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.
Controlling the dissemination of malaria requires the development of new drugs against its etiological agent, a protozoan of the Plasmodium genus. Angiotensin II and its analog peptides exhibit activity against the development of immature and mature sporozoites of Plasmodium gallinaceum. In this study, we report the synthesis and characterization of angiotensin II linear and cyclic analogs with anti‐plasmodium activity. The peptides were synthesized by a conventional solid‐phase method on Merrifield's resin using the t‐Boc strategy, purified by RP‐HPLC and characterized by liquid chromatography/ESI (+) MS (LC‐ESI(+)/MS), amino acid analysis, and capillary electrophoresis. Anti‐plasmodium activity was measured in vitro by fluorescence microscopy using propidium iodine uptake as an indicator of cellular damage. The activities of the linear and cyclic peptides are not significantly different (p < 0.05). Kinetics studies indicate that the effects of these peptides on plasmodium viability overtime exhibit a sigmoidal profile and that the system stabilizes after a period of 1 h for all peptides examined. The results were rationalized by partial least‐square analysis, assessing the position‐wise contribution of each amino acid. The highest contribution of polar amino acids and a Lys residue proximal to the C‐terminus, as well as that of hydrophobic amino acids in the N‐terminus, suggests that the mechanism underlying the anti‐malarial activity of these peptides is attributed to its amphiphilic character. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The peptide‐based quantitation accuracy and precision of LC‐ESI (QSTAR Elite) and LC‐MALDI (4800 MALDI TOF/TOF) were compared by analyzing identical Escherichia coli tryptic digests containing iTRAQ‐labeled peptides of defined abundances (1:1, 2.5:1, 5:1, and 10:1). Only 51.4% of QSTAR spectra were used for quantitation by ProteinPilot Software versus 66.7% of LC‐MALDI spectra. The average protein sequence coverages for LC‐ESI and LC‐MALDI were 24.0 and 18.2% (14.9 and 8.4 peptides per protein), respectively. The iTRAQ‐based expression ratios determined by ProteinPilot from the 57 467 ESI‐MS/MS and 26 085 MALDI‐MS/MS spectra were analyzed for measurement accuracy and reproducibility. When the relative abundances of peptides within a sample were increased from 1:1 to 10:1, the mean ratios calculated on both instruments differed by only 0.7–6.7% between platforms. In the 10:1 experiment, up to 64.7% of iTRAQ ratios from LC‐ESI MS/MS spectra failed S/N thresholds and were excluded from quantitation, while only 0.1% of the equivalent LC‐MALDI iTRAQ ratios were rejected. Re‐analysis of an archived LC‐MALDI sample set stored for 5 months generated 3715 MS/MS spectra for quantitation, compared with 3845 acquired originally, and the average ratios differed by only 3.1%. Overall, MS/MS‐based peptide quantitation performance of offline LC‐MALDI was comparable with on‐line LC‐ESI, which required threefold less time. However, offline LC‐MALDI allows the re‐analysis of archived HPLC‐separated samples.  相似文献   

3.
Tachykinin perform multiple physiological functions such as smoothing muscle contraction, vasodilation, inflammation, the processing of nerve signal, neuroprotection and neurodegeneration. Two novel tachykinin‐like peptides named tachykinin‐DR1 and ‐DR2 were identified from skin secretions of Danio rerio in current work. Their amino acid sequences were determined as SKSQHFHGLM‐NH2 and NKGEIFVGLM‐NH2, respectively. They share a conserved FXGLM‐NH2C‐terminal consensus motif. By cDNA cloning, the precursor encoding both tachykinin‐DR1 and ‐DR2 was screened from the skin cDNA library of D. rerio. Tachykinin‐DR1 and ‐DR2 share the same precursor, which is composed of 108 amino acid (aa) residues. Regarding the biological activity, tachykinin‐DRs could induce the contraction of isolated strips of guinea pig ileum just like other tackykinins. To our best knowledge, this is the first report of tachykinin from fish skin. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Aim: The study aimed for the complete purification and recharacterization of the highly hydrophobic circular bacteriocins, gassericin A and reutericin 6. Methods and Results: Gassericin A and reutericin 6 were purified to homogeneity using previously described method and reverse‐phase HPLC with an octyl column and eluents of aqueous acetonitrile and 2‐propanol. Mass analysis, N‐terminal sequencing and bacteriocin assay of the HPLC‐purified bacteriocins showed the two bacteriocins had identical seamless circular structures with the same m/z value (5651) of [M + H]+ and both had the same specific activity. d/l‐ amino acid composition analysis using two distinct methods with the chiral fluorescent derivatization reagents (+)‐1‐(9‐fluorenyl)ethyl chloroformate and o‐phthalaldehyde/N‐acetyl‐l ‐cystein revealed neither gassericin A nor reutericin 6 contained d ‐alanine residues contrary to our previous results. Conclusion: Purified gassericin A and reutericin 6 are chemically identical circular molecules containing no d ‐alanine residues. Significance and Impact of the Study: The HPLC conditions developed in this study will facilitate advanced purification and correct characterization of other highly hydrophobic bacteriocins.  相似文献   

5.
Two new oleanane‐type saponins: β‐d ‐xylopyranosyl‐(1 → 4)‐6‐deoxy‐α‐l ‐mannopyranosyl‐(1 → 2)‐1‐O‐{(3β)‐28‐oxo‐3‐[(2‐Oβ‐d ‐xylopyranosyl‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐yl}‐β‐d ‐glucopyranose ( 1 ) and 1‐O‐[(3β)‐28‐oxo‐3‐{[β‐d ‐xylopyranosyl‐(1 → 2)‐α‐l ‐arabinopyranosyl‐(1 → 6)‐2‐acetamido‐2‐deoxy‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐yl]β‐d ‐glucopyranose ( 2 ), along with two known saponins: (3β)‐3‐[(β‐d ‐Glucopyranosyl‐(1 → 2)‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐oic acid ( 3 ) and (3β)‐3‐{[α‐l ‐arabinopyranosyl‐(1 → 6)‐[β‐d ‐glucopyranosyl‐(1 → 2)]‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐oic acid ( 4 ) were isolated from the acetone‐insoluble fraction obtained from the 80% aqueous MeOH extract of Albizia anthelmintica Brongn . leaves. Their structures were identified using different NMR experiments including: 1H‐ and 13C‐NMR, HSQC, HMBC and 1H,1H‐COSY, together with HR‐ESI‐MS/MS, as well as by acid hydrolysis. The four isolated saponins and the fractions of the extract exhibited cytotoxic activity against HepG‐2 and HCT‐116 cell lines. Compound 2 showed the most potent cytotoxic activity among the other tested compounds against the HepG2 cell line with an IC50 value of 3.60μm . Whereas, compound 1 showed the most potent cytotoxic effect with an IC50 value of 4.75μm on HCT‐116 cells.  相似文献   

6.
Introduction –  Callistemon lanceolatus DC. (Myrtaceae) is a plant rich in polyphenols, and is used as anticough, antibronchitis and insecticide in folk medicine. Because of the biological importance of plant polyphenols, particularly tannins, a phytochemical study was of interest to investigate the constitutive poyphenols in the extracts of flowers and leaves. Objective –  To avoid time‐consuming methodology for isolation of a complex mixture of known metabolites, HPLC‐ESI/MS was employed for fast picking up of the new compounds followed by identification of the structures with UV and one‐ and two‐dimensional NMR. Methodology –  Flowers and leaves were separately extracted with hot aqueous methanol under reflux (70°C). Pre‐isolation of the total extracts was achieved through column chromatographic fractionation on polyamide with water–methanol for gradient elution. The main fractions were purified using repeated column chromatography on cellulose and/or Sephadex LH‐20 with suitable eluents. HPLC‐ESI/MS analyses were carried out in the single ion monitoring (SIM) and negative ion modes. The pure compounds in methanol–water (1:1) were analysed by direct infusion ESI/MS. Final structure elucidation was obtained by one‐ and two‐dimensional NMR. Results –  Two new metabolites namely quercetin 3‐O‐β‐D‐glucuronopyranoside n‐butyl ester ( 1 ) and n‐butylgallate 4‐O‐(2′,6′‐di‐O‐galloyl)‐β‐d ‐glucopyranoside ( 4 ) along with nine known ones were identified from the aqueous methanol extracts of flowers and leaves. Conclusion –  The study has shown that Callistemon lanceolatus is rich in polyphenols. HPLC‐ESI/MS may be used, in negative ion mode, as an efficient and rapid analytical tool for investigating complex plant extracts. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

7.
A novel defensin‐like antimicrobial peptide named longicornsin was isolated from the salivary glands of the hard tick, Haemaphysalis longicornis, using a 10‐kDa cut‐off Centriprep filter and reversed‐phase high‐performance liquid chromatography (RP‐HPLC). Its amino acid sequence was determined as DFGCGQGMIFMCQRRCMRLYPGSTGFCRGFRCMCDTHIPLRPPFMVG by Edman degradation. The cDNA encoding longicornsin was cloned by cDNA library screening. The predicted protein from the cDNA sequence was composed of 78 amino acids including a mature longicornsin. It showed similarity with defensin‐like peptides from other ticks by BLAST search. Different from most other tick defensin‐like peptides, longicornsin had a C‐terminal extension. Purified longicornsin exerted potent antimicrobial activities against bacteria and fungi. Interestingly, it even showed strong antimicrobial ability against drug‐resistant microorganisms and Helicobacter pylori. The results of this study indicated that longicornsin is a potential candidate for novel antimicrobial drug design.  相似文献   

8.
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   

9.
Filamentous fungi of the genus Stilbella are recognized as an abundant source of naturally occurring α‐aminoisobutyric acid‐containing peptides. The culture broth of Stilbella (Trichoderma) flavipes CBS 146.81 yielded a mixture of peptides named stilboflavins (SF), and these were isolated and separated by preparative TLC into groups named SF‐A, SF‐B, and SF‐C. Although all three of these groups resolved as single spots on thin‐layer chromatograms, HPLC analysis revealed that each of the groups represents very microheterogeneous mixtures of closely related peptides. Here, we report on the sequence analysis of SF‐C peptides, formerly isolated by preparative TLC. HPLC coupled to QqTOF‐ESI‐HRMS provided the sequences of 10 16‐residue peptides and five 19‐residue peptides, all of which were N‐terminally acetylated. In contrast to the previously described SF‐A and SF‐B peptaibols, SF‐C peptaibols contain Ser‐Alaol or Ser‐Leuol, which are rarely found as C‐termini, and repetitive Leu‐Aib‐Gly sequences, which have not been detected in peptaibols before. Taking the previously determined sequences of SF‐A and SF‐B into account, the entirety of peptides produced by S. flavipes (the ‘peptaibiome’) approaches or exceeds 100 non‐ribosomally biosynthesized peptaibiotics. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Four new steroidal glycosides, protolinckiosides A – D ( 1 – 4 , resp.), were isolated along with four previously known glycosides, 5 – 8 , from the MeOH/EtOH extract of the starfish Protoreaster lincki. The structures of 1 – 4 were elucidated by extensive NMR and ESI‐MS techniques as (3β,4β,5α,6β,7α,15α,16β,25S)‐4,6,7,8,15,16,26‐heptahydroxycholestan‐3‐yl 2‐O‐methyl‐β‐d ‐xylopyranoside ( 1 ), (3β,5α,6β,15α,24S)‐3,5,6,8,15‐pentahydroxycholestan‐24‐yl α‐l ‐arabinofuranoside ( 2 ), sodium (3β,6β,15α,16β,24R)‐29‐(β‐d ‐galactofuranosyloxy)‐6,8,16‐trihydroxy‐3‐[(2‐O‐methyl‐β‐d ‐xylopyranosyl)oxy]stigmast‐4‐en‐15‐yl sulfate ( 3 ), and sodium (3β,6β,15α,16β,22E,24R)‐28‐(β‐d ‐galactofuranosyloxy)‐6,8,16‐trihydroxy‐3‐[(2‐O‐methyl‐β‐d ‐xylopyranosyl)oxy]ergosta‐4,22‐dien‐15‐yl sulfate ( 4 ). The unsubstituted β‐d ‐galactofuranose residue at C(28) or C(29) of the side chains was found in starfish steroidal glycosides for the first time. Compounds 1 – 4 significantly decreased the intracellular reactive oxygen species (ROS) content in RAW 264.7 murine macrophages at induction by proinflammatory endotoxic lipopolysaccharide (LPS) from E. coli.  相似文献   

11.
Nonenzymatic protein glycation is caused by a Schiff's base reaction between the aldehyde groups of reducing sugars and the primary amines of proteins. These structures may undergo further Amadori rearrangement and free radical‐mediated oxidation to finally generate irreversible advanced glycation end products (AGEs). One of the factors known to modulate the glycation of proteins is glutathione, the most abundant nonprotein thiol tripeptide with the γ‐linkage, H‐Glu(Cys‐Gly‐OH)‐OH (GSH). Screening for products formed by GSH with D ‐glucose is an essential step in understanding the participation of GSH in glycation (the Maillard) reaction. Under the conditions used in these studies we observed N‐(1‐deoxy‐D ‐fructos‐1‐yl)‐pyroglutamic acid as the major glycation product formed in the mixtures of GSH and glucose in vitro. A RP HPLC/MS and tandem MS analyses of the GSH/glucose mixtures revealed that cleavage of the N‐terminal glutamic acid and the formation of pyroglutamic acid‐related Amadori product were accompanied by generation of Cys‐Gly‐derived Amadori and thiazolidine compounds. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Representative members of a group of linear, N‐acylated polypeptide antibiotics (peptaibols) containing α‐aminoisobutyric acid (Aib) and, in part, isovaline (Iva), as well as proteinogenic amino acids and a C‐terminal‐bonded 2‐amino alcohol, were treated with anhydrous trifluoroacetic acid (TFA) at 37° for 0.5–26 h. The resulting fragments were separated by HPLC and characterized by electrospray ionization collision‐induced dissociation mass spectrometry (ESI‐CID‐MS). The following 16–20‐residue peptaibols were investigated: natural, microheterogeneous mixtures of antiamoebins and alamethicin F50, uniform paracelsin A, and synthetic trichotoxin A50/E. In the natural peptides, bonds formed between Aib (Iva) and Pro (Hyp) were rapidly and selectively cleaved within 0.5 h. Furthermore, TFA esters of the C‐terminal amino alcohols were formed. Depending on time, release of C‐terminal tri‐ and tetrapeptides as well as amino acids from the major fragments was observed. Synthetic homooligopeptides, namely Z‐ and Ac‐(Aib)10‐OtBu and Z‐(Aib)7‐OtBu, were analyzed for comparison. On treatment with TFA, a regular series of Z‐(Aib)10–5‐OH from Z‐(Aib)10‐OtBu were detected within 0.5 h, and, after 3 h, release of a regular series of Z‐(Aib)7–3‐OH from Z‐(Aib)7‐OtBu were observed. Moreover, concomitant release of the series of H‐(Aib)10–3‐OH from the decapeptide occurred. From these data, a repetitive cleavage mechanism via intermediate formation of C‐terminal oxazolones on trifluoroacetolysis is proposed. Furthermore, their formation and stability in native peptaibols are correlated with subtle structural differences in protein amino acids linked to Aib. From the conspicuous concordance of the formation and abundance of regular series of trifluoroacetolytic fragments and of positive ions of the b‐series in CID‐MS, the generation of intermediate oxazolonium ions in both gas and liquid phase is concluded.  相似文献   

14.
4‐[Bis(2‐chloroethyl)amino]benzenebutanoic acid (=chlorambucil, 1 ; 2.5 mM ) was allowed to react with single‐ and double‐stranded calf thymus DNA at physiological pH (cacodylic acid, 50% base) at 37°. The DNA–chlorambucil adducts were identified by analyzing the DNA hydrolysates by NMR, UV, HPLC, LC/ESI‐MS/MS techniques as well as by spiking with authentic materials. ssDNA was more reactive than dsDNA, and the order of reactivity in ssDNA was Ade‐N1>Gua‐N7>Cyt‐N3>Ade‐N3. The most reactive site in dsDNA was Ade‐N3. The Gua‐N7 and Ade‐N3 adducts were hydrolytically labile. Ade‐N7 adduct could not be identified in the hydrolysates of ssDNA or dsDNA. The adduct Gua‐N7,N7, which consists of two units of Gua bound together with a unit derived from chlorambucil, is a cross‐linking adduct, and it was detected in the hydrolysates of ssDNA and dsDNA. Also several other adducts were detected which could be characterized by spiking with previously isolated authentic adducts or tentatively by MS. The role of chlorambucil–DNA adducts on the cytotoxicity and mutagenity of 1 is also discussed.  相似文献   

15.
MS was used to characterize the 24 kDa human growth hormone (hGH) glycoprotein isoform and determine the locus of O‐linked oligosaccharide attachment, the oligosaccharide branching topology, and the monosaccharide sequence. MALDI‐TOF/MS and ESI‐MS/MS analyses of glycosylated 24 kDa hGH tryptic peptides showed that this hGH isoform is a product of the hGH normal gene. Analysis of the glycoprotein hydrolysate by high‐performance anion‐exchange chromatography with pulsed amperometric detection and HPLC with fluorescent detection for N‐acetyl neuraminic acid (NeuAc) yielded the oligosaccharide composition (NeuAc2, N‐acetyl galactosamine1, Gal1). After β‐elimination to release the oligosaccharide from glycosylated 24 kDa hGH, collision‐induced dissociation of tryptic glycopeptide T6 indicated that there had been an O‐linked oligosaccharide attached to Thr‐60. The sequence and branching structure of the oligosaccharide were determined by ESI‐MS/MS analysis of tryptic glycopeptide T6. The mucin‐like O‐oligosaccharide sequence linked to Thr‐60 begins with N‐acetyl galactosamine and branches in a bifurcated topology with one appendage consisting of galactose followed by NeuAc and the other consisting of a single NeuAc. The oligosaccharide moiety lies in the high‐affinity binding site 1 structural epitope of hGH that interfaces with both the growth hormone and the prolactin receptors and is predicted to sterically affect receptor interactions and alter the biological actions of hGH.  相似文献   

16.
Two novel antimicrobial peptides with similarity to brevinin‐2 family are purified and characterized from the skin secretions of the frog, Rana nigrovittata. Their amino acid sequences were determined as GAFGNFLKGVAKKAGLKILSIAQCKLSGTC (brevinin‐2‐RN1) and GAFGNFLKGVAKKAGLKILSIAQCKLFGTC (brevinin‐2‐RN2), respectively, by Edman degradation. Different from brevinin‐2, which is composed of 33 amino acid residues (aa), both brevinin‐2‐RN1 and ‐RN2 contain 30 aa. Five cDNA sequences (Genbank accession numbers, EU136465‐9) encoding precursors of brevinin‐2‐RN1 and ‐RN2 were screened from the skin cDNA library of R. nigrovittata. These precursors are composed of 72 aa including a predicted signal peptide, an acidic spacer peptide, and a mature brevinin‐2‐RN. Both brevinin‐2‐RN1 and ‐RN2 showed strong antimicrobial activities against gram‐positive and gram‐negative bacteria and fungi. The current work identified and characterized two novel antimicrobial peptides with unique primary structure. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
DKP formation is a serious side reaction during the solid‐phase synthesis of peptide acids containing either Pro or Gly at the C‐terminus. This side reaction not only leads to a lower overall yield, but also to the presence in the reaction crude of several deletion peptides lacking the first amino acids. For the preparation of protected peptides using the Fmoc/tBu strategy, the use of a ClTrt‐Cl‐resin with a limited incorporation of the C‐terminal amino acid is the method of choice. The use of resins with higher loading levels leads to more impure peptide crudes. The use of HPLC‐ESMS is a useful method for analysing complex samples, such as those formed when C‐terminal Pro peptides are prepared by non‐optimized solid‐phase strategies. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
A single chiral cyclic α,α‐disubstituted amino acid, (3S,4S)‐1‐amino‐(3,4‐dimethoxy)cyclopentanecarboxylic acid [(S,S)‐Ac5cdOM], was placed at the N‐terminal or C‐terminal positions of achiral α‐aminoisobutyric acid (Aib) peptide segments. The IR and 1H NMR spectra indicated that the dominant conformations of two peptides Cbz‐[(S,S)‐Ac5cdOM]‐(Aib)4‐OEt ( 1) and Cbz‐(Aib)4‐[(S,S)‐Ac5cdOM]‐OMe (2) in solution were helical structures. X‐ray crystallographic analysis of 1 and 2 revealed that a left‐handed (M) 310‐helical structure was present in 1 and that a right‐handed (P) 310‐helical structure was present in 2 in their crystalline states. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
20.
PA3535 (EprS), an autotransporter (AT) protein of Pseudomonas aeruginosa, is predicted to contain a serine protease motif. The eprS encodes a 104.5 kDa protein with a 30‐amino‐acid‐long signal peptide, a 51.2 kDa amino‐terminal secreted passenger domain and a 50.1 kDa carboxyl‐terminal outer membrane channel formed translocator. Although the majority of AT proteins have been reported to be virulence factors, little is known about the functions of EprS in the pathogenicity of P. aeruginosa. In this study, we performed functional analyses of recombinant EprS secreted by Escherichia coli. The proteolytic activity of EprS was markedly decreased by changing Ser to Ala at position 308 or by serine protease inhibitors. EprS preferred to cleave substrates that terminated with arginine or lysine residues. Thus, these results indicate that EprS, a serine protease, displays the substrate specificity, cleaving after basic residues. We demonstrated that EprS activates NF‐κB‐driven promoters through protease‐activated receptor (PAR)‐1, ‐2 or ‐4 and induces IL‐8 production through PAR‐2 in a human bronchiole epithelial cell line. Moreover, EprS cleaved the peptides corresponding to the tethered ligand region of PAR‐1, ‐2 and ‐4 at a specific site with exposure oftheir tethered ligands. Collectively, these results suggest that EprS activates host inflammatory responses through PARs.  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