首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
To examine the role of both Arg residues at positions 5 and 7 of dog neuromedin U-8 (d-NMU-8; pGlu1-Phe-Leu-Phe-Arg5-Pro-Arg7-Asn8-NH2) for smooth muscle contractile activity on isolated chicken crop, d-NMU-8 analogs were synthesized where either Arg residue was systematically replaced by various amino acids [X: Ala, Thr, Glu, Gln, Lys, Orn, His, citrulline (Cit) or homoarginine (Har)]. All [X5]-d-NMU-8, except for [Glu5]- and [Des-Arg5]-d-NMU-8, were full agonists, although their affinities to NMU receptors were decreased. No [X7]-d-NMU-8 showed contractile activity even at concentrations of 10(-5) mol/l, except for [Har7]-d-NMU-8, which retained weak biological activity. These analogs had no antagonistic activity against porcine neuromedin U-8 (p-NMU-8). The results revealed that Arg7 of d-NMU-8 is indispensable for receptor binding and activation to induce smooth muscle contraction, and the guanidino group of the side chain at position 7, but not at position 5, is strictly recognized by NMU receptors in the chicken crop.  相似文献   

2.
The surface-induced dissociation (SID) of six model peptides containing either methionine sulfoxide or aspartic acid (GAILM(O)GAILR, GAILM(O)GAILK, GAILM(O)GAILA, GAILDGAILR, GAILDGAILK, and GAILDGAILA) have been studied using a specially configured Fourier transform ion-cyclotron resonance mass spectrometer (FT-ICR MS). In particular, we have investigated the energetics and dynamics associated with (i) preferential cleavage of the methionine sulfoxide side chain via the loss of CH3SOH (64 Da), and (ii) preferential cleavage of the amide bond C-terminal to aspartic acid. The role of proton mobility in these selective bond cleavage reactions was examined by changing the C-terminal residue of the peptide from arginine (nonmobile proton conditions) to lysine (partially mobile proton conditions) to alanine (mobile proton conditions). Time- and energy-resolved fragmentation efficiency curves (TFECs) reveal that selective cleavages due to the methionine sulfoxide and aspartic acid residues are characterized by slow fragmentation kinetics. RRKM modeling of the experimental data suggests that the slow kinetics is associated with large negative entropy effects and these may be due to the presence of rearrangements prior to fragmentation. It was found that the Arrhenius pre-exponential factor (A) for peptide fragmentations occurring via selective bond cleavages are 1-2 orders of magnitude lower than nonselective peptide fragmentation reactions, while the dissociation threshold (E0) is relatively invariant. This means that selective bond cleavage is kinetically disfavored compared to nonselective amide bond cleavage. It was also found that the energetics and dynamics for the preferential loss of CH3SOH from peptide ions containing methionine sulfoxide are very similar to selective C-terminal amide bond cleavage at the aspartic acid residue. These results suggest that while preferential cleavage can compete with amide bond cleavage energetically, dynamically, these processes are much slower compared to amide bond cleavage, explaining why these selective bond cleavages are not observed if fragmentation is performed under mobile proton conditions. This study further affirms that fragmentation of peptide ions in the gas phase are predominantly governed by entropic effects.  相似文献   

3.
We examined the fragmentation of the electrospray-produced [M-H]- and [M-2H]2- ions of a number of peptides containing two acidic amino acid residues, one being aspartic acid (Asp) or glutamic acid (Glu), and the other being cysteine sulfinic acid [C(SO2H)] or cysteine sulfonic acid [C(SO3H)], on an ion-trap mass spectrometer. We observed facile neutral losses of H2S and H2SO2 from the side chains of cysteine and C(SO2H), respectively, whereas the corresponding elimination of H2SO3 from the side chain of C(SO3H) was undetectable for most peptides that we investigated. In addition, the collisional activation of the [M-H]- ions of the C(SO2H)-containing peptides resulted in the cleavage of the amide bond on the C-terminal side of the C(SO2H) residue. Moreover, collisional activation of the [M-2H]2- ions of the above Asp-containing peptides led to the cleavage of the backbone N-Calpha bond of the Asp residue to give cn and/or its complementary [zn-H2O] ions. Similar cleavage also occurred for the singly deprotonated ions of the otherwise identical peptides with a C-terminal amide functionality, but not for the [M-H]- ions of same peptides with a free C-terminal carboxylic acid. Furthermore, ab initio calculation results for model cleavage reactions are consistent with the selective cleavage of the backbone N-Calpha bond in the Asp residue.  相似文献   

4.
The X-Pro peptide bond (in which X represents any amino acid residue) in peptides and proteins is resistant to cleavage by most proteolytic enzymes. We show that [Pd(H(2)O)(4)](2+) ion can selectively hydrolyze this tertiary peptide bond within the X-Pro-Met and X-Pro-His sequence segments. The hydrolysis requires an equimolar amount of the Pd(II) reagent and occurs under mild conditions-at temperature as low as 20 degrees C (with half-life of 1.0 h at pH 2.0) and at pH as high as 7.0 (with half-life of 4.2 h at pH 7.0 and 40 degrees C). The secondary peptide bond, exemplified by X-Gly in the X-Gly-Met and X-Gly-His sequence segments, however, is cleaved only in weakly acidic solution (pH < 4.0) and more slowly (half-life is 4.2 h at pH 2.0 and 60 degrees C). We explain the sequence-specificity of X-Pro cleavage by NMR spectroscopic analysis of the coordination of the X-Pro-Met segment to the Pd(II) ion. We give indirect evidence for the mechanism of cleavage by analyzing the conformation of the scissile X-Pro peptide bond, and by comparing the rate constants for the cleavage of the tertiary X-Pro peptide bond, the tertiary X-Sar peptide bond (Sar is N-methyl glycine), and the typical secondary X-Gly peptide bond in a set of analogous oligopeptides. Methionine and histidine side chains provide the recognition by selectively binding (anchoring) the Pd(II) ion. The proline residue provides the enhanced activity because its tertiary X-Pro peptide bond favors the cleavage-enhancing binding of the Pd(II) ion to the peptide oxygen atom and prevents the cleavage-inhibiting binding of the Pd(II) ion upstream of the anchoring (histidine or methionine) residue. Cleavage can be switched from the residue-selective to the sequence-specific mode by simply adjusting the pH of the aqueous solution. In acidic solutions, any X-Y bond in X-Y-Met and X-Y-His segments is cleaved because the cleavage is directed by anchoring methionine and histidine residues. In mildly acidic and neutral solutions, only the X-Pro bond in X-Pro-Met and X-Pro-His sequences is cleaved because of an interplay between the anchoring residue and the proline residue preceding it. Because Pro-Met and Pro-His sequences are rare in proteins, this sequence-specific cleavage is potentially useful for the removal of the fusion tags from the bioengineered fusion proteins.  相似文献   

5.
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   

6.
The oxidation of the sulfhydryl group in cysteine to sulfenic acid, sulfinic acid, and sulfonic acid in proteins is important in a number of enzymatic processes. In this study we examined the fragmentation of four peptides containing cysteine, cysteine sulfinic acid (Cys-SO(2)H), and cysteine sulfonic acid (Cys-SO(3)H) in an ion-trap mass spectrometer. Our results show that the presence of a Cys-SO(2)H in a peptide leads to preferential cleavage of the amide bond at the C-terminal side of the oxidized cysteine residue. The results are important for the determination of the site of the cysteine oxidation and might be useful for the sequencing of cysteine-containing peptides.  相似文献   

7.
A molecular dynamics (MD) simulation study has been carried out to understand the stability of the triple helical collagen models. The calculations show that the presence of the aspartic acid residue in different positions leads to the local variation in the structure. Analyses of root-mean-square deviation (RMSD), radial distribution function (RDF), puckering effect, dihedral angle variation, hydrogen bond (H-bond), and conformational changes during molecular dynamics simulation reveal that the local perturbation in the sequences, increase in chain flexibility due to removal of five membered rings in the collagen by aspartic acid, change of intermolecular H-bonding pattern, and differences in the association of water are mainly influencing the nature of stabilization of collagen by aspartic acid.  相似文献   

8.
As part of a series of peptides designed to have binding ability selective for each of the nucleic acid bases, five tripeptides consisting of N-acetyl-Trp-X-Trp-NHCH3 (X = Gly, Asn, Asp, Gln and Glu) were synthesized, and their abilities to form complexes with four different nucleotides were examined by the fluorescence and phase distribution methods. The association constants obtained indicated that, depending on the sort of X residue, the peptides showed a variation in their interaction with guanosine monophosphate (GMP), while no noticeable selectivity was observed for other nucleotides adenosine monophosphate (AMP), uridine monophosphate (UMP) and cytidine monophosphate (CMP). The binding mode of N-acetyl-Trp-Asp-Trp-NHCH3 for the guanine base was further investigated using the proton nuclear magnetic resonance (1H-NMR) method. The mode was suggested to involve intimate cooperation of (1) the hydrogen bond formation between the carboxyl group of the Asp side chain and the guanine C2-amino group, and (2) the stacking interaction of the base with two terminal Trp residues of the peptide. Such interaction was strengthened by the protonation of the guanine base. A tentative binding mode is proposed based on these results.  相似文献   

9.
A novel cysteine derivative, N(alpha)-trityl-S-(9H-xanthen-9-yl)-l-cysteine [Trt-Cys(Xan)-OH] has been introduced for peptide synthesis, specifically for application to a new strategy for the preparation of cyclic peptides. The following steps were carried out to synthesize the cyclic model peptide cyclo(Cys-Thr-Abu-Gly-Gly-Ala-Arg-Pro-Asp-Phe): (i). side-chain anchoring of Fmoc-Asp-OAl via its free beta-carboxyl as a p-alkoxybenzyl ester to a solid support; (ii). stepwise chain elongation of the peptide by standard Fmoc/tBu solid-phase chemistry; (iii). removal of the N-terminal Fmoc group; (iv). coupling of Trt-Cys(Xan)-OH; (v). selective Pd(0)-promoted cleavage of the C-terminal allyl ester; (vi). coupling of the C-terminal residue, i.e., H-Phe-SBzl, preactivated as a thioester; (vii). selective removal of the N(alpha)-Trt and S-Xan protecting groups under very mild acid conditions; (viii). on-resin cyclization by native chemical ligation in an aqueous milieu; and (ix). final acidolytic cleavage of the cyclic peptide from the resin. The strategy was evaluated for three supports: poly[N,N-dimethacrylamide-co-poly(ethylene glycol)] (PEGA), cross-linked ethoxylate acrylate resin (CLEAR), and poly(ethylene glycol)-polystyrene (PEG-PS) graft resin supports. For PEGA and CLEAR, the desired cyclic product was obtained in 76-86% overall yield with initial purities of approximately 70%, whereas for PEG-PS (which does not swell nearly as well in water), results were inferior. Solid-phase native chemical ligation/cyclization methodology appears to have advantages of convenience and specificity, which make it promising for further generalization.  相似文献   

10.
Fragmentation of radical cationic peptides [R(G)(n-2)X(G)(7-n)]˙(+) and [R(G)(m-2)XG]˙(+) (X = Phe or Tyr; m = 2-5; n = 2-7) leads selectively to a(n)(+) product ions through in situ C(α)-C peptide backbone cleavage at the aromatic amino acid residues. In contrast, substituting the arginine residue with a less-basic lysine residue, forming [K(G)(n-2)X(G)(7-n)]˙(+) (X = Phe or Tyr; n = 2-7) analogs, generates abundant b-y product ions; no site-selective C(α)-C peptide bond cleavage was observed. Studying the prototypical radical cationic tripeptides [RFG]˙(+) and [KFG]˙(+) using low-energy collision-induced dissociation and density functional theory, we have examined the influence of the basicity of the N-terminal amino acid residue on the competition between the isomerization and dissociation channels, particularly the selective C(α)-C bond cleavage viaβ-hydrogen atom migration. The dissociation barriers for the formation of a(2)(+) ions from [RFG]˙(+) and [KFG]˙(+)via their β-radical isomers are comparable (33.1 and 35.0 kcal mol(-1), respectively); the dissociation barrier for the charge-induced formation of the [b(2)- H]˙(+) radical cation from [RFG]˙(+)via its α-radical isomer (39.8 kcal mol(-1)) was considerably higher than that from [KFG]˙(+) (27.2 kcal mol(-1)). Thus, the basic arginine residue sequesters the mobile proton to promote the charge-remote selective C(α)-C bond cleavage by energetically hindering the competing charge-induced pathways.  相似文献   

11.
By the peptide map method, a phosphorylated peptide has been isolated from a tryptic hydrolysate of phosphorylated yeast inorganic pyrophosphatase (I), and this is a direct proof of the formation of a covalent bond between (I) and phosphate in the course of this reaction. The isolation and analysis of the peptide from the tryptic hydrolysate shows that the phosphate acceptor is probably the aspartic acid residue 240 or 248. Analysis of a tryptic hydrolysate of (I) modified with O-phosphoethanolamine has shown that O-phosphoethanolamine forms an amide bond with the carboxy group of the same aspartic acid residue. In an alkaline medium, the phosphate residue migrates to the imidazole ring of a histidine residue, apparently that present in position 222.  相似文献   

12.
We report that the cis/trans ratio of the proline peptide bond can be strongly influenced by the chirality of the acyl residue preceding proline. Acyl moieties derived from (2S)-2,6-dimethyl-3-oxo-3,4-dihydro-2H-1,4-benzoxazine-2-carboxylic acid (8) and (2R)-3-methoxy-2-methyl-2-(4-methyl-2-nitrophenoxy)-3-oxopropanoic acid (5) in acyl-Pro molecules influence isomerization of the proline peptide bond constraining the omega dihedral angle to the trans orientation. Structures of benzyl (2S)-1-([(2S)-2,6-dimethyl-3-oxo-3,4-dihydro-2H-1,4-benzoxazin-2-yl]carbonyl)-2-pyrrolidinecarboxylate (3) derived from 2D (1)H NMR conformational analysis and crystallographic data exhibit only the trans conformation of proline peptide bond. On the other hand the diastereomer 4, which contains an (R) acyl moiety, exhibits two sets of signals in (1)H NMR spectra. The signals were assigned to trans (72%) and cis (28%) conformers. Crystallographic analysis of 4 showed that only the cis conformation is present in the crystalline state. The (1)H NMR chemical shift pattern of three sets of signals observed in 2 was observed also in benzyl (2S)-1-[(2R/S)-3-methoxy-2-methyl-2-(4-methyl-2-nitrophenoxy)-3-oxopropanoyl]-2-pyrrolidinecarboxylate. (R)-Carboxylic acid 5, after coupling with (S)-ProOBn, yielded benzyl (2S)-1-[(2R)-3-methoxy-2-methyl-2-(4-methyl-2-nitrophenoxy)-3-oxopropanoyl]-2-pyrrolidinecarboxylate (6), which in DMSO-d(6) exhibited only the trans conformation of the proline peptide bond. These results suggest that in these particular cases acyl-Pro peptide bond isomerization is strongly influenced by the stereochemistry of the acyl residue preceding proline. (2S)-2,6-Dimethyl-3-oxo-3,4-dihydro-2H-1,4-benzoxazine-2-carboxylic acid (8) and (2R)-3-methoxy-2-methyl-2-(4-methyl-2-nitrophenoxy)-3-oxopropanoic acid (5) are promising chiral peptidomimetic building blocks that can be used as acyl moieties to force the proline peptide bond into the trans conformation in a variety of acyl-Pro molecules.  相似文献   

13.
As an extension of our investigation of peptidyl-resin linkage stability towards different cleavage procedures used in the solid-phase peptide synthesis (SPPS) technique, the present paper evaluated the trifluoromethanesulfonic acid (TFMSA)/trifluoroacetic acid (TFA)/thioanisole method, varying the type of resin (benzhydrylamine-resin, BHAR; methylbenzhydrylamine-resin, MBHAR and 4-(oxymethyl)-phenylacetamidomethyl-resin, PAMR) and peptide resin-bound residue (Gly and Phe). The vasoactive angiotensin II (AII, DRVYIHPF) and its [Gly8]-AII analogue linked to those resins used routinely in tert-butyloxycarbonyl (Boc)-SPPS chemistry were submitted comparatively to a time course study towards TFMSA/TFA cleavage. At 0 degrees C, [Gly8]-AII was completely removed from all resins in less than 6 h, but the hydrophobic Phe8 moiety-containing AII sequence was only partially cleaved (not more than 15%) from BHAR or MBHAR in this period. At 25 degrees C, [Gly8]-AII cleavage time decreased to less than 2 h irrespective of the solid support, and quantitative removal of AII from PAMR and MBHAR occurred in less than 3 h. However, about 10-15 h seemed to be necessary for cleavage of AII from BHAR, and in this extended cleavage reaction a significant increase in peptide degradation rate was observed. Regardless of the cleavage temperature used, the decreasing order of acid stability measured for resins was BHAR>MBHAR>PAMR. Collectively, these findings demonstrated the feasibility of applying TFMSA/TFA solution as a substitute for anhydrous HF at the cleavage step in Boc-SPPS methodology. Care should be taken however, as the cleavage efficacy depends on multiple factors including the resin, peptide sequence, the time and temperature of reaction.  相似文献   

14.
Multiple Attenuated Total Reflectance (MATR) IR spectra of solutions of methanesulfonic acid (MSA) in 2-pyrrolidone (Pyr) (0—100 % MSA) were examined at 30 °C. Depending on the ratio of components, two types of species with strong symmetric H bonds are formed: uncharged quasi-ion pairs of 1 : 1 composition and anions (AHA). The solvation of quasi-ion pairs by base molecules and of the (AHA) anions by acid molecules was detected. The data on the influence of the hydrogen bond donor and acceptor on the IR spectra of the quasi-ion pairs and (AHA) anions are presented.  相似文献   

15.
Ab initio and density functional methods have been used to examine the structures and energetics of the hydrated clusters of methane sulfonic acid (MSA), CH3SO3H.(H2O)n (n = 1-5). For small clusters with one or two water molecules, the most stable clusters have strong cyclic hydrogen bonds between the proton of OH group in MSA and the water molecules. With three or more water molecules, the proton transfer from MSA to water becomes possible, forming ion-pair structures between CH3SO3- and H3O+ moieties. For MSA.(H2O)3, the energy difference between the most stable ion pair and neutral structures are less than 1 kJ/mol, thus coexistence of neutral and ion-pair isomers are expected. For larger clusters with four and five water molecules, the ion-pair isomers are more stable (>10 kJ/mol) than the neutral ones; thus, proton transfer takes place. The ion-pair clusters can have direct hydrogen bond between CH3SO3- and H3O+ or indirect one through water molecule. For MSA.(H2O)5, the energy difference between ion pairs with direct and indirect hydrogen bonds are less than 1 kJ/mol; namely, the charge separation and acid ionization is energetically possible. The calculated IR spectra of stable isomers of MSA.(H2O)n clusters clearly demonstrate the significant red shift of OH stretching of MSA and hydrogen-bonded OH stretching of water molecules as the size of cluster increases.  相似文献   

16.
Direct liquid-phase sulfonation of methane to methanesulfonic acid (MSA) with SO2 has been achieved in triflic acid using K2S2O8 as the oxidant and a small amount of a Ca salt as the promoter. The effects of reaction conditions on the conversion of SO2 to MSA were examined. Included were the influence of solvent acidity, reaction duration, reaction temperature, amount of K2S2O8, and composition and amount of promoters.  相似文献   

17.
Amino acid thiohydantoins were identified using electron impact and chemical ionization mass spectrometry. Some fragmentations of thiohydantoin ring were also identified. These results suggest that the thiohydantoin method should be useful for the stepwise peptide sequence analysis starting from the carboxyl terminus. The sequence of a model tripeptide was determined by the combination of the thiohydantoin method and mass spectrometry. In the case of peptide which contained a proline residue as their C-terminal peptide, cleavage of the peptide bond cannot be achieved.  相似文献   

18.
A 4,5-disubstituted-9,9-dimethylxanthene-based amino acid (10) has been synthesized for incorporation into peptide sequences which have a propensity to adopt beta-sheet structure. Molecular dynamics studies support the FT-IR and NMR results which demonstrate that amides based on this residue utilize the NH and the C=O from the xanthene residue to form an intramolecular hydrogen bond (13-membered ring), unlike the previously studied dibenzofuran-based amino acid residues in which the NH and the C=O of the attached amide groups participate in intramolecular hydrogen bonding (15-membered ring). Interestingly, residue 10 derivatized as a simple amide prefers to adopt a trans conformation where the aliphatic side chains are placed on opposite sides of the plane of the 9,9-dimethylxanthene ring system. This is different than the conformational preferences of the dibenzofuran-based amino acids which adopt a cis conformation that is preorganized to nucleate beta-sheet formation. It will be interesting to see how these conformational differences effect nucleation in aqueous solution.  相似文献   

19.
The positive ion electrospray ionization (ESI+) mass spectra of peptides usually display only protonated molecules provided that soft ionization conditions are applied (low cone voltage to prevent in-source dissociations). Such ions can be multiply charged depending on the molecular weight of the studied compounds. We have experienced an unexpected behavior during the ESI analysis of a modified peptide of relatively high mass (3079 Da). A specific fragmentation occurred even under soft energetic conditions, leading to a mass spectrum containing multiply charged molecular and fragment ions. The selective rupture involved the amide bond between the glutamic acid and proline residues (E-P sequence). The successive replacement of each amino acid by an alanine residue (positional scanning study) was undertaken to assess which part of the sequence induced such selective and abundant fragmentation on multiply charged species. The succession P-P was evidenced as the minimum unit giving rise to the first peptide bond rupture in the sequence X-P-P. Any acidic amino acid at the X position (X = D, E) favored the fragmentation by an intramolecular interaction. Such proline-induced fragmentation occurring readily in the source differed from the literature data on the specific behavior of proline-containing peptides where bond ruptures occur solely in dissociation conditions.  相似文献   

20.
Pseudo-peptide analogs of trans-4-aminomethylcyclohexanecarbonyl-L-phenylalanyl-4-aminopheny l acetic acid (PKSI-527, plasma kallikrein selective inhibitor), in which an amide bond (peptide bond) has been replaced by a CH2-NH bond, i.e., trans-4-aminomethylcyclohexanecarbonyl-L-phenylalanyl-psi (CH2-NH)-4-aminophenyl acetic acid (I), trans-4-aminomethylcyclohexanecarbonyl-psi (CH2-NH)-L-phenylalanyl-4-aminophenyl acetic acid (II) and trans-4-aminomethylcyclohexanecarbonyl-D-phenylalanyl-psi (CH2-NH)-4-aminophenyl acetic acid (III) were synthesized. These pseudo-peptide analogs did not exhibit any detectable inhibitory activity against plasma kallikrein (PK), plasmin (PL), urokinase (UK), thrombine (TH) or trypsin (TRY). These results indicate that both carbonyl groups in the PKSI-527 are important for the manifestation of potent inhibitory activity against plasma kallikrein.  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