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.
Electrospray ionization (ESI) was combined with ultra-high-resolution Fourier transform-ion cyclotron resonance mass spectrometry (FTICR MS) to characterize complex humic and fulvic acid mixtures. Lower than expected molecular weight distributions previously observed for humics when analyzed by ESI-MS have fueled speculation about a bias in favor of low molecular weight. Multiply charged ions, ionization suppression, and sample fragmentation have all been suggested as sources of this low molecular weight bias. In this work, resolution of the individual components of humic mixtures within a 1 mass-to-charge unit window was accomplished by FTICR MS at 9.4 T. At mass resolving powers between 60,000 (high mass) and 120,000 (low mass), it was possible to determine that virtually all ions present in spectra of Suwannee River fulvic and humic acid are singly charged, thus eliminating inadequate accounting for multiply charged ions as a primary source of any low molecular weight bias. The high-resolution mass spectra also revealed the presence of molecular families containing ions that differ from each other in degree of saturation, functional group substitution (primarily CH vs N and CH4 vs O), and number of CH2 groups. Ionization suppression and ion fragmentation were addressed for humic and fulvic acid mixtures and well-characterized poly(ethylene glycol) (PEG) mixtures with average molecular weights of 8000 and 10,000. Although these high molecular weight PEG mixtures fragment extensively under traditional positive-ion mode ESI conditions, similar fragmentation could not be confirmed for humic and fulvic acid mixtures.  相似文献   

2.
The multiple charging of large molecules in electrospray ionization provides key advantages for obtaining accurate molecular weights by mass spectrometry and for obtaining structural information by tandem mass spectrometry and MS(n) experiments. Addition of glycerol or m-nitrobenzyl alcohol into the electrospray solutions dramatically increases both the maximum observed charge state and the abundances of the high charge states of protein and peptide ions. Adding glycerol to acidified aqueous solutions of cytochrome c shifts the most abundant charge state from 17+ to 21+, shifts the maximum charge state from 20+ to 23+, and shifts the average charge state from 16.6+ to 20.9+. Much less m-nitrobenzyl alcohol (<1%) is required to produce similar results. With just 0.7% m-nitrobenzyl alcohol, even the 24+ charge state of cytochrome c is readily observed. Similar results are obtained with myoglobin and (Lys)4. For the latter molecule, the 5+ charge state is observed in the electrospray mass spectrum obtained from solutions containing 6.7% m-nitrobenzyl alcohol. This charge state corresponds to protonation of all basic sites in this peptide. Although the mechanism for enhanced charging is unclear, it does not appear to be a consequence of conformational changes of the analyte molecules. This method of producing highly charged protein ions should be useful for improving the performance of mass measurements on mass spectrometers with performances that decrease with increasing m/z. This should also be particularly useful for tandem mass spectrometry experiments, such as electron capture dissociation, for which highly charged ions are desired.  相似文献   

3.
This study focuses on the molecular level interpretation of the selective gas-phase cleavage at aspartic acid residues (Asp) in protonated peptides. A phi3P+CH2C(=O)group (phi = 2,4,6-trimethoxyphenyl) is attached to the N-terminal nitrogen of the selected peptides LDIFSDF and LDIFSDFR, via solid-phase synthesis, to "mimic" the tightly held charge of a protonated arginine (Arg) residue. Collision-induced dissociation in a quadrupole ion trap instrument and surface-induced dissociation in a dual quadrupole instrument were performed for electrospray-generated ions of the fixed-charge peptide derivatives. Selective cleavages at Asp-Xxx are observed for those ions with charge provided only by the fixed charge or for those with a fixed charge and one Arg plus one added proton. This supports a previously proposed mechanism which suggests that the cleavages at Asp-Xxx, initiated by the acidic hydrogen of the Asp residue, become significant when ionizing protons are strongly bound by Arg in the protonated peptides. It is clear that the fixed charge is indeed serving as a "mimic" of protonated Arg and that a protonated Arg side chain is not required to interact with the Asp to induce cleavage at Asp-Xxx. When the number of protons exceeds the number of Arg in a peptide containing Arg and Asp, nonselective cleavages occur. The fragmentation efficiency of the peptides is consistent with the idea that these nonselective cleavages are promoted by a mobile proton. The peptide with a fixed charge and one added proton, [phi3P+CH2C(=O)-LDIFSDF + H]2+, fragments much more efficiently than the corresponding peptide with a fixed charge, an Arg and one added proton, [phi3P+CH2C(=O)-LDIFSDFR + H]2+; both of these fragment more efficiently than the peptide with a fixed charge and no added proton, phi3P+CH2C(=O)-LDIFSDF. MS/MS/MS (i.e., MS3) experimental results for bn ions formed at Asp-Xxx from phi3P+CH2C(=O)-LDIFSDF and its H/D exchange derivative, phi3P+CH2C(=O)-LDIFSDF-d11, are consistent with the bn ions formed at Asp-Xxx having a succinic anhydride cyclic structure. MS/MS experiments were also carried out for phi3P+CH2C(=O)-AAAA, a peptide derivative containing active hydrogens only at amide nitrogens plus the C-terminus, and its active H/D exchange product, phi3P+CH2C(=O)-AAAA-d5. The results show that a hydrogen originally located at an amide nitrogen is transferred away in the formation of a cyclic charge remote b ion.  相似文献   

4.
A new technique for studying the time dependence of conformational changes of gas-phase protein ions is described. In this approach, a short pulse of electrosprayed protein ions is introduced into an ion trap and stored. After a defined time period, the distribution of ions is ejected from the trap into an ion mobility/time-of-flight mass spectrometer. Combined measurements of mobilities and flight times in the mass spectrometer provide information about the abundances of different conformer types and charge-state distributions. By varying the storage time in the trap, it is possible to monitor changes in ion conformation that occur over extended time periods (approximately 10-200 ms). The method is demonstrated by examining changes in cytochrome c ion conformations for the +7 to +10 charge states.  相似文献   

5.
The negative-ion mass spectra produced by kiloelectronvolt energy (CsI)nCs+ (n = 0-2) and megaelectronvolt energy 252Cf fission fragment projectile impacts on NaNO3 and NaNO2 were collected and compared. The mass spectra generated by impacts of the kiloelectronvolt polyatomic primary ions on NaNO3 were markedly different from those derived from the fission fragment impacts, featuring higher relative intensities of nitrate (NO3-) specific secondary ions (those that reflect the sample stoichiometry). The most prominent secondary ion (SI) peaks produced from NaNO3 by the kiloelectronvolt energy projectiles were NO3- and Na(NO3)2-, both of which relate directly back to the chemical composition of the staring material. Likewise, the most prominent peaks produced by the kiloelectronvolt energy polyatomic projectile impacts on NaNO2 were NO2- and Na(NO2)2-. The fission fragment projectiles produced SI spectra from NaNO3 that were dominated by signals characteristic more of NaNO2, indicating that the megaelectronvolt energy ions induce considerable degradation of the nitrate solid. In addition, the fission fragment projectile produced relative negative SI intensity distributions that are remarkably similar to those reported in earlier studies of the use of laser desorption to produce SI signals from NaNO3. Of the projectiles examined in this study, the 20 keV (CsI)Cs+ projectile generated negative-ion mass spectra that best differentiated NaNO3 and NaNO2, primarily by producing a base peak in the NaNO3 spectrum that was unambiguously representative of the original sample stoichiometry.  相似文献   

6.
We have investigated ultrasonic degradations of poly(ethylene oxide) (PEG) and poly(methyl methacrylate) (PMMA) in aqueous media by means of matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOFMS). The ultrasonic degradation of polymers was monitored as a function of ultrasonication duration to examine the structural details of ultrasonic degradation polymers. PEG solution ultrasonication produced five types of oligomers (M approximately 1000 Da) with different end groups, irrespective of the initial average molecular masses (M=2, 6, 20, and 2000 kDa). Several degradation pathways with free radical reactions have been suggested to explain these degradation products: the ultrasonic degradation of PEG is initiated by breaking of the C-O bond in the PEG chain, generating polymeric radicals with two terminal groups, i.e., X*( approximately CH2CH2*) and Y*( approximately CH2CH2O*), followed by termination with extraction or release of a hydrogen atom. However, PMMA (M=1630 Da) ultrasonication generated only one type of degradation oligomer, which has a hydrogen group at both ends, the same as that of the original oligomer. It has been suggested that the presence of the radical terminal groups X*( approximately CH2*) and Y*( approximately (CH3)CCOO(CH3)C*) is due to selective C-C bond breaking in the chain during the ultrasonic degradation of PMMA. The MALDI-TOFMS combined with the ultrasonic degradation technique (UD/MALDI-TOFMS) developed in this study could be extended to the analysis of synthetic polymer structures with high molecular weights.  相似文献   

7.
A number of tertiary amine and quaternary ammonium cations spanning a mass range of 60-146 amu (trimethylamine, tetramethylammonium, trimethylethylammonium, N,N-dimethylaminoethanol, choline, N,N-dimethylglycine, betaine, acetylcholine, (3-carboxypropyl)trimethylammonium) were investigated using electrospray ionization ion mobility spectrometry. Measured ion mobilities demonstrate a high correlation between mass and mobility in N(2). In addition, identical mobilities within experimental uncertainties are observed for structurally dissimilar ions with similar ion masses. For example, dimethylethylammonium (88 amu) cations and protonated N,N-dimethylaminoethanol cations (90 amu) show identical mobilities (1.93 cm(2) V(-1) s(-1)) though N,N-dimethylaminoethanol contains a hydroxyl functional group while dimethylethylammonium only contains alkyl groups. Computational analysis was performed using the modified trajectory (TJ) method with nonspherical N(2) molecules as the drift gas. The sensitivity of the ammonium cation collision cross sections to the details of the ion-neutral interactions was investigated and compared to other classes of organic molecules (carboxylic acids and abiotic amino acids). The specific charge distribution of the molecular ions in the investigated mass range has an insignificant affect on the collision cross section.  相似文献   

8.
Gas-phase ion/molecule chemistry has been combined with ion mobility separation and time-of-flight mass spectrometry to enable the characterization of large poly(ethylene glycol)s (PEGs) and PEGylated molecules (>40 kDa). A facile method is presented in which gas-phase superbases are reacted in the high-pressure source region of commercial TOF mass spectrometers to manipulate the charge states of large ions generated by electrospray ionization (ESI). Charge stripping decreases the spectral congestion typically observed in ESI mass spectra of high molecular weight polydisperse PEGylated molecules. From these data, accurate average molecular weights and molecular weight distributions for synthetic polymers and PEGylated proteins are determined. The average MW measured for PEGylated Granulocyte colony-stimulating factor (rh-GCSF, 40 726.2 Da) is in good agreement with the theoretical value, and a 16 Da mass shift is easily observed in the spectrum of an oxidized form of the heterogeneous PEGylated protein. Ion mobility separations can fractionate PEGs of different chain length; when coupled with charge stripping ion/molecule reactions, ion mobility mass spectrometry (IMMS) offers several analytical advantages over mass spectrometry alone for the characterization of large PEGylated molecules including enhanced dynamic range, increased sensitivity, and specificity. Low abundance free PEG in a PEGylated peptide preparation, which is not directly detectable by mass spectrometry, can be easily observed and accurately quantified with gas-phase ion/molecule chemistry combined with ion mobility mass spectrometry.  相似文献   

9.
Aqueous solutions of poly(ethylene glycol) (PEG) in a 10 mM ammonium acetate buffer are electrosprayed, and the maximum charge state on the resulting gas-phase ions is reduced to unity using a radioactive source. The mobility distribution of these charged particles is then measured in air in a differential mobility analyzer of unusually high resolution. The relation Z(m) between the mobility Z of a polymer molecule and its mass m is determined by means of narrowly distributed PEG mass standards. The molecular weight range of available standards is extended by generating clusters containing from one up to six molecules of the primary PEG standard. The mass at the peak of the distribution of the lowest standard (PEG-4k) is determined by MALDI mass spectrometry and agrees with the manufacturer's value and previous MALDI literature data. The masses for the 50K and 120K standards are found to differ by 8.6 and 6.6%, respectively, from the manufacturer's value. Using known relationships, the particle diameter d of the ions is calculated from the measured mobility. Plots of d versus m(1/3) give straight lines over the full mass range studied (4000-700 000 Da, particle diameter from 3 to 12 nm), indicating that these PEG particles are indeed spherical and have a density rho independent of size. The slope of the d versus m(1/3) curve provides a density rho = 1.25 g/cm(3), close to the known bulk density, rho(PEG) = 1.21 g/cm(3).  相似文献   

10.
Laser-induced acoustic desorption (LIAD), combined with chemical ionization with the ClMn(H(2)O)(+) ion, is demonstrated to facilitate the analysis of base oils by Fourier transform ion cyclotron resonance mass spectrometry. The LIAD/ClMn(H(2)O)(+) method produces only one product ion, [ClMn + M](+), for each component (M) in base oils, thus providing molecular weight (MW) information for the analytes. With the exception of one sample, no fragmentation was observed. The mass spectra indicate the presence of homologous series of ions differing in mass by multiples of 14 Da (i.e., CH(2)). All peaks in the spectra correspond to ions with even m/z values and hence are formed from hydrocarbons with no nitrogen atoms, in agreement with the compositional nature of base oils. The MW distributions measured for two groups of base oil samples cover the range 350-600 Da, which is in excellent agreement with the values determined by gas chromatography. Moreover, the hydrocarbon types (i.e., paraffin and cycloparaffins with different numbers of rings) present in each base oil sample can be determined based on the m/z values of the product ions. Finally, the results obtained by using LIAD/ClMn(H(2)O)(+) indicate that the efficiency of the technique (combined desorption and ionization efficiency) is similar for different hydrocarbon types and fairly uniform over a wide molecular weight range, thus allowing quantitative analysis of the base oils. Hence, the product ions' relative abundances were used to determine the percentage of each type of hydrocarbon in the base oil. In summary, three important parameters (MW distributions, hydrocarbon types, and their relative concentrations) can be obtained in a single experiment. This mass spectrometric technique therefore provides detailed molecular-level information for base oils, which cannot be obtained by other analytical methods.  相似文献   

11.
High-field asymmetric waveform ion mobility spectrometry (FAIMS) is a new technique that separates gas-phase ions at atmospheric pressure (760 Torr) and room temperature. A FAIMS instrument acts as an ion filter and can be set to continuously transmit one type of ion. Despite the stringent requirement for a flow of clean, dry gas in the FAIMS analyzer region, a method of coupling electrospray to FAIMS has been developed. The identity of the electrospray ions separated by FAIMS was determined using mass spectrometry (FAIMS-MS). The theory of FAIMS is discussed, and electrospray FAIMS-MS spectra of several compounds in modes P1, P2, N1, and N2 are presented. Ions appearing in P1 and N1 modes tend to have mobilities that increase as a function of increasing electric field strength, whereas ions appearing in P2 and N2 modes tend to have mobilities that decrease. In general, low-mass ions are focused in P1 and N1 modes, whereas larger ions (e.g., proteins) are focused in P2 and N2 modes. Short-chain peptides, (Gly)(n) where n = 1-6, are shown to cross over from P1 mode into P2 mode as the chain length increases. The removal of the low-mass solvent cluster ions, combined with a reduction of the background noise in electrospray FAIMS-MS, results in an improved signal-to-noise ratio for mass spectra of larger ions (e.g., cyctochrome c) when compared with conventional electrospray-MS. Preliminary results also suggest that various charge states of cytochrome c can be distinguished by FAIMS, implying that the ion mobility of these species at high electric field strength is sensitive to the structure of the protein ion. The linearity of response of electrospray FAIMS-MS was investigated using leucine enkephalin and shows the calibration curve to be linear for ~3 orders of magnitude.  相似文献   

12.
The blackbody infrared radiative dissociation technique was used to study the thermal decomposition of the gaseous B5 pentamer of the Shiga-like toxin I and its complexes with the Pk trisaccharide and a decavalent Pk-based oligosaccharide ligand (STARFISH, S). Dissociation of the protonated pentamer, (B5 + nH)n+ triple bond B5n+ where n = 11-14, proceeds almost exclusively by the loss of a single subunit (B) with a disproportionately large fraction (30-50%) of the parent ion charge. The degree of charge enrichment of the leaving subunit increases with increasing parent ion charge state. For n = 12-14, a distribution of product ion charge states is observed. The yields of the complementary pairs of product ions are sensitive to the reaction temperature, with higher temperatures favoring greater charge enrichment of the leaving subunit for +13 and +14, and the opposite effect for +12. These results indicate that some of the protons are rapidly exchanged between subunits in the gas phase. Dissociation of B5(14+) x S proceeds exclusively by the loss of one subunit, although the ligand increases the stability of the complex and also reduces the degree of charge enrichment in the ejected monomer. For B5(12+)(Pk)1-3, the loss of neutral Pk competes with loss of a subunit at low temperatures. Linear Arrhenius plots were obtained from the temperature-dependent dissociation rate constants measured for the loss of B from B5n+ and B514+ x S. The magnitude of the Arrhenius parameters is highly dependent on the charge state of the pentamer: Ea = 35 kcal/mol and A = 1,019 s(-1) (+14), 46 kcal/mol and 1,023 S(-1) (+13), 50 kcal/mol and 1026 s(-1) (+12), and 80 kcal/mol and 10(39) (+11). The Ea and A for B5(14+) x S are 59 kcal/mol and 10(30) s(-1), respectively. The reaction pathways leading to greater charge enrichment of the subunit lost from the B5(14+) and B5(13+) ions correspond to higher energy processes, however, these pathways are kinetically preferred at higher temperatures due to their large A factors. A simple electrostatic model, whereby charge enrichment leads to Coulombic repulsion-induced denaturation of the subunits and disruption of the intersubunit interactions, provides an explanation for the magnitude of the Arrhenius parameters and the origin of the asymmetric dissociation behavior of the complexes.  相似文献   

13.
A novel image charge detection mass spectrometer (CDMS) with improved sensitivity and mass accuracy is described. The improved detector design and method of data analysis allow us to measure a reliable mass for a single macroion that is an order of magnitude smaller than previously achieved with CDMS. The apparatus employs an image charge detector array consisting of 22 detectors. The detectors are divided into two groups that can be floated at different potentials. The signals from the detector array are analyzed using a correlation approach to yield the velocities in the two groups of detectors and the charge. These quantities, together with the voltage difference between the two groups of detectors, provide a value for the mass. The mass, m/z, and charge distributions recorded for 300 kDa poly(ethylene oxide) (PEG) are presented. The mass distribution shows a peak at around 300 kDa with a width close to that expected from the polymer size distribution. In addition, there are broad peaks in the mass distribution at around 100 and 500 MDa. The 300 kDa ions have m/z ratios of ~2 kDa/e, and the 100 and 500 MDa ions have m/z ratios of ~40 kDa/e. The 100 and 500 MDa ions probably result from PEG aggregates that are either present in solution or the residue of large electrospray droplets.  相似文献   

14.
High-resolution ion mobility/time-of-flight techniques were used to measure collision cross sections for 968 tryptic digest peptide ions obtained from digestion of common proteins. Here, we report a mobility signature that aids in identifying proline-containing peptides containing 4-10 residues. Of 129 peptides (< or = 10 residues in length) in the database that contain proline residues, 57% show multiple resolved features in the ion mobility distribution for at least one of the [M + H]+ or [M + 2H]2+ ions. These multiple features are attributed to different conformations that arise from populations of cis and trans forms of proline. The number of resolved peaks in the ion mobility distribution appears to be correlated with the peptide ion charge state and the number of proline residues in the peptide.  相似文献   

15.
Li D  Fu S  Lucy CA 《Analytical chemistry》1999,71(3):687-699
Plots of mobility versus the square root of ionic strength (I(1/2)) do not show the linear behavior predicted by Kohlrausch's law. Classical electrolyte theory states that such deviations are to be expected due to the finite size of the ions. This paper uses the Pitts equation to account for the effect of ionic size on the ionic strength dependence of mobilities in CZE. Experimental mobilities for carboxylates, phenols, and sulfonates of -1 to -6 charge in aqueous buffers ranging from 0.001 to 0.1 M ionic strength were described by μ(-) = μ(0) - Az (I(1/2)/(1 + 2.4I(1/2))), where the constant in the denominator is empirically determined. Infinite dilution mobilities (μ(0)) determined by extrapolation of mobility data to zero ionic strength based on this expression yielded excellent agreement (100.3 ± 3.3%) with literature values for 14 compounds in a variety of buffers. The Pitts equation provides a reasonable estimate of the constant A for solutes up to a charge of -5. However, this constant also depends on temperature and the nature of the buffer counterion, presumably due to ion association. Thus it is most appropriate to determine the constant A empirically for a given buffer system.  相似文献   

16.
The formation of a range of precursor ion charge states from a single concentrated and purified charge state, followed by activation of each charge state, is introduced as a means to obtain more protein structural information than is available from dissociation of a single charge state alone. This approach is illustrated using off-resonance collisional activation of the [M + 8H]8+ to [M + 6H]6+ precursor ions of the bacteriophage MS2 viral coat protein following concentration and purification of the [M + 8H]8+ charge state. This range of charge states was selected on the basis of an ion trap collisional activation study of the effects of precursor ion charge state on the dissociation of the [M + 12H]12+ to [M + 5H]5+ ions. Gas-phase ion/ion proton-transfer reactions and the ion parking technique were applied to purify and concentrate selected precursor ion charge states as well as to simplify the product ion spectra. The high-charge-state ions fragment preferentially at the N-terminal side of proline residues while the product ion spectra of the lowest charge states investigated are dominated by C-terminal aspartic acid cleavages. Maximum structural information is obtained by fragmentation of the intermediate-charge states.  相似文献   

17.
Electron capture dissociation (ECD) is a promising method for de novo sequencing proteins and peptides and for locating the positions of labile posttranslational modifications and binding sites of noncovalently bound species. We report the ECD of a synthetic peptide containing 10 alanine residues and 6 lysine residues uniformly distributed across the sequence. ECD of the (M + 2H)(2+) produces a limited range of c (c(7)-c(15)) and z (z(9)-z(15)) fragment ions, but ECD of higher charge states produces a wider range of c (c(2)-c(15)) and z (z(2)-z(6), z(9)-z(15)) ions. Fragmentation efficiency increases with increasing precursor charge state, and efficiencies up to 88% are achieved. Heating the (M + 2H)(2+) to 150 degrees C does not increase the observed range of ECD fragment ions, indicating that the limited products are due to backbone cleavages occurring near charges and not due to effects of tertiary structure. ECD of the (M + 2Li)(2+) and (M + 2Cs)(2+) produces di- and monometalated analogues of the same c and z ions observed from the (M + 2H)(2+), with the abundance of dimetalated fragment ions increasing with fragment ion mass, a result consistent with the metal cations being located near the peptide termini to minimize Coulombic repulsion. In stark contrast to the ECD results, collisional activation of cesiated dications overwhelmingly results in ejection of Cs(+). The abundance of cesiated fragment ions formed from ECD of the (M + Cs + Li)(2+) exceeds that of lithiated fragment ions by 10:1. ECD of the (M + H + Li)(2+) results in exclusively lithiated c and z ions, indicating an overwhelming preference for neutralization and cleavage at protonated sites over metalated sites. These results are consistent with preferential neutralization of the cation with the highest recombination energy.  相似文献   

18.
The molecular ion of the nanoparticle Au 25(SCH 2CH 2Ph) 18 (A 25(SR) 18) is observed at 7394 Da in fast atom bombardment (FAB, Xe atoms) ionization mass spectrometry using a 3-nitrobenzyl alcohol matrix. A distinctive pattern of positive fragment ions is evident in the mass interval 5225-7394 Da, where peaks are seen for successive mass losses equivalent to R 2S entities. Because the Au 25(SCH 2CH 2Ph) 18 nanoparticle structure is crystallographically known to consist of a centered Au 13 icosahedral core surrounded by six Au 2(SR) 3 semirings, the R 2S loses are proposed to represent serial rearrangements and decompositions of the semiring structures. Mass losses equivalent to R 2S 2 and R 2 entities also appear at the lower end of this mass interval. The most intense spectral peak, at m/ z = 5246 Da, is assigned to the fragment Au 25S 10, from which all of the CH 2CH 2Ph organic units have been cleaved but from which no gold atoms have been lost. A different pattern of fragmentation is observed at lower masses, producing ions corresponding to serial losses of one gold atom and varied numbers of sulfur atoms, which continues down to a Au 9S 2 fragment. FAB mass spectra of the Au nanoparticle are much easier to interpret than laser desorption/ionization spectra, but they show more extensive fragmentation than do electrospray and low laser pulse intensity MALDI spectra. The loss of R 2S fragmentation in FAB is distinctive and unlike that seen in the other ionization modes. The FAB spectrum for the nanoparticle Au 25(S(CH 2) 9CH 3) 18 is also reported; its fragmentation parallels that for Au 25(SCH 2CH 2Ph) 18, implying that this nanoparticle has the same surprising stellated (staples) structure.  相似文献   

19.
The [M + Ag](+) ions of polystyrene (PS) oligomers are formed by matrix-assisted laser desorption/ionization, and their fragmentation characteristics are determined by tandem mass spectrometry experiments in a quadrupole/time-of-flight mass spectrometer. Collisionally activated dissociation (CAD) of [M + Ag](+) starts with random homolytic C-C bond cleavages in the PS chain, which generate radical ions carrying either the initiating (a(n*), b(n*)) or the terminating (y(n*), z(n*)) chain end and primary (a(n*), y(n*)) or benzylic (b(n*), z(n*)) radical centers. The fragments ultimately observed arise by consecutive, radical-induced dissociations. The primary radical ions mainly decompose by monomer evaporation and, to a lesser extent, by beta-H(*) loss. The benzylic radical ions primarily decompose by 1,5-H rearrangement (backbiting) followed by beta C-C bond scissions; this pathway leads to either closed-shell fragments with CH(2) end groups, internal fragments with 2-3 repeat units, or truncated benzylic b(n*)/z(n*) radical ions that can undergo anew backbiting. The same internal fragments are produced in all backbiting steps; hence, these fragments and small benzylic radical ions (which cannot undergo backbiting) dominate the low-mass region of the CAD spectra, while the less abundant closed-shell fragments with CH(2) end groups (a(n)/y(n)) dominate the medium- and high-mass regions. The latter fragments are suitable for determining the individual initiating and terminating end groups, whereas the internal ions could be valuable in sequence analyses of styrene copolymers.  相似文献   

20.
We have tested the effect of m-nitrobenzyl alcohol (m-NBA) as a method to increase the average charge state of protonated gas-phase molecular ions generated by ESI from tryptic peptides and phosphopeptides. Various concentrations of m-NBA were added to the mobile phases of a liquid chromatography system coupled to an ESI tandem mass spectrometer. Addition of just 0.1% m-NBA changed the average charge state for the identified tryptic BSA peptides from 2.2+ to 2.6+. As a result, the predominant charge states for BSA peptides were changed from 2+ to > or =3+. To evaluate the benefits of peptide charge enhancement, the ETD fragmentation efficiency and Mascot peptide score were compared for BSA peptides in charge states 2+ and 3+. In all cases but one, triply charged peptides fragmented more efficiently than the analogues 2+ peptide ions. On average, triply charged peptides received a 68% higher Mascot score (24 units) than doubly charged peptides. m-NBA also increased the average charge state of phosphopeptides by up to 0.5 charge unit. The ease of implementation and the analytical benefits of charge enhancement of tryptic peptides by addition of m-NBA to the LC solvents suggest the general application of this reagent in proteomic studies that employ ETD-MS/MS and related techniques.  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