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.
Scolopendra subspinipes mutilans, also known as Chinese red‐headed centipede, is a venomous centipede from East Asia and Australasia. Venom from this animal has not been researched as thoroughly as venom from snakes, snails, scorpions, and spiders. In this study, we isolated and characterized SsmTx‐I, a novel neurotoxin from the venom of S. subspinipes mutilans. SsmTx‐I contains 36 residues with four cysteines forming two disulfide bonds. It had low sequence similarity (<10%) with other identified peptide toxins. By whole‐cell recording, SsmTx‐I significantly blocked voltage‐gated K+ channels in dorsal root ganglion neurons with an IC50 value of 200 nM, but it had no effect on voltage‐gated Na+ channels. Among the nine K+ channel subtypes expressed in human embryonic kidney 293 cells, SsmTx‐I selectively blocked the Kv2.1 current with an IC50 value of 41.7 nM, but it had little effect on currents mediated by other K+ channel subtypes. Blockage of Kv2.1 by SsmTx‐I was not associated with significant alteration of steady‐state activation, suggesting that SsmTx‐I might act as a simple inhibitor or channel blocker rather than a gating modifier. Our study reported a specific Kv2.1‐blocker from centipede venom and provided a basis for future investigations of SsmTx‐I, for example on structure–function relationships, mechanism of action, and pharmacological potential. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The antiviral lectin scytovirin (SVN) contains a total of five disulfide bonds in two structurally similar domains. Previous reports provided contradictory results on the disulfide pairing in each individual domain, and we have now re‐examined the disulfide topology. N‐terminal sequencing and mass spectrometry were used to analyze proteolytic fragments of native SVN obtained at acidic pH, yielding the assignment as Cys7–Cys55, Cys20–Cys32, Cys26–Cys38, Cys68–Cys80, and Cys74–Cys86. We also analyzed the N‐terminal domain of SVN (SD1, residues 1–48) prepared by expression/oxidative folding of the recombinant protein and by chemical synthesis. The disulfide pairing in the chemically synthesized SD1 was forced into predetermined topologies: SD1A (Cys20–Cys26, Cys32–Cys38) or SD1B (Cys20–Cys32, Cys26–Cys38). The topology of native SVN was found to be in agreement with the SD1B and the one determined for the recombinant SD1 domain. Although the two synthetic forms of SD1 were distinct when subjected to chromatography, their antiviral properties were indistinguishable, having low nM activity against HIV. Tryptic fragments, the “cystine clusters” [Cys20–Cys32/Cys26–Cys38; SD1] and [Cys68–Cys80/Cys74–C‐86; SD2], were found to undergo rapid disulfide interchange at pH 8. This interchange resulted in accumulation of artifactual fragments in alkaline pH digests that are structurally unrelated to the original topology, providing a rational explanation for the differences between the topology reported herein and the one reported earlier (Bokesh et al., Biochemistry 2003;42:2578–2584). Our observations emphasize the fact that proteins such as SVN, with disulfide bonds in close proximity, require considerable precautions when being fragmented for the purpose of disulfide assignment.  相似文献   

3.
4.
α‐Conotoxins are peptide neurotoxins that selectively inhibit various subtypes of nicotinic acetylcholine receptors. They are important research tools for studying numerous pharmacological disorders, with profound potential for developing drug leads for treating pain, tobacco addiction, and other conditions. They are characterized by the presence of two disulfide bonds connected in a globular arrangement, which stabilizes a bioactive helical conformation. Despite extensive structure–activity relationship studies that have produced α‐conotoxin analogs with increased potency and selectivity towards specific nicotinic acetylcholine receptor subtypes, the efficient production of diversity‐oriented α‐conotoxin combinatorial libraries has been limited by inefficient folding and purification procedures. We have investigated the optimized conditions for the reliable folding of α‐conotoxins using simplified oxidation procedures for use in the accelerated production of synthetic combinatorial libraries of α‐conotoxins. To this end, the effect of co‐solvent, redox reagents, pH, and temperature on the proportion of disulfide bond isomers was determined for α‐conotoxins exhibiting commonly known Cys loop spacing frameworks. In addition, we have developed high‐throughput ‘semi‐purification’ methods for the quick and efficient parallel preparation of α‐conotoxin libraries for use in accelerated structure–activity relationship studies. Our simplified procedures represent an effective strategy for the preparation of large arrays of correctly folded α‐conotoxin analogs and permit the rapid identification of active hits directly from high‐throughput pharmacological screening assays. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
Chicken cystatin (cC) mutant I66Q is located in the hydrophobic core of the protein and increases the propensity for amyloid formation. Here, we demonstrate that under physiological conditions, the replacement of Ile with the Gln in the I66Q mutant increases the susceptibility for the disulfide bond Cys71–Cys81 to be reduced when compared to the wild type (WT) cC. Molecular dynamics (MD) simulations under conditions favoring cC amyloid fibril formation are in agreement with the experimental results. MD simulations were also performed to investigate the impact of disrupting the Cys71–Cys81 disulfide bond on the conformational stability of cC at the atomic level, and highlighted major disruption to the cC appendant structure. Domain swapping and extensive unfolding has been proposed as one of the possible mechanisms initiating amyloid fibril formation by cystatin. Our in silico studies suggest that disulfide bond formation between residues Cys95 and Cys115 is necessary to maintain conformational stability of the I66Q mutant following breakage of the Cys71–Cys81 disulfide bridge. Subsequent breakage of disulfide bond Cys95–Cys115 resulted in large structural destabilization of the I66Q mutant, which increased the α–β interface distance and expanded the hydrophobic core. These experimental and computational studies provide molecular-level insight into the relationship between disulfide bond formation and progressive unfolding of amyloidogenic cC mutant I66Q.

An animated Interactive 3D Complement (I3DC) is available in Proteopedia at http://proteopedia.org/w/Journal:JBSD:23  相似文献   

6.
There are many examples of bioactive, disulfide‐rich peptides and proteins whose biological activity relies on proper disulfide connectivity. Regioselective disulfide bond formation is a strategy for the synthesis of these bioactive peptides, but many of these methods suffer from a lack of orthogonality between pairs of protected cysteine (Cys) residues, efficiency, and high yields. Here, we show the utilization of 2,2′‐dipyridyl diselenide (PySeSePy) as a chemical tool for the removal of Cys‐protecting groups and regioselective formation of disulfide bonds in peptides. We found that peptides containing either Cys(Mob) or Cys(Acm) groups treated with PySeSePy in trifluoroacetic acid (TFA) (with or without triisopropylsilane (TIS) were converted to Cys‐S–SePy adducts at 37 °C and various incubation times. This novel Cys‐S–SePy adduct is able to be chemoselectively reduced by five‐fold excess ascorbate at pH 4.5, a condition that should spare already installed peptide disulfide bonds from reduction. This chemoselective reduction by ascorbate will undoubtedly find utility in numerous biotechnological applications. We applied our new chemistry to the iodine‐free synthesis of the human intestinal hormone guanylin, which contains two disulfide bonds. While we originally envisioned using ascorbate to chemoselectively reduce one of the formed Cys‐S–SePy adducts to catalyze disulfide bond formation, we found that when pairs of Cys(Acm) residues were treated with PySeSePy in TFA, the second disulfide bond formed spontaneously. Spontaneous formation of the second disulfide is most likely driven by the formation of the thermodynamically favored diselenide (PySeSePy) from the two Cys‐S–SePy adducts. Thus, we have developed a one‐pot method for concomitant deprotection and disulfide bond formation of Cys(Acm) pairs in the presence of an existing disulfide bond.  相似文献   

7.
Kaerner A  Rabenstein DL 《Biochemistry》1999,38(17):5459-5470
alpha-Conotoxin GI is a 13 residue snail toxin peptide cross-linked by Cys2-Cys7 and Cys3-Cys13 disulfide bridges. The formation of the two disulfide bonds by thiol/disulfide exchange with oxidized glutathione (GSSG) has been characterized. To characterize formation of the first disulfide bond in each of the two pathways by which the two disulfide bonds can form, two model peptides were synthesized in which Cys3 and Cys13 (Cono-1) or Cys2 and Cys7 (Cono-2) were replaced by alanines. Equilibrium constants were determined for formation of the single disulfide bonds of Cono-1 and Cono-2, and an overall equilibrium constant was measured for formation of the two disulfide bonds of alpha-conotoxin GI in pH 7.00 buffer and in pH 7. 00 buffer plus 8 M urea using concentrations obtained by HPLC analysis of equilibrium thiol/disulfide exchange reaction mixtures. The results indicate a modest amount of cooperativity in the formation of the second disulfide bond in both of the two-step pathways by which alpha-conotoxin GI folds into its native structure at pH 7.00. However, when considered in terms of the reactive thiolate species, the results indicate substantial cooperativity in formation of the second disulfide bond. The solution conformational and structural properties of Cono-1, Cono-2, and alpha-conotoxin GI were studied by 1H NMR to identify structural features which might facilitate formation of the disulfide bonds or are induced by formation of the disulfide bonds. The NMR data indicate that both Cono-1 and Cono-2 have some secondary structure in solution, including some of the same secondary structure as alpha-conotoxin GI, which facilitates formation of the second disulfide bond by thiol/disulfide exchange. However, both Cono-1 and Cono-2 are considerably less structured than alpha-conotoxin GI, which indicates that formation of the second disulfide bond to give the Cys2-Cys7, Cys3-Cys13 pairing induces considerable structure into the backbone of the peptide.  相似文献   

8.
We have determined the three-dimensional structure of a two-disulfide intermediate (Cys(8)-Cys(20), Cys(14)-Cys(26)) on the oxidative folding pathway of the cyclotide MCoTI-II. Cyclotides have a range of bioactivities and, because of their exceptional stability, have been proposed as potential molecular scaffolds for drug design applications. The three-dimensional structure of the stable two-disulfide intermediate shows for the most part identical secondary and tertiary structure to the native state. The only exception is a flexible loop, which is collapsed onto the protein core in the native state, whereas in the intermediate it is more loosely associated with the remainder of the protein. The results suggest that the native fold of the peptide does not represent the free energy minimum in the absence of the Cys(1)-Cys(18) disulfide bridge and that although there is not a large energy barrier, the peptide must transiently adopt an energetically unfavorable state before the final disulfide can form.  相似文献   

9.
Aqualysin I is a heat-stable alkaline serine protease produced by Thermus aquaticus YT-1. Aqualysin I comprises 281 amino acid residues and contains four cysteine residues. The cysteine residues seemed to form disulfide bonds in the molecule. Thus, the positions of the disulfide bonds were investigated. Disulfide bond-containing peptides were identified by peptide mapping with HPLC before and after carboxymethylation of chymotryptic peptides of aqualysin I. The disulfide bond-containing peptides were isolated and then carboxymethylated. Carboxymethylcysteine-containing peptides were purified, and their amino acid compositions and sequences were determined. Based on the data obtained and the primary structure of aqualysin I, it was concluded that two disulfide bonds were formed between Cys67 and Cys99, and between Cys163 and Cys194.  相似文献   

10.
We examined the effect of a novel disulfide bond engineered in subtilisin E from Bacillus subtilis based on the structure of a thermophilic subtilisin-type serine protease aqualysin I. Four sites (Ser163/Ser194, Lys170/Ser194, Lys170/Glu195, and Pro172/Glu195) in subtilisin E were chosen as candidates for Cys substitutions by site-directed mutagenesis. The Cys170/Cys195 mutant subtilisin formed a disulfide bond in B. subtilis, and showed a 5-10-fold increase in specific activity for an authentic peptide substrate for subtilisin, N-succinyl-L-Ala-L-Ala-L-Pro-L-Phe-p-nitroanilide, compared with the single-Cys mutants. However, the disulfide mutant had a 50% decrease in catalytic efficiency due to a smaller k(cat) and was thermolabile relative to the wild-type enzyme, whereas it was greatly stabilized relative to its reduced form. These results suggest that an electrostatic interaction between Lys170 and Glu195 is important for catalysis and stability in subtilisin E. Interestingly, the disulfide mutant was found to be more stable in polar organic solvents, such as dimethylformamide and ethanol, than the wild-type enzyme, even under reducing conditions; this is probably due to the substitution of uncharged Cys by charged surface residues (Lys170 and Glu195). Further, the amino-terminal engineered disulfide bond (Gly61Cys/Ser98Cys) and the mutation Ile31Leu were introduced to enhance the stability and catalytic activity. A prominent 3-4-fold increase in the catalytic efficiency occurred in the quintet mutant enzyme over the range of dimethylformamide concentration (up to 40%).  相似文献   

11.
The mechanism by which a disordered peptide nucleates and forms amyloid is incompletely understood. A central domain of β‐amyloid (Aβ21–30) has been proposed to have intrinsic structural propensities that guide the limited formation of structure in the process of fibrillization. In order to test this hypothesis, we examine several internal fragments of Aβ, and variants of these either cyclized or with an N‐terminal Cys. While Aβ21–30 and variants were always monomeric and unstructured (circular dichroism (CD) and nuclear magnetic resonance spectroscopy (NMRS)), we found that the addition of flanking hydrophobic residues in Aβ16–34 led to formation of typical amyloid fibrils. NMR showed no long‐range nuclear overhauser effect (nOes) in Aβ21–30, Aβ16–34, or their variants, however. Serial 1H‐15N‐heteronuclear single quantum coherence spectroscopy, 1H‐1H nuclear overhauser effect spectroscopy, and 1H‐1H total correlational spectroscopy spectra were used to follow aggregation of Aβ16–34 and Cys‐Aβ16–34 at a site‐specific level. The addition of an N‐terminal Cys residue (in Cys‐Aβ16–34) increased the rate of fibrillization which was attributable to disulfide bond formation. We propose a scheme comparing the aggregation pathways for Aβ16–34 and Cys‐Aβ16–34, according to which Cys‐Aβ16–34 dimerizes, which accelerates fibril formation. In this context, cysteine residues form a focal point that guides fibrillization, a role which, in native peptides, can be assumed by heterogeneous nucleators of aggregation.  相似文献   

12.
The membrane fusion protein of murine leukemia virus is a trimer of a disulfide-linked peripheral-transmembrane (SU-TM) subunit complex. The intersubunit disulfide bond is in SU linked to a disulfide bond isomerization motif, CXXC, with which the virus controls its fusion reaction (M. Wallin, M. Ekstr?m, and H. Garoff, EMBO J. 23:54-65, 2004). Upon receptor binding the isomerase rearranges the intersubunit disulfide bond into a disulfide bond isomer within the motif. This facilitates SU dissociation and fusion activation in the TM subunit. In the present study we have asked whether furin cleavage of the Env precursor potentiates the isomerase to be triggered. To this end we accumulated the late form of the precursor, gp90, in the cell by incubation in the presence of a furin-inhibiting peptide. The isomerization was done by NP-40 incubation or by a heat pulse under alkylation-free conditions. The cells were lysed in the presence of alkylator, and the precursor was immunoprecipitated, gel isolated, deglycosylated, and subjected to complete trypsin digestion. Disulfide-linked peptide complexes were separated by sodium dodecyl sulfate-tricine-polyacrylamide gel electrophoresis under nonreducing conditions. This assay revealed the size of the characteristic major disulfide-linked peptide complex that differentiates the two isomers of the disulfide bond between Cys336 (or Cys339) and Cys563, i.e., the bond corresponding to the intersubunit disulfide bond. The analyses showed that the isomerase was five- to eightfold more resistant to triggering in the precursor than in the mature, cleaved form. This suggests that the isomerase becomes potentiated for triggering by a structural change in Env that is induced by furin cleavage in the cell.  相似文献   

13.
Many studies have been performed on venomous peptides derived from animals. However, little of this research has focused on peptides from centipede venoms. Here, a venom gland cDNA library was suc-cessfully constructed for the centipede Scolopendra subspinipes mutilans. A new cDNA encoding the precursor of a venom peptide, named SsmTx, was cloned from the venomous gland cDNA library of the centipede S. subspinipes mutilans. The full-length SsmTx cDNA sequence is 465 nt, including a 249 nt ORF, a 45 nt 5′ UTR and a 171 nt 3′ UTR. There is a signal tail AATAAA 31 nt upstream of the poly (A) tail. The precursor nucleotide sequence of SsmTx encodes a signal peptide of 25 residues and a mature peptide of 57 residues, which is bridged by two pairs of disulfide bonds. SsmTx displays a unique cysteine motif that is completely different from that of other venomous animal toxins. This is the first reported cDNA sequence encoding a venom peptide from the centipede S. subspinipes mutilans.  相似文献   

14.
The primary structure of human IGF-I, except for the disulfide bond system, has been reported by Rinderknecht and Humbel. IGF-I afforded the corresponding characteristic peptide fragment on V8 protease digestion, which contained Cys6, Cys47, Cys48, and Cys52. Two possible fragments, Type I with Cys6-Cys47 and Cys48-Cys52, and Type II with Cys6-Cys48 and Cys47-Cys52, were synthesized. The disulfide bond system of IGF-I was unequivocally determined to be the Type II form along with Cys18-Cys61. Interestingly, the Type I system was included in the disulfide bond isomer produced as the main by-product in the refolding step on IGF-I synthesis by the recombinant DNA method.  相似文献   

15.
We prepared two dissected fragments of hen lysozyme and examined whether or not these two fragments associated to form a native-like structure. One (Fragment I) is the peptide fragment Asn59-homoserine-105 containing Cys64-Cys80 and Cys76-Cys94. The other (Fragment II) is the peptide fragment Lys1-homoserine-58 connected by two disulfide bridges, Cys6-Cys127 and Cys30-Cys115, to the peptide fragment Asn106-Leu129. It was found that the Fragment I immobilized in the cuvette formed an equimolar complex with Fragment II (K(d) = 3.3x10(-4) M at pH 8 and 25 degrees C) by means of surface plasmon resonance. Moreover, from analyses by circular dichroism spectroscopy and ion-exchange chromatography of the mixture of Fragments I and II at pH 8 under non-reducing conditions, it was suggested that these fragments associated to give the native-like structure. However, the mutant Fragment I in which Cys64-Cys80 and Cys76-Cys94 are lacking owing to the mutation of Cys to Ala, or the mutant fragment in which Trp62 is mutated to Gly, did not form the native-like species with Fragment II, because the mutant Fragment I derived from mutant lysozymes had no local conformation due to mutations. Considering our previous results where the preferential oxidation of two inside disulfide bonds, Cys64-Cys80 and Cys76-Cys94, occurred in the refolding of the fully reduced Fragment I, we suggest that the peptide region corresponding to Fragment I is an initiation site for hen lysozyme folding.  相似文献   

16.
17.
The attachment protein or G protein of the A2 strain of human respiratory syncytial virus (RSV) was digested with trypsin and the resultant peptides separated by reverse-phase high-performance liquid chromatography (HPLC). One tryptic peptide produced a mass by matrix-assisted laser desorption/ionization (MALDI) time-of-flight (TOF) mass spectrometry (MS) corresponding to residues 152-187 with the four Cys residues of the ectodomain (residues 173, 176, 182, and 186) in disulfide linkage and absence of glycosylation. Sub-digestion of this tryptic peptide with pepsin and thermolysin produced peptides consistent with disulfide bonds between Cys173 and Cys186 and between Cys176 and Cys182. Analysis of ions produced by post-source decay of a peptic peptide during MALDI-TOF-MS revealed fragmentation of peptide bonds with minimal fission of an inter-chain disulfide bond. Ions produced by this unprecedented MALDI-induced post-source fragmentation corroborated the existence of the disulfide arrangement deduced from mass analysis of proteolysis products. These findings indicate that the ectodomain of the G protein has a non-glycosylated subdomain containing a "cystine noose."  相似文献   

18.
The Class A family of guanine nucleotide-binding protein (G protein)-coupled receptors that includes receptors for motilin, ghrelin, and growth hormone secretagogue (GHS) has substantial potential importance as drug targets. Understanding of the molecular basis of hormone binding and receptor activation should provide insights helpful in the development of such drugs. We previously reported that Cys residues and the perimembranous residues in the extracellular loops and amino-terminal tail of the motilin receptor are critical for peptide ligand, motilin, binding and biological activity. In the current work, we focused on the predicted extracellular domains of the human GHS receptor 1a, and identified functionally important residues by using sequential deletions ranging from one to twelve amino acid residues and site-directed replacement mutagenesis approach. Each construct was transiently expressed in COS cells, and characterized for ghrelin- and growth hormone releasing peptide (GHRP)-6-stimulated intracellular calcium responses and ghrelin radioligand binding. Cys residues in positions 116 and 198 in the first and second extracellular loops and the perimembranous Glu187 residue in the second extracellular loop were critical for ghrelin and GHRP-6 biological activity. These results suggest that Cys residues in the extracellular domains in this family of Class A G protein-coupled receptor is likely involved in the highly conserved and functionally important disulfide bond, and that the perimembranous residues contribute peptide ligand binding and signaling.  相似文献   

19.
Selective reduction on the Cys28-Cys32 disulfide of Ophiophagus hannah neurotoxins, Oh-4 and Oh-5, revealed that isomerization of this disulfide linkage caused the two toxins to have distinct conformation and different retention time on a reversed-phase column. The Cys28-Cys32 disulfide of Oh-4 and Oh-5 was prone to form mixed disulfides with glutathione following pseudo-first-order kinetics. In addition to glutathionylated proteins, Oh-4 could be promoted to convert into Oh-5 by thiol compounds. Isomerization of Oh-5 into Oh-4 was not observed in the presence of thiol compounds. Dethiolation of glutathionylated proteins produced Oh-4 and Oh-5. Oxidation of the partially reduced toxin with reduced Cys28 and Cys32 was exclusively converted into Oh-5 regardless of the absence or presence of GSH/GSSG. Acrylamide quenching studies revealed difference in degree of exposure of the single Trp27 between Oh-4 and Oh-5. Synthesized peptides with substitution of Trp27 or Phe31 with Gly abolished entirely the formation of disulfide-linked dimeric product noted with the peptide of wild-type sequence. These results suggest that disulfide formation and isomerization of Cys28-Cys32 could be regulated by thiolation, and that the bulky aromatic residues Trp27 and Phe31 facilitate favorably the occurrence of disulfide isomerization of Cys28-Cys32.  相似文献   

20.
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   

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