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.
An investigation of phosphate loss from sodium-cationized phosphotyrosine containing peptide ions was conducted using liquid infrared (2.94 microm) atmospheric pressure matrix-assisted laser desorption/ionization (AP MALDI) coupled to an ion trap mass spectrometer (ITMS). Previous experiments in our laboratory explored the fragmentation patterns of protonated phosphotyrosine containing peptides, which experience a loss of 98 Da under CID conditions in the ITMS. This loss of 98 Da is unexpected for phosphotyrosine, given the structure of its side chain. Phosphate loss from phosphotyrosine residues seems to be dependent on the presence of arginine or lysine residues in the peptide sequence. In the absence of a basic residue, the protonated phosphotyrosine peptides do not undergo losses of HPO(3) (Delta 80 Da) nor HPO(3) + H(2)O (Delta 98 Da) in their CID spectra. However, sodium cationized phosphotyrosine containing peptides that do not contain arginine or lysine residues within their sequences do undergo losses of HPO(3) (Delta 80 Da) and HPO(3) + H(2)O (Delta 98 Da) in their CID spectra.  相似文献   

2.
Deprotonated peptides containing C-terminal glutamic acid, aspartic acid, or serine residues were studied by sustained off-resonance irradiation collision-induced dissociation (SORI-CID) in a Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometer with ion production by electrospray ionization (ESI). Additional studies were performed by post source decay (PSD) in a matrix-assisted laser desorption ionization/time-of-flight (MALDI/TOF) mass spectrometer. This work included both model peptides synthesized in our laboratory and bioactive peptides with more complex sequences. During SORI-CID and PSD, [M - H]- and [M - 2H]2- underwent an unusual cleavage corresponding to the elimination of the C-terminal residue. Two mechanisms are proposed to occur. They involve nucleophilic attack on the carbonyl carbon of the adjacent residue by either the carboxylate group of the C-terminus or the side chain carboxylate group of C-terminal glutamic acid and aspartic acid residues. To confirm the proposed mechanisms, AAAAAD was labelled by 18O specifically on the side chain of the aspartic acid residue. For peptides that contain multiple C-terminal glutamic acid residues, each of these residues can be sequentially eliminated from the deprotonated ions; a driving force may be the formation of a very stable pyroglutamatic acid neutral. For peptides with multiple aspartic acid residues at the C-terminus, aspartic acid residue loss is not sequential. For peptides with multiple serine residues at the C-terminus, C-terminal residue loss is sequential; however, abundant loss of other neutral molecules also occurs. In addition, the presence of basic residues (arginine or lysine) in the sequence has no effect on C-terminal residue elimination in the negative ion mode.  相似文献   

3.
[M + Cu]+ peptide ions formed by matrix-assisted laser desorption/ionization from direct desorption off a copper sample stage have sufficient internal energy to undergo metastable ion dissociation in a time-of-flight mass spectrometer. On the basis of fragmentation chemistry of peptides containing an N-terminal arginine, we propose the primary Cu+ ion binding site is the N-terminal arginine with Cu+ binding to the guanidine group of arginine and the N-terminal amine. The principal decay products of [M + Cu]+ peptide ions containing an N-terminal arginine are [a(n) + Cu - H]+ and [b(n) + Cu - H]+ fragments. We show evidence to suggest that [a(n) + Cu - H]+ fragment ions are formed by elimination of CO from [b(n) + Cu - H]+ ions and by direct backbone cleavage. We conclude that Cu+ ionizes the peptide by attaching to the N-terminal arginine residue; however, fragmentation occurs remote from the Cu+ ion attachment site involving metal ion promoted deprotonation to generate a new site of protonation. That is, the fragmentation reactions of [M + Cu]+ ions can be described in terms of a "mobile proton" model. Furthermore, proline residues that are adjacent to the N-terminal arginine do not inhibit formation of [b(n) + Cu - H]+ ion, whereas proline residues that are distant to the charge carrying arginine inhibit formation of [b(n) + Cu - H]+ ions. An unusual fragment ion, [c(n) + Cu + H]+, is also observed for peptides containing lysine, glutamine, or asparagine in close proximity to the Cu+ carrying N-terminal arginine. Mechanisms for formation of this fragment ion are also proposed.  相似文献   

4.
Phosphorylation of proteins is an important post-translational protein modification in cellular response to environmental change and occurs in both prokaryotes and eukaryotes. Identification of the amino acid on individual proteins that become phosphorylated in response to extracellular stimulus is essential for understanding the mechanisms involved in the intracellular signals that these modifications facilitate. Most protein kinases catalyze the phosphorylation of proteins on serine, threonine or tyrosine. Although tyrosine phosphorylation is often the least abundant of the three major phosphorylation sites, it is important owing to its role in signal pathways. Currently available methods for the identification of phosphorylation sites can often miss low levels of tyrosine phosphorylations. This paper describes a method for the identification of phosphotyrosine-containing peptides using electrospray ionization on an ion trap mass spectrometer. Skimmer-activated collision-induced dissociation (CID) was used to generate the phosphotyrosine immonium ion at m/z 216. This method is gentle enough that the protonated molecule of the intact peptide is still observed. In-trap CID was employed for the verification of the phosphotyrosine immonium ion. Using this technique, low levels of phosphotyrosine-containing peptides can be identified from peptide mixtures separated by nanoflow micro liquid chromatography/mass spectrometry.  相似文献   

5.
An investigation of phosphate loss from phosphopeptide ions was conducted, using both atmospheric pressure matrix-assisted laser desorption/ionization (AP MALDI) and electrospray ionization (ESI) coupled to an ion trap mass spectrometer (ITMS). These experiments were carried out on a number of phosphorylated peptides in order to investigate gas phase dephosphorylation patterns associated with phosphoserine, phosphothreonine, and phosphotyrosine residues. In particular, we explored the fragmentation patterns of phosphotyrosine containing peptides, which experience a loss of 98 Da under collision induced dissociation (CID) conditions in the ITMS. The loss of 98 Da is unexpected for phosphotyrosine, given the structure of its side chain. The fragmentation of phosphoserine and phosphothreonine containing peptides was also investigated. While phosphoserine and phosphothreonine residues undergo a loss of 98 Da under CID conditions regardless of peptide amino acid composition, phosphate loss from phosphotyrosine residues seems to be dependent on the presence of arginine or lysine residues in the peptide sequence.  相似文献   

6.
Peptides and proteins may contain post-translationally modified phosphorylated amino acid residues, in particular phosphorylated serine (pSer), threonine (pThr) and tyrosine (pTyr). Following earlier work by Lehmann et al., the [M-H]- anions of peptides containing pSer and pThr functionality show loss of the elements of H3PO4. This process, illustrated for Ser (and using a model system), is CH3CONH-C(CH2OPO3H2)CONHCH(3) --> [CH3CONHC(==CH2)CONHCH3 (-OPO3H2)] (a) --> [CH3CONHC(==CH2)CONHCH3-H]- + H3PO4, a process endothermic by 83 kJ mol(-1) at the MP2/6-31++G(d,p)//HF/6-31++G(d,p) level of theory. In addition, intermediate (a) may decompose to yield CH3CONHC(==CH2)CONHCH3 + H2PO4 - in a process exothermic by 3 kJ mol(-1). The barrier to the transition state for these two processes is 49 kJ mol(-1). Characteristic cleavages of pSer and pThr are more energetically favourable than the negative ion backbone cleavages of peptides described previously. In contrast, loss of HPO3 from [M-H]- is characteristic of pTyr. The cleavage [NH2CH(CH2-C6H4-OPO3H-)CO2H] --> [NH2C(CH2-C6H4-O-)CO2H (HPO3)] (b) --> NH2CH(CH2-C6H4-O-)CO2H + HPO3 is endothermic by 318 kJ mol(-1) at the HF/6-31+G(d)//AM1 level of theory. In addition, intermediate (b) also yields NH2CH(CH2-C6H4-OH)CO2H + PO3 - (reaction endothermic by 137 kJ mol(-1)). The two negative ion cleavages of pTyr have a barrier to the transition state of 198 kJ mol(-1) (at the HF/6-31+G(d)//AM1 level of theory) comparable with those already reported for negative ion backbone cleavages.  相似文献   

7.
Amino acid profiles, established by direct infusion mass spectrometry, have been used to classify vegetable oils according to their botanical origin. The proteins present in hazelnut, sunflower, corn, soybean, olive, avocado, peanut and grapeseed oils were precipitated with acetone, and the residue was hydrolyzed in acid medium, diluted in a hydrochloric acid/ethanol mixture, and infused into the mass spectrometer. The spectra of the hydrolyzed protein extracts showed [M+H]+ ions of the following amino acids: glycine, alanine, serine, proline, valine, threonine, cysteine, isoleucine + leucine, aspartic acid, lysine, glutamic acid, methionine, histidine, phenylalanine, arginine and tyrosine. These ions were used to construct linear discriminant analysis (LDA) models. The ratios of the ion signal intensities selected by pairs were used as predictors. With the sequential application of three LDA models, the eight botanical origin categories of the samples were well resolved.  相似文献   

8.
Factors governing the molecular ion yields of amino acids and peptides have been studied using fast atom bombardment (FAB) and matrix-assisted laser desorption/ionization (MALDI) mass spectrometry (MS) in positive-ion mode. The ion yields of protonated amino acids under FAB conditions are dependent on proton affinity (PA), hydrophobicity, and aromaticity of amino acids. Both PA and hydrophobicity contribute to an increase in the ion yields, while aromaticity contributes to a decrease. In MALDI, the ion yields increase linearly with the increase of PA of amino acids with the exception of lysine. In both FAB and MALDI experiments with peptides, the presence of arginine residues is essential for producing abundant protonated peptides. In FAB, the presence of aliphatic and hydrophobic amino acids (leucine and isoleucine) increases the ion yields of protonated peptides, while some hydrophilic amino acids (aspartic acid and asparagines) decrease the ion yields. The presence of two or more arginine residues does not give higher ion yields in FAB. In MALDI, the presence of aromatic amino acids (phenylalanine and tyrosine) enhances the signals for protonated peptides. Thus, physicochemical factors of individual amino acids cooperatively affect the ion yields of protonated amino acids and peptides. These factors governing the ion yields in FAB and MALDI affect two processes, desorption and ionization, that can be considered independently.  相似文献   

9.
The complexes formed between chromium(III) and synthetic acidic peptides were studied by sustained off-resonance irradiation collision-induced dissociation (SORI-CID) in a Fourier transform ion-cyclotron resonance (FT-ICR) mass spectrometer equipped with electrospray ionization (ESI). Neutral peptides and peptides containing one, two, and multiple acidic residues were studied. Formation of [M + Cr-2H]+ occurred for all peptides. Three noteworthy features were found in the CID spectra of [M + Cr-2H]+. The first is that fewer fragment ions were produced from [M + Cr-2H]+ than from [M + H]+. The reason may be that multiple coordination between chromium(III) and carboxylate or carbonyl groups hinders the production of fragment ions by continuing to bind pieces of the peptide to chromium(III) after cleavage of bonds within the peptide. The second feature is loss of CO from [M + Cr-2H]+ and [y(n) + Cr-H]+. A mechanism involving coordination of chromium(III) with carboxylate groups is proposed to rationalize elimination of CO. The third feature is that chromium(III) is retained in all fragment ions, indicating strong binding of the metal ion to the peptides. The complex [M + 2Cr-5H]+ is formed as the peptide chain length and number of acidic residues increases. Longer peptides have more sites to coordinate with chromium(III) and more conformational flexibility. In addition, formation of [M + Cr-2H]+ from AGGAAAA-OCH(3), which has no carboxylic acid groups, suggests that chromium(III) can coordinate with sites on the peptide backbone, albeit in low abundance. In the negative mode, [M + Cr-4H](-) was only found for peptides containing four or more carboxylic acid groups. This is consistent with deprotonated carboxylic acid groups being involved in chromium(III) coordination and with chromium existing in the 3 + state in the gas-phase ions.  相似文献   

10.
迟明  毕炜  卢庄  宋丽娜  贾伟  张养军  钱小红  蔡耘 《色谱》2010,28(2):152-157
二氧化钛富集法作为目前使用最为广泛的金属氧化物富集磷酸肽的方法,在富集过程中常常对富含天冬氨酸和谷氨酸的酸性非磷酸化肽段存在一定的非特异性吸附作用。这些肽段与磷酸化肽段一同被富集,降低了磷酸肽富集的选择性。传统方法中使用的非特异性吸附抑制剂常会对质谱的电喷雾离子源造成污染,因而限制了其在液相色谱-质谱联用(LC-MS)系统中的应用。本研究将天冬氨酸作为一种新型的非特异性吸附抑制剂加入到二氧化钛富集体系中,并分别对3种和9种标准蛋白质酶切肽段混合物进行富集实验,同时与添加另一种非特异性吸附抑制剂——谷氨酸以及不添加任何非特异性吸附抑制剂的富集体系进行了富集效果的比较。结果表明,天冬氨酸可以有效地提高二氧化钛对磷酸肽富集的选择性。将添加天冬氨酸的二氧化钛富集体系应用于鼠肝全蛋白质磷酸肽的富集中,同样取得了很好的效果,表明天冬氨酸在复杂的生物样本的磷酸肽富集中也同样具有良好的应用前景。此外,由于天冬氨酸在反相色谱中极易被洗脱去除,从而避免了传统抑制剂对LC-MS系统离子源的污染问题。  相似文献   

11.
The gas-phase dissociation of phosphorylated peptides was modelled using a combination of quantum mechanics and the Rice-Ramsperger-Kassel-Marcus theory. Potential energy surfaces and unimolecular reaction rates for several low-energy fragmentation and rearrangement pathways were estimated, and a general mechanism was proposed. The neutral loss of the phosphoric acid was mainly an outcome of the intramolecular nucleophilic substitution mechanism. The mechanism involves a nucleophilic attack of the phosphorylated amino acid N-terminal carbonyl oxygen on β-carbon, yielding a cyclic five-membered oxazoline product ion. Regardless of the proton mobility, the pathway was charge directed either by a mobile proton or by a positively charged side chain of some basic residue. Although the mechanistic aspects of the phosphate loss are not influenced by the proton mobility environment, it does affect ion abundances. Results suggest that under the mobile proton environment, the interplay between phosphoric acid neutral loss product ion and backbone cleavage fragments should occur. On the other hand, when proton mobility is limited, neutral loss product ion may predominate. The fragmentation dynamics of phosphoserine versus phosphothreonine containing peptides suggests that H(3)PO(4) neutral loss from phosphothreonine containing peptides is less abundant than that from their phosphoserine containing analogs. During the low-energy CID of phosphorylated peptides in the millisecond time range, typical for ion trap instruments, a phosphate group rearrangement may happen, resulting in an interchange between the phosphorylated and the hydroxylated residues. Unimolecular dissociation rate constants imply the low abundance of such scrambled product ions.  相似文献   

12.
In the era of complete genome sequences, biochemical and medical research will focus more on the dynamic proteome of a cell. Regulation of proteins by post-translational modifications, which are not determined by the gene sequence, are already intensively studied. One example is phosphorylation of serines and threonines, probably the single most common cellular regulatory mechanism. In this paper we describe the sequencing of mono- and bisphosphorylated peptides, including identification of the phosphorylation sites, by post-source decay (PSD) in matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. In addition to dephosphorylation of the parent ions, we studied the influence of the phosphate group on the fragmentation of peptides. Generally, peptides phosphorylated on serine and threonine residues displayed no difference in their fragmentation patterns. The intensities of the resulting fragment ion signals depend only on the peptide sequence and not on either the phosphorylated amino acid or its position in the peptide chain. Phosphorylation increased the bond cleavage C-terminal to the phosphorylation site more than 10-fold, resulting in abundant signals, which typically dominated the PSD spectra. The produced C-terminally phosphorylated b-type fragment ions showed characteristic dephosphorylated fragment ions b(n) -H(3)PO(4) (-98 Da) and b(n) -HPO(3) (-80 Da) of higher abundances than the phosphorylated fragment ion. As a second layer to identify the phosphorylation site, all internally phosphorylated fragment ions were accompanied by minor, but always detectable, signals of the dephosphorylated fragment ions. Interpretation of PSD spectra of phosphopeptides was not more complicated than for unphosphorylated peptides, despite the increased number of obtained fragment ion signals.  相似文献   

13.
Phosphoric acid forms two distinct coordination compounds with manganese salts in aqueous media, a two-dimensional layered structure, [Mn(HPO4).(H2O)3], 1, under ambient conditions, and a three-dimensional synthetic mineral, [Mn5(mu-OH2)2(HPO4)2(PO4)2(H2O)2], 2, under hydrothermal procedures, at 120 degrees C. In compound 1, the oxygen atom of the doubly deprotonated phosphoric acid interconnects the metal centers to form a layered structure. The neutral hydrophilic layers of 1 are separated by 5.5 A and may potentially intercalate hydrophilic organic guest molecules. The metal centers in 2 are octahedral and bridged by PO4(3-), HPO4(2-), and OH2 groups to form a complex three-dimensional network. XPS analysis on 1 and 2 confirms that manganese exists in the +2 oxidation state. Compound 2 is a poor ion exchanger, but some improvement is observed on partial dehydration. The magnetic properties of both 1 and 2 were studied in detail to examine the amplitude of the magnetic interactions through phosphate ligand bridges. While 1 reveals dominant antiferromagnetic interactions between the spin carriers, a complete investigation of the magnetic properties of 2 revealed its true nature to be a glassy magnet.  相似文献   

14.
Collision-induced dissociation (CID) was performed on multiply deprotonated ions from three commercial peptides: hirudin (54-65), fibrinopeptide B, and oxidized insulin chain A. Ions were produced by electrospray ionization in a Fourier transform ion cyclotron resonance mass spectrometer. Each of these peptides contains multiple acidic residues, which makes them very difficult to ionize in the positive mode. However, the peptides deprotonate readily making negative ion studies a viable alternative. The CID spectra indicated that the likely deprotonation sites are acidic residues (aspartic, glutamic, and cysteic acids) and the C-terminus. The spectra are rife with c, y, and internal ions, although some a, b, x, and z ions form. Many of the fragment ions were formed from cleavage adjacent to acidic residues, both N- and C-terminal to the acidic site. In addition, neutral loss (e.g., NH3, CH3, H2O, and CO2) was prevalent from both the parent ions and from fragment ions. These neutral eliminations were often indicative of specific amino acid residues. The fragmentation patterns from several charge states of the parent ions, when combined, provide significant primary sequence information. These results suggest that negative mode CID of multiply deprotonated ions provides useful structural information and can be worthwhile for highly acidic peptides that do not form positive ions in abundance.  相似文献   

15.
Facile cleavage C‐terminal to ornithine residues in gas phase peptides has been observed and termed the ornithine effect. Peptides containing internal or C‐terminal ornithine residues, which are formed from deguanidination of arginine in solution, were fragmented to produce either a y‐ion or water loss, respectively, and the complementary b‐ion. The fragmentation patterns of several peptides containing arginine were compared to those of the ornithine analogues. Conversion of arginine to ornithine results in a decrease of the gas phase proton affinity of the residue, thereby increasing the mobility of the ionizing proton. This alteration allows the nucleophilic amine to facilitate a neighboring group reaction to induce a cleavage of the adjacent amide bond. The selective cleavage at the ornithine residue is proposed to result from the highly favorable generation of a six‐membered lactam ring. The ornithine effect was compared with the well‐known proline and aspartic acid effects in peptide fragmentation using angiotensin II, DRVYIHPF and the ornithine analogue, DOVYIHPF. Under conditions favorable to either the aspartic acid (i.e. singly protonated peptide) or proline effect (i.e. doubly protonated peptide), the ornithine effect was consistently observed to be the more favorable fragmentation pathway. The highly selective nature of the ornithine effect opens up the possibility for conversion of arginine to ornithine residues to induce selective cleavages in polypeptide ions. Such an approach may complement strategies that seek to generate non‐selective cleavages of the related peptides. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

16.
Five macrocycle-oxoanion adducts have been isolated from aqueous solutions containing 1,4,7,10,13,16-hexaazacyclooctadecane ([18]aneN(6), L) and phosphoric acid whose pH had been adjusted to selected values in the 1-8 range. Four products, (H(6)L)(H(2)PO(4))(6).2H(3)PO(4) (1), (H(6)L)(H(2)PO(4))(6) (2), (H(4)L)(H(2)PO(4))(4).2H(2)O (4), and (H(4)L)(HPO(4))(2).7H(2)O (5) crystallized from aqueous solutions at pH 1, 3, 6, and 8, respectively, while (H(4)L)(H(2)PO(4))(4) (3) crystallized on diffusion of EtOH into an aqueous reaction mixture at pH 6. Single-crystal X-ray structure determinations enabled an examination of supramolecular interactions between protonated forms of [18]aneN(6), phosphoric acid and its conjugate bases, and water of solvation. The macrocycle adopts a variety of conformations in order to accommodate the supramolecular constructs formed by the oxoanions and solvent molecules as the relative proportions of interacting species are altered. At pH 1 and 3, the fully protonated macrocycle, [LH(6)](6+), is found with six H(2)PO(4)(-) anions. At pH 6 and 8, the tetraprotonated macrocycle, [LH(4)](4+), crystallizes with four H(2)PO(4)(-) and two HPO(4)(2)(-), respectively. Variations in the solute of crystallization are evident, with phosphoric acid being present at the lowest pH and water at pH 6 and 8. In 5, the seven unique water molecules form a string-of-pearls motif within which a new heptameric isomer, consisting of a water pentamer that uses a single water to interact with the other two unique water molecules, is found. Structures 1, 2, 4, and 5 exhibit eta-3 H-bonding of ammonium protons to a single oxygen of the guest phosphates located above and below the macrocyclic ring. In 3, two phosphate oxygens of the cavity anion interact with the macrocycle, one of which participates in eta-2 H-bonding with ammonium groups.  相似文献   

17.
Positive- and negative-ion fast-atom bombardment (FAB) mass spectrometry and linked-field scan techniques at constant B/E are used to characterize phosphorylated serine, threonine, and tyrosine amino acids. Abundant molecular ions are formed for all three amino acids in both modes of ionization. The dominant fragmentation is cleavage of the phosphate ester bond with charge retention in positive-ion FAB by the amino acid backbone and in the negative-ion mode by the phosphate group. The unique feature of positive-ion FAB mass spectra of phosphoserine and -threonine is the loss, from the ion [M + H]+, of a molecule of phosphoric acid (98 Da), whereas the corresponding tyrosine expels a HPO4 (96 Da) moiety to yield a stable phenylalanine ion.  相似文献   

18.
A prototype matrix-assisted laser desorption/ionization quadrupole time-of-flight (MALDI-TOF) tandem mass spectrometer was used to sequence a series of phosphotyrosine-, phosphothreonine- and phosphoserine-containing peptides. The high mass resolution and mass accuracy of the instrument allowed the localization of one, three or four phosphorylated amino acid residues in phosphopeptides up to 3.1 kDa. Tandem mass spectra of two different phosphotyrosine peptides permitted amino acid sequence determination and localization of one and three phosphorylation sites, respectively. The phosphotyrosine immonium ion at m/z 216.04 was observed in these MALDI low-energy CID tandem mass spectra. Elimination of phosphate groups was evident from the triphosphorylated peptide but not from the monophosphorylated species. The main fragmentation pathway for the synthetic phosphothreonine-containing peptide and for phosphoserine-containing peptides derived from beta-casein and ovalbumin was the beta-elimination of phosphoric acid with concomitant conversion of phosphoserine to dehydroalanine and phosphothreonine to 2-aminodehydrobutyric acid. Peptide fragment ions of the b- and y-type allowed, in all cases, the localization of phosphorylation sites. Ion signals corresponding to (b-17), (b-18) and (y-17) fragment ions were also observed. The abundant neutral loss of phosphoric acid (-98 Da) is useful for femtomole level detection of phosphoserine-peptides in crude peptide mixtures generated by gel in situ digestion of phosphoproteins.  相似文献   

19.
Deamidation of asparagine and spontaneous isomerization of aspartic acid in proteins and peptides occur frequently. These modifications result in a mixture of peptide variants containing all three residues in the sequences. Identification and isomer quantification for these systems are challenging tasks for tandem mass spectrometry commonly utilized in protein analysis. Chromatographic data provide a set of sequence-specific information complementary to mass spectrometry. In order to compare measured retention times (RTs) with those calculated from the sequences derived from protein databases, it is necessary to develop chromatographic models and tools allowing the prediction of RT and elution order for peptides with modified residues. In this work we extended recently introduced critical liquid chromatography of biomacromolecule model for prediction of RTs for peptides containing asparagines, aspartic acid, and isoaspartic acid residues.  相似文献   

20.
Formation of S-carbamidomethylmethionine (camMet) occurs as a side reaction during cysteine alkylation with iodoacetamide (IAA). In collision-induced dissociation, peptides with camMet show an abundant neutral loss of 2-(methylthio)acetamide (C3H7NOS = 105.025 Da) at moderate collision offset values which are similar to those optimal for loss of phosphoric acid (H3PO4 = 97.977 Da). Neutral loss analysis is used for spotting of phosphopeptides which contain phosphoserine (pSer) or phosphothreonine (pThr) residues. In the case where precursor ions cannot be accurately assigned in the survey spectrum (e.g. due to low ion abundance or signal overlap), the mass accuracy of a neutral loss tandem mass spectrometry (MS/MS) analysis depends on the precursor ion isolation window. For the charge states 2+, 3+ or 4+, a typical 3.5 Da precursor isolation window results in neutral loss windows of 7, 10.5 or 14 Da, respectively. Consequently, neutral loss of 105 Da from alkylated methionine residues can mimic the phosphoserine/phosphothreonine-specific neutral loss of 98 Da. In the evaluation of quadrupole time-of-flight (QTOF) parent ion scan data for neutral loss of H3PO4, this interference was frequently observed. It is illustrated in this study using the analysis of ovalbumin phosphorylation as an example. The +80 Da molecular weight shift connected with phosphorylation at serine or threonine may also be mimicked by carbamidomethylation of methionine through a combination with sodium adduction (+57 Da +22 Da = +79 Da). For highly sensitive neutral loss analysis of serine and threonine phosphorylation, careful data inspection is recommended if reduction and alkylation by IAA is employed.  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