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 31 毫秒
1.
The translational energy releases accompanying fromation of H˙ and [CHO]+ from [CHOH], H˙ and [CDO]+ from [CDOH], D˙ and [CHO]+ from [CHOD] and D˙ and [CDO]+ from [CDOD] have been calculated from the widths at half-height of the respective metastable peaks. The energy releases for [CHOD] and [CDOD] are similar to each other in magnitude, but are both significantly larger than those for [CHOH] and [CDOH]. These differences are discussed in terms of the dynamical theory of energy partition. The larger energy releases for decompositions giving D˙ are attributed to larger proportions of the kinetic energy being associated with the departing D˙ in the transition states, as compared to H˙ in the transition states for the decompositions yielding H˙.  相似文献   

2.
Utilization of sodium [1-14C]-, [2–14C]-, and [1,2-13C]-acetates, [1-14C]-, [1-13C]-, or [2-14C]-propionates, [1-14C]-or [2-14C]-malonates, of [1-14C]- or of [1-14C]-myristic acid, or of [1-14C]- and [1-14C]-palmitic acid in the biosynthesis of cytochalasin D ( 1 ) by Zygosporium masonii was determined by degradation studies or by carbon magnetic resonance spectroscopy. The precursors were incorporated primarily via the acetate-malonate pathway to generate 1 from nine intact acetate units, eight of which are coupled in a head to tail fashion to form the C16-polyketide moiety.  相似文献   

3.
Coupling of uronium salts with in situ generated N‐heterocyclic carbenes provides straightforward access to symmetrical [ 4 ]2+ and unsymmetrical bis‐imidazolium salts [ 6 ]2+ and [ 9 ]2+. As indicated by cyclic and square‐wave voltammetry, [ 6 ]2+ and [ 9 ]2+ can be (irreversibly) reduced by one electron. The initially formed radicals [ 6 ].+ and [ 9 ].+ undergo further reactions, which were probed by EPR spectroscopy and density functional calculations. The final products of the two‐electron reduction are the two carbenes. Upon irradiation with UV light both [ 6 ]2+ and [ 9 ]2+ emit at room temperature in solution but with dramatically different characteristics. The different fluorescence behavior is analyzed by emission spectroscopy and interpreted by using time‐dependent density functional calculations as largely due to different excited‐state dynamics of [ 6 ]2+ and [ 9 ]2+. The geometries of both radicals [ 6 ].+ and [ 9 ].+ and excited states {[ 6 ]2+} * and {[ 9 ]2+}* are substantially different from those of the parent ground‐state molecules.  相似文献   

4.
The CI mass spectra of aryl ketones, πCOR, were studied and found to give primarily [M + 29]+, [M + 1]+, [M ? 1]+, [πCO]+ and [RCO]+ ions. The major change in the spectra with increasing length of the aliphatic side chain was an increase in the [M ? 1]+/[M + 1]+ ratio. Increasing sample size was reflected primarily in the formation of [2M + 1]+ ions and a decrease in [M + 1]+ ions. Small amounts of water in the reactant gas reduced the extent of fragmentation action.  相似文献   

5.
The gas-phase H/D exchange of bradykinin [M + H]+, [M + Na]+, [M + 2H]2+, and [M + H + Na]2+ ions; des-Arg1-bradykinin, des-Arg9-bradykinin, and bradykinin fragment 2-7 [M + H]+ ions; and O-methylbradykinin [M + H]+ and [M + 2H]2+ ions with D2O have been examined by electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry at 9.4 T. The different peptides vary widely in exchange rate and extent of deuterium incorporation. H/D exchange was slowest and deuterium incorporation was least for bradykinin [M + H]+, [M + H + Na]2+ and bradykinin methyl ester [M + 2H]2+ ions. In contrast, H/D exchange and extent of deuteration are higher for des-Arg1-bradykinin, des-Arg9-bradykinin, and bradykinin fragment 2-7 [M + H]+ ions; and highest for bradykinin [M + Na]+ and [M + 2H]2+, and O-methylbradykinin [M + H]+. Because the most likely site of protonation is the guanidino group of arginine, the above reactivity pattern strongly supports a zwitterion form for protonated gas-phase bradykinin.  相似文献   

6.
A computational study is performed to identify the origin of the room‐temperature stability, in aqueous solution, of functionalized π‐[R‐TTF]22+ dimers (TTF=tetrathiafulvalene; R=(CH2OCH2)5CH2OH) included in the cavity of a cucurbit[8]uril (CB[8]) molecule. π‐[R‐TTF]22+ dimers in pure water are weakly stable, and are mostly dissociated at room temperature. Upon addition of CB[8] to an aqueous π‐[R‐TTF]22+ solution, a (π‐[R‐TTF]2?CB[8])2+ inclusion complex is formed. The same complex is obtained after the sequential inclusion of two [R‐TTF].+ monomers in the CB[8] molecule. Both processes are thermodynamically and kinetically allowed. π‐[R‐TTF]22+ dimers dissolved in pure water present a [TTF].+???[TTF].+ long, multicenter bond, similar to that already identified in π‐[TTF]22+ dimers dissolved in organic solvents. Upon their inclusion in CB[8], the strength and other features of the [TTF].+???[TTF].+ long, multicenter bond are preserved. The room temperature stability of the π‐[R‐TTF]22+ dimers included in CB[8] is shown to originate in the π‐[R‐TTF]22+???CB[8] interaction, the strength of which comes from a strongly attractive electrostatic component and a dispersion component. Such a dominant electrostatic term is caused by the strongly polarized charge distribution in CB[8], the geometrical complementarity of the π‐[R‐TTF]22+ and CB[8] geometries, and the amplifying effect of the 2+ charge in π‐[R‐TTF]22+.  相似文献   

7.
Carbon-11-labeled nimesulide analogs, N-[11C]methyl-N-(2-benzyloxy-4-nitrophenyl)methanesulfonamide ([11C]4a), N-[11C]methyl-N-[2-(4′-methylbenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]4b), N-[11C]methyl-N-[2-(4′-fluorobenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]4c), N-[11C]methyl-N-[2-(4′-nitrobenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]8a), N-[11C]methyl-N-[2-(β-naphthylmethoxy)-4-nitrophenyl]methanesulfonamide ([11C]8b), and N-[11C]methyl-N-[2-(2′-phenylbenzyloxy)-4-nitrophenyl]methanesulfonamide ([11C]8c), have been synthesized as new potential positron emission tomography (PET) selective aromatase expression regulator (SAER) radiotracers for imaging of aromatase expression in breast cancer. The target tracers were prepared by N-[11C]methylation of their corresponding precursors using [11C]CH3OTf under basic conditions (NaH) and isolated by reversed-phase high-pressure liquid chromatography (HPLC) method in 30–50% radiochemical yields decay corrected to end of bombardment (EOB) with 25–30 min overall synthesis time and 111–148 GBq/μmol specific activity at end of synthesis (EOS).  相似文献   

8.
Mass spectra of isotope-labeled triptycenes, triphenylmethanes and diphenylmethanes rule out the bulk of postulated decomposition mechanisms and fragment-ion structures. The formation of [M ? H]+ and [M ? 2H]2+ from triptycene, of [M ? H]+, [M ? CH3]+ and [M ? CH4]+ from triphenylmethane, and of [M ? H]2+ and [M ? 2H]2+-as well as the previously reported [M ? H]+ and [M ? CH3]+-from diphenylmethane all seem to be preceded or accompanied by complete loss of position identity of the α and ring hydrogens in the original molecules. A statistical preference for loss of α hydrogens is found in the process leading to [M ? 2H]+ and [M ? H]2+ from triptycene, as in the formation of [M ? H]2+ from toluene.  相似文献   

9.
The mass spectra of 30 sulfinamide derivatives (RSONHR', R' alkyl or p-XC6H4) are reported. Most of the spectra had peaks attributable to thermal decomposition products. For some compounds these were identified by pyrolysis under similar conditions to be: RSO2NHR', RSO2SR, RSSR and NH2R' (in all kinds of sulfinyl amides); RSNHR' (in the case of arylsulfinyl arylamides); RSO2C6H4NH2, RSOC6H4NH2 and RSC6H4NH2 (in the case of arylsulfinyl arylamides of the type of X = H) The mass spectra of the three thermally stable compounds showed that there are several kinds of common fragment ions. The mass spectra of the thermally labile compounds had two groups of ions; (i) characteristic fragment ions of the intact molecules and (ii) the molecular ions of the thermal decomposition products. It was concluded that the sulfinamides give the following ions after electron impact: [M]+, [M ? R]+, [M ? R + H]+, [M ? SO]+, [RS]+, [NHR']+, [NHR' + H]+, [RSO]+, [RSO + H]+, [R]+, [R + H]+, [R']+ and [M ? OH]+, and that the thermal decomposition products give the following ions: [RSO2SR]+, [RSSR]+, [M ? O]+, [M + O]+ and [RSOC6H4NH2]+.  相似文献   

10.
Collision-induced dissociation of the ions [ArS]?, [ArSO]? and [ArSO2]? has uncovered a rich and varied ion chemistry. The major fragmentations of [ArS]? are complex and occur without prior ring hydrogen scrambling: for example, [C6H5S]?→[C2HS]? and [HS]?; [p-CD3C6H4S]?→[C6H4S]?˙, [CD3C4S]? and [C2HS]?. In contrast, all decompositions of [C6H5CH2S]? are preceded by specific benzylic and phenyl hydrogen interchange reactions. [ArSO2]? and [ArSO2]? ions undergo rearrangement, e.g. [C6H5SO]?→[C6H5O]? and [C6H5S]?; [C6H5SO2]?→[C6H5O] ?. The ion [C6H5CH2SO]? eliminates water, this decomposition is preceded by benzylic and phenyl hydrogen exchange.  相似文献   

11.
(1) Thiocarboxylate anions [RCOS]? formed by dissociative secondary electron capture are either stable or fragment to yield [R]?. (2) Thiocarboxylate cations [RCOS]+ formed by charge stripping from [RCOS]?, fragment to form [R]+, [COS]+. and [RCO]+ (not when R=CF3). (3) Aryl hydrogen scrambling is observed in the case of the thiobenzoate cation. Aliphatic hydrogen scrambling is not detected for the thiopropionate cation.  相似文献   

12.
The thermal gas‐phase catalytic reduction of N2O by CO, mediated by the transition‐metal nitride cluster ion [NbN]+, has been explored by using FT‐ICR mass spectrometry and complemented by high‐level quantum chemical calculations. In contrast to the [Nb]+/[NbO]+ and [NbO]+/[Nb(O)2]+ systems, in which the oxidation of [Nb]+ and [NbO]+ with N2O is facile, but in which neither [NbO]+ nor [Nb(O)2]+ react with CO at room temperature, the [NbN]+/[ONbN]+ system at ambient temperature mediates the catalytic oxidation of CO. The origins of the distinctly different reactivities upon nitrogen ligation are addressed by quantum chemical calculations.  相似文献   

13.
Abstract— Carbon-14-labeled hematoporphyrin ([14C]HP) was synthesized by two methods, (i) Using an in vitro avian whole-blood system, [14C]protoheme was obtained biosynthetically by incorporating [4C]aminolevulinic acid into the porphyrin ring structure. Subsequently, the [14C]protoheme was converted to [4C]HP by standard procedures, (ii) By adopting several well-characterized chemical reactions, deuteroporphyrin was treated with [14C]acetyl chloride, giving [14C]diacetyl deuteroporphy-rin which was readily reduced and hydrolyzed to [14C]HP (with thecarbon–14 label on the hydroxyethyl side-chains). These two methods are simple and afford good yields of [14C]HP with moderate to high specific activities. The [14C]HP was then treated with acetic acid/sulfuric acid followed by sodium hydroxide to give [14C]HPD. Upon gel- and ultra-filtration, the [14C]HPD was enriched in the so-called tumor-localizing fraction of HPD, giving [14C]PII with specific activities of 0.4 Ci/mol (biosynthesis) and 10 Ci/mol (chemical synthesis). These [14C]PII preparations were equivalent with respect to chromatographic and spectrophotometric characteristics, as well as tumoricidal photodynamic activity in the DBA/2 Ha-DD mouse: SMT-F tumor system, to the unlabeled commercial product Photofrin? II. The distribution of [14C]PII in mouse tissues was in close agreement to that previously reported, after adjustment for dose, for [14C]HPD biosynthetically labeled in vivo (Gomer and Dougherty, 1979), as well as for Photofrin? II, where tissue levels were determined spectrophotometrically after extraction (Dougherty and Mang, unpublished).  相似文献   

14.
Secondary ion mass spectra obtained by [Xe]+ bombardment are compared with those obtained by [Ar]+ bombardment. Although [Ar]+ ions are commonly used as primary ions in secondary ion mass spectrometry for organic compounds, [Xe]+ ions seem better as primary ions because they give a larger sputtering yield for a metal substrate than [Ar]+ ions. Cationized molecular intensities of sucrose, raffinose and stachyose, and quasimolecular ion intensities of tuftsin and eledoisin related peptide are investigated using [Xe]+ and [Ar]+ bombardments. The observed molecular species are 2–4 times more intense for [Xe]+ bombardment than for [Ar]+ bombardment, although the secondary ion mass spectra are almost the same in both cases.  相似文献   

15.
The correlation with substituent constants reported previously for [YØCO]+/[ØCO]+ ratios in the electron ionization mass spectra of substituted benzophenones and acetophenones has also been observed in the electron ionization spectra of substituted benzils. The [YØCO]+/[ØCO]+ ratio for the substituted benzils varied with energy of the ionizing electrons according to predictions from a simple kinetic and thermochemical analysis. [YØCO]+/[ØCO]+ ratios in the charge exchange spectra of benzophenones obtained with Xe, Kr, CO, N2 and Ar gave good correlations with sub-stituent constants in agreement with the same analysis. Good correlations were also noted for [YØCO]+/[ØCO]+ ratios with substituent constants for [M]+ ions of the benzophenones of the same excess energy (5.5 eV). [YØCO]+/[ØCO]+ ratios for benzils obtained by charge exchange with [CO]+ also showed good correlations with substituent constant. It is suggested that [Ø]+ and [YØ]+ ions from the benzophenones may be formed primarily by one step decompositions of the molecular ions, but that the [Ø]+ and [YØ]+ ions from the benzils are formed primarily by decomposition of [ØCO]+ and [YØCO]+ ions.  相似文献   

16.
Centrosymmetric linear [Ar-H-Ar]+ and asymmetric linear [Ar---Ar-H]+ are two stable configurations of [Ar2H]+. Based on the global potential energy surface of [Ar2H]+ provided by our group recently, we calculated the vibrational spectra of [Ar---Ar-H]+ with total angular momentum J = 0 by time-dependent quantum mechanical method, and the influence of quantum tunneling effect on vibrational spectra was found. With the help of the observation on the eigenstate functions and the modified potential energy surface, assignments were made to the spectra. The strong coupling between the excited bending mode of [Ar-H-Ar]+ and the vibrational states of [Ar---Ar-H]+ was discussed.  相似文献   

17.
A stable cyclic (alkyl)(amino)carbene (CAAC) 1 inserts into the para‐CF bond of pentafluoropyridine, and after fluoride abstraction, the iminium‐pyridyl adduct [ 3 ]+ was isolated. A cyclic voltammetry study shows a reversible three‐state redox system involving [ 3 ]+, [ 3 ] ? , and [ 3 ] ? . The CAAC‐pyridyl radical [ 3 ] ? , obtained by reduction of [ 3 ]+ with magnesium, has been spectroscopically and crystallographically characterized. In contrast to the lack of π communication between the CAAC and the pyridine units in cation [ 3 ]+, the unpaired electron of [ 3 ] ? is delocalized over an extended π system involving both heterocycles.  相似文献   

18.
The isomeric ions [H2NC(H)O]+˙, [H2NCOH]+˙, [H3CNO]+˙ and [H2CNOH]+˙ were examined in the gas phase by mass spectrometry. Ab initio molecular orbital theory was used to calculate the relative stabilities of [H2NC(H)O]+˙, [H2NCOH]+˙, [H3NCO]+˙ and their neutral counterparts. Theory predicted [H2NC(H)O]+˙ to be the most stable ion. [H2NCOH]+˙ ions were generated via a 1,4-hydrogen transfer in [H2NC(O)OCH3]+˙, [H2NC(O)C(O)OH]+˙ and [H2NC(O)CH2CH3]+˙. Its metastable dissociation takes place via [H3NCO]+˙ with the isomerization as the rate-determining step. [H2CNOH]+˙ undergoes a rate-determining isomerization into [H3CNO]+˙ prior to metastable fragmentation. Neutralization-reionization mass spectrometry was used to identify the neutral counterparts of these [H3,C,N,O]+˙ ions as stable species in the gas phase. The ion [H3NCO]+˙ was not independently generated in these experiments; its neutral counterpart was predicted by theory to be only weakly bound.  相似文献   

19.
Metal complex formation was investigated for di‐exo‐, di‐endo‐ and trans‐2,3‐ and 2,5‐disubstituted trinorbornanediols, and di‐exo‐ and di‐endo‐ 2,3‐disubstituted camphanediols using different divalent transition metals (Co2+, Ni2+, Cu2+) and electrospray ionization quadrupole ion trap mass spectrometry. Many metal‐coordinated complex ions were formed for cobalt and nickel: [2M+Met]2+, [3M+Met]2+, [M–H+Met]+, [2M–H+Met]+, [M+MetX]+, [2M+MetX]+ and [3M–H+Co]+, where M is the diol, Met is the metal used and X is the counter ion (acetate, chloride, nitrate). Copper showed the weakest formation of metal complexes with di‐exo‐2,3‐disubstituted trinorbornanediol yielding only the minor singly charged ions [M–H+Cu]+, [2M–H+Cu]+ and [2M+CuX]+. No clear differences were noted for cobalt complex formation, especially for cis‐2,3‐disubstituted isomers. However, 2,5‐disubstituted trinorbornanediols showed moderate diastereomeric differentiation because of the unidentate nature of the sterically more hindered exo‐isomer. trans‐Isomers gave rise to abundant [3M–H+Co]+ ion products, which may be considered a characteristic ion for bicyclo[221]heptane trans‐2,3‐ and trans‐2,5‐diols. To differentiate cis‐2,3‐isomers, the collision‐induced dissociation (CID) products for [3M+Co]2+, [M+CoOAc]+, [2M–H+Co]+ and [2M+CoOAc]+ cobalt complexes were investigated. The results of the CID of the monomeric and dimeric metal adduct complexes [M+CoOAc]+ and [2M–H+Co]+ were stereochemically controlled and could be used for stereochemical differentiation of the compounds investigated. In addition, the structures and relative energies of some complex ions were studied using hybrid density functional theory calculations. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
Treatment of [Ir(bpa)(cod)]+ complex [ 1 ]+ with a strong base (e.g., tBuO?) led to unexpected double deprotonation to form the anionic [Ir(bpa?2H)(cod)]? species [ 3 ]?, via the mono‐deprotonated neutral amido complex [Ir(bpa?H)(cod)] as an isolable intermediate. A certain degree of aromaticity of the obtained metal–chelate ring may explain the favourable double deprotonation. The rhodium analogue [ 4 ]? was prepared in situ. The new species [M(bpa?2H)(cod)]? (M=Rh, Ir) are best described as two‐electron reduced analogues of the cationic imine complexes [MI(cod)(Py‐CH2‐N?CH‐Py)]+. One‐electron oxidation of [ 3 ]? and [ 4 ]? produced the ligand radical complexes [ 3 ]. and [ 4 ].. Oxygenation of [ 3 ]? with O2 gave the neutral carboxamido complex [Ir(cod)(py‐CH2N‐CO‐py)] via the ligand radical complex [ 3 ]. as a detectable intermediate.  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