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.
Abstract

The purpose of this work was to study the conformation of cyclic peptide 1, cyclo(1,12)- Pen1-Ile2-Thr3-Asp4-Gly5-Glu6-Ala7-Thr8-Asp9-Ser10-Gly11-Cys12-OH, derived from the I-domain of the LFA-1 α-subunit. We found that cyclic peptide 1 can bind to the D1- domain of ICAM-1 and inhibit ICAM-1/LFA-1-mediated homotypic and heterotypic T-cell adhesion. To understand the bioactive conformation and binding requirements for cyclic peptide 1, its solution structure was studied using NMR, CD, and molecular dynamics simulations. Furthermore, possible binding properties between the cyclic peptide and the D1- domain of ICAM-1 were evaluated using docking experiments. This cyclic peptide has a stable βII'-turn at Asp4-Gly5-Glu6-Ala7 and a βI-turn at Pen1-Ile2-Thr3-Asp4; a less stable βV-turn is found at the C-terminal region. The β-turn at Asp4-Gly5-Glu6-Ala7 was also found in the X-ray structure of the I-domain of LFA-1. Our CD studies showed that the peptide binds to calcium/magnesium and forms a 1:1 (peptide:calcium/magnesium) complex with low cation concentrations and multiple types of complexes with higher cation concentrations. Binding to divalent cations causes a conformational change in peptide 1; this is consistent with our previous study that binding of peptide 1 to ICAM-1 was influenced by divalent cations. Docking studies show the interaction between cyclic peptide 1 and the D1- domain of ICAM-1; it indicates that the Ile2-Thr3-Asp4-Gly4-Glu6-Ala7-Thr8 sequence interacts with the F and C strands of the D1-domain. Finally, these studies will help us design a new generation of selective peptides that may bind better to the D1-domain of ICAM-1.  相似文献   

2.
The objective of this study was to explain the increased propensity for the conversion of cyclo-(1,7)-Gly-Arg-Gly-Asp-Ser-Pro-Asp-Gly-OH (1), a vitronectin-selective inhibitor, to its cyclic imide counterpart cyclo-(1,7)-Gly-Arg-Gly-Asu-Ser-Pro-Asp-Gly-OH (2). Therefore, we present the conformational analysis of peptides 1 and 2 by NMR and molecular dynamic simulations (MD). Several different NMR experiments, including COSY, COSY-Relay, HOHAHA, NOESY, ROESY, DQF-COSY and HMQC, were used to: (a) identify each proton in the peptides; (b) determine the sequential assignments; (c) determine the cis-trans isomerization of X-Pro peptide bond; and (d) measure the NH-HCalpha coupling constants. NOE- or ROE-constraints were used in the MD simulations and energy minimizations to determine the preferred conformations of cyclic peptides 1 and 2. Both cyclic peptides 1 and 2 have a stable solution conformation; MD simulations suggest that cyclic peptide 1 has a distorted type I beta-turn at Arg2-Gly3-Asp4-Ser5 and cyclic peptide 2 has a pseudo-type I beta-turn at Ser5-Pro6-Asp7-Gly1. A shift in position of the type I beta-turn at Arg2-Gly3-Asp4-Ser5 in peptide 1 to Ser5-Pro6-Asp7-Gly1 in peptide 2 occurs upon formation of the cyclic imide at the Asp4 residue. Although the secondary structure of cyclic peptide 1 is not conducive to succinimide formation, the reaction proceeds via neighbouring group catalysis by the Ser5 side chain. This mechanism is also supported by the intramolecular hydrogen bond network between the hydroxyl side chain and the backbone nitrogen of Ser5. Based on these results, the stability of Asp-containing peptides cannot be predicted by conformational analysis alone; the influence of anchimeric assistance by surrounding residues must also be considered.  相似文献   

3.
The purpose of this work was to study the conformation of cyclic peptide 1, cyclo(1,12)-Pen1-Ile2-Thr3-Asp4-Gly5-Glu6-Ala7- Thr8-Asp9-Ser10-Gly11-Cys12-OH, derived from the I-domain of the LFA-1 alpha-subunit. We found that cyclic peptide 1 can bind to the D1-domain of ICAM-1 and inhibit ICAM-1/LFA-1-mediated homotypic and heterotypic T-cell adhesion. To understand the bioactive conformation and binding requirements for cyclic peptide 1, its solution structure was studied using NMR, CD, and molecular dynamics simulations. Furthermore, possible binding properties between the cyclic peptide and the D1-domain of ICAM-1 were evaluated using docking experiments. This cyclic peptide has a stable betaII -turn at Asp4- Gly5-Glu6-Ala7 and a betaI-turn at Pen1-Ile2-Thr3-Asp4; a less stable betaV-turn is found at the C-terminal region. The beta-turn at Asp4- Gly5-Glu6-Ala7 was also found in the X-ray structure of the I-domain of LFA-1. Our CD studies showed that the peptide binds to calcium/magnesium and forms a 1:1 (peptide:calcium/magnesium) complex with low cation concentrations and multiple types of complexes with higher cation concentrations. Binding to divalent cations causes a conformational change in peptide 1; this is consistent with our previous study that binding of peptide 1 to ICAM-1 was influenced by divalent cations. Docking studies show the interaction between cyclic peptide 1 and the D1-domain of ICAM-1; it indicates that the Ile2-Thr3-Asp4-Gly4-Glu6-Ala7-Thr8 sequence interacts with the F and C strands of the D1-domain. Finally, these studies will help us design a new generation of selective peptides that may bind better to the D1-domain of ICAM-1.  相似文献   

4.
M D Bruch  J Rizo  L M Gierasch 《Biopolymers》1992,32(12):1741-1754
In an effort to explore the influence of interfacial environments on reverse turns, we have performed a detailed analysis by nmr of the solution conformations of two cyclic pentapeptides in sodium dodecyl sulfate (SDS) micelles. The first peptide, cyclo (D-Phe1-Pro2-Gly3-D-Ala4-Pro5), adopts a single rigid conformation in solution (either chloroform or dimethylsulfoxide) and in crystals, whereas the second, cyclo (Gly1-Pro2-D-Phe3-Gly4-Val5), is much more flexible and adopts different conformations in the crystal and in solution. Both of these peptides are solubilized by SDS micelles, and nmr relaxation rates indicate that they are both partially immobilized by interaction with the micelles. Furthermore, some amide protons in both peptides participate in hydrogen bonds with water. In the presence of micelles, the former peptide retains a conformation essentially the same as that found in crystals and in solution, which consists of a beta turn and an inverse gamma turn. However, the micellar environment has a significant effect on the latter peptide. In particular, the population of a conformer containing a cis Gly-Pro peptide bond is increased significantly. The most likely conformation of the cis isomer, determined by a combination of nmr and restrained molecular dynamics, contains a Gly1-Pro2 delta turn and a gamma turn about D-Phe3. The nmr data on the trans isomer indicate that this isomer is averaging between two conformations that differ mainly in the orientation of the D-Phe3-Gly4 peptide bond.  相似文献   

5.
In an effort to explore the residue preferences in three-residue reverse turns (so-called gamma-turns), two cyclic pentapeptides--cyclo(Gly1-Pro2-D-Phe3-Gly4-Ala5) (I) and cyclo(Gly1-Pro2-D-Phe3-Gly4-Val5) (II)--have been synthesized and analyzed by nmr. It was anticipated that the Gly-Pro-D-Phe-Gly portions of these molecules would favor a beta-turn conformation, leaving the remainder of the molecule to adopt a gamma turn, as seen in several previously studied model cyclic pentapeptides. The nmr data for both peptides in CDCl3 (5% DMSO-d6) and in neat DMSO-d6 indicate that the most populated conformation contains a distorted beta turn around Pro2-D-Phe3, which includes a gamma turn around D-Phe3. The distortion in the beta turn does not impede the formation of an inverse gamma turn around residue 5, and indeed, this conformation is observed in both peptides. Both the alanine and the bulkier valine residues are therefore found to be compatible with an inverse gamma turn. Molecular dynamics simulations on the title peptides are reported in the following paper. These simulations indicate that there is conformational flexibility around the D-Phe3-Gly4 peptide bond, which enables the formation of the gamma turn around D-Phe3. The third paper in this series explores the impact of a micellar environment on conformational equilibria in II.  相似文献   

6.
A bicyclic undecapeptide of sequence cyclo-(Ala(1)-Pro(2)-Asp(3)-Glu(4)-Lys(5)-Ala(6)-Pro(7)-Asp(8)-Ser(9) -Glu(10))-cyclo-(10gamma --> 5varepsilon)-Gly(11), designed to mimic the calcium coordination site I of Calmodulin, has been synthesized and its conformation and calcium binding properties have been investigated by means of CD and nmr spectroscopy. The nmr analysis of the free peptide, carried out in DMSO and in TFE/H(2)O at different pH values, shows the presence in solution of one stable conformer, exhibiting trans configuration around both Proline residues. The nmr results in both solvents suggest for the molecule a rectangular shape constituted by two antiparallel beta-strands connected by two beta-turns. Interproton distances, evaluated by NOE contacts, have been used to obtain feasible models by means of Restrained Molecular Dynamic (RMD). The average models from RMD calculations, for both solvents, exhibit good analogies with Calmodulin site I. The model system, when compared with the reference system (Asp(20)-Glu(31) segment in CaM), shows similar dimensions and an effective superimposition of the respective sequence segments Ala(1)-Glu(4) and Thr(28)-Glu(31). The remaining segments of the model peptide exhibit a bending that is intermediate between that of the free and Ca(2+)-coordinated site I. CD spectra, recorded in TFE solutions, point to a 1:1 stoichiometry for the Ca(2+)-peptide complex, with an association constant of at least 1 x 10(5) M(-1).  相似文献   

7.
Peptides occur in solution as ensembles of conformations rather than in a fixed conformation. The existing energy functions are usually inadequate to predict the conformational equilibrium in solution, because of failure to account properly for solvation, if the solvent is not considered explicitly (which is usually prohibitively expensive). NMR data are therefore widely incorporated into theoretical conformational analysis. Because of conformational flexibility, restrained molecular dynamics (with restraints derived from NMR data), which is usually applied to determine protein conformation is of limited use in the case of peptides. Instead, (a) the restraints are averaged within predefined time windows during molecular dynamics (MD) simulations (time averaging), (b) multiple-copy MD simulations are carried out and the restraints are averaged over the copies (ensemble averaging), or (c) a representative ensemble of sterically feasible conformations is generated and the weights of the conformations are then fitted so that the computed average observables match the experimental data (weight fitting). All these approaches are briefly discussed in this article. If an adequate force field is used, conformations with large statistical weights obtained from the weight-fitting procedure should also have low energies, which can be implemented in force field calibration. Such a procedure is particularly attractive regarding the parameterization of the solvation energy in nonaqueous solvents, e.g., dimethyl sulfoxide, for which thermodynamic solvation data are scarce. A method for calibration of solvation parameters in dimethyl sulfoxide, which is based on this principle was recently proposed by C. Baysal and H. Meirovitch (Journal of the American Chemical Society, 1998, Vol. 120, pp. 800--812), in which the energy gap between the conformations compatible with NMR data and the alternative conformations is maximized. In this work we propose an alternative method based on the principle that the best-fitting statistical weights of conformations should match the Boltzmann weights computed with the force field applied. Preliminary results obtained using three test peptides of varying conformational mobility: H-Ser(1)-Pro(2)-Lys(3)-Leu(4)-OH, Ac-Tyr(1)-D-Phe(2)-Ser(3)-Pro(4)-Lys(5)-Leu(6)-NH(2), and cyclo(Tyr(1)-D-Phe(2)-Ser(3)-Pro(4)-Lys(5)-Leu(6)) are presented.  相似文献   

8.
Two new cyclic peptides, dianthins G-H (1 and 2), together with the known dianthin E (3), were isolated from the traditional Chinese medicinal plant Dianthus superbus. The sequences of cyclic peptides 1 and 2 were elucidated as cyclo (-Gly(1)-Pro(2)-Leu(3)-Thr(4)-Leu(5)-Phe(6)-) and cyclo (-Gly(1)-Pro(2)-Val(3)-Thr(4)-Ile(5)-Phe(6)-), on the basis of ESI tandem mass fragmentation analysis, extensive 2D NMR methods and X-ray diffraction. The isolated three compounds all increase proliferation of MC3T3-E1 cells in vitro using MTT method.  相似文献   

9.
Dai J  Wang X  Feng Y  Fan G  Wang J 《Biopolymers》2004,75(3):229-241
The N-terminal short fragments of staphylococcal nuclease (SNase), SNase20, SNase28, and SNase36, corresponding to the sequence regions, Ala1-Gly20, Ala1-Lys28, and Ala1-Leu36, respectively, as well as an 8-residue peptide (Ala17-Ile18-Asp19-Gly20-Asp21-Thr22-Val23-Lys24) have been synthesized. The conformational states of these fragments were investigated using CD and NMR spectroscopy in aqueous solution and in trifluoroethanol (TFE)-H(2)O mixture. SNase20 containing a sequence corresponding to a bent peptide in native SNase shows a transient population of bend-like conformation around Ala12-Thr13-Leu14 in TFE-H(2)O mixture. The sequence region of Ala17-Thr22 of SNase28 displays a localized propensity for turn-like conformation in both aqueous solution and TFE-H(2)O mixture. The conformational ensemble of SNase36 in aqueous solution includes populated turn-like conformations localized in sequence regions Ala17-Thr22 and Tyr27-Gln30. The analysis suggests that these sequence regions, which form the regular secondary structures in native protein, may serve as the folding nucleation sites of SNase fragments of different chain lengths starting from the N-terminal end. Thus, the formation of bend- and turn-like conformations of these sequence regions may be involved in the early folding events of the SNase polypeptide chain in vitro.  相似文献   

10.
The ribonucleoprotein complex between 5-S RNA and its binding protein (5-S RNA . protein complex) of yeast ribosomes was released from 60-S subunits with 25 mM EDTA and the protein component was purified by chromatography on DEAE-cellulose. This protein, designated YL3 (Mr = 36000 on dodecylsulfate gels), was relatively insoluble in neutral solutions (pH 4--9) and migrated as one of four acidic 60-S subunit proteins when analyzed by the Kaltschmidt and Wittman two-dimensional gel system. Amino acid analyses indicated lower amounts of lysine and arginine than most ribosomal proteins. Sequence homology was observed in the N terminus of YL3, and two prokaryotic 5-S RNA binding proteins, EL18 from Escherichia coli and HL13 from Halobacterium cutirubrum: Ala1-Phe2-Gln3-Lys4-Asp5-Ala6-Lys7-Ser8-Ser9-Ala10-Tyr11-Ser12-Ser13-Arg14-Phe15-Gln16-Tyr17-Pro18-Phe19-Arg20-Arg21-Arg22-Arg23-Glu24-Gly25-Lys26-Thr27-Asp28-Tyr29-Tyr35; of particular interest was homology in the cluster of basic residues (18--23). Since the protein contained one methionine residue it could be split into two fragments, CN1 (Mr = 24700) and CN2 (Mr = 11300) by CNBr treatment; the larger fragment originated from the N terminus. The N-terminal amino acid sequence of CN2 shared a limited sequence homology with an internal portion of a second 5-S RNA binding protein from E. coli, EL5, and, based also on the molecular weights of the proteins and studies on the protein binding sites in 5-S RNAs, a model for the evolution of the eukaryotic 5-S RNA binding protein is suggested in which a fusion of the prokaryotic sequences may have occurred. Unlike the native 5-S RNA . protein complex, a variety of RNAs interacted with the smaller CN2 fragment to form homogeneous ribonucleoprotein complexes; the results suggest that the CN1 fragment may confer specificity on the natural 5-S RNA-protein interaction.  相似文献   

11.
MUC1 mucin is a large transmembrane glycoprotein, of which the extracellular domain is formed by a repeating 20 amino acid sequence, GVTSAPDTRPAPGSTAPPAH. In normal breast epithelial cells, the extracellular domain is densely covered with highly branched complex carbohydrate structures. However, in neoplastic breast tissue, the extracellular domain is underglycosylated, resulting in the exposure of a highly immunogenic core peptide epitope (PDTRP in bold above) as well as the normally cryptic core Tn (GalNAc), STn (sialyl alpha2-6 GalNAc), and TF (Gal beta1-3 GalNAc) carbohydrates. In the present study, NMR methods were used to correlate the effects of cryptic glycosylation outside of the PDTRP core epitope region to the recognition and binding of a monoclonal antibody, Mab B27.29, raised against the intact tumor-associated MUC1 mucin. Four peptides were studied: a MUC1 16mer peptide of the sequence Gly1-Val2-Thr3-Ser4-Ala5-Pro6-Asp7-Thr8-Arg9-Pro10-Ala11-Pro12-Gly13-Ser14-Thr15-Ala16, two singly Tn-glycosylated versions of this peptide at either Thr3 or Ser4, and a doubly Tn-glycosylated version at both Thr3 and Ser4. The results of these studies showed that the B27.29 MUC1 B-cell epitope maps to two separate parts of the glycopeptide, the core peptide epitope spanning the PDTRP sequence and a second (carbohydrate) epitope comprised of the Tn moieties attached at Thr3 and Ser4. The implications of these results are discussed within the framework of developing a glycosylated second-generation MUC1 glycopeptide vaccine.  相似文献   

12.
The conformation and calcium binding properties of the bicyclic nonapeptide BCP2, cyclo-(Glu(1)-Ala(2)-Pro(3)-Gly(4)-Lys(5)-Ala(6)-Pro(7)-Gly(8))-cyclo-(1gamma --> 5epsilon) Gly(9), have been investigated by means of NMR spectroscopy. Interproton distances, evaluated by nuclear Overhauser effect (NOE) contacts, and straight phi angles, from (3)J(NH-alphaCH), have been used to obtain a feasible model for the BCP2-Ca(2+) (BCP: bicyclic peptide) complex by means of restrained molecular dynamics (RMD). The NMR analysis of the free peptide, carried out in CD(3)CN, shows the presence in solution of at least four conformers in intermediate exchange rate. The addition of calcium ions caused the appearance of a new set of resonances, differing from those observed for the free BCP2. A comparison with published data about the conformational behavior of the closely analogous peptide BCP3, differing from BCP2 for two Leu residues instead of two Ala residues in positions 2 and 6, shows that this simple substitution dramatically increases the peptide flexibility. On the contrary, upon calcium ion addition, both BCP2 and BCP3 reach a strictly close conformation, as strongly testified by the almost identical (1)H-NMR spectra exhibited by both peptides. The RMD molecular model of the BCP2-Ca(2+) complex, here reported, is a quite symmetric structure, presenting a three-dimensional cavity ideal for the binding of spherical cations. Four carbonyls from the main ring (Ala(2), Gly(4), Ala(6) and Gly(8)) point toward it, offering, together with the two carbonyls of the peptide bridge (Gly(9) and gammaGlu(1)), putative coordinations to the cation.  相似文献   

13.
The conformational features of Pam-Lys(0)-Arg(1)-Pro(2)-Pro(3)-Gly(4)-Phe(5)-Ser(6)-Pro(7)-Phe(8)-Arg(9)-OH (PKD) and Pam-Gly(-1)-Lys(0)-Arg(1)-Pro(2)-Pro(3)-Gly(4)-Phe(5)-Ser(6)-Pro(7)-Phe(8)-Arg(9)-OH (PGKD), the Pam-Lys and Pam-Gly-Lys analogues of bradykinin, have been determined by high-resolution NMR in a zwitterionic lipoid environment. Radical-induced relaxation of the (1)H NMR signals was used to probe the topological orientation of the peptides with respect to the zwitterionic lipid interface. The radical-induced relaxation and molecular dynamics (MD) data indicated that the palmitic acid and N-terminal amino acid residues embed into the micelles, while the rest of the polypeptide chain is closely associated with the water-micelle interface. Throughout the entire nuclear Overhauser effect restrained MD simulation, a nonideal type I beta-turn was observed in the C-terminus of PKD between residues 6 and 9, and a gamma-turn was observed in the C-terminus of PGKD between residues 6 and 7. Therefore, the additional glycine has a dramatic effect on the structural preferences of the biologically important C-terminus, an effect brought about by the interaction with the lipid environment. These structural features are correlated to the biological activity at the bradykinin B2 receptor.  相似文献   

14.
MUC1 mucin is a large transmembrane glycoprotein, the extracellular domain of which is formed by a repeating 20 amino acid sequence, GVTSAPDTRPAPGSTAPPAH. In normal breast epithelial cells, the extracellular domain is densely covered with highly branched complex carbohydrate structures. However, in neoplastic breast tissue, the extracellular domain is under-glycosylated, resulting in the exposure of a highly immunogenic core peptide epitope (PDTRP in bold above), as well as in the exposure of normally cryptic core Tn (GalNAc), STn (sialyl alpha2-6 GalNAc) and TF (Gal beta1-3 GalNAc) carbohydrates. Here, we report the results of 1H NMR structural studies, natural abundance 13C NMR relaxation measurements and distance-restrained MD simulations designed to probe the structural and dynamical effects of Tn-glycosylation within the PDTRP core peptide epitope. Two synthetic peptides were studied: a nine-residue MUC1 peptide of the sequence, Thr1-Ser2-Ala3-Pro4-Asp5-Thr6-Arg7-Pro8-Ala9, and a Tn-glycosylated version of this peptide, Thr1-Ser2-Ala3-Pro4-Asp5-Thr6(alphaGalNAc)-Arg7-Pro8-Ala9. The results of these studies show that a type I beta-turn conformation is adopted by residues PDTR within the PDTRP region of the unglycosylated MUC1 sequence. The existence of a similar beta-turn within the PDTRP core peptide epitope of the under-glycosylated cancer-associated MUC1 mucin protein might explain the immunodominance of this region in vivo, as the presence of defined secondary structure within peptide epitope regions has been correlated with increased immunogenicity in other systems. Our results have also shown that Tn glycosylation at the central threonine within the PDTRP core epitope region shifts the conformational equilibrium away from the type I beta-turn conformation and toward a more rigid and extended state. The significance of these results are discussed in relation to the possible roles that peptide epitope secondary structure and glycosylation state may play in MUC1 tumor immunogenicity.  相似文献   

15.
Baysal C  Meirovitch H 《Biopolymers》2000,54(6):416-428
A statistical mechanics methodology for predicting the solution structures and populations of peptides developed recently is based on a novel method for optimizing implicit solvation models, which was applied initially to a cyclic hexapeptide in DMSO (C. Baysal and H. Meirovitch, Journal of American Chemical Society, 1998, vol. 120, pp. 800-812). Thus, the molecule has been described by the simplified energy function E(tot) = E(GRO) + summation operator(k) sigma(k)A(k), where E(GRO) is the GROMOS force-field energy, sigma(k) and A(k) are the atomic solvation parameter (ASP) and the solvent accessible surface area of atom k, respectively. In a more recent study, these ASPs have been found to be transferable to the cyclic pentapeptide cyclo(D-Pro(1)-Ala(2)-Ala(3)-Ala(4)-Ala(5)) in DMSO (C. Baysal and H. Meirovitch, Biopolymers, 2000, vol. 53, pp. 423-433). In the present paper, our methodology is applied to the cyclic heptapeptides axinastatin 2 [cyclo(Asn(1)-Pro(2)-Phe(3)-Val(4)-Leu(5)-Pro(6)-Val(7))] and axinastatin 3 [cyclo(Asn(1)-Pro(2)-Phe(3)-Ile(4)-Leu(5)-Pro(6)-Val(7))], in DMSO, which were studied by nmr by Mechnich et al. (Helvetica Chimica Acta, 1997, vol. 80, pp. 1338-1354). The calculations for axinastatin 2 show that special ASPs should be optimized for the partially charged side-chain atoms of Asn while the rest of the atoms take their values derived in our previous work; this suggests that similar optimization might be needed for other side chains as well. The solution structures of these peptides are obtained ab initio (i.e., without using experimental restraints) by an extensive conformational search based on E(GRO) alone and E(*)(tot), which consists of the new set of ASPs. For E(*)(tot), the theoretical values of proton-proton distances, (3)J coupling constants, and other properties are found to agree very well with the nmr results, and they are always better than those based on E(GRO).  相似文献   

16.
The secondary structure of a bradykinin B(1)receptor antagonist B-10324 (F5C-Lys-(1)- Lys(0)-Arg(1)-Pro(2)- Hyp(3)-Gly(4)-CpG(5)- Ser(6)-DTic(7)-CpG(8)) was determined by NMR at 800MHz. The conformational data are compared with those obtained previously for two bradykinin B(1) receptor antagonists, namely B-9858 (Lys-(1)- Lys(0)-Arg(1)-Pro(2)- Hyp(3)-Gly(4)-Igl(5)- Ser(6)-DIgl(7)-Oic(8)) and B-10148 (Lys-(1)-Lys(0)-Arg(1)- Pro(2)-Hyp(3)-Gly(4)- Igl(5)-Ser(6)-DF5F(7)- Oic(8)). The abnormal amino acids are: Hyp, trans-4- hydroxyproline; Tic, 1, 2, 3, 4-tetrahydroisoquinoline-3-carboxylic acid; Oic, (2S, 3aS, 7aS)-octahydroindole-2-carboxylic acid; Igl, alpha(2- indanyl)glycine; F5F, 2,3,4,5,6-pentafluorophenylalanine; CpG, alpha- cyclopentylglycine. F5C, pentafluorocinnamoyl, is the N-terminal protecting group and is not involved in the peptide secondary structure. B-10324 contains an N-terminal Pro(2)- CpG(5) distorted type II beta-turn whereas the rest of the peptide is random. A salt bridge is not observed between the carboxylate group at the C-terminal end and the Arg(1) side chain, in contrast to that previously observed for B-9858 and B- 10148. The conformations are correlated with the measured B(1) receptor antagonist activities (J.-F. Larrivée, L. Gera, S. Houle, J. Bouthillier, D. R. Bachvarov, J. M. Stewart and F. Marc au, Br. J. Pharmacol. 131, 885-892 (2000)). The importance of the N-terminal beta-turn is highlighted.  相似文献   

17.
One carbonyl oxygen of the cyclic hexapeptide cyclo(-Gly1-Pro2-Phe3-Val4-Phe5-Phe6-) (A) can be selectively exchanged with sulphur using Yokoyama's reagent. Surprisingly it was not the C=] of Gly1 but that of Phe5 which was substituted and cyclo(-Gly1-Pro2-Phe3-Val4-Phe5 psi [CS-NH]Phe6-) (B) was obtained. Thionation results in a conformational change of the peptide backbone although the C=O of Phe5 and the corresponding C=S are not involved in internal hydrogen bonds. Two isomers in slow exchange, containing a cis Gly1-Pro2 bond in a beta VIa-turn (minor) and a trans Gly-Pro bond in a beta II'-turn (major), were analyzed by restrained molecular dynamics in vacuo and in DMSO as well as using time dependent distance constraints. It is impossible to fit all experimental data to a static structure of each isomer. Interpreting the conflicting NOEs, local segment flexibility is found. MD simulations lead to a dynamic model for each structure with evidence of an equilibrium between a beta I- and beta II-turn about the Val4-Phe5 amide bond in both the cis and trans isomers. Additionally proton relaxation rates in the rotating frame (R1 rho) were measured to verify the assumption of this fast beta I/beta II equilibrium within each isomer. Significant contributions to R1 rho-rates from intramolecular motions were found for both isomers. Therefore it is possible to distinguish between at least four conformers interconverting on different time scales based on NMR data and MD refinement. This work shows that thionation is a useful modification of peptides for conformation-activity investigations.  相似文献   

18.
The cyclododecapeptide, (Ala1-Pro2-Gly3-Val4-Gly5-Val6)2, was synthesized and its secondary structure was evaluated from extensive studies in dimethyl sulphoxide, trifluoroethanol and water using NMR methods. A selective decoupling technique in 13C-NMR has been utilized in order to assign the C=O carbon resonances. Temperature dependence of the peptide NH protons and the solvent perturbation of the peptide NH and C=O resonances show the occurrence in all solvents of a beta-turn (a 10-membered H-bond between the Val4 NH and Ala1 C=O) and a gamma-turn, an 11-membered H-bond between the Gly3 NH and the Gly5 C=O; and a possible 14-membered H-bond between the Ala1 NH and the Val4 C=O in dimethyl sulphoxide and trifluoroethanol. These secondary structural features are compared with the linear polyhexapeptide and found the the beta-turn and the gamma-turn are the common conformational features of these peptide systems.  相似文献   

19.
Three peptides, B-10148 (Lys-1-Lys0-Arg1-Pro2-Hyp3-Gly4-Igl5-Ser6- DF5F7-Oic8; where Hyp is trans-4-hydroxyproline, Igl is alpha-(2-indanyl)glycine, F5F is 2,3,4,5,6-pentafluorophenylalanine and Oic is (3aS,7aS)-octahydroindole-2-carboxylic acid), B-10206 (DArg0-Arg1-Pro2-Hyp3-Gly4-Igl5-Ser6-DF 5F7-Nc7G8-Arg9; where Nc7G is N-cycloheptylglycine) and B- 10284 (Arg1-Pro2-Pro3-Gly4-Phe5-Thr6-DTic7-Oic8- NH2; where Tic is 1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid), were studied in detail by NMR spectroscopy in 60% CD3OH /40% H2O and modeled by a simulated annealing protocol to determine their solution structure. B-10148, an extremely potent BK B1 receptor antagonist with very high BK B2 receptor antagonist activity, despite lacking a C-terminal Arg, displayed an ideal type II beta-turn from Pro2 to Igl5, as well as a salt bridge between the guanidino group of Arg1 and the carboXylate group of Oic8. B-10206, the most potent B2 antagonist, also displayed an ideal type II beta-turn from Pro2 to Igl5 but secondary structure was not observed at the C-terminal end. The third peptide, B-10284, a des-Arg9 analog with a C-terminal amide and a very potent B2 antagonist, had no definite solution structure. The high activity of these peptides emphasizes the importance of the N-terminal beta-turn and the hydrophobic character at the C-terminus in determining the activity of bradykinin antagonists.  相似文献   

20.
Cyclic hexapeptide analogues representing the modified retro sequence of the amino acid residues 7-11 of natural somatostatin are known to protect liver cells from phalloidin poisoning. To determine the influence of steric, lipophilic, and charge effects on (a) the conformation of the backbone and the aromatic side chains and (b) the biological response, the side chains of Phe2, Lys4, and Phe6 of cyclo(-D-Pro1-Phe2-Thr3-Lys(Z)4-Trp5-Phe6-), 1a, one of the most active peptides found so far, were modified by various residues. The discussion of conformationally relevant parameters proves that neither backbone conformations nor populations of aromatic side chain rotamers were altered by these substitutions. The potency of these derivatives in a cytoprotection assay varies by at most one order of magnitude (more or less active than the parent peptide 1a). A qualitative evaluation of lipophilic, steric, and charge effects reveals the dominance of lipophilic effects of aromatic residues; the most potent compounds contain aromatic substructures in the side chain of Lys4.  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