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.
IsCT1‐NH2 is a cationic antimicrobial peptide isolated from the venom of the scorpion Opisthacanthus madagascariensis that has a tendency to form an α‐helical structure and shows potent antimicrobial activity and also inopportunely shows hemolytic effects. In this study, five IsCT1 (ILGKIWEGIKSLF)‐based analogs with amino acid modifications at positions 1, 3, 5, or 8 and one analog with three simultaneous substitutions at the 1, 5, and 8 positions were designed. The net charge of each analog was between +2 and +3. The peptides obtained were characterized by mass spectrometry and analyzed by circular dichroism for their structure in different media. Studies of antimicrobial activity, hemolytic activity, and stability against proteases were also carried out. Peptides with a substitution at position 3 or 5 ([L]3‐IsCT1‐NH2, [K]3‐IsCT1‐NH2, or [F]5‐IsCT1‐NH2) showed no significant change in an activity relative to IsCT1‐NH2. The addition of a proline residue at position 8 ([P]8‐IsCT1‐NH2) reduced the hemolytic activity as well as the antimicrobial activity (MIC ranging 3.13‐50 μmol L?1), and the addition of a tryptophan residue at position 1 ([W]1‐IsCT1‐NH2) increased the hemolytic activity (MHC = 1.56 μmol L?1) without an improvement in antimicrobial activity. The analog [A]1[F]5[K]8‐IsCT1‐NH2, which carries three simultaneous modifications, presented increasing or equivalent values in antimicrobial activity (MIC approximately 0.38 and 12.5 μmol L?1) with a reduction in hemolytic activity. In addition, this analog presented the best resistance against proteases. This kind of strategy can find functional hotspots in peptide molecules in an attempt to generate novel potent peptide antibiotics.  相似文献   

2.
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   

3.
N‐[1‐(4‐(4‐fluorophenyl)‐2,6‐dioxocyclohexylidene)ethyl] (Fde) protected amino acids have been prepared and applied in solid‐phase peptide synthesis monitored by gel‐phase 19F NMR spectroscopy. The Fde protective group could be cleaved with 2% hydrazine or 5% hydroxylamine solution in DMF as determined with gel‐phase 19F NMR spectroscopy. The dipeptide Ac‐L ‐Val‐L ‐Val‐NH2 12 was constructed using Fde‐L ‐Val‐OH and no noticeable racemization took place during the amino acid coupling with N,N′‐diisopropylcarbodiimide and 1‐hydroxy‐7‐azabenzotriazole or Fde deblocking. To extend the scope of Fde protection, the hydrophobic nonapeptide LLLLTVLTV from the signal sequence of mucin MUC1 was successfully prepared using Fde‐L ‐Leu‐OH at diagnostic positions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Linear cationic α‐helical antimicrobial peptides are promising chemotherapeutics. Most of them act by different mechanisms, making it difficult to microorganisms acquiring resistance. Decoralin is an example of antimicrobial peptide; it was described by Konno et al. and presented activity against microorganisms, but with pronounced hemolytic activity. We synthesized leucine‐substituted decoralin analogs designed based on important physicochemical properties, which depend on the maintenance of the amphiphilic α‐helical tendency of the native molecule. Peptides were synthesized, purified, and characterized, and the conformational studies were performed. The results indicated that the analogs presented both higher therapeutic indexes, but with antagonistic behavior. While [Leu]10‐Dec‐NH2 analog showed similar activity against different microorganisms (c.a. 0.4–0.8 μmol L?1), helical structuration, and some hemolytic activity, [Leu]8‐Dec‐NH2 analog did not tend to helical structure and presented antimicrobial activities two orders higher than the other two peptides analyzed. On the other hand, this analog showed to be the less hemolytic (MHC value = 50.0 μmol L?1). This approach provided insight for understanding the effects of the leucine substitution in the amphiphilic balance. They led to changes on the conformational tendency, which showed to be important for the mechanism of action and affecting antimicrobial and hemolytic activities. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
The broad‐spectrum activity of antimicrobial peptides (AMPs) and low probability of development of host resistance make them excellent candidates as novel bio‐control agents. A number of AMPs are found to be cationic, and a small proportion of these are tryptophan‐rich. The puroindolines (PIN) are small, basic proteins found in wheat grains with proposed roles in biotic defence of seeds and seedlings. Synthetic peptides based on their unique tryptophan‐rich domain (TRD) display antimicrobial properties. Bacterial endospores and biofilms are highly resistant cells, with significant implications in both medical and food industries. In this study, the cationic PIN TRD‐based peptides PuroA (FPVTWRWWKWWKG‐NH2) and Pina‐M (FSVTWRWWKWWKG‐NH2) and the related barley hordoindoline (HIN) based Hina (FPVTWRWWTWWKG‐NH2) were tested for effects on planktonic cells and biofilms of the common human pathogens including Pseudomonas aeruginosa, Listeria monocytogenes and the non‐pathogenic Listeria innocua. All peptides showed significant bactericidal activity. Further, PuroA and Pina‐M at 2 × MIC prevented initial biomass attachment by 85–90% and inhibited >90% of 6‐h preformed biofilms of all three organisms. However Hina, with a substitution of Lys‐9 with uncharged Thr, particularly inhibited Listeria biofilms. The PIN based peptides were also tested against vegetative cells and endospores of Bacillus subtilis. The results provided evidence that these tryptophan‐rich peptides could kill B. subtilis even in sporulated state, reducing the number of viable spores by 4 log units. The treated spores appeared withered under scanning electron microscopy. The results establish the potential of these tryptophan‐rich peptides in controlling persistent pathogens of relevance to food industries and human health. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Urotensin II (U‐II) is a disulfide bridged peptide hormone identified as the ligand of a G‐protein‐coupled receptor. Human U‐II (H‐Glu‐Thr‐Pro‐Asp‐c[Cys‐Phe‐Trp‐Lys‐Tyr‐Cys]‐Val‐OH) has been described as the most potent vasoconstrictor compound identified to date. We have recently identified both a superagonist of human U‐II termed P5U (H‐Asp‐c[Pen‐Phe‐Trp‐Lys‐Tyr‐Cys]‐Val‐OH) and the compound termed urantide (H‐Asp‐c[Pen‐Phe‐d ‐Trp‐Orn‐Tyr‐Cys]‐Val‐OH), which is the most potent UT receptor peptide antagonist described to date. In the present study, we have synthesized four analogues of P5U and urantide in which the Trp7 residue was replaced by the highly constrained l ‐Tpi and d ‐Tpi residues. The replacement of the Trp7 by Tpi led to active analogues. Solution NMR analysis allowed improving the knowledge on conformation–activity relationships previously reported on UT receptor ligands. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The interaction with model membranes of a peptide, EqtII1–32, corresponding to the N‐terminal region of the pore‐forming toxin equinatoxin II (EqtII) has been studied using solid‐state NMR and molecular dynamics (MD) simulations. The distances between specifically labeled nuclei in [19F‐para]Phe16‐[1‐13C]Leu19 and [19F‐para]Phe16‐[15N]Leu23 analogs of EqtII1–32 measured by REDOR in lyophilized peptide were in agreement with published crystal and solution structures. However, in both DMPC and mixed DMPC:SM membrane environments, significant changes in the distances between the labeled amino acid pairs were observed, suggesting changes in helical content around the experimentally studied region, 16–23, in the presence of bilayers. 19F‐31P REDOR experiments indicated that the aromatic ring of Phe16 is in contact with lipid headgroups in both membrane environments. For the DMPC:SM mixed bilayers, a closer interaction between Phe16 side chains and lipid headgroups was observed, but an increase in distances was observed for both labeled amino acid pairs compared with those measured for EqtII1–32 in pure DMPC bilayers. The observed differences between DMPC and DMPC:SM bilayers may be due to the greater affinity of EqtII for the latter. MD simulations of EqtII1–32 in water, on a pure DMPC bilayer, and on a mixed DMPC:SM bilayer indicate significant peptide secondary structural differences in the different environments, with the DMPC‐bound peptide adopting helical formations at residues 16–24, whereas the DMPC:SM‐bound peptide exhibits a longer helical stretch, which may contribute to its enhanced activity against PC:SM compared with pure PC bilayers. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
The effects of hydrophobic thickness and the molar phosphatidylglycerol (PG) content of lipid bilayers on the structure and membrane interaction of three cationic antimicrobial peptides were examined: aurein 2.2, aurein 2.3 (almost identical to aurein 2.2, except for a point mutation at residue 13), and a carboxy C-terminal analog of aurein 2.3. Circular dichroism results indicated that all three peptides adopt an α-helical structure in the presence of a 3:1 molar mixture of 1,2-dimyristoyl-sn-glycero-3-phosphocholine/1,2-dimyristoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (DMPC/DMPG), and 1:1 and 3:1 molar mixtures of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine/1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (POPC/POPG). Oriented circular dichroism data for three different lipid compositions showed that all three peptides were surface-adsorbed at low peptide concentrations, but were inserted into the membrane at higher peptide concentrations. The 31P solid-state NMR data of the three peptides in the DMPC/DMPG and POPC/POPG bilayers showed that all three peptides significantly perturbed lipid headgroups, in a peptide or lipid composition-dependent manner. Differential scanning calorimetry results demonstrated that both amidated aurein peptides perturbed the overall phase structure of DMPC/DMPG bilayers, but perturbed the POPC/POPG chains less. The nature of the perturbation of DMPC/DMPG bilayers was most likely micellization, and for the POPC/POPG bilayers, distorted toroidal pores or localized membrane aggregate formation. Calcein release assay results showed that aurein peptide-induced membrane leakage was more severe in DMPC/DMPG liposomes than in POPC/POPG liposomes, and that aurein 2.2 induced higher calcein release than aurein 2.3 and aurein 2.3-COOH from 1:1 and 3:1 POPC/POPG liposomes. Finally, DiSC35 assay data further delineated aurein 2.2 from the others by showing that it perturbed the lipid membranes of intact S. aureus C622 most efficiently, whereas aurein 2.3 had the same efficiency as gramicidin S, and aurein 2.3-COOH was the least efficient. Taken together, these data show that the membrane interactions of aurein peptides are affected by the hydrophobic thickness of the lipid bilayers and the PG content.  相似文献   

9.
The cell‐penetrating peptide Tat (48–60) (GRKKRRQRRRPPQ) derived from HIV‐1 Tat protein showed potent antibacterial activity (MIC: 2–8 µM ). To investigate the effect of dimerization of Tat (48–60) analog, [Tat(W): GRKKRRQRRRPWQ‐NH2], on antimicrobial activity and mechanism of bactericidal action, its dimeric peptides, di‐Tat(W)‐C and di‐Tat(W)‐K, were synthesized by a disulfide bond linkage and lysine linkage of monomeric Tat(W), respectively. From the viewpoint of a weight basis and the monomer concentration, these dimeric peptides displayed almost similar antimicrobial activity against six bacterial strains tested but acted more rapidly against Staphylococcus aureus on kinetics of bactericidal activity, compared with monomeric Tat(W). Unlike monomeric Tat(W), these dimeric peptides significantly depolarized the cytoplasmic membrane of intact S. aureus cells at MIC and induced dye leakage from bacterial‐membrane‐mimicking egg yolk L ‐α‐phosphatidylethanolamine/egg yolk L ‐α‐phosphatidyl‐DL ‐glycerol (7:3, w/w) vesicles. Furthermore, these dimeric peptides were less effective to translocate across lipid bilayers than monomeric Tat(W). These results indicated that the dimerization of Tat analog induces a partial change in the mode of its bactericidal action from intracellular target mechanism to membrane‐targeting mechanism. Collectively, our designed dimeric Tat peptides with high antimicrobial activity and rapid bactericidal activity appear to be excellent candidates for future development as novel antimicrobial agents. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The synthesis of new dermorphin analogues is described. The (R)‐alanine or phenylalanine residues of natural dermorphin were substituted by the corresponding α‐methyl‐β‐azidoalanine or α‐benzyl‐β‐azido(1‐piperidinyl)alanine residues. The potency and selectivity of the new analogues were evaluated by a competitive receptor binding assay in rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). The most active analogue in this series, Tyr‐(R)‐Ala‐(R)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 and its epimer were analysed by 1H and 13C NMR spectroscopy and restrained molecular dynamics simulations. The dominant conformation of the investigated peptides depended on the absolute configuration around Cα in the α‐benzyl‐β‐azidoAla residue in position 3. The (R) configuration led to the formation of a type I β‐turn, whilst switching to the (S) configuration gave rise to an inverse β‐turn of type I′, followed by the formation of a very short β‐sheet. The selectivity of Tyr‐(R)‐Ala‐(R) and (S)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 was shown to be very similar; nevertheless, the two analogues exhibited different conformational preferences. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
We have previously shown that leucine to lysine substitution(s) in neutral synthetic crown ether containing 14-mer peptide affect the peptide structure and its ability to permeabilize bilayers. Depending on the substitution position, the peptides adopt mainly either a α-helical structure able to permeabilize dimyristoylphosphatidylcholine (DMPC) and dimyristoylphosphatidylglycerol (DMPG) vesicles (nonselective peptides) or an intermolecular β-sheet structure only able to permeabilize DMPG vesicles (selective peptides). In this study, we have used a combination of solid-state NMR and Fourier transform infrared spectroscopy to investigate the effects of nonselective α-helical and selective intermolecular β-sheet peptides on both types of bilayers. 31P NMR results indicate that both types of peptides interact with the headgroups of DMPC and DMPG bilayers. 2H NMR and Fourier transform infrared results reveal an ordering of the hydrophobic core of bilayers when leakage is noted, i.e., for DMPG vesicles in the presence of both types of peptides and DMPC vesicles in the presence of nonselective peptides. However, selective peptides have no significant effect on the ordering of DMPC acyl chains. The ability of these 14-mer peptides to permeabilize lipid vesicles therefore appears to be related to their ability to increase the order of the bilayer hydrophobic core.  相似文献   

12.
In an attempt to develop non‐ATP‐competitive inhibitors of the autophosphorylation of IR, the effects of the synthetic peptides, Ac‐DIY1158ET‐NH2 and Ac‐DY1162Y1163RK‐NH2, on the phosphorylation of IR were studied in vitro. The peptides were derived from the amino‐acid sequence in the activation loop of IR. They inhibited the autophosphorylation of IR to 20.5 and 40.7%, respectively, at 4000 µM . The Asp/Asn‐ and Glu/Gln‐substituted peptides, Ac‐NIYQT‐NH2 and Ac‐NYYRK‐NH2, more potently inhibited the autophosphorylation than did the corresponding parent peptides. The inhibitory potencies of the substituted peptides were decreased with increasing concentrations of ATP, indicating that these peptides employ an ATP‐competitive mechanism in inhibiting the autophosphorylation of IR. In contrast, those of the parent peptides were not affected. Mass spectrometry showed that the parent peptides were phosphorylated by IR, suggesting that they interact with the catalytic loop. Moreover, docking simulations predicted that the substituted peptides would interact with the ATP‐binding region of IR, whereas their parent peptides would interact with the catalytic loop of IR. Thus, Ac‐DIYET‐NH2 and Ac‐DYYRK‐NH2 are expected to be non‐ATP‐competitive inhibitors. These peptides could contribute to the development of a drug employing a novel mechanism. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
The binding of spermine and ifenprodil to the amino terminal regulatory (R) domain of the N‐methyl‐D ‐aspartate receptor was studied using purified regulatory domains of the NR1, NR2A and NR2B subunits, termed NR1‐R, NR2A‐R and NR2B‐R. The R domains were over‐expressed in Escherichia coli and purified to near homogeneity. The Kd values for binding of [14C]spermine to NR1‐R, NR2A‐R and NR2B‐R were 19, 140, and 33 μM, respectively. [3H]Ifenprodil bound to NR1‐R (Kd, 0.18 μM) and NR2B‐R (Kd, 0.21 μM), but not to NR2A‐R at the concentrations tested (0.1–0.8 μM). These Kd values were confirmed by circular dichroism measurements. The Kd values reflected their effective concentrations at intact NR1/NR2A and NR1/NR2B receptors. The results suggest that effects of spermine and ifenprodil on NMDA receptors occur through binding to the regulatory domains of the NR1, NR2A and NR2B subunits. The binding capacity of spermine or ifenprodil to a mixture of NR1‐R and NR2A‐R or NR1‐R and NR2B‐R was additive with that of each individual R domain. Binding of spermine to NR1‐R and NR2B‐R was not inhibited by ifenprodil and vice versa, indicating that the binding sites for spermine and ifenprodil on NR1‐R and NR2B‐R are distinct.  相似文献   

14.
Overexpression of gonadotropin‐releasing hormone (GnRH) receptor in many tumors but not in normal tissues makes it possible to use GnRH analogs as targeting peptides for selective delivery of cytotoxic agents, which may help to enhance the uptake of anticancer drugs by cancer cells and reduce toxicity to normal cells. The GnRH analogs [d ‐Cys6, desGly10, Pro9‐NH2]‐GnRH, [d ‐Cys6, desGly10, Pro9‐NHEt]‐GnRH, and [d ‐Cys6, α‐aza‐Gly10‐NH2]‐GnRH were conjugated with doxorubicin (Dox), respectively, through N‐succinimidyl‐3‐maleimidopropionate as a linker to afford three new GnRH‐Dox conjugates. The metabolic stability of these conjugates in human serum was determined by RP‐HPLC. The antiproliferative activity of the conjugates was examined in GnRH receptor‐positive MCF‐7 human breast cancer cell line by MTT assay. The three GnRH‐Dox conjugates showed improved metabolic stability in human serum in comparison with AN‐152. The antiproliferative effect of conjugate II ([d ‐Cys6, desGly10, Pro9‐NHEt]‐GnRH‐Dox) on MCF‐7 cells was higher than that of conjugate I ([d ‐Cys6, desGly10, Pro9‐NH2]‐GnRH‐Dox) and conjugate III ([d ‐Cys6, α‐aza‐Gly10‐NH2]‐GnRH‐Dox), and the cytotoxicity of conjugate II against GnRH receptor‐negative 3T3 mouse embryo fibroblast cells was decreased in comparison with free Dox. GnRH receptor inhibition test suggested that the antiproliferative activity of conjugate II might be due to the cellular uptake mediated by the targeting binding of [d ‐Cys6‐des‐Gly10‐Pro9‐NHEt]‐GnRH to GnRH receptors. Our study indicates that targeting delivery of conjugate II mediated by [d ‐Cys6‐des‐Gly10‐Pro9‐NHEt]‐GnRH is a promising strategy for chemotherapy of tumors that overexpress GnRH receptors.  相似文献   

15.
A procedure for the synthesis of a11C‐labeled oligopeptide containing [1‐11C]1,2,3,4‐tetrahydro‐β‐carboline‐3‐carboxylic acid ([1‐11C]Tpi) from the corresponding Trp?HCl‐containing peptides has been developed involving a Pictet‐Spengler reaction with [11C]formaldehyde. The synthesis of [1‐11C]Tpi from Trp and [11C]formaldehyde was examined as a model reaction with the aim of developing a facile and effective method for the labeling of peptides with carbon‐11. The Pictet‐Spengler reaction of Trp and [11C]formaldehyde in acidic media (TsOH or HCl) afforded the desired [1‐11C]Tpi in a moderate radiochemical yield. Herein, the application of a Pictet‐Spengler reaction to an aqueous solution of Trp?HCl gave the desired product with a radiochemical yield of 45.2%. The RGD peptide cyclo[Arg‐Gly‐Asp‐D‐Tyr‐Lys] was then selected as a substrate for the labeling reaction with [11C]formaldehyde. The radiolabeling of a Trp?HCl‐containing RGD peptide using the Pictet‐Spengler reaction was successful. Furthermore, the remote‐controlled synthesis of a [1‐11C]Tpi‐containing RGD peptide was attempted by using an automatic production system to generate [11C]CH3I. The radiochemical yield of the [1‐11C]Tpi‐containing RGD at the end of synthesis (EOS) was 5.9 ± 1.9% (n = 4), for a total synthesis time of about 35 min. The specific activity was 85.7 ± 9.4 GBq/µmol at the EOS. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Glucose‐dependent insulinotropic polypeptide (gastric inhibitory polypeptide, or GIP), a 42‐amino acid incretin hormone, modulates insulin secretion in a glucose‐concentration‐dependent manner. Its insulinotropic action is highly dependent on glucose concentration that surmounts the hypoglycemia side effects associated with current therapy. In order to develop a GIP‐based anti‐diabetic therapy, it is essential to establish the 3D structure of the peptide and study its interaction with the GIP receptor (GIPR) in detail. This will give an insight into the GIP‐mediated insulin release process. In this article, we report the solution structure of GIP(1–42, human)NH2 deduced by NMR and the interaction of the peptide with the N‐terminus of GIPR using molecular modelling methods. The structure of GIP(1–42, human)NH2 in H2O has been investigated using 2D‐NMR (DQF‐COSY, TOCSY, NOESY, 1H‐13C HSQC) experiments, and its conformation was built by constrained MD simulations with the NMR data as constraints. The peptide in H2O exhibits an α‐helical structure between residues Ser8 and Asn39 with some discontinuity at residues Gln29 to Asp35; the helix is bent at Gln29. This bent gives the peptide an ‘L’ shape that becomes more pronounced upon binding to the receptor. The interaction of GIP with the N‐terminus of GIPR was modelled by allowing GIP to interact with the N‐terminus of GIPR under a series of decreasing constraints in a molecular dynamics simulation, culminating with energy minimization without application of any constraints on the system. The canonical ensemble obtained from the simulation was subjected to a detailed energy analysis to identify the peptide–protein interaction patterns at the individual residue level. These interaction energies shed some light on the binding of GIP with the GIPR N‐terminus in a quantitative manner. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
M2 transmembrane domain channel (M2‐TMD) permeation properties are studied using molecular dynamics simulations of M2‐TMD (1NYJ) embedded in a lipid bilayer (DMPC) with 1 mol/kg NaCl or KCl saline solution. This study allows examination of spontaneous cation and anion entry into the selectivity filter. Three titration states of the M2‐TMD tetramer are modeled for which the four His37 residues, forming the selectivity filter, are net uncharged, +2 charged, or +3 charged. M2‐TMD structural properties from our simulations are compared with the properties of other models extracted from NMR and X‐ray studies. During 10 ns simulations, chloride ions occasionally occupy the positively‐charged selectivity filter region, and from umbrella sampling simulations, Cl? has a lower free‐energy barrier in the selectivity‐filter region than either Na+ or NH, and NH has a lower free‐energy barrier than Na+. For Na+ and Cl?, the free‐energy barriers are less than 5 kcal/mol, suggesting that the 1NYJ conformation would probably not be exquisitely proton selective. We also point out a rotameric configuration of Trp41 that could fully occlude the channel. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

18.
The interaction of the synthetic antimicrobial peptide P5 (KWKKLLKKPLLKKLLKKL-NH2) with model phospholipid membranes was studied using solid-state NMR and circular dichroism (CD) spectroscopy. P5 peptide had little secondary structure in buffer, but addition of large unilamellar vesicles (LUV) composed of dimyristoylphosphatidylcholine (DMPC) increased the β-sheet content to ~20%. Addition of negatively charged LUV, DMPC–dimyristoylphosphatidylglycerol (DMPG) 2:1, led to a substantial (~40%) increase of the α-helical conformation. The peptide structure did not change significantly above and below the phospholipid phase transition temperature. P5 peptide interacted differently with DMPC bilayers with deuterated acyl chains (d54-DMPC) and mixed d54-DMPC–DMPG bilayers, used to mimic eukaryotic and prokaryotic membranes, respectively. In DMPC vesicles, P5 peptide had no significant interaction apart from slightly perturbing the upper region of the lipid acyl chain with minimum effect at the terminal methyl groups. By contrast, in the DMPC–DMPG vesicles the peptide increased disorder throughout the entire acyl chain of DMPC in the mixed bilayer. P5 promoted disordering of the headgroup of neutral membranes, observed by 31P NMR. However, no perturbations in the T 1 relaxation nor the T 2- values were observed at 30°C, although a slight change in the dynamics of the headgroup at 20°C was noticeable compared with peptide-free vesicles. However, the P5 peptide caused similar perturbations of the headgroup of negatively charged vesicles at both temperatures. These data correlate with the non-haemolytic activity of the P5 peptide against red blood cells (neutral membranes) while inhibiting bacterial growth (negatively charged membranes).  相似文献   

19.
Protaetiamycine is an insect defensin, derived from the larvae of the beetle Protaetia brevitarsis. In our previous work, we designed 9‐mer peptide analogs of protaetiamycine, including 9Pbw2 (RLWLAIKRR‐NH2), 9Pbw3 (RLWLAIWRR‐NH2), and 9Pbw4 (RLWLAWKRR‐NH2). 9Pbw2 and 9Pbw4 showed high antimicrobial activity without cytotoxicity, while 9Pbw3 with higher hydrophobicity compared to 9Pbw2 and 9Pbw4 showed high cytotoxicity as well as high antimicrobial activity (Shin et al., J. Pept. Sci. 2009; 15: 559–568). In this study, we investigated the anti‐inflammatory activities of 9Pbw2, 9Pbw3, and 9Pbw4 by quantitation of NO production in LPS‐stimulated RAW264.7 cells. The results showed that only 9Pbw3 has strong inhibition of NO production, implying that Trp7 as well as optimum level of hydrophobicity may play key roles in the anti‐inflammatory activity of 9Pbw3. In order to design potent anti‐inflammatory peptide with lower cytotoxicity as well as high stability from cleavage by protease compared to 9Pbw3, we synthesized 9Pbw3‐D , the all‐D ‐amino acid analog of 9Pbw3. 9Pbw3‐D showed less cytotoxicity against RAW264.7 cells as well as considerably stronger inhibition of NO production and inflammation‐induced cytokine production in LPS‐stimulated RAW264.7 cells than 9Pbw3. 9Pbw3‐D inhibited the gene expression of inflammatory‐induced cytokine significantly more than 9Pbw3 and showed high resistance to proteolytic digestion. Binding of 9Pbw3‐D with LPS caused higher enhancement of the FITC fluorescence as a result of its stronger interaction with LPS compared to that of 9Pbw3 and this result is in good agreement with their anti‐inflammatory activities. 9Pbw3‐D with higher anti‐inflammatory activity as well as lower cytotoxicity against mammalian cell compared to 9Pbw3 can be a potent noncytotoxic antibiotic candidates. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Gomesin (Gm) has a broad antimicrobial activity making it of great interest for development of drugs. In this study, we analyzed three Gm analogs, [Trp1]‐Gm, [Trp7]‐Gm, and [Trp9]‐Gm, in an attempt to gain insight into the contributions of different regions of the peptide sequence to its activity. The incorporation of the tryptophan residue in different positions has no effect on the antimicrobial and hemolytic activities of the Gm analogs in relation to Gm. Spectroscopic studies (circular dichroism, fluorescence and absorbance) of Gm and its analogs were performed in the presence of SDS, below and above its critical micelle concentration (CMC) (~8 mM), in order to monitor structural changes induced by the interaction with this anionic surfactant (0–15 mM). Interestingly, we found that the analogs interact more strongly with SDS at low concentrations (0.3‐6.0 mM) than close to or above its CMC. This suggests that SDS monomers are able to cover the whole peptide, forming large detergent‐peptide aggregates. On the other hand, the peptides interact differently with SDS micelles, inserting partially into the micelle core. Among the peptides, Trp in position 1 becomes more motionally‐restricted in the presence of SDS, probably because this residue is located at the N‐terminal region, which presents higher conformational freedom to interact stronger with SDS molecules. Trp residues in positions 7 and 9, close to and in the region of the turn of the molecule, respectively, induced a more constrained structure and the compounds cannot insert deeper into the micelle core or be completely buried by SDS monomers. Copyright © 2014 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