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.
A kinetic study of the reaction of [M(C10H12 · OCH3)(P)]+ complexes (M = Pd, Pt; PP = 1,2-bisdiphenylphosphinoethane; C10H12 = endo-dicyclopentadiene) with hydrogen halides, HX (X = Cl, Br) in aqueous methanol at 35° C is described. The proposed mechanism involves slow formation of the solvato species [M(C10H12)(solv.)(P)]2+ followed by fast reaction with X- to give M(PP)X2.  相似文献   

2.
Interaction of hydrazines N2H4X+ (X=H/Ph) with [Ru(HL)(OH2)] (1) (L=1,2-diaminopropanetetraacetate, PDTA) has been investigated by potentiometry, spectrophotometry and electrochemistry in aqueous solution at 25°C. The deprotonation and hydrolysis constants of 1 and its hydrazinium adducts formed in 0.1 M Na2SO4 solution were determined by potentiometry, while the second order rate constant k1 and k2 for the formation of [Ru(HL)(N2H4X)]+ and [Ru(L)(N2H4X)] were determined kinetically by spectrophotometry. At pH 2.8, the complex 1 exhibited a quasi-reversible one-electron reduction step at (E1/2) −0.251 V vs. SCE and the hydrazinium adducts obtained in situ in the presence of excess (100 equiv) N2H4X+ showed an additional multi-electron (two-electron per metal at a time) reduction step at (E1/2) −0.046/−0.158 V (vs. SCE) in the case of X=H/Ph, respectively in sampled-dc. The hydrazines, N2H4X+ were reduced electrolytically by holding the potential at −0.150 and −0.250 V (Hg-pool) vs. SCE, respectively, and chemically by using H2 at atmospheric pressure, ascorbic acid, catechol, Zn-dust and NaBH4 in the presence of these hydrazinium adducts. The turnover numbers, moles of ammonia formed per mole of metal per hour, have been calculated, discussed and compared with those of EDTA (ethylenediaminetetraacetate) analogue. The plausible reaction mechanism for the chemical and the electrochemical reduction of N2H4X+ to ammonia and/or aniline and the implication of these results on the possible function of nitrogenases have also been proposed. The complex [Ru(HL)(N2H5)]HSO4 and [Ru(HL)(N2H4Ph)]Cl were synthesized and characterized.  相似文献   

3.
The major metal-containing species formed upon fast atom bombardment of amino acid/Ni+2 mixtures is the [M + Ni]+ adduct, involving reduction of the Ni+2 to the +1 oxidation state. By contrast, electrospray ionization of amino acid/Ni+2 mixtures produces predominantly [Ni(M ? H)M]+; this species, on collisional activation, produces predominantly [M + Ni]+ by elimination of [M - H], presumably a carboxylate radical. The unimolecular fragmentation reactions occurring on the metastable ion time scale for the [M + Ni]+ adducts of a variety of α-amino acids have been recorded. The adducts with phenylalanine, α-aminoisobutyric acid and α-aminobutyric acid fragment by elimination of H2O, H2O + CO and, to a minor extent, by elimination of CO2. These reactions are similar to those observed for the [M + Cu]+ adducts of α-amino acids. A reaction distinctive for the [M + Ni]+ adducts involves formation of the immonium ion RCH=NH 2 + . By contrast, the [M + Ni]+ adducts with leucine, isoleucine, and norleucine show extensive metastable ion fragmentation by elimination of H2, CH4, C2H4, C3H6, and C4H8, with the relative importance of the different fragmentation channels depending on the configuration of the C4H9 side chain. These results are interpreted in terms of C-C and C-H bond activation of the C4H9 side chain by the Ni+. The adducts with valine and norvaline fragment in a fashion similar to the adduct with phenylalanine, except that minor elimination of C3H6 is observed.  相似文献   

4.
The heat capacities of Ln(Me2dtc)3(C12H8N2) (Ln = La, Pr, Nd, Sm, Me2dtc = dimethyldithiocarbamate) have been measured by the adiabatic method within the temperature range 78–404 K. The temperature dependencies of the heat capacities, C p,m [La(Me2dtc)3(C12H8N2)] = 542.097 + 229.576 X ? 27.169 X 2 + 14.596 X 3 ? 7.135 X 4 (J K?1 mol?1), C p,m [Pr(Me2dtc)3(C12H8N2)] = 500.252 + 314.114 X ? 17.596 X 2 ? 0.131 X 3 + 16.627 X 4 (J K?1 mol?1), C p,m [Nd(Me2dtc)3(C12H8N2)] = 543.586 + 213.876 X ? 68.040 X 2 + 1.173 X 3 + 2.563 X 4 (J K?1 mol?1) and C p,m [Sm(Me2dtc)3(C12H8N2)] = 528.650 + 216.408 X ? 16.492 X 2 + 12.076 X 3 + 4.912 X 4 (J K?1 mol?1), were derived by the least-squares method from the experimental data. The heat capacities of Ce(Me2dtc)3(C12H8N2) and Pm(Me2dtc)3(C12H8N2) at 298.15 K were evaluated to be 617.99 and 610.09 J K?1 mol?1, respectively. Furthermore, the thermodynamic functions (entropy, enthalpy and Gibbs free energy) have been calculated using the obtained experimental heat capacity data.  相似文献   

5.
In strychninium 4‐chloro­benzoate, C21H23N2O2+·C7H4ClO2, (I), and strychninium 4‐nitro­benzoate, C21H23N2O2+·C7H4NO4, (II), the strychninium cations form pillars stabilized by C—H⋯O and C—H⋯π hydrogen bonds. Channels between the pillars are occupied by anions linked to one another by C—H⋯π hydrogen bonds. The cations and anions are linked by ionic N—H+⋯O and C—H⋯X hydrogen bonds, where X = O, π and Cl in (I), and O and π in (II).  相似文献   

6.
The two title proton‐transfer compounds, 5‐methylimidazolium 3‐carboxy‐4‐hydroxybenzenesulfonate, C4H7N2+·C7H5O6S, (I), and bis(5‐methylimidazolium) 3‐carboxylato‐4‐hydroxybenzenesulfonate, 2C4H7N2+·C7H5O6S2−, (II), are each organized into a three‐dimensional network by a combination of X—H...O (X = O, N or C) hydrogen bonds, and π–π and C—H...π interactions.  相似文献   

7.
The distonic ions HO+?CHCH2C˙H2 (1) and CH3C(?O+H)CH2C˙H2 (2) were directly generated, their decompositions characterized and their appearance energies determined by photoionization. Heats of formation derived from the appearance energies were 757 kJ mol?1 for 1 and 692 kJ mol?1 for 2. Deuterium labeling demonstrates that both ions decompose at low energies in the same ways as their isomers with the same skeletal structures, consistent with proposals that 1 and 2 are intermediates in the decompositions of those systems. Surprisingly, the values of the translational energy releases accompanying the formation of CH3CO+ and C2H5CO+ from 2 appear to be inversely proportional to the available excess energy. The 1,2-H-shift RC(?O+H)CH2C˙H2 → RC(?O+H)C˙HCH3 is compared to the corresponding, non-occurring 1,2-H-shift in alkyl free radicals.  相似文献   

8.
The determination of gas-phase reactivity of a series of polycyclic aromatic hydrocarbons (PAHs) with nucleophiles is directed at achieving isomer differentiation through ion-molecule reactions and collisionally activated decomposition spectra. A series of PAH isomers form gas-phase [adduci — H]+ ions with the reagent nucleophiles pyridine and N-methylimidazole. Collisionally activated decomposition spectra of the [adduct — H]+ ions of the pyridine/PAH systems are dominated by products formed by losses of C5H4N, C5H5N (presumably neutral pyridine), and C5H6N. Collisional activation of PAH/N-methylimidazole [adduct — H]+ ions causes analogous losses of C4H5N2, C4H6N2 (presumably neutral N-methylimidazole), and C4H7N2. The relative abundances of the ions that result from these losses are highly isomer specific for N-methylimidazole but less so for pyridine. Furthermore, PAH/N-methylimidazole [adduct — H]+ ions undergo a series of metastableion decompositions that also provide highly isomer-specific information. The C4H7N2 (from PAH/N-methylimidazole product ions) and C5H6N (from PAH/pyridine product ions) losses tend to increase with the ΔH f of the PAH radical cation. In addition, it is shown that the fragmentation patterns of these gas-phase PAH/nucleophile adducts are similar to fragmentation patterns of PAH/nucleoside adducts generated in solution, which suggests that the structures of products formed in gas-phase reactions are similar to those produced in solution.  相似文献   

9.
Fast atom bombardment (FAB) mass spectrometry of the gold(I) and gold(III) derivatives, {Au[C(Y)–NHAr]2}+X? and {Au[C(Y)–NHAr]2I2} + X? (Y =? OC2H5 or ? NHAr; X? = CIO4? or BF4?; Ar = p-CH3? C6H4) has led to the detection, for the alkoxyamino derivatives only, of [M–H]+˙ molecular species. The mechanism of the formation of these unusual species has been studied with respect to the oxidation state of gold, nature of the matrix, matrix acidity and ligand structure. The energetics of two possible alternative mechanisms has been studied by means of ab initio theoretical calculations. Both experimental and theoretical data indicate that [M–H]+˙ formation is due to the reaction of M+ with H+-philic and/or H˙-philic species produced from the matrix by FAB. Whatever the operative mechanism, the [M–H]+˙ formation is to be considered a FAB-induced oxidative process.  相似文献   

10.
In the ion/molecule reactions of the cyclometalated platinum complexes [Pt(L? H)]+ (L=2,2′‐bipyridine (bipy), 2‐phenylpyridine (phpy), and 7,8‐benzoquinoline (bq)) with linear and branched alkanes CnH2n+2 (n=2–4), the main reaction channels correspond to the eliminations of dihydrogen and the respective alkenes in varying ratios. For all three couples [Pt(L? H)]+/C2H6, loss of C2H4 dominates clearly over H2 elimination; however, the mechanisms significantly differs for the reactions of the “rollover”‐cyclometalated bipy complex and the classically cyclometalated phpy and bq complexes. While double hydrogen‐atom transfer from C2H6 to [Pt(bipy? H)]+, followed by ring rotation, gives rise to the formation of [Pt(H)(bipy)]+, for the phpy and bq complexes [Pt(L? H)]+, the cyclometalated motif is conserved; rather, according to DFT calculations, formation of [Pt(L? H)(H2)]+ as the ionic product accounts for C2H4 liberation. In the latter process, [Pt(L? H)(H2)(C2H4)]+ (that carries H2 trans to the nitrogen atom of the heterocyclic ligand) serves, according to DFT calculation, as a precursor from which, due to the electronic peculiarities of the cyclometalated ligand, C2H4 rather than H2 is ejected. For both product‐ion types, [Pt(H)(bipy)]+ and [Pt(L? H)(H2)]+ (L=phpy, bq), H2 loss to close a catalytic dehydrogenation cycle is feasible. In the reactions of [Pt(bipy? H)]+ with the higher alkanes CnH2n+2 (n=3, 4), H2 elimination dominates over alkene formation; most probably, this observation is a consequence of the generation of allyl complexes, such as [Pt(C3H5)(bipy)]+. In the reactions of [Pt(L? H)]+ (L=phpy, bq) with propane and n‐butane, the losses of the alkenes and dihydrogen are of comparable intensities. While in the reactions of “rollover”‐cyclometalated [Pt(bipy? H)]+ with CnH2n+2 (n=2–4) less than 15 % of the generated product ions are formed by C? C bond‐cleavage processes, this value is about 60 % for the reaction with neo‐pentane. The result that C? C bond cleavage gains in importance for this substrate is a consequence of the fact that 1,2‐elimination of two hydrogen atoms is no option; this observation may suggest that in the reactions with the smaller alkanes, 1,1‐ and 1,3‐elimination pathways are only of minor importance.  相似文献   

11.
A theoretical study of the halogenated cations of mono-, di-, tri- and tetramethyl-substituted ethylenes, C3H6X+, C4H8X+, C5H10X+ and C6H12X+, X=F, Cl, Br, have been studied at the ab initio MP2 and density functional B3LYP levels of theory implementing 6-311++G(d,p) basis set. The potential energy surfaces of all molecules under investigation have been scanned and the 13C and 1H NMR chemical shifts for all the bridged halonium ions studied have been calculated using the GIAO method at the B3LYP level. The calculated halogen binding energies in the halonium ions have been correlated with the experimental rates of chlorination and bromination of the corresponding alkenes. The computed hydride affinities and the NICS values for the bridged cations show that the bromo cations are more stable than the analogous chloro and fluoro cations.  相似文献   

12.
《Polyhedron》1987,6(10):1885-1899
The synthesis and characterization, chiefly as salts of the anions [X(ONO2)2] (X = H+ or Ag+) (by analysis, X-ray powder photography, vibrational spectra and thermogravimetry) of adducts of the nitrates trans-[M(L)4X2](NO3) (M =Rh or Ir; L = pyridine, perdeuteriopyridine or 4-methylpyridine; X = Cl or Br) with hydrogen nitrate and silver nitrate are described.  相似文献   

13.
The formation of diethyl halonium ions (C2H5)2X+ (X = Cl, Br, I) by a variety of ion-molecule reactions is described. The dissociation characteristics (metastable and collision-induced dissociation mass spectra) of these ions and their isomers were studied in detail. Some of the neutral fragmentation products were examined by their collision-induced dissociative ionization mass spectra. The participation of classical (1, CH3CH2X+CH2CH3) and nonclassical forms of the ions was considered. Dissociation reactions for which loss of positional identity of H-D atoms took place, for example C2H4 loss (a common fragmentation of metastable ions) and C2H5 + formation, were interpreted as involving nonclassical ions, 2. It was concluded that the ion-molecule reactions produced both ion structures, but in different halogen-dependent proportions. For (C2H5)2C1+ ions, 2 is the major species, for (C2H5)2Br+ both 1- and 2-type ions are generated, whereas for (C2H5)2I+ the classical form 1 must be the predominant structure.  相似文献   

14.
The unimolecular dissociations of C5 epoxides ions mono- or disubstituted at C1 give exclusive loss of CH3 and exclusive formation of methoxy vinyl carbenium ion, both in the source and in the 2nd field-free region. In the case of the 1,2-disubstituted ion in the 2nd field-free region the loss of ethene is the only pathway, while a competition occurs for the trisubstituted ion leading to [C3H6O]+˙ and [C4H7O]+˙ ions, the structure of which are demonstrated. The first step of the different mechanisms is the cleavage of the heterocyclic C? C bond.  相似文献   

15.
The metastable molecular ion of 2-hexanone loses a methyl radical mainly (~80%) from positions C(4) and C(6), in equal proportions, as indicated by 13C labelling. The necessary skeletal rearrangement of the butyl chain is interpreted in terms of a 1,2-[enol-olefin] +˙ shift. The results and the mechanisms concerning the minor eliminations of C2H4, C2H5˙, C3H5˙ and C3H6 neutrals are also discussed.  相似文献   

16.
The dihalomethanes CH2X2 (X=Cl, Br, I) were co‐crystallized with the isocyanide complexes trans‐[MXM2(CNC6H4‐4‐XC)2] (M=Pd, Pt; XM=Br, I; XC=F, Cl, Br) to give an extended series comprising 15 X‐ray structures of isostructural adducts featuring 1D metal‐involving hexagon‐like arrays. In these structures, CH2X2 behave as bent bifunctional XB/XB‐donating building blocks, whereas trans‐[MXM2(CNC6H4‐4‐XC)2] act as a linear XB/XB acceptors. Results of DFT calculations indicate that all XCH2–X???XM–M contacts are typical noncovalent interactions with estimated strengths in the range of 1.3–3.2 kcal mol?1. A CCDC search reveals that hexagon‐like arrays are rather common but previously overlooked structural motives for adducts of trans‐bis(halide) complexes and halomethanes.  相似文献   

17.
The formation of the compound RSnX(acac)2 (acac = 2,4-pentanedionato) by reaction of bis(2,4-pentanedionato)tin(II) on a halide RX with R = CH3, C2H5, C4H9, C6H5, CH2I, (C6H5)3SnCH2, (C2H5)3SnCH2 and X = I, Br has been studied by polarography. At 25°C, it is in fact an equilibrium whose constant has been measured. The intermediate formation of the ion-pair [RSn(acac)2+X?] has allowed us to explain the experimental results.  相似文献   

18.
Fragments of the potential energy surfaces (PES) of the systems [C3H8 + CBr3 +] and [C3H8 + Br2CBr+·Br2AlBr2 ] were simulated by the MNDO/PM3 method. Energy minima corresponding to weakly bound adducts of propane molecule with the CBr3 + cation or neutral complex CBr3 +·AlBr4 were found on the PES's of both systems. These are adducts with the coordination of a H atom of the methylene group of the propane molecule to the electrophile at the Br atom carrying the largest positive charge. As the fragments of the adducts are brought close together, the coordinated H atom migrates to the C atom of the CBr3 + fragment. The potential barriers of these migrations were found to be low for both systems. The reactions proceed without formation of cyclic intermediates or transition states typical of the Olah mechanism.  相似文献   

19.
The reaction of N-nitro-O-(4-nitrophenyl)hydroxylamine (1) with conc. H2SO4 affords 4-nitropyrocatechol and that with conc. sulfonic acids (RSO3H where R = Me, CF3) affords 2-hydroxy-5-nitrophenyl-R-sulfonates in yields of 80?C85%. These reactions are assumed to proceed through an intermediate (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+, which eliminates the N2O molecule to form the aryloxenium ion [NO2C6H4O]+. The latter reacts with acid anions at the ortho-carbon atom of the phenyl ring. The thermodynamical parameters of the elementary reactions resulting in the formation of the (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+ and aryloxenium ion [NO2C6H4O]+ were calculated in the B3LYP/6?311+G(d) study of the combined molecular system (nitrohydroxylamine 1 + [H3SO4]+). The reaction of nitrohydroxylamine 1 with aqueous solutions of strong acids (??70% H2SO4, CF3SO3H) affords mainly 4-nitrophenol. It appears that the mechanism of this reaction does not involve the formation of the aryloxenium ion.  相似文献   

20.
Phosphonium adduct formation via attack of tri-n-butylphosphine on the cations [(C7H7)M(CO)3]+ (M = Cr, Mo, W) obeys the rate law, Rate = k [complex] [PBu3]. The very similar rate constants for the Cr, Mo and W complexes confirm the similar electrophilicities of the tropylium rings in these cations, and also support the view that there is direct addition to the rings. The related complexes [(C6H7)Fe(CO)3]BF4 and [(C6H6)Mn(CO)3]BF4 also form adducts with PBu3, and the quantitative reactivity order [(C6H7)Fe(CO)3]+ > [(C7H7)Cr(CO)3]+ » [(C6H6)Mn(CO)3]+ (160:60:1) has been established.  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