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.
A natural proline‐rich tetrapeptide cyclo‐prolyl‐leucyl‐prolyl‐phenylalanyl was prepared employing solution‐phase method of peptide synthesis through coupling of dipeptide fragments Boc‐l‐Pro‐l‐Leu‐OH and l‐Pro‐l‐Phe‐OMe which utilizes diisopropylcarbodiimide (DIPC) as coupling agent and N‐methylmorpholine (NMM) as the base. Deprotection of linear tetrapeptide unit followed by its cyclization provided a cyclopeptide, identical in all aspects to the natural molecule. Pharmacological evaluation showed cytotoxic, antifungal and antihelmintic potential of synthesized peptide against Dalton's Lymphoma Ascites (DLA) and Ehrlich's Ascites Carcinoma (EAC) cell lines, pathogenic dermatophytes and earthworms.  相似文献   

2.
A functional model complex of nickel superoxide dismutase (NiSOD) with a non‐peptide ligand which mimics the full catalytic cycle of NiSOD is unknown. Similarly, it has not been fully elucidated whether NiSOD activity is a result of an outer‐ or inner‐sphere electron‐transfer mechanism. With this in mind, two octahedral nickel(II)/(III) complexes of a bis‐tridentate N2S donor carboxamide ligand, N‐2‐phenylthiophenyl‐2′‐pyridinecarboxamide (HLPh), have been synthesized, structurally characterized, and their SOD activities examined. These complexes mimic the full catalytic cycle of NiSOD. Electrochemical experiments support an outer‐sphere electron‐transfer mechanism for their SOD activity.  相似文献   

3.
Fooling enzymes with mock amides : Analogues of apicidin, a cyclic‐tetrapeptide inhibitor of histone deacetylase (HDAC), were designed with a 1,4‐ or 1,5‐disubstituted 1,2,3‐triazole in place of a backbone amide bond to fix the bond in question in either a trans‐like or a cis‐like configuration. Thus, the binding affinity of distinct peptide conformations (see picture) could be probed. One analogue proved in some cases to be superior to apicidin as an HDAC inhibitor.

  相似文献   


4.
α‐Methyl‐L ‐proline is an α‐substituted analog of proline that has been previously employed to constrain prolyl peptide bonds in a trans conformation. Here, we revisit the cistrans prolyl peptide bond equilibrium in derivatives of α‐methyl‐L ‐proline, such as N‐Boc‐protected α‐methyl‐L ‐proline and the hexapeptide H‐Ala‐Tyr‐αMePro‐Tyr‐Asp‐Val‐OH. In Boc‐α‐methyl‐L ‐proline, we found that both cis and trans conformers were populated, whereas, in the short peptide, only the trans conformer was detected. The energy barrier for the cistrans isomerization in Boc‐α‐methyl‐L ‐proline was determined by line‐shape analysis of NMR spectra obtained at different temperatures and found to be 1.24 kcal/mol (at 298 K) higher than the corresponding value for Boc‐L ‐proline. These findings further illuminate the conformationally constraining properties of α‐methyl‐L ‐proline.  相似文献   

5.
A would‐be amide : A 1,4‐disubstituted 1,2,3‐triazole was used as a surrogate for a trans amide bond to create a library of 16 diastereomeric pseudotetrapeptides as β‐turn mimetics. High‐resolution structural analysis indicated that these scaffolds adopt distinct, rigid, conformationally homogeneous β‐turn‐like structures (see example), some of which bind somatostatin receptor subtypes selectively, and some of which show broad‐spectrum activity.

  相似文献   


6.
The incorporation of the β‐amino acid residues into specific positions in the strands and β‐turn segments of peptide hairpins is being systematically explored. The presence of an additional torsion variable about the C(α) C(β) bond (θ) enhances the conformational repertoire in β‐residues. The conformational analysis of three designed peptide hairpins composed of α/β‐hybrid segments is described: Boc‐Leu‐Val‐Val‐DPro‐β Phe ‐Leu‐Val‐Val‐OMe ( 1 ), Boc‐Leu‐Val‐β Val ‐DPro‐Gly‐β Leu ‐Val‐Val‐OMe ( 2 ), and Boc‐Leu‐Val‐β Phe ‐Val‐DPro‐Gly‐Leu‐β Phe ‐Val‐Val‐OMe ( 3 ). 500‐MHz 1H‐NMR Analysis supports a preponderance of β‐hairpin conformation in solution for all three peptides, with critical cross‐strand NOEs providing evidence for the proposed structures. The crystal structure of peptide 2 reveals a β‐hairpin conformation with two β‐residues occupying facing, non‐H‐bonded positions in antiparallel β‐strands. Notably, βVal(3) adopts a gauche conformation about the C(α) C(β) bond (θ=+65°) without disturbing cross‐strand H‐bonding. The crystal structure of 2 , together with previously published crystal structures of peptides 3 and Boc‐β Phe ‐β Phe ‐DPro‐Gly‐β Phe ‐β Phe ‐OMe, provide an opportunity to visualize the packing of peptide sheets with local ‘polar segments' formed as a consequence of reversal peptide‐bond orientation. The available structural evidence for hairpins suggests that β‐residues can be accommodated into nucleating turn segments and into both the H‐bonding and non‐H‐bonding positions on the strands.  相似文献   

7.
8.
For the first time, the existence of a substrate adduct of a nickel superoxide dismutase (NiSOD) model, based on the first nine residues from the N terminus of the active form of Streptomyces coelicolor NiSOD, has been proven and the adduct has been isolated. This adduct is based on the cyanide anion (CN?), as a substrate analogue of the superoxide anion (O2.?), and the nickel metallopeptide H‐HCDLPCGVY‐NH2‐Ni. Spectroscopic studies, including IR, UV/Vis, and liquid‐ and solid‐state NMR spectroscopy, show a single nickel‐bound cyanide anion, which is embedded in the metallopeptide structure. This complex sheds new light on the question of whether the mode of action of the NiSOD enzyme is an inner‐ or outer‐sphere mechanism. Whereas discussion was previously biased in favor of an outer‐sphere electron‐transfer mechanism due to the fact that binding of cyanide or azide moieties to the nickel active site had never been observed, our results are a clear indication in favor of the inner‐sphere electron‐transfer mechanism for the disproportionation of the O2.? ion, whereby the substrate is attached to the Ni atom in the active site of the NiSOD.  相似文献   

9.
A triple‐helix‐forming collagen model peptide, (prolyl‐trans‐4‐hydroxyprolyl‐glycyl)10 [(Pro‐Hyp‐Gly)10], and a thermosensitive elastin‐derived pentapeptide, valyl‐prolyl‐glycyl‐valyl‐glycyl (Val‐Pro‐Gly‐Val‐Gly), were copolymerized in various mole ratios using 1‐ethyl‐3‐(3‐dimethylaminopropyl)‐carbodiimide hydrochloride and 1‐hydroxybenzotriazole in dimethyl sulfoxide at 20 °C. All of the obtained polypeptides have molecular weight higher than 103 and contain a triple‐helical structure, and showed an inverse phase transition from transparent solution to turbid suspension in response to a rise in temperature. The lower critical solution temperature of the polypeptide solution decreased upon increasing the content of Val‐Pro‐Gly‐Val‐Gly. Furthermore, polypeptides containing 82–86 mol % of Val‐Pro‐Gly‐Val‐Gly in composition showed reversible gel formation, suggesting that (Pro‐Hyp‐Gly)10 acts as a hydrated unit and Val‐Pro‐Gly‐Val‐Gly acts as a thermosensitive crosslinking point. These biodegradable thermosensitive polypeptides may be useful for biomedical applications, including, as a scaffold for tissue regeneration. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 6048–6056, 2005  相似文献   

10.
The development of new bifunctional linchpins that permit the union of diverse building blocks is essential for the synthetic utility of anion relay chemistry (ARC). The design, synthesis, and validation of three vinylepoxide linchpins for through‐bond/through‐space ARC are now reported. For negative charge migration, this class of bifunctional linchpins employs initial through‐bond ARC by an SN2′ reaction, followed by through‐space ARC exploiting a 1,4‐Brook rearrangement. The trans‐disubstituted vinylepoxide linchpin yields a mixture of E/Z isomers, whereas the cis‐disubstituted and the trans‐trisubstituted vinylepoxide linchpins proceed to deliver three‐component adducts with excellent E selectivity.  相似文献   

11.
According to the earlier published synthesis of the C‐terminal nonapeptide of Trichovirin I 1B, Z‐Ser(tBu)‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 5 ), the complete tetradecapeptide Z‐Aib‐Asn(Trt)‐Leu‐Aib‐Pro‐Ser(tBu)‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 11b ), a protected Trichovirin I 1B, has now been prepared by means of the ‘azirine/oxazolone method’. With the exception of the N‐terminal Aib(1), all Aib residues were introduced by the coupling of the corresponding amino or peptide acids with 2,2‐dimethyl‐2H‐azirine‐3‐(N‐methyl‐N‐phenylamine) ( 1a ) and methyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate ( 3a ) as the Aib and Aib‐Pro synthons, respectively. Single crystals of two segments, i.e., the N‐terminal hexapeptide Z‐Aib‐Asn(Trt)‐Leu‐Aib‐Pro‐Ser(tBu)‐OMe ( 23 ) and the C‐terminal octapeptide Z‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 17 ), were obtained and their structures have been established by X‐ray crystallography. Following the same strategy, the C‐terminal nonapeptide of Trichovirin I 4A, Z‐Ala‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 26 ), was also synthesized and characterized by X‐ray crystallography.  相似文献   

12.
Electrochemical and photochemical bond‐activation steps are important for a variety of chemical transformations. We present here four new complexes, [Ru(Ln)(dmso)(Cl)]PF6 ( 1 – 4 ), where Ln is a tripodal amine ligand with 4?n pyridylmethyl arms and n?1 triazolylmethyl arms. Structural comparisons show that the triazoles bind closer to the Ru center than the pyridines. For L2, two isomers (with respect to the position of the triazole arm, equatorial or axial), trans‐ 2 sym and trans‐ 2 un, could be separated and compared. The increase in the number of the triazole arms in the ligand has almost no effect on the RuII/RuIII oxidation potentials, but it increases the stability of the Ru?Sdmso bond. Hence, the oxidation waves become more reversible from trans‐ 1 to trans‐ 4 , and whereas the dmso ligand readily dissociates from trans‐ 1 upon heating or irradiation with UV light, the Ru?S bond of trans‐ 4 remains perfectly stable under the same conditions. The strength of the Ru?S bond is not only influenced by the number of triazole arms but also by their position, as evidenced by the difference in redox behavior and reactivity of the two isomers, trans‐ 2 sym and trans‐ 2 un. A mechanistic picture for the electrochemical, thermal, and photochemical bond activation is discussed with data from NMR spectroscopy, cyclic voltammetry, and spectroelectrochemistry.  相似文献   

13.
We synthesized Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu, the peptide contained in lactoferrin (Lf), to identify the angiotensin converting enzyme (ACE) inhibition. In an attempt to know the structure‐activity relationship of this peptide, we replaced Pro (the third amino acid residues from N‐terminal) or Val (the fourth amino acid residues from N‐terminal) with Ala (neutral amino acid), Glu (acidic amino acid) or Lys (basic amino acid) to produce six peptides. From the in vitro ACE inhibition (IC50) of these synthesized peptides, the original peptide (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu) showed higher ACE inhibition than the replaced six peptides. Thus, replacement of Pro at the third amino acid residues or Val at the fourth position with Ala, Glu or Lys revealed the ACE inhibition to be lower than the original form of Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu. Otherwise, we added one peptide at the C‐terminal of Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu and found both products with an addition of Val (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu‐Val) or Ile (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu‐Ile) showing a lower ACE inhibition than the original one. The ACE inhibitions produced by both replaced peptides were without significance. Also, deletion of the last peptide at the C‐terminal (Leu‐Arg‐Pro‐Val‐Ala‐Ala) failed to produce a marked change of ACE inhibition as compared to the original one. These results suggest that Pro and Val are essential in the peptide for inhibition of ACE activity.  相似文献   

14.
An Escherichia coli whole‐cell biocatalyst for the direct hydroxylation of benzene to phenol has been developed. By adding amino acid derivatives as decoy molecules to the culture medium, wild‐type cytochrome P450BM3 (P450BM3) expressed in E.coli can be activated and non‐native substrates hydroxylated, without supplementing with NADPH. The yield of phenol reached 59 % when N‐heptyl‐l ‐prolyl‐l ‐phenylalanine (C7‐Pro‐Phe) was employed as the decoy molecule. It was shown that decoy molecules, especially those lacking fluorination, reached the cytosol of E. coli, thus imparting in vivo catalytic activity for the oxyfunctionalisation of non‐native substrates to intracellular P450BM3.  相似文献   

15.
The four TeII complexes, cis‐[TeCl2{(iPrNH)2CS}2] ( 1 ), cis‐[TeCl2{(iBuNH)2CS}2] ( 2 ), trans‐[TeCl2{(PhNMe)2CS}2] ( 3 ), and trans‐[TeCl2{(Et2N)2CS}2] ( 4 ), have been synthesised and their molecular structures solved by means of X‐ray crystallography. All four complexes are square planar, those with disubstituted thiourea ligands have a cis configuration, those with tetrasubstituted thioureas have a trans configuration. The Te–S bond lengths in 1 and 2 average 2.4994 and 2.5213 Å, respectively. The Te–Cl bonds trans to the Te–S bonds have average lengths of 2.8754 and 2.8334 Å, reflecting the trans influence of the two disubstituted thioureas. In 3 and 4 with identical ligands trans to each other, the average Te–S and Te–Cl bond lengths are 2.6834 and 2.5964 Å, respectively.  相似文献   

16.
The Cu+ catalyzed, 1,3‐dipolar cycloaddition of polyoxyethylene di(azidoalkynes), yields a mixture of the polyoxyethylene 1,5‐disubstituted fused di(1,2,3‐triazole‐1,4‐oxazines) as the major product, and the 1,4‐disubstituted mono‐(1,2,3‐triazolo) azidoalkyne crown ether.  相似文献   

17.
The crystal structure of the lipoundecapeptide amphisin, presented here as the tetrahydrate, C66H114N12O20·4H2O, originating from non‐ribosomal biosynthesis by Pseudomonas sp. strain DSS73, has been solved to a resolution of 0.65 Å. The primary structure of amphisin is β‐hydroxy­decanoyl‐d ‐Leu‐d ‐Asp‐d ‐allo‐Thr‐d ‐Leu‐d ‐Leu‐d ‐Ser‐l ‐Leu‐d ‐Gln‐l ‐Leu‐l ‐Ile‐l ‐Asp (Leu is leucine, Asp is aspartic acid, Thr is threonine, Ser is serine, Gln is glut­amine and Ile is isoleucine). The peptide is a lactone, linking Thr4 Oγ to the C‐terminal. The stereochemistry of the β‐hydroxy acid is R. The peptide is a close analogue of the cyclic lipopeptides tensin and pholipeptin produced by Pseudomonas fluorescens. The structure of amphisin is mainly helical (310‐helix), with the cyclic peptide wrapping around a hydrogen‐bonded water mol­ecule. This lipopeptide is amphiphilic and has biosurfactant and antifungal properties.  相似文献   

18.
The conformations of peptides and proteins are often influenced by glycans O‐linked to serine (Ser) or threonine (Thr). (2S,4R)‐4‐Hydroxyproline (Hyp), together with L ‐proline (Pro), are interesting targets for O‐glycosylation because they have a unique influence on peptide and protein conformation. In previous work we found that glycosylation of Hyp does not affect the N‐terminal amide trans/cis ratios (Ktrans/cis) or the rates of amide isomerization in model amides. The stereoisomer of Hyp—(2S,4S)‐4‐hydroxyproline (hyp)—is rarely found in nature, and has a different influence both on the conformation of the pyrrolidine ring and on Ktrans/cis. Glycans attached to hyp would be expected to be projected from the opposite face of the prolyl side chain relative to Hyp; the impact this would have on Ktrans/cis was unknown. Measurements of 3J coupling constants indicate that the glycan has little impact on the Cγendo conformation produced by hyp. As a result, it was found that the D ‐galactose residue extending from a Cγendo pucker affects both Ktrans/cis and the rate of isomerization, which is not found to occur when it is projected from a Cγexo pucker; this reflects the different environments delineated by the proline side chain. The enthalpic contributions to the stabilization of the trans amide isomer may be due to disruption of intramolecular interactions present in hyp; the change in enthalpy is balanced by a decrease in entropy incurred upon glycosylation. Because the different stereoisomers—Hyp and hyp—project the O‐linked carbohydrates in opposite spatial orientations, these glycosylated amino acids may be useful for understanding of how the projection of a glycan from the peptide or protein backbone exerts its influence.  相似文献   

19.
We report non‐chiral amino acid residues cis‐ and trans‐1,4‐diaminocyclohexane‐1‐carboxylic acid (cyclo‐ornithine, cO) that exhibit unprecedented stereospecific control of backbone dissociations of singly charged peptide cations and hydrogen‐rich cation radicals produced by electron‐transfer dissociation. Upon collision‐induced dissociation (CID) in the slow heating regime, peptide cations containing trans‐cO residues undergo facile backbone cleavages of amide bonds C‐terminal to trans‐cO. By contrast, peptides with cis‐cO residues undergo dissociations at several amide bonds along the peptide ion backbone. Diastereoisomeric cO‐containing peptides thus provide remarkably distinct tandem mass spectra. The stereospecific effect in CID of the trans‐cO residue is explained by syn‐facially directed proton transfer from the 4‐ammonium group at cO to the C‐terminal amide followed by neighboring group participation in the cleavage of the CO―NH bond, analogous to the aspartic acid and ornithine effects. Backbone dissociations of diastereoisomeric cO‐containing peptide ions generate distinct [bn]+‐type fragment ions that were characterized by CID‐MS3 spectra. Stereospecific control is also reported for electron‐transfer dissociation of cis‐ and trans‐cO containing doubly charged peptide ions. The stereospecific effect upon electron transfer is related to the different conformations of doubly charged peptide ions that affect the electron attachment sites and ensuing N―Cα bond dissociations.  相似文献   

20.
The design of disulfide bond mimetics is an important strategy for optimising cysteine‐rich peptides in drug development. Mimetics of the drug lead conotoxin MrIA, in which one disulfide bond is selectively replaced of by a 1,4‐disubstituted‐1,2,3‐triazole bridge, are described. Sequential copper‐catalyzed azide–alkyne cycloaddition (CuAAC; click reaction) followed by disulfide formation resulted in the regioselective syntheses of triazole–disulfide hybrid MrIA analogues. Mimetics with a triazole replacing the Cys4–Cys13 disulfide bond retained tertiary structure and full in vitro and in vivo activity as norepinephrine reuptake inhibitors. Importantly, these mimetics are resistant to reduction in the presence of glutathione, thus resulting in improved plasma stability and increased suitability for drug development.  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