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 483 毫秒
1.
The dissociation of the amide (peptide) bond in protonated peptides, [M + H](+), is discussed in terms of the structures and energetics of the resulting N-terminal b(n) and C-terminal y(n) sequence ions. The combined data provide strong evidence that dissociation proceeds with no reverse barriers through interconverting proton-bound complexes between the segments emerging upon cleavage of the protonated peptide bond. These complexes contain the C-terminal part as a smaller linear peptide (amino acid if one residue) and the N-terminal part either as an oxazolone or a cyclic peptide (cyclic amide if one residue). Owing to the higher thermodynamic stability but substantially lower gas-phase basicity of cyclic peptides vs isomeric oxazolones, the N-terminus is cleaved as a protonated oxazolone when ionic (b(n) series) but as a cyclic peptide when neutral (accompanying the C-terminal y(n) series). It is demonstrated that free energy correlations can be used to derive thermochemical data about sequence ions. In this context, the dependence of the logarithm of the abundance ratio log[y(1)/b(2)], from protonated GGX (G, glycine; X, varying amino acid) on the gas-phase basicity of X is used to obtain a first experimental estimate of the gas-phase basicity of the simplest b-type oxazolone, viz. 2-aminomethyl-5-oxazolone (b(2) ion with two glycyl residues).  相似文献   

2.
We report a study on encapsulation of various amino acids into gas-phase sodium bis(2-ethylhexyl) sulfosuccinate (NaAOT) reverse micelles, using electrospray ionization guided-ion-beam tandem mass spectrometry. Collision-induced dissociation of mass-selected reverse micellar ions with Xe was performed to probe structures of gas-phase micellar assemblies, identify solute-surfactant interactions, and determine preferential incorporation sites of amino acids. Integration into gas-phase reverse micelles depends upon amino acid hydrophobicity and charge state. For examples, glycine and protonated amino acids (such as protonated tryptophan) are encapsulated within the micellar core via electrostatic interactions; while neutral tryptophan is adsorbed in the surfactant layer. As verified using model polar hydrophobic compounds, the hydrophobic effect and solute-interface hydrogen-bonding do not provide sufficient driving force needed for interfacial solubilization of neutral tryptophan. Neutral tryptophan, with a zwitterionic structure, is intercalated at the micellar interface between surfactant molecules through complementary effects of electrostatic interactions between tryptophan backbone and AOT polar heads, and hydrophobic interactions between tryptophan side chain and AOT alkyl tails. Protonation of tryptophan could significantly improve its incorporation capacity into gas-phase reverse micelles, and displace its incorporation site from the micellar interfacial zone to the core; protonation of glycine, on the other hand, has little effect on its encapsulation capacity. Another interesting observation is that amino acids of different isoelectric points could be selectively encapsulated into, and transported by, reverse micelles from solution to the gas phase, based upon their competition for protonation and subsequent encapsulation within the micellar core.  相似文献   

3.
We report the first gas-phase spectroscopic study of a three-residue model of a peptide chain, Ac-Phe-Gly-Gly-NH2 (Ac = acetyl), using the IR/UV double resonance technique. The existence of at least five different conformers under supersonic expansion conditions is established, most of them exhibiting rather strong intramolecular H-bonds. One of the most populated conformers, however, exhibits a different H-bonding network characterized by two weak H-bonds. Comparison of the amide A and I/II experimental data with density functional theory calculations carried out on a series of selected conformations enables us to assign this conformer to two successive beta-turns along the peptide chain, the two H-bonds being of C10 type, i.e., each of them closing a 10-atom ring in the molecule. The corresponding form is found to be more stable than the 310 helix secondary structure (not observed), presumably because of specific effects due to the glycine residues.  相似文献   

4.
The role of benzodiazepine derivatives (BZD) as a privileged scaffold that mimics beta-turn structures (Ripka et al. (1993) Tetrahedron 49:3593-3608) in peptide/protein recognition was reexamined in detail. Stable BZD ring conformers were determined with MM3, and experimental reverse-turn structures were extracted from the basis set of protein crystal structures previously defined by Ripka et al. Ideal beta-turns were also modeled and similarly compared with BZD conformers. Huge numbers of conformers were generated by systematically scanning the torsional degrees of freedom for BZDs, as well as those of ideal beta-turns for comparison. Using these structures, conformers of BZDs were fit to experimental structures as suggested by Ripka et al., or modeled classical beta-turn conformers, and the root-mean-square deviation (RMSD) values were calculated for each pairwise comparison. Pairs of conformers with the smallest RMSD values for overlap of the four alpha-beta side-chain orientations were selected. All overlaps of BZD conformers with experimental beta-turns yielded one or more comparisons where the least RMSD was significantly small, 0.48-0.86 angstroms, as previously suggested. Utilizing a different methodology, the overall conclusion that benzodiazepines could serve as reverse-turn mimetics of Ripka et al. is justified. The least RMSD values for the overlap of BZDs and modeled classical beta-turns were also less than 1 angstrom. When comparing BZDs with experimental or classical beta-turns, the set of experimental beta-turns selected by Ripka et al. fit the BZD scaffolds better than modeled classical beta-turns; however, all the experimental beta-turns did not fit a particular BZD scaffold better. A single BZD ring conformation, and/or chiral orientation, can mimic some, but not all, of the experimental beta-turn structures. BZD has two central ring conformations and one chiral center that explains why the four variations of the BZD scaffold can mimic all types of beta-turn structure examined. It was found, moreover, that the BZD scaffold also mimics each of the nine clusters of experimental orientations of side chains of reverse turns in the Protein Data Bank, when the new classification scheme for the four side-chain directions (the relative orientations of alpha-beta vectors of residues i through i+3) was considered (Tran et al. (2005) J Comput-Aided Mol Des 19:551-566).  相似文献   

5.
All-atom molecular dynamics simulations for a single molecule of Leu-Enkephalin in aqueous solution have been used to study the role of the water network during the formation of beta-turns. We give a detailed account of the intramolecular hydrogen bonding, the water-peptide hydrogen bonding, and the orientation and residence times of water molecules focusing on the short critical periods of transition to the stable beta-turns. These studies suggest that, when intramolecular hydrogen bonding between the first and fourth residue of the beta-turn is not present, the disruption of the water network and the establishment of water bridges constitute decisive factors in the formation and stability of the beta-turn. Finally, we provide possible explanations and mechanisms for the formations of different kinds of beta-turns.  相似文献   

6.
The difficulty with integrating solution-phase hydrogen/deuterium exchange (HDX) and tandem mass spectrometry is that the energy added to cause fragmentation might promote gas-phase migration of the added deuterium atoms. Here, we compare the solution-phase HDX profiles generated from a- b- and y-type fragment ion series originating from capillary-skimmer dissociation. The isotopic distributions of fragments from the different fragment ion types were used to determine the isotopic state of the amide hydrogen within a specific residue. Even though the same amide hydrogen was examined, the result was different for different fragment ion types. This observation indicates that different fragment series are not equally subjected to inter-molecular migration during collision-induced dissociation (CID). We also investigated the gas-phase reactivity of originally undeuterated CID fragments of penta-phenylalanine using gas-phase HDX in an external accumulation hexapole. The incorporation of deuterium into the different fragments was studied as a function of hexapole pressure. It was found that different b- and y-ions from the same peptide had different gas-phase reactivity. However, the a-ions did not display significant gas-phase reactivity. The observed behavior has significant impact on any method that involves comparing the isotopic distributions of different fragment ions. Great care has to be taken in the interpretation of the HDX data using CID to increase the spatial resolution. The isotopic state observed after solution-phase exchange might be more preserved for some CID-fragment types.  相似文献   

7.
Semiempirical SCFMO computations of the MNDO and AM1 varieties have been employed to model possible interaction processes for cocaine with the biological substrate glycine. It was found for the gas-phase species that the most likely interactions occurred as nucleophilic attack of the nitrogen lone pair of glycine on the two carbonyl groups of cocaine. These reactions led to intermediates which further decomposed exothermically to amide and alcohol species. The activation enthalpies for the gas-phase reactions were predicted to be high (39–46 kcal/mol), but it is believed that such processes could possibly occur by reaction pathways with considerably lower activation energies in the liquid state in the presence of mucus containing olfactory binding protein in vertebrate olfactory reception.  相似文献   

8.
The incremental free energies of aqueous solution for acetyl(ala)NNH2 in its extended unfolded and alpha-helical conformations are compared using the SM5.2 solvation method of Cramer and Truhlar. A combination of density functional theory (DFT) at the B3LYP/D95(d,p) and AM1 has been employed using the ONIOM method. The incremental solvation energies of alpha-helical structures are very similar for both ONIOM and AM1 optimized structures as these structures do not significantly change upon solution. However, the conformations of the unfolded peptides change from extended beta-strand to polyproline II conformations upon aqueous solution. The incremental solvation free energy per residue of the polyproline II structure is about 2 kcal/mol/residue greater than that for the alpha-helix, representing an upper limit for the difference between the solvation energies. However, most of this difference disappears when the energy required to distort the optimized gas-phase extended beta-strand structure to the optimized polyproline II solution structure is included in the analysis, leaving an estimated difference in incremental solvation free energy of 0.3-0.5 kcal/mol favoring the unfolded structure. The solution structure sacrifices the stability derived from the intramolecular C5 H-bonds for more favorable interactions with the aqueous solvent.  相似文献   

9.
The issue of the influence of the side chain/backbone interaction on the local conformational preferences of a phenylalanine residue in a peptide chain is addressed. A synergetic approach is used, which combines gas-phase UV spectroscopy as well as gas-phase IR/UV double-resonance experiments with DFT and post Hartree-Fock calculations. N-Acetyl-Phe-amide was chosen as a model system for which three different conformers were observed. The most stable conformer has been identified as an extended beta(L) conformation of the peptide backbone. It is stabilized by a weak but significant NH-pi interaction bridging the aromatic ring on the residue (i) with the NH group on residue (i+1), with the aromatic side chain being in an anti conformation. This stable conformation corresponds to the common NH(i+1)-aromatic(i) interaction encountered in proteins for the three aromatic residues (phenylalanine, tyrosine, and tryptophan), which illustrates the relevance of gas-phase investigations to structural biology issues. The two other less abundant conformers have been assigned to two gamma-folded backbone conformations that differ by the orientation of the side chain. In all cases, the IR data provided spectroscopic fingerprints of these interactions. Finally, the strong conformational dependence of the fluorescence yield found for N-acetyl-Phe-amide illustrates the role of the environment on the excited-state dynamics of these species, which is often exploited by biochemists to monitor protein structural changes from tryptophan lifetime measurements.  相似文献   

10.
We previously reported that the peptide containing a Cys-Pro ester (CPE) moiety is spontaneously transformed into a peptide thioester via an N to S acyl shift followed by diketopiperazine formation. In an attempt to identify more reactive structures for the formation of a peptide thioester, we modified the CPE structure, in which the Pro residue in the CPE moiety was replaced with N-substituted glycine derivatives. These peptides were transformed into a peptide thioester more rapidly. Alternatively, the addition of an amino acid residue at the C-terminus of the CPE moiety also accelerated thioester formation.  相似文献   

11.
A rigid pyrrolidine based scaffold comprising of 2,5-dideoxy-2,5-imino-D-idaric acid (1) is developed. Attachment of peptide strands to the carboxylic groups at both ends of this novel template led to the peptidomimetics 2 and 3. Conformational analysis by NMR studies revealed that compounds 2b, 3b and 2c, 3c take interesting turn structures (C(2) symmetric for 2c and 3c) in DMSO-d(6) consisting of identical intramolecular hydrogen bonds at two ends between LeuNH --> sugar-OH as depicted in structure A, whereas 2a and 3a display structures with regular beta-turns with hydrogen bonds between LeuNH --> Boc-C=O in one-half of their molecular frameworks (structure B), characteristic of the turn structures commonly observed in "D-Pro-Gly"-containing peptides. These results suggest that a cis hydroxyl group at the 3-position of the proline residue favors a pseudo beta-turn-like nine-membered ring structure in hydroxyproline-containing peptides involving an intramolecular hydrogen bond between the hydroxyl and the i + 2 backbone amide.  相似文献   

12.
13.
We present a method based on the first-order Markov models for predicting simple beta-turns and loops containing multiple turns in proteins. Sequences of 338 proteins in a database are divided using the published turn criteria into the following three regions, namely, the turn, the boundary, and the nonturn ones. A transition probability matrix is constructed for either the turn or the nonturn region using the weighted transition probabilities computed for dipeptides identified from each region. There are two such matrices constructed for the boundary region since the transition probabilities for dipeptides immediately preceding or following a turn are different. The window used for scanning a protein sequence from amino (N-) to carboxyl (C-) terminal is a hexapeptide since the transition probability computed for a turn tetrapeptide is capped at both the N- and C- termini with a boundary transition probability indexed respectively from the two boundary transition matrices. A sum of the averaged product of the transition probabilities of all the hexapeptides involving each residue is computed. This is then weighted with a probability computed from assuming that all the hexapeptides are from the nonturn region to give the final prediction quantity. Both simple beta-turns and loops containing multiple turns in a protein are then identified by the rising of the prediction quantity computed. The performance of the prediction scheme or the percentage (%) of correct prediction is evaluated through computation of Matthews correlation coefficients for each protein predicted. It is found that the prediction method is capable of giving prediction results with better correlation between the percent of correct prediction and the Matthews correlation coefficients for a group of test proteins as compared with those predicted using some secondary structural prediction methods. The prediction accuracy for about 40% of proteins in the database or 50% of proteins in the test set is better than 70%. Such a percentage for the test set is reduced to 30 if the structures of all the proteins in the set are treated as unknown.  相似文献   

14.
The most stable structures for the gas-phase complexes of minor tautomers of uracil (U) with glycine (G) were characterized at the density functional B3LYP/6-31++G level of theory. These are cyclic structures stabilized by two hydrogen bonds. The relative stability of isolated tautomers of uracil was rationalized by using thermodynamic and structural arguments. The stabilization energies for complexes between the tautomers of U and G result from interplay between the stabilizing two-body interaction energies and destabilizing one-body terms. The latter are related to the energies of (i) tautomerization of the unperturbed moieties and (ii) distortions of the resulting rare tautomers in the complex. The two-body term describes the interaction energy between distorted tautomers. The two-body interaction energy term correlates with perturbations of length of the proton-donor bonds as well as with deprotonation enthalpies and proton affinities of the appropriate monomer sites. It was demonstrated that the relative instability of rare tautomers of uracil is diminished due to their interactions with glycine. In particular, the instability of the third most stable tautomer (U(III)) is decreased from 11.9 kcal/mol for non-interacting uracil to 6.7 kcal/mol for uracil in a complex with the zwitterionic tautomer of glycine. A decrease of instability by 5.2 kcal/mol could result in an increase of concentration of U(III) by almost 5 orders of magnitude. This is the tautomer with proton donor and acceptor sites matching guanine rather than adenine. Moreover, kinetic characteristics obtained for the glycine-assisted conversion of the most stable tautomer of uracil (U(I)) to U(III) indicate that the U(I)<-->U(III) thermodynamic equilibrium could be easily attained at room temperature. The resulting concentration of this tautomer falls in a mutationally significant range.  相似文献   

15.
Arrhenius activation energies in the zero-pressure limit for dissociation of gas-phase proton-bound homodimers of N,N-dimethylacetamide (N,N-DMA), glycine, alanine, and lysine and the heterodimer alanine.glycine were measured using blackbody infrared radiative dissociation (BIRD). In combination with master equation modeling of the kinetic data, binding energies of these dimers were determined. A value of 1.25 +/- 0.05 eV is obtained for N,N-DMA and is in excellent agreement with that reported in the literature. The value obtained from the truncated Boltzmann model is significantly higher, indicating that the assumptions of this model do not apply to these ions. This is due to the competitive rates of photon emission and dissociation for these relatively large ions. The binding energies of the amino acid dimers are ~1.15 +/- 0.05 eV and are indistinguishable despite the difference in their gas-phase basicity and structure. The threshold dissociation energies can be accurately modeled using a range of dissociation parameters and absorption/emission rates. However, the absolute values of the dissociation rates depend more strongly on the absorption/emission rates. For N,N-DMA and glycine, an accurate fit was obtained using frequencies and transition dipole moments calculated at the ab initio RHF/2-31G* and MP2/2-31G* level, respectively. In order to obtain a similar accuracy using values obtained from AM1 semiempirical calculations, it was necessary to multiply the transition dipole moments by a factor of 3. These results demonstrate that in combination with master equation modeling, BIRD can be used to obtain accurate threshold dissociation energies of relatively small ions of biological interest.  相似文献   

16.
Matrix-assisted laser desorption/ionisation (MALDI) mechanisms and the factors that influence the intensity of the ion signal in the mass spectrum remain imperfectly understood. In proteomics, it is often necessary to maximise the peptide response in the mass spectrum, especially for low abundant proteins or for proteolytic peptides of particular significance. We set out to determine which of the common proteolytic enzymes give rise to peptides with the best response factors under MALDI conditions. Standard proteins were enzymatically digested using four common proteases. We assessed relative response factors by coanalyzing the resulting digests. Thus, when tryptic peptides were added in equimolar quantities to their corresponding Asp-N, chymotrypsin and Glu-C digests, tryptic peptide signals were always predominant in the resulting MALDI mass spectra. Observable peaks attributable to non-tryptic peptides generally contained a terminal basic residue. It was proposed that a terminal basic residue has a disproportionate influence upon gas-phase basicity, and this hypothesis was supported by experiments with model isotopically labelled peptides. Experiments applying Cook's kinetic method showed that the peptide with a C-terminal arginine residue was more basic than the equivalent peptide with an N-terminal arginine, which was more basic than the peptide in which the arginine was mid-chain. Thus, the observation of the higher MALDI mass spectrometry response factors of tryptic peptides in comparison with peptides derived using other proteolytic enzymes corresponds with higher gas-phase basicities and may, along with other factors such as the complexity of the digest, influence the choice of enzyme in "bottom-up" proteomic experiments.  相似文献   

17.
Geometries, as well as bond and atomic properties obtained with the atoms-in-molecules theory applied on B3LYP/6-31++G//B3LYP/6-31G charge densities, of the N-formyl amides of the nine tripeptides obtained by combining glycine, alanine, and serine around a central glycine residue were analyzed to check how the properties of the central residue are modified by other amino acids bonded to it. All of the molecules were optimized from an alpha-helix conformation that was also displayed by the optimized structure. Significant variations of the geometry (especially remarkable for dihedral angles) and atomic properties of the central glycine residue are observed when it is attached to a serine residue whose side chain is involved in a hydrogen bond.  相似文献   

18.
Combining laser desorption with a supersonic expansion together with the selectivity of IR/UV double resonance spectroscopy makes it possible to isolate and characterise the gas phase of remarkable backbone conformations of short peptide chains mimicking protein segments. A systematic bottom-up approach involving a conformer-specific IR study of peptide sequences of increasing sizes has enabled us to map the spectral signatures of the intramolecular interactions, which shape the peptide backbone, in particular H-bonds. The precise data collected are directly comparable to the most sophisticated quantum chemistry calculations of these species and therefore constitute a stringent test for the theoretical methods used. One-residue chains reveal the local conformational preference of the backbone and its dependence upon the nature of the residue. The investigation of longer chains provides evidence for a competition between simple successions of local conformational preferences along the chain and more folded structures, in which a new H-bonding network, involving distant H-bonding sites along the backbone, takes place. From three residues, the issue of helical folding can also be addressed. The present review of the gas phase literature data emphasizes the observation of remarkable secondary structures of biology, including short segments of beta-strands, gamma- and beta-turns, combinations of turns, including a 3(10) helix. It also provides evidence for the flexibility of the peptide chains, i.e., a critical influence of rather minor interactions (like side-chain/backbone interactions) on the conformational stability. Finally, the paper will discuss future promising directions of the present approach.  相似文献   

19.
Intramolecularly hydrogen-bonded structures of proline-containing model peptides with a sequence of N-tert-butoxycarbonyl-prolyl-Xaa-NHCH3 [Xaa = Gly (glycyl), Ala (alanyl), Phe (phenylalanyl), Leu (leucyl), Ile (isoleucyl), and Val (valyl)] were studied by proton nuclear magnetic resonance and infrared spectroscopy. Variation of chemical shifts of amide protons with composition change of DMSO-d6/CDCl3 mixed solvents were found to be a good measure of intramolecular hydrogen bonding of peptides in CDCl3 solution. It has been shown that 10- and 7-membered hydrogen-bonded rings, which should have the beta- and gamma-turn like structures in proteins, respectively, form competitively with each other. It is suggested that the equilibrium between the two hydrogen-bonded rings is determined by steric hindrance due to a side chain of the Xaa residue. Free energies for formation of the 10- and 7-membered hydrogen-bonded rings, deltaG10 and deltaG7, were estimated from the solvent composition-dependent change of the chemical shifts. A good correlation between deltaG10 and the occurrence frequencies of residues Xaa at the (i + 2)th position for the beta-turns in proteins has been found.  相似文献   

20.
The gas-phase basicities of oligomers of alanine and valine have been determined by bracketing measurements in an external source Fourier transform mass spectrometer. The results are compared to the oligomers of glycine, which were reported in an earlier publication, to observe the effect of the alkyl group and the increasing gas-phase basicity of the monomer units on the rates of proton transfer reactions. Molecular orbital calculations were performed on protonated triglycine and trialanine to determine how the alkyl groups affect intramolecular interactions. The results show that a high degree of ordering of the carbonyl groups is present in the protonated species. The carbonyl groups in turn order the side chain alkyl groups and decrease the rates of proton transfer reactions in, for example, the oligomers of valine.  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