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.
The unimolecular and collision-induced fragmentation reactions of the enolate ion of 2,3-butanedione, [CH3COCOCH2]?, have been studied, Unimolecular fragmentation on the metastable ion time-scale forms [HCCO]?, [C2H3O]?, [C3H5O]? and [CH3CO2]?. Charge inversion mass spectrometry shows that the [C2H3O]? ion is the acetyl anion while the [C3H5O]? product is the acetone enolate ion; formation of the latter product involves a large release of kinetic energy (T 1/2 = 0.99 eV). The fragmentation reactions occurring following collisional activation have been determined for 8 keV collisions and over the range 1.5–30 eV center-of-mass collision energy. Formation of [HCCO]? and [CH3CO]? are of the most important reactions following collisional activation and it is concluded that the two reactions have similar critical reaction energies even though formation of [HCCO]? is favored thermochemically.  相似文献   

2.
The energy dependence of the fragmentation of a selection of ester enolate ions has been studied by variable, low-energy collision-induced dissociation experiments in the quadrupole collision cell of a hybrid BEQQ mass spectrometer. The dominant fragmentation reactions observed are where ΔH1 ? ΔH2=PA([RCCO]?) ? PA([?O]?) (PA=proton affinity). The anion of lowest proton affinity is formed preferentially at low internal energies with the yield of the anion of higher proton affinity increasing with increasing internal energy. The [CH3OCOCOCH2]? anion derived from methyl pyruvate forms [CH3OCO]? by reaction (2); this anion readily fragments to [CH3G]?+ CO consistent with a structure represented by a dipole-stabilized cluster of [CH3O]? and CO. Comparison of the 8-keV with the 50-eV collision-induced dissociation mass spectra indicated that the average internal energy of the fragmenting ions is considerably lower in the high-energy collisional experiments than it is in the low-energy collisional experiments.  相似文献   

3.
Specific ion/molecule reactions are demonstrated that distinguish the structures of the following isomeric organosilylenium ions: Si(CH3) 3 + and SiH(CH3)(C2H5)+; Si(CH3)2(C2H5)+ and SiH(C2H5) 2 + ; and Si(CH3)2(i?C3H7)+, Si(CH3)2(n?C3H7)+, Si(CH3)(C2H5) 2 + , and Si(CH3)3(π?C2H4)+. Both methanol and isotopically labeled ethene yield structure-specific reactions with these ions. Methanol reacts with alkylsilylenium ions by competitive elimination of a corresponding alkane or dehydrogenation and yields a methoxysilylenium ion. Isotopically labeled ethene reacts specifically with alkylsilylenium ions containing a two-carbon or larger alkyl substituent by displacement of the corresponding olefin and yields an ethylsilylenium ion. Methanol reactions were found to be efficient for all systems, whereas isotopically labeled ethene reaction efficiencies were quite variable, with dialkylsilylenium ions reacting rapidly and trialkylsilylenium ions reacting much more slowly. Mechanisms for these reactions and differences in the kinetics are discussed.  相似文献   

4.
The collisional charge inversion and neutralization-reionization (?NR) mass spectra of the enolate ions of m/z 115 derived from the four butyl acetates, the two propyl propionates, ethyl butyrate, ethyl isobutyrate, methyl valerate, methyl 2-methylbutyrate and methyl 3-methylbutyrate were recorded. The major primary fragmentation reactions of the unstable carbenium ion formed by charge inversion involve elimination of an alkoxy radical to form a ketene or alkylketene molecular ion and formation of an alkyl ion consisting of the R1 group of RCOOR1. A minor fragmentation reaction involves elimination of an alkyl radical by cleavage of a C? C bond α to the ether oxygen. The alkylketene ions fragment by β-cleavage eliminating an alkyl radical to form an olefinic acylium ion. In most cases the charge inversion mass spectra of the enolate ions allow identification of the ester.  相似文献   

5.
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   

6.
Deprotonation of methyl acetoacetate yields two enolate ions MeCOC?HCO2Me (a) and C?H2COCH2CO2Me (b). On collisional activation, ions a and b fragment differently. The major fragmentation of a is specific loss of MeOH through a four-centred transition state to form ?O(Me)C?C?C?O. In contrast, ion b eliminates CH2CO to give ?CH2CO2Me. Some rearrangement of b to a is also noted. Rearrangement of a to b is very minor under single collision conditions but at high collision gas pressure rearrangement of a to b is strongly promoted. Similar effects are observed in the collisional activation spectra of MeCOC?(Me)CO2Me (c) and ?CH2COC(Me)CO2Me (d). The loss of MeOH from (c) proceeds via a six membered transition state to ?CH2? CO? C(Me)?C?O; this is a stepwise process in which the deprotonation (step two) is not rate determining. A number of other decompositions occur, these have also been studied by deuterium labelling.  相似文献   

7.
The mass spectrometric investigation of specifically deuterium and 13C labelled 2-trimethylsilyl-l-phenoxyethanes proves that the dissociative ionization of β-silyl-substituted ethane derivatives (loss of PhO?; p-CH3C6H4O?; and C4H?9 from PhOCH2CH2SiMe3, p-MeC6H4OCH2CH2SiMe3 and CH3CH2CH(CH3)CH2-CH2SiMe3, respectively) yields the non-classical bridge ethylene trimethylsilanium ion and not the open-chain isomer. Other stable C5H13Si+? ions, characterised by collisional activation mass spectrometry, are the dimethyl n-propyl silicenium ion and the l-trimethylsilyl ethyl cation, both generated from the molecular ions of CH3CH2CH2Si(Cl)Me2 and CH3CH(Cl)SiMe3 via unimolecular loss of Cl?.  相似文献   

8.
Nucleophilic attack by the acetonyl anion (CH3COCH) on highly electron‐deficient nitroaromatics such as trinitrotoluene (TNT) results in the carbon‐bonded anionic σ‐complexes (Meisenheimer complexes). The complexes are generated at atmospheric pressure with relatively low internal energy and do not dissociate under atmospheric pressure chemical ionization conditions, but can be dissociated upon collisional activation to produce characteristic nitrobenzyl‐substituted acetonyl anions via elimination of HNO2. Further fragmentation produces deprotonated nitrotoluene anions through the loss of CH2CO, while loss of CH2COCH3 and CH3COCH3 yields nitroaromatic molecular anions and their H‐loss counterparts. Hydrogen/deuterium (H/D) scrambling is observed in the fragmentation products of the [TNT · CD3COCD2]? complex to extents which vary for different fragmentation pathways. Selective Meisenheimer complex formation, together with its distinctive fragmentation pattern, supplies a highly discriminatory method for detection of TNT and related compounds. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

9.
Methylcobalt Compounds with Non-chelating Ligands, IV. Monoolefin Complexes Tris(trimethylphosphane)cobalt(I) halides in ether solvents saturated with olefin at low temperatures from monoolefin complexes which are prone to dissociation. Upon reaction with methyl-or phenyllithium more stable compounds are formed of the composition CoR(CC)L3 ( 1 – 4 ) (R  CH3; CC  C2H4, C3H6, cyclo-C5H8; R  C6H5; CC  C2H4; L  P(CH3)3). In solution the fluctional molecules adopt a ground state structure containing a σ-bonded group and an olefin ligand in adjacent positions (trigonal-bipyramidal: CH3 axial and C2H4 equatorial or C6H5 and C2H4 equatorial). The latter arrangement is confirmed for the crystalline state by an X-ray structure determination of (ethene)phenyltris(trimethylphosphane)cobalt ( 4 ). An equatorial plane of coordination along a Co P bond not only contains both ethene-C atoms but also all the atoms of the phenyl group. The compound is thermally decomposed to give biphenyl and (ethene)tris-(trimethylphosphane)cobalt(0). No products of an olefin insertion reaction are observed.  相似文献   

10.
The elimination of ethene from CH3CH2NH=CH 2 + is characterized by ab initio procedures. This reaction occurs through several asynchronous stages, but without passing through formal intermediates. A potential energy barrier to hydrogen migration from the β carbon to N is largely determined by the energy required to cleave the CN bond, but is lowered slightly by H transfer from the β to the α carbon and then to N. The complex [C2H 5 + NH=CH2] is bypassed, even though that complex could exist at energies only slightly above that of the transition state for ethene elimination. Furthermore, conversion of a substantial reverse activation energy into energy of motion causes CH2=NH 2 + and CH2=CH2 to dissociate faster than they can form [CH2=NH 2 + CH2=CH2]. Comparison of results for CH3CH2NH=CH 2 + to ab initio ones for methane from CH3CH2CH 3 + and elimination of ethene from CH3CH2O=CH 2 + and CH3CH2CH=OH+ reveals that these dissociations occur in a similar but, in each case, a distinct series of asynchronous steps or stages, and that there is no sharp demarcation between concerted and stepwise eliminations as presently defined. In dissociations of CH3CH2NH=CH 2 + , loss of electron density at the C in the breaking N bond leads the transfer of electron density to that carbon by migration of a hydrogen from the adjacent C. We attribute this to a requirement for the moving H to be close to Cα before the moving H can start to develop covalent bonding to Cα. It is also concluded that elimination of ethene from CH3CH2NH=CH 2 + avoids a Woodward-Hoffmann symmetry-imposed barrier by H migrating sufficiently from the β to the α carbon on the way to N, so that the dissociation is essentially a 1,1 rather than a 1,2 elimination.  相似文献   

11.
The mass spectra of the compounds Rh2 (RCOO)4[R=C(CH3)3 (I), CH(CH3)2 (II), CF3 (III)], Pd3(RCOO)6 [R=C(CH3)3 (IV), CH(CH3)2 (V), CF3 (VI)], Os2(RCOO)4Cl2 [R=C(CH3)3 (VII)], and Ru2(RCOO)4 [R=C(CH3)3 (VIII)] have been investigated. It has been shown that in the gaseous state compounds I–III, VII, and VIII have a dimeric structure, while compounds IV–VI have a trimeric structure. The mass spectra of compounds I–VIII show peaks of the molecular ions [M]+; and the fragmentation of the molecular ions takes place mainly with the elimination of the RCOO groups. Rearrangements with the elimination of F2O have been discovered for compounds III and VI, and rearrangements with the elimination of O=C--OH have been discovered for IV and V. The migration of a fluorine atom to the metal in compound III and its absence in compound V have been explained in the framework of the principle of hard and soft acids and bases. A scheme for the fragmentation of [M]+. under the action of electron impact has been proposed.Translated from Teoreticheskaya i Eksperimental'naya Khimiya, Vol. 22, No. 3, pp. 322–330, May–June, 1986.  相似文献   

12.
By combining results from a variety of mass spectrometric techniques (metastatle ion, collisional activation, collision-induced dissociative ionization, neutralization–reionization spectrometry and appearance energy measurements) and the classical method of isotopic labelling, a unified mechanism is proposed for the complex unimolecular chemistry of ionized 1,2-propanediol. The key intermediates involved are the stable hydrogen-bridged radical cations [CH2?C(H)? H…?O…?O(H)CH3]+˙, which were generated independently from [4-methoxy, 1-butanol]+˙ (loss of C2H4) and [1-methoxyglycerol]+˙ (loss of CH2O), [CH3? C?O…?H…?O(H)CH3]+˙ and the related ion-dipole complex [CH2?C(OH)CH3/H2O]+˙. The latter species serves as the precursor for the loss of CH3˙ and in this reaction the same non-ergodic behaviour is observed as in the loss of CH3˙ from the ionized enol of acetone.  相似文献   

13.
Different Mechanisms of the Cyclisation of Aminofluorosilanes The reaction of aminofluorosilanes of the type RR′SiFNHR″ (R = H, F, CH3, C2H3, C6H5, C(CH3)3; R′ = C(CH3)3, NiC3H7Si(CH3)3, NC(CH3)3Si(CH3)3, N[Si(CH3)3]2; R″ = iC3H7, C(CH3)3, C6H5) with butyllithium depends on the steric influence of the ligands. With increasing size of the ligands the reaction takes its pathway from the substitution under LiF elimination via dimerisation with additional elimination of butan to the C? H cleavage and cyclisation via a methylen group. A further increase of the size of the substituted groups leads through the intermediate formation of a silicenium-ylid to ring closure reactions. These occure by migration of a methanid ion leading to intermolecular nucleophilic substitution. The isolated acyclic and heterocyclic compounds are described and the mass and 1H-n.m.r. spectra are reported.  相似文献   

14.
Dissociative ionization of 1,2-epoxy n-alkanes gives rise to abundant [C4H7O]+ ions of structure [CH3OCHCHCH2]+. This conclusion is drawn from metastable ion analysis and from collisional activation spectra. This fragmentation involves the C? C ring opening and a 1,4-H migration leading to the corresponding enol ether [CH3OCHCHCH2R]+. precursor of [CH3OCHCHCH2]+ fragment. The same isomerization scheme applies to 1,2-epoxy methyl substituted alkanes and 2,3-epoxy n-alkanes.  相似文献   

15.
Cyclohexanone with the pMeOC6H4 and CH2?C(Me) substituents at the C3 and C4‐positions was prepared from (+)‐β‐pinene and converted to the allylic picolinate by a Masamune–Wittig reaction followed by reduction and esterification. Allylic substitution of this picolinate with Me2CuMgBr ? MgBr2 in the presence of ZnI2 proceeded with γ regio‐ and stereoselectively to afford the quaternary carbon center on the cyclohexane ring with the CH2?CH and Me groups in axial and equatorial positions, respectively. This product was converted to cyclobakuchiol A by demethylation and to cyclobakuchiol C by epoxidation of the CH2?C(Me) group. For the synthesis of cyclobakuchiol B, the enantiomer of the above cyclohexanone derived from (?)‐β‐pinene was converted to the cyclohexane‐carboxylate, and the derived enolate was subjected to the reaction with CH2?CHSOPh followed by sulfoxide elimination to afford the intermediate with the quaternary carbon center with MeOC(?O) and CH2?CH groups in axial and equatorial positions. The MeOC(?O) group was transformed to the Me group to complete the synthesis of cyclobakuchiol B.  相似文献   

16.
Loss of H2S is the characteristic Cys side‐chain fragmentation of the [M? H]? anions of Cys‐containing peptides. A combination of experiment and theory suggests that this reaction is initiated from the Cys enolate anion as follows: RNH‐?C(CH2SH)CONHR′ Ø [RNHC(?CH2)CONHR′ (HS?)] Ø [RNHC(?CH2)CO‐HNR′‐H]?+H2S. This process is facile. Calculations at the HF/6‐31G(d)//AM1 level of theory indicate that the initial anion needs only ≥20.1 kcal mol?1 of excess energy to effect loss of H2S. Loss of CH2S is a minor process, RNHCH(CH2SH)CON?‐R′ Ø RNHCH(CH2S?)CONHR′ Ø RNH ?CHCONHR+CH2S, requiring an excess energy of ≥50.2 kcal mol?1. When Cys occupies the C‐terminal end of a peptide, the major fragmentation from the [M–H]? species involves loss of (H2S+CO2). A deuterium‐labelling study suggests that this could either be a charge‐remote reaction (a process which occurs remote from and uninfluenced by the charged centre in the molecule), or an anionic reaction initiated from the C‐terminal CO2? group. These processes have barriers requiring the starting material to have an excess energy of ≥79.6 (charge‐remote) or ≥67.1 (anion‐directed) kcal mol?1, respectively, at the HF/6‐31G(d)//AM1 level of theory. The corresponding losses of CH2O and H2O from the [M? H]? anions of Ser‐containing peptides require ≥35.6 and ≥44.4 kcal mol?1 of excess energy (calculated at the AM1 level of theory), explaining why loss of CH2O is the characteristic side‐chain loss of Ser in the negative ion mode. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

17.
Using a relative kinetic technique, rate coefficients have been measured, at 296 ± 2 K and 740 Torr total pressure of synthetic air, for the gas‐phase reaction of OH radicals with the dibasic esters dimethyl succinate [CH3OC(O)CH2CH2C(O)OCH3], dimethyl glutarate [CH3OC(O)CH2CH2CH2C(O)OCH3], and dimethyl adipate [CH3OC(O)CH2CH2CH2CH2C(O)OCH3]. The rate coefficients obtained were (in units of cm3 molecule?1 s?1): dimethyl succinate (1.89 ± 0.26) × 10?12; dimethyl glutarate (2.13 ± 0.28) × 10?12; and dimethyl adipate (3.64 ± 0.66) × 10?12. Rate coefficients have been also measured for the reaction of chlorine atoms with the three dibasic esters; the rate coefficients obtained were (in units of cm3 molecule?1 s?1): dimethyl succinate (6.79 ± 0.93) × 10?12; dimethyl glutarate (1.90 ± 0.33) × 10?11; and dimethyl adipate (6.08 ± 0.86) × 10?11. Dibasic esters are industrial solvents, and their increased use will lead to their possible release into the atmosphere, where they may contribute to the formation of photochemical air pollution in urban and regional areas. Consequently, the products formed from the oxidation of dimethyl succinate have been investigated in a 405‐L Pyrex glass reactor using Cl‐atom–initiated oxidation as a surrogate for the OH radical. The products observed using in situ Fourier transform infrared (FT‐IR) absorption spectroscopy and their fractional molar yields were: succinic formic anhydride (0.341 ± 0.068), monomethyl succinate (0.447 ± 0.111), carbon monoxide (0.307 ± 0.061), dimethyl oxaloacetate (0.176 ± 0.044), and methoxy formylperoxynitrate (0.032–0.084). These products account for 82.4 ± 16.4% C of the total reaction products. Although there are large uncertainties in the quantification of monomethyl succinate and dimethyl oxaloacetate, the product study allows the elucidation of an oxidation mechanism for dimethyl succinate. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 431–439, 2001  相似文献   

18.
Ab initio molecular orbital calculations with split-valence plus polarization basis sets and incorporating valence-electron correlation have been performed to determine the equilibrium structure of ethyloxonium ([CH3CH2OH2]+) and examine its modes of unimolecular dissociation. An asymmetric structure (1) is predicted to be the most stable form of ethyloxonium, but a second conformational isomer of Cs symmetry lies only 1.4 kJ mol?1 higher in energy than 1. Four unimolecular decomposition pathways for 1 have been examined involving loss of H2, CH4, H2O or C2H4. The most stable fragmentation products, lying 65 kJ mol?1 above 1, are associated with the H2 elimination reaction. However, large barriers of 257 and 223 kJ mol?1 have to be surmounted for H2 and CH4 loss, respectively. On the other hand, elimination of either C2H4 or H2O from ethyloxonium can proceed without a barrier to the reverse associations and, with total endothermicities of 130 and 160 kJ mol?1, respectively, these reactions are expected to dominate at lower energies. A second important equilibrium structure on the surface is a hydrogen-bridged complex, lying 53 kJ mol?1 above 1. This complex is involved in the C2H4 elimination reaction, acts as an intermediate in the proton-transfer reaction connecting [C2H5]+ +H2O and C2H4 + [H3O]+ and plays an important role in the isotopic scrambling that has been observed experimentally in the elimination of either H2O or C2H4 from ethyloxonium. The proton affinity of ethanol was calculated as 799 kJ mol?1, in close agreement with the experimental value of 794 kJ mol?1.  相似文献   

19.
Unstable 2-hydroxpropene was prepared by retro-Diels-Alder decomposition of 5-exo-methyl-5-norbornenol at 800°C/2 × 10?6 Torr. The ionization energy of 2-hydroxypropene was measured as 8.67±0.05 eV. Formation of [C2H3O]+ and [CH3]+ ions originating from different parts of the parent ion was examined by means of 13C and deuterium labelling. Threshold-energy [H2C?C(OH)? CH3] ions decompose to CH3CO++CH3˙ with appearance energy AE(CH3CO+) = 11.03 ± 0.03 eV. Higher energy ions also form CH2?C?OH+ + CH3 with appearance energy AE(CH2?C?OH+) = 12.2–12.3 eV. The fragmentation competes with hydrogen migration between C(1) and C(3) in the parent ion. [C2H3O]+ ions containing the original methyl group and [CH3]+ ions incorporating the former methylene and the hydroxyl hydrogen atom are formed preferentially, compared with their corresponding counterparts. This behaviour is due to rate-determining isomerization [H2C?C(OH)? CH3] →[CH3COCH3], followed by asymmetrical fragmentation of the latter ions. Effects of internal energy and isotope substitution are discussed.  相似文献   

20.
The reactions of ionized di-n-butyl ether are reported and compared with those of ionized n-butyl sec-butyl and di-sec-butyl ether. The main fragmentation of metastable (CH3CH2CH2CH2)2O+. is C2H5? loss (?85%), but minor amounts (2–4%) of CH3?, C4H7?, C4H9?, C4H10 and C4H10O are also eliminated. In contrast, C2H5? elimination is of much lower abundance (20 and 4%, respectively) from metastable CH3CH2CH2CH2OCH(CH3)CH2CH3+. and [CH3CH2(CH3)CH]2O+., which expel mainly C2H6 and CH3? (35–55%). Studies on collisional activation spectra of the C6H13O+ oxonium ions reveal that C2H5? loss from (CH3CH2CH2CH2)2O+. gives the same product, (CH3CH2CH2CH2 +O?CHCH3) as that formed by direct cleavage of CH3CH2CH2CH2OCH(CH3)CH2CH3+.. Elimination of C2H5? from (CH3CH2CH2CH2)2O+. is interpreted by means of a mechanism in which a 1,4-H shift to the oxygen atom initiates a unidirectional skeletal rearrangement to CH3CH2CH2CH2OCH(CH3)CH2CH3+., which then undergoes cleavage to CH3CH2CH2CH2+O?CHCH3 and C2H5?. Further support for this mechanism is obtained from considering the collisional activation and neutralization-reionization mass spectra of the (C4H9)2O+. species and the behaviour of labelled analogues of (CH3CH2CH2CH2)2O+.. The rate of ethyl radical loss is suppressed relative to those of alternative dissociations by deuteriation at the γ-position of either or both butyl substituents. Moreover, C2H5? loss via skeletal rearrangement and fragmentation of the unlabelled butyl group in CH3CH2CH2CH2OCH2CH2CD2CH3+. occurs approximately five times more rapidly than C2H4D? expulsion via isomerization and fission of the labelled butyl substituent. These findings indicate that the initial 1,4-hydrogen shift is influenced by a significant isotope effect, as would be expected if this step is rate limiting in ethyl radical loss.  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