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 78 毫秒
1.
The α-ketoacid–hydroxylamine (KAHA) ligation with 5-oxaproline enables the direct cyclization of peptides upon cleavage from a solid support, without coupling reagents, protecting groups, or purification of the linear precursors. This Fmoc SPPS-based method was applied to the synthesis of a library of 24 homoserine-containing cyclic peptides and was compared side-by-side with the synthesis of the same products using a standard method for cyclizing side-chain protected substrates. A detailed mechanistic study including 2H and 18O labeling experiments and the characterization of reaction intermediates by NMR and mass spectrometry is reported.  相似文献   

2.
A new base-catalyzed elimination reaction employing the hindered, non-nucleophilic bases tetramethylguanidine or 1,8-diazabicyclo-[5.4.0.]undec-7-ene has been developed for the removal of protected peptides from a 2-[4-(hydroxymethyl)phenyl-acetoxy]propionyl-resin. The proposed reaction mechanism involved cleavage of the ester bond between the peptide and resin via a base-catalyzed elimination. The protected peptide-resin cleavage reaction is mild, rapid and proceeds in good yield with a very simple work- up procedure. Four protected peptide-resins varying in size from seven to sixteen residues were prepared using the 2-[4-(hydroxymethyl)phenyl-acetoxy]- propionyl-resin and then cleaved in the protected form to demonstrate the utility of the new cleavage technique. The protected peptide cleavage products can be used in the preparation of larger peptides by fragment condensation.  相似文献   

3.
This paper reports the study of backbone cleavages in the collision-induced negative-ion mass spectra of the [M - H](-) anions of some synthetic modifications of the bioactive amphibian peptide citropin 1 (GLFDVIKKVASVIGGL-NH(2)). The peptides chosen for study contain no amino acid residues which could effect facile side-chain cleavage, i.e. Ser (-CH(2)O, side-chain cleavage) and Asp (-H(2)O) are replaced by Ala or Lys. We expected that such peptides should exhibit standard and pronounced peaks due to alpha cleavage ions (and to a lesser extent beta cleavage ions) in their collision-induced negative-ion spectra. This expectation was realised, but the spectra also contained peaks formed by a new series of cleavage anions. These are produced following cyclisation of the C-terminal CONH(-) moiety at carbonyl functions of amide groups along the peptide backbone; effectively transferring the NH of the C-terminal CONH(-) group to other amino acid residues. We have called the product anions of these processes beta' ions, in order to distinguish them from standard beta ions. Some beta' ions also fragment directly to some other beta' ions of smaller mass. The reaction coordinates of alpha,beta and beta' backbone processes have been calculated at the HF/6-31G*//AM1 level theory for simple model systems. The initial cyclisation step of the beta' sequence is barrierless and exothermic. Subsequent steps have a maximum barrier of +40 kcal mol(-1), with the overall reaction being endothermic by some 30 kcal mol(-1) at the level of theory used. These calculations take no account of the complexity of the conformationally flexible peptide system, and it is surprising that each of the two reacting centres can 'find' each other in such a large system.  相似文献   

4.
Here we examined the fragmentation, on a quadrupole ion-trap mass spectrometer, of the protonated ions of a group of peptides containing one arginine and two different acidic amino acids, one being aspartic acid (Asp) or glutamic acid (Glu) and the other being cysteine sulfinic acid [C(SO2H)] or cysteine sulfonic acid [C(SO3H)]. Our results showed that, upon collisional activation, the cleavage of the peptide bond C-terminal to C(SO2H) is much more facile than that of the peptide bond C-terminal to Asp, Glu, or C(SO3H). There is no significant difference, however, in susceptibility to cleavage of peptide bonds that are C-terminal to Asp, Glu, and C(SO3H). To understand these experimental observations, we carried out B3LYP/6-31G* density functional theory calculations for a model cleavage reaction of GXG --> b2 + Gly, in which X is Asp, Glu, C(SO2H), or C(SO3H). Our calculation results showed that the cleavage reaction is thermodynamically more favorable when X = C(SO2H) than when X = Asp or C(SO3H). We attributed the less facile cleavage of the amide bond after Glu to that the formation of a six-membered ring b ion for Glu-bearing peptides is kinetically not as favorable as the formation of a five-membered ring b ion for peptides containing the other three acidic amino acids. The results from this study may provide useful tools for peptide sequencing.  相似文献   

5.
The hydroperoxide anion cleaves unactivated amides and peptides although it is completely unreactive toward ethyl esters. The cleavage by HO2(-) proceeds faster than by OH(-) and involves additional routes with general acid assistance by H2O2 and general base assistance by OH(-) and HO2(-). Cleavage of polypeptides occurs at the N-terminal peptide bond. [reaction: see text]  相似文献   

6.
An on-line electrochemistry/electrospray mass spectrometry system (EC/MS) is described that allows fast analysis of the oxidation products of peptides. A range of peptides was oxidized in an electrochemical cell by application of a potential ramp from 0 to 1.5 V during passage of the sample. Electrochemical oxidation of peptides was found to occur readily when tyrosine was present. Tyrosine was found to be oxidized between 0.5 and 1.0 V to various oxidation products, including peptide fragments formed by hydrolysis at the C-terminal side of tyrosine. The results confirm earlier knowledge on the mechanisms and reaction products of chemical and electrochemical peptide oxidation. Methionine residues are also readily oxidized, but do not induce peptide cleavage. At potentials higher than about 1.1 V, additional oxidation products were observed in some peptides, including loss of 28 Da from the C-terminus and dimerization. The tyrosine-specific cleavage reaction suggests a possible use of the EC/MS system as an on-line protein digestion and peptide mapping system. In addition, the system can be used to distinguish phosphorylated from unphosphorylated tyrosine residues. Four forms of the ZAP-70 peptide ALGADDSYYTAR with both, either or neither tyrosine phosphorylated were subjected to a 0-1.5 V potential ramp. Oxidation of, and cleavage adjacent to, tyrosine was observed exclusively at unphosphorylated tyrosine residues.  相似文献   

7.
The specificity of three neutral endopeptidases toward several biologically active peptides was determined by combined high-performance liquid chromatography and amino acid analysis of the degradation products. Incubation mixtures were chromatographed on a reversed-phase column equilibrated with a mixture of acetonitrile and potassium phosphate buffer (0.05 M; pH 2.0). Reaction products were eluted with a linear gradient of acetonitrile and the absorbance of the effluent monitored at 210 nm. Fractions corresponding to discrete peaks were subjected to quantitative amino acid analysis. The peptide bond undergoing cleavage is readily assigned from the knowledge of the primary structure of the peptide and the amino acid composition of the reaction products.  相似文献   

8.
The reaction of [Mn(TF(4)TMAP)](CF(3)SO(3))(5) (TF(4)TMAP=meso-tetrakis(2,3,5,6-tetrafluoro-N,N,N-trimethyl-4-aniliniumyl)porphinato dianion) with H(2)O(2) (2 equiv) at pH 10.5 and 0 degrees C yielded an oxomanganese(V) porphyrin complex 1 in aqueous solution, whereas an oxomanganese(IV) porphyrin complex 2 was generated in the reactions of tert-alkyl hydroperoxides such as tert-butyl hydroperoxide and 2-methyl-1-phenyl-2-propyl hydroperoxide. Complex 1 was capable of epoxidizing olefins and exchanging its oxygen with H(2) (18)O, whereas 2 did not epoxidize olefins. From the reactions of [Mn(TF(4)TMAP)](5+) with various oxidants in the pH range 3-11, the O-O bond cleavage of hydroperoxides was found to be sensitive to the hydroperoxide substituent and the pH of the reaction solution. Whereas the O-O bond of hydroperoxides containing an electron-donating tert-alkyl group is cleaved homolytically, an electron-withdrawing substituent such as an acyl group in m-chloroperoxybenzoic acid (m-CPBA) facilitates O-O bond heterolysis. The mechanism of the O-O bond cleavage of H(2)O(2) depends on the pH of the reaction solution: O-O bond homolysis prevails at low pH and O-O bond heterolysis becomes a predominant pathway at high pH. The effect of pH on (18)O incorporation from H(2) (18)O into oxygenated products was examined over a wide pH range, by carrying out the epoxidation of carbamazepine (CBZ) with [Mn(TF(4)TMAP)](5+) and KHSO(5) in buffered H(2) (18)O solutions. A high proportion of (18)O was incorporated into the CBZ-10,11-oxide product at all pH values but this proportion was not affected significantly by the pH of the reaction solution.  相似文献   

9.
Complicated electropherograms are produced in the separation of fluorescently labeled peptides. Incomplete labeling of epsilon-amino groups on lysine residues results in the production of 2n-1 reaction products, where n is the number of alpha and epsilon amino groups in the peptide. A single label is attached to the peptide by first taking the peptide through one cycle of the Edman degradation reaction. All epsilon-amino groups are converted to the phenyl thiocarbamyl and the cleavage step exposes one alpha-amino group at the N-terminus of the peptide; the fluorescent label is attached to the N-terminus.  相似文献   

10.
Elution time shifts between 33 different peptides and their corresponding phosphopeptides ranging from 4 amino acid residues to 35 amino acids in length were systematically investigated using high-resolution reversed-phase liquid chromatography (RPLC)-tandem mass spectrometry (MS/MS) analysis with trifluoroacetic acid as the ion pairing agent. Observed peptide elution time shifts for a single phosphorylation ranged from -5.28 min (for pYVPML) to +0.59 min (for HRDpSGLLDSLGR). Peptides containing a phosphotyrosine residue displayed a significant decrease in elution time following phosphorylation compared to their similar-sized peptides with phosphoserine or phosphothreonine residues. While peptide phosphorylation generally led to a decrease in the observed elution time, five peptides displayed increased elution times as a result of phosphorylation. For large peptides (> or =18 amino acids), the elution time shifts due to single phosphorylation were limited (ranging between -0.48 and +0.03 min), while the elution time shifts for small peptides (<18 amino acids) were characterized by a larger deviation (ranging between -5.28 and +0.59 min). The predictive capability for the observed RPLC elution time change due to phosphorylation has been suggested, which will aid in assigning confident phosphopeptide identifications and their subsequent confirmation.  相似文献   

11.
A method is described for the identification and relative quantification of proteomes using accurate mass tags (AMT) generated by nLC-dual ESI-FT-ICR-MS on a 7T instrument in conjunction with stable isotope labeling using 16O/18O ratios. AMTs were used for putative peptide identification, followed by confirmation of peptide identity by tandem mass spectrometry. For a combined set of 58 tryptic peptides from bovine serum albumin (BSA) and human transferrin, a mean mass measurement accuracy of 1.9 ppm +/-0.94 ppm (CIM99%) was obtained. This subset of tryptic peptides was used to measure 16O/18O ratios of 0.36 +/- 0.09 (CIM99%) for BSA (micro = 0.33) and 1.48 +/- 0.47 (CIM99%) for transferrin (micro = 1.0) using a method for calculating 16O/18O ratios from overlapping isotopic multiplets arising from mixtures of 16O, 18O1, and 18O2 labeled C-termini. The model amino acid averagine was used to calculate a representative molecular formula for estimating and subtracting the contributions of naturally occurring isotopes solely as a function of peptide molecular weight. The method was tested against simulated composite 16O/18O spectra where peptide molecular weight, 16O/18O ratio, 18O1/18O2 ratios, and number of sulfur atoms were varied. Relative errors of 20% or less were incurred when the 16O/18O ratios were less than three, even for peptides where the number of sulfur atoms was over- or under-estimated. These data demonstrate that for biomarker discovery, it is advantageous to label the proteome representing the disease state with 18O; and the method is not sensitive to variations in 18O1/18O2 ratio. This approach allows a comprehensive differentiation of expression levels and tentative identification via AMTs, followed by targeted analysis of over- and under-expressed peptides using tandem mass spectrometry, for applications such as the discovery of disease biomarkers.  相似文献   

12.
A potential pitfall in 18O-based N-linked glycosylation site mapping   总被引:1,自引:0,他引:1  
A common procedure for identifying N-linked glycosylation sites involves tryptic digestion of the glycoprotein, followed by the conversion of glycosylated asparagine residues into (18)O-labeled aspartic acids by PNGase F digestion in (18)O water. The 3 Da mass tag created by this process is readily observable by liquid chromatography/tandem mass spectrometry (LC/MS/MS) analysis, and is often used to identify the sites of N-linked glycosylation. While using this procedure, we noticed that 60% of the asparagines identified as being glycosylated were not part of the consensus sequence required for N-linked glycosylation, and thus were not biologically possible. Investigation into the source of this unacceptably high false positive rate demonstrated that even after reversed-phase cleanup and heat denaturation, the trypsin used for proteolysis was still active and led to the incorporation of (18)O into the C-termini of the peptides during the deglycosylation step. The resulting mass shift accounted for most of the false positive sites, as the database search algorithm confused it with an (18)O-labeled Asp residue near the C-terminus of a peptide. This problem can be overcome by eliminating trypsin from the solution prior to performing the deglycosylation process, by resuspending the peptides in natural abundance water following deglycosylation, or by allowing (18)O incorporation into the C-terminus as a variable modification during the database search. These methods have been demonstrated on a model protein, and are applicable to the analyses of glycoproteins that are digested with trypsin or another serine protease prior to enzymatic release of the carbohydrate side chains. This study should alert investigators in the field to this potential and unexpected pitfall and provide strategies to overcome this phenomenon.  相似文献   

13.
Incorporation of omega-amino acids into peptide sequences plays an important role in designing peptides with modified backbone conformation and enhanced stability against proteolysis. The present study establishes the presence of unusual turns involving 12-membered hydrogen bonded rings in terminally blocked tri- and tetrapeptides. X-ray diffraction analysis of single crystals and NMR studies have been used to probe the three-dimensional structures of two terminally protected short peptides, Boc-gamma-Abu(1)-Aib(2)-Ala(3)-OMe 1 and Boc-gamma-Abu(1)-Aib(2)-Ala(3)-Aib(4)-OMe 2 (gamma-Abu = gamma-aminobutyric acid), in which conformationally flexible omega-amino acids (gamma-Abu) and conformationally restricted alpha-aminoisobutyric acid (Aib) residues are positioned contiguously. The crystal structures of both peptides 1 and 2 exhibit unusual turns composed of 12-membered hydrogen bonded rings involving C [double bond] O from the Boc-group and Ala(3) NH. A type I' beta-turn was observed in the structure of peptide 2 adjacent to the unusual turn with a hydrogen bond between gamma-Abu(1) C [double bond] O and Aib(4) NH. The crystals of peptide 1 are in the space group P2(1), a = 9.3020(10) A, b = 23.785(2) A, c = 10.022(3) A, beta = 101.35 degrees(4), Z = 4, R = 5.7%, and R(w) = 14.5%. Similarly, the crystals of peptide 2 are in the space group C2, a = 19.0772(6) A, b = 8.7883(2) A, c = 16.7758(3) A, beta = 110.7910 degrees(10), Z = 4, R = 6.71%, and R(w) = 15.11%. The unusual turn in both peptides 1 and 2 are retained in solution as is evident from NMR studies in CDCl(3). The role of the adjacently located Aib residue to nucleate the 12-membered hydrogen bonded ring is also addressed.  相似文献   

14.
以2709碱性蛋白酶酶解牛血清白蛋白为研究对象,采用液相色谱-飞行时间质谱联用分析方法和位点非特异性酶切肽谱鉴定方法,分析水解过程中肽谱的动态变化。蛋白电泳分析结果显示,2709碱性蛋白酶在0.1%的添加比例条件下,0.5 h内可将全部蛋白降解,表明其具有极强的水解能力。肽谱分析结果显示,该酶酶解BSA蛋白过程中肽谱变化复杂多样,不同序列多肽具有不同的动态变化特征。酶切位点分析结果显示,2709碱性蛋白酶几乎能在所有种类的氨基酸位点发生酶切,但酶切的频率并不相同,这表明该酶在水解位点上具有宽泛的选择性和一定的倾向性,其中在肽键C端氨基酸种类的选择上具有明显的亮氨酸倾向性。该文可为研究其他工业蛋白酶水解肽谱的规律提供借鉴,并有助于提高我国工业蛋白酶制剂的应用水平和蛋白质加工水平。  相似文献   

15.
Peptide and protein lipoxidation is a deleterious process which has been related to several degenerative conditions. In the present study, the interaction of lipid secondary oxidation products with peptides was investigated by evaluating the modifications occurring to angiotensin II (Ang-II) in the presence of an oxidizing polyunsaturated glycerophospholipid (1-palmitoyl-2-arachidonoyl-glycerophosphatidylcholine, PAPC). PAPC oxidation was promoted by Fenton chemistry and the oxidation products were incubated with Ang-II. The reaction products were finally analysed by off-line nanospray high-performance liquid chromatography/matrix-assisted laser desorption/ionization tandem mass spectrometry (nano-HPLC/MALDI-TOF-MS/MS). Ang-II was found to form adducts with 26 different aldehydes, leading to 37 distinct reaction products. Modification of Ang-II occurred through reaction with reactive carbonyl species (RCS) originating from fatty acyl chain cleavage, while interactions with the oxidized phospholipid could not be detected. Adduction was observed to occur both by Michael and Schiff base mechanisms, most prevalently taking place at the peptide N-terminus or the arginine residue. Histidine modification could only be demonstrated to occur via Michael addition with two aldehydes: 4-hydroxy-2-nonenal (HNE) and 2-octenal. The highly reactive 4-oxo-2-nonenal (ONE) was shown to react preferentially with the arginine side chain, while malondialdehyde addition could only be confirmed at the N-terminus. Aspartic acid oxidative decarboxylation, amino acid side chain oxidation, multiple adduction or peptide cross-links could not be perceived. The inability to detect these reaction products is indicative of their low abundance or non-existence in competitive reaction conditions. The multiplicity of peptide modifications described emphasizes the complexity of lipoxidation, the effects of which are not possible to fully understand by the evaluation of independent reaction products.  相似文献   

16.
Disulfide bonds are a post-translational modification (PTM) that can be scrambled or shuffled to non-native bonds during recombinant expression, sample handling, or sample purification. Currently, mapping of disulfide bonds is not easy because of various sample requirements and data analysis difficulties. One step towards facilitating this difficult work is developing a better understanding of how disulfide-bonded peptides fragment during collision induced dissociation (CID). Most automated analysis algorithms function based on the assumption that the preponderance of product ions observed during the dissociation of disulfide-bonded peptides result from the cleavage of just one peptide bond, and in this report we tested that assumption by extensively analyzing the product ions generated when several disulfide-bonded peptides are subjected to CID on a quadrupole time of flight (QTOF) instrument. We found that one of the most common types of product ions generated resulted from two peptide bond cleavages, or a double cleavage. We found that for several of the disulfide-bonded peptides analyzed, the number of double cleavage product ions outnumbered those of single cleavages. The influence of charge state and precursor ion size was investigated, to determine if those parameters dictated the amount of double cleavage product ions formed. It was found in this sample set that no strong correlation existed between the charge state or peptide size and the portion of product ions assigned as double cleavages. These data show that these ions could account for many of the product ions detected in CID data of disulfide bonded peptides. We also showed the utility of double cleavage product ions on a peptide with multiple cysteines present. Double cleavage products were able to fully characterize the bonding pattern of each cysteine where typical single b/y cleavage products could not.  相似文献   

17.
In this study, bovine beta-lactoglobulin A (beta-Lg A) was hydrolysed with thermolysin under non-denaturing and heat-denaturing conditions. The peptides released during hydrolysis were identified by HPLC-MS/MS. A total of 25 peptides were identified in the hydrolysate obtained at 37 degrees C for 5 min. Some of these peptides survived to further proteolysis even at higher incubation temperatures. Furthermore, novel cleavage sites localised in the most buried zones of beta-Lg and available for thermolysin were recognised when the incubation temperature increased in the range between 60 and 80 degrees C. Three new peptides, LDA, LKPTPEGD, and LQKW, appeared after 30 min hydrolysis at these incubation temperatures, but they were not identified in the 30-min hydrolysates obtained at 37 and 50 degrees C. Of special interest was the peptide LQKW, corresponding to the fragment f(58-61) that had been previously described as a potent angiotensin-converting enzyme-inhibitor (IC50 value of 34.7 microM).  相似文献   

18.
Phosphonamidate peptides were prepared in good yields by reaction of ethoxy(phenyl)phosphinate with free peptides in a mixed solution of H2O, C2H5OH, Et3N and CCl4 at room temp.; their in situ electrospray ionization mass spectra exhibited high sensitivity compared with the free peptides, and sequential loss of amino acid residues of the sodiated phosphonamidate peptides from the C-terminus was found in multistage ESI mass spectrometry. The results show that N-phosphonyl derivatization combined with multistage ESI-MS is a powerful method for peptide sequencing.  相似文献   

19.
The mature form of green fluorescent protein (GFP) is generated by a spontaneous self-modification process that is essentially irreversible. A key step in chromophore biosynthesis involves slow air oxidation of an intermediate species, in which the backbone atoms of residues 65-67 have condensed to form a five-membered heterocycle. We have investigated the kinetics of hydrogen peroxide evolution during in vitro GFP maturation and found that the H2O2 coproduct is generated prior to the acquisition of green fluorescence at a stoichiometry of 1:1 (peroxide/chromophore). The experimental progress curves were computer-fitted to a three-step mechanism, in which the first step proceeds with a time constant of 1.5 (+/-1.1) min and includes protein folding and peptide cyclization. Kinetic data obtained by HPLC analysis support a rapid cyclization reaction that can be reversed upon acid denaturation. The second step proceeds with a time constant of 34.0 (+/-1.5) min and entails rate-limiting protein oxidation, as supported by a mass loss of 2 Da observed for tryptic peptides derived from species that accumulate during the reaction. The final step in GFP maturation proceeds with a time constant of 10.6 (+/-1.2) min, suggesting that this step may contribute to overall rate retardation. We propose that under highly aerobic conditions, the dominant reaction path follows a cyclization-oxidation-dehydration mechanism, in which dehydration of the heterocycle is facilitated by slow proton abstraction from the Tyr66 beta-carbon. In combination, the results presented here suggest a role for molecular oxygen in trapping the cyclized form of GFP.  相似文献   

20.
The specificity and rate of cleavage of adrenocorticotrophic hormone (ACTH) peptide bonds by endoproteinase Arg C were analyzed using capillary electrophoresis (CE) and reversed-phase (C18) high-performance liquid chromatography (HPLC). Acidic cleavage products were readily resolved by CE in uncoated capillaries using low ionic strength electrolytes. However, products predicted to have a net positive charge greater than 2 or more than 4 positively charged groups per peptide did not migrate out from the capillary at low ionic strength. Addition of salts and zwitterions to the electrolyte decreased capillary-peptide interactions such that all of the ACTH peptides examined were eluted with high efficiency separation by CE. Commercially obtained endoproteinase Arg C preparations exhibited peptidase activity at Lys-15-Lys16 and at Lys16-Arg17 in addition to the expected cleavage at Arg-X bonds. ACTH peptide bond cleavage rates for Arg8-Trp9, Arg17-Arg-18, Lys15-Lys16, and Lys16-Arg17 were 1.46, 0.096, 0.57, and 0.029 mumol min-1 mg-1 respectively. CE separations generally exhibited better resolution and were accomplished in shorter times than C18 HPLC separations. These properties make CE a particularly appropriate method for kinetic analysis of proteolytic enzyme action on peptide substrates.  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