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 187 毫秒
1.
This paper describes how weakly bound adduct ions form when the precursor ions used in selected ion flow mass spectrometry, SIFT-MS, analyses, viz. H3O+, NO+ and O2+, associate with the major components of air and exhaled breath, N2, O2 and CO2. These adduct ions, which include H3O+N2, H3O+CO2, NO+CO2, O2+O2 and O2+CO2, are clearly seen when dry air containing 5% CO2 (typical of that in exhaled breath) is analysed using SIFT-MS. These adduct ions must not be misinterpreted as characteristic product ions of trace gases; if so, serious analytical errors can result. However, when exhaled breath is analysed these adduct ions are partly removed by ligand switching reactions with the abundant water molecules and the problems they represent are alleviated. But the small fractions of the adduct ions that remain in the SIFT-MS spectra, and especially when they are isobaric with genuine characteristic product ion of breath trace gases, can result in erroneous quantifications; such is the case for H3O+N2 interfering with breath ethanol analysis and H3O+CO2 with breath acetaldehyde analysis. However, these difficulties can be overcome when the isobaric adduct ions are properly recognised and excluded from the analyses; then these two important compounds can be properly quantified in breath. The presence of O2+CO2 in the product ion spectra interferes with the analysis of CS2 present at low levels in exhaled breath. It is likely that similar problems will occur as other trace compounds are detected in exhaled breath when consideration will have to be given to the possibility of overlapping between their characteristic product ions and ions produced by hitherto unknown reactions. Similar problems are evident in other systems; for example, H3O+CH4 adduct ions are observed in both SIFT-MS analyses of methane rich mixtures like biologically generated waste gases and in model planetary atmospheres.  相似文献   

2.
The reactions of carbon dioxide, CO2, with the precursor ions used for selected ion flow tube mass spectrometry, SIFT‐MS, analyses, viz. H3O+, NO+ and O, are so slow that the presence of CO2 in exhaled breath has, until recently, not had to be accounted for in SIFT‐MS analyses of breath. This has, however, to be accounted for in the analysis of acetaldehyde in breath, because an overlap occurs of the monohydrate of protonated acetaldehyde and the weakly bound adduct ion, H3O+CO2, formed by the slow association reaction of the precursor ion H3O+ with CO2 molecules. The understanding of the kinetics of formation and the loss rates of the relevant ions gained from experimentation using the new generation of more sensitive SIFT‐MS instruments now allows accurate quantification of CO2 in breath using the level of the H3O+CO2 adduct ion. However, this is complicated by the rapid reaction of H3O+CO2 with water vapour molecules, H2O, that are in abundance in exhaled breath. Thus, a study has been carried out of the formation of this adduct ion by the slow three‐body association reaction of H3O+ with CO2 and its rapid loss in the two‐body reaction with H2O molecules. It is seen that the signal level of the H3O+CO2 adduct ion is sensitively dependent on the humidity (H2O concentration) of the sample to be analysed and a functional form of this dependence has been obtained. This has resulted in an appropriate extension of the SIFT‐MS software and kinetics library that allows accurate measurement of CO2 levels in air samples, ranging from very low percentage levels (0.03% typical of tropospheric air) to the 6% level that is about the upper limit in exhaled breath. Thus, the level of CO2 can be traced through single time exhalation cycles along with that of water vapour, also close to the 6% level, and of trace gas metabolites that are present at only a few parts‐per‐billion. This has added a further dimension to the analysis of major and trace compounds in breath using SIFT‐MS. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
The chemical and physical structure of ion-implanted silicone rubbers has been studied in order to analyze their blood compatibility such as reduction of platelet accumulation owing to ion implantation. H+2, He+, C+, O+, O+2, N+, N+2, Ne+, Na+, Ar+, K+, and Kr+ ion implantations were performed at an energy of 150 keV with fluences between 1 × 1017 and 3 × 1017 ions/cm2 at room temperature. Results of FT-IR-ATR showed that ion implantation broke the original chemical bond to form new radicals such as OH, >C = O, SiH, and CH2. The formation of these radicals depended on the ion species employed: >C = O formation by O+ or O+2 implantation and formation of amines by N+ or N+2 implantation. The results of Raman spectroscopy showed that ion implantation always produced a peak at near 1500 cm−1, although the intensity of this peak was dependent on the ion species. The light ions like H+2 and He+ were more effective than heavy ions in producing this peak, and O+2 implantation was the most effective on producing amorphous carbon. These results indicated that >C = O and amorphous carbon, generated by O+2 implantation, may improve the antithrombogenicity. The antithrombogenicity was tested by the superior vena cava (SVC) indwelling method for two days in rats with in-111-tropolone-platelets, and by the inferior vena cava (IVC) indwelling method for periods of 1–4 weeks in dogs. Results of the SVC indwelling method showed that platelet accumulation on H+2 and O+2 implanted specimens decreased. In particular 1 × 1017 O+2/cm2 implantation caused both accumulation onto specimens and the SVC to decrease. Macroscopic views of the ion-implanted IVC specimens in dogs revealed little thrombus formation. It is concluded that ion implantation into silicone rod is a useful technique to improve its antithrombogenicity.  相似文献   

4.
The rate constants for proton transfer from H3+ ions to N2, O2, and CO have been measured as function of hydrogen buffer gas partial pressure. The rate constant for proton transfer from H3+ to N2 shows a very large pressure dependence, increasing from 1.0 × 10?9 cm3/s at low H2 partial pressures to 1.7 × 10?9 cm3/s at high H2 partial pressures. The rate constants for proton transfer from H3+ to O2 and CO are constant with partial pressure of H2; giving values of 6.4 × 10?10 cm3/s and 1.7 × 10?9 cm3/s, respectively. The roles of excess vibrational energy in H3+ ions and of equilibrium between forward and back reaction are discussed. Back reaction is observed only for the reaction of H3+ ions with O2, and an equilibrium constant of K = 2.0 ± 0.4 at 298 K has been determined. From these data the proton affinity of O2 is deduced to be 0.47 ± 0.11 kcal/mole higher than that of H2.  相似文献   

5.
《Chemical physics letters》1986,124(5):409-413
Intravalence and Rydberg-type electronic transitions from excited projectile atoms and ions have been observed in the spectral range 30-200 nm as a result of 3 keV H+2, N+2, and CO+ bombardment of magnesium and graphite surfaces. Population of both types of excited states is consistent with an electron promotion mechanism resulting from close atomic encounters.  相似文献   

6.
The dissociative photoionization of molecular‐beam cooled CH2CO in a region of ?10–20 eV was investigated with photoionization mass spectrometry using a synchrotron radiation as the light source. Photoionization efficiency curves of CH2CO+ and of observed fragment ions CH2+, CHCO+, HCO+, C2O+, CO+, and C2H2+ were measured to determine their appearance energies. Relative branching ratios as a function of photon energy were determined. Energies for formation of these observed fragment ions and their neutral counterparts upon ionization of CH2CO are computed with the Gaussian‐3 method. Dissociative photoionization channels associated with six observed fragment ions are proposed based on comparison of determined appearance energies and predicted energies. The principal dissociative processes are direct breaking of C=C and C‐H bonds to form CH2+ + CO and CHCO+ + H, respectively; at greater energies, dissociation involving H migration takes place.  相似文献   

7.
Several polyatomic ions in inductively coupled plasma–mass spectrometry are studied experimentally and by computational methods. Novel calculations based on spin-restricted open shell second order perturbation theory (ZAPT2) and coupled cluster (CCSD(T)) theory are performed to determine the energies, structures and partition functions of the ions. These values are combined with experimental data to evaluate a dissociation constant and gas kinetic temperature (Tgas) value. In our opinion, the resulting Tgas value can sometimes be interpreted to deduce the location where the polyatomic ion of interest is generated. The dissociation of N2H+ to N2+ leads to a calculated Tgas of 4550 to 4900 K, depending on the computational data used. The COH+ to CO+ system yields a similar temperature, which is not surprising considering the similar energies and structures of COH+ and N2H+. The dissociation of H2CO+ to HCO+ leads to a much lower Tgas (< 1000 to 2000 K). Finally, the dissociation of H2COH+ to HCOH+ generates a Tgas value between those from the other HxCO+ ions studied here. All of these measured Tgas values correspond to formation of extra polyatomic ion in the interface or extraction region. The computations reveal the existence of isomers such as HCO+ and COH+, and H2CO+ and HCOH+, which have virtually the same m/z values and need to be considered in the interpretation of results.  相似文献   

8.
The rate coefficients k for the ternary association reactions of CH3+ and CD3+ with H2, N2, O2 CO and CO2 N2+ with N2, and C+ with H2 and D2 have been measured within the temperature range 80–520 K in helium buffer gas. In every case, kAT?n and the magnitudes of k are greater when the deuterated species are involved.  相似文献   

9.
Product distributions and rate constants for the reaction of ground state C+ ions with O2, NO, HCl, CO2, H2S, H2O, HCN, NH3, CH4, H2CO, CH3OH, and CH3NH2 have been measured. Rate constants were obtained using ion cyclotron resonance trapped ion methods at JPL, and product distributions were obtained using a tandem (Dempster-ICR) mass spectrometer at the University of Utah. Rapid carbon isotope exchange has also been observed in C+-CO collisions.  相似文献   

10.
A method is described for measuring wavelength-resolved fluorescence lifetimes of ions formed by HeIα photoionization, using electric-field drift of the ions in competition with fluorescence. Fluorescence lifetimes of 216 and 232 ns for the 357 nm and 378 nm peaks of N2O+, and 3.5 μs for the à → X? band of CS2+, were measured. The wavelength-resolved à → X? band of CO2+ showed no significant dependence of fluorescence lifetime on V′, but there is some indication of variation in the CS2+ lifetimes in the à → X? band.  相似文献   

11.
Sabo M  Matúška J  Matejčík S 《Talanta》2011,85(1):400-405
This study deals with O2 generation in corona discharge (CD) in point to plane geometry for single flow ion mobility spectrometry (IMS) with gas outlet located behind the ionization source. We have designed CD of special geometry in order to achieve the high O2 yield. Using this ion source we have achieved in zero air conditions that up to 74% all negative ions were O2 or O2(H2O). It has been demonstrated that the non-electronegative nitrogen positively influences the efficiency of O2 generation in O2/N2 mixtures. The reduced ion mobility of 2.27 cm2 V−1 s−1 has been measured for O2/O2(H2O) ions in zero air. Additional ions detected in zero air (less than 200 ppb CO2) using the mass spectrometric and IMS technique were, NO2, N2O2 (2.37 cm2 V−1 s−1), NO3, N2O3 and N2O3(H2O). The CO3 and CO4 ions have been detected after the introduction of 5 ppm CO2 into zero air.  相似文献   

12.
Ion clusters were formed in a temperature-variable high-pressure ion source from neat acetone and acetone/water mixtures and subjected to tandem mass spectrometry studies-unimolecular and collisionally activated mass-analyzed ion kinetic energy spectroscopy. The predominance of water loss from H+(H20)(A) l=3, where A = acetone, suggests that the solvation sphere around H3O+ does not close at l = 3, contrary to the case of acetonitrile or dimethyl ether. The results may be interpreted in terms of suggested ion structures which involve isomerization enroute to dissociation. The virtual absence of H/D scrambling in the collisionally activated dissociation of H3O+(DA)3, DA =acetone-d 6, and of D3O+(A)3 means that if enolization takes place, it is a rate-determining step in an irreversible isomerization. The stability of H+(H2O)(A)3 is a dominant factor in the observation of acetone loss from H+(H20)(A)4.  相似文献   

13.
The Coulomb explosion process of N2O in an intense laser-field (∼5 PW/cm2) has been investigated by the high-resolution time-of-flight (TOF) spectroscopy. Six two-body explosion pathways involving the NO+, NO2+, N2 + molecular ions have been securely identified from the momentum-scaled TOF spectra of the fragment ions. Assuming a linear geometry, three-body explosion pathways were investigated by sequential and concerted explosion models. When the concerted model is adopted, the observed momentum distributions of six atomic ion channels; N+, N2+, N3+, O+, O2+ and O3+, were well fitted using the Gaussian momentum distribution with the optimized bond elongation factor of 2.2(3). From the yields of individual Coulomb explosion pathways determined by the fit, the abundance of the parent ions, N2Oz+ (z=2–8), prior to the two- body and three-body explosion processes was found to have a smooth distribution with a maximum at z∼3.  相似文献   

14.
Ionization-fragmentation of uranium(IV) tetraborohydride, U(BH4)4, by He+ and by N+/N2+ yields, predominantly, U(BH5)+ and U(B2H8)+, respectively. Attachment of thermal electrons yields U(BH4)4? and ions of 1, 2, and 3 mass units less. Fluoride transfer with SF6?, BF4?, and UFn? (n = 5–7) and reactions with other small ions (O?, O2?, NO2?, F?, Cl?, O2+) are described.  相似文献   

15.
The isomers 4‐methylethcathinone and N‐ethylbuphedrone are substitutes for the recently banned drug mephedrone. We find that with conventional proton transfer reaction mass spectrometry (PTR‐MS), it is not possible to distinguish between these two isomers, because essentially for both substances, only the protonated molecules are observed at a mass‐to‐charge ratio of 192 (C12H18NO+). However, when utilising an advanced PTR‐MS instrument that allows us to switch the reagent ions (selective reagent ionisation) from H3O+ (which is commonly used in PTR‐MS) to NO+, O2+ and Kr+, characteristic product (fragment) ions are detected: C4H10N+ (72 Da) for 4‐methylethcathinone and C5H12N+ (86 Da) for N‐ethylbuphedrone; thus, selective reagent ionisation MS proves to be a powerful tool for fast detection and identification of these compounds. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

16.
The distonic ions HO+?CHCH2C˙H2 (1) and CH3C(?O+H)CH2C˙H2 (2) were directly generated, their decompositions characterized and their appearance energies determined by photoionization. Heats of formation derived from the appearance energies were 757 kJ mol?1 for 1 and 692 kJ mol?1 for 2. Deuterium labeling demonstrates that both ions decompose at low energies in the same ways as their isomers with the same skeletal structures, consistent with proposals that 1 and 2 are intermediates in the decompositions of those systems. Surprisingly, the values of the translational energy releases accompanying the formation of CH3CO+ and C2H5CO+ from 2 appear to be inversely proportional to the available excess energy. The 1,2-H-shift RC(?O+H)CH2C˙H2 → RC(?O+H)C˙HCH3 is compared to the corresponding, non-occurring 1,2-H-shift in alkyl free radicals.  相似文献   

17.
The structure of title compound, C6H16N+·C14H8N3O7S, comprises discrete ions which are inter­connected by N—H⋯O and N—H+⋯O hydrogen bonds, leading to a neutral one‐dimensional network along [100]. These hydrogen bonds appear to complement the Coulombic inter­action and help to stabilize the structure further.  相似文献   

18.
The kinetics of the deactivation of O2(1Σg+) is studied in real time. O2(1Σg+) is generated in this system by the O(1D) + O2 reaction following O3laser flash photolysis in the presence of excess O2, and it is monitored by its characteristic emission band at 762 nm. Quenching rate constants were obtained for O2, O3, N2, CO2, H2O, CF4and the rare gases. Since O(1D) is the precursor for the formation of O2(1Σg+), the addition of an O(1D) quencher effectively lowers the initial concentration of O2(1Σg+). By measuring the initial intensity of the 762 nm fluorescence signal, the relative quenching efficiencies were determined for O(1D) quenching by N2, CO2, Xe, and Kr with respect to O2; the results are in good agreement with literature values.  相似文献   

19.
20.
Methylation is an essential metabolic process in the biological systems, and it is significant for several biological reactions in living organisms. Methylated compounds are known to be involved in most of the bodily functions, and some of them serve as biomarkers. Theoretically, all α‐amino acids can be methylated, and it is possible to encounter them in most animal/plant samples. But the analytical data, especially the mass spectral data, are available only for a few of the methylated amino acids. Thus, it is essential to generate mass spectral data and to develop mass spectrometry methods for the identification of all possible methylated amino acids for future metabolomic studies. In this study, all N‐methyl and N,N‐dimethyl amino acids were synthesized by the methylation of α‐amino acids and characterized by a GC‐MS method. The methylated amino acids were derivatized with ethyl chloroformate and analyzed by GC‐MS under EI and methane/CI conditions. The EI mass spectra of ethyl chloroformate derivatives of N‐methyl ( 1–18 ) and N,N‐dimethyl amino acids ( 19–35 ) showed abundant [M‐COOC2H5]+ ions. The fragment ions due to loss of C2H4, CO2, (CO2 + C2H4) from [M‐COOC2H5]+ were of structure indicative for 1–18 . The EI spectra of 19–35 showed less number of fragment ions when compared with those of 1–18 . The side chain group (R) caused specific fragment ions characteristic to its structure. The methane/CI spectra of the studied compounds showed [M + H]+ ions to substantiate their molecular weights. The detected EI fragment ions were characteristic of the structure that made easy identification of the studied compounds, including isomeric/isobaric compounds. Fragmentation patterns of the studied compounds ( 1–35 ) were confirmed by high‐resolution mass spectra data and further substantiated by the data obtained from 13C2‐labeled glycines and N‐ethoxycarbonyl methoxy esters. The method was applied to human plasma samples for the identification of amino acids and methylated amino acids. Copyright © 2016 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