(CO)2Mn; R = Br) which contain trigonally planar coordinated Sb(+1) with the stibanediyl ligand stabilized by M ? Sb(R) ? M-π-bonding have been obtained. Their synthesis and properties as well as an X-ray structure determination of [CH3C5H4(CO)2Mn]2SbBr are described.  相似文献   

17.
The compound Mo(η-C5H4(CH2)2SPrn)2(SPrn)2 acts as a bidentate ligand giving the heteronuclear bi-metallic compounds [Mo(η-C5H4CH2CH2SPrn)2-(SPrn)2(PtCl2)],[Mo(η-C5H4CH2CH2SPrn)2(SPrn)2(PdCl2)2], [Mo(η-C5C4CH2CH2SPrn)2(SPrn)2(RhCl3)2], [Mo(η-C5H4CH2CH2SPrn)2(μ-SPrn)2Rh(dppe)]BF4, [Mo(η-C5H4CH2CH2SPrn)2(μ-SPrn)2(COD)Rh]Cl, [Mo(η-C5H4CH2CH2SPrn)2-(μ-SPrn)2Pt(PPh3)2](PF6)2, and the compound [Mo(η-C5H4(CH2)2-μ-SPh)2Cl2Rh(COD)]Cl bonds via the ring-sulphur substituents giving [Mo(η-C5H4(CH2)2-μ-SPh)2-Cl2Rh(COD)]Cl.  相似文献   

18.
The first hypercoordinate sila[1]ferrocenophanes [fcSiMe(2‐C6H4CH2NMe2)] ( 5 a ) and [fcSi(CH2Cl)(2‐C6H4CH2NMe2)] ( 5 b ) (fc=(η5‐C5H4)Fe(η5‐C5H4)) were synthesized by low‐temperature (?78 °C) reactions of Li[2‐C6H4CH2NMe2] with the appropriate chlorinated sila[1]ferrocenophanes ([fcSiMeCl] ( 1 a ) and [fcSi(CH2Cl)Cl] ( 1 d ), respectively). Single‐crystal Xray diffraction studies revealed pseudo‐trigonal bipyramidal structures for both 5 a and 5 b , with one of the shortest reported Si???N distances for an sp3‐hybridized nitrogen atom interacting with a tetraorganosilane detected for 5 a (2.776(2) Å). Elongated Si? Cipso bonds trans to the donating NMe2 arms (1.919(2) and 1.909(2) Å for 5 a and 5 b , respectively) were observed relative to both the non‐trans bonds ( 5 a : 1.891(2); 5 b : 1.879(2) Å) and the Si? Cipso bonds of the non‐hypercoordinate analogues ([fcSiMePh] ( 1 b ): 1.879(4), 1.880(4) Å; [fcSi(CH2Cl)Ph] ( 1 e ): 1.881(2), 1.884(2)). Solution‐state fluxionality of 5 a and 5 b , suggestive of reversible coordination of the NMe2 group to silicon, was demonstrated by means of variable‐temperature NMR studies. The ΔG of the fluxional processes for 5 a and 5 b in CD2Cl2 were estimated to be 35.0 and 37.6 kJ mol?1, respectively (35.8 and 38.3 kJ mol?1 in [D8]toluene). The quaternization of 5 a and 5 b by MeOTf, to give [fcSiMe(2‐C6H4CH2NMe3)][OTf] ( 7 a‐ OTf) and [fcSi(CH2Cl)(2‐C6H4CH2NMe3)][OTf] ( 7 b‐ OTf), respectively, supported the reversibility of NMe2 coordination at the silicon center as the source of fluxionality for 5 a and 5 b . Surprisingly, low room‐temperature stability was detected for 5 b due to its tendency to intramolecularly cyclize and form the spirocyclic [fcSi(cyclo‐CH2NMe2CH2C6H4)]Cl ( 9 ‐Cl). This process was observed in both solution and the solid state, and isolation and Xray characterization of 9 ‐Cl was achieved. The model compound, [Fc2Si(2‐C6H4CH2NMe2)2] ( 8 ), synthesized through reaction of [Fc2SiCl2] with two equivalents of Li[2‐C6H4CH2NMe2] at ?78 °C, showed a lack of hypercoordination in both the solid state and in solution (down to ?80 °C). This suggests that either the reduced steric hindrance around Si or the unique electronics of the strained sila[1]ferrocenophanes is necessary for hypercoordination to occur.  相似文献   

19.
Solubility was studied for the first time in ternary aqueous phase-separating systems containing synthanol DS-10 or ALM-10 (polyethylene glycol monoalkyl ethers based on primary fatty alcohols, C n H2n ? 1O(C2H4O) m H, where m = 8–10 and n = 10–18 (synthanol DS-10) or 12–14 (synthanol ALM-10)) and inorganic salt (NH4)2SO4, Na2SO4, or Li2SO4 at 25°C. The boundaries of two-phase liquid equilibrium regions were determined. It was proposed to use the studied phase-separating systems for liquid extraction of metal ions.  相似文献   

20.
The photolysis of azomethane in the near UV has been studied at room temperature and pressures from 10 mtorr to 10 torr. The main products, C2H6 and N2, accounted for more than 99% of the reaction. Minor hydrocarbon products observed were (with quantum yields) C3H8 (3.5 × 10?3), C2H4 (3.2 × 10?4), CH4 (3 × 10?3), and n-C4H10 (trace). Quantum yields of H2 of 4 × 10?5 and 2 × 10?5 were measured at azomethane pressures of 0.1 and 1.0 torr, respectively. The minor hydrocarbon products can be accounted for by reactions of CH3 and C2H5 radicals following hydrogen abstraction from azomethane by CH3. The H2 product observed represents an upper limit for the H2 elimination from vibrationally excited C2H6 formed by CH3 combination in the system, corresponding to a rate of elimination ca. 5 × 10?5 times the competing rate of dissociation to 2CH3. Assuming a frequency factor of 1013 s?1 for the H2 elimination, a lower limit of about 90 kcal mol?1 was estimated for the energy barrier.  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