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.
Helix α4 of Bacillus thuringiensis Cry toxins is thought to play a critical role in the toxins'' mode of action. Accordingly, single-site substitutions of many Cry1Aa helix α4 amino acid residues have previously been shown to cause substantial reductions in the protein''s pore-forming activity. Changes in protein structure and formation of intermolecular disulfide bonds were investigated as possible factors responsible for the inactivity of these mutants. Incubation of each mutant with trypsin and chymotrypsin for 12 h did not reveal overt structural differences with Cry1Aa, although circular dichroism was slightly decreased in the 190- to 210-nm region for the I132C, S139C, and V150C mutants. The addition of dithiothreitol stimulated pore formation by the E128C, I132C, S139C, T142C, I145C, P146C, and V150C mutants. However, in the presence of these mutants, the membrane permeability never reached that measured for Cry1Aa, indicating that the formation of disulfide bridges could only partially explain their loss of activity. The ability of a number of inactive mutants to compete with wild-type Cry1Aa for pore formation in brush border membrane vesicles isolated from Manduca sexta was also investigated with an osmotic swelling assay. With the exception of the L147C mutant, all mutants tested could inhibit the formation of pores by Cry1Aa, indicating that they retained receptor binding ability. These results strongly suggest that helix α4 is involved mainly in the postbinding steps of pore formation.During the last few decades, the insecticidal toxins produced by Bacillus thuringiensis have been used increasingly in the forms of formulated sprays and transgenic plants for the highly focused biological control of insect pests (29). At the same time, the mechanism by which these proteins form pores in the apical membrane of midgut epithelial cells of targeted insects has been studied extensively (7, 29). In the case of the three-domain Cry toxins, specificity is mostly attributable to their capacity to bind to certain proteins located on the surface of the intestinal membrane through specific segments of domains II and III, composed mainly of β sheets (16, 27). On the other hand, membrane insertion and pore formation are thought to occur through elements of domain I, composed of a bundle of six amphipathic α-helices surrounding the highly hydrophobic helix α5 (17, 20).Several lines of evidence indicate that helices α4 and α5 play a particularly important role in these processes (3). Spectroscopic studies with synthetic peptides corresponding to domain I helices revealed that α4 and α5 have the greatest propensity for insertion into artificial membranes, although insertion and pore formation were most efficient when α4 and α5 were connected by a segment corresponding to the α4-α5 loop of the toxin (13, 14). A particularly large number of single-site mutations with altered amino acids from these helices, which lead to a strong reduction in the toxicity and pore-forming ability of the toxin, have been characterized (2, 9, 10, 15, 18, 23, 25, 30, 31, 33). Finally, a site-directed chemical modification study has provided strong evidence that α4 lines the lumens of the pores formed by the toxin (23).Recent studies have established that toxin activity is especially sensitive to modifications not only in the charged residues of α4 (31) but in most of its hydrophilic residues (15). Furthermore, the loss of activity of most of these mutants did not result from an altered selectivity or size of the pores but from a reduced pore-forming capacity of the toxin (15, 31). In order to better understand the role of α4 in the mechanism of pore formation, the present study was carried out with a series of previously characterized Cry1Aa mutants in which most of the residues from this helix were replaced by cysteines (15). By subjecting these mutants to circular dichroism (CD), protease sensitivity, pore formation inhibition, and electrophoretic mobility analyses, our data suggest that the mutations in α4 which alter the pore-forming ability of Cry1Aa do so mainly by preventing the proper oligomerization or membrane insertion of the toxin.  相似文献   

2.
Pore formation in the apical membrane of the midgut epithelial cells of susceptible insects constitutes a key step in the mode of action of Bacillus thuringiensis insecticidal toxins. In order to study the mechanism of toxin insertion into the membrane, at least one residue in each of the pore-forming-domain (domain I) interhelical loops of Cry1Aa was replaced individually by cysteine, an amino acid which is normally absent from the activated Cry1Aa toxin, using site-directed mutagenesis. The toxicity of most mutants to Manduca sexta neonate larvae was comparable to that of Cry1Aa. The ability of each of the activated mutant toxins to permeabilize M. sexta midgut brush border membrane vesicles was examined with an osmotic swelling assay. Following a 1-h preincubation, all mutants except the V150C mutant were able to form pores at pH 7.5, although the W182C mutant had a weaker activity than the other toxins. Increasing the pH to 10.5, a procedure which introduces a negative charge on the thiol group of the cysteine residues, caused a significant reduction in the pore-forming abilities of most mutants without affecting those of Cry1Aa or the I88C, T122C, Y153C, or S252C mutant. The rate of pore formation was significantly lower for the F50C, Q151C, Y153C, W182C, and S252C mutants than for Cry1Aa at pH 7.5. At the higher pH, all mutants formed pores significantly more slowly than Cry1Aa, except the I88C mutant, which formed pores significantly faster, and the T122C mutant. These results indicate that domain I interhelical loop residues play an important role in the conformational changes leading to toxin insertion and pore formation.Once ingested by susceptible insect larvae, the insecticidal crystal proteins of Bacillus thuringiensis are solubilized and converted to their toxic form by midgut proteases. The activated toxins bind to specific receptors on the surface of the luminal membrane of midgut columnar cells, insert into the membrane, and form pores that abolish transmembrane ionic gradients and osmotic balance, leading to the disruption of the epithelium and death of the insect (47, 51). Members of the B. thuringiensis Cry toxin family for which the atomic structure has been reported share a similar three-domain organization in which domain I is composed of a bundle of six amphipathic α-helices surrounding a hydrophobic helix (α5), and domains II and III are formed mostly of β-sheets (7, 8, 18, 26, 37, 38, 43). While domains II and III are thought to be involved in receptor binding and toxin specificity (47), domain I is believed to play a major role in membrane insertion and pore formation (51). Toxin fragments corresponding to domain I of Cry1Ac (62), Cry3Aa (53), and Cry3Ba (61) or to the first five α-helices of Cry4B (48) have been shown to form pores in model membranes. Pore formation in artificial membranes has also been demonstrated with synthetic peptides corresponding to α5 of Cry1Ac (13) and Cry3Aa (19, 21) and to the α4-loop-α5 segment of Cry3Aa (23). Spectroscopic studies have also revealed that while synthetic peptides corresponding to α4 and α5 can coassemble within a lipid bilayer, those corresponding to α2, α3, α6, and α7 adopt a membrane surface orientation (20, 22). In agreement with these findings, α4 was shown to line the lumen of the pores (42). On the other hand, convincing evidence supporting previous suggestions that most of the toxin molecule may become imbedded in the membrane (3, 39, 60) has recently been reported (44, 45).Thus, several models have been proposed for the mechanism of toxin insertion and pore formation (4, 9, 28, 32, 39, 44, 52, 56). Although these models differ in the identities of the toxin segments that are suggested to insert into the membrane, they all imply that the toxin undergoes conformational changes following binding to the membrane surface. Even though such changes imply rotations about the polypeptide backbone in domain I interhelical loops, little attention has been devoted so far to the role of domain I loop residues in pore formation.In the present study, amino acid residues strategically located within each of these loops in Cry1Aa were replaced by a cysteine using site-directed mutagenesis. The resulting mutant toxins were assayed with Manduca sexta midgut brush border membrane vesicles using a light-scattering technique. Mutations mapping within several of these loops altered the functional properties of Cry1Aa, suggesting the involvement of most domain I α-helices in the pore-forming process.  相似文献   

3.
The insecticidal Cry toxins produced by the bacterium Bacillus thuringiensis are comprised of three structural domains. Domain I, a seven-helix bundle, is thought to penetrate the insect epithelial cell plasma membrane through a hairpin composed of α-helices 4 and 5, followed by the oligomerization of four hairpin monomers. The α-helix 4 has been proposed to line the lumen of the pore, whereas some residues in α-helix 5 have been shown to be responsible for oligomerization. Mutation of the Cry1Ac1 α-helix 4 amino acid Asn135 to Gln resulted in the loss of toxicity to Manduca sexta, yet binding was still observed. In this study, the equivalent mutation was made in the Cry1Ab5 toxin, and the properties of both wild-type and mutant toxin counterparts were analyzed. Both mutants appeared to bind to M. sexta membrane vesicles, but they were not able to form pores. The ability of both N135Q mutants to oligomerize was also disrupted, providing the first evidence that a residue in α-helix 4 can contribute to toxin oligomerization.  相似文献   

4.
Disulfide bridges were introduced into Cry1Aa, a Bacillus thuringiensis lepidopteran toxin, to stabilize different protein domains including domain I α-helical regions thought to be involved in membrane integration and permeation. Bridged mutants could not form functional ion channels in lipid bilayers in the oxidized state, but upon reduction with β-mercaptoethanol, regained parental toxin channel activity. Our results show that unfolding of the protein around a hinge region linking domain I and II is a necessary step for pore formation. They also suggest that membrane insertion of the hydrophobic hairpin made of α-helices 4 and 5 in domain I plays a critical role in the formation of a functional pore.  相似文献   

5.
The insecticidal Cry11Aa and Cyt1Aa proteins are produced by Bacillus thuringiensis as crystal inclusions. They work synergistically inducing high toxicity against mosquito larvae. It was proposed that these crystal inclusions are rapidly solubilized and activated in the gut lumen, followed by pore formation in midgut cells killing the larvae. In addition, Cyt1Aa functions as a Cry11Aa binding receptor, inducing Cry11Aa oligomerization and membrane insertion. Here, we used fluorescent labeled crystals, protoxins or activated toxins for in vivo localization at nano-scale resolution. We show that after larvae were fed solubilized proteins, these proteins were not accumulated inside the gut and larvae were not killed. In contrast, if larvae were fed soluble non-toxic mutant proteins, these proteins were found inside the gut bound to gut-microvilli. Only feeding with crystal inclusions resulted in high larval mortality, suggesting that they have a role for an optimal intoxication process. At the macroscopic level, Cry11Aa completely degraded the gastric caeca structure and, in the presence of Cyt1Aa, this effect was observed at lower toxin-concentrations and at shorter periods. The labeled Cry11Aa crystal protein, after midgut processing, binds to the gastric caeca and posterior midgut regions, and also to anterior and medium regions where it is internalized in ordered “net like” structures, leading finally to cell break down. During synergism both Cry11Aa and Cyt1Aa toxins showed a dynamic layered array at the surface of apical microvilli, where Cry11Aa is localized in the lower layer closer to the cell cytoplasm, and Cyt1Aa is layered over Cry11Aa. This array depends on the pore formation activity of Cry11Aa, since the non-toxic mutant Cry11Aa-E97A, which is unable to oligomerize, inverted this array. Internalization of Cry11Aa was also observed during synergism. These data indicate that the mechanism of action of Cry11Aa is more complex than previously anticipated, and may involve additional steps besides pore-formation activity.  相似文献   

6.
Bacillus thuringiensis Cry1AbMod toxins are engineered versions of Cry1Ab that lack the amino-terminal end, including domain I helix α-1 and part of helix α-2. This deletion improves oligomerization of these toxins in solution in the absence of cadherin receptor and counters resistance to Cry1A toxins in different lepidopteran insects, suggesting that oligomerization plays a major role in their toxicity. However, Cry1AbMod toxins are toxic to Escherichia coli cells, since the cry1A promoter that drives its expression in B. thuringiensis has readthrough expression activity in E. coli, making difficult the construction of these CryMod toxins. In this work, we show that Cry1AbMod and Cry1AcMod toxins can be cloned efficiently under regulation of the cry3A promoter region to drive its expression in B. thuringiensis without expression in E. coli cells. However, p3A-Cry1Ab(c)Mod construction promotes the formation of Cry1AMod crystals in B. thuringiensis cells that were not soluble at pH 10.5 and showed no toxicity to Plutella xylostella larvae. Cysteine residues in the protoxin carboxyl-terminal end of Cry1A toxins have been shown to be involved in disulfide bond formation, which is important for crystallization. Six individual cysteine substitutions for serine residues were constructed in the carboxyl-terminal protoxin end of the p3A-Cry1AbMod construct and one in the carboxyl-terminal protoxin end of p3A-Cry1AcMod. Interestingly, p3A-Cry1AbMod C654S and C729S and p3A-Cry1AcMod C730S recover crystal solubility at pH 10.5 and toxicity to P. xylostella. These results show that combining the cry3A promoter expression system with single cysteine mutations is a useful system for efficient expression of Cry1AMod toxins in B. thuringiensis.  相似文献   

7.
Cry11Aa and Cyt1Aa of Bacillus thuringiensis are active against mosquitoes and show synergism. Cyt1Aa functions as a membrane receptor inducing Cry11Aa oligomerization. Here we characterized Cry11Aa helix α-3 mutants impaired in oligomerization and toxicity against Aedes aegypti, indicating that oligomerization of Cry11Aa is important for toxin action. Cyt1Aa did not recover the insecticidal activity of Cry11Aa mutants.Bacillus thuringiensis subsp. israelensis has been used worldwide for the control of different mosquitoes that are vectors of several human diseases (10, 11). This bacterium produces different toxins that individually show activity against mosquitoes, i.e., Cry4Aa, Cry4Ba, Cry11Aa, and Cyt1Aa (2). The toxicity of Cry11Aa and Cry4 toxins against Aedes aegypti is greatly increased in the presence of sublethal concentrations of Cyt1Aa (14). Also, Cyt1Aa overcomes the resistance of the Culex quinquefasciatus population to Cry11Aa (12, 13). Cyt1Aa synergizes the toxic activity of Cry11Aa by functioning as a Cry11Aa receptor, facilitating the oligomerization of Cry11Aa and its pore formation activity (7, 8). Oligomerization is a complex event that involves interaction with a toxin receptor and further proteolysis of helix α-1 (3). In the case of the Cry1Ab toxin, helix α-3 of domain I contains coiled-coil structures that are important for oligomerization (4). Some point mutations in helix α-3 do not affect interaction with receptors but severely affected oligomerization, influencing pore formation and toxicity against Manduca sexta larvae (4).Since binding with Cyt1Aa facilitates Cry11Aa oligomerization, we hypothesize that Cry11Aa mutants unable to oligomerize would be affected in synergism with Cyt1Aa and in toxicity. In this report, we analyzed the effect of point mutations in helix α-3 of Cry11Aa on oligomerization, synergism with Cyt1Aa, and toxicity against A. aegypti larvae.Helix α-3 of Cry11Aa potentially forms coiled-coil structures, as determined by the program COILS, which calculates the probability that a sequence will adopt a coiled-coil conformation (6). The coiled-coil structures are characterized by heptads of residues (abcdefg), where positions a and d are occupied mostly by apolar residues and g and e by charged residues. Here we mutagenized some residues located at positions g and a of the predicted coiled-coil (Fig. (Fig.1).1). Substitutions R90E, E97A, Y98E, V104E, and S105E were produced by site-directed mutagenesis (Quick Change; Stratagene, La Jolla, CA) using the pCG6 plasmid (1) as a template and appropriate mutagenic oligonucleotides. Point mutations were verified by automated DNA sequencing at Instituto de Biotecnología-UNAM and transformed into the acrystalliferous B. thuringiensis 407 strain. B. thuringiensis strains were grown in solid nutrient broth sporulation medium supplemented with 10 μg/ml erythromycin (5). Crystal inclusions were purified as described previously (8) and solubilized in 100 mM NaOH for 1 h at 4°C. After solubilization, the Cry11Aa protoxins were dialyzed for 12 h against 50 mM Na2CO3, pH 10.5. The pH was equilibrated at pH 8.6 with equal volumes of 1 M Tris-HCl, pH 8, and protoxins were activated with trypsin (1:50, wt/wt) for 2 h at 25°C. All mutants, with the exception of the V104E mutant, which was not analyzed further, produced crystal inclusions similar to those for the wild-type toxin, composed of a 70-kDa protoxin (Fig. (Fig.2A).2A). After trypsin activation, all mutants produced two polypeptides of 32 and 36 kDa, similarly to the Cry11Aa toxin, suggesting that these mutations did not cause a major structural disturbance (Fig. (Fig.2B).2B). The Cry11Aa and mutant activated toxins were analyzed by circular dichroism spectroscopy (Fig. (Fig.2C).2C). The activated toxins were dialyzed against 10 mM Na2HPO4, 50 mM NaF, pH 9, and then purified by anion-exchange chromatography with HiTrap Q-Sepharose (Pharmacia LKB Biotechnology) in the same buffer, using a linear NaF gradient from 50 to 400 mM. The similarities among the curves indicate that the mutant toxins have a structure similar to that of the wild-type toxin.Open in a separate windowFIG. 1.Schematic representation of the coiled-coil structures of the α-3 helices of Cry1Ab and Cry11Aa toxins. The positions of residues a, b, c, d, e, f, and g of the heptads are presented. The mutated residues in both toxins that affected oligomerization and toxicity are shown in boldface type (reference 4 and this work).Open in a separate windowFIG. 2.SDS-PAGE analysis and circular dichroism spectra of Cry11Aa mutant toxins. (A) The Cry11Aa protoxins were solubilized at pH 10.5 and analyzed by SDS-PAGE (15% acrylamide). (B) SDS-PAGE analysis (15% acrylamide) of the activated toxins with trypsin. Both SDS-polyacrylamide gels were stained with Coomassie blue. Lanes 1, Cry11Aa; lanes 2, E97A mutant; lanes 3, Y98E mutant; lanes 4, R90E mutant; lanes 5, S105E mutant. (C) Analysis of the secondary-structure compositions of the mutants and Cry11Aa activated toxins. Circular dichroism spectra were recorded with a Jasco model J-715 spectropolarimeter equipped with a Peltier temperature control supplied by Jasco. Spectra were collected from 190 to 250 nm. Eight replicate spectra were collected for each sample to improve the signal-to-noise ratios. The final purified-protein concentration was 0.3 mg/ml, and spectra were collected in a 0.1-cm-pathlength cell. The secondary-structure prediction was performed using the CDSSTR algorithm (1a, 11a). Solid black line, Cry11Aa; dotted black line, E97A mutant; dashed black line, Y98E mutant; solid gray line, R90E mutant; dotted gray line, S105E mutant; MRE, mean residue ellipticity; [θ], ellipticity.The toxicity of spore/crystal suspensions of Cry11Aa or the individual mutants (75 to 10,000 ng/ml) was analyzed with bioassays against 10 fourth-instar A. aegypti larvae reared at 28°C, 87% humidity, and 12:12 light-dark conditions in 100 ml dechlorinated water, and mortality was scored after 24 h (four independent assays). The Cry11Aa toxin showed a mean lethal concentration of 355 ng/ml, with 95% confidence limits of 265 to 446 (Probit analysis using Polo-PC LeOra Software). In contrast, the R90E, E97A, Y98E, and S105E mutants were severely affected in toxicity against A. aegypti larvae, since no mortality was observed at the highest concentration used (10,000 ng/ml).We then analyzed the oligomerization of Cry11Aa toxins as previously described (8). Small unilamelar vesicles (SUV), composed of egg yolk phosphatidyl choline, cholesterol (Avanti Polar Lipids, Alabaster, AL), and stearylamine (Sigma, St. Louis, MO) at a 10:3:1 proportion, respectively, were used (8). Cyt1Aa was purified from the 4Q7/pWF45 strain (14) grown as described above. Cyt1Aa inclusions were purified by sucrose gradients, solubilized in 50 mM Na2CO3, 10 mM dithiothreitol, pH 10.5 (2 h at 30°C), and activated with 1:100 proteinase K (Sigma-Aldrich Co.), wt/wt, for 20 min at 30°C.For oligomerization assays, 2.5 μg soluble Cry11Aa or mutant protoxin was incubated for 2 h at 37°C in a 100-μl final volume of 50 mM Na2CO3, pH 10.5, with 200 μM SUV, 1:50 trypsin (wt/wt), and 0.5 μg Cyt1Aa activated toxin. After 2 h of incubation, 1 mM phenylmethylsulfonyl fluoride was added to stop the reaction, and the membrane fraction was separated by centrifugation (1 h at 100,000 × g). The pellet was suspended in the same buffer solution. Oligomeric structures of Cry toxins are highly stable after boiling as well as after urea denaturation (9). The suspension was boiled for 4 min, analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) (8% acrylamide), and electrotransferred onto polyvinylidene difluoride membranes (Millipore, Bedford, MA). The oligomeric and monomeric structures of Cry11Aa were detected using polyclonal anti-Cry11Aa antibody (1/15,000; 1 h) and a secondary antibody coupled with horseradish peroxidase (Sigma, St. Louis, MO) (1/5,000; 1 h) followed by luminol (ECL; Amersham Pharmacia Biotech) as described by the manufacturers. Figure Figure33 shows that only the Cry11Aa wild-type toxin was able to oligomerize, while the mutants were severely impaired in oligomerization.Open in a separate windowFIG. 3.Analysis of Cry11Aa oligomer formation. Soluble Cry11Aa protoxin was activated with trypsin for 2 h at 37°C in the presence of SUV and Cyt1Aa activated toxin. The membrane fraction was separated by ultracentrifugation, and the Cry11Aa protein was analyzed by Western blotting of the membrane pellet with polyclonal anti-Cry11A antibody. The sizes of the proteins were estimated from a molecular prestained plus standard, all blue (Bio-Rad). Lane 1, Cry11Aa; lane 2, R90E mutant; lane 3, Y98E mutant; lane 4, E97A mutant; lane 5, S105E mutant.Finally, the synergistic activity between Cyt1Aa and Cry11Aa was analyzed. A concentration of Cyt1Aa that produced 10% mortality was assayed in the presence of a protein concentration of wild-type Cry11A that produced 20% mortality. Larvae were examined 24 h after treatment, in three repetitions. This particular protein mixture produced a synergism factor of 8. Under these conditions, mortality was more than 80%, due to the synergistic activities of both toxins. Similar experiments were performed with the mutant toxins, using the same concentration of Cyt1Aa toxin and different concentrations (up to 6,000 ng/ml) of the mutant toxins. Cyt1A did not increase the toxicity of the Cry11Aa mutants, since only 10% mortality was observed, even at the highest concentration of the mutant toxins.Previously, helix α-3 of a lepidopteran-specific toxin (Cry1Ab) was subjected to mutagenesis. The R99E and Y107E mutants of the Cry1Ab toxin were severely impaired in oligomerization and toxicity, showing that oligomer formation is a necessary step to kill the larvae (4). The data presented here indicate that oligomer formation is also an essential step in the mechanism of toxicity of the mosquitocidal Cry11Aa toxin and that helix α-3 is involved in this process.  相似文献   

8.
The role played by alpha-helix 4 of the Bacillus thuringiensis toxin Cry1Aa in pore formation was investigated by individually replacing each of its charged residues with either a neutral or an oppositely charged amino acid by using site-directed mutagenesis. The majority of the resulting mutant proteins were considerably less toxic to Manduca sexta larvae than Cry1Aa. Most mutants also had a considerably reduced ability to form pores in midgut brush border membrane vesicles isolated from this insect, with the notable exception of those with alterations at amino acid position 127 (R127N and R127E), located near the N-terminal end of the helix. Introducing a negatively charged amino acid near the C-terminal end of the helix (T142D and T143D), a region normally devoid of charged residues, completely abolished pore formation. For each mutant that retained detectable pore-forming activity, reduced membrane permeability to KCl was accompanied by an approximately equivalent reduction in permeability to N-methyl-D-glucamine hydrochloride, potassium gluconate, sucrose, and raffinose and by a reduced rate of pore formation. These results indicate that the main effect of the mutations was to decrease the toxin's ability to form pores. They provide further evidence that alpha-helix 4 plays a crucial role in the mechanism of pore formation.  相似文献   

9.
Little information is available on the systemic effects of Bacillus thuringiensis toxins in the hemocoel of insects. In order to test whether B. thuringiensis-activated toxins elicit a toxic response in the hemocoel, we measured the effect of intrahemocoelic injections of several Cry1 toxins on the food intake, growth, and survival of Lymantria dispar (Lepidoptera) and Neobellieria bullata (Diptera) larvae. Injection of Cry1C was highly toxic to the Lymantria larvae and resulted in the complete inhibition of food intake, growth arrest, and death in a dose-dependent manner. Cry1Aa and Cry1Ab (5 μg/0.2 g [fresh weight] [g fresh wt]) also affected growth and food intake but were less toxic than Cry1C (0.5 μg/0.2 g fresh wt). Cry1E and Cry1Ac (5 μg/0.2 g fresh wt) had no toxic effect upon injection. Cry1C was also highly toxic to N. bullata larvae upon injection. Injection of 5 μg/0.2 g fresh wt resulted in rapid paralysis, followed by hemocytic melanization and death. Lower concentrations delayed pupariation or gave rise to malformation of the puparium. Finally, Cry1C was toxic to brain cells of Lymantria in vitro. The addition of Cry1C (20 μg/ml) to primary cultures of Lymantria brain cells resulted in rapid lysis of the cultured neurons.  相似文献   

10.
Bacillus thuringiensis Cry toxins exert their toxic effect by specific recognition of larval midgut proteins leading to oligomerization of the toxin, membrane insertion and pore formation. The exposed domain II loop regions of Cry toxins have been shown to be involved in receptor binding. Insect cadherins have shown to be functionally involved in toxin binding facilitating toxin oligomerization. Here, we isolated a VHH (VHHA5) antibody by phage display that binds Cry3Aa loop 1 and competed with the binding of Cry3Aa to Tenebrio molitor brush border membranes. VHHA5 also competed with the binding of Cry3Aa to a cadherin fragment (CR12) that was previously shown to be involved in binding and toxicity of Cry3Aa, indicating that Cry3Aa binds CR12 through domain II loop 1. Moreover, we show that a loop 1 mutant, previously characterized to have increased toxicity to T. molitor, displayed a correlative enhanced binding affinity to T. molitor CR12 and to VHHA5. These results show that Cry3Aa domain II loop 1 is a binding site of CR12 T. molitor cadherin.  相似文献   

11.
During sporulation, Bacillus thuringiensis produces crystalline inclusions comprised of a mixture of δ-endotoxins. Following ingestion by insect larvae, these inclusion proteins are solubilized, and the protoxins are converted to toxins. These bind specifically to receptors on the surfaces of midgut apical cells and are then incorporated into the membrane to form ion channels. The steps required for toxin insertion into the membrane and possible oligomerization to form a channel have been examined. When bound to vesicles from the midguts of Manduca sexta larvae, the Cry1Ac toxin was largely resistant to digestion with protease K. Only about 60 amino acids were removed from the Cry1Ac amino terminus, which included primarily helix α1. Following incubation of the Cry1Ab or Cry1Ac toxins with vesicles, the preparations were solubilized by relatively mild conditions, and the toxin antigens were analyzed by immunoblotting. In both cases, most of the toxin formed a large, antigenic aggregate of ca. 200 kDa. These toxin aggregates did not include the toxin receptor aminopeptidase N, but interactions with other vesicle components were not excluded. No oligomerization occurred when inactive toxins with mutations in amphipathic helices (α5) and known to insert into the membrane were tested. Active toxins with other mutations in this helix did form oligomers. There was one exception; a very active helix α5 mutant toxin bound very well to membranes, but no oligomers were detected. Toxins with mutations in the loop connecting helices α2 and α3, which affected the irreversible binding to vesicles, also did not oligomerize. There was a greater extent of oligomerization of the Cry1Ac toxin with vesicles from the Heliothis virescens midgut than with those from the M. sexta midgut, which correlated with observed differences in toxicity. Tight binding of virtually the entire toxin molecule to the membrane and the subsequent oligomerization are both important steps in toxicity.  相似文献   

12.
The mode of action of Bacillus thuringiensis insecticidal proteins is not well understood. Based on analogies with other bacterial toxins and ion channels, we hypothesized that charged amino acids in helix 4 of the Cry1Aa toxin are critical for toxicity and ion channel function. Using Plutella xylostella as a model target, we analyzed responses to Cry1Aa and eight proteins with altered helix 4 residues. Toxicity was abolished in five charged residue mutants (E129K, R131Q, R131D, D136N, D136C), however, two charged (R127E and R127N) and one polar (N138C) residue mutant retained wild-type toxicity. Compared with Cry1Aa and toxic mutants, nontoxic mutants did not show greatly reduced binding to brush border membrane vesicles, but their ion channel conductance was greatly reduced in planar lipid bilayers. Substituted cysteine accessibility tests showed that in situ restoration of the negative charge of D136C restored conductance to wild-type levels. The results imply that charged amino acids on the Asp-136 side of helix 4 are essential for toxicity and passage of ions through the channel. These results also support a refined version of the umbrella model of membrane integration in which the side of helix 4 containing Asp-136 faces the aqueous lumen of the ion channel.  相似文献   

13.

Background

Bacillus thuringiensis Cry toxins, that are used worldwide in insect control, kill insects by a mechanism that depends on their ability to form oligomeric pores that insert into the insect-midgut cells. These toxins are being used worldwide in transgenic plants or spray to control insect pests in agriculture. However, a major concern has been the possible effects of these insecticidal proteins on non-target organisms mainly in ecosystems adjacent to agricultural fields.

Methodology/Principal Findings

We isolated and characterized 11 non-toxic mutants of Cry1Ab toxin affected in different steps of the mechanism of action namely binding to receptors, oligomerization and pore-formation. These mutant toxins were analyzed for their capacity to block wild type toxin activity, presenting a dominant negative phenotype. The dominant negative phenotype was analyzed at two levels, in vivo by toxicity bioassays against susceptible Manduca sexta larvae and in vitro by pore formation activity in black lipid bilayers. We demonstrate that some mutations located in helix α-4 completely block the wild type toxin activity at sub-stoichiometric level confirming a dominant negative phenotype, thereby functioning as potent antitoxins.

Conclusions/Significance

This is the first reported case of a Cry toxin dominant inhibitor. These data demonstrate that oligomerization is a fundamental step in Cry toxin action and represent a potential mechanism to protect special ecosystems from the possible effect of Cry toxins on non-target organisms.  相似文献   

14.
Because Bombyx mori ABC transporter C2 (BmABCC2) has 1000-fold higher potential than B. mori cadherin-like protein as a receptor for Bacillus thuringiensis Cry1Aa toxin (Tanaka et al., 2013), the gate-opening ability of the latent pore under six extracellular loops (ECLs) of BmABCC2 was expected to be the reason for its higher potential (Heckel, 2012). In this study, cell swelling assays in Sf9 cells showed that BmABCC2 mutants lacking substrate-excreting activity retained receptor activity, indicating that the gate-opening activity of BmABCC2 is not responsible for Cry1Aa toxicity. The analysis of 29 BmABCC2 mutants demonstrated that 770DYWL773 of ECL 4 comprise a putative binding site to Cry1Aa. This suggests that specific toxicity of Cry1Aa toxin to a restricted range of lepidopteran insects is dependent on conservation and variation in the amino acid residues around 770DYWL773 of ECL 4 in the ABCC2.  相似文献   

15.
Development of targeted biological agents against agricultural insect pests is of prime importance for the elaboration and implementation of integrated pest management strategies that are environment-friendly, respectful of bio-diversity and safer to human health through reduced use of chemical pesticides. A major goal to understand how Bt toxins work is to elucidate the functions of their three domains. Domains II and III are involved in binding specificity and structural integrity, but the function of Domain I remains poorly understood. Using aManduca sexta BBMV (brush border membrane vesicles) system, we analyzed its responses to Cry1Aa 15 single-point mutations with altered Domain I helix 4 residues. Light scattering assay showed that toxicity was almost lost in 3 mutants, and we observed significantly reduced toxicity in other 7 mutants. However, 5 mutants retained wild-type toxicity. Using computer software, we simulated the three-dimensional structures of helix 4. Both experimental and bioinformatic analysis showed that residues in Cry1Aa Domain I helix 4 were involved in the formation of ion channels that is critical for its insect toxicity.  相似文献   

16.
Bacillus thuringiensis subsp. israelensis (Bti) produces at least four different crystal proteins that are specifically toxic to different mosquito species and that belong to two non-related family of toxins, Cry and Cyt named Cry4Aa, Cry4Ba, Cry11Aa and Cyt1Aa. Cyt1Aa enhances the activity of Cry4Aa, Cry4Ba or Cry11Aa and overcomes resistance of Culex quinquefasciatus populations resistant to Cry11Aa, Cry4Aa or Cry4Ba. Cyt1Aa synergized Cry11Aa by their specific interaction since single point mutants on both Cyt1Aa and Cry11Aa that affected their binding interaction affected their synergistic insecticidal activity. In this work we show that Cyt1Aa loop β6-αE K198A, E204A and β7 K225A mutants affected binding and synergism with Cry4Ba. In addition, site directed mutagenesis showed that Cry4Ba domain II loop α-8 is involved in binding and in synergism with Cyt1Aa since Cry4Ba SI303-304AA double mutant showed decreased binding and synergism with Cyt1Aa. These data suggest that similarly to the synergism between Cry11Aa and Cyt1Aa toxins, the Cyt1Aa also functions as a receptor for Cry4Ba explaining the mechanism of synergism between these two Bti toxins.  相似文献   

17.
Bacillus thuringiensis insecticidal proteins, Cry toxins, following ingestion by insect larvae, induce insecticidal effect by penetrating the brush border membranes (BBM) of midgut epithelial cells. Purified, activated B. thuringiensis Cry1Aa bound to Bombyx mori BBMV or unbound Cry1Aa were vigorously digested with Pronase. Both digests were compared by Western blotting. Free Cry1Aa was digested to α-helix and/or to amino acids at 1 mg Pronase/mL within 2.4 h at 37 °C. Whereas, BBMV-bound Cry1Aa was very resistant to Pronase digestion and even at 2 mg for 24 h, 7.5 kDa and 30 kDa peptide were detected by α-2,3 antiserum, and α-4,5 and α-6,7 antisera, respectively. Another 30 kDa peptide was also detected by β-6-11 and domain III antisera. These fragments are believed either to be embedded in or to strongly interact with the BBMV. The 7.5 and former 30 kDa peptides are thought to be derived from α-2,3 helix and stretch of α-4 to α-7 helices. Furthermore the latter 30 kDa was thought to include the stretch of β-6 to domain III. Moreover, the embedded Cry1Aa molecule appears to be segregated in some areas of β-1-5 sheets, resulting in the above two 30 kDa peptides. From these digestion patterns, we proposed new membrane insertion model for single Cry1Aa molecule. On the other hand, in digestion of BBMV-bound Cry1Aa, 15 kDa peptide which was recognized only by α-4,5 antiserum was observed. This fragment must be dimeric α-4,5 helices and we discussed the origin of this peptide.  相似文献   

18.
19.
Development of targeted biological agents against agricultural insect pests is of prime importance for the elaboration and implementation of integrated pest management strategies that are environment-friendly, respectful of bio-diversity and safer to human health through reduced use of chemical pesticides. A major goal to understand how Bt toxins work is to elucidate the functions of their three domains. Domains II and III are involved in binding specificity and structural integrity, but the function of Domain I remains poorly understood. Using a Manduca sexta BBMV (brush border membrane vesicles) system, we analyzed its responses to Cry1Aa 15 single-point mutations with altered Domain I helix 4 residues. Light scattering assay showed that toxicity was almost lost in 3 mutants, and we observed significantly reduced toxicity in other 7 mutants. However, 5 mutants retained wild-type toxicity. Using computer software, we simulated the three-dimensional structures of helix 4. Both experimental and bioinform  相似文献   

20.
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