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.
Crustacean insulin‐like androgenic gland factor (IAG) of Macrobrachium rosenbergii, a heterodimeric peptide having both four disulfide bonds and an N‐linked glycan, was synthesized by the combination of solid‐phase peptide synthesis and the regioselective disulfide formation reactions. The disulfide isomer of IAG could also be synthesized by the same manner. The conformational analysis of these peptides by circular dichroism (CD) spectral measurement indicated that the disulfide bond arrangement affected the peptide conformation in IAG. On the other hand, the N‐linked glycan attached at A chain showed no effect on CD spectra of IAG. This is the first report for the chemical synthesis of insulin‐like heterodimeric glycopeptide having three interchain disulfides, and the synthetic strategy shown here might be useful for the synthesis of other glycosylated four‐disulfide insulin‐like peptides.  相似文献   

2.
The cytochrome c maturation process is carried out in the bacterial periplasm, where some specialized thiol‐disulfide oxidoreductases work in close synergy for the correct reduction of oxidized apocytochrome before covalent heme attachment. We present a structural and functional characterization of the soluble periplasmic domain of CcmG from the opportunistic pathogen P. aeruginosa (Pa‐CcmG), a component of the protein machinery involved in cyt c maturation in gram‐negative bacteria. X‐ray crystallography reveals that Pa‐CcmG is a TRX‐like protein; high‐resolution crystal structures show that the oxidized and the reduced forms of the enzyme are identical except for the active‐site disulfide. The standard redox potential was calculated to be E0′ = ?0.213 V at pH 7.0; the pKa of the active site thiols were pKa = 6.13 ± 0.05 for the N‐terminal Cys74 and pKa = 10.5 ± 0.17 for the C‐terminal Cys77. Experiments were carried out to characterize and isolate the mixed disulfide complex between Pa‐CcmG and Pa‐CcmH (the other redox active component of System I in P. aeruginosa). Our data indicate that the target disulfide of this TRX‐like protein is not the intramolecular disulfide of oxidized Pa‐CcmH, but the intermolecular disulfide formed between Cys28 of Pa‐CcmH and DTNB used for the in vitro experiments. This observation suggests that, in vivo, the physiological substrate of Pa‐CcmG may be the mixed‐disulfide complex between Pa‐CcmH and apo‐cyt. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

3.
A toxin with four disulfide bridges from Tityus serrulatus venom was able to compete with 125I‐kaliotoxin on rat brain synaptosomal preparations, with an IC50 of 46 nM . The obtained amino acid sequence and molecular mass are identical to the previously described butantoxin. Enzymatic cleavages in the native peptide followed by mass spectrometry peptide mapping analysis were used to determine the disulfide bridge pattern of α‐KTx12?1. Also, after the cleavage of the first six N‐terminal residues, including the unusual disulfide bridge which forms an N‐terminus ring, the potency of the cleaved peptide was found to decrease about 100 fold compared with the native protein. Copyright © 2003 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
AA‐NADase from Agkistrodon acutus venom is a unique multicatalytic enzyme with both NADase and AT(D)Pase activities. Among all identified NADases, only AA‐NADase contains Cu(II) and has disulfide‐bond linkages between two peptide chains. The effects of the reduction of the disulfide‐bonds and Cu(II) in AA‐NADase by small‐molecule reductants on its NADase and ADPase activities have been investigated by polyacrylamide gel electrophoresis, high performance liquid chromatography, electron paramagnetic resonance spectroscopy and isothermal titration calorimetry. The results show that AA‐NADase has six disulfide‐bonds and fifteen free cysteine residues. L‐ascorbate inhibits AA‐NADase on both NADase and ADPase activities through the reduction of Cu(II) in AA‐NADase to Cu(I), while other reductants, dithiothreitol, glutathione and tris(2‐carboxyethyl)phosphine inhibit both NADase and ADPase activities through the reduction of Cu(II) to Cu(I) and the cleavage of disulfide‐bonds in AA‐NADase. Apo‐AA‐NADase can recover its NADase and ADPase activities in the presence of 1 mM Zn(II). However, apo‐AA‐NADase does not recover any NADase or ADPase activity in the presence of 1 mM Zn(II) and 2 mM TCEP. The multicatalytic activity relies on both disulfide‐bonds and Cu(II), while Cu(I) can not activate the enzyme activities. AA‐NADase is probably only active as a dimer. The inhibition curves for both ADPase and NADase activities by each reductant share a similar trend, suggesting both ADPase and NADase activities probably occur at the same site. In addition, we also find that glutathione and L‐ascorbate are endogenous inhibitors to the multicatalytic activity of AA‐NADase. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 141–149, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

5.
Incorporation of noncanonical disulfide linkages into single‐domain antibodies (sdAbs) has been shown to enhance thermostability and other properties. Here, we evaluated the effects of introducing a novel disulfide linkage formed between Cys residues at IMGT positions 40 and 55 on the melting temperatures (T ms), reversibility of thermal unfolding, solubility, and antigen‐binding affinities of three types of sdAbs (VHH, VH, and VL domains). The Cys40‐Cys55 disulfide linkage was tolerated by 9/9 VHHs, 12/12 VHs, and 2/11 VLs tested and its formation was confirmed by mass spectrometry. Using circular dichroism, we found that the Cys40‐Cys55 disulfide linkage increased sdAb T m by an average of 10.0°C (range: 0–21.8°C). However, enhanced thermostability came at the cost of a partial loss of refolding ability upon thermal denaturation as well as, for some sdAbs, significantly decreased solubility and antigen‐binding affinity. Thus, Cys40/Cys55 can be added to the panel of known locations for introducing stabilizing noncanonical disulfide linkages into antibody variable domains, although its effects should be tested empirically for individual sdAbs.  相似文献   

6.
P‐selectin glycoprotein ligand‐1 (PSGL‐1) is a homodimeric mucin ligand that is important to mediate the earliest adhesive event during an inflammatory response by rapidly forming and dissociating the selectin‐ligand adhesive bonds. Recent research indicates that the noncovalent associations between the PSGL‐1 transmembrane domains (TMDs) can substitute for the C320‐dependent covalent bond to mediate the dimerization of PSGL‐1. In this article, we combined TOXCAT assays and molecular dynamics (MD) simulations to probe the mechanism of PSGL‐1 dimerization. The results of TOXCAT assays and Martini coarse‐grained molecular dynamics (CG MD) simulations demonstrated that PSGL‐1 TMDs strongly dimerized in a natural membrane and a leucine zipper motif was responsible for the noncovalent dimerization of PSGL‐1 TMD since mutations of the residues that occupied a or d positions in an (abcdefg)n leucine heptad repeat motif significantly reduced the dimer activity. Furthermore, we studied the effects of the disulfide bond on the PSGL‐1 dimer using MD simulations. The disulfide bond was critical to form the leucine zipper structure, by which the disulfide bond further improved the stability of the PSGL‐1 dimer. These findings provide insights to understand the transmembrane association of PSGL‐1 that is an important structural basis for PSGL‐1 preferentially binding to P‐selectin to achieve its biochemical and biophysical functions.  相似文献   

7.
Proteins that are modified by chemical conjugation require at least two separate purification processes. First the bulk protein is purified, and then after chemical conjugation, a second purification process is required to obtain the modified protein. In an effort to develop new enabling technologies to integrate bioprocessing and protein modification, we describe the use of disulfide‐bridging conjugation to conduct PEGylation during protein refolding. Preliminary experiments using a PEG‐mono‐sulfone reagent with partially unfolded leptin and unfolded RNAse T1 indicated that the cysteine thiols underwent disulfide‐bridging conjugation to give the PEGylated proteins. Interferon‐β1b (IFN‐β1b) was then expressed in E.coli as inclusion bodies and found to undergo disulfide bridging‐conjugation during refolding. The PEG‐IFN‐β1b was isolated by ion‐exchange chromatography and displayed in vitro biological activity. In the absence of the PEGylation reagent, IFN‐β1b refolding was less efficient and yielded protein aggregates. No PEGylation was observed if the cysteines on IFN‐β1b were first modified with iodoacetamide prior to refolding. Our results demonstrate that the simultaneous refolding and disulfide bridging PEGylation of proteins could be a useful strategy in the development of affordable modified protein therapeutics.  相似文献   

8.
The nematocyst walls of Hydra are formed by proteins containing small cysteine‐rich domains (CRDs) of ~25 amino acids. The first CRD of nematocyst outer all antigen (NW1) and the C‐terminal CRD of minicollagen‐1 (Mcol1C) contain six cysteines at identical sequence positions, however adopt different disulfide bonded structures. NW1 shows the disulfide connectivities C2‐C14/C6‐C19/C10‐C18 and Mcol1C C2‐C18/C6‐C14/C10‐C19. To analyze if both show structural preferences in the open, non‐disulfide bonded form, which explain the formation of either disulfide connectivity pattern, molecular dynamics (MD) simulations at different temperatures were performed. NW1 maintained in the 100‐ns MD simulations at 283 K a rather compact fold that is stabilized by specific hydrogen bonds. The Mcol1C structure fluctuated overall more, however stayed most of the time also rather compact. The analysis of the backbone Φ/ψ angles indicated different turn propensities for NW1 and Mcol1C, which mostly can be explained based on published data about the influence of different amino acid side chains on the local backbone conformation. Whereas a folded precursor mechanism may be considered for NW1, Mcol1C may fold according to the quasi‐stochastic folding model involving disulfide bond reshuffling and conformational changes, locking the native disulfide conformations. The study further demonstrates the power of MD simulations to detect local structural preferences in rather dynamic systems such as the open, non‐disulfide bonded forms of NW1 and Mcol1C, which complement published information from NMR backbone residual dipolar couplings. Because the backbone structural preferences encoded by the amino acid sequence embedding the cysteines influence which disulfide connectivities are formed, the data are generally interesting for a better understanding of oxidative folding and the design of disulfide stabilized therapeutics. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Granulins (GRNs) are a family of small (~6 kDa) proteins generated by the proteolytic processing of their precursor, progranulin (PGRN), in many cell types. Both PGRN and GRNs are implicated in a plethora of biological functions, often in opposing roles to each other. Lately, GRNs have generated significant attention due to their implicated roles in neurodegenerative disorders. Despite their physiological and pathological significance, the structure‐function relationships of GRNs are poorly defined. GRNs contain 12 conserved cysteines forming six intramolecular disulfide bonds, making them rather exceptional, even among a few proteins with high disulfide bond density. Solution NMR investigations in the past have revealed a unique structure containing putative interdigitated disulfide bonds for several GRNs, but GRN‐3 was unsolvable due to its heterogeneity and disorder. In our previous report, we showed that abrogation of disulfide bonds in GRN‐3 renders the protein completely disordered (Ghag et al., Prot Eng Des Sel 2016). In this study, we report the cellular expression and biophysical analysis of fully oxidized, native GRN‐3. Our results indicate that both E. coli and human embryonic kidney (HEK) cells do not exclusively make GRN‐3 with homogenous disulfide bonds, likely due to the high cysteine density within the protein. Biophysical analysis suggests that GRN‐3 structure is dominated by irregular loops held together only by disulfide bonds, which induced remarkable thermal stability to the protein despite the lack of regular secondary structure. This unusual handshake between disulfide bonds and disorder within GRN‐3 could suggest a unique adaptation of intrinsically disordered proteins towards structural stability.  相似文献   

10.
CIGB‐300 is a first‐in‐class synthetic peptide‐based drug of 25 amino acids currently undergoing clinical trials in cancer patients. It contains an amidated disulfide cyclic undecapeptide fused to the TAT cell‐penetrating peptide through a beta‐alanine spacer. CIGB‐300 inhibits the CK2‐mediated phosphorylation leading to apoptosis of tumor cells in vitro, and in vivo in cancer patients. Despite the clinical development of CIGB‐300, the characterization of peptide‐related impurities present in the active pharmaceutical ingredient has not been reported earlier. In the decision tree of ICHQ3A(R2) guidelines, the daily doses intake, the abundance, and the identity of the peptide‐related species are pivotal nodes that define actions to be taken (reporting, identification, and qualification). For this, purity was first assessed by reverse‐phase chromatography (>97%) and low‐abundance impurities (≤0.27%) were collected and identified by mass spectrometry. Most of the impurities were generated during peptide synthesis, the spontaneous air oxidation of the reduced peptide, and the lyophilization step. The most abundant impurity, with no biological activity, was the full‐length peptide containing Met17 transformed into a sulfoxide residue. Interestingly, parallel and antiparallel dimers of CIGB‐300 linked by 2 intermolecular disulfide bonds exhibited a higher antiproliferative activity than the CIGB‐300 monomer. Likewise, very low abundance trimers and tetramers of CIGB‐300 linked by disulfide bonds (≤0.01%) were also detected. Here we describe for the first time the presence of active dimeric species whose feasibility as novel CIGB‐300 derived entities merits further investigation.  相似文献   

11.
Thioredoxin superfamily proteins introduce disulfide bonds into substrates, catalyze the removal of disulfides, and operate in electron relays. These functions rely on one or more dithiol/disulfide exchange reactions. The flavoenzyme quiescin sulfhydryl oxidase (QSOX), a catalyst of disulfide bond formation with an interdomain electron transfer step in its catalytic cycle, provides a unique opportunity for exploring the structural environment of enzymatic dithiol/disulfide exchange. Wild‐type Rattus norvegicus QSOX1 (RnQSOX1) was crystallized in a conformation that juxtaposes the two redox‐active di‐cysteine motifs in the enzyme, presenting the entire electron‐transfer pathway and proton‐transfer participants in their native configurations. As such a state cannot generally be enriched and stabilized for analysis, RnQSOX1 gives unprecedented insight into the functional group environments of the four cysteines involved in dithiol/disulfide exchange and provides the framework for analysis of the energetics of electron transfer in the presence of the bound flavin adenine dinucleotide cofactor. Hybrid quantum mechanics/molecular mechanics (QM/MM) free energy simulations based on the X‐ray crystal structure suggest that formation of the interdomain disulfide intermediate is highly favorable and secures the flexible enzyme in a state from which further electron transfer via the flavin can occur.  相似文献   

12.
ScFv‐h3D6 is a single chain variable fragment that precludes Aβ peptide‐induced cytotoxicity by withdrawing Aβ oligomers from the amyloid pathway to the worm‐like pathway. Production of scFv molecules is not a straightforward procedure because of the occurrence of disulfide scrambled conformations generated in the refolding process. Here, we separately removed the disulfide bond of each domain and solved the scrambling problem; and then, we intended to compensate the loss of thermodynamic stability by adding three C‐terminal elongation mutations, previously described to stabilize the native fold of scFv‐h3D6. Such stabilization occurred through stabilization of the intermediate state in the folding pathway and destabilization of a different, β‐rich, intermediate state driving to worm‐like fibrils. Elimination of the disulfide bridge of the less stable domain, VL, deeply compromised the yield and increased the aggregation tendency, but elimination of the disulfide bridge of the more stable domain, VH, solved the scrambling problem and doubled the production yield. Notably, it also changed the aggregation pathway from the protective worm‐like morphology to an amyloid one. This was so because a partially unfolded intermediate driving to amyloid aggregation was present, instead of the β‐rich intermediate driving to worm‐like fibrils. When combining with the elongation mutants, stabilization of the partially unfolded intermediate driving to amyloid fibrils was the only effect observed. Therefore, the same mutations drove to completely different scenarios depending on the presence of disulfide bridges and this illustrates the relevance of such linkages in the stability of different intermediate states for folding and misfolding.  相似文献   

13.
The MarR/DUF24‐type repressor YodB controls the azoreductase AzoR1, the nitroreductase YodC and the redox‐sensing regulator Spx in response to quinones and diamide in Bacillus subtilis. Previously, we showed using a yodBCys6‐Ala mutant that the conserved Cys6 apparently contributes to the DNA‐binding activity of YodB in vivo. Here, we present data that mutation of Cys6 to Ser led to a form of the protein that was reduced in redox‐sensing in response to diamide and 2‐methylhydroquinone (MHQ) in vivo. DNA‐binding experiments indicate that YodB is regulated by a reversible thiol‐modification in response to diamide and MHQ in vitro. Redox‐regulation of YodB involves Cys6‐Cys101' intermolecular disulfide formation by diamide and quinones in vitro. Diagonal Western blot analyses confirm the formation of intersubunit disulfides in YodB in vivo that require the conserved Cys6 and either of the C‐terminal Cys101' or Cys108' residues. This study reveals a thiol‐disulfide switch model of redox‐regulation for the YodB repressor to sense electrophilic compounds in vivo.  相似文献   

14.
A relaxin‐like gonad‐stimulating peptide (RGP) in starfish was the first identified invertebrate gonadotropin responsible for final gamete maturation. An RGP ortholog was newly identified from Astropecten scoparius of the order Paxillosida. The A. scoparius RGP (AscRGP) precursor is encoded by a 354 base pair open reading frame and is a 118 amino acid (aa) protein consisting of a signal peptide (26 aa), B‐chain (21 aa), C‐peptide (47 aa), and A‐chain (24 aa). There are three putative processing sites (Lys‐Arg) between the B‐chain and C‐peptide, between the C‐peptide and A‐chain, and within the C‐peptide. This structural organization revealed that the mature AscRGP is composed of A‐ and B‐chains with two interchain disulfide bonds and one intrachain disulfide bond. The C‐terminal residues of the B‐chain are Gln‐Gly‐Arg, which is a potential substrate for formation of an amidated C‐terminal Gln residue. Non‐amidated (AscRGP‐GR) and amidated (AscRGP‐NH2) peptides were chemically synthesized and their effect on gamete shedding activity was examined using A. scoparius ovaries. Both AscRGP‐GR and AscRGP‐NH2 induced oocyte maturation and ovulation in similar dose‐dependent manners. This is the first report on a C‐terminally amidated functional RGP. Collectively, these results suggest that AscRGP‐GR and AscRGP‐NH2 act as a natural gonadotropic hormone in A. scoparius.  相似文献   

15.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
In previous parts of this study we developed procedures for the high‐efficiency chemical extraction of soluble and insoluble protein from intact Escherichia coli cells. Although high yields were obtained, extraction of recombinant protein directly from cytoplasmic inclusion bodies led to low product purity due to coextraction of soluble contaminants. In this work, a two‐stage procedure for the selective extraction of recombinant protein at high efficiency and high purity is reported. In the first stage, inclusion‐body stability is promoted by the addition of 15 mM 2‐hydroxyethyldisulfide (2‐HEDS), also known as oxidized β‐mercaptoethanol, to the permeabilization buffer (6 M urea + 3 mM ethylenediaminetetraacetate [EDTA]). 2‐HEDS is an oxidizing agent believed to promote disulfide bond formation, rendering the inclusion body resistant to solubilization in 6 M urea. Contaminating proteins are separated from the inclusion‐body fraction by centrifugation. In the second stage, disulfide bonds are readily eliminated by including reducing agent (20 mM dithiothreitol [DTT]) into the permeabilization buffer. Extraction using this selective two‐stage process yielded an 81% (w/w) recovery of the recombinant protein Long‐R3‐IGF‐I from inclusion bodies located in the cytoplasm of intact E. coli, at a purity of 46% (w/w). This was comparable to that achieved by conventional extraction (mechanical disruption followed by centrifugation and solubilization). A pilot‐scale procedure was also demonstrated using a stirred reactor and diafiltration. This is the first reported study that achieves both high extraction efficiency and selectivity by the chemical treatment of cytoplasmic inclusion bodies in intact bacterial cells. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 62: 455–460, 1999.  相似文献   

17.
We have introduced two disulfide crosslinks into the loop regions on opposite ends of the beta barrel in superfolder green fluorescent protein (GFP) in order to better understand the nature of its folding pathway. When the disulfide on the side opposite the N/C‐termini is formed, folding is 2× faster, unfolding is 2000× slower, and the protein is stabilized by 16 kJ/mol. But when the disulfide bond on the side of the termini is formed we see little change in the kinetics and stability. The stabilization upon combining the two crosslinks is approximately additive. When the kinetic effects are broken down into multiple phases, we observe Hammond behavior in the upward shift of the kinetic m‐value of unfolding. We use these results in conjunction with structural analysis to assign folding intermediates to two parallel folding pathways. The data are consistent with a view that the two fastest transition states of folding are "barrel closing" steps. The slower of the two phases passes through an intermediate with the barrel opening occurring between strands 7 and 8, while the faster phase opens between 9 and 4. We conclude that disulfide crosslink‐induced perturbations in kinetics are useful for mapping the protein folding pathway.  相似文献   

18.
Site‐specific labeling of synthetic peptides carrying N‐methoxyglycine (MeOGly) by isothiocyanate is demonstrated. A nonapeptide having MeOGly at its N‐terminus was synthesized by the solid‐phase method and reacted with phenylisothiocyanate under various conditions. In acidic solution, the reaction specifically gave a peptide having phenylthiourea structure at its N‐terminus, leaving side chain amino group intact. The synthetic human β‐defensin‐2 carrying MeOGly at its N‐terminus or the side chain amino group of Lys10 reacted with phenylisothiocyanate or fluorescein isothiocyanate also at the N‐methoxyamino group under the same conditions, demonstrating that this method is generally useful for the site‐specific labeling of linear synthetic peptides as well as disulfide‐containing peptides. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
PGR5‐LIKE PHOTOSYNTHETIC PHENOTYPE1 (PGRL1) regulates photosystem I cyclic electron flow which transiently activates non‐photochemical quenching at the onset of light. Here, we show that a disulfide‐based mechanism of PGRL1 regulated this process in vivo at the onset of low light levels. We found that PGRL1 regulation depended on active formation of key regulatory disulfides in the dark, and that PGR5 was required for this activity. The disulfide state of PGRL1 was modulated in plants by counteracting reductive and oxidative components and reached a balanced state that depended on the light level. We propose that the redox regulation of PGRL1 fine‐tunes a timely activation of photosynthesis at the onset of low light.  相似文献   

20.
Gomesin (Gm) was the first antimicrobial peptide (AMP) isolated from the hemocytes of a spider, the Brazilian mygalomorph Acanthoscurria gomesiana. We have been studying the properties of this interesting AMP, which also displays anticancer, antimalarial, anticryptococcal and anti‐Leishmania activities. In the present study, the total syntheses of backbone‐cyclized analogues of Gm (two disulfide bonds), [Cys(Acm)2,15]‐Gm (one disulfide bond) and [Thr2,6,11,15,d ‐Pro9]‐Gm (no disulfide bonds) were accomplished, and the impact of cyclization on their properties was examined. The consequence of simultaneous deletion of pGlu1 and Arg16‐Glu‐Arg18‐NH2 on Gm antimicrobial activity and structure was also analyzed. The results obtained showed that the synthetic route that includes peptide backbone cyclization on resin was advantageous and that a combination of 20% DMSO/NMP, EDC/HOBt, 60 °C and conventional heating appears to be particularly suitable for backbone cyclization of bioactive peptides. The biological properties of the Gm analogues clearly revealed that the N‐terminal amino acid pGlu1 and the amidated C‐terminal tripeptide Arg16‐Glu‐Arg18‐NH2 play a major role in the interaction of Gm with the target membranes. Moreover, backbone cyclization practically did not affect the stability of the peptides in human serum; it also did not affect or enhanced hemolytic activity, but induced selectivity and, in some cases, discrete enhancements of antimicrobial activity and salt tolerance. Because of its high therapeutic index, easy synthesis and lower cost, the [Thr2,6,11,15,d ‐Pro9]‐Gm analogue remains the best active Gm‐derived AMP developed so far; nevertheless, its elevated instability in human serum may limit its therapeutic potential. Copyright © 2012 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