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.
Several secretory and nonsecretory enzymes were localized histochemically in the main venom gland of 13 viperid snakes. All secretory cells show the intracellular oxidative enzymes succinate dehydrogenase and monoamine oxidase. The granular reactions obtained for both enzymes resemble mitochondria in distribution. Distinctive cells with a very high succinate dehydrogenase activity are dispersed among the secretory cells of all species except Atractaspis. Nonspecific acid phosphatase activity is found in the supranuclear region of the secretory cells in species that do not secrete this enzyme and throughout the cytoplasm in snakes that secrete the enzyme. Nonspecific alkaline phosphatase activity occurs in the secretory cells of those snakes whose venom shows this activity. Leucine amino peptidase (aryl amidase) activity is found in the venom and in the secretory cells of all the species. In Vipera palaestinae both the venom and the secretory cells of the main venom gland contain nonspecific esterase, L-amino acid oxidase and phosphodiesterase activities. The localization of phosphodiesterase and L-amino acid oxidase do not show major differences between glands at different intervals from an initial milking. Adenosine-monophosphate phosphatase activity is localized in the supranuclear region of the secretory cells in the glands of Vipera palaestinae and Aspis cerastes. Its activity is found in the venom of Aspis only.  相似文献   

2.
Viperid snakes show the most complex snake‐venom proteomes and offer an intriguing challenge in terms of understanding the nature of their components and the pathological outcomes of envenomation characterized by local and systemic effects. In this work, the venom complexity of eight Bothrops species was analyzed by 2‐DE, and their subproteomes of proteinases were explored by 2‐D immunostaining and 2‐D gelatin zymography, demonstrating the diversity of their profiles. Heparin, a highly sulfated glycosaminoglycan released from mast cells, is involved in anti‐coagulant and anti‐inflammatory processes. Here, we explored the hypothesis that heparin released upon envenomation could interact with toxins and interfere with venom pathogenesis. We first identified the Bothrops venom subproteome of toxins that bind with high‐affinity for heparin as composed of mainly serine proteinases and C‐type lectins. Next, we explored the Bothrops jararaca toxins that bind to heparin under physiological conditions and identified a relationship between the subproteomes of proteinases, and that of heparin‐binding toxins. Only the non‐bound fraction, composed mainly of metalloproteinases, showed lethal and hemorrhagic activities, whereas the heparin‐bound fraction contained mainly serine proteinases associated with coagulant and fibrinogenolytic activities. These data suggest that heparin binding to B. jararaca venom components in vivo has a minor protective effect to venom toxicity.  相似文献   

3.
Since the introduction of antivenom administration 120 years ago to treat venomous snake bit, it has been the gold standard for saving life and limb. However, this therapeutic approach is not always effective and not without potential life-threatening side effects. We tested a new paradigm to abrogate the plasmatic anticoagulant effects of fibrinogenolytic snake venom metalloproteinases by modification of fibrinogen with iron and carbon monoxide and by inhibiting these Zn2+ dependent metalloproteinases directly with carbon monoxide exposure. Assessment of the fibrinogenolytic effects of venoms collected from Puff adder, Gaboon viper and Indian cobra snakes on plasmatic coagulation kinetics was performed with thrombelastography. Pretreatment of plasma with iron and carbon monoxide exposure markedly attenuated the effects of all three venoms, and direct pretreatment of each venom with carbon monoxide also significantly decreased the ability to compromise coagulation. These results demonstrated that the introduction of a transition metal (e.g., modulation of the α-chain of fibrinogen with iron), modulation of transition metal in heme (e.g., carbon monoxide modulation of fibrinogen-bound heme iron), and direct inhibition of transition metal containing venom enzymes (e.g., CO binding to Zn2+ or displacing Zn2+ from the catalytic site) significantly decreased fibrinogenolytic activity. This biometal modulation strategy to attenuate the anticoagulant effects of snake venom metalloproteinases could potentially diminish hemostatic injury in envenomed patients until antivenom can be administered.  相似文献   

4.
It has been recently demonstrated that the hemotoxic venom activity of several species of snakes can be inhibited by carbon monoxide (CO) or a metheme forming agent. These and other data suggest that the biometal, heme, may be attached to venom enzymes and may be modulated by CO. A novel fibrinogenolytic metalloproteinase, named CatroxMP-II, was isolated and purified from the venom of a Crotalus atrox viper, and subjected to proteolysis and mass spectroscopy. An ion similar to the predicted singly charged m/z of heme at 617.18 was identified. Lastly, CORM-2 (tricarbonyldichlororuthenium (II) dimer, a CO releasing molecule) inhibited the fibrinogenolytic effects of CatroxMP-II on coagulation kinetics in human plasma. In conclusion, we present the first example of a snake venom metalloproteinase that is heme-bound and CO-inhibited.  相似文献   

5.
Snake venom thrombin-like enzymes (SVTLEs) are widely applied in the treatment of thrombotic diseases, however, the molecular mechanism of its inhibition by synthetic and natural proteinaceous inhibitors is not yet understood. Here we investigated effects of protease inhibitors including phenylmethylsulfonil fluoride (PMSF), benzamidine (BMD) and its derivates on the activity of recombinant gloshedobin, a SVTLE from the snake Gloydius shedaoensis. The molecular inhibition mechanism was postulated by separately docking inhibitors into three-dimensional model of gloshedobin using protein C activator from Agkistrodon contortrix contortrix venom (ACC-C, which bear 78% identity with gloshedobin) as template. The analysis indicated that the strongest inhibitor, PMSF, was via a covalent bond with the catalytic Ser195, while other inhibitors showing weaker inhibitory activity were via hydrogen bond with Ser195 or non-catalytic residues.  相似文献   

6.
Serine proteinases and Kunitz-type inhibitors are widely represented in the venoms of snakes belonging to different genera. During the studies of the venoms of snakes inhabiting Russia, we have cloned cDNAs coding for novel proteins of these families. A novel serine proteinase that we named nikobin was identified in the venom gland of the Nikolsky viper. The amino acid sequence of nikobin deduced from the cDNA sequence slightly differs from those of the serine proteinases found in other snakes, displaying 15 unique amino acid substitutions. This is the first serine proteinase from a viper of the Vipera genus for which the complete amino acid sequence has been determined. A cDNA coding for a Kunitz-type inhibitor has also been cloned. The deduced amino acid sequence of the inhibitor displays overall homology to the already known sequences of analogous proteins from vipers of the Vipera genus. However, several unusual amino acid substitutions that can cause a change of the inhibitor activity have been detected.  相似文献   

7.
Antibiotic resistance presents a real problem in which new antibacterial molecules from natural secretions could be beneficial in the development of new drugs. In this study, Cerastes cerastes venom was investigated for its antibacterial activity against Gram‐positive and Gram‐negative bacteria. The antibacterial activity was evaluated by measuring the halo inhibition and minimum inhibitory concentration (MIC). An l ‐amino acid oxidase (CcLAAO) was purified from this venom using three chromatographic steps; its homogeneity (60 kDa) was confirmed by SDS‐PAGE. LC–MS/MS analysis of CcLAAO showed similarities with other LAAO enzymes from Echis ocellatus and Viridovipera stejnegeri venoms. CcLAAO presents an antibacterial activity against three bacterial strains (Staphylococcus aureus, Methicillin‐resistant S. aureus, and Pseudomonas aeruginosa) with MIC values of 10, 10, and 20 μg/mL, respectively. However, no effect was observed against Escherichia coli and yeast strains. Kinetic parameters of CcLAAO evaluated on l ‐leucine at pH 8.0 and 20°C were Km = 0.06 mmol and Vmax = 164 mmol/min.  相似文献   

8.
The venom glands and related muscles of sea snakes conform in their general structure to those of the terrestrial elapids. The venom gland, however, is smaller in size and the accessory gland is considerably reduced. A similar pattern is found in the Australian elapid Notechis. The musculus compressor glandulae is well developed in the sea snakes and in some species its posterior-medial portion runs uninterruptedly from the origin to the insertion of the muscle. This might be considered as a primitive condition suggesting an early divergence of the sea snakes from an ancestral elapid stock. Three species of sea snakes, Aipysurus eydouxi, Emydocephalus annulatus, and E. ijimae, feed on fish eggs and have very small, but still functioning, venom glands. The reduced accessory gland of the sea snakes is apparently connected with their aquatic environment, as a similar condition is found also in the elapine Boulengerina annulata which lives in large lakes of Central Africa. The similarity in structure of the venom gland between sea snakes and Notechis scutatus may point to a possible phylogenetic relationship between this group of Australian elapids and hydrophiine snakes.  相似文献   

9.
A fibrinogenolytic proteinase from the venom of Naja nigricollis was purified by chromatography on Bio-Rex 70 and Phenyl-Sepharose. The purified enzyme, designated proteinase F1, was homogeneous by the criterion of SDS-polyacrylamide gel electrophoresis, and consisted of a single chain with a molecular weight of 58 000. Purified proteinase F1 had approximately 15-fold more proteinase activity than the crude venom, based on its ability to inactive α2-macroglobulin. The enzyme acted on only the Aα-chain of fibrinogen and left the Bβ- and γ-chains intact. The pH optimum for this fibrinogenolytic activity was in the range of pH 8 to 10. In addition to its activity on fibrinogen, proteinase F1 was active on α2-macroglobulin and fibronectin, but did not degrade casein, hemoglobin or bovine serum albumin. The enzyme was not inhibited by inhibitors of serine proteinases, cysteine proteinases or acid proteinases, but only by the metalloproteinase inhibitor, EDTA. The inhibition by EDTA could be prevented by Zn2+, but not by Ca2+ or Mg2+.  相似文献   

10.
The ants use their venom for predation, defense, and communication. The venom of these insects is rich in peptides and proteins, and compared with other animal venoms, ant venoms remain poorly explored. The objective of this study was to evaluate the protein content of the venom in the Ponerinae ant Pachycondyla striata. Venom samples were collected by manual gland reservoir dissection, and samples were submitted to two‐dimensional gel electrophoresis and separation by ion‐exchange and reverse‐phase high‐performance liquid chromatography followed by mass spectrometry using tanden matrix‐assisted laser desorption/ionization with time‐of‐flight (MALDI‐TOF/TOF) mass spectrometry and electrospray ionization‐quadrupole with time‐of‐flight (ESI‐Q/TOF) mass spectrometry for obtaining amino acid sequence. Spectra obtained were searched against the NCBInr and SwissProt database. Additional analysis was performed using PEAKS Studio 7.0 (Sequencing de novo). The venom of P. striata has a complex mixture of proteins from which 43 were identified. Within the identified proteins are classical venom proteins (phospholipase A, hyaluronidase, and aminopeptidase N), allergenic proteins (different venom allergens), and bioactive peptides (U10‐ctenitoxin Pn1a). Venom allergens are among the most expressed proteins, suggesting that P. striata venom has high allergenic potential. This study discusses the possible functions of the proteins identified in the venom of P. striata.  相似文献   

11.
One novel venom factor was isolated and purified from the venom of Taiwan habu (Trimeresurus mucrosquamatus) using two consecutive anion-exchange and gel-filtration chromatographies followed by cation-exchange HPLC. Further characterization of the purified protein indicated that it lacks the proteolytic activity toward fibrinogen molecules, suggesting that this protein factor does not belong to the familes of metalloproteinases and thrombin-like serine proteases commonly found in the crude venoms of various crotalid snakes. The purified protein exists as a native dimeric protein of 26 kDa, consisting of two closely similar subunits of 16 and 13 kDa, held together by disulfide linkage. N-Terminal sequence analysis revealed that both chains are homologous to each other at the N-terminal fragment and also similar to the factors IX/X-binding protein isolated fromTrimeresurus flavoviridis and botrocetin fromBothrops jararaca. This study points to the existence of one new two-chain venom factor without fibrinogenase activity from Taiwan habu, which, in contrast to botrocetin, promotes platelet agglutination even in the absence of von Willebrand factor. Unlike factors IX/X-binding proteins, it did not show affinity to coagulation factors IX and X in the presence of Ca2+ ion. It also shows no inhibition on thrombin, in contrast with bothrojaracin, a thrombin inhibitor isolated fromBothrops jararaca venom. We have therefore named this novel venom factortrimecetin to distinguish it from some structurally related venom factors present in various crotalid and viperid snakes.  相似文献   

12.
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   

13.
Gloshedobin, a thrombin-like enzyme from the venom of Gloydius shedaoensis, is usually produced as inclusion bodies in Escherichia coli cell. In this work, gloshedobin was separately fused with three fusion partners NusA, GST, and TrxA at its N terminus and then was expressed as fusion proteins in E. coli. The results showed that the NusA was the most efficient fusion partner to improve the solubility of recombinant gloshedobin. The purified NusA-fused gloshedobin with an overall yield of 64.6% was resolved as one band in the SDS-PAGE gel with molecular mass of about 90 kDa. Both fibrinogen clotting and fibrinogenolytic activities were found for the recombinant product. The purified NusA-fused gloshedobin exhibited amidolytic activity of 506 U/mg under optimal conditions of pH of 8.0 and 40°C. The inhibition study of NusA-fused gloshedobin by various inhibitors showed that serine protease inhibitors, phenylmethylsulphonyl fluoride, and N-tosyl-l-phenylalanine chloromethyl ketone, strongly inhibited its admidolytic activity, whereas ethylenediaminetetraacetic acid as well as heparin and hirudin did not, suggesting that NusA-fused gloshedobin exhibited the same characteristics as the native form of gloshedobin. The strategy of this work may contribute to improve the soluble expression level of other thrombin-like enzymes from snake venom in E. coli.  相似文献   

14.
A venom protein from the endoparasitic wasp, Pimpla hypochondriaca, was recently biochemically isolated. This protein possessed haemocyte anti‐aggregation activity in vitro and shares the same N‐terminal amino acid sequence as that deduced from a gene termed vpr3. The vpr3 gene was identified by sequence analysis of randomly isolated cDNAs from a P. hypochondriaca venom gland library. Presently, the gene for the full‐length sequence of mature VPr3 protein was amplified from the P. hypochondriaca venom gland cDNA library by PCR. The amplicon was directionally cloned into a pET expression vector so that recombinant VPr3 (rVPr3) would have an N‐terminal polyhistidine (His) tag. High levels of target protein expression were obtained following addition of IPTG (1 mM) and growth of the bacteria at 37°C for 5 h, or at 24°C for 20 h. Following lysis of bacteria grown at 37°C, the target protein partitioned into the insoluble fraction. However, at 24°C, a small amount of soluble protein was consistently detected. The amount of soluble rVPr3 was subsequently increased when the transformed bacteria were grown in Overnight Express Instant TB medium at 24°C. Soluble rVPr3 was purified utilizing the MagneHis Protein Purification System. Recombinant VPr3 was determined to have adverse effects on the cytoskeleton of Lacanobia oleracea haemocytes and to inhibit the ability of these cells to form aggregates in vitro. © 2009 Wiley Periodicals, Inc.  相似文献   

15.
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   

16.
Abstract Trimeresurus flavoviridis snakes inhabit the southwestern islands of Japan. A phospholipase A2 (PLA2), named PL-Y, was isolated from Okinawa T. flavoviridis venom and its amino acid sequence was determined from both protein and cDNA. PL-Y was unable to induce edema. In contrast, PLA-B, a PLA2 from Tokunoshima T. flavoviridis venom, which is different at only three positions from PL-Y, is known to induce edema. A new PLA2, named PLA-B′, which is similar to PLA-B, was cloned from Amami-Oshima T. flavoviridis venom gland. Three T. flavoviridis venom basic [Asp49]PLA2 isozymes, PL-Y (Okinawa), PLA-B (Tokunoshima), and PLA-B′ (Amami-Oshima), are identical in the N-terminal half but have one to four amino acid substitutions in the β1-sheet and its vicinity. Such interisland sequence diversities among them are due to isolation in the different environments over 1 to 2 million years and appear to have been brought about by natural selection for point mutation in their genes. Otherwise, a major PLA2, named PLA2, ubiquitously exists in the venoms of T. flavoviridis snakes from the three islands with one to three synonymous substitutions in their cDNAs. It is assumed that the PLA2 gene is a prototype among T. flavoviridis venom PLA2 isozyme genes and has hardly undergone nonsynonymous mutation as a principal toxic component. Phylogenetic analysis based on the amino acid sequences revealed that T. flavoviridis PLA2 isozymes are clearly separated into three groups, PLA2 type, basic [Asp49]PLA2 type, and [Lys49]PLA2 type. Basic [Asp49]PLA2-type isozymes may manifest their own particular toxic functions different from those of the isozymes of the PLA2 type and [Lys49]PLA2 type.  相似文献   

17.
IMP-hydrolyzing activity, which is reactive with goose anti-pig lung IMP-GMP 5′-nucleotidase (c-N-II: EC.3.1.3.5) serum, was detected in extracts from liver, heart, kidney, spleen, stomach, skeletal muscle and lung from several species of reptiles: The values found in liver (U/mg protein) of one animal were: 4.5 for an ammono-ureotelic turtle (Trionyx sinensis japonicus); 3.7 for an ureo-uricotelic tortoise (Testudo elegans); 13–23 for three species of uricotelic snakes: Elaphe quadrivirgata, Elaphe conspicillata and Elaphe climacophora. These findings suggest that in the liver of snakes, c-N-II may participate in the production of uric acid as an end product of amino acid metabolism.  相似文献   

18.
The amino acid sequence of a short-chain neurotoxin Acanthophis antarcticus c (toxin Aa c) from the venom of an Australian elapid snake, the common death adder (Acanthophis antarcticus, subfamily Acanthophiinae) was elucidated. Toxin Aa c is composed of 62 amino acid residues, including eight half-cystine residues and a cysteine residue. The amino acid sequence of toxin Aa c is homologous with those of other short-chain neurotoxins found in snakes of the family Elapidae, especially with those from snakes of the subfamily Hydrophiinae. The single cysteine residue was located in position 4. Toxin Aa c has a lethal dose (LD50) of 0.08 micrograms/g body weight of mouse on intramuscular injection.  相似文献   

19.
Aims: Alcaligenes sp. NBRC 14130 was found as a strain hydrolysing a mixture of (±)‐trans‐ and (±)‐cis ethyl chrysanthemates to (1R,3R)‐(+)‐trans‐chrysanthemic acid. The Alcaligenes cells also have hydrolytic activity for 6‐aminohexanoate‐cyclic dimer (6‐AHCD, 1,8‐diazacyclotetradecane‐2,9‐dione). The correlation of function on the enzyme from the Alcaligenes strain with hydrolysis activities for both ethyl chrysanthemate and 6‐AHCD was demonstrated. Methods and Results: The esterase was purified to homogeneity. The purified esterase hydrolysed 20 mmol l?1 ester including the four stereoisomers to the corresponding (+)‐trans acid with a 37% molar conversion of ethyl (+)‐trans chrysanthemate. The esterase showed high hydrolytic activity for various short‐chain fatty acid esters, n‐hexane amide and 6‐AHCD. The amino acid sequence of the Alcaligenes esterase was identical to that of Arthrobacter 6‐AHCD hydrolase (EC 3.5.2.12) and similar to that of fatty acid amide hydrolase (EC 3.5.1.4) from Rattus norvegicus, having both serine and lysine residues of the catalytic site and the consensus motif Gly‐X‐Ser‐X‐Gly. Conclusion: The stereo‐selective hydrolytic activity was found in Alcaligenes sp. NBRC14130 by screening of ethyl chrysanthemate‐hydrolysing activity in micro‐organisms, and the purified esterase also acted on fatty acid esters and amides. Significance and Impact of the Study: This study has demonstrated that there are great differences in the enzymatic properties, amino acid sequence and catalytic motif of esterases in both Alcaligenes and Arthrobacter globiformis with excellent stereo‐selectivity for (+)‐trans‐ethyl chrysanthemate, but the amino acid sequence of Alcaligenes esterase is identical to that of Arthrobacter 6‐AHCD hydrolase.  相似文献   

20.
Al3+‐resistant cultivars of wheat (Triticum aestivum L.) release malate through the Al3+‐activated anion transport protein Triticum aestivum aluminum‐activated malate transporter 1 (TaALMT1). Expression of TaALMT1 in Xenopus oocytes and tobacco suspension cells enhances the basal transport activity (inward and outward currents present in the absence of external Al3+), and generates the same Al3+‐activated currents (reflecting the Al3+‐dependent transport function) as observed in wheat cells. We investigated the amino acid residues involved in this Al3+‐dependent transport activity by generating a series of mutations to the TaALMT1 protein. We targeted the acidic residues on the hydrophilic C‐terminal domain of TaALMT1 and changed them to uncharged residues by site‐directed mutagenesis. These mutant proteins were expressed in Xenopus oocytes and their transport activity was measured before and after Al3+ addition. Three mutations (E274Q, D275N and E284Q) abolished the Al3+‐activated transport activity without affecting the basal transport activity. Truncation of the hydrophilic C‐terminal domain abolished both basal and Al3+‐activated transport activities. Al3+‐dependent transport activity was recovered by fusing the N‐terminal region of TaALMT1 with the C‐terminal region of AtALMT1, a homolog from Arabidopsis. These findings demonstrate that the extracellular C‐terminal domain is required for both basal and Al3+‐dependent TaALMT1 activity. Furthermore, we identified three acidic amino acids within this domain that are specifically required for the activation of transport function by external Al3+.  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