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.
Negative ion CID spectra of N‐linked glycans released from glycoproteins contain many ions that are diagnostic for specific structural features such as the detailed arrangement of antennae and the location of fucose residues. Identification of such ions requires reference glycans that are often difficult to acquire in a pure state. The recent acquisition of a sample of N‐glycans from a patient lacking the enzyme N‐acetylglucosaminyltransferase‐2 provided an opportunity to investigate fragmentation of glycans lacking a 6‐antenna. These glycans contained one or two galactose‐N‐acetylglucosamine‐chains attached to the 3‐linked mannose residue of the trimannosyl‐chitobiose core with and without fucose substitution. The spectra from the patient sample clearly defined the antenna distribution and showed striking differences from the spectra of isomeric compounds obtained from normal subjects. Furthermore, they provided additional information on previously identified antenna‐specific fragment ions and indicated the presence of additional ions that were diagnostic of fucose substitution. Glycans obtained from such enzyme‐deficient patients can, thus, be a valuable way of obtaining spectra of specific isomers in a relatively pure state for interpretation of mass spectra. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

2.
In the course of systematic studies on surface layer (S-layer) glycoproteins of bacilli, the chloroform/methanol extract from whole cells of Geobacillus stearothermophilus NRS 2004/3a has been submitted to MS analysis. Glucosylated cardiolipins were found as minor components of the total lipid and phospholipid mixture by de novo identification. After purification of the crude extract using a combined column chromatography/2D TLC protocol, structural investigations of components in the lipid fraction by high resolution ESI-QTOF MS analysis provided evidence about homologous molecules attributable to the cardiolipin species containing a glycosylated backbone, and about a diversity of ester-linked fatty acid substituents. In comparative studies by positive and negative ion nanoESI-QTOF-CID-MS, maps of cardiolipin molecular ions were obtained, followed by MS/MS of the most abundant species, to provide structural details of D-glucopyranosylcardiolipin and the fatty acid substituent patterns. Experiments of the parent ion scan type revealed the presence of fatty acid moieties as isobaric combinations, represented in single molecular ion species.  相似文献   

3.
Electrospray‐generated precursor ions usually follow the ‘even‐electron rule’ and yield ‘closed shell’ fragment ions. We characterize an exception to the ‘even‐electron rule.’ In negative ion electrospray mass spectrometry (ES‐MS), 2‐(ethoxymethoxy)‐3‐hydroxyphenol (2‐hydroxyl protected pyrogallol) easily formed a deprotonated molecular ion (M‐H)? at m/z 183. Upon low‐energy collision induced decomposition (CID), the m/z 183 precursor yielded a radical ion at m/z 124 as the base peak. The radical anion at m/z 124 was still the major fragment at all tested collision energies between 0 and 50 eV (Elab). Supported by computational studies, the appearance of the radical anion at m/z 124 as the major product ion can be attributed to the combination of a low reverse activation barrier and resonance stabilization of the product ions. Furthermore, our data lead to the proposal of a novel alternative radical formation pathway in the protection group removal of pyrogallol. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
The use of a novel electrophoric derivatisation reagent, o-(pentafluorobenzyloxycarbonyl)-benzoyl chloride is described for the quantitative determination of amphetamine in plasma. Amphetamine can be quantitatively measured down to 49 pg/mL plasma using only 250 μL of sample due to the extraordinary sensitivity of the derivatives under negative ion chemical ionisation mass spectrometry. Plasma samples were made alkaline with carbonate buffer and treated with n-hexane and reagent solution for 20 min, which, after concentration was measured by negative ion chemical ionisation gas chromatography-mass spectrometry. The method is rapid as extraction and derivatisation occur in one single step. [(2) H(5) ]-Amphetamine was used as an internal standard. Validation data are given to demonstrate the usefulness of the assay, including specificity, linearity, accuracy and precision, benchtop stability, freeze-thaw stability, autosampler stability, aliquot analysis and prospective analytical batch size accuracy.  相似文献   

5.
Positive and negative ion electrospray ionization (ESI) tandem mass spectral study of a new series of hybrid peptides, viz, BocN-alpha,beta-peptides and BocN-beta,alpha-peptides, synthesized from C-linked carbo-beta3-amino acids [Caa (S)] and L-Ala has been carried out. The alpha,beta-peptides have been differentiated from beta,alpha-peptides by the collision-induced dissociation (CID) of [M + H]+ and [M - H]- ions in positive and negative ion ESI-MS respectively. The fragment ion [M + H - C(CH3)3 + H]+ formed from [M + H]+ ions by the loss of 2-methyl-prop-2-ene in alpha,beta-peptides with L-Ala at the N-terminus is insignificant or totally absent for beta,alpha-peptides which have the Caa (S) at N-terminus. The fragment ion [M - H-C(CH3)3OH - HNCO]- formed from [M - H]- of beta,alpha-peptide acids is totally absent for alpha,beta-peptide acids. This has been attributed to the absence of the beta-methylene group in alpha,beta-peptides, and the participation of the beta-methylene group in the loss of HNCO in beta,alpha-peptide acids is confirmed by the deuteration experiments. The CID of [M + H-Boc + H]+ ions of these peptides also produce characteristic fragmentation. In the CID spectra of alpha,beta-peptides, the b(n)+ ions and the resulting y(n)+ ions occur at a mass difference of 243 and 71 Da corresponding to the successive losses of Caa and L-Ala, whereas a mass difference of 71 and 243 Da is observed for beta,alpha-peptides. In contrast to the CID of protonated peptides, the CID of [M - H]- ions of the alpha,beta- and beta,alpha-peptide acids do not give b(n)- ions and show abundant z(n) (-) ions. Further, a pair of diastereomeric dipeptide esters and acids have been distinguished by the CID of [M + H]+ ions. The loss of 2-methyl-prop-2-ene is more pronounced for Boc-NH-Caa(R)-D-Ala-OCH3 (21) and Boc-NH-Caa(R)-D-Ala-OH (23) with Caa (R) at the N-terminus, whereas it is totally absent for Boc-NH-Caa (S)-D-Ala-OCH3 (22) and Boc-NH-Caa(S)-D-Ala-OH (24) peptides, which have Caa (S) at the N-terminus. Thus, on the basis of our previous and present studies, we propose that the CID of [M + H]+ ions provides a simple and useful method for distinguishing the configuration of Caa (S or R) at the N-terminus of BocN-carbo beta,alpha- and beta,beta-dipeptides.  相似文献   

6.
A series of meso-dialkyl, alkyl aryl and cycloalkyl calix(4)pyrroles (1-15) are studied under positive and negative ion electrospray ionization (ESI) conditions. The positive ion spectra show abundant [M + H](+) and [M + Na](+) ions and the negative ion spectra show the [M + Cl](-) (the Cl(-) ions from the solvent) and [M - H](-) ions. The collision induced dissociation (CID) spectra of [M + H](+), [M + Na](+), [M + Cl](-) and [M - H](-) ions are studied to understand their dissociation pathway and compared to that reported for M(+) under electron ionization (EI) conditions. The beta-cleavage process that was diagnostic to M(+) is absent in all the CID spectra of the ions studied under ESI. Dissociation of all the studied ions resulted in the fragment ions formed by sequential elimination of pyrrole (A) and/or dialkyl/alkyl aryl/cycloalkyl (B) groups involving hydrogen migration to pyrrole ring at each cleavage of A--B bond, which clearly reveals the arrangement of A and B groups in the calix(4)pyrroles. The source of hydrogen that migrates to pyrrole ring during A--B bond cleavage is investigated by the experiments on deuterated compounds and [M + D](+) ions; and confirmed that the hydrogen attached to pyrrole nitrogen, hydrogen on alpha-carbon of alkyl group and the H(+)/Na(+) ion that added during ESI process to generate [M + H](+)/[M + Na](+) ions involve in the migration. The yields of [M + Na](+) ions are found to be different for the isomeric meso-cycloalkyl compounds (cycloheptyl, and 2-, 3- and 4-methyl cyclohexyl) and for normal and N-confused calix(4)pyrroles. The isomeric methyl and 3-hydroxy/4-hydroxy phenyl calix(4)pyrroles show specific fragmentation pattern during the dissociation of their [M - H](-) ions.  相似文献   

7.
6-oxo-PGF(1alpha) and TXB(2) are the metabolites of the prostanglandin PGI(2) and of the thromboxane TXA(2), respectively. PGI(2) and TXA(2) are arachidonic acid-derived compounds which regulate the blood hemostasis. Their quick metabolism leads to the 6-oxo-PGF(1alpha) and TXB(2) metabolites in plasma. In order to study on a large base the external factors influencing the hemostatic conditions, there is a need for a fast and reliable assay for quantitating these metabolites. Some methods have been published for the analysis of the arachidonic acid-derived compounds and some are dealing with mass spectrometry but nonspecifically centered on these specific compounds with a fast and cheap protocol, amenable for large-scale studies. Here we describe an analytical strategy that incorporates a two-dimensional chromatography running coupled to tandem mass spectrometry that minimizes the sample preparation and addresses the presence of the TXB(2) anomers for a robust quantitation measurement. After a protein precipitation, 100 microl of the supernatant (corresponding to 50 microl of the original plasma) was injected in a two-dimensional chromatographic system which operates an on-line clean-up and a subsequent chromatographic separation of the targeted analytes with a limit of quantitation (LOQ) of 22 pg/ml for 6-oxo-PGF(1alpha), and and a LOQ of 25 pg/ml for TXB(2).  相似文献   

8.
Summary The mass spectra of 1:1 complexes ofEDTA with lanthanide cations (Ln=Sm, Eu, Gd, Tb or Dy) upon positive/negative LSIMS are presented. In glycerol used as a matrix, adduct-ions such as [M+H]+, [M+H+nGly]+, [2M+H]+, [2M+H+Gly]+ (positive LSIMS) or [M-H], [M-H+nGly], [2M-H], [2M-H+Gly] (negative LSIMS), wheren=1–3, are formed. Reactions leading to the formation of adduct-ions are suggested.
Positive/negative Liquid-Sekundärionen-Massenspektrometrie vonLn-EDTA-(1:1)-Komplexen. Bildung von molekularen Ionenaddukten mit neutralen Spezies aus der Matrix oderLn-EDTA
Zusammenfassung Die Massenspektren von 1:1-Komplexen vonEDTA mit Lanthanidenkationen (Ln=Sm, Eu, Gd, Tb oder Dy) mittels positiver/negativer LSIMS werden präsentiert. In Glycerin als Matrix bilden sich Adduktionen wie [M+H]+, [M+H+nGly]+, [2M+H]+, [2M+H+Gly]+ (positive LSIMS) oder [M-H], [M-H+nGly], [2M-H], [2M-H+Gly] (negative LSIMS), wobein=1–3. Es werden Reaktionen vorgeschlagen, die zur Bildung von Adduktionen führen.
  相似文献   

9.
The aim of this study was to present integrated mass spectrometric methods for the structural characterization and identification of flavonoid glycoconjugates. During the liquid chromatography/mass spectrometry analyses, TriVersa NanoMate chip‐based system with nanoelectrospray ionization and fraction collection was combined to a quadrupole time‐of‐flight mass spectrometer. In the extract samples prepared from green leaves of wheat plantlets, 41 flavonoid derivatives were recognized. Part of the target natural products had the full structure being characterized after the registration of mass spectra, where m/z values for protonated [M + H]+ and deprotonated molecules [M ? H]? were annotated. MS2 and pseudo‐MS3 experiments were performed for [M + H]+ or [M ? H]? and aglycone ions (Y0+/?‐type), respectively. It should be underlined that pseudo‐MS3 mass spectra were registered for aglycone product ions in the mass spectra of O‐glycosides present in the extract samples. In many cases, only tentative structural identification of aglycones was possible, mainly because of the presence of numerous C‐monoglycoside or C‐diglycoside in the samples. Acylation of the sugar moiety and/or methylation of the aglycone in the flavonoid glycosides under study was observed. The existence of isobaric and/or isomeric compounds was demonstrated in the extract studied. The collision‐induced dissociation mass spectra registered for C,O‐diglycosides and C,C‐diglycosides did not permit to draw complete structural conclusions about the compounds studied. For the investigated class of natural products, unambiguous classification of sugar moieties linked to the aglycones from the recorded mass spectra was not possible. Registration of the positive and negative ion mass spectra did not lead to the precise conclusion about the glycosylation position at C‐6 or C‐8, and O‐4′ or O‐7 atoms. It was possible, on the basis of the collected MS2 spectra, to differentiate between O‐glycosides and C‐glycosides present in the samples analyzed. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

10.
Gynura procumbens (Lour.) Merr. is traditionally used as a raw material for making dumplings or steamed stuffed buns, and its fresh leaves are boiled with water for tea. Herein, we established an ultra‐high–performance liquid chromatography–quadrupole time‐of‐flight mass spectrometry (UHPLC–Q‐TOF–MS/MS) combined with characteristic ion filtration (CIF) strategy to rapidly screen active ingredients with antioxidant and anti‐inflammatory properties in G. procumbens. This strategy involved screening the active part of G. procumbens using antioxidation and anti‐inflammatory activity assays; discovering the active compounds by speculating on the active site's chemical composition by UHPLC–Q‐TOF–MS/MS plus CIF; and verifying the active compounds' activities. The ethyl acetate extract (EEAF) of G. procumbens was the major active site. Eighty‐one compounds were identified from the EEAF using UHPLC–Q‐TOF–MS/MS plus CIF. Furthermore, polyphenols such as cynarine, isochlorogenic acids A and isochlorogenic acids C have excellent antioxidizing and anti‐inflammatory activities. This study provides a practical strategy for rapid in vitro screening of the antioxidizing and anti‐inflammatory activities of traditional vegetables and herbs and identification of active ingredients.  相似文献   

11.
The factors determining the spectroscopic characteristics of the v(1)-SO4(2-) band of the MgSO4 ion pairs are discussed via ab initio calculation, including coupling effect, hydrogen bonding effect, and direct contact effect of Mg2+ with SO4(2-). With the calculation of the heavy water hydrated contact ion pairs (CIP), the overlap between the librations of water and the v(1)-SO4(2-) band can be separated, and thus the coupling effect is abstracted, and this coupling effect leads to a blue shift for the v(1)-SO4(2-) band of 5.6 cm(-1) in the monodentate CIP and 3.6 cm(-1) in the bidentate CIP. The hydrogen bonding between each water molecule without relation to Mg2+ and the sulfate ion makes the v(1)-SO4(2-) band blue shift of 3.7 cm(-1). When the outer-sphere water around Mg2+ are hydrogen bonded between SO4(2-) and Mg2+, it will make the largest disturbance to the v(1)-SO4(2-) band. Moreover, the inner-sphere water can affect the v(1)-SO4(2-) band conjunct with the direct contact of Mg2+ with SO4(2-), showing a blue shift of 14.4 cm(-1) in the solvent-shared ion pair, 22.6 cm(-1) in the monodentate CIP, 4.3 cm(-1) in the bidentate CIP, and 21.4 cm(-1) in the tridentate CIP. At last, the Raman spectral evolution in the efflorescence production process is tried to be rationalized. The shoulder at 995 cm(-1) is attributed to the monodentate CIP with 2-3 outer-sphere water molecules, whereas the new peak at 1021 cm(-1) at high concentration is assigned to the formation of aqueous triple ion.  相似文献   

12.
The disposition of the cannabimimetic naphthalen‐1‐yl‐(1‐pentylindol‐3‐yl)methanone (JWH‐018) in mice following inhalation of the smoke of the herbal incense product (HIP) ‘Buzz’ is presented. A high‐pressure liquid chromatography with electrospray ionization triple quadrupole mass spectrometer (HPLC/MS/MS) method was validated for the analysis of JWH‐018 in the specimens using deuterated Δ9‐tetrahydrocannabinol (d3‐THC) as the internal standard. JWH‐018 was isolated by cold acetonitrile liquid–liquid extraction. Chromatographic separation was performed on a Zorbaz eclipse XDB‐C18 column. The assay was linear from 1 to 1000 ng/mL. Six C57BL6 mice were sacrificed 20 min after exposure to the smoke of 200 mg ‘Buzz’ containing 5.4% JWH‐018. Specimen concentrations of JWH‐018 were: blood, 54–166 ng/mL (mean 82 ± 42 ng/mL); brain, 316–708 ng/g (mean 510 ± 166 ng/g); and liver, 1370–3220 ng/mL (mean 1990 ± 752 ng/mL). The mean blood to brain ratio for JWH‐018 was 6.8 and ranged from 4.2 to 10.9. After exposure, the responses of the mice were consistent with cannabinoid receptor type 1 activity: body temperatures dropped 7.3 ± 1.1 °C, and catalepsy, hyperreflexia, straub tail and ptosis were observed. The brain concentrations and physiological responses are consistent with the hypothesis that the behavioral effects of ‘Buzz’ are attributable to JWH‐018. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

13.
Negative ion tandem mass spectrometry (MS/MS) spectra of three isomeric triantennary N-linked glycans provided clear differentiation between the isomers and confirmed the occurrence of an isomer that was substituted with galactose on a bisecting GlcNAc (1 --> 4-substituted on the core mannose) residue recently reported by Takegawa et al. from N-glycans released from human immunoglobulin G (IgG). We extend this analysis of human serum IgG to reveal an analogue of the fucosylated triantennary glycan reported by Takegawa et al. together with a third compound that lacked both the sialic acid and the fucose residues. In addition, we demonstrate the biosynthesis of bisected hybrid-type glycans with the galactose modification, with and without core fucose, on the stem cell marker glycoprotein, 19A, expressed in a partially ricin-resistant human embryonic kidney cell line. It would appear, therefore, that this modification of N-linked glycans containing a galactosylated bisecting GlcNAc residue may be more common than originally thought. Negative ion MS/MS analysis of glycans is likely to prove an invaluable tool in the analysis and monitoring of therapeutic glycoproteins.  相似文献   

14.
15.
In a previous report we showed that certain binary Ag(+)-amino acid complexes formed adduct ions by the attachment of a single water and methanol molecule when stored in an ion trap mass spectrometer: complexes with aliphatic amino acids and with 4-fluorophenylalanine formed the adduct ions whereas complexes with phenylalanine and tryptophan did not. In this study we compared the tendency of the Ag(+) complexes derived from phenylalanine, 4-fluorophenylalanine, 4-hydroxyphenylalanine (tyrosine), 4-bromophenylalanine, 4-nitrophenylalanine and aminocyclohexanepropionic acid to form water adducts when stored, without further activation, in the ion trap for times ranging from 1 to 500 ms. Because the donation of pi electron density to the Ag(+) ion is a likely determining factor in complex reactivity, our aim in the present study was to determine qualitatively the influence of para-position substituents on the aromatic ring on the formation of the water adducts. Our results show that the reactivity of the complexes is influenced significantly by the presence of the various substituents. Decreases in [M + Ag](+) ion abundance, and increases in adduct ion abundance, both measured as a function of storage time, follow the trend -NO(2) > -Br > -F > -OH > -H. The complex of Ag(+) with 4-nitrophenylalanine was nearly as reactive towards water as the Ag(+) complex with aminocyclohexanepropionic acid, the last being an amino acid devoid of pi character in the ring system. Collision induced dissociation of the [M + Ag](+) species derived from the amino acids produces, among other products, Ag(+) complexes with a para-substituted phenylacetaldehyde: complexes that also form adduct species when stored in the ion trap. The trends in adduct ion formation exhibited by the aldehyde-Ag(+) complex ions were similar to those observed for the precursor complexes of Ag(+) and the amino acids, confirming the influence of the ring substituent.  相似文献   

16.
《Electrophoresis》2017,38(3-4):540-546
In this work, we present a novel electrophoretic method that was developed for the determination of lithium and transition metals in LiNi1/3Co1/3Mn1/3O2 cathode material after microwave digestion. The cations in the digested LiNi1/3Co1/3Mn1/3O2 material were separated by CE and the element content was determined by UV/Vis detection. To characterize the precision of the measurements, the RSDs and concentrations were calculated and compared to those obtained with ICP‐optical emission spectrometry (ICP‐OES). Furthermore, a certified reference material (BCR 176R —fly ash) was investigated for all techniques. For active material components, the LOD and LOQ were determined. The LODs and LOQs for the metals determined by CE were as follows: lithium (LOD/LOQ): 17.41/62.70 μg/L, cobalt (LOD/LOQ): 348.4/1283 μg/L, manganese (LOD/LOQ): 540.2/2095 μg/L, and nickel (LOD/LOQ): 838.0/2982 μg/L. Recovery rates for lithium were in the range of 95–103%. It could be proven that with the new technique, the results for the determination of the lithium content of active material were comparable with those obtained by ICP‐OES and ion chromatography. Furthermore, the recovery rates of the transition metals were determined to be between 96 and 110% by CE and ICP‐OES.  相似文献   

17.
Metal oxides have a large storage capacity when employed as anode materials for lithium‐ion batteries (LIBs). However, they often suffer from poor capacity retention due to their low electrical conductivity and huge volume variation during the charge–discharge process. To overcome these limitations, fabrication of metal oxides/carbon hybrids with hollow structures can be expected to further improve their electrochemical properties. Herein, ZnO‐Co3O4 nanocomposites embedded in N‐doped carbon (ZnO‐Co3O4@N‐C) nanocages with hollow dodecahedral shapes have been prepared successfully by the simple carbonizing and oxidizing of metal–organic frameworks (MOFs). Benefiting from the advantages of the structural features, i.e. the conductive N‐doped carbon coating, the porous structure of the nanocages and the synergistic effects of different components, the as‐prepared ZnO‐Co3O4@N‐C not only avoids particle aggregation and nanostructure cracking but also facilitates the transport of ions and electrons. As a result, the resultant ZnO‐Co3O4@N‐C shows a discharge capacity of 2373 mAh g?1 at the first cycle and exhibits a retention capacity of 1305 mAh g?1 even after 300 cycles at 0.1 A g?1. In addition, a reversible capacity of 948 mAh g?1 is obtained at a current density of 2 A g?1, which delivers an excellent high‐rate cycle ability.  相似文献   

18.
《Analytical letters》2012,45(11):1157-1167
Abstract

Temperature profiling of MIKE spectra allows identification of protonated psilocin in the presence of other compounds with the same mass. The use of multiple ionizing methods to give ions (M+H)+, (M+ML4)+, and M is shown to assist in determining the molecular weights of new natural products. Structural information is obtained from the MIKE spectra. All these determinations can be made on the intact mushroom or its simple alcoholic extracts.  相似文献   

19.
The basic study on the determination of tetrafluoroborate ion (BF4) by ion chromatography, and total boron by conversion of boric acid to BF4 followed by ion chromatography of BF4 has been carried out. The results of thermodynamic calculations for the system of boric acid (H3BO3)-F-H+ showed that the mole fraction of BF4 was higher than 99% at pH lower than 3.5 and 4.5 when the total free fluoride concentration (2[H2F2] + 2[HF2] + [HF] + [F]) was as high as 0.1 and 1.0 M, respectively. The fraction of BF4 increased with increasing total free fluoride concentration. BF4 fraction values were higher than 99% at pH 0.75 and at total free fluoride concentration of 0.05 M or higher. BF4 was hardly formed at pH > 7 even when the total free fluoride concentration was as high as 1.0 M. According to the experimental results, the fraction of BF4 at pH 0.7-0.8 was 51.2, 95.6 and 96.7% when the total fluoride concentration (2[H2F2] + 2[HF2] + [HF] + [F] + 3[BF3OH] + 4[BF4]) was 0.2, 1.0 and 3.3 M, respectively. The formation reaction of BF4 from boric acid reached an equilibrium state within 20 min regardless of reaction temperature, in the range of 20-50 °C, when the total boron and total fluoride concentrations were 66.7 mM and 1.0 M, respectively. Although BF4 was formed only under acidic conditions, BF4, once formed, was very stable under alkaline conditions at least for several hours. We have concluded that BF4 could be analyzed by ion chromatography using sodium hydroxide solution as an eluent because BF4 was stable under chromatographic conditions. BF4 solution prepared from boric acid could be used as a standard solution in the ion chromatographic analysis of BF4 instead of the sodium tetrafluoroborate (NaBF4) reagent available commercially, if a discrepancy of about 4-5% was allowed.  相似文献   

20.
Although series of N1, N1‐dimethyl‐N2‐arylformamidines and of 1,1,3,3‐tetraalkyl‐2‐arylguanidines are structurally analogous and similar electron‐ionization mass spectral fragmentation may be expected, they display important differences in the favored routes of fragmentation and consequently in substituent effects on ion abundances. In the case of formamidines, the cyclization‐elimination process (initiated by nucleophilic attack of the N‐amino atom on the 2‐position of the phenyl ring) and formation of the cyclic benzimidazolium [M‐H]+ ions dominates, whereas the loss of the NR2 group is more favored for guanidines. In order to gain information on the most probable structures of the principal fragments, quantum‐chemical calculations were performed on a selected set. A good linear relation between log{I[M‐H]+I [M]+?} and σR+ constants of substituent at para position in the phenyl ring occurs solely for formamidines (r = 0.989). In the case of guanidines, this relation is not significant (r = 0.659). A good linear relation is found between log{I[M‐NMe2]+/I [M]+?} and σp+ constants (r = 0.993). Copyright © 2010 John Wiley & Sons, Ltd.  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