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.
An ion‐mobility mass spectrometry study showed that the preferred O‐protonated form of p‐aminobenzoic in the gas phase can be converted to the thermodynamically less favored N‐protomer by in‐source collision‐induced ion activation during the ion transfer process from the atmospheric region to the first vacuum region if the humidity is high in the ion source. Upon the addition of water vapor to the nitrogen gas used to promote the solid analyte to the gas phase under helium‐plasma ionization conditions, the intensity of the ion‐mobility arrival‐time peak for the N‐protomer increased dramatically. Evidently, the ion‐activation process in the first vacuum region is able to provide the energy required to surmount the barrier to isomerize the O‐protomer to the more energetic N‐protomer. The transfer of the proton attached to the carbonyl oxygen atom of the O‐protomer to the amino group takes place by a water‐bridge mechanism. Apparently, the postionization transformations that take place during the transmission of ions from the atmospheric‐pressure ion source to the detector, via different physical compartments of low to high vacuum, play an eminent role in determining the population ratios eventually manifested at the detector.  相似文献   

2.
The role of water vapor in transforming the thermodynamically preferred species of protonated benzocaine to the less favored protomer was investigated using helium-plasma ionization (HePI) in conjunction with ion-mobility mass spectrometry (IM-MS). The IM arrival-time distribution (ATD) recorded from a neat benzocaine sample desorbed to the gas phase by a stream of dry nitrogen and ionized by HePI showed essentially one peak for the O-protonated species. However, when water vapor was introduced to the enclosed ion source, within a span of about 150 ms the ATD profile changed completely to one dominated by the N-protonated species. Under spray-based ionization conditions, the nature and composition of the solvents have been postulated to play a decisive role in defining the manifested protomer ratios. In reality, the solvent vapors present in the ion source (particularly the ambient humidity) indirectly dictate the gas-phase ratio of the protomers. Evidently, the gas-phase protomer ratio established at the confinement of the ions is readjusted by the ion-activation that takes place during the transmission of ions to the vacuum. Although it has been repeatedly stated that ions can retain a “memory” of their solution structures because they can be kinetically trapped, and thereby represent their solution-based stabilities, we show that the initial airborne ions can undergo significant transformations in the transit through the intermediate vacuum zones between the ion source and the mass detector. In this context, we demonstrate that the kinetically trapped N-protomer of benzocaine can be untrapped by reducing the humidity of the enclosed ion source.
Graphical Abstract ?
  相似文献   

3.
In this study, ion mobility separation coupled with tandem mass spectrometry (IM‐tandem MS) was utilized to investigate the ionization behaviors of two amino acids including leucine and isoleucine. Under the electrospray ionization (ESI) mode, two protonation sites in each molecular sturcture caused two forms of protomer. One arose from the amino being protonated (amino‐protomer) and the other from the carboxyl being protonated (carboxyl‐protomer). In the two‐dimensional (drift time, m/z) spectrum, the protomers had the same mass, but the distinguishable drift times and fragmentation patterns. For the characterization purpose, the theoretical collision cross section (CCS) values of the protomers were calculated and proven to be consistent with the experimental. Moreover, the quantified relationship between the amino acids and their protomers was evaluated. It showed that the abundance of the carboxyl‐protomer was proportional to the concentration of the amino acid, whereas that of the amino‐protomer did not have the same trend. Under the atmospheric pressure chemical ionization (APCI) mode, only the carboxyl‐protomer was observed. In addition, the amino‐protomer and the cluster ions observed under ESI were absent completely. The results demonstrate that the ionization mode impacts heavily on the ionization behaviors of leucine and isoleucine not only on the form of therir protomers but also on the quantified relationship.  相似文献   

4.
The gas‐phase chemistry of deprotonated benzyl N‐phenylcarbamates was investigated by electrospray ionization tandem mass spectrometry. Characteristic losses of a substituted phenylcarbinol and a benzaldehyde from the precursor ion were proposed to be derived from an ion‐neutral complex (INC)‐mediated competitive proton and hydride transfer reactions. The intermediacy of the INC consisting of a substituted benzyloxy anion and a phenyl isocyanate was supported by both ortho‐site‐blocking experiments and density functional theory calculations. Within the INC, the benzyloxy anion played the role of either a proton abstractor or a hydride donor toward its neutral counterpart. Relative abundances of the product ions were influenced by the nature of the substituents. Electron‐withdrawing groups at the N‐phenyl ring favored the hydrogen transfer process (including proton and hydride transfer), whereas electron‐donating groups favored direct decomposition to generate the benzyloxy anion (or substituted benzyloxy anion). By contrast, electron‐withdrawing and electron‐donating substitutions at the O‐benzyl ring exhibited opposite effects. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
Travelling wave ion mobility mass spectrometry (TWIM‐MS) with post‐TWIM and pre‐TWIM collision‐induced dissociation (CID) experiments were used to form, separate and characterize protomers sampled directly from solutions or generated in the gas phase via CID. When in solution equilibria, these species were transferred to the gas phase via electrospray ionization, and then separated by TWIM‐MS. CID performed after TWIM separation (post‐TWIM) allowed the characterization of both protomers via structurally diagnostic fragments. Protonated aniline (1) sampled from solution was found to be constituted of a ca. 5:1 mixture of two gaseous protomers, that is, the N‐protonated (1a) and ring protonated (1b) molecules, respectively. When dissociated, 1a nearly exclusively loses NH3, whereas 1b displays a much diverse set of fragments. When formed via CID, varying populations of 1a and 1b were detected. Two co‐existing protomers of two isomeric porphyrins were also separated and characterized via post‐TWIM CID. A deprotonated porphyrin sampled from a basic methanolic solution was found to be constituted predominantly of the protomer arising from deprotonation at the carboxyl group, which dissociates promptly by CO2 loss, but a CID‐resistant protomer arising from deprotonation at a porphyrinic ring NH was also detected and characterized. The doubly deprotonated porphyrin was found to be constituted predominantly of a single protomer arising from deprotonation of two carboxyl groups. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

6.
Helium Plasma Ionization (HePI) generates gaseous negative ions upon exposure of vapors emanating from organic nitro compounds. A simple adaptation converts any electrospray ionization source to a HePI source by passing helium through the sample delivery metal capillary held at a negative potential. Compared with the demands of other He‐requiring ambient pressure ionization sources, the consumption of helium by the HePI source is minimal (20–30 ml/min). Quantification experiments conducted by exposing solid deposits to a HePI source revealed that 1 ng of 2,4,6‐trinitrotoluene (TNT) on a filter paper (about 0.01 ng/mm2) could be detected by this method. When vapor emanating from a 1,3,5‐trinitroperhydro‐1,3,5‐triazine (RDX) sample was subjected to helium plasma ionization mass spectrometry (HePI‐MS), a peak was observed at m/z 268 for (RDX●NO2)?. This facile formation of NO2? adducts was noted without the need of any extra additives as dopants. Quantitative evaluations showed RDX detection by HePI‐MS to be linear over at least three orders of magnitude. TNT samples placed even 5 m away from the source were detected when the sample headspace vapor was swept by a stream of argon or nitrogen and delivered to the helium plasma ion source via a metal tube. Among the tubing materials investigated, stainless steel showed the best performance for sample delivery. A system with a copper tube, and air as the carrier gas, for example, failed to deliver any detectable amount of TNT to the source. In fact, passing over hot copper appears to be a practical way of removing TNT or other nitroaromatics from ambient air. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

7.
In this paper, we report nano‐electrospray ionization‐ion mobility mass spectrometry (nano‐ESI‐IM‐MS) characterization of bovine superoxide dismutase (SOD‐1) and human SOD‐1 purified from erythrocytes. SOD‐1 aggregates are characteristic of amyotrophic lateral sclerosis (ALS), a fatal neurodegenerative disease in humans that could be triggered by dissociation of the native dimeric enzyme (Cu2,Zn2‐dimer SOD‐1). In contrast to ESI‐MS, nano‐ESI‐IM‐MS allowed an extra dimension for ion separation, yielding three‐way mass spectra (drift time, mass‐to‐charge ratio and intensity). Drift time provided valuable structural information related to ion size, which proved useful to differentiate between the dimeric and monomeric forms of SOD‐1 under non denaturing conditions. In order to obtain detailed structural information, including the most relevant post‐translational modifications, we evaluated several parameters of the IM method, such as sample composition (10 mM ammonium acetate, pH 7) and activation voltages (trap collision energy and cone voltage). Neutral pH and a careful selection of the most appropriate activation voltages were necessary to minimize dimer dissociation, although human enzyme resulted less prone to dissociation. Under optimum conditions, a comparison between monomer‐to‐dimer abundance ratios of two small sets of blood samples from healthy control and ALS patients demonstrated the presence of a higher relative abundance of Cu,Zn‐monomer SOD‐1 in patient samples. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

8.
A promising replacement for the radioactive sources commonly encountered in ion mobility spectrometers is a miniaturized, energy‐efficient photoionization source that produce the reactant ions via soft X‐radiation (2.8 keV). In order to successfully apply the photoionization source, it is imperative to know the spectrum of reactant ions and the subsequent ionization reactions leading to the detection of analytes. To that end, an ionization chamber based on the photoionization source that reproduces the ionization processes in the ion mobility spectrometer and facilitates efficient transfer of the product ions into a mass spectrometer was developed. Photoionization of pure gasses and gas mixtures containing air, N2, CO2 and N2O and the dopant CH2Cl2 is discussed. The main product ions of photoionization are identified and compared with the spectrum of reactant ions formed by radioactive and corona discharge sources on the basis of literature data. The results suggest that photoionization by soft X‐radiation in the negative mode is more selective than the other sources. In air, adduct ions of O2 with H2O and CO2 were exclusively detected. Traces of CO2 impact the formation of adduct ions of O2 and Cl (upon addition of dopant) and are capable of suppressing them almost completely at high CO2 concentrations. Additionally, the ionization products of four alkyl nitrates (ethylene glycol dinitrate, nitroglycerin, erythritol tetranitrate and pentaerythritol tetranitrate) formed by atmospheric pressure chemical ionization induced by X‐ray photoionization in different gasses (air, N2 and N2O) and dopants (CH2Cl2, C2H5Br and CH3I) are investigated. The experimental studies are complemented by density functional theory calculations of the most important adduct ions of the alkyl nitrates (M) used for their spectrometric identification. In addition to the adduct ions [M + NO3] and [M + Cl], adduct ions such as [M + N2O2], [M + Br] and [M + I] were detected, and their gas‐phase structures and energetics are investigated by density functional theory calculations. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

9.
The zwitterionic drug 3‐methyl‐9‐(2‐oxa‐2λ5‐2H‐1,3,2‐oxazaphosphorine‐2‐cyclohexyl)‐3,6,9‐triazaspiro[5,5]undecane chloride (SLXM‐2) is a novel synthetic compound which has shown anticancer activity and low toxicity in vivo. In this study, the various gas‐phase fragmentation routes were analyzed by electrospray ionization mass spectrometry (positive ion mode) in conjunction with tandem mass spectrometry (ESI‐MSn) for the first time. In ESI‐MS the fragment ion at m/z 289 (base peak) was formed by loss of the chlorine anion from the zwitterionic precursor SLXM‐2. The fragment ion at m/z 232 was formed from the ion at m/z 289 by loss of 1‐methylaziridine. The detailed gas‐phase collision‐induced dissociation (CID) fragmentation mechanisms obtained from the various precursor ions extracted from the zwitterionic SLXM‐2 drug was obtained by tandem mass spectrometry analyses. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
The isomeric structure of high‐mannose N‐glycans can significantly impact biological recognition events. Here, the utility of travelling‐wave ion mobility mass spectrometry for isomer separation of high‐mannose N‐glycans is investigated. Negative ion fragmentation using collision‐induced dissociation gave more informative spectra than positive ion spectra with mass‐different fragment ions characterizing many of the isomers. Isomer separation by ion mobility in both ionization modes was generally limited, with the arrival time distributions (ATD) often showing little sign of isomers. However, isomers could be partially resolved by plotting extracted fragment ATDs of the diagnostic fragment ions from the negative ion spectra, and the fragmentation spectra of the isomers could be extracted by using ions from limited areas of the ATD peak. In some cases, asymmetric ATDs were observed, but no isomers could be detected by fragmentation. In these cases, it was assumed that conformers or anomers were being separated. Collision cross sections of the isomers in positive and negative fragmentation mode were estimated from travelling‐wave ion mobility mass spectrometry data using dextran glycans as calibrant. More complete collision cross section data were achieved in negative ion mode by utilizing the diagnostic fragment ions. Examples of isomer separations are shown for N‐glycans released from the well‐characterized glycoproteins chicken ovalbumin, porcine thyroglobulin and gp120 from the human immunodeficiency virus. In addition to the cross‐sectional data, details of the negative ion collision‐induced dissociation spectra of all resolved isomers are discussed. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

11.
The fragmentation reactions of N‐monoalkyloxyphosphoryl amino acids (N‐MAP‐AAs) were studied by electrospray ionization tandem mass spectrometry (ESI‐MS). The sodiated cyclic acylphosphoramidates (CAPAs) were formed through a characteristic pentacoordinate phosphate participated rearrangement reaction in the positive‐ion ESI‐MS/MS and HR‐MS/MS of N‐MAP‐AAs, in which the fragmentation patterns were clearly different from those observed in the corresponding ESI‐MS/MS of N‐dialkyloxyphosphoryl amino acids/peptides and N‐phosphono amino acids. The formation of CAPAs depended on the chemical structures of N‐terminal phosphoryl groups, such as alkyloxy group, negative charge and alkali metal ion. A possible integrated rearrangement mechanism for both PN to PO phosphoryl group migration and formation of CAPAs was proposed. The fragmentation patterns of CAPAs as novel intermediates in gas phase were also investigated. In addition, it was found that the formation of α‐amino acid CAPAs was more favorable than β‐ or γ‐CAPAs in gas phase, which was consistent with previous solution‐phase experiments. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

12.
The synthesis, recrystallization, and X‐ray deterimination of N,N,N‐trimethyl‐5‐[(2,3,5,6‐tetrafluorophenoxy)carbonyl]pyridin‐2‐aminium trifluoromethanesulfonate (PyTFP‐precursor), C15H13F4N2O2+·CF3SO3, is described. This triflate salt precursor is required for the synthesis of 2,3,5,6‐tetrafluorophenyl 6‐[18F]‐fluoronicotinate ([18F]FPyTFP), a prosthetic group used to radiolabel peptides for positron emission tomography (PET), as peptides are increasingly being used as PET‐imaging probes in nuclear medicine. Radiolabeling of peptides is typically done using a `prosthetic group', a small synthon to which the radioisotope is attached in the first step, followed by attachment to the peptide in the second step. During the synthesis of the PyTFP‐precursor, displacement of a Cl atom with trimethylamine gas and anion replacement with a triflate counter‐ion is critical, as incomplete replacement would hinder radioisotopic incorporation of nucleophilic fluorine‐18 and result in diminished radiochemical yields. The structural determination of the PyTFP‐precursor by X‐ray crystallography helped confirm the anion exchange of chloride with triflate.  相似文献   

13.
The mononuclear complex Pd(1‐TosC‐N3)2Cl2 (2) containing 1‐(p‐toluenesulfonyl)cytosine (1) as a ligand, as well as dinuclear complexes Pd2(1‐TosC?N3,N4)4 (3) and Pd2(1‐TosC?N3,N4)2DMSO2Cl2 (4) containing the ligand anion (1‐TosC?), was mass analyzed by electrospray ionization ion trap MS/MS and high resolution MS. Complexes 3 and 4 were obtained by recrystallization of 2 from DMF and DMSO, respectively. The behavior of complex 2 in different solutions was monitored by electrospray ionization mass spectrometry (ESI‐MS). Under the applied ESI‐MS conditions, complex 2 in methanol reorganized itself dominantly as new complex 3 and the solvent did not coordinate the formed species. In H2O/DMSO, CH3CN/DMSO and CH3OH/DMSO solutions, complex 2 formed several new species with solvent molecules involved in their structure, e.g. complex 4 was formed as the major product. The newly formed species were also examined by LC‐MS‐DAD, confirming the solvent induced reorganization and the solution instability of complex 2. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
A novel ion/molecule reaction was observed to occur under electrospray ionization (ESI), atmospheric pressure chemical ionization (APCI), and atmospheric pressure photo ionization (APPI) conditions, leading to dimerization of ionized 4‐(methyl mercapto)‐phenol followed by fast H· loss. The reaction is particularly favored during ESI, which suggests that this ion/molecule reaction can occur both in the solution inside the ESI‐charged droplets and in the gas‐phase environment of most other atmospheric pressure ionization techniques. The dimerization reaction is inherent to the electrolytic process during ESI, whereas it is more by ion/molecule chemistry in nature during APCI and APPI. From the tandem mass spectrometry (MS/MS) data, accurate mass measurements, hydrogen/deuterium (H/D) exchange experiments and density functional theory (DFT) calculations, two methyl sulfonium ions appear to be the most likely products of this electrophilic aromatic substitution reaction. The possible occurrence of this unexpected reaction complicates mass spectral data interpretation and can be misleading in terms of structural assignment as reported herein for 4‐(methyl mercapto)‐phenol. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

15.
The ability of positively charged aggregates of the surfactant (1R ,2S )‐dodecyl(2‐hydroxy‐1‐methyl‐2‐phenylethyl)dimethylammonium bromide (DMEB) to incorporate D‐tryptophan or L‐tryptophan in the gas phase has been investigated by electrospray ion mobility mass spectrometry (ESI‐IM‐MS). Strongly impacted by the pH of the electrosprayed solutions, both protonated (T+) and deprotonated (T) tryptophan are effectively included into the aggregates, whereas, tryptophan in zwitterionic (T0) form is practically absent in singly charged DMEB aggregates but can be found in multiply charged ones. The ability to incorporate tryptophan increases with the aggregation number and charge state of aggregates. More than 1 tryptophan species can be entrapped (aggregates including up to 5 tryptophan are observed). Collision induced dissociation experiments performed on the positively singly charged DMEB hexamer containing 1 T show that at low collision energies the loss of a DMEB molecule is preferred with respect to the loss of the DMEB cation plus T species which, in turn, is preferred with respect to the loss of mere tryptophan, suggesting that the deprotonated amino acid is preferentially located in proximity of a DMEB head group and with the ionic moiety pointing towards the core of the aggregate. The analysis of the collision cross sections (CCS) of bare and tryptophan containing aggregates allowed evaluating the contributions of tryptophan and bromide ions to the total aggregate CCS. No significant discrimination between D‐tryptophan and L‐tryptophan by the chiral DMEB aggregates has been evidenced by mass spectra data, CID experiments, and CCS values.  相似文献   

16.
A method of reversed‐phase ion‐pair solid‐phase extraction combined with ion chromatography for determination of pyrrolidinium ionic liquid cations (N‐methyl‐N‐ethyl pyrrolidinium, N‐methyl‐N‐propyl pyrrolidinium, and N‐methyl‐N‐butyl pyrrolidinium) in water samples was developed in this study. First, ion‐pair reagent sodium heptanesulfonate was added to the water samples after static, centrifugation and filteration. Then, pyrrolidinium cations in the samples were enriched and purified by a reversed‐phase solid‐phase extraction column, and eluted from the column with methanol aqueous solution as eluent. Finally, the eluate collected was analyzed by ion chromatography. The separation and direct conductivity detection of these pyrrolidinium cations by ion‐exchange column using 1.0 mM methanesulfonic acid (in water)/acetonitrile (97:3, v:v) as mobile phase was achieved within 10 min. By using this method, pyrrolidinium cations in Songhua River and Hulan River were successfully extracted with the recoveries ranging from 74.2 to 97.1% and the enrichment factor assessed as 60. Pyrrolidinium cations with the concentration of 0.001?0.03 mg/L can be enriched and detected in the water samples. The developed method for the determination of pyrrolidinium ionic liquid cations in water samples is simple and reliable, which provides a reference for the study of the potential impact of ionic liquids on the environment.  相似文献   

17.
The thermal reduction of N2O by CO mediated by the metal‐free cluster cations [Si2Ox].+ (x =2–5) has been examined in the gas phase using Fourier transform ion cyclotron resonance (FT‐ICR) mass spectrometry in conjunction with quantum chemical calculations. Three successive oxidation/reduction steps occur starting from [Si2O2].+ and N2O to form eventually [Si2O5].+; the latter as well as the intermediate oxide cluster ions react sequentially with CO molecules to regenerate [Si2O2].+. Thus, full catalytic cycles occur at ambient conditions in the gas phase. Mechanistic aspects of these sequential redox processes have been addressed to reveal the electronic origins of these unparalleled reactions.  相似文献   

18.
Procyanidins are polymeric flavan‐3‐ones occurring in many plants with antioxidant and other beneficial bioactivities. They are composed of catechin and epicatechin monomeric units connected by single carbon‐carbon B‐type linkages or A‐type linkages containing both carbon‐carbon and carbon‐oxygen‐carbon bonds. Their polymeric structure makes analysis of procyanidin mixtures always difficult. Evaluation of procyanidins according to degree of polymerization (DP) using high‐performance liquid chromatography (HPLC) is time‐consuming and at best has resolved polymeric families up to DP‐17. To expedite studies of procyanidins, the utility of positive ion electrospray ion mobility‐mass spectrometry (IM‐MS) was investigated for the rapid separation and characterization of procyanidins in mixtures. Applying IM‐MS to analyse structurally defined standards containing up to five subunits, procyanidins could be resolved in less than 6 ms not only by degree of polymerization but also by linkage type. A‐type procyanidins could be resolved from B‐type and both could be at least partially resolved from mixed‐type procyanidins of the same DP. IM‐MS separated higher order procyanidins with DP of at least 24 from extracts of cranberry. As DP increased, the abundances of multiply‐charged procyanidins also increased. During IM‐MS of ions of similar m/z, the ion drift times decreased inversely with increasing charge state. Therefore, IM‐MS was shown to separate mixtures of procyanidins containing at least 24 interconnected subunits in less than 16 ms, not only according to DP, but also according to linkage type between subunits and charge state.  相似文献   

19.
The formation of crystalline CdSe particles in the thermal degradation of Cd(SePh)2·TMEDA (TMEDA = tetramethylethylenediamine) as a single‐source‐precursor was investigated by in‐situ powder X‐ray diffraction. It was shown that the primary grains were formed in the cubic zinc blende modification. After an increase in particle size by further annealing a phase transition to the thermodynamically favored hexagonal wurtzite type was detected. This behaviour indicates that, according to Ostwald's rule, the primary grains consist of the less stable polymorph due to the lower activation barrier of its formation. When the volume energy of the particles gets dominant over the surface energy, the metastable form is transformed and the system adopts the modification of lowest energy.  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