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 62 毫秒
1.
The mass spectrum of o-picolinotoluidide gives rise to three major fragments at m/z 184, m/z 169 and m/z 168, corresponding to the loss of CO from the molecular ion followed by the loss of ?H2 and ?H3 by independent pathways. It has been shown that the ortho methyl group and the nitrogen of the pyridine ring in the 2-position are involved in the formation of these three major fragments observed in the mass spectrum of o-picolinotoluidide. The mass spectrum of 2-(o-toluidino) pyridine, the molecular ion of which can resemble the [M? CO]+ ion in o-picolinotoluidide, also shows loss of CH3 and NH2 radicals from the molecular ions. Based on these observations coupled with the high resolution data, the mass analysed ion kinetic energy spectrometry and high voltage scans of these fragments in both the compounds, two mechanistic pathways have been proposed for the formation of these ions in o-picolinotoluidide.  相似文献   

2.
Abstract— This study was undertaken to further investigate the way in which the counter anion controls the Λmax of the absorption spectrum of compounds similar to N-retinylidene-n-butylammonium salts (NRBA). The following relationship had been found: ΔE =ΔEo -F d0e2/εd2; here ΔE is the observed excitation energy, e the charge on the electron, ε the dielectric constant, d0 a constant and d the distance between centers of opposite charge as estimated from crystallographic radii. Resonance theory implies that ΔEo should be of the same numerical value as the corresponding carbonium ion which can be generated readily from the corresponding alcohol. The C22SB analog of NRBA was prepared and then converted to the halide salts. The Δmax of these salts was determined in several halohydrocar-bon solvents, and ΔEo was determined by least squares for each solvent. The average value of ΔEo was found to be 653 nm, while the Λmax, for the carbonium ion was previously found to be 644 nm. The results are supportive of previous work.  相似文献   

3.
The reactions of metal carbonyl anions (M(CO)n?; M = Cr, Mn and Fe; n = 1–3) with n-heptane, water and methanol were studied with use of a Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometer equipped with an external ion source. The M(CO)n? ions were formed in the FT-ICR cell by collision-induced dissociation of the most abundant primary ion generated by electron impact of the appropriate metal carbonyl compound present in the external ion source. The M(CO)n? ions were allowed subsequently to undergo non-reactive collisions with argon in order to remove possible excess internal/translational energy prior to the ion/molecule reaction. Only the Cr(CO)3?, Mn(CO)3? and Fe(CO)2? ions react with n-heptane. This reaction proceeds by loss of H2 from the collision complex and the Cr(CO)3? and Fe(CO)2? ions react about three times more efficiently than the Mn(CO)3? ion. With water, Mn(CO)? and Fe(CO)3? are unreactive, whereas the other ions react by loss of one or two CO molecules from the collision complex. The rate of the reaction with water decreases in the order Cr(CO)3?, Fe(CO)2?, Cr(CO)2?, Fe(CO)?, Mn(CO)3? and Mn(CO)2?. With methanol, the Cr(CO)2? ion reacts by loss of two CO molecules from the collision complex, whereas loss of one CO molecule and elimination of CO + H2 occur in the reaction with Cr(CO)3?. Competing loss of CO and one or two H2 molecules occurs in the reactions of Mn(CO)3? and Fe(CO)2? with methanol. The rate of the reaction with methanol decreases in the order Cr(CO)3?, Fe(CO)2?, Cr(CO)2? and Mn(CO)3?.  相似文献   

4.
Hydrogen/deuterium exchange and rearrangements in the molecular ion of o-(methyl-d3-thio)benzoic acid lead to fragment ions [M? OD]+ as well as [M? OH]+ and m/z 106 and 107, just as in the molecular ion of o-methoxybenzoic acid. However, the fragment ion m/z 108 has the composition C6H4S rather than C6H2D3CO as it does in the case of o-methoxy-d3-benzoic acid. By varing the repeller potential at 10 eV (and thus the residence time in the ion source), the corresponding fragments are seen to be formed more slowly from the methylthio acid than from the methoxy acid, which leads to the conclusion that H/D exchange between carboxyl and labelled methylthio is slower than it is between carboxyl and labelled methoxyl.  相似文献   

5.
A standard procedure for recording and correcting collisionally activated dissociation mass spectra is proposed, and used to distinguish between the C4H5N ions formed from hydroxy- and amino-pyridines after loss of CO and HCN, respectively. It is concluded that these ions are cyclic. From the 4-isomers the 3H-pyrrole ion is formed whereas from 2-hydroxypyridine the 1H-pyrrole ion is formed. In the other cases, mixtures of 2H- and either 1H- or 3H-pyrrole ions are generated, depending on the nature of the precursor.  相似文献   

6.
Loss of CO from the molecular ions ([CH3OC6H4COF]+˙) of o-, m- and p-anisoyl fluorides has been investigated by mass-analysed ion kinetic energy (MIKE) spectrometry. This reaction involves fluorine atom migration from the carbonyl group to the benzene ring. In the cases of o- and p-anisoyl fluorides, the fluorine atom migrates via a three-membered transition state to form the molecular ions ([CH3OC6H4F]+˙) of o- and p-fluoroanisoles, respectively. On the other hand, in the case of m-anisoyl fluoride, the fluorine atom migrates from the carbonyl group to the benzene ring via a three- or four-membered transition state.  相似文献   

7.
Intrinsic ionic heats of transport q o * (ion) and ionic heats of transport Q o * (ion) have been evaluated for 53 aqueous ions at infinite dilution at 25°C using the reduction rule proposed by the authors and the limiting laws of Agar, and of Helfand and Kirkwood without electrophoretic terms. q o * (ion) have been found to correlate linearly with the standard ionic entropies of hydration for the 38 ions investigated. The correlation yields three distinctive proportionality constants indicating that the ions may be divided into three distinctive groups. Although the sign of Q o * (ion) is not definite, all values of q o * (ion) are positive. For 17 ions Q o * (ion) are in good agreement with TS o * (ion). Here, S o * (ion) is the absolute standard ionic entropy of transport which can be obtained from potentiometric measurements on cells. The values of S o * (ion) were determined by Agar, and recently by Lin and coworkers.  相似文献   

8.
Unimolecular hydroxyl (OD˙) loss from regio- and stereo-specifically labelled o-nitrostyrenes 1a, 1c and 1d results in the formation of an ion which upon collisional activation gives identical mass spectra. Suggestions are made which aim at explaining: (i) the loss of stereochemical integrity of the diastereotopic methylene hydrogens in the course of hydroxyl elimination; and (ii) to account for the collision induced losses of CO and HNC from the [M—hydroxyl]+ ion.  相似文献   

9.
[C8H6O]+˙ ions with o-quinonoid ketene, benzocyclobutenone, phenyl ketene and benzofuran structures have been generated from various precursors. Their collisionally induced decompositions in both field free regions of a double focusing mass spectrometer with so-called reversed geometry have been studied using mass analysed ion kinetic energy scans and B/E linked scans. In both cases the abovementioned [C8H6O]+˙ structures can be distinguished–except the benzocyclobutenone ion which gives very similar spectra to the o-quinonoid ion–on the basis of the intensity ratios [m/z 77]/[m/z 76] and [m/z 104]/[m/z 102]. The stable [C8H6O]+˙ ions generated from the molecular ions of 7 -phenylbicyclo[3.1.1]heptan-6-one appear to have the phenyl ketene structure, as was suspected from previous kinetic energy release measurements.  相似文献   

10.
Mass spectra and ion kinetic energy (IKE) spectra of o-, m- and p-d1 ethyl benzoates have given further information on the loss of OH˙ and OD˙ from the [M ? C2H4]+˙ ions. The ‘metastable peaks’ in the mass spectra give information on fragmentations in the field-free region following the electric sector; the IKE spectra give information on fragmentations in the field-free region preceding this sector. Transfer of hydrogen and deuterium from the ortho-positions on the ring to the carboxyl group can occur, but scrambling of ring hydrogens does not take place. A sample of o-d1 benzoic acid was also examined and confirmed that similar transfer reactions occur in this compound too.  相似文献   

11.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

12.
Anodic stripping voltammetry was made in AgNO3 solution, here Ag was deposited under long term potentiostatic conditions to evaluate the reduction charge, qr, and then was stripped by linear sweep voltammetry to determine the oxidation charge, qo. The charges were unbalanced, satisfying ca. qo=0.7|qr|, where other possible reduction charge such as by dioxygen and dichlorosilver were subtracted. The 30 % loss of the anodic charge can be ascribed to the negative capacitance by the potential sweep generation of Ag+. The generated Ag+ forms a dipole with a counter ion, of which orientation is the same as the direction of the externally applied electric field and opposite to the dipoles of solvent. The redox dipole decreases the conventional double layer capacitance caused by solvent dipoles, and high concentrations of Ag+ takes the capacitance to be negative values. The unbalanced charge, however, has no influence on quantitative determination of concentrations Ag+ by use of a calibration line.  相似文献   

13.
Fragmentation patterns resulting from electron impact ionization of 3-(2′-hydroxyethyl)quinolin-2(1H)-one, three of its monosubstituted derivatives and four of its disubstituted derivatives were studied. The molecular ion of quinolinone-2-etbanol undergoes initial fragmentation with the loss of OH·, H2O, CO, ·CHO, CH2O, CH2OH·, CH2?CHOH and HCNO species. The [M – CHO]+ ion is tentatively suggested to have been formed by the expulsion of H· from the [M – CO] ion and the [M - CHO]+ peak may be considered as diagnostic of a 2-quinolone-3-ethanol.  相似文献   

14.
It has been noticed that the major part of the loss of ?H from the molecular ion of most of the o-methoxythioamides results from an ortho effect of the methoxy group. Comparison of the MIKE spectra of the [M? SH]+ of 1-(2-methoxyphenylthioxomethyl)piperidine and 1-(2-methoxyphenylthioxomethyl)pyrrolidine with the MIKE spectra of [M? SH]+ of the corresponding unsubstituted compounds, reported earlier, indicated two parallel pathways for the formation of [M? SH]+ in the o-methoxy compounds. In the first pathway, as has been noticed in thioamides in general, the loss of ?H involves the migration of either the α-hydrogen in the amine moiety or the hydrogen attached to nitrogen. In the second pathway, the migration of a hydrogen from the o-methoxy group to the sulphur atom followed by ejection of SH from the molecular ion leads to a stable cyclized ion. Interesting secondary fragmentations as a consequence of this ortho effect have also been noticed.  相似文献   

15.
The ortho, meta, and para isomers of hydroxybenzyl alcohol can be unequivocally distinguished by the collision-induced dissociation mass spectra of their anions. The presence of a prominent peak at m/z 121 for an elimination of a dihydrogen molecule renders the ortho-isomer spectrum markedly different from those of its meta and para congeners. Investigations carried out with deuterium-labeled isotopologues of the ortho isomer verified that the labile hydrogen atom on the hydroxyl group and one of the benzylic hydrogen atoms are specifically removed in the formation of the m/z 121 ion. The ortho-isomer spectrum also showed a prominent peak at m/z 93. Experimental data indicated that the m/z 93 product ion originates either from a two-step H2and CO elimination mechanism or from a direct loss of a HCHO molecule from the precursor anion. The intensity ratio of the m/z 93 and 94 peaks in the spectrum recorded from the m/z 124 ion generated from a sample of o-hydroxybenzyl alcohol dissolved in D2O supported the notion that the direct HCHO loss is the more dominant pathway for the generation of the phenolate ion under low activation conditions. In contrast, the two-step mechanism becomes the more dominant pathway under high collisional activation conditions. The spectrum also showed a weak peak at m/z 105 for a water loss. Based on computational data, the m/z 105 ion generated in this way appears to be a composite generated from a common ion-neutral complex intermediate in which a hydroxyl anion is positioned equidistantly between one of the benzylic hydrogens and a nearby hydrogen atom of the benzene ring. Upon activation, the complex dissociates to form either a phenide or a quinone methide anion. The reaction forming a carbon dioxide adduct under ion-mobility conditions was used to support the proposed water-loss mechanism.  相似文献   

16.
The principal feature of the mass spectra of o-nitroanils, ArCH?NC6H4NO2(o-), is an intense peak corresponding to the [ArCO]+ ion; this implies oxygen transfer from the nitro group to the azomethine carbon during the fragmentation process. In this series of anils, loss of OH from the molecular ion is not apparently an important fragmentation pathway, in contrast to the fragmentation of o-nitrobenzylideneanilines. Benzylideneaniline derivatives with an o-nitro substituent in both rings have mass spectra which indicate interaction of both nitro groups with the ? CH?N? group, but in this series of spectra the [M—17]+ ion is again of low intensity.  相似文献   

17.
Tertiary α-carbomethoxy-α,α-dimethyl-methyl cations a have been generated by electron impact induced fragmentation from the appropriately α-substituted methyl isobutyrates 1–4. The destabilized carbenium ions a can be distinguished from their more stable isomers protonated methyl methacrylate c and protonated methyl crotonate d by MIKE and CA spectra. The loss of I and Br˙ from the molecular ions of 1 and 2, respectively, predominantly gives rise to the destabilized ions a, whereas loss of Cl˙ from [3]+ ˙ results in a mixture of ions a and c. The loss of CH3˙ from [4]+˙ favours skeletal rearrangement leading to ions d. The characteristic reactions of the destabilized ions a are the loss of CO and elimination of methanol. The loss of CO is associated by a very large KER and non-statistical kinetic energy release (T50 = 920 meV). Specific deuterium labelling experiments indicate that the α-carbomethoxy-α,α-dimethyl-methyl cations a rearrange via a 1,4-H shift into the carbonyl protonated methyl methacrylate c and eventually into the alkyl-O protonated methyl methacrylate before the loss of methanol. The hydrogen rearrangements exhibit a deuterium isotope effect indicating substantial energy barriers between the [C5H9O2]+ isomers. Thus the destabilized carbenium ion a exists as a kinetically stable species within a potential energy well.  相似文献   

18.
[o-(Trimethylgermyl)phenyl]acetylene was polymerized in the presence of WCl6, W(CO)6-hv, etc., to give polymers whose weight-average molecular weights reached ca. 7.0 X 105 at the highest. When the MoOCl4-n-Bu4Sn-EtOH (1 : 1 : 1) catalyst was used, the polydispersity ratio of the polymer obtained was 1.08, and the number-average molecular weight increased in direct proportion to monomer conversion; these indicate that this polymerization is a living polymerization. The polymer had the structure ? [CH?C(C6H4-o-GeMe3)]n ? and was a dark purple solid (λmax = 551 nm, εmax = 6100 M-1 cm-1 in THF) soluble in organic solvents such as toluene and chloroform. The onset temperature of weight loss of the polymer in TGA in air was ca. 230°C, and the glass transition temperature was above 180°C. The Po2 of the present polymer is 105 barrers—larger than the value of natural rubber and fairly close to that of poly(dimethylsiloxane). © 1993 John Wiley & Sons, Inc.  相似文献   

19.
Summary Solvolysis of the macrocyclic trithioether complexes fac-[Mo(CO)3([9]aneS3)] ([9]aneS3 = 1,4,7-trithiacyclononane) and fac-[Mo(CO)3(ttob)] (ttob = 2,5,8-trithia[9]-o-benzenophane), and the acyclic trithioether complex fac-[Mo(CO)3(ttn)] (ttn = 2,5,8-trithiononane) in dimethylsulfoxide (DMSO) solution gives the free trithioether in each case. The kinetics of these reactions were studied by 1H-n.m.r. spectroscopy. In the absence of acid or base, the rate of solvolysis of [Mo(CO)3([9]aneS3)] is at least 4.6 × 105 slower than that of [Mo(CO)3(ttn)]; larger macrocycles give intermediate rates. [Mo(CO)3-([9]aneS3) and [Mo(CO)3(ttob)] undergo stoichiometric reactions with acid and base in DMSO which in all cases lead to more rapid loss of the intact macrocycle than in DMSO alone. The results are discussed in terms of the structure and bonding of likely intermediate complexes.  相似文献   

20.
Equilibria concerning picrates of tetraalkylammonium ions (Me4N+, Et4N+, Pr4N+, Bu4N+, Bu3MeN+) in a dichloromethane−water system have been investigated at 25 C. The 1:1 ion-pair formation constants (K IP,o o) in dichloromethane at infinite dilution were conductometrically determined. The distribution constants (K D o) of the ion pairs and the free cations between the solvents were determined by a batch-extraction method. The K IP,o o value varies in the cation sequence, Bu4N+ ≈ Pr4N+ ≈ Et4N+ < Bu3MeN+ < < Me4N+; this trend is explained by the electrostatic cation−anion interaction taking into account the structures of the ion pairs determined by density functional theory calculations. For the ion pairs of the symmetric R4N+ cations, there is a linear positive relationship between log10 K D o and the number of methylene groups in the cation (N CH 2). The ion pair of asymmetric Bu3MeN+ has a higher distribution constant than that expected from the above log10 K D o versus N CH 2 relationship. These cation dependencies of log10 K D o for the ion pairs are explained theoretically by using the Hildebrand-Scatchard equation. For all the cations, the log10 K D o value of the free cation increases linearly with N CH 2; the variation of log10 K D o is discussed by decomposing the distribution constant into the Born-type electrostatic contribution and the non-Born one, and attributed to the latter that is governed by the differences in the molar volumes of the cations. The cation dependencies of the ion-pair extractability and ion pairing in water are also discussed. An erratum to this article can be found at  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