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 390 毫秒
1.
Mass spectra from collisionally activated dissociation (CAD) of [C2H3O]+ ions, including isotopically labeled analogs, provide further information on the isomers [CH3C?O+] (a), [CH2?C?O+H] (b), [+CH2CH?O] (c) and (d). Our data generally support the recent conclusions from theory by Radom and coworkers and from experiment by Terlouw, Holmes and coworkers. Most acetyl-containing molecular ions form a ions in high purity only at low energies, consistent with isomerization of higher energy molecular ions to form the more stable enol which dissociates to b. Isomer d, prepared from (CICH2)2CHOH, undergoes facile hydrogen scrambling, presumably through a degenerate 1,2-hydrogen shift. Theory suggests that c undergoes spontaneous isomerization to a and d; although [C2H3O]+ ions from BrCH2CHO appear to consist of a and ~15% d, the latter are formed without substantial hydrogen scrambling.  相似文献   

2.
Scrambling data for the three observed [C2H3O]+ isomers, namely [CH3CO]+ (a), [CH2COH]+ (b) and (c), are rationalized by using ab initio molecular orbital calculations. For ions a and c, processes leading to scrambling of the carbon atoms require substantially more energy than the threshold for decomposition to [CH3]+ + CO. Accordingly, little or no carbon scrambling is predicted nor is any observed in the metastable dissociation of a and c. The observed carbon scrambling in b prior to metastable dissociation to [CH3]+ + CO has previously been explained in terms of a mechanism involving the oxiranyl cation (c). However, this mechanism is shown to be unlikely because of the high energies involved. An alternative lower-energy pathway involving the intermediacy of protonated oxirene (h) is proposed. Such a mechanism is fully compatible with the experimental data.  相似文献   

3.
The relative energies of 11 [C3H3O]+ ions are calculated by different molecular orbital methods (MINDO/3, MNDO, ab initio with 3-21G and 4-31G* basis set and configuration interaction). The four most stable structures are: a ([CH2?CH? CO]+), b c ([CH?C? CHOH]+) and d ([CH2?C?COH]+); their relative energies at the CI/4-31G*//3-21G level are 0, 117, 171 and 218 kJ mol?1, respectively. The isomerizations c→[CH?CH? CHO]+→[CH2?C? CHO]+a and dissociations into [C2H3]++CO and [HCO]++C2H2 are explored. The calculated potential energy profile reveals that the energy-determining step is the 1,3-H migration c→[CH?CH? CHO]+. This explains the value of unity of the branching ratio and the spread of kinetic energy released for the two dissociation channels.  相似文献   

4.
The proposed formation of [CH3C(OH)OCH2]+˙ (b) as the intermediate in the isomerization [CH2?C(OH)OCH3]+˙ (c)?b?[CH3COOCH3]+˙ (c has been confirmed by preparation of b from CH3COOCH2OCH3. For the three isomers a–c the dominant metastable ion (MI) dissociation, CH3O˙ loss, involves identical kinetic energy release values. The kinetic barriers for a?b and b?c must be nearly as high as that for CH3O˙ loss from c, as shown by the insensitivity of the mass spectra from collisionally activated dissociation (CAD) of a–c to ionizing electron energy. The H/D scrambling of metastable [CH2?C(OD)OCH3]+˙ and c–D3 ions confirm this, indicating that the barrier for a?b is slightly below that for b?c. Minor low-energy dissociations include losses of CH4 and CH3OH from a and losses of ˙CHO and CH2O from b. Comparison of MI and CAD spectra of a–c with those from [CH3(OH)CH2O]+˙ (d) and [CH3COCH2OH]+˙ (e) give no evidence for skeletal rearrangement of a–c to d or e.  相似文献   

5.
Threshold photoelectron-photoion coincidence (TPEPICO) spectroscopy has been used to investigate the unimolecular chemistry of gas-phase methyl 2-methyl butanoate ions [CH3CH2CH(CH3)COOCH3·+]. This ester ion isomerizes to a lower energy distonic ion [CH2CH2CH(CH3)COHOCH3·+] prior to dissociating by the loss of C2H4. The asymmetric time of flight distributions, which arise from the slow rate of dissociation at low ion energies, provide information about the ion dissociation rates. By modeling these rates with assumed k(E) functions, the thermal energy distribution for room temperature sample, and the analyzer function for threshold electrons, it was possible to extract the dissociative photoionization threshold for methyl 2-methyl butanoate which at 0 K is 9.80 ± 0.01 eV as well as the dissociation barrier of the distonic ion of 0.86 ± 0.01 eV. By combining these with an estimated heat of formation of methyl 2-methyl butanoate, we derive a 0 K heat of formation of the distonic ion CH2CH2CH(CH3)COHOCH3·+ of 101.0 ± 2.0 kcal/mol. The product ion is the enol of methyl propionate, CH3CHCOHOCH3·+, which has a derived heat of formation at 0 K of 106.0 ± 2.0 kcal/mol.  相似文献   

6.
4-Methoxymethylbenzaldimmonium ions (a) and the corresponding N-methylated ions (b) and N,N-dimethylated ions (c) were easily generated in the ion source by electron impact-induced dissociation from 1-(4-methoxymethylphenyl)ethylamine and its N-methylated derivatives. The spontaneous fragmentations of metastable ions a-c and of specifically deuterated derivatives in the second field-free region of a VG ZAB-2F mass spectrometer were studied by mass-analysed ion kinetic energy Spectrometry. The formation of an amino-p-quinodimethane radical cation by loss of the methoxy group is observed for all ions. In the case of a and b carrying at least one proton at the immonium group, competing fragmentations are the loss of CH2O and CH3OH, respectively, and the formation of ions CH3OCH2 +, m/z 45, and C7H7 +, m/z 91. Deuterium-labelling experiments indicated the migration of a proton from the protonated imino group of a and b to the aromatic ring followed by the loss of methanol from the methoxymethyl side-chain or protolysis of the bond to either side-chain to form ion-neutral complexes, in close analogy with the reactions of the corresponding protonated benzaldehydes. The intermediate ion-neutral complexes dissociate eventually by internal ion-neutral reactions resulting in the loss of CH2 O and the formation of C7H7 +, respectively.  相似文献   

7.
Two CrIII–picolinato complexes were obtained and characterized in solution. The [Cr(C2O4)(pyac)2] and [Cr(C2O4)2(pyac)]2– ions (pyac = picolinic acid anion) in acidic solutions undergo a reversible one-end CrIII–picolinato chelate ring opening via CrIII—N bond breaking. The reaction rate was determined spectrophotometrically in the 0.1–1.0 M HClO4 range at I = 1.0 M. The observed pseudo-first order rate constant depends on [H+] according to the equation: k obs = a + b[H+] + c/[H+]. A reaction mechanism, which assumes participation of the protonated and unprotonated forms of the reactants, has been proposed. The kinetic parameters a, b, c have been defined as a = k 1, b = k 2 Q 1, c = k –1/Q 2, where k 1, k –1,k 2 are rate constants for the forward and reverse processes and Q 1, Q 2 are the protolytic equilibrium constants in the term of the proposed mechanism. The activation parameters have been determined and discussed.  相似文献   

8.
The implementation of surface-induced dissociation (SID) to study the fast dissociation kinetics (sub-microsecond dissociation) of peptides in a MALDI TOF instrument has been reported previously. Silicon nanoparticle assisted laser desorption/ionization (SPALDI) now allows the study of small molecule dissociation kinetics for ions formed with low initial source internal energy and without MALDI matrix interference. The dissociation kinetics of N(CH3)4+ and N(CD3)4+ were chosen for investigation because the dissociation mechanisms of N(CH3)4+ have been studied extensively, providing well-characterized systems to investigate by collision with a surface. With changes in laboratory collision energy, changes in fragmentation timescale and dominant fragment ions were observed, verifying that these ions dissociate via unimolecular decay. At lower collision energies, methyl radical (CH3) loss with a sub-microsecond dissociation rate is dominant, but consecutive H loss after CH3 loss becomes dominant at higher collision energies. These observations are consistent with the known dissociation pathways. The dissociation rate of CH3 loss from N(CH3)4+ formed by SPALDI and dissociated by an SID lab collision energy of 15 eV corresponds to log k = 8.1, a value achieved by laser desorption ionization (LDI) and SID at 5 eV. The results obtained with SPALDI SID and LDI SID confirm that (1) the dissociation follows unimolecular decay as predicted by RRKM calculations; (2) the SPALDI process deposits less initial energy than LDI, which has advantages for kinetics studies; and (3) fluorinated self-assembled monolayers convert about 18% of laboratory collision energy into internal energy. SID TOF experiments combined with SPALDI and peak shape analysis enable the measurement of dissociation rates for fast dissociation of small molecules.  相似文献   

9.
Dissociative electron ionization (70eV) of selenophene (C4H4Se) generates m/z 106 ions of composition [H2, C2, 80Se]+? and m/z 105 ions of [H, C2, 80Se]+. From tandem mass spectrometric experiments, Density Functional Theory (DFT) and ab initio calculations, it is concluded that these ions have the structure of selenoketene H2C?C?Se+? (1a+? )and selenoketyl HC?C?Se+ (2a+) ions respectively. The calculations predict that selenoketene ion 1a+? is separated by high energy barriers from its isomers selenirene (H e)+? 1b+?, ethyne selenol (HCCSeH)+? 1c+?, (CCHSeH)+? 1d+? and (CCSeH2)+? 1e+?. The selenoketyl ion 2a+ is separated by high barriers from its isomers (CCHSe)+ 2b+, and (CCSeH)+ 2c+. Neutralization‐reionization mass spectra (NRMS) of these structurally characterized ions confirmed that the corresponding neutral analogues, selenoketene H2CCSe 1a and selenoketyl radical HCCSe 2a? are stable in the rarefied gas phase. The relative, dissociation, and isomerization energies for selenoketene and selenoketyl ions and neutrals studied at B3LYP/6–31G(d,p) and G2/G2(MP2) levels are used to support and interpret the experimental results. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

10.
The kinetics and mechanism for the reaction of NH2 with HNO have been investigated by ab initio calculations with rate constant prediction. The potential energy surface of this reaction has been computed by single‐point calculations at the CCSD(T)/6‐311+G(3df, 2p) level based on geometries optimized at the CCSD/6‐311++G(d, p) level. The major products of this reaction were found to be NH3 + NO formed by H‐abstraction via a long‐lived H2N???HNO complex and the H2NN(H)O radical intermediate formed by association with 26.9 kcal/mol binding energy. The rate constants for formation of primary products in the temperature range of 300–3000 K were predicted by variational transition state or RRKM theories. The predicted total rate constants at the 760 Torr Ar pressure can be represented by ktotal = 3.83 × 10?20 × T+2.47exp(1450/T) at T = 300–600 K; 2.58 × 10?22 × T+3.15 exp(1831/T) cm3 molecule?1 s?1 at T = 600?3000 K. The branching ratios of major channels at 760 Torr Ar pressure are predicted: k1 + k3 + k4 producing NH3 + NO accounts for 0.59–0.90 at T = 300–3000 K peaking around 1000 K, k2 accounts for 0.41–0.03 at T = 300–600 K decreasing with temperature, and k5 accounts for 0.07–0.27 at T > 600 K increasing gradually with temperature. The NH3 + NO formation rate constant was found to be a factor of 3–10 smaller than that of the isoelectronic reaction CH3 + HNO producing CH4 + NO, which has been shown to take place by barrierless H‐abstraction without involving a hydrogen‐bonding complex as in the NH2 case. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 677–677, 2009  相似文献   

11.
Ionization energies of substituted trimethylamines (CH3)2NCH2X, together with appearance energies of [M ? X]+ ions, have been measured by the photoionization technique. The heats of formation for the following radicals have been derived (in kJ mol?1): C2H3˙ (+ 314), Ph. (+ 347), p-CH3C6H4˙ (+ 326), HC2˙ (+ 494), (CH3)2N?H2 (+ 134), NC˙ (+ 452), CF3. (? 435), (CH3)3Si˙ (+ 54.5). The heats of formation obtained for [(CH3)2NCH2]+ (+ 649) and (CH3)3Si+ (+ 610) lead to ionization energies of 5.35 eV for (CH3)2N?H2 and 5.75 eV for (CH3)3Si˙.  相似文献   

12.
Zwitterionic diazeniumdiolates of the form RN[N(O)NO?](CH2)2NH2+R, where R=CH3 ( 1 ), (CH2)3CH3 ( 2 ), (CH2)5CH3 ( 3 ), and (CH2)7CH3 ( 4 ) were synthesized by reaction of the corresponding diamines with nitric oxide. Spectrophotometrically determined pKa(O) values, attributed to protonation at the terminal oxygen of the diazeniumdiolate group, show shifts to higher values in dependence of the chain lengths of R. The pH dependence of the decomposition of NO donors 1 – 3 was studied in buffered solution between pH 5 and 8 at 22 °C, from which pKa(N) values for protonation at the amino nitrogen, leading to release of NO, were estimated. It is shown that the decomposition of these diazeniumdiolates is markedly catalyzed by anionic SDS micelles. First‐order rate constants for the decay of 1 – 4 were determined in phosphate buffer pH 7.4 at 22 °C as a function of SDS concentration. Micellar binding constants, KSM, for the association of diazeniumdiolates 1 – 3 with the SDS micelles were also determined, again showing a significant increase with increasing length of the alkyl side chains. The decomposition of 1 – 3 in micellar solution is quantitatively described by using the pseudo‐phase ion‐exchange (PIE) model, in which the degree of micellar catalysis is taken into account through the ratio of the second‐order rate constants (k2m/k2w) for decay in the micelles and in the bulk aqueous phase. The decay kinetics of 1 – 3 were further studied in the presence of cosolvents and nonionic surfactants, but no effect on the rate of NO release was observed. The kinetic data are discussed in terms of association to the micelle–aqueous phase interface of the negatively charged micelles. The apparent interfacial pH value of SDS micelles was evaluated from comparison of the pH dependence of the first‐order decay rate constants of 2 and 3 in neat buffer and the rate data obtained for the surfactant‐mediated decay. For a bulk phase of pH 7.4, an interfacial pH of 5.7–5.8 was determined, consistent with the distribution of H+ in the vicinity of the negatively charged micelles. The data demonstrate the utility of 2 and 3 as probes for the determination of the apparent pH value in the Stern region of anionic micelles.  相似文献   

13.
Rate constants for the gas‐phase reactions of CH3OCH2CF3 (k1), CH3OCH3 (k2), CH3OCH2CH3 (k3), and CH3CH2OCH2CH3 (k4) with NO3 radicals were determined by means of a relative rate method at 298 K. NO3 radicals were prepared by thermal decomposition of N2O5 in a 700–750 Torr N2O5/NO2/NO3/air gas mixture in a 1‐m3 temperature‐controlled chamber. The measured rate constants at 298 K were k1 = (5.3 ± 0.9) × 10?18, k2 = (1.07 ± 0.10) × 10?16, k3 = (7.81 ± 0.36) × 10?16, and k4 = (2.80 ± 0.10) × 10?15 cm3 molecule?1 s?1. Potential energy surfaces for the NO3 radical reactions were computationally explored, and the rate constants of k1k5 were calculated according to the transition state theory. The calculated values of rate constants k1k4 were in reasonable agreement with the experimentally determined values. The calculated value of k5 was compared with the estimate (k5 < 5.3 × 10?21 cm3 molecule?1 s?1) derived from the correlation between the rate constants for reactions with NO3 radicals (k1k4) and the corresponding rate constants for reactions with OH radicals. We estimated the tropospheric lifetimes of CH3OCH2CF3 and CHF2CF2OCH2CF3 to be 240 and >2.4 × 105 years, respectively, with respect to reaction with NO3 radicals. The tropospheric lifetimes of these compounds are much shorter with respect to the OH reaction. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 490–497, 2009  相似文献   

14.
The dissociative photoionization of molecular‐beam cooled CH2CO in a region of ?10–20 eV was investigated with photoionization mass spectrometry using a synchrotron radiation as the light source. Photoionization efficiency curves of CH2CO+ and of observed fragment ions CH2+, CHCO+, HCO+, C2O+, CO+, and C2H2+ were measured to determine their appearance energies. Relative branching ratios as a function of photon energy were determined. Energies for formation of these observed fragment ions and their neutral counterparts upon ionization of CH2CO are computed with the Gaussian‐3 method. Dissociative photoionization channels associated with six observed fragment ions are proposed based on comparison of determined appearance energies and predicted energies. The principal dissociative processes are direct breaking of C=C and C‐H bonds to form CH2+ + CO and CHCO+ + H, respectively; at greater energies, dissociation involving H migration takes place.  相似文献   

15.
Absolute rate constants for H-atom abstraction by OH radicals from cyclopropane, cyclopentane, and cycloheptane have been determined in the gas phase at 298 K. Hydroxyl radicals were generated by flash photolysis of H2O vapor in the vacuum UV, and monitored by time-resolved resonance absorption at 308.2 nm [OH(A2Σ+X2Π)]. The rate constants in units of cm3 mol−1 s−1 at the 95% confidence limits were as follows: k(c C3H6) = (3.74 ± 0.83) × 1010, k(c C5H10) = (3.12 ± 0.23) × 1012, k(c C7H14) = (7.88 ± 1.38) × 1012. A linear correlation was found to exist between the logarithm of the rate constant per C H bond and the corresponding bond dissociation energy for several classes of organic compounds with equivalent C H bonds. The correlation favors a value of D(c C3H5–H) = (101 ± 2) kcal mol−1.  相似文献   

16.
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   

17.
The electron impact (EI) ionization-induced fragmentation pathways of the new 1,9-bis(dimethylamino) phenalenium cation [1]+ were investigated. The peri-dimethylamino substituents of [1]+ are incorporated in a trimethine cyanine substructure and show strong steric interactions. A mechanism is proposed for the unusual elimination of CH3N?CH2, HN(CH3)2 and (CH3)3N from [1]+ and for the accompanying cyclizations to heterocyclic ions: prior to fragmentation, the intact cation [1]+ rearranges, by reciprocal CH3 and H transfers, to new isomeric cations which decompose subsequently in a characteristic way. A wealth of consistent information on dissociation pathways and fragment structures is provided by collision-induced dissociation tandem mass spectra, collision-induced dissociation mass-analysed ion kinetic energy spectra and exact mass measurements of the salt cation and of its primary fragment ions. The liquid secondary ion mass spectrum of [1]+ is very similar to its EI mass spectrum.  相似文献   

18.
The neutral counterparts of the C2H7O+ isomers CH3O+ (H)CH3, CH3CH2OH2+ and $ {\rm C}_2 \,{\rm H}_4 \,\, \cdot \cdot \cdot \mathop {\rm H}\limits^ + \, \cdot \cdot \cdot {\rm OH}_2 $ were studied by neutralization-reionization mass spectrometry. Protonated dimethyl ether and its —O(D)+ analogue were produced by protonation (deuteration) of dimethyl ether and also generated as a fragment ion from (labeled) ionized CH3OCH2CH(OH)CH3 by loss of CH3CO?. It was observed that the dissociation characteristics of the ions and the stability of their neutral counterpart depended on the internal energy of the protonated ether ions. Stable neutral CH3?(H)CH3 was only produced from energy-rich ions. The classical protonated ethanol ion CH3CH2OH2+ (a) was produced at threshold by the loss of CH3CO?. from ionized butane-2,3-diol. Mixtures of a with the non-classical ion $ {\rm C}_2 \,{\rm H}_4 \,\, \cdot \cdot \cdot \mathop {\rm H}\limits^ + \, \cdot \cdot \cdot {\rm OH}_2 $ (b) were produced by reaction of C2H5+ ions with H2O. As for the protonated ether, only high-energy a and/or b ions yielded stable hypervalent radicals. It is suggested that the stable C2H7?O radicals are Rydberg states.  相似文献   

19.
CEPA-PNO and PNO-CI calculations have been performed for the potential energy curves of the He 2 + ground state and the six lowest excited states of He2 in the range of 1.4 a0R ≤ 3.5 a0. The calculated equilibrium distances as well as the spectroscopic constants are in very good agreement with molecular constants as derived experimentally from the rotation-vibration spectrum of He2 by Ginter, except for thec 3g + state. This latter discrepancy is probably due to an “obligatory” hump in thec 3g + state occurring at 3.5 a0 which cannot be properly treated in our calculation. The relative energetic positions of the six lowest states and their ionization energies are reproduced by our calculations with an accuracy of 0–400 cm−1. Extrapolation of our results to infinite basis sets leads to estimates of the dissociation energies of He2 excited states which cannot be measured spectroscopically because of the humps in all these states.  相似文献   

20.
Theoretical investigations are carried out on the multichannel reactions CH3COCH3 + F (R1) and CH3COCH3 + Br (R2) by means of direct dynamics methods. The minimum energy path (MEP) is obtained at the MP2/6-31 + G(d,p) level, and energetic information is further refined at the MC-QCISD (single-point) level. The rate constants are calculated by the improved canonical variational transition-state theory (ICVT) with the small-curvature tunneling (SCT) contributions in a wide temperature range 200–1,500 K for the title reactions, H-abstraction channel is favored for the two reactions. The theoretical overall rate constants are in good agreement with the available experimental data and are found to be k 1a  = 3.22 × 10−15 T 1.51exp(1,190.91/T) cmmolecule−1 s−1, k 2  = 5.95 × 10−18 T 1.98exp(−4,622.45/T) cmmolecule−1 s−1. Furthermore, the rate constants of reaction Cl + CH3COCH3 (R3) calculated in the other paper are added to discuss the reactivity trend of different halogen reaction with acetone on the rate constants of this class of hydrogen abstraction 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