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.
Hydrogen isotope exchange reaction rate in tritium and methane mixed gas, as induced by tritium decay and beta radiation, has been experimentally measured. Initially T2 gas was filled to 40 kPa and 20 kPa of CH4 gas was added. The mixed gas spectrum was analyzed periodically by laser Raman spectrometry. The first order HT and H2 formation rates and T2 and CH4 decay rates by hydrogen isotope exchange reaction were observed between 2.9·10–3 h–1 and 4.8·10–3 h–1. Although the estimated hydrogen isotope exchange reaction rate was 1/20–1/10 slower than the rate of H2+T2 mixed gases, it was nearly equivalent to the ion formation rate by tritium beta radiation. This suggested that isotopic hydrogen radicals formed via ionization would disappear in the presence of methane.  相似文献   

2.
The rate constant for the exchange between Hg2+ and CH3Hg+ ions in a Cl-free 1NH2SO4 solution at 25°C was measured with203Hg as a radiotracer and separation of the two compounds by liquid extraction of the organic species into benzene from a NaCl- or NaI-solution.  相似文献   

3.
Fluorinated Organoelements: Oxidative Liquid-Phase Direct Fluorination. X. Organyloxyfluorophosphoranes: Direct Synthesis by F2-Addition to Phosphinic-, Phosphonic-, and Phosphoric-Acid Ester(fluorides) and Thermal Behaviour The phenoxyfluorophosphoranes (PhO)2PF2R ( 2a: R = CH3, 2b: R = Ph) and (PhO)3?nPFn+2 ( 4a: n = 0, 4b: n = 1, 4c: n = 2) were obtained in reasonable yield by direct fluorination of the corresponding organyloxy(fluoro)phosphanes for the first time. Contrary to 4c the intramolecular ligand exchange can be frozen up in 4b. The up to now unknown thermally unstable alkoxy-substituted difluorides (AlkO)3?PF2Rn ( 6a: Alk = CH3, R = Ph), n = 2; 6b: Alk = CH2CF3, R = Ph, n = 2; 6c: Alk = CH3, R = Ph, n = 1; 6d: Alk = CH3, n = 0) were isolated by low temperature F2-addition in pure substance, too. Their thermal decomposition (scrambling, CH3F-elimination) was cleared up for 6a as model substance and transferred to (CH3O)3PF2 6d Here the splitting of (CH3)2O under “Arbusov-conditions” is very surprising. The trigonal bipyramidal covalent structure of all organyloxyphosphoranes was confirmed by multinuclear 19F, 31P{1H}, 13C{1H}) NMR experiments.  相似文献   

4.
The 220 MHz 1H NMR spectrum of an ether solution of CH3Li and LiBr in 10–1 ratio has been examined as a function of temperature. At low temperature distinct resonances, assignable to Li4(CH3)4 and Li4(CH3)3Br, are seen. Methyl group exchange between the two tetramers is observed in the NMR spectra in the temperature interval ?32 to 0°. The exchange is shown to be much slower than the dissociation of Li4(CH3)4 tetramer, measured in other work. It is proposed that the rate-determining step is dissociation of Li4(CH3)3Br to form Li2(CH3)2 and Li2(CH3)Br. The rate constant for dissociation, k2, obeys the equation ln k2 = 36.0?83303/T.  相似文献   

5.
Product distributions and rate constants for the reaction of ground state C+ ions with O2, NO, HCl, CO2, H2S, H2O, HCN, NH3, CH4, H2CO, CH3OH, and CH3NH2 have been measured. Rate constants were obtained using ion cyclotron resonance trapped ion methods at JPL, and product distributions were obtained using a tandem (Dempster-ICR) mass spectrometer at the University of Utah. Rapid carbon isotope exchange has also been observed in C+-CO collisions.  相似文献   

6.
The thermal ion‐molecule reactions NiX++CH4→Ni(CH3)++HX (X=H, CH3, OH, F) have been studied by mass spectrometric methods, and the experimental data are complemented by density functional theory (DFT)‐based computations. With regard to mechanistic aspects, a rather coherent picture emerges such that, for none of the systems studied, oxidative addition/reductive elimination pathways are involved. Rather, the energetically most favored variant corresponds to a σ‐complex‐assisted metathesis (σ‐CAM). For X=H and CH3, the ligand exchange follows a ‘two‐state reactivity (TSR)’ scenario such that, in the course of the thermal reaction, a twofold spin inversion, i.e., triplet→singlet→triplet, is involved. This TSR feature bypasses the energetically high‐lying transition state of the adiabatic ground‐state triplet surface. In contrast, for X=F, the exothermic ligand exchange proceeds adiabatically on the triplet ground state, and some arguments are proposed to account for the different behavior of NiX+/Ni(CH3)+ (X=H, CH3) vs. NiF+. While the couple Ni(OH)+/CH4 does not undergo a thermal ligand switch, the DFT computations suggest a potential‐energy surface that is mechanistically comparable to the NiF+/CH4 system. Obviously, the ligands X act as a mechanistic distributor to switch between single vs. two‐state reactivity patterns.  相似文献   

7.
The deposition of diamondlike carbon (DLC) film and the measurements of ionic species by means of mass spectrometry were carried out in a CH4/N2 RF (13.56 MHz) plasma at 0.1 Torr. The film deposition rate greatly depended on both CH4/N2 composition ratio and RF power input. It was decreased monotonically as CH4 content decreased in the plasma and then rapidly diminished to negligible amounts at a critical CH4 content, which became large for higher RF power. The rate increased with increasing RF power, reaching a maximum value in 40% CH4 plasma. The predominant ionic products in CH4/N2 plasma were NH+ 4 and CH4N+ ions, which were produced by reactions of hydrocarbon ions, such as CH+ 3, CH+ 2, CH+ 5, and C2H+ 5 with NH3 molecules in the plasma. It was speculated that the production of NH+ 4 ion induced the decrease of C2H+ 5 ion density in the plasma, which caused a reduction in higher hydrocarbon ions densities and, accordingly, in film deposition rate. The N+ 2 ion sputtering also plays a major role in a reduction of film deposition rate for relatively large RF powers. The incorporation of nitrogen atoms into the bonding network of the DLC film deposited was greatly suppressed at present gas pressure conditions.  相似文献   

8.
The rate constant for the reaction I(2P1/2) + CH3I → I2 + CH3 has been reevaluated taking into account both collisional deactivation of excited iodine atoms and loss of I2 by I2 + CH3 → I + CH3I. The reevaluation is based upon data obtained (R. T. Meyer), J. Chem. Phys., 46 , 4146 (1967) from the flash photolysis of CH3I using time-resolved mass spectrometry to measure the rate of I2 formation. Computer simulations of the complete kinetic system and a closed-form solution of a simplified set of the differential equations yielded a value of 6(± 4) × 106 1./mole-sec for the excited iodine atom reaction in the temperature region of 316 to 447 K. A slight temperature dependence was observed, but an activation energy could not be evaluated quantitatively due to the small temperature range studied. An upper limit for the collisional deactivation of I(2P1/2) with CH3I was also determined (2.4 × 107 1./mole-sec).  相似文献   

9.
The relative rate technique has been used to measure the hydroxyl radical (OH) reaction rate constant of +2-butanol (2BU, CH3CH2CH(OH)CH3) and 2-pentanol (2PE, CH3CH2CH2CH(OH)CH3). 2BU and 2PE react with OH yielding bimolecular rate constants of (8.1±2.0)×10−12 cm3molecule−1s−1 and (11.9±3.0)×10−12 cm3molecule−1s−1, respectively, at 297±3 K and 1 atmosphere total pressure. Both 2BU and 2PE OH rate constants reported here are in agreement with previously reported values [1–4]. In order to more clearly define these alcohols' atmospheric reaction mechanisms, an investigation into the OH+alcohol reaction products was also conducted. The OH+2BU reaction products and yields observed were: methyl ethyl ketone (MEK, (60±2)%, CH3CH2C((DOUBLEBOND)O)CH3) and acetaldehyde ((29±4)% HC((DOUBLEBOND)O)CH3). The OH+2PE reaction products and yields observed were: 2-pentanone (2PO, (41±4)%, CH3C((DOUBLEBOND)O)CH2CH2CH3), propionaldehyde ((14±2)% HC((DOUBLEBOND)O)CH2CH3), and acetaldehyde ((40±4)%, HC((DOUBLEBOND)O)CH3). The alcohols' reaction mechanisms are discussed in light of current understanding of oxygenated hydrocarbon atmospheric chemistry. Labeled (18O) 2BU/OH reactions were conducted to investigate 2BU's atmospheric transformation mechanism details. The findings reported here can be related to other structurally similar alcohols and may impact regulatory tools such as ground level ozone-forming potential calculations (incremental reactivity) [5]. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 745–752, 1998  相似文献   

10.
Evidence is presented that the dimeric π-allylic species [(η3-allyl)PdCl]2 is not intermediate in the Li2Pd2Cl6-catalysed allylic H/D exchange in alkenes. Neither H/D exchange in α-methylstyrene, nor enrichment of [(η3-2-PhC3H4)PdCl]2, was observed when the latter complex was incubated at 100°C in D3CCOOD either in the presence or in the absence of PhC(CH3)?CH2, respectively. The kinetics of H/D exchange in α-methylstyrene catalysed by Li2Pd2Cl6 were studied in some detail. The exchange proceeds at highest rates when reduction of palladium(II) takes place and is much slower in the presence of 1,4-benzoquinone as a palladium reoxidant. The exchange rate is directly proportional to the alkene and catalyst concentrations and independent of the reoxidant concentration. It is suggested that the palladium(II)-catalysed exchange involves an intermediate hydrid-allyl species where palladium has a formal oxidation state of IV.  相似文献   

11.
The fluxional behaviour of triphenylmethyl substituted cyclopentadienyl metal compounds 1H-NMR. spectroscopy has been used to study the influence of the triphenylmethyl substituent on the metallotropic rearrangement in C5H4(CPh3)Si(CH3)3 and C5H4 (CPh3)Sn(CH3)3. The most likely mechanism corresponds to a degenerate metal exchange between two neighbouring ring positions. The AA'XX' spectrum of the cyclopentadienyl ring protons in C5H4(CPh3)Sn(CH3)3 has been analysed under rapid exchange conditions. The free energy of activation for the sigmatropic [1,5]-Sn shift has been measured by comparison with computer simulated spectra for slow exchange.  相似文献   

12.
Ab initio calculations of fragments of the potential energy surfaces of hydrogen exchange reactions between H2, CH4, and alanine molecules and the H3O+ ion were performed by the restricted Hartree-Fock method, at the second-order Møller-Plesset level of perturbation theory, and by the method of coupled clusters using the 6–31G* and aug-cc-pVDZ basis sets. The one-center synchronous mechanism of hydrogen exchange reaction was studied and the activation energies and structures of transition states were determined. It was found that the geometric parameters of the H2 and CH4 molecules in the transition states are close to those of the H3 + and CH5 + ions. The higher the proton affinity of the reacting molecule in the reaction studied the lower the activiation energy of hydrogen exchange. The one-center mechanism studied can be used to describe the high-temperature solid-state catalytic isotope exchange (HSCIE) reaction. The results ofab initio calculations of synchronous hydrogen exchange between the H3O+ ion and hydrogen atoms in different positions of the alanine molecule are in good agreement with experimental data on the regioselectivity and stereoselectivity of the HSCIE reaction with spillover-tritium.  相似文献   

13.
Triplet methylene, CH2(3B1), and methyl radicals were produced by flash photolysis of a mixture of ketene and azomethane. A computer fit of the product ratios, using the known rate constants for CH2 + CH2, and CH3 + CH3, requires a rate constant of 5.0 × 10?11 cm3 molecule?1s?1 for the reaction CH2 + CH3 ? C2H4 + H.  相似文献   

14.
The ligand exchange MX5·L + *L?MX5·*L + L for the octahedral adducts MX5·L, in an inert solvent (CH2Cl2 or CHCl3) with neutral ligands, proceeds via a dissociative D mechanism when M = Nb, X = Cl and L = phosphoryl compound. A dissociative interchange Id mechanism is suggested when M = Nb or Ta, and X = F. A first order rate law and positive values for ΔS* (+4 to +14 cal K?1 mol?1) are observed for the exchanges on the pentachloride adducts. However, a second order rate law and large negative values for ΔS* (-15 to -24 cal K?1 mol?1) are found for the intermolecular neutral ligand exchange (measured by 1H-NMR.) and for the intramolecular fluorine exchange (measured by 19F-NMR.) reactions on the pentafluoride adducts. The fluorine exchange is 2 to 5 times faster than the ligand exchange. The exchanges, on the pentachloride and on the pentafluoride adducts, are slowed down with increasing donor strength of the phosphoryl compound.  相似文献   

15.
The rate coefficients and partial rate factors for the hydrogen-deuterium exchange of some 1-substituted pyrroles (1-substituents = -COCH3, - COC6H5, -SO2CH3, -SO2CF3, -N(CH3)3, -NH(CH3)2, -SO2C6H5) in deuterated trifluoroacetic acid have been determined. In all cases the rate of exchange is faster at the 2-position. Similarly, the hydrogen-deuterium exchange of some pyrroles with electron withdrawing substituents in the 2-position indicate a relative reactivity of 4→ 5→ 3-position with the selectivity being greatest for the more electron withdrawing groups. A nitro group in the 3-position of pyrrole shows a relative reactivity of 5→2→4-position for the hydrogen-deuterium exchange in deuterated trifluoroacetic acid. A linear correlation is observed for the log of the rate coefficient of the hydrogen-deuterium exchange at the 4-position of some 2-substituted pyrroles and the difference in the calculated energy of formation of the pyrrole and the corresponding 4-deuterated cation.  相似文献   

16.
We study dynamics of the CH3 + OH reaction over the temperature range of 300–2500 K using a quasiclassical method for the potential energy composed of explicit forms of short‐range and long‐range interactions. The explicit potential energy used in the study gives minimum energy paths on potential energy surfaces showing barrier heights, channel energies, and van der Waals well, which are consistent with ab initio calculations. Approximately, 20% of CH3 + OH collisions undergo OH dissociation in a direct‐mode mechanism on a subpicosecond scale (<50 fs) with the rate coefficient as high as ~10?10 cm3 molecule?1 s?1. Less than 10% leads to the formation of excited intermediates CH3OH? with excess vibrational energies in CO and OH bonds. CH3OH? stabilizes to CH3OH, redissociates back to reactants, or forms one of various products after intramolecular energy redistribution via bond dissociation and formation on the time scale of 50–200 fs. The principal product is 1CH2 (k being ~10?11), whereas ks for CH2OH, CH2O, and CH3O are ~10?12. The minor products are HCOH and CH4 (k~10?13). The total rate coefficient for CH3 + OH → CH3OH? → products is ~10?11 and is weakly dependent on temperature. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 455–466, 2011  相似文献   

17.
The electrochemical reduction of dichlorodifluoromethane (CFC-12) at Ag, Pd, Cu and Au electrodes (which were deposited on Nafion® 117 (H+ form) membranes, by reduction with NaBH4 (10% w/v) solution) was studied. The products of the reduction were CHClF2, CH2F2, CH3F and CH4, as well as small amounts of dimers, CF2=CF2 and CHF2CHF2. The silver electrode gave the highest current efficiency (CE) and reduction rate. The rate of reduction at the silver electrode was almost 10–160 times higher than that measured for the other electrodes, under the same conditions. Selectivity of CH4 production increased for all metals with increasing negative potential, except for CHClF2 where it decreased. For the other products, a maximum in the selectivity–potential curve appeared. This fact led us to the conclusion that the reduction proceeds by the following mechanism: CCl2F2→CHClF2→CH2F2→CH3F→CH4. The rate of reduction of CFC-12 and the product distribution also depend on the pH of the solution, which is in contact with the membrane. The rate of reduction at the silver electrode was about 4000 times higher at pH 14 than at pH 1. The cation of the supporting electrolyte was also important: the rate of reduction was lowered in the order K+>Na+>Li+, and this was attributed to the size of the cations, which influenced the structure of the double layer.  相似文献   

18.
We present an approach to couple ab initio quantum mechanical geometry optimiuzations with molecular mechanical optimizations, with the added capability to carry out molecular dynamics simulations of the systems to earch for new local minima. The approach is applied to the aqueous solution CH3Cl + Cl? exchange reaction and the gas phase protonation of polyethers.  相似文献   

19.
《Polyhedron》2003,22(14-17):1727-1733
The syntheses and physical properties are reported for three single-molecule magnets (SMMs) with the composition [Ni(hmp)(ROH)Cl]4, where R is CH3 (complex 1), CH2CH3 (complex 2) or CH2CH2C(CH3)3 (complex 3) and hmp is the monoanion of 2-hydroxymethylpyridine. The core of each complex is a distorted cube formed by four NiII ions and four alkoxide hmp oxygen atoms at alternating corners. Ferromagnetic exchange interactions give a S=4 ground state. Single crystal high-frequency EPR spectra clearly indicate that each of the complexes has a S=4 ground state and that there is negative magnetoanisotropy, where D is negative for the axial zero-field splitting z2. Magnetization versus magnetic field measurements made on single crystals with a micro-SQUID magnetometer indicate these Ni4 complexes are SMMs. Exchange bias is seen in the magnetization hysteresis loops for complexes 1 and 2.  相似文献   

20.
Abstract

The ligand exchange reaction between [M(phen)3]2+ and [M(DIP)3]2+ (where M is the same and M = FeII or NiII, phen = 1,10-phenanthroline, DIP = 4,7-diphenyl-1,10-phenanthroline) has been investigated by reversed phase ion-paired chromatography (RP-IPC). The effect of pH and solvent on the ligand-exchange reaction is studied by monitoring the variation in chromatograms with time after mixing. The results have shown that the ligand exchange reaction between [M(phen)3]2+ and [M(DIP)3]2+ takes place in the pH range of 3–8 and the rate of reaction for nickel(II) complexes is about two times slower than that for iron(II) complexes. Experiments on the effect of various solvents on the ligand-exchange reaction have revealed that the rate of reaction is enhanced by the solvent in the following order: (CH3)2CO > CHCl3 ≥ CH2Cl2 > CH3CN > CH3OH. Elemental analysis and UV-visible spectroscopy confirmed that the products obtained from the ligand-exchange reaction are mixed-ligand complexes containing phen and DIP ligands, i.e., [M(phen)2(DIP)]2+ and [M(phen)(DIP)2]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