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.
We report the synthesis, binding affinities to the recombinant human somatostatin receptors, and structure‐activity relationship studies of compounds related to the cyclic hexapeptide, c‐[Pro6‐Phe7‐D‐Trp8‐Lys9‐Thr10‐Phe11], L‐363,301 (the numbering in the sequence refers to the position of the residues in native somatostatin). The Pro residue in this compound is replaced with the arylalkyl peptoid residues Nphe (N‐benzylglycine), (S)βMeNphe [(S)‐N‐[(α‐methyl)benzyl]glycine] or (R)βMeNphe [(R)‐N‐[(α‐methyl)benzyl]glycine] and l ‐1‐naphthylalanine is incorporated into either position 7 or 11 of the parent compound. The synthesis and binding data of the Nnal6 ([N‐naphthylmethyl]glycine) analog of L‐363,301 is also reported. The incorporation of the Nnal residue into position 6 of L‐363,301 resulted in an analog with weaker binding affinities to all hsst receptors but enhanced selectivity towards the hsst2 receptor compared with the parent compound. The other compounds bind effectively to the hsst2 receptor but show some variations in the binding to the hsst3 and hsst5 receptors resulting in different ratios of binding affinities to the hsst5 and hsst2 or hsst3 and hsst2, respectively. The incorporation of the Nphe residue into position 6 and the Nal residue into position 7 of L‐363,301 led to a compound which binds potently to the hsst2 and has increased selectivity towards this receptor (weaker binding to hsst3 and hsst5 receptors) compared with the parent compound. The analogs with β‐methyl chiral substitutions in the aromatic peptoid side chain and Nal in position 7 or 11 bind effectively to the hsst2 and hsst5 receptors. They exhibit similar ratios of binding affinities to the hsst5 and hsst2 receptors as observed for L‐363,301. There are however minor differences in binding to the hsst3 receptor among these analogs. These studies allow us to investigate the influence of additional hydrophobic groups on the binding activity to the isolated human somatostatin receptors and the results are important for the design of other somatostatin analogs. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The synthesis, binding affinity, and structure-activity relationships of compounds related to the cyclic hexapeptide, c[Pro6-Phe7-D-Trp8-Lys9-Thr10-Phe11], L-363,301 (the numbering in the sequence refers to the position of the residue in native somatostatin) is reported. The Pro residue in this compound is replaced with the peptoid residues Nasp [N-(2-carboxyethyl) glycine], Ndab [N-(2-aminoethyl) glycine] and Nlys [N-(4-aminobutyl) glycine]. This series of compounds enables us to draw conclusions about the influence of positively or negatively charged residues in the bridging region on the binding affinity towards the isolated human somatostatin receptors. A loss of binding to the recombinant human somatostatin (hsst) receptors in the Nasp analog compared with L-363,301 and compared with the Ndab and Nlys analogs clearly demonstrates that the presence of an acidic residue in the bridging region is unfavorable for binding to the hsst receptors. Comparison between the Ndab analog and the Nlys analog suggests that the presence of a basic residue in the bridging region might be advantageous for binding to the hsst5 receptor provided that the residue bearing the basic group extends far enough to allow for interaction with the receptor, while the length of the basic peptoid residue does not influence binding to the hsst2 receptor. These results are useful for the design of hsst5 selective somatostatin analogs.  相似文献   

3.
We report the conformational analysis by 1H‐nmr and computer simulations of five potent sweet molecules, N‐(3,3‐dimethylbutyl)‐L ‐aspartyl‐S‐(α‐methyl)phenylalanine methylester (1; 5000 times more potent than sucrose), L ‐aspartyl‐D ‐valine (S)‐α‐methoxycarbonylmethylbenzylamide (2; 1400 times more potent than sucrose), L ‐aspartyl‐D ‐valine α‐phenylcyclopentylamide (3; 1200 times more potent than sucrose), L ‐aspartyl‐D ‐α‐aminobutyric acid (S)‐α‐cyclohexylpropylamide (4; 2300 times more potent than sucrose), and L ‐aspartyl‐D ‐valine (R)‐α‐methylthiomethylbenzylamide (5; 3000 times more potent than sucrose). The “L‐shaped” structure, which we believe to be responsible for sweet taste, is accessible to all five sweet compounds in solution. This structure is characterized by a zwitterionic ring formed by the A‐H and B containing moieties located in the +y axis and by the hydrophobic group X pointing into the +x axis. Other accessible conformations of these flexible molecules are extended conformations with the A‐H and B containing moieties in the +y axis and the hydrophobic group X pointing in the −y axis and reversed L‐shaped structures with the hydrophobic group X projecting along the −x axis. The remarkable potency of the N‐alkylated compound 1 supports our recent hypothesis that a second hydrophobic binding domain in addition to interactions arising from the L‐shaped structure leads to an enhancement of sweetness potency. © 1999 John Wiley & Sons, Inc. Biopoly 49: 525–539, 1999  相似文献   

4.
We report the conformational analysis by 1H NMR in DMSO and computer simulations involving distance geometry and molecular dynamics simulations at 300K of peptoid analogs of the cyclic hexapeptide c-[Phe11-Pro6-Phe7-D-Trp8-Lys9-Thr10]. The analogs c-[Phe11-Nasp6-Phe7-D-Trp8-Lys9-Thr10](1), c-[Phe11-Ndab6Phe7-D-Trp8-Lys9-Thr10] (2) and c-[Phen11-Nlys6-Phe7-D-Trp8-Lys9-Thr10](3) where Nasp denotes N-(2-carboxyethyl) glycine, Ndab N-(2-aminoethyl) glycine and Nlys N-(4-aminobutyl) glycine are subject to conformational studies. The results of free and restrained molecular dynamics simulations at 300K are reported and give insight into the conformational behaviour of these analogs. The compounds show two sets of nuclear magnetic resonance signals corresponding to the cis and trans orientations of the peptide bond between residues 11 and 6. The backbone conformation of the cis isomers that we believe are the bioactive isomers of the three compounds are very similar to each other while there are larger variations amongst the trans isomers. The binding data to the isolated receptors show that the introduction of the Nlys residue in analog 3 leads to an enhancement of binding potency to the hsst5 receptor compared with analog 2 while maintaining identical binding potency to the hsst2 receptor. The Nasp6 analog 1 binds weakly to the hsst2 and is essentially inactive towards the other receptors. Comparison of the conformations and binding activities of these three analogs indicates that the Nlys residue extends sufficiently far to allow binding to a negatively charged binding domain on the hsst5 receptor. According to this model, the Ndab analog 2 cannot extend far enough to allow for binding to the receptor pocket. The loss of activity observed for the Nasp6 compound 1 indicates that the presence of a negatively charged residue in position 6 is unfavorable for binding to the hsst receptors.  相似文献   

5.
Five linear analogs of GnRH containing a p-aminophenylalanine (Pap) residue in their sequence and their six corresponding azo-bridged cyclic derivatives were synthesized. The precyclic peptides were prepared on solid-support, while azo-cyclization was performed in solution by diazotization of the p-aminophenylalanine residue followed by intramolecular coupling of the formed diazo salt with either tyrosine or histidine side chains present in the sequence. All peptides were examined for their binding ability to the GnRH receptor expressed on rat pituitary membranes and for their LH-release activity from dispersed rat pituitary cells. Linear analogs 1 i.e [Pap(5)] GnRH and 3, i.e. [Tyr(3), Pap(5)] GnRH, were found to bind to the GnRH receptors only slightly less avidly than native GnRH. Their cyclization, however, led to a marked reduction in the binding capacity, i.e. from IC(50) of 10(-9) M to the 10(-7) M range, and in biopotency, i.e. LH-release. All other linear and cyclic peptides were found to bind selectively to the GnRH receptor only in the low microM range. Only peptide 1 was found comparable to native GnRH in respect to LH-release activity and thus may potentially be a good agonist of the parent peptide. Peptides 1-4, the most potent GnRH receptor binders, were examined for their conformational properties using CD. Cyclic-azo peptides 2 and 4 were further evaluated by NMR spectroscopy in solution combined with molecular modeling. The structural information obtained explains in part the GnRH-like biological activity observed.  相似文献   

6.
High affinity and saturable binding of [125I-Tyr11]somatostatin (SS) is described in membrane homogenates from a pituitary transplantable tumor (GH4C1) rich in somatotrophs (KD for SS = 0.67 nM; Bmax = 30 fmol/mg protein). Binding characteristics and pharmacology are similar to those measured on normal pituitary membranes. The potency of various SS analogs highly correlates with that measured in in vitro bioassay for growth hormone. This suggests that those GH4C1 membranes are a good model for SS receptors on somatotrophs. Interestingly however, analogs in which the Asn5 is deleted (Des-Asn5) or D-Ser replaces Ser13 show dissociated potencies between the various assays: [D-Ser13] analogs are more potent in pituitary than in GH4C1 membranes binding assay. Des-Asn5-modified analogs are much more potent in both pituitary binding assays than in the bioassay. This could reflect a multiplicity of SS receptor subtypes in pituitary.  相似文献   

7.
Peptide T (ASTTTNYT), a segment corresponding to residues 185-192 of gp120, the coat protein of HIV, has several important biological properties in vitro that have stimulated the search for simpler and possibly more active analogs. We have previously shown that pseudocyclic hexapeptide analogs containing the central residues of peptide T retain considerable chemotactic activity. We have now extended the design of this type of analogs to peptides containing different aromatic residues and/or Ser in lieu of Thr. The complex conformation-activity relationship of these analogs called for a reexamination of the basic conformational tendencies of peptide T itself. Here, we present an exhaustive NMR conformational study of peptide T in different media. Peptide T assumes a gamma-turn in aqueous mixtures of ethylene glycol, a type-IV beta-turn conformation in aqueous mixtures of DMF, and a type-II beta-turn conformation in aqueous mixtures of DMSO. The preferred conformations for the analogs were derived from modeling, starting from the preferred conformations of peptide T. The best models derived from the gamma-turn conformation of peptide T are those of peptides XII (DSNYSR), XIII (ETNYTK) and XVI (ESNYSR). The best models derived from the type-IV beta-turn conformation of peptide T are those of peptides XIV (KTTNYE) and XV (DSSNYR). No low-energy models could be derived starting from the type-II beta-turn conformation of peptide T. The analogs with the most favored conformations are also the most active in the chemotactic test.  相似文献   

8.
We designed five ascidiacyclamide analogues [cyclo(‐Xxx1‐oxazoline2‐d ‐Val3‐thiazole4‐l ‐Ile5‐oxazoline6‐d ‐Val7‐thiazole8‐)] incorporating l ‐1‐naphthylalanine (l ‐1Nal), l ‐2‐naphthylalanine (l ‐2Nal), d ‐phenylalanine (d ‐Phe), d ‐1‐naphthylalanine (d ‐1Nal) or d ‐2‐naphthylalanine (d ‐2Nal) into the Xxx1 position of the peptide. The conformations of these analogues were then examined using 1H NMR, CD spectroscopy, and X‐ray diffraction. These analyses suggested that d ‐enantiomer‐incorporated ASCs [(d ‐Phe), (d ‐1Nal), and (d ‐2Nal)ASC] transformed from the folded to the open structure in solution more easily than l ‐enantiomer‐incorporated ASCs [(l ‐Phe), (l ‐1Nal), and (l ‐2Nal)ASC]. Structural comparison of the two analogues containing isomeric naphthyl groups showed that the 1‐naphthyl isomer induced a more stable open structure than the 2‐naphthyl isomer. In particular, [d ‐1Nal]ASC showed the most significant transformation from the folded to the open structure in solution, and exhibited the strongest cytotoxicity toward HL‐60 cells. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Potent, cyclic hexapeptide analogues of somatostatin are generally believed to adopt some common secondary structural features: a II′ β turn at one end of the cycle, and a type VI turn with a cis amide bond at the other. A proposed cis amide surrogate, the 1,5-disubstituted tetrazole, has been placed into a cyclic hexapeptide analog of somatostatin in order to constrain the putative cis amide bond. The final cyclization was done by either chemical or enzymatic means. The product, cyclo(Ala6-Tyr7-D -Trp8-Lys9-Val10-Phe11-Ψ[CN4]), was found to have 83% of the activity of somatostatin. Solution nmr analysis in DMSO/water revealed that the backbone as well as side chain χ1 and χ2 were well ordered. Relaxation matrix methods were used to extract distance restraints from the nuclear Overhauser effect spectroscopy data set, and these were used in a systematic search of torsional space to identify structures consistent with the nmr data. Restrained minimizations of these structures using a number of different force fields produced structures having the expected βII′ turn at D -Trp8-Lys9 and αβVIa turn in the Phe11-Ψ[CN4]-Ala6 portion of the molecule. The similarity of the minimized structures to those previously reported for cyclic hexapeptide analogues of somatostatin confirms the similarity of the tetrazole geometry to that of the cis amide in solution. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
Antimicrobial peptides are recognized candidates with pharmaceutical potential against epidemic emerging multi‐drug resistant bacteria. In this study, we use nuclear magnetic resonance spectroscopy and molecular dynamics simulations to determine the unknown structure and evaluate the interaction with dodecylphosphatidylcholine (DPC) and sodium dodecylsulphate (SDS) micelles with three W6‐Hylin‐a1 analogs antimicrobial peptides (HyAc, HyK, and HyD). The HyAc, HyK, and HyD bound to DPC micelles are all formed by a unique α‐helix structure. Moreover, all peptides reach the DPC micelles' core, which thus suggests that the N‐terminal modifications do not influence the interaction with zwiterionic surfaces. On the other hand, only HyAc and HyK peptides are able to penetrate the SDS micelle core while HyD remains always at its surface. The stability of the α‐helical structure, after peptide‐membrane interaction, can also be important to the second step of peptide insertion into the membrane hydrophobic core during permeabilization. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The conformation of a series of cyclic enkephalin analogues of a general formula X(1)-cyclo[Y(2)-Z(3)-Nal(4)-Leu(5)] (Nal: beta-(2-naphthyl)alanine), where X = Tyr, Phe, or Phe(NO(2)), Y = D-Dab or L-Dab (Dab: 2,4-diaminobutyric acid), and Z = D-Pro or L-Pro, was studied by means of NMR spectroscopy and theoretical conformational analysis with the Empirical Conformational Energy Program for Peptides and Proteins force field plus solvation. The NMR measurements were performed in dimethyl sulfoxide solution. The nuclear Overhauser effect intensities and coupling constants were used to compute the statistical weights of the conformations of the ensemble generated in global conformational searches. The purpose of this study was to determine whether introducing the D- or L-proline residue in position 3 can produce peptides with both rigid backbone and significant separation of the pharmacophore groups in position 1 and 4 (as required for high affinity for the mu-type opioid receptors). It was found that the analogues with D-Dab in position 2 and D-Pro in position 3 possess a stable type II' beta-turn at positions 3 and 4, which rigidifies the cyclic backbone; this finding was confirmed by independent measurements of the temperature coefficients of the amide protons, which indicated very significant screening of the Leu(5) amide proton from the solvent. However, these analogues were found to possess a short interchromophore distance. The analogues containing both Dab and Pro in the L-configuration are characterized by a larger interchromophore distance; however, they do not possess a stable beta-turn and have therefore a higher conformational flexibility. The modifications proposed in this work are therefore not likely to lead to enkephalin analogues with a high affinity for the mu-receptors.  相似文献   

12.
The three-dimensional structure of a potent SSTR3-selective analogue of somatostatin, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-Agl(8)(N(beta) Me, 2-naphthoyl)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-Agl(8)(N(beta) Me, 2-naphthoyl)]-SRIF) (peptide 1) has been determined by (1)H NMR in water and molecular dynamics (MD) simulations. The peptide exists in two conformational isomers differing mainly by the cis/trans isomerization of the side chain in residue 8. The structure of 1 is compared with the consensus structural motifs of other somatostatin analogues that bind predominantly to SSTR1, SSTR2/SSTR5 and SSTR4 receptors, and to the 3D structure of a non-selective SRIF analogue, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-2Nal(8)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-2Nal(8)]-SRIF) (peptide 2). The structural determinant factors that could explain selectivity of peptide 1 for SSTR3 receptors are discussed.  相似文献   

13.
A dipeptide taste ligand L -aspartyl-D -2-aminobutyric acid-(S)-α-ethylbenzylamide was found to be about 2000 times more potent than sucrose. To investigate the molecular basis of its potent sweet taste, we carried out conformational analysis of this molecule and several related analogues by NMR spectroscopy, computer simulations and X-ray crystallographic studies. The results of the studies support our earlier model that an ‘L’-shape molecular array is essential for eliciting sweet taste. In addition, we have identified an aromatic group located between the stem and the base of the ‘L-shape’, which is responsible for enhancement of sweetness potency. In this study, we also assessed the optimal size of the essential hydrophobic group (X) and the effects of the chirality of the second residue toward taste. ©1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
A set of terminally protected tripeptoids containing a residue of either N-methylglycine or N-isobutylglycine in position i + 1/i + 2 were synthesized and tested for intramolecularly H-bonded beta-turn formation. By exploiting FT-IR absorption and 1H NMR techniques, their folding tendencies were compared with those of a variety of reference peptides. The amount of beta-turn induction and the relative extent of the various types of intramolecularly H-bonded beta-turn conformers were determined in chloroform solution.  相似文献   

15.
The conformational analysis of N-tosyl-substituted diaza crown ethers 1 - 4 by means of NMR spectroscopy and molecular modelling is described. The structures of 1 - 4 were investigated theoretically by empirical and semiempirical methods using MSI/DISCOVER97 (ESFF force field) and MOPAC (PM3). Conclusions about preferred conformations are drawn.Electronic Supplementary Material available.  相似文献   

16.
Three linear Thr6-bradykinin analogues in which either one or both the two phenylalanine residues in the peptide sequence have been substituted by N-benzylglycine (BzlGly) and their head-to-tail cyclic analogues were synthesized and tested on an isolated rat duodenum preparation. The linear (BzlGly5,Thr6-BK, BzlGly8,Thr6-BK and BzlGly(5,8),Thr6-BK) and the cyclic (cyclo BzlGly5,Thr6-BK, cyclo BzlGly8,Thr6-BK and cyclo BzlGly(5,8),Thr6-BK) peptoid-like analogues were characterized by amino acid analysis, optical rotation, analytical HPLC and MALDI-TOF mass spectroscopy. The conformational features of both the linear and cyclic derivatives were investigated by FT-IR and CD measurements. Preliminary molecular mechanics calculations were also performed on some synthetic peptides. Pharmacological screening using the relaxation of the isolated rat duodenum preparation showed that incorporation of N-benzylglycine at positions 5 and/or 8 in the linear Thr6-BK causes a substantial decrease in potency. Comparable incorporation in cyclo Thr6-BK, at position 8, or 5 and 8, resulted in nearly inactive analogues. However, cyclo BzlGly5,Thr6-BK showed a potency which is of the same order of magnitude as for cyclo-BK and cyclo Thr6-BK.  相似文献   

17.
18.
N‐substituted glycines constitute mimics of natural amino acids that are of great interest in the peptide‐based drug development. Peptoids‐oligo(N‐substituted glycines) have been recently demonstrated to be highly active peptidomimetics in biological systems, resistant to proteolytic degradation. We developed a method of the deuterium labeling of peptidomimetics containing N‐substituted glycine residues via H/D exchange of their α‐carbon hydrogen atoms. The labeling was shown to be easy, inexpensive, and without the use of derivatization reagents or the need for a further purification. The deuterons introduced at the α‐carbon atoms do not undergo a back exchange under acidic conditions during liquid chromatography mass spectrometry (LC‐MS) analysis. The LC‐MS analysis of a mixture of isotopologues revealed a co‐elution of deuterated and nondeuterated forms of the peptidomimetics, which may be useful in the quantitative isotope dilution analysis of peptoids and other derivatives of N‐substituted glycines. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The NMR and molecular dynamics methods are used to study the conformations of a hexapeptide, GRGDTP, which has been shown to be accessible to various types of cell‐adhesion based cellular behaviors such as cell‐to‐matrix interactions, cell differentiation, immunogenicity development, gene expression, angiogenesis, metastasis, sex determination and gamete fusion. 1H‐NMR results indicate the existence of weak 5→2 hydrogen bonded β‐turn type‐III. Molecular simulation studies using a mixed protocol of distance geometry, constrained minimization, restrained molecular dynamics followed by energy minimization resulted additional conformations that include about 64% of population of inverse γ‐turn (HB, 3→1) and about 35% population of γ‐turn (HB, 4→2). The inter‐proton distances observed in γ‐and inverse γ‐turns are also consistent with the NMR constraints. The variable internal hydrogen bonding due to γ‐turns initiated at Gly 1 and Arg 2 , and its tendency to inter‐convert between γ‐and inverse γ‐turn conformations imply that the peptide is flexible in nature. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 460–471, 2013.  相似文献   

20.
The N‐terminal 1–34 segments of both parathyroid hormone (PTH) and parathyroid hormone‐related protein (PTHrP) bind and activate the same membrane receptor in spite of major differences in their amino acid sequence. The hypothesis was made that they share the same bioactive conformation when bound to the receptor. A common structural motif in all bioactive fragments of the hormone in water/trifluoroethanol mixtures or in aqueous solution containing detergent micelles is the presence of two helical segments at the N‐ and C‐termini of the sequence. In order to stabilize the helical structures, we have recently synthesized and studied the PTHrP(1–34) analog [(Lys13–As p17, Lys26–As p30)]PTHrP(1–34)NH2, which contains lactam‐constrained Lys‐Asp side chains at positions i, i+4. This very potent agonist exhibits enhanced helix stability with respect to the corresponding linear peptide and also two flexible sites at positions 12 and 19 in 1:1 trifluoroethanol/water. These structural elements have been suggested to play a critical role in bioactivity. In the present work we have extended our conformational studies on the bicyclic lactam‐constrained analog to aqueous solution. By CD, 2D‐NMR and structure calculations we have shown that in water two helical segments are present in the region of the lactam bridges (13–18, and 26–31) with high flexibility around Gly12 and Arg19. Thus, the essential structural features observed in the aqueous‐organic medium are maintained in water even if, in this solvent, the overall structure is more flexible. Our findings confirm the stabilizing effect of side‐chain lactam constraints on the α‐helical structure. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  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