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.
Precursor in cotranslational secretion of diphtheria toxin.   总被引:14,自引:6,他引:8       下载免费PDF全文
By extracellular labeling of peptides of intact Corynebacterium diphtheriae, followed by fractionation of the cells and chain completion by isolated polysomes, it is shown that diphtheria toxin is formed and secreted cotranslationally by membrane-bound polysomes; free polysomes from none. Moreover, when the chains on these polysomes were completed in vitro, in the absence of membrane they were found to include not only diphtheria toxin of a molecular weight of 62,000, but also a larger precursor of a molecular weight of 68,000. The precursor was identified by several properties: immune precipitation; conversion into toxin fragments A and B; adenosine diphosphate ribosyl-transferase activity after activation with trypsin; and cleavage to 62,000 daltons by membrane enzymes. The precursor yields an N-terminal A fragment with a broadened molecular weight distribution, compared with that from authentic toxin, thus supporting the expectation that the extra segment of the precursor is N-terminal.  相似文献   

2.
Tryptic cleavage of EF-2, molecular mass 93 kDa, produced an 82-kDa polypeptide and a 10-kDa fragment, which was further degraded. By a slower reaction the 82-kDa polypeptide was gradually split into a 48-kDa and a 34-kDa fragment. Similarly, treatment with chymotrypsin resulted in the formation of an 82-kDa polypeptide and a small fragment. In contrast to the tryptic 82-kDa polypeptide the corresponding chymotryptic cleavage product was relatively resistant to further attack. The degradation of the 82-kDa polypeptide with either trypsin or chymotrypsin was facilitated by the presence of guanosine nucleotides, indicating a conformational shift in native EF-2 upon nucleotide binding. No effect was observed in the presence of ATP, indicating that the effect was specific for guanosine nucleotides. After affinity labelling of native EF-2 with oxidized [3H]GTP and subsequent trypsin treatment the radioactivity was recovered in the 48-kDa polypeptide showing that the GTP-binding site was located within this part of the factor. Correspondingly, tryptic degradation of EF-2 labelled with [14C]NAD+ in the presence of diphtheria toxin showed that the site of ADP-ribosylation was within the 34-kDa polypeptide. By cleavage with the tryptophan-specific reagent N-chlorosuccinimide the site of ADP-ribosylation could be located at a distance of 40-60 kDa from the GTP-binding site and about 4-11 kDa from the nearest terminus.  相似文献   

3.
Photoaffinity labeling experiments with diphtheria toxin fragment A have implicated glutamic acid 148 as a constituent of the NAD binding site. To evaluate the role of this residue in ADP-ribosylation of elongation factor 2, we replaced it with aspartic acid by in vitro mutagenesis of a toxin gene fragment cloned in Escherichia coli. Fragment A containing aspartic acid at position 148 had less than 0.6% the ADP-ribosylation activity of wild-type fragment A. The mutation produced no change in sensitivity of fragment A to trypsin and little, if any, reduction in affinity of fragment A for NAD. These results indicate that glutamic acid 148 is essential for the ADP-ribosylation of elongation factor 2 and are consistent with other data suggesting that this residue may be at or near the catalytic center of the toxin.  相似文献   

4.
The complete amino acid sequence of fragment B from diphtheria toxin has been determined. The polypeptide chain was split with cyanogen bromide, o-iodosobenzoic acid, clostripain and trypsin; all amino acid sequence analyses were made by automated Edman degradation. Fragment B, which corresponds to the carboxy terminus of the toxin molecule, contains 342 amino acids and has an Mr of 37240. The proposed amino acid sequence fully confirms the structure recently deduced from the nucleotide sequence of the structural gene. The complete sequence is analyzed in relationship with the role of fragment B in the transfer of diphtheria toxin fragment A from the extracellular medium into the cell cytoplasm.  相似文献   

5.
Monoclonal antibodies to the diphtheria toxin were produced without cross reactivity with the thermolabile toxin (LT) from Escherichia coli; ricin; choleraic toxin; the SeA, SeB, SeE, SeI, and SeG toxins of staphylococcus; the lethal factor of the anthrax toxin; and the protective antigen of the anthrax toxin. A pair of antibodies for the quantitative determination of the diphtheria toxin in the sandwich variation of enzyme-linked immunosorbent assay (ELISA) was chosen. The determination limit of the toxin was 0.7 ng/ml in plate and 1.6 ng/ml in microchip ELISA. The presence of a secretion from the nasopharynx lavage did not decrease the sensitivity of the toxin determination by sandwich ELISA. The immunization of mice with the diphtheria toxin and with a conjugate of the diphtheria toxin with polystyrene microspheres demonstrated that the conjugate immunization resulted in the formation of hybridoma clones which produced antibodies only to the epitopes of the A fragment of the diphtheria toxin. The immunization with the native toxin caused the production of hybridoma clones which predominantly produced antibodies to the epitopes of the B fragment.  相似文献   

6.
The role of specific receptors in the translocation of diphtheria toxin A fragment to the cytosol and for the insertion of the B fragment into the cell membrane was studied. To induce nonspecific binding to cells, toxin was either added at low pH, or biotinylated toxin was added at neutral pH to cells that had been treated with avidin. In both cases large amounts of diphtheria toxin became associated with the cells, but there was no increase in the toxic effect. There was also no increase in the amount of A fragment that was translocated to the cytosol, as estimated from protection against externally added Pronase E. In cells where specific binding was abolished by treatment with 12-O-tetradecanoyl-phorbol 13-acetate, trypsin, or 4,4'-diisothiocyano-2,2'-stilbenedisulfonic acid, unspecific binding did not induce intoxication or protection against protease. This was also the case in untreated L cells, which showed no specific binding of the toxin. When Vero cells with diphtheria toxin bound to specific receptors were exposed to low pH, the cells were permeabilized to K+, whereas this was not the case when the toxin was bound nonspecifically at low pH or via avidin-biotin. The data indicate that the cell-surface receptor for diphtheria toxin facilitates both insertion of the B fragment into the cell membrane and translocation of the A fragment to the cytosol.  相似文献   

7.
Treatment with phospholipase C strongly protected monkey kidney (Vero) cells against diphtheria toxin and reduced the ability of the cells to bind 125I-labelled toxin. Treatment with phospholipase D and with trypsin also protected the cells, although to a lesser extent. Phospholipase A2 had no protective effect. Phospholipase C also protected fetal hamster kidney cells against the toxin. After removal of the enzymes, as well as after treatment of the cells with 4-acetamide 4'-isothiocyanostilbene 2,2'-disulfonic acid, diphtheria toxin binding capability was restored slowly, apparently by a process requiring protein synthesis, since cycloheximide blocked the restoration. The data indicate that both phospholipids and protein are involved in the binding sites for diphtheria toxin.  相似文献   

8.
The pH-dependent conformational change of diphtheria toxin   总被引:8,自引:0,他引:8  
Labeling by a hydrophobic photoactivatable reagent and limited proteolysis have been used to study conformational changes of diphtheria toxin related to its pH-dependent membrane insertion and translocation. TID (3-(trifluoromethyl)-3-(m-[125I]iodophenyl)diazirine) labels diphtheria toxin at pH 5 much more efficiently than at pH 7, both in the presence and absence of lipid vesicles. In the absence of membranes, the extent of labeling is greater and the pH dependence is stronger. As analyzed on sodium dodecyl sulfate-polyacrylamide gels and by high pressure liquid chromatography, both the A- and B-subunits and most of the cyanogen bromide fragments of the toxin are labeled by TID at acid pH. The products of trypsin cleavage of diphtheria toxin at pH 5 are different from those seen at neutral pH. Trypsin-susceptible sites were identified by gel electrophoresis of the trypsin fragments, combined with electrophoresis and high pressure liquid chromatography of CNBr digests of trypsin-treated toxin. At neutral pH, the main sites of digestion are at the junction between the A- and B-fragments and near the NH2 terminus of the A-fragment. At pH 5.2, these sites are less efficiently cut, and new sites appear near the NH2 terminus of the B-fragment, in an amphipathic portion of the sequence. Thus, even in the absence of membranes, acid pH induces a significant conformational change in diphtheria toxin. This change involves burial of some previously accessible sites, exposure of previously inaccessible sites, and the formation of hydrophobic regions over an extensive portion of the polypeptide chain.  相似文献   

9.
Chimeric proteins composed of ricin toxin A chain (RTA) and staphylococcal protein A (PA) have been produced in E. coli. Constructs consisting of N-terminal RTA and C-terminal PA (RTA-PA) or N-terminal PA and C-terminal (PA-RTA) were capable of binding to immunoglobulin G (via PA) and of specifically depurinating 28 S ribosomal RNA (via RTA). However, neither fusion protein was cytotoxic to antigen-bearing target cells in the presence of an appropriate monoclonal antibody presumably because the RTA could not be released from the PA within the cytosol where the ribosomal substrate of RTA is located. The overcome this, a short amino acid sequence from diphtheria toxin was engineered between the RTA and PA to produce a disulfide-linked loop containing a trypsin sensitive cleavage site. Cleavage of this fusion protein with trypsin converted the RTA-DT-PA to the two chain form consisting of RTA linked by a disulfide bond to PA. The cleaved fusion protein was highly toxic to Daudi cells coated with anti-immunoglobulin antibody suggesting that the RTA could be released from the PA by reduction within the cytosol.  相似文献   

10.
Active diphtheria toxin consists of two disulphide-linked fragments, termed A and B. Fragment B, which contains an internal disulphide bridge, facilitates translocation of the enzymatically active fragment A to the cytosol of eukaryotic cells. In this process cation-selective channels are formed. An in vitro translated full-length mutant lacking the internal disulphide bridge (A-58**) was functionally indistinguishable from its disulphide-containing counterpart (A-58) with respect to trypsin sensitivity, receptor binding, A-fragment translocation, and channel formation. In contrast, the B fragment of A-58** (B-36**) was slightly less trypsin resistant than the S-S-containing B fragment, B-36, and was approximately 300-fold less efficient than B-36 in permeabilizing cells. When first dialysed and then reconstituted with A fragment, B fragment without disulphide bridge yielded a less-active toxin than did wild-type B fragment. We conclude that the disulphide bridge in fragment B is not necessary for toxicity, as earlier believed, and that channel formation may play a role in membrane translocation.  相似文献   

11.
CRM197, CRM176, and CRM228 are products of single or multiple missense mutations in the diphtheria toxin gene. CRM197 differs from wild-type toxin in 1 amino acid residue of the fragment A region, and also CRM176 and CRM228 have amino acid substitution(s) in fragment A. We compared the binding properties of CRM197 to toxin-sensitive Vero cells with those of diphtheria toxin and other CRMs. Nicked CRM197 is about 50 times more effective than intact CRM197 in inhibiting the action of diphtheria toxin on sensitive cells, as shown by inhibition of diphtheria toxin cytotoxicity or inhibition of binding of 125I-diphtheria toxin. The binding of native toxin or other CRMs was not significantly affected by nicking. Moreover, the binding of CRM197 to cells was unaffected by ATP, although ATP clearly inhibits binding of diphtheria toxin, CRM176, and CRM228. Two kinds of hybrid protein were formed using fragment B of CRM197: one with fragment A of diphtheria toxin and one with fragment A of CRM228. ATP inhibited the binding of these hybrid proteins. Furthermore, the affinities of these hybrid proteins for diphtheria toxin-sensitive cells were the same as that of native toxin. Thus, it was concluded that the altered binding properties of CRM197 were due to alteration of fragment A and what the interaction of diphtheria toxin with ATP involves both fragments. The results also suggest that fragment A plays a role in diphtheria toxin-receptor interaction.  相似文献   

12.
Homogeneous fragment B, obtained through nicking of diphtheria toxin with insoluble trypsin, was cleaved with cyanogen bromide in 70% formic acid. After citraconylation, the cleavage products were separated by gel filtration on Sephadex G--75 and purified by gel filtration, ion-exchange and thin-layer or paper chromatography. Six CNBr peptides were characterized, the composition of which account for the total amino acid content of fragment B. Their apparent molecular weights are: CB 1, 12 000; CB 2, 14 000; CB 3, 8000; CB 4a, 2400; CB 4b, 2200; CB 5, 2200. CB 4a is the NH2--terminal peptide; it contains the cysteine residue of the disulfide bridge linking fragment B to fragment A. CB 3 is the COOH--terminal peptide; it bears the disulfide bridge of fragment B. Characterization of two CNBr--derived overlapping peptides provided the positioning of CB 4b and CB 2 and allowed an alignment of the CNBr peptides of fragment B to be proposed.  相似文献   

13.
To be toxic for mammalian cells, Pseudomonas exotoxin (PE) requires proteolytic cleavage between Arg-279 and Gly-280. Cleavage, which is mediated by the cellular protease furin, generates an active C-terminal fragment which translocates to the cytosol and inhibits protein synthesis. In vitro , furin-mediated cleavage is optimal at pH 5.5 with a relatively slow turnover rate. Within cells, only 5–10% of cell-associated PE is cleaved. To investigate the reasons for this inefficient cleavage, the amino acid composition near the cleavage site was altered to resemble more closely the arginine-rich sequence from the functionally similar region of diphtheria toxin (DT). Four PE-DT mutants were generated, whereby 1, 5, 6 or 8 amino acids at the PE-cleavage site were changed to amino acids found at the DT-cleavage site. Mutant proteins were expressed in Escherichia coli , purified and then analysed for their susceptibility to cleavage by furin and trypsin, susceptibility to cell-mediated cleavage, and cytotoxic activity relative to wild-type PE. At pH 5.5, the rate of both furin-mediated cleavage and trypsin-mediated cleavage increased dramatically when amino acids in PE were altered to resemble the DT sequence. This increase did not alter the pH optimum for furin-mediated cleavage of PE toxins, which remained at pH 5.0–5.5. When radioactive versions of selected PE-DT proteins were added to intact cells, an increase in the percentage of molecules that were cleaved relative to wild-type PE was also seen. However, changes that favoured increased proteolysis apparently interfered with other important toxin functions because none of the PE-DT proteins exhibited enhanced toxicity for cells when compared with the activity of wild-type PE.  相似文献   

14.
Two restriction fragments from corynebacteriophage beta vir tox+ that encode peptides similar to diphtheria toxin fragment A and the chain termination fragment, CRM45, have been cloned into Escherichia coli in plasmid pBR322. Clones containing the recombinant plasmids produced gene products that were active in catalyzing the ADP ribosylation of elongation factor 2 and were reactive with diphtheria toxin antiserum. Toxin-related peptides were found primarily in the periplasmic compartment and were degraded to nonimmunoreactive forms within 1 to 2 h of synthesis. The expression of both gene fragments appears to have originated from the diphtheria toxin promoter.  相似文献   

15.
A hybrid protein of ricin and the enzymatically active fragment A of diphtheria toxin (toxin A) has been synthesized and purified. The diphtheria toxin A fragment of the hybrid protein is shown to enter the cytosol compartment of HeLa cells, its presence assayed by the fall of intracellular elongation factor II (EF-2) and the rise of ADP-ribosylated EF-2. Hybrid entrance to HeLa cells is blocked by lactose which blocks receptor-mediated entry of ricin but not by NH4Cl which blocks the transport of diphtheria toxin. It is concluded that the diphtheria toxin fragment A moiety of the hybrid enters the cell cytosol via the ricin receptor-mediated transport system. The kinetics of intracellular ADP-ribosylation of EF-2 by diphtheria toxin have also been studied. Ribosylation is preceded by a toxin dose-dependent lag period. The data suggest that the time constant responsible for the lag period is in the transport step. Models consistent with these data are discussed.  相似文献   

16.
Diphtheria toxin and fragment B bind to hydrocarbon-coated agaroses. Fragment A of the toxin is not adsorbed to such resins. Using Seph-C4, the toxin and fragment B can be eluted from the column after adsorption by increasing the ionic strength of the eluent. The toxin is also eluted from the Seph-C6 column, but fragment B is eluted only in the denatured form. Purification of the toxin can be achieved simply by passing the growth medium supernatant through a small size Seph-C6 column and eluting the toxin by 0.1 m NaCl. The fragments of diphtheria toxin obtained after mild trypsin treatment can be separated purely on a Seph-C4 column. The hydrophobic chromatography system may thus serve as a tool for purification of the toxin and its fragments: it may also be useful in large-scale preparations.  相似文献   

17.
Ehrlich ascites tumor cells were found to be very insensitive to diphtheria toxin. We formed 37 hybrids from Ehrlich tumor cells and diphtheria toxin-sensitive human fibroblasts. The effects of diphtheria toxin on protein synthesis in those hybrids were examined. The hybrids were divided into three groups on the basis of toxin sensitivity. Group A hybrids were as sensitive to diphtheria toxin as human fibroblasts, Group C were as resistant as Ehrlich tumor cells, and Group B had intermediate sensitivity. Group A hybrids had diphtheria toxin-binding sites but Group B and C had no detectable binding sites. Elongation factor-2 of all the hybrids was susceptible to ADP-ribosylation by fragment A of diphtheria toxin. Cells of Group A and B became more sensitive to CRM 45 (cross-reacting material 45 of diphtheria toxin) after they were exposed to low pH (pH = 4.5). The resistance of Group C to CRM 45 was not affected by the same treatment. Group A and B hybrids and human fibroblasts had similar sensitivities to a hybrid toxin composed of wheat germ agglutinin and fragment A of diphtheria toxin, but Group C and Ehrlich tumor cells were resistant to this hybrid toxin. All the hybrids and Ehrlich tumor cells were more sensitive to a hybrid toxin composed of wheat germ agglutinin and subunit A of ricin than were human fibroblasts. On subcloning of Group B hybrids, one Group C hybrid was obtained, but no Group A hybrid. These facts suggest that Ehrlich ascites tumor cells differ from human fibroblasts in the expression of a factor(s) that is involved in entry of fragment A of diphtheria toxin into the cytoplasm after the toxin binds to its surface receptors.  相似文献   

18.
Anthrax toxin is the only protein secreted by Bacillus anthracis that contributes to the virulence of this bacterium. An obligatory step in the action of anthrax toxin on eukaryotic cells is cleavage of the receptor-bound protective antigen (PA) protein (83 kilodaltons) to produce a 63-kilodalton, receptor-bound COOH-terminal fragment. A similar fragment can be obtained by limited treatment with trypsin. This proteolytic processing event exposes a site with high affinity for the other two anthrax toxin proteins, lethal factor and edema factor. Terminal sequencing of the purified fragment showed that the activating cleavage occurred in the sequence Arg164-Lys165-Lys166-Arg167. The gene encoding PA was mutagenized to delete residues 163-168, and the deleted PA was purified from a Bacillus subtilis host. The deleted PA was not cleaved by either trypsin or the cell-surface protease, and was non-toxic when administered with lethal factor. Purified, deleted PA protected rats when administered 90 min before injection of 20 minimum lethal doses of toxin. This mutant PA may be useful as a replacement for the PA that is the major active ingredient in the current human anthrax vaccine, because deleted PA is expected to have normal immunogenicity, but would not combine with trace amounts of LF and EF to cause toxicity.  相似文献   

19.
Activation of Cry protoxins is carried out by midgut proteases. This process is important for toxicity and in some cases for specificity. Commercial proteases have been used for in vitro protoxin activation. In the case of Cry1A protoxins, trypsin digestion generates a toxic fragment of 60–65 kDa. Here, we have analyzed the in vitro and in vivo activation of Cry1Ab. We found differences in the processing of Cry1Ab protoxin by Manduca sexta and Spodoptera frugiperda midgut proteases as compared to trypsin. Midgut juice proteases produced two additional nicks at the N-terminal end removing helices 1 and 2a to produce a 58 kDa protein. A further cleavage within domain II splits the toxin into two fragments of 30 kDa. The resulting fragments were not separated, but instead coeluted with the 58 kDa monomer, in size-exclusion chromatography. To examine if this processing was involved in the activation or degradation of Cry1Ab toxin, binding, pore formation, and toxicity assays were performed. Pore formation assays showed that midgut juice treatment produced a more active toxin than trypsin treatment. In addition, it was determined that the 1 helix is dispensable for Cry1Ab activity. In contrast, the appearance of the 30 kDa fragments correlates with a decrease in pore formation and insecticidal activities. Our results suggest that the cleavage in domain II may be involved in toxin inactivation, and that the 30 kDa fragments are stable intermediates in the degradation pathway.  相似文献   

20.
The native crystal delta-endotoxin produced by Bacillus thuringiensis var. colmeri, serotype 21, is toxic to both lepidopteran (Pieris brassicae) and dipteran (Aedes aegypti) larvae. Solubilization of the crystal delta-endotoxin in alkaline reducing conditions and activation with trypsin and gut extracts from susceptible insects yielded a preparation whose toxicity could be assayed in vitro against a range of insect cell lines. After activation with Aedes aegypti gut extract the preparation was toxic to all of the mosquito cell lines but only one lepidopteran line (Spodoptera frugiperda), whereas an activated preparation produced by treatment with P. brassicae gut enzymes or trypsin was toxic only to lepidopteran cell lines. These in vitro results were paralleled by the results of in vivo bioassays. Gel electrophoretic analysis of the products of these different activation regimes suggested that a 130-kDa protoxin in the native crystal is converted to a 55-kDa lepidopteran-specific toxin by trypsin or P. brassicae enzymes and to a 52-kDa dipteran toxin by A. aegypti enzymes. Two-step activation of the 130-kDa protoxin by successive treatment with trypsin and A. aegypti enzymes further suggested that the 52-kDa dipteran toxin is derived from the 55-kDa lepidopteran toxin by enzymes specific to the mosquito gut. Confirmation of this suggestion was obtained by peptide mapping of these two polypeptides. The native crystal 130 kDa delta-endotoxin and the two insect-specific toxins all cross-reacted with antiserum to B. thuringiensis var. kurstaki P1 lepidopteran toxin. Preincubation of the two activated colmeri toxins with P1 antiserum neutralized their cytotoxicity to both lepidopteran and dipteran cell lines.  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