首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Abstract: We describe here a systematic study to determine the effect on secondary structure of d ‐amino acid substitutions in the nonpolar face of an amphipathic α‐helical peptide. The helix‐destabilizing ability of 19 d ‐amino acid residues in an amphipathic α‐helical model peptide was evaluated by reversed‐phase HPLC and CD spectroscopy. l ‐Amino acid and d ‐amino acid residues show a wide range of helix‐destabilizing effects relative to Gly, as evidenced in melting temperatures (ΔTm) ranging from ?8.5°C to 30.5°C for the l ‐amino acids and ?9.5°C to 9.0°C for the d ‐amino acids. Helix stereochemistry stability coefficients defined as the difference in Tm values for the l ‐ and d ‐amino acid substitutions [(ΔTm′ = TmL and TmD)] ranging from 1°C to 34.5°C. HPLC retention times [ΔtR(XL?XD)] also had values ranging from ?0.52 to 7.31 min at pH 7.0. The helix‐destabilizing ability of a specific d ‐amino acid is highly dependent on its side‐chain, with no clear relationship to the helical propensity of its corresponding l ‐enantiomers. In both CD and reversed‐phase HPLC studies, d ‐amino acids with β‐branched side‐chains destabilize α‐helical structure to the greatest extent. A series of helix stability coefficients was subsequently determined, which should prove valuable both for protein structure‐activity studies and de novo design of novel biologically active peptides.  相似文献   

2.
Abstract: Indolizidin‐2‐one amino acids (I2aas, 6S‐ and 6R‐ 1 ) possessing 6S‐ and 6R‐ring‐fusion stereochemistry were introduced into the antimicrobial peptide gramicidin S (GS) to explore the relationships between configuration, peptide conformation and biological activity. Solution‐phase and solid‐phase techniques were used to synthesize three analogs with I2aa residues in place of the d ‐Phe‐Pro residues at the turn regions of GS: [(6S)‐I2aa4?5,4′?5′]GS ( 2 ), [Lys2,2′,(6S)‐I2aa4?5,4′?5′]GS ( 3 ) and [(6R)‐I2aa4?5,4′?5′]GS ( 4 ). Although conformational analysis of [I2aa4?5,4′?5′]GS analogs 2?4 indicated that both ring‐fusion stereoisomers of I2aa gave peptides with CD and NMR spectral data characteristic of GS, the (6S)‐I2aa analogs 2 and 3 exhibited more intense CD curve shapes, as well as greater numbers of nonsequential NOE between opposing Val and Leu residues, relative to the (6R)‐I2aa analog 4 , suggesting a greater propensity for the (6S)‐diastereomer to adopt the β‐turn/antiparallel β‐pleated sheet conformation. In measurements of antibacterial and antifungal activity, the (6S)‐I2aa analog 2 exhibited significantly better potency than the (6R)‐I2aa diastereomer 4 . Relative to GS, [(6S)‐I2aa4?5,4′?5′]GS ( 2 ) exhibited usually 1/2 to 1/4 antimicrobial activity as well as 1/4 hemolytic activity. In certain cases, antimicrobial and hemolytic activities of GS were shown to be dissociated through modification at the peptide turn regions with the (6S)‐I2aa diastereomer. The synthesis and evaluation of GS analogs 2?4 has furnished new insight into the importance of ring‐fusion stereochemistry for turn mimicry by indolizidin‐2‐one amino acids as well as novel antimicrobial peptides.  相似文献   

3.
Abstract: Novel polycationic analogs of the cyclic decapeptide antibiotic, gramicidin S, possessing NH2, d /l ‐Phe‐NH or l ‐Lys‐NH groups at the 4α‐ or 4β‐positions of the l ‐Pro residues, were synthesized. While l ‐Pro(4α/β‐NH2)‐containing analogs exhibited much weaker antibacterial activity, the d /l ‐Phe and l ‐Lys‐substituted analogs exhibited higher antibacterial activity against Gram‐negative bacteria than the parent gramicidin S. All of these additional amino group‐containing analogs showed substantially reduced toxicity against human blood cells.  相似文献   

4.
Abstract: As a step towards understanding the mechanism of the biological activity of cyclic antimicrobial peptides, the biophysical properties and conformations of four membrane‐active cyclic peptide antibiotics, based on gramicidin S (GS), were examined in aqueous environments. These cyclic peptides, GS10 [cyclo(VKLdY P)2], GS12 [cyclo(VKLKdY PKVKLdY P)], GS14 [cyclo(VKLKVdY PLKVKLdY P)] and [d ‐Lys]4GS14 [cyclo(VKLdK VdY PLKVKLdY P)] (d ‐amino acid residues are denoted by d and are underlined) had different ring sizes of 10, 12 and 14 residues, were different in structure and amphipathicity, and covered a broad spectrum of hemolytic and antimicrobial activities. GS10, GS12 and [d ‐Lys]4GS14 were shown to be monomeric in buffer systems with ionic strength biological environments. GS14 was also monomeric at low concentrations, but aggregated at concentrations > 50 µm . The affinity of peptides for self‐assembly and interaction with hydrophobic surfaces was related to their free energy of intermolecular interaction. The effects of variations in salt and organic solvent (trifluoroethanol) concentration and temperature on peptide conformation were also examined. Similar to GS, GS10 proved to have a stable and rather rigid conformation in different environments and over a broad range of temperatures, whereas GS12, GS14 and [d ‐Lys]4GS14 had more flexible conformations. Despite its conformational similarity to GS10, GS14 had unique physicochemical properties due to its tendency to aggregate at relatively low concentrations. The biophysical data explain the direct relation between structure, amphipathicity and hydrophobicity of the cyclic peptides and their hemolytic activity. However, this relation with the antimicrobial activity of the peptides is of a more complex nature due to the diversity in membrane structures of microorganisms.  相似文献   

5.
Abstract: Many antimicrobial peptides bear arginine (R)‐ and tryptophan (W)‐rich sequence motifs. Based on the sequence Ac‐RRWWRF‐NH2, sets of linear and cyclic peptides were generated by changes in the amino acid sequence, l ‐d ‐amino acid exchange and naphthylalanine substituted for tryptophan. Linear RW‐peptides displayed moderate activity towards Gram‐positive Bacillus subtilis (15 < MIC < 31 μm ) and were inactive against Gram‐negative Escherichia coli at peptide concentrations <100 μm . Cyclization induced high antimicrobial activity. The effect of cyclization was most pronounced for peptides with three adjacent aromatic residues. Incorporation of d ‐amino acid residues had minor influence on the biological activity. The haemolytic activity of all RW‐peptides at 100 μm concentration was low (<7% lysis for linear R/W‐rich peptides and <28% for the cyclic analogues). Introduction of naphthylalanine enhanced the biological activities of both the linear and cyclic peptides. All peptides induced permeabilization of large unilamellar vesicles (LUVs) composed of lipids of the membrane of B. subtilis and erythrocytes, but surprisingly had no effect on LUVs composed of lipids of the E. coli inner membrane. The profiles of peptide activity against B. subtilis and red blood cells correlated with the permeabilizing effects on the corresponding model membranes and were related to hydrophobicity parameters as derived from reversed phase high‐performance liquid chromatography (HPLC). The results underlined the importance of amphipathicity as a driving force for cell lytic activity and suggest that conformational constraints and an appropriate position of aromatic residues allowing the formation of hydrophobic clusters are highly favourable for antimicrobial activity and selectivity.  相似文献   

6.
Abstract: A synthetic peptide that inhibits the growth of estrogen receptor positive (ER+) human breast cancers, growing as xenografts in mice, has been reported. The cyclic 9‐mer peptide, cyclo[EMTOVNOGQ], is derived from α‐fetoprotein (AFP), a safe, naturally occurring human protein produced during pregnancy, which itself has anti‐estrogenic and anti‐breast cancer activity. To determine the pharmacophore of the peptide, a series of analogs was prepared using solid‐phase peptide synthesis. Analogs were screened in a 1‐day bioassay, which assessed their ability to inhibit the estrogen‐stimulated growth of uterus in immature mice. Deletion of glutamic acid, Glu1, abolished activity of the peptide, but glutamine (Gln) or asparagine (Asn) could be substituted for Glu1 without loss of activity. Methionine (Met2) was replaced with lysine (Lys) or tyrosine (Tyr) with retention of activity. Substitution of Lys for Met2 in the cyclic molecule resulted in a compound with activity comparable with the Met2‐containing cyclic molecule, but with a greater than twofold increase in purity and corresponding increase in yield. This Lys analog demonstrated anti‐breast cancer activity equivalent to that of the original Met‐containing peptide. Therefore, Met2 is not essential for biologic activity and substitution of Lys is synthetically advantageous. Threonine (Thr3) is a nonessential site, and can be substituted with serine (Ser), valine (Val), or alanine (Ala) without significant loss of activity. Hydroxyproline (Hyp), substituted in place of the naturally occurring prolines (Pro4, Pro7), allowed retention of activity and increased stability of the peptide during storage. Replacement of the first Pro (Pro4) with Ser maintains the activity of the peptide, but substitution of Ser for the second Pro (Pro7) abolishes the activity of the peptide. This suggests that the imino acid at residue 7 is important for conformation of the peptide, and the backbone atoms are part of the pharmacophore, but Pro4 is not essential. Valine (Val5) can be substituted only with branched‐chain amino acids (isoleucine, leucine or Thr); replacement by d ‐valine or Ala resulted in loss of biologic activity. Thus, for this site, the bulky branched side chain is essential. Asparagine (Asn6) is essential for activity. Substitution with Gln or aspartic acid (Asp), resulted in reduction of biologic activity. Removal of glycine (Gly8) resulted in a loss of activity but nonconservative substitutions can be made at this site without a loss of activity indicating that it is not part of the pharmacophore. Cyclization of the peptide is facilitated by addition of Gln9, but this residue does not occur in AFP nor is it necessary for activity. Gln9 can be replaced with Asn, resulting in a molecule with similar activity. These data indicate that the pharmacophore of the peptide includes side chains of Val5 and Asn6 and backbone atoms contributed by Thr3, Val5, Asn6, Hyp7 and Gly8. Met2 and Gln9 can be modified or replaced. Glu1 can be replaced with charged amino acids, and is not likely to be part of the binding site of the peptide. The results of this study provide information that will be helpful in the rational modification of cyclo[EMTOVNOGQ] to yield peptide analogs and peptidomimetics with advantages in synthesis, pharmacologic properties, and biologic activity.  相似文献   

7.
Abstract: To investigate the role of peptide–membrane interactions in the biological activity of cyclic cationic peptides, the conformations and interactions of four membrane‐active antimicrobial peptides [based on Gramicidin S (GS)] were examined in neutral and negatively charged micelles and phospholipid vesicles, using CD and fluorescence spectroscopy and ultracentrifugation techniques. Moreover, the effects of these peptides on the release of entrapped fluorescent dye from unilamellar vesicles of phosphatidylcholine (PC) and phosphatidylethanolamine/phosphatidylglycerol (PE/PG) were studied. The cyclic peptides include GS10 [Cyclo(VKLdY P)2], GS12 [Cyclo(VKLKdY PKVKLdY P)], GS14 [Cyclo(VKLKVdY PLKVKLdY P)] and [d ‐Lys]4GS14 [Cyclo(VKLdK VdY PLKVKLdY P)] (underlined residues are d ‐amino acids), were different in their ring size, structure and amphipathicity, and covered a broad spectrum of hemolytic and antimicrobial activities. Interaction of the peptides with the zwitterionic PC and negatively charged PE/PG vesicles were distinct from each other. The hydrophobic interaction seems to be the dominant factor in the hemolytic activity of the peptides, as well as their interaction with the PC vesicles. A combination of electrostatic and hydrophobic interactions of the peptides induces aggregation and fusion in PE/PG vesicles with different propensities in the order: [d ‐Lys]4GS14 > GS14 > GS12 > GS10. GS10 and GS14 are apparently located in the deeper levels of the membrane interfaces and closer to the hydrophobic core of the bilayers, whereas GS12 and [d ‐Lys]4GS14 reside closer to the outer boundary of the interface. Because of differing modes of interaction of the cyclic cationic peptides with lipid bilayers, the mechanism of their biological activity (and its relation to peptide–lipid interaction) proved to be versatile and complex, and dependent on the biophysical properties of both the peptides and membranes.  相似文献   

8.
Abstract: In the 11‐residue long dark‐color‐inducing neurohormone (DCIN = [His7]‐corazonin), of locusts, from residue 2 to residue 11, one amino acid at each time was substituted by d ‐phenylalanine (d ‐Phe). The dark‐color‐inducing effect of these peptides was investigated in comparison with unaltered DCIN by a bioassay based on nymphs of a DCIN‐deficient albino mutant of the migratory locust, Locusta migratoria. Substitution of any single amino acid by d ‐Phe always reduced the activity, but did not abolish it completely. Maximum inactivation was obtained after substitution of Gln4, Ser6, or Trp9. The latter two residues are within the partial sequence ‐Ser‐Xxx‐Gly‐Trp‐ (Xxx = His in the DCIN) that seems to be important for the dark‐color‐inducing activity, as found also in another study (Insect Biochem. Mol. Biol. 32, 2002, 909). Gln4, however, is outside of this partial sequence. Minimal, although still considerable, inactivation occurred after substitution of Gly8, Phe3, or Asn11, despite the fact that Gly8 is within the ‐Ser‐Xxx‐Gly‐Trp‐ partial sequence. In conclusion, no single active core was found, indicating that the whole sequence of the DCIN is necessary to induce maximum darkening effect. No difference was found in the activity of the peptides in which Gly8 was substituted by d ‐Phe or by l ‐Phe. Therefore the ‐Ser‐Xxx‐Gly‐Trp‐ partial sequence does not seem to be stabilized by a type II β‐turn. Nevertheless, existence of another kind of turn that includes this partial sequence is feasible. A single unsuccessful attempt was made to discover an antagonist to the DCIN.  相似文献   

9.
Abstract: Using results from our previously reported cyclic opioid peptide series and reliable models for μ‐, δ‐, and κ‐opioid receptors (MOR, DOR, and KOR, respectively) and their complexes with peptide ligands, we have designed and synthesized a series of cyclic pentapeptides of structure Tyr‐c[d ‐Cys‐Phe‐Phe‐X]‐NH2, cyclized via disulfide, methylene, or ethylene dithioethers, and where X = d ‐ or l ‐Cys; or d ‐ or l ‐penicillamine (Pen; β,β‐dimethylcysteine). Determination of binding affinities to MOR, DOR, and KOR revealed that members of this series with X = d ‐ or l ‐Cys display KOR affinities in the low nanomolar range, demonstrating that a ‘DPDPE‐like’ tetrapeptide scaffold is suitable not only for DOR and MOR ligands, but also for KOR ligands. The cyclic pentapeptides reported here are not, however, selective for KOR, rather they display significant selectivity and high affinity for MOR. Indeed, peptide 8 , Tyr‐c[d ‐Cys‐Phe‐Phe‐Cys]‐NH2‐cyclized via a methylene dithioether, shows picomolar binding affinity for MOR ( = 16 pm ) with more than 100‐fold selectivity for MOR vs. DOR or KOR, and may be of interest as a high affinity, high selectivity MOR ligand. Nonetheless, the high affinity KOR peptides in this series represent excellent leads for the development of structurally related, selective KOR ligands designed to exploit structurally specific features of KOR, MOR, and DOR.  相似文献   

10.
Abstract: A series of cyclic, disulfide‐ or dithioether‐containing tetrapeptides based on previously reported potent μ‐ and δ‐selective analogs has been explored with the aim of improving their poor affinity to the κ‐opioid receptor. Specifically targeted were modifications of tetrapeptide residues 3 and 4, as they presumably interact with residues from transmembrane helices 6 and 7 and extracellular loop 3 that differ among the three receptors. Accordingly, tetrapeptides were synthesized with Phe3 replaced by aliphatic (Gly, Ala, Aib, Cha), basic (Lys, Arg, homo‐Arg), or aromatic sides chains (Trp, Tyr, p‐NH2Phe), and with d ‐Pen4 replaced by d ‐Cys4, and binding affinities to stably expressed μ‐, δ‐, and κ‐receptors were determined. In general, the resulting analogs failed to exhibit appreciable affinity for the κ‐receptor, with the exception of the tetrapeptide Tyr‐c[d ‐Cys‐Phe‐d ‐Cys]‐NH2, cyclized via a disulfide bond, which demonstrated high binding affinity toward all opioid receptors (Kiμ = 1.26 nm , Kiδ = 16.1 nm , Kiκ = 38.7 nm ). Modeling of the κ‐receptor/ligand complex in the active state reveals that the receptor‐binding pocket for residues 3 and 4 of the tetrapeptide ligands is smaller than that in the μ‐receptor and requires, for optimal fit, that the tripeptide cycle of the ligand assume a higher energy conformation. The magnitude of this energy penalty depends on the nature of the fourth residue of the peptide (d ‐Pen or d ‐Cys) and correlates well with the observed κ‐receptor binding affinity.  相似文献   

11.
In the 11‐residue long dark‐color‐inducing neurohormone (DCIN = [His7]‐corazonin), of locusts, from residue 2 to residue 11, one amino acid at each time was substituted by d ‐phenylalanine (d ‐Phe). The dark‐color‐inducing effect of these peptides was investigated in comparison with unaltered DCIN by a bioassay based on nymphs of a DCIN‐deficient albino mutant of the migratory locust, Locusta migratoria. Substitution of any single amino acid by d ‐Phe always reduced the activity, but did not abolish it completely. Maximum inactivation was obtained after substitution of Gln4, Ser6, or Trp9. The latter two residues are within the partial sequence ‐Ser‐Xxx‐Gly‐Trp‐ (Xxx = His in the DCIN) that seems to be important for the dark‐color‐inducing activity, as found also in another study (Insect Biochem. Mol. Biol. 32, 2002, 909). Gln4, however, is outside of this partial sequence. Minimal, although still considerable, inactivation occurred after substitution of Gly8, Phe3, or Asn11, despite the fact that Gly8 is within the ‐Ser‐Xxx‐Gly‐Trp‐ partial sequence. In conclusion, no single active core was found, indicating that the whole sequence of the DCIN is necessary to induce maximum darkening effect. No difference was found in the activity of the peptides in which Gly8 was substituted by d ‐Phe or by L ‐Phe. Therefore the ‐Ser‐Xxx‐Gly‐Trp‐ partial sequence does not seem to be stabilized by a type II β‐turn. Nevertheless, existence of another kind of turn that includes this partial sequence is feasible. A single unsuccessful attempt was made to discover an antagonist to the DCIN.  相似文献   

12.
Abstract: The presence of l ‐5,5‐dimethylproline (dmP) within an amino acid sequence results in the formation of an X‐dmP peptide bond predominantly locked in a cis conformation. However, the common use of this unnatural amino acid has been hampered by the difficulty of the economical incorporation of the dmP residue into longer peptide segments due to the steric hindrance imposed by the dimethyl moieties. Here, we describe synthesis of the C‐terminal 36‐residue peptide, corresponding to the 89–124 sequence of bovine pancreatic ribonuclease A (RNase A), in which dmP is incorporated as a substitute for Pro93. The peptide was assembled by condensation of protected 5‐ and 31‐residue peptide fragments, which were synthesized by solid‐phase peptide methodology using fluorenylmethyloxycarbonyl chemistry. We focused on optimizing the synthesis of the Fmoc‐Ser(tBu)‐Ser(tBu)‐Lys(Boc)‐Tyr(tBu)‐dmP‐OH pentapeptide (residues 89–93) with efficient acylation of the sterically hindered dmP residue. In a comparative study, the reagent O‐(7‐azabenzotriazol‐1‐yl)‐1,1,3,3‐tetramethyluronium hexafluorophosphate was found to be superior to bromo‐tris‐pyrrolidino‐phosphonium hexafluorophosphate and tetramethylfluoroformamidinium hexafluorophosphate for the synthesis of the ‐Tyr(tBu)‐dmP‐ peptide bond in solution as well as on a resin.  相似文献   

13.
We have designed de novo and synthesized eight 26‐residue all D‐conformation amphipathic α‐helical cationic antimicrobial peptides (AMPs), four with “specificity determinants” which provide specificity for prokaryotic cells over eukaryotic cells and four AMPs without specificity determinants. The eight AMPs contain six positively charged Lys residues on the polar face in four different arrangements to understand the role of these residues have on antimicrobial activity against 14 Acinetobacter baumannii strains, seven of which were resistant to polymyxin B and colistin; six diverse Pseudomonas aeruginosa strains and 17 Staphylococcus aureus strains, nine of which were methicillin‐sensitive, and eight of which were methicillin‐resistant. The four AMPs without specificity determinants are extremely hemolytic. In contrast, the four AMPs with specificity determinants had dramatic improvements in therapeutic indices showing the importance of specificity determinants in removing eukaryotic cell toxicity. The specificity determinants combined with the location of positively charged residues on the polar face provide Gram‐negative pathogen selectivity between A. baumannii and S. aureus. Specificity determinants maintain excellent antimicrobial activity in the presence of human sera, whereas the AMPs without specificity determinants were inactive. This study clearly shows the potential of amphipathic α‐helical AMPs with specificity determinants as therapeutics to replace existing antibiotics.  相似文献   

14.
Abstract: Bolaform amides were designed from N,N'‐bis(carboethoxy‐l ‐valinyl)‐diaminoethane ( 1 ) by linking t‐butyloxycarbonyl‐valine through ethylenediamine (EDA) to enable spectroscopic and X‐ray diffraction analyses. N,N'‐Bis(Boc‐l ‐valinyl)‐diaminoethane ( 2 ) and N,N'‐bis(Boc‐d ‐valinyl)‐diaminoethane ( 3 ) were composed of l ‐Val and d ‐Val, respectively. N‐(Boc‐l ‐valinyl)‐N'‐(Boc‐d ‐valinyl)‐diaminoethane ( 4 ) was composed of both l ‐Val and d ‐Val, and was achiral (meso‐peptide). Peptide 5 was a 1 : 1 mixture of 2 and 3 , and was also achiral (racemate). These peptides mediated gelation of corn oil at a concentration of approximately 1%. Within crystals, the peptides formed β‐sheet ribbons, but differences were observed in hydrogen‐bonding patterns and side‐chain arrangements. These differences were also deduced from temperature dependence of amide protons. Force‐field calculations based on the crystal structures indicated that association of β‐sheet ribbons had energy benefits, and it was assumed that molecular aggregation progressed spontaneously. These structural studies indicated the chirality of amino acids affected for the properties of bolaform amides.  相似文献   

15.
The synthesis of selectively 14C‐labeled l ‐tryptophan and its derivative 5‐hydroxy‐l ‐tryptophan using chemical and multienzymatic methods is reported. The mixture containing [1‐14C[‐dl ‐alanine, indole or 5‐hydroxyindole has been converted to [1‐14C]‐l ‐tryptophan or 5′‐hydroxy‐[1‐14C]‐l ‐tryptophan, respectively, in a one‐pot multienzymatic reaction using four enzymes: d ‐amino acid oxidase, catalase, glutamic‐pyruvic transaminase and tryptophanase. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

16.
Abstract: Dynorphin A (Dyn A), a 17 amino acid peptide H‐Tyr‐Gly‐Gly‐Phe‐Leu‐Arg‐Arg‐Ile‐Arg‐Pro‐Lys‐Leu‐Lys‐Trp‐Asp‐Asn‐Gln‐OH, is a potent opioid peptide which interacts preferentially with κ‐opioid receptors. Research in the development of selective and potent opioid peptide ligands for the κ‐receptor is important in mediating analgesia. Several cyclic disulphide bridge‐containing peptide analogues of Dyn A, which were conformationally constrained in the putative message or address segment of the opioid ligand, were designed, synthesized and assayed. To further investigate the conformational and topographical requirements for the residues in positions 5 and 11 of these analogues, a systematic series of Dyn A1?11‐NH2 cyclic analogues incorporating the sulphydryl‐containing amino acids l ‐ and d ‐Cys and l ‐ and d ‐Pen in positions 5 and 11 were synthesized and assayed. Cyclic lactam peptide analogues were also synthesized and assayed. Several of these cyclic analogues, retained the same affinity and selectivity (vs. the μ‐ and δ‐receptors) as the parent Dyn A1?11‐NH2 peptide in the guinea‐pig brain (GPB), but exhibited a much lower activity in the guinea‐pig ileum (GPI), thus leading to centrally vs. peripherally selective peptides. Studies of the structure–activity relationship of Dyn A peptide provide new insights into the importance of each amino acid residue (and their configurations) in Dyn A analogues for high potency and good selectivity at κ‐opioid receptors. We report herein the progress towards the development of Dyn A peptide ligands, which can act as agonists or antagonists at cell surface receptors that modulate cell function and animal behaviour using various approaches to rational peptide ligand‐based drug design.  相似文献   

17.
Analogues of the opioid peptides H‐Tyr‐c[d ‐Cys‐Gly‐Phe(pNO2)‐d ‐Cys]NH2 (non‐selective), H‐Tyr‐d ‐Arg‐Phe‐Lys‐NH2 (μ‐selective) and dynorphin A(1‐11)‐NH2 (κ‐selective) containing 4′‐[N‐((4′‐phenyl)‐phenethyl)carboxamido]phenylanine (Bcp) in place of Tyr1 were synthesized. All three Bcp1‐opioid peptides retained high μ opioid receptor binding affinity, but showed very significant differences in the opioid receptor selectivity profiles as compared with the corresponding Tyr1‐containing parent peptides. The cyclic peptide H‐Bcp‐c[d ‐Cys‐Gly‐Phe(pNO2)‐d ‐Cys]NH2 turned out to be an extraordinarily potent, μ‐selective opioid agonist, whereas the Bcp1‐analogue of dynorphin A(1‐11)‐NH2 displayed partial agonism at the μ receptor. The obtained results suggest that the large biphenylethyl substituent contained in these compounds may engage in a hydrophobic interaction with a receptor subsite and thereby may play a role in the ligand’s ability to induce a specific receptor conformation or to bind to a distinct receptor conformation in a situation of conformational receptor heterogeneity.  相似文献   

18.
The use of peptides as drugs in pharmaceutical applications is hindered by their susceptibility to proteolysis and therefore low bioavailability. β‐Peptides that contain an additional methylene group in the backbone, are gaining recognition from a pharmaceutical stand point as they are considerably more resilient to proteolysis and metabolism. Recently, we reported two new classes of β ‐peptides, β 3‐ and β2‐peptides derived from l ‐aspartic acid and l ‐diaminopropionic acid, respectively. Here, we report the proteolytic stability of these β‐peptidic compounds and a mixed α /β‐peptide against three enzymes (pronase, trypsin and elastase), as well as, human serum. The stability of these peptides was compared to an α‐peptide. Peptides containing β‐linkages were resistant to all conditions. The mixed α /β‐peptide, however, exhibited proteolysis in the presence of trypsin and pronase but not elastase. The rate of degradation of the mixed α /β‐peptide was slower than that would be expected for an α‐peptide. In addition, these β‐peptides were not toxic to HeLa and COS‐1 cell lines as observed by MTT cytotoxicity assay. These results expand the scope of mixed α /β‐peptides containing β‐amino acids or small β‐peptide fragments as therapeutic peptides.  相似文献   

19.
Abstract: The ability of certain θ‐defensins, including retrocyclin‐1, to protect human cells from infection by HIV‐1 marks them as potentially useful molecules. θ‐Defensins composed of l ‐amino acids are likely to be unstable in environments that contain host and microbial proteases. This study compared the properties of two enantiomeric θ‐defensins, retrocyclin‐1, and RC‐112. Although these peptides have identical sequences, RC‐112 is composed exclusively of d ‐amino acids, whereas retrocyclin‐1 contains only l ‐amino acids. We compared the ability of these peptides to protect JC53‐BL human cells from infection by 30 primary HIV‐1 isolates. JC53‐BL cells are modified HeLa cells that express surface CD4, CXCR4, and CCR5. They also contain reporter cassettes that are driven by the HIV‐1 LTR, and express β‐galactosidase and luciferase. The HIV‐1 isolates varied in co‐receptor specificity and included subtypes A, B, C, D, CRF01‐AE, and G. RC‐112 was several fold more potent than retrocyclin‐1 across the entire HIV‐1 panel. Although RC‐112 bound immobilized gp120 and CD4 with lower affinity than did retrocyclin‐1, surface plasmon resonance experiments performed with 1 μg/mL of RC‐112 and retrocyclin‐1 revealed that both glycoproteins were bound to a similar extent. The superior antiviral performance of RC‐112 most likely reflected its resistance to degradation by surface‐associated or secreted proteases of the JC53‐BL target cells. θ‐Defensins composed exclusively of d ‐amino acids merit consideration as starting points for designing microbicides for topical application to the vagina or rectum.  相似文献   

20.
Abstract: To investigate the molecular basis for the interaction of the χ‐constrained conformation of melanotropin peptide with the human melanocortin receptors, a series of β‐substituted proline analogs were synthesized and incorporated into the Ac‐Nle‐c[Asp‐His‐d ‐Phe‐Arg‐Trp‐Lys]‐NH2 (MT‐II) template at the His6 and d ‐Phe7 positions. It was found that the binding affinities generally diminished as the steric bulk of the p‐substituents of the 3‐phenylproline residues increased. From (2S, 3R)‐3‐phenyl‐Pro6 to (2S, 3R)‐3‐(p‐methoxyphenyl)‐Pro6 analogs the binding affinity decreased 23‐fold at the human melanocortin‐3 receptor (hMC3R), 17‐fold at the hMC4R, and eight‐fold at the hMC5R, but selectivity for the hMC5R increased. In addition, the substitution of the d ‐Phe7 residue with a (2R, 3S)‐3‐phenyl‐Pro resulted in greatly reduced binding affinity (103–105) at these melanocortin receptors. Macromodel's Large Scale Low Mode (LLMOD) with OPLS‐AA force field simulations revealed that both MT‐II and SHU‐9119 share a similar backbone conformation and topography with the exception of the orientation of the side chains of d ‐Phe7/d ‐Nal (2′)7 in χ space. Introduction of the dihedrally constrained phenylproline analogs into the His6 position (analogs 2 – 6 ) caused topographical changes that might be responsible for the lower binding affinities. Our findings indicate that hMC3 and hMC4 receptors are more sensitive to steric effects and conformational constraints than the hMC5 receptor. This is the first example for melanocortin receptor selectivity where the propensity of steric interactions in χ space of β‐modified Pro6 analogs of MT‐II has been shown to play a critical role for binding as well as bioefficacy of melanotropins at hMC3 and hMC4 receptors, but not at the hMC5 receptor.  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