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 15 毫秒
1.
We report the results of a selected ion flow tube (SIFT) study of the reactions of H3O+, NO+ and O+2 with some nine carboxylic acids and eight esters. We assume that all the exothermic proton transfer reactions of H3O+ with all the acid and esters molecules occur at the collisional rate, i.e. the rate coefficients, k, are equal to kc; then it is seen that k values for most of the NO+ and O+2 reactions also are equal to or close to kc. The major ionic products of the H3O+ reactions with both the acids and esters are the protonated parent molecules, MH+, but minor channels are also evident, these being the result of H2O elimination from the excited (MH+)1 in some of the acid reactions and an alcohol molecule elimination (CH3OH or C2H5OH) in some of the ester reactions. The NO+ reactions with the acids and esters result in both ion-molecule association producing NO+M in parallel with hydroxide ion (OH) transfer with some of the acids, and parallel methoxide ion (CH3O) and ethoxide ion (C2H5O) transfer as appropriate with some of the esters. The O+2 reactions proceed by dissociative charge transfer with the production of two or more ionic fragments of the parent molecules, the different isomeric forms of both the acid and the ester molecules resulting in different product ions.  相似文献   

2.
Abstract

The reactions of Mo(CO)6 and W(CO)6 with HCl(g) in the presence of 12-crown-4 and H2O have been investigated in toluene. For both reactions, two products were isolated, depending on the oxidation of the metal center. For molybdenum, the MoIII species, [H3O+ · 12-crown-4]3[Mo2Cl9 3-], 1, was obtained from the liquid clathrate layer in the reaction mixture. Upon air oxidation of the reaction mixture, the Mov complex, [H7O3 ? · H4O2 + · (12-crown-4)2][MoOCl4(H2O)?]2, 2, rapidly formed. For tungsten, the WII species, [(H5O2 +)2 · 12-crown-4][W(CO)4Cl? 3]2, 3, deposited from the liquid clathrate layer which upon oxidation formed the Wv complex, [H3O+· 12-crown-4][WOCl4(H2O)?], 4. These reactions were promoted by UV radiation and formed liquid clathrates almost immediately upon reaction. X-ray crystal structures were performed on each compound. Complexes 1 and 4 have H3O+ oxonium ions involved in complex hydrogen bonded arrays with the 12-crown-4 acceptor molecules. The H5O2 + oxonium ions in 2 and 3 contain extremely short O…O separations, equivalent to the shortest O-H…O bonds known. Also isolated in complex 2 was the H7O3 + oxonium ion which contains an unusual linear O…O…O core.  相似文献   

3.
The reactions of the cyclic molecules C6H6 (benzene), c-C3H6 (cyclopropane) and c-C6H12 (cyclohexane) with ArH+ (ArD+), H3+, N2H+, CH5+, HCO+, OCSH+, C2H3+, CS2H+ and H3O+ have been studied at 300 K using a SIFT apparatus. All the reactions except those of C2H3+ proceed via proton transfer and all are fast except the H3O+ and CS2H+ reactions with c-C6H12 which are endothermic and which establish that the proton affinity of c-C6H12 is 160 ± 1 kcal mol−1, which is considerably lower than the published value. In the c-C3H6 and the c-C6H12 reactions multiple products are observed and hence “breakdown curves” for the protonated molecules are constructed and the appearance energies of the various ion products are consistent with available thermochemical data. The reactions of C2H3+ with these cyclic molecules are atypical within this series of reactions in that they appear to proceed largely via hydride ion transfer. The implications of the results of this study to interstellar chemistry are alluded to.  相似文献   

4.
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   

5.
The iodine–sulfur (IS) thermochemical process for hydrogen production is one of the most promising approaches in using high‐temperature process heat supplied by a nuclear reactor. This process includes three reactions that form a closed cycle: the Bunsen reaction, in which iodine, water, and sulfur dioxide react to form sulfuric acid and hydriodic acid (HI); HI decomposition; and sulfuric acid decomposition. However, the side reactions between H2SO4 and HI may disturb the operation of the IS closed cycle. For optimal process conditions, the reaction kinetics between H2SO4 and HI should be examined. In this work, a preliminary kinetic study was conducted. Using the initial reaction rate method, the kinetic parameters of the reaction between sulfuric acid and HI, such as the apparent reaction orders and rate constant were determined. For I?, the apparent reaction order was approximately 1.77, whereas the orders for H+ and SO42? were 7.78 and 1.29, respectively. The apparent rate constant at 85 ± 1°C was approximately 2.949 × 10?11 min?1 (mol/L)?9.84. The H+ concentration had more significant influence on the reaction rate than those of SO42? and I?. Such basic data provide useful information for related process design and further kinetics study.  相似文献   

6.
Thermal gas-phase reactions of the ruthenium-oxide clusters [RuOx]+ (x=1–3) with methane and dihydrogen have been explored by using FT-ICR mass spectrometry complemented by high-level quantum chemical calculations. For methane activation, as compared to the previously studied [RuO]+/CH4 couple, the higher oxidized Ru systems give rise to completely different product distributions. [RuO2]+ brings about the generations of [Ru,O,C,H2]+/H2O, [Ru,O,C]+/H2/H2O, and [Ru,O,H2]+/CH2O, whereas [RuO3]+ exhibits a higher selectivity and efficiency in producing formaldehyde and syngas (CO+H2). Regarding the reactions with H2, as compared to CH4, both [RuO]+ and [RuO2]+ react similarly inefficiently with oxygen-atom transfer being the main reaction channel; in contrast, [RuO3]+ is inert toward dihydrogen. Theoretical analysis reveals that the reduction of the metal center drives the overall oxidation of methane, whereas the back-bonding orbital interactions between the cluster ions and dihydrogen control the H−H bond activation. Furthermore, the reactivity patterns of [RuOx]+ (x=1–3) with CH4 and H2 have been compared with the previously reported results of Group 8 analogues [OsOx]+/CH4/H2 (x=1–3) and the [FeO]+/H2 system. The electronic origins for their distinctly different reaction behaviors have been addressed.  相似文献   

7.
We present here a study of the collision induced dissociation (CID) of deprotonated cysteic acid containing peptides produced by MALDI. The effect of cysteic acid (Cox) position is interrogated by considering the positional isomers, CoxLVINVLSQG, LVINVLSQGCox, and LVINVCoxLSQG. Although considerable variation between the CID spectra is observed, the mechanistic picture that emerges involves charge retention at the deprotonated cysteic acid side chain. Fragmentation occurs in the proximity of the cysteic acid group by charge directed mechanisms as well as remote from this group to form ions, which may be rationalized by charge remote mechanisms. Additionally, the formation of the SO3–• ion is observed in all cases. Fragmentation of CoxLVINVLSQCox provides both N- and C-terminal, y and b ions, respectively indicating that the negative charge may be retained at either of the cysteic acids; however, there is some evidence that charge retention at the C-terminal cysteic acid may be preferred. Fragmentation of tryptic type peptides containing a C-terminal arginine or lysine residue is considered through comparison of three peptides CoxLVINKLSQG, CoxLVINVLSQK, and CoxLVINVLSQR. Lastly, we rationalize the formation of b n–1 + H2O and a n–1 ions through a mechanism involving rearrangement of the C-terminal residue to form a mixed anhydride intermediate.  相似文献   

8.
Four new low-dimensional phenylarsonates, A(HO3AsC6H5)(H2O3AsC6H5) (A = Tl(1), Na(2), K(3) and Rb(4)), have been synthesized and characterized by X-ray diffraction, spectroscopic and thermal studies. They crystallize in triclinic unit cells and have approximately planar arrangement of A+ ions, coordinated to oxygen atoms of phenylarsonates, on both sides. Structure of thallium phenylarsonate as determined by single crystal X-ray diffraction, is one-dimensional, whereas those of alkalimetal analogues are two-dimensional. Successful intercalation reactions of compounds1 and2 with primaryn-alkyl amines have been demonstrated. Dedicated to Prof J Gopalakrishnan on his 62nd birthday.  相似文献   

9.
H4BOPTC reacts with IMI to yield a supramolecular organic framework, formulated as [IMIH+]2√[H2BOPTC2 ? ]√0.5H2O (1) (IMI = imidazole, H4BOPTC = benzophenone-3,3′,4,4′-tetracarboxylic acid). Single-crystal X-ray diffraction analysis reveals that 1 shows a novel architecture with two-level hierarchical entanglement. H2BOPTC2 ?  connects to IMIH+ through hydrogen bonds, providing 1D ribbon. The basic ribbons are entangled into a 3D net with ant topology. Then the ant nets further interpenetrate to give the final entangled framework. The thermal stability, optical band gap energy and photoluminescent property of 1 have also been investigated.  相似文献   

10.
The reactions of ?-C3H3+ (propargylium cation) with acetylene and diacetylene have been modeled kinetically. Data were obtained from Fourier Transform Ion Cyclotron Resonance (FTICR) experiments on these systems, which are themselves models for soot particle initiation. Acetylene forms an encounter complex with ?-C3H3+, but, in the absence of a third body collision, the complex decomposes to acetylene and c-C3H3+ (cyclopropenylium cation) at about 1/3 the rate it decomposes to acetylene and ?-C3H3+, in spite of the fact that c-C3H3+ is ca. 115 kJ/mol more stable than ?-C3H3+. The encounter complex is long enough lived, and energetic enough, to scramble deuterium in reactions between ?-C3H3+ and C2D2. These reactions have been successfully modeled, yielding a nearly statistical distribution of deuterium, and a rather large kinetic isotope effect. The more complex reactions of ?-C3H3+ with diacetylene have also been modeled.  相似文献   

11.
Abstract

The crystal structures of two hexaaza macrocycles 1,4,7,12,15,18-hexaazacyclodocosane ([22]N6:1-6H+, 6C1?) and 1,13-dioxa-4,7,10,16,19,22-hexaazacyclotetraeicosane ([24]N6O2:2-6H+,6CI?) as their hexa-hydrochloride salts have been determined. 1-6H+ binds specifically two CI? anions above and below the almost planar hexaammonium macrocycle yielding a dinuclear CI? complex. The hexacation 2-6H+ on the other hand interacts preferentially with three CI? anions of the six present in the solid state. Among the three closest anions, one of them, interacting with four ammonium groups, is located in the centre of the macrocycle which adopts a “pocket-like” conformation. Potentiometric and 35CI NMR experiments demonstrate that 2-6H+ also binds CI? in aqueous solution. Subsequent extensive molecular dynamics computational studies starting from X-ray coordinates show that the solid state structure is representative of the solution conformations for I-6H+, whereas the conformations of 2-6H+ are strongly affected by intramolecular interactions between the ammonium centres and O-atoms of the ether linkage as well as by intermolecular interactions with H2O molecules and CI? counterions.  相似文献   

12.
The probable fragmentation channels of hydroxymethyl radical cation were studied through the H‐and H2‐abstraction and C‐O bond breaking reactions including their related isomerization reactions. The energy barriers for hydroxymethyl cation undergoing isomerization reactions are generally higher than those undergoing the concerted 1,2‐elimination reactions to generate CHO+ and H2. The fragmentation reaction to form CHO+ and H2 through the 1,2‐elimination pathways is the major fragmentation channel for hydroxymethyl cation, consistent with the experimental observation. H abstraction from the hydroxyl group of CH2OH+ is more difficult than that from the methylene group. The feasible path to lose H is to generate CHOH2+ through hydrogen transfer reaction as the first step and then to undergo H‐elimination to generate trans‐CHOH+. Among all the reactions found in this study, the OH‐elimination to generate CH2+ has the highest energy barrier. Our calculation results indicate that the major signals contributed from the related species of hydroxymethyl cation found in the mass spectrum should be m/e 29, m/e 30.  相似文献   

13.
Density functional theory was used to study gas-phase reactions between the Cp2*ZrMe+ cations, where Cp* = C5H5 (1), Me5Cp = C5Me5 (2), and Flu = C13H9 (3), and the ethylene molecule, Cp2*ZrMe+ + C2H4 → Cp2*ZrPr+ → Cp2*ZrAllyl+ + H2. The reactivity of the Cp2*ZrMe+ cations with respect to the ethylene molecule decreased in the series 1 > 32. Substitution in the Cp ring decreased the reactivity of the Cp2*ZrMe+ cations toward ethylene, in agreement with the experimental data on the comparative reactivities of complexes 1 and 3. The two main energy barriers along the reaction path (the formation of the C-C bond leading to the primary product Cp2*ZrPr+ and hydride shift leading to the secondary product Cp2*Zr(H2)Allyl+) vary in opposite directions in the series of the compounds studied. For Flu (3), these barriers are close to each other, and for the other compounds, the formation of the C-C bond requires the overcoming of a higher energy barrier. A comparison of the results obtained with the data on the activity of zirconocene catalysts in real catalytic systems for the polymerization of ethylene led us to conclude that the properties of the catalytic center changed drastically in the passage from the model reaction in the gas phase to real catalytic systems.  相似文献   

14.
The complexation reactions of iron(III) with 2-pyridine carboxylic acia (picolinic acid) and 2,6-pyridine dicarboxylic acid (dipicolinic acid) in aqueous solutions have been studied by spectrophotometric and stopped flow techniques. Equilibrium constants were determined for the 1 : 1 complexes at temperatures between 25 and 80°C. The values obtained are: Picolinic Acid (HL): Fe3++ H2L+? FeHL3++H+(K1 = 2.8,ΔH = 2 kcal mole?1 at 25°C, μ = 2.67 M) Dipicolinic Acid (H2D): Fe3++H2D? FeD++2H+(K1K1A= 227 M, ΔH = 3.4 kcal mole?1 at 25°C,μ = 1.0 M). The rate constants for the formation of these complexes are also given. The results are used to evaluate the effects of these two acids upon the rate of dissolution of iron(III) from its oxides.  相似文献   

15.
Abstract

A sensitive specific assay for taurine using high performance liquid chromatograpy and fluorescence measurement is described. The method employs precolumn derivatization with o-phthalaldehyde in the presence of ethanethiol. Taurine is clearly separated from other amino acids including its precursors hypotaurine and cysteine sulfinic acid. The fluorescence peak height is linear between 1 and 100 picomoles of taurine. There is clear separation of taurine from a contaminant of other taurine assays, α-glycerophosphoryl ethanolamine.  相似文献   

16.
The kinetics and mechanism of reduction of aqueous toluidine blue (TB+) by phenyl hydrazine (Pz), which exhibits nonlinear behavior, is studied spectrophotometrically at 630 nm. Typical kinetic curves exhibited autocatalytic characteristics. The role of H+ as an autocatalyst is established. Rate constants for the uncatalyzed and acid catalyzed reactions are determined. The forward rate constants for the uncatalyzed and acid catalyzed reactions were 1.4 × 10−2 M−1 s−1 and 60 M−1 s−1. Reaction products are toluidine white, phenol, and an azo dye. From the stoichiometric ratios, the major reaction is Pz + 2 TB+ + H2O = PhOH + 2 TBH + 2 H+ + N2. The rate expression and a detailed 12‐step reaction mechanism supported by simulations are proposed. ©1999 John Wiley & Sons, Inc. Int J Chem Kinet: 31: 83–88, 1999  相似文献   

17.
The rate constants and modes of reaction of NO2+ and C2H5ONO2NO2+ with aromatic compounds and alkanes have been determined in a pulsed ion cyclotron resonance mass spectrometer. Both ions undergo competing charge transfer and substitution reactions (NO2+ + M → MO+ + NO; C2H5ONO2NO2+ + M → MNO2+ + C2H5ONO2) with aromatic molecules. In both cases, the probability that a collision results in charge transfer increases with increasing exothermicity of that process. The C2H5ONO2NO2+ ion does not undergo charge transfer with molecules having an ionization potential greater than about 212 kcal/mol (9.2 eV); this observation leads to an estimate of 13 kcal/mol for the binding energy between NO2+ and C2H5ONO2. The importance of the substitution reaction depends on the number of substituents on the aromatic ring and the molecular structure, and, in the case of C2H5ONO2NO2+ ions, on the energetics of the competing charge transfer process. Both NO2+ and C2H5ONO2NO2+ undergo hydride transfer reactions with alkanes. For both these ions, k(hydride transfer)/k (collision) increases with increasing exothermicity of reaction, but in both cases the rate constants of reaction are unusually low when compared with other hydride transfer reactions of comparable exothermicity which have been reported in the literature. This is interpreted as evidence that the attack on the alkane preferentially involves the nitrogen atom (where the charge is localized) rather than one of the oxygen atoms of NO2+.  相似文献   

18.
The principal fragmentation reactions of metastable [C3H7S]+ ions are loss of H2S and C2H4. These reactions and the preceding isomerizations of [C3H7S]+ ions with six different initial structures were studied by means of labelling with 13C or D. From the results it is concluded that the loss of H2S and C2H4 both occur at least mainly from ions with the structure [CH3CH2CH? SH]+ or from ions with the same carbon sulfur skeleton, with the exception of the ions with the initial structure [CH3CH2S? CH2]+, which partly lose C2H4 without a preceding isomerization. For all ions, more than one reaction route leads to [CH3CH2CH?SH]+. It is concluded that the loss of H2S is at least mainly a 1,3-elimination from the [CH3CH2CH?SH]+ ions. Both decomposition reactions are preceded by extensive but incomplete hydrogen exchange.  相似文献   

19.
《Journal of Coordination Chemistry》2012,65(16-18):2905-2912
Abstract

During an effort to synthesize the trans-III-copper(II) complex with 1,4,8,11-tetramethyl-pyro-phosphonate-1,4,8,11-tetra-aza-cyclo-tetradecane, using only perchlorate salts, it was noted that the perchlorate is reduced to chloride. Analysis of the reactions leading to this surprising result points out that Cu(H2O)42+ catalyzes the reduction of perchlorate by H2 and by CH2O. These reactions are slow at room temperature and ambient pressures. A plausible mechanism, supported by DFT calculations, is proposed pointing out that the role of CuH+ under mild conditions cannot be ignored.  相似文献   

20.
A computational study with the M06/B3LYP density functional is carried out to explore the effects of additives C5H5NO vs. PhNO on the gold-catalyzed dehydrogenative heterocyclization of 2-(1-alkynyl)-2-alken-1-ones to form 2,3-furan-fused carbocycles. The following three conclusions are obtained based on our theoretical calculations. (a) The Au(I) catalyst plays a crucial role on the intramolecular cyclization reaction. (b) Both additives C5H5NO and PhNO as the proton shuttle can assist proton-transfer through a two-step proton-transfer mechanism including the protonation of additive and the deprotonation of additive-H+, whereas the catalytic capability of PhNO is weaker than that of C5H5NO (energy barrier: 90.6 vs. 33.2 kJ/mol). (c) C5H5NO-H+ has stronger stability comparing with PhNO-H+ because the basicity of C5H5NO is stronger than that of PhNO, which cause that the energy barrier of ts3 + PhNO-H+ (131.5 kJ/mol) is higher than that of ts3 + C5H5NO-H+ (60.5 kJ/mol) in the intermolecular addition. Therefore, the base strength is the primary factor that controls the catalytic capability of additives C5H5NO vs. PhNO. These studies are expected to improve our understanding of Au(I)-catalyzed reactions involving additive as the cocatalyst and to provide guidance for the future design of new catalysts and new reactions.  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