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.
The Cry48Aa/Cry49Aa mosquitocidal two-component toxin was recently characterized from Bacillus sphaericus strain IAB59 and is uniquely composed of a three-domain Cry protein toxin (Cry48Aa) and a binary (Bin) toxin-like protein (Cry49Aa). Its mode of action has not been elucidated, but a remarkable feature of this protein is the high toxicity against species from the Culex complex, besides its capacity to overcome Culex resistance to the Bin toxin, the major insecticidal factor in B. sphaericus-based larvicides. The goal of this work was to investigate the ultrastructural effects of Cry48Aa/Cry49Aa on midgut cells of Bin-toxin-susceptible and -resistant Culex quinquefasciatus larvae. The major cytopathological effects observed after Cry48Aa/Cry49Aa treatment were intense mitochondrial vacuolation, breakdown of endoplasmic reticulum, production of cytoplasmic vacuoles, and microvillus disruption. These effects were similar in Bin-toxin-susceptible and -resistant larvae and demonstrated that Cry48Aa/Cry49Aa toxin interacts with and displays toxic effects on cells lacking receptors for the Bin toxin, while B. sphaericus IAB59-resistant larvae did not show mortality after treatment with Cry48Aa/Cry49Aa toxin. The cytopathological alterations in Bin-toxin-resistant larvae provoked by Cry48Aa/Cry49Aa treatment were similar to those observed when larvae were exposed to a synergistic mixture of Bin/Cry11Aa toxins. Such effects seemed to result from a combined action of Cry-like and Bin-like toxins. The complex effects caused by Cry48Aa/Cry49Aa provide evidence for the potential of these toxins as active ingredients of a new generation of biolarvicides that conjugate insecticidal factors with distinct sites of action, in order to manage mosquito resistance.Bacillus sphaericus is considered an important entomopathogen due to its capacity to produce insecticidal proteins with specific action against mosquitoes (Diptera: Culicidae). The binary (Bin) toxin, which is produced during bacterial sporulation and deposited in parasporal crystalline inclusions, is the most important larvicidal factor. Other proteins characterized, such as mosquitocidal toxins (Mtx proteins), can be produced during vegetative growth, and although these proteins may have larvicidal potential, they play a minor role in the toxicity of the native strains since they are produced by vegetative cells and are degraded by B. sphaericus proteinases (20, 30), and do not form components of the spore-crystal preparations that are used in control programs. Recently, a new two-component toxin was characterized from B. sphaericus strain IAB59. This is formed by the proteins Cry48Aa (135 kDa) and Cry49Aa (53 kDa), which are produced as crystalline inclusions (13). The toxin has a unique composition since the Cry48Aa component belongs to the three-domain family of Cry proteins with 30% similarity to the mosquitocidal Cry4Aa protein from Bacillus thuringiensis serovar israelensis, while Cry49Aa is one of the Bin-toxin-like proteins, a family that comprises the Bin toxin from B. sphaericus, in addition to the Cry36 and Cry35 proteins from B. thuringiensis (9, 13).Cry48Aa/Cry49Aa is considered a two-component toxin because neither component shows toxicity alone, whereas both can act in synergy and the optimum level of toxicity to Culex species is achieved when the two are present at an equimolar ratio. The 50% lethal concentration for third-instar larvae equates to 15.9 ng/ml Cry48Aa and 6.3 ng/ml Cry49Aa of purified toxins, which is a level of toxicity comparable to that of the Bin toxin (13). However, in contrast to the Bin toxin, which is naturally produced in an equimolar ratio, Cry48Aa production is low in native strains and does not confer high toxicity (13). The initial steps of the mode of action of Bin and Cry48Aa/Cry49Aa crystals are similar and comprise the ingestion of crystals, solubilization under alkaline pH, and activation of protoxins into toxins by midgut proteases. After processing, Bin toxin recognizes and binds to specific receptors in the midgut of Bin-toxin-susceptible species through its subunit BinB (51 kDa), while the component BinA (42 kDa) confers toxicity and is likely to form pores in the cell membrane (7, 25). The membrane-bound receptors of Bin toxin on the midgut of Culex quinquefasciatus larvae, Cqm1, were characterized as 60-kDa α-glucosidases (24). The mode of action of Cry48Aa/Cry49Aa is still unknown, but a remarkable feature of this new two-component toxin is the capacity to overcome C. quinquefasciatus resistance to the Bin toxin (13, 19, 21). Resistance of Culex larvae to the Bin-toxin-based larvicides often relies on the absence of functional Cqm1 receptors in the midgut (19, 24, 26). As a consequence, toxins with a distinct mode of action, such as Cry48Aa/Cry49Aa as well as B. thuringiensis serovar israelensis toxins (Cry11Aa, Cry4Aa, Cry4Ba, and Cyt1Aa), do not experience cross-resistance in the Bin-toxin-resistant larvae (12, 21, 32). Such toxins can play a strategic role in the management of resistance, and the major goal of this study was to investigate the ultrastructural effects of the Cry48Aa/Cry49Aa toxin on Bin-toxin-susceptible and -resistant C. quinquefasciatus larvae and to compare these with the effects of a synergistic mixture of Bin/Cry11Aa used to overcome Bin toxin resistance.  相似文献   

2.
Cry4Aa produced by Bacillus thuringiensis is a dipteran-specific toxin and is of great interest for developing a bioinsecticide to control mosquitoes. Therefore, it is very important to characterize the functional motif of Cry4Aa that is responsible for its mosquitocidal activity. In this study, to characterize a potential receptor binding site, namely, loops 1, 2, and 3 in domain II, we constructed a series of Cry4Aa mutants in which a residue in these three loops was replaced with alanine. A bioassay using Culex pipiens larvae revealed that replacement of some residues affected the mosquitocidal activity of Cry4Aa, but the effect was limited. This finding was partially inconsistent with previous results which suggested that replacement of the Cry4Aa loop 2 results in a significant loss of mosquitocidal activity. Therefore, we constructed additional mutants in which multiple (five or six) residues in loop 2 were replaced with alanine. Although the replacement of multiple residues also resulted in some decrease in mosquitocidal activity, the mutants still showed relatively high activity. Since the insecticidal spectrum of Cry4Aa is specific, Cry4Aa must have a specific receptor on the surface of the target tissue, and loss of binding to the receptor should result in a complete loss of mosquitocidal activity. Our results suggested that, unlike the receptor binding site of the well-characterized molecule Cry1, the receptor binding site of Cry4Aa is different from loops 1, 2, and 3 or that there are multiple binding sites that work cooperatively for receptor binding.Bacillus thuringiensis subsp. israelensis has received considerable attention for mosquito control because of its specific and potent toxicity (15). B. thuringiensis subsp. israelensis-based microbial insecticides have been widely used as active components for integrated management of mosquitoes (11, 13, 33, 34). B. thuringiensis subsp. israelensis produces at least four major crystal toxins (Cry toxins), namely, Cry4Aa, Cry4Ba, Cry11Aa, and Cyt1Aa (5). Cry4Aa exhibits specific toxicity against Anopheles, Aedes, and Culex mosquito larvae (15, 27). The 130-kDa Cry4Aa protoxin is released from the protein crystal upon ingestion by susceptible mosquito larvae and is activated by gut proteases into two protease-resistant fragments with molecular masses of 20 and 45 kDa through intramolecular cleavage of a 60-kDa intermediate (39). The three-dimensional structure of Cry4Aa has been determined by X-ray crystallography at a resolution of 2.8 Å (6). The structure of Cry4Aa is similar to the structures of previously characterized Cry toxins (24, 26, 31) that are composed of three domains (domains I, II, and III). In general, domain I, which is located in the N-terminal region, is composed of seven amphipathic α-helices and is thought to participate in membrane insertion. Domain II, which consists of three antiparallel β-sheets, is a putative receptor binding domain (Fig. (Fig.1).1). In particular, the loops in domain II that are exposed on the surface of the toxin molecule vary significantly in length and amino acid sequence among Cry toxins (31) and are thought to be receptor binding sites. Domain III in the C-terminal region contains two antiparallel β-sheets that form a β-sandwich fold with a jellyroll topology (31). Domain III is assumed to be involved in structural integrity, membrane protein recognition, or both (23, 24, 30).Open in a separate windowFIG. 1.Three-dimensional structure of Cry4Aa domain II. The structure was generated with PyMOL software (8) using the Cry4Aa PDB code (6). The amino acid sequences and corresponding regions of loops 1, 2, and 3 are indicated by blue, red, and yellow, respectively. The amino acid sequences of β-strands adjacent to the loops are underlined.The insecticidal mechanism of Cry toxin involves multiple steps, including ingestion by susceptible insects, solubilization in the alkaline midgut juice, activation by trypsin-like midgut proteases, binding to specific receptors on midgut epithelial cells, and then insertion into the plasma membrane followed by the formation of cation-selective channels or pores (26, 31, 34, 41). According to the colloid-osmotic lysis model, these channels or pores allow ions and water to pass into the cells, resulting in destruction of the membrane potential, cell swelling, cell lysis, and eventual death of the host (20, 21). Thus, the mechanism seems to be very complicated and is affected by multiple factors. The binding of the toxin to the specific receptor is considered a vital step for specific insecticidal activity (35). In fact, modification of the receptor molecules has been reported for insects resistant to certain Cry toxins (12, 22, 36).In a search for the functional structures of Cry4Aa, we previously constructed various loop replacement mutants with mutations in the three major loops in domain II and showed that the replacement of loop 2 resulted in a significant loss of mosquitocidal activity. Replacement of loops 1 and 3 of Cry4Aa also affected mosquitocidal activity, but it did not eliminate it (17). In this study, to further characterize the loops, we constructed Cry4Aa mutants in which individual amino acids in the loops were replaced with alanine and analyzed the mutants to determine their mosquitocidal activity against Culex pipiens larvae. We also analyzed the structural integrity of the Cry4Aa mutant proteins subjected to proteolytic digestion and their binding affinity to brush border membrane vesicles (BBMV) prepared from C. pipiens larvae.  相似文献   

3.
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   

4.
The Cry3Aa and Cry3Bb insecticidal proteins of Bacillus thuringiensis are used in biopesticides and transgenic crops to control larvae of leaf-feeding beetles and rootworms. Cadherins localized in the midgut epithelium are identified as receptors for Cry toxins in lepidopteran and dipteran larvae. Previously, we discovered that a peptide of a toxin-binding cadherin expressed in Escherichia coli functions as a synergist for Cry1A toxicity against lepidopteran larvae and Cry4 toxicity against dipteran larvae. Here we report that the fragment containing the three most C-terminal cadherin repeats (CR) from the cadherin of the western corn rootworm binds toxin and enhances Cry3 toxicity to larvae of naturally susceptible species. The cadherin fragment (CR8 to CR10 [CR8-10]) of western corn rootworm Diabrotica virgifera virgifera was expressed in E. coli as an inclusion body. By an enzyme-linked immunosorbent microplate assay, we demonstrated that the CR8-10 peptide binds α-chymotrypsin-treated Cry3Aa and Cry3Bb toxins at high affinity (11.8 nM and 1.4 nM, respectively). Coleopteran larvae ingesting CR8-10 inclusions had increased susceptibility to Cry3Aa or Cry3Bb toxin. The Cry3 toxin-enhancing effect of CR8-10 was demonstrated for Colorado potato beetle Leptinotarsa decemlineata, southern corn rootworm Diabrotica undecimpunctata howardi, and western corn rootworm. The extent of Cry3 toxin enhancement, which ranged from 3- to 13-fold, may have practical applications for insect control. Cry3-containing biopesticides that include a cadherin fragment could be more efficacious. And Bt corn (i.e., corn treated with B. thuringiensis to make it resistant to pests) coexpressing Cry3Bb and CR8-10 could increase the functional dose level of the insect toxic activity, reducing the overall resistance risk.The Cry3 class of Bacillus thuringiensis Cry proteins is known for toxicity to coleopteran larvae in the family Chrysomelidae. Cry3Aa and Cry3Bb proteins are highly toxic to Colorado potato beetle (CPB) Leptinotarsa decemlineata (Coleoptera: Chrysomelidae), and both were used for the development of Bt crops (crops treated with B. thuringiensis to make them resistant to pests) and Bt biopesticides. Due to the limited efficacy of Cry3-based biopesticides/plants and the success of competing chemical pesticides, these biopesticides have had limited usage and sales (12). Cry3Bb is toxic to corn rootworms (8, 17), and a modified version is expressed in commercialized MON863 corn hybrids (26).Cry3 toxins have a mode of action that is similar to, yet distinct from, the action of lepidopteran-active Cry1 toxins. The Cry3A protoxin (73 kDa) lacks the large C-terminal region of the 130-kDa Cry1 protoxins, which is removed by proteases during activation to toxin. The Cry3A protoxin is activated to a 55-kDa toxin and then further cleaved within the toxin molecule (5, 18). Activated Cry3A toxin binds to brush border membrane vesicles with a Kd (dissociation constant) of ∼37 nM (19) and recognizes a 144-kDa binding protein in brush border membrane vesicles prepared from the yellow mealworm Tenebrio molitor (Coleoptera: Tenebrionidae) (2). Recently, Ochoa-Campuzano et al. (20) identified an ADAM metalloprotease as a receptor for Cry3Aa toxin in CPB larvae.Structural differences between Cry3Bb and Cry3Aa toxins must underlie the unique rootworm activities of Cry3Bb toxin. As noted by Galitsky et al. (11), differences in toxin solubility, oligomerization, and binding are reported for these Cry3 toxins. Recently, Cry3Aa was modified to have activity against western corn rootworm (WCRW) Diabrotica virgifera virgifera (Coleoptera: Chrysomelidae) (27). Those authors introduced a chymotrypsin/cathepsin G cleavage site into domain 1 of Cry3Aa that allowed the processing of the 65-kDa form to a 55-kDa toxin that bound rootworm midgut.Cadherins function as receptors for Cry toxins in lepidopteran and dipteran larvae. A critical Cry1 toxin binding site is localized within the final cadherin repeat (CR), CR12, of cadherins from tobacco hornworm Manduca sexta (Lepidoptera: Sphingidae) and tobacco budworm Heliothis virescens (Lepidoptera: Noctuidae) (14, 28). Unexpectedly, a fragment of B. thuringiensis R1 cadherin, the Cry1A receptor from M. sexta, not only bound toxin but enhanced Cry1A toxicity against lepidopteran larvae (6). If the binding residues within CR12 were removed, the resulting peptide lost the ability to bind toxin and lost its function as a toxin synergist. Recently, we identified a cadherin from mosquito Anopheles gambiae (Diptera: Culicidae) that binds Cry4Ba toxin and probably functions as a receptor. We discovered a similar effect where a fragment of a cadherin from A. gambiae enhanced the toxicity of the mosquitocidal toxin Cry4Ba to mosquito larvae (15). Sayed et al. (22) identified a novel cadherin-like gene in WCRW and proposed this protein as a candidate Bt toxin receptor. The cadherin-like gene is highly expressed in the midgut tissue of larval stages. The encoded protein is conserved in structure relative to that of other insect midgut cadherins.In this study, we hypothesized that a fragment from a beetle cadherin that contains a putative Bt toxin binding region might enhance the insecticidal toxicities of Cry3Aa and Cry3Bb toxins. The region spanning CR8 to CR10 (CR8-10) of the WCRW cadherin (22) was cloned and expressed in E. coli. This cadherin fragment significantly enhanced the toxicities of Cry3Aa and Cry3Bb toxins to CPB and rootworms.  相似文献   

5.
6.
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   

7.
8.
9.
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   

10.
On the basis of the known cry gene sequences of Bacillus thuringiensis, three sets of primers were designed from four conserved blocks found in the delta-endotoxin-coding region. The primer pairs designed amplify the regions between blocks 1 and 5, 2 and 5, and 1 and 4. In silico analyses indicated that 100% of the known three-domain cry gene sequences can be amplified by these sets of primers. To test their ability to amplify known and unknown cry gene sequences, 27 strains from the CINVESTAV (LBIT series) collection showing atypical crystal morphology were selected. Their DNA was used as the template with the new primer system, and after a systematic amplification and sequencing of the amplicons, each strain showed one or more cry-related sequences, totaling 54 different sequences harbored by the 27 strains. Seven sequences were selected on the basis of their low level of identity to the known cry sequences, and once cloning and sequencing of the complete open reading frames were done, three new cry-type genes (primary ranks) were identified and the toxins that they encode were designated Cry57Aa1, Cry58Aa1, and Cry59Aa1 by the B. thuringiensis Toxin Nomenclature Committee. The rest of the seven sequences were classified Cry8Ka2, Cry8-like, Cry20Ba1, and Cry1Ma1 by the committee. The crystal morphology of the selected strains and analysis of the new Cry protein sequences showed interesting peculiarities.Bacillus thuringiensis has been safely used for the control of insect pests within the orders Lepidoptera, Coleoptera, and Diptera for the last 50 years (19). It is an aerobic, Gram-positive bacterium whose insecticidal activity is based on the presence of parasporal crystalline inclusions formed during the sporulation process. These parasporal bodies or crystals are assembled by the so-called Cry proteins, expressed by the cry genes (21).B. thuringiensis shows great variability, as has been demonstrated by the huge number of strains isolated around the world (19), by the number of serotypes known to date (a total of 84) (20), and by the great number of different cry gene sequences accumulated so far (a total of 492) (8), as well as by the number of molecular characterization tools that have been developed, such as sequencing of the flagellin gene and of the gyrB and aroE genes, the band patterns from repetitive extragenic palindromic-PCR analyses, and the plasmid patterns, among others (17, 18, 25, 27), all indicating the great variability within this species.In spite of this variability, some similarity has been found in the three-domain Cry proteins, evidenced by the presence of three conserved domains in the tertiary structure of the active toxin (delta-endotoxin), even from Cry toxins showing low levels of identity (i.e., the Cry1-type, Cry3-type, and Cry4-type toxins): one domain with a bundle of α helices and two domains of β sheets (3). This uniformity is, at least in part, a reflection of the five conserved blocks in the gene structure, which are present in almost all the cry genes (10).The great number of sequences known to date is mostly a result of the strong interest in finding novel Cry proteins, with the foci being on three main purposes: (i) the search for a new range of activities, (ii) the search for higher levels of toxicity, and (iii) the search for alternative toxins in case of resistance development. The search for novel Cry toxins has followed different strategies, such as (i) the development of DNA libraries and their expression in acrystalliferous mutants of B. thuringiensis (24); (ii) hybridization with degenerate oligonucleotides, partial sequences, or complete cry genes used as probes (5, 12); and (iii) PCR amplification of putative novel cry genes by using multiple primers (14) or a combination of general primers, followed either by restriction analysis (26) or by a second amplification using a mixture of general and specific primers (13).These strategies have shown certain levels of efficiency in detecting new cry genes, but they also have some limitations. For example, the screening of a large library is time-consuming and requires a high level serendipity, and the use of Southern analysis with either degenerate or specific probes is limited to finding only sequences related to known cry genes. The amplification of putatively new cry genes by PCR has been widely used in a series of different strategies. However, most of them are limited to finding genes with sequences similar to those used in the primer design. Previous attempts to design universal primers used only nine cry types, which significantly limits the search for actual new genes. Besides, the new sequences described were limited to the amplicons and no new cry genes were reported (6).This report describes the design of a system of three sets of primers, based on conserved regions of the cry family, potentially able to amplify all the known three-domain cry genes. The use of this strategy with B. thuringiensis strains showing atypical crystal morphology allowed the identification of three new cry types and four more new cry holotypes.  相似文献   

11.
12.
Tegument is a unique structure of herpesvirus, which surrounds the capsid and interacts with the envelope. Morphogenesis of gammaherpesvirus is poorly understood due to lack of efficient lytic replication for Epstein-Barr virus and Kaposi''s sarcoma-associated herpesvirus/human herpesvirus 8, which are etiologically associated with several types of human malignancies. Murine gammaherpesvirus 68 (MHV-68) is genetically related to the human gammaherpesviruses and presents an excellent model for studying de novo lytic replication of gammaherpesviruses. MHV-68 open reading frame 33 (ORF33) is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. However, the specific role of ORF33 in gammaherpesvirus replication has not yet been characterized. We describe here that ORF33 is a true late gene and encodes a tegument protein. By constructing an ORF33-null MHV-68 mutant, we demonstrated that ORF33 is not required for viral DNA replication, early and late gene expression, viral DNA packaging or capsid assembly but is required for virion morphogenesis and egress. Although the ORF33-null virus was deficient in release of infectious virions, partially tegumented capsids produced by the ORF33-null mutant accumulated in the cytoplasm, containing conserved capsid proteins, ORF52 tegument protein, but virtually no ORF45 tegument protein and the 65-kDa glycoprotein B. Finally, we found that the defect of ORF33-null MHV-68 could be rescued by providing ORF33 in trans or in an ORF33-null revertant virus. Taken together, our results indicate that ORF33 is a tegument protein required for viral lytic replication and functions in virion morphogenesis and egress.Gammaherpesviruses are associated with tumorigenesis. Like other herpesviruses, they are characterized as having two distinct stages in their life cycle: lytic replication and latency (15, 16, 18, 21, 54). Latency provides the viruses with advantages to escape host immune surveillance and to establish lifelong persistent infection and contributes to transformation and development of malignancies. However, it is through lytic replication that viruses propagate and transmit among hosts to maintain viral reservoirs. Both viral latency and lytic replication play important roles in tumorigenesis. The gammaherpesvirus subfamily includes Epstein-Barr virus (EBV), Kaposi''s sarcoma-associated herpesvirus (KSHV)/human herpesvirus 8 and murine gammaherpesvirus 68 (MHV-68), among others. EBV is associated with Burkitt''s lymphoma, nasopharyngeal carcinoma, Hodgkin''s disease, and lymphoproliferative diseases in immunodeficient patients (28). KSHV is etiologically linked with Kaposi''s sarcoma, primary effusion lymphoma, and multicentric Castleman''s disease (11-13, 22, 52). Neither in vivo nor in vitro studies of EBV and KSHV are convenient due to their propensity to establish latency in cell culture and their limited host ranges.MHV-68 is genetically related to these two human gammaherpesviruses, especially to KSHV, based on the alignment of their genomic sequences and other biological properties (55). As a natural pathogen of wild rodents, MHV-68 also infects laboratory mice (6, 40, 46) and replicates to a high titer in a variety of fibroblast and epithelial cell lines. These advantages make MHV-68 an excellent model for studying the lytic replication of gammaherpesviruses in vitro and certain aspects of virus-host interactions in vivo. In addition, the MHV-68 genome has been cloned as a bacterial artificial chromosome (BAC) that can propagate in Escherichia coli (1, 2, 36, 51), making it convenient to study the function of each open reading frame (ORF) by genetic methods. Exploring the functions of MHV-68 ORFs will likely shed light on the functions of their homologues in human gammaherpesviruses.Gammaherpesviral particles have a characteristic multilayered architecture. An infectious virion contains a double-stranded DNA genome, an icosahedral capsid shell, a thick, proteinaceous tegument compartment, and a lipid bilayer envelope spiked with glycoproteins (14, 30, 47, 49). As a unique structure of herpesviruses, the tegument plays important roles in multiple aspects of the viral life cycle, including virion assembly and egress (38, 48, 53), translocation of nucleocapsids into the nucleus, transactivation of viral immediate-early genes, and modulation of host cell gene expression, innate immunity, and signal transduction (9, 10, 23, 60). Some components of MHV-68 tegument have been identified by a mass spectrometric study (8), and the functions of some tegument proteins have been revealed, such as ORF45, ORF52, and ORF75c (7, 24, 29).MHV-68 ORF33 is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. Its homologues include human herpes simplex virus type 1 (HSV-1) UL16, human herpes simplex virus type 2 (HSV-2) UL16, human cytomegalovirus (HCMV) UL94, EBV BGLF2, KSHV ORF33, and rhesus monkey rhadinovirus (RRV) ORF33. HSV-1 UL16 has been identified as a tegument protein and may function in viral DNA packaging, virion assembly, budding, and egress (5, 32, 35, 41, 44). HCMV UL94 is a virion associated protein and might function in virion assembly and budding (31, 57). EBV BGLF2, KSHV ORF33, and RRV ORF33 are also virion-associated proteins, but their functions are not clear (26, 43, 59). The mass spectrometric study of MHV-68 did not identify ORF33 as a virion component (8), although ORF33 is found to be essential for viral lytic replication by transposon mutagenesis of the MHV-68 genome cloned as a BAC (51). However, insertion of the 1.2-kbp Mu transposon in that study may influence the expression of ORFs approximate to ORF33. Consequently, the role ORF33 plays in viral replication needs to be confirmed, preferably through site-directed mutagenesis. Whether ORF33 is a tegument protein and the exact viral replication stage in which it functions also need to be investigated.We determined that MHV-68 ORF33 encodes a tegument protein and is expressed with true late kinetics. To explore the function of ORF33 in viral lytic phase, we used site-directed mutagenesis and generated an ORF33-null mutant, taking advantage of the MHV-68 BAC system. We showed that the ORF33-null mutant is capable of viral DNA replication, early and late gene expression, capsid assembly, and DNA packaging, but incapable of virion release. The defect of ORF33-null mutant can be rescued in trans by an ORF33 expression plasmid.  相似文献   

13.
Bacillus thuringiensis has been used as a bioinsecticide to control agricultural insects. Bacillus cereus group genomes were found to have a Bacillus enhancin-like (bel) gene, encoding a peptide with 20 to 30% identity to viral enhancin protein, which can enhance viral infection by degradation of the peritrophic matrix (PM) of the insect midgut. In this study, the bel gene was found to have an activity similar to that of the viral enhancin gene. A bel knockout mutant was constructed by using a plasmid-free B. thuringiensis derivative, BMB171. The 50% lethal concentrations of this mutant plus the cry1Ac insecticidal protein gene were about 5.8-fold higher than those of the BMB171 strain. When purified Bel was mixed with the Cry1Ac protein and fed to Helicoverpa armigera larvae, 3 μg/ml Cry1Ac alone induced 34.2% mortality. Meanwhile, the mortality rate rose to 74.4% when the same amount of Cry1Ac was mixed with 0.8 μg/ml of Bel. Microscopic observation showed a significant disruption detected on the midgut PM of H. armigera larvae after they were fed Bel. In vitro degradation assays showed that Bel digested the intestinal mucin (IIM) of Trichoplusia ni and H. armigera larvae to various degrading products, similar to findings for viral enhancin. These results imply Bel toxicity enhancement depends on the destruction of midgut PM and IIM, similar to the case with viral enhancin. This discovery showed that Bel has the potential to enhance insecticidal activity of B. thuringiensis-based biopesticides and transgenic crops.Bacillus thuringiensis is a ubiquitous gram-positive, spore-forming soil bacterium and produces insecticidal crystal proteins during the sporulation phase of its growth cycle. Because these insecticidal crystal proteins have activity against certain insect species, B. thuringiensis has been extensively used as a biopesticide to control crop pests in commercial agriculture and forest management. It is also a key source of genes for transgenic expression and provides pest resistance in plants (2, 20, 30).The viral enhancin protein was originally described for granuloviruses (GVs) as a 126-kDa protein that showed an ability to enhance the infectivity of nucleopolyhedroviruses (NPVs) (36, 37, 39). It has also been found in several other GVs (13) and NPVs (19, 27). Considered a pathogenicity factor, it is not essential for growth of viruses in cell culture or infected insects but has the function of facilitating GV and NPV infection and decreasing larval survival time (14, 17, 19, 27).The widely accepted action mode of the viral enhancin protein, which has been identified as a metalloprotease (17), is that it can disrupt the protective peritrophic matrix (PM), allowing virion access to the underlying epithelial cells of the insect gut (17). The PM has a lattice structure formed by chitin and insect intestinal mucin (IIM), and the viral enhancin protein targets the IIM for degradation (33).Enhancin-like genes with 24 to 25% nucleotide identity to viral enhancin genes have been found in Yersinia pestis, Bacillus anthracis, Bacillus thuringiensis, and Bacillus cereus genome sequences (16, 25, 28). When B. cereus enhancin-like protein was expressed in recombinant Autographa californica multicapsid nucleopolyhedrovirus (AcMNPV) budded viruses and polyhedral inclusion bodies, it was found to be cytotoxic compared to viral enhancin protein. However, larval bioassays indicated that this enhancin-like protein did not enhance infection (8). Hajaij-Ellouze et al. (12) isolated a B. thuringiensis enhancin-like gene from a 407 crystal-minus strain and found that this enhancin-like protein has a typical metalloprotease zinc-binding domain (HEIAH) and belongs to the PlcR regulon. When the enhancin-like mutant was fed to Galleria mellonella larvae, no significant reduction in virulence was observed.In the present study, we report a B. thuringiensis enhancin-like gene (bel) encoding a protein (Bel) that has 20 to 30% identity to the viral enhancin protein and 95% identity to bacterial enhancin-like proteins. Therefore, Bel function may have a synergistic action similar to that of the virus enhancin protein. To understand the biochemical activity of this novel bacterial gene, bel was knocked out in the plasmid-free strain BMB171. We expected that this bel mutant would have no significant reduction in toxicity according to the reports of Galloway et al. (8) and Hajaij-Ellouze et al. (12). However, the bel mutant surprisingly resulted in dramatically reduced Cry1Ac toxicity to Helicoverpa armigera larvae. To further confirm this result, purified Bel was fed together with the Cry1Ac protein to H. armigera larvae. We found that Bel can function as a synergist of Cry1Ac toxicity against H. armigera. In vivo and in vitro observations showed that Bel can disrupt the insect midgut PM and degrade IIM of insect midgut PM. The target of Bel is the IIM of PM, similar to the results found with viral enhancin.  相似文献   

14.
The entomopathogenic fungus Beauveria bassiana acts slowly on insect pests through cuticle infection. Vegetative insecticidal proteins (Vip3A) of Bacillus thuringiensis kill lepidopteran pests rapidly, via per os infection, but their use for pest control is restricted to integration into transgenic plants. A transgenic B. bassiana strain (BbV28) expressing Vip3Aa1 (a Vip3A toxin) was thus created to infect the larvae of the oriental leafworm moth Spodoptera litura through conidial ingestion and cuticle adhesion. Vip3Aa1 (∼88 kDa) was highly expressed in the conidial cytoplasm of BbV28 and was detected as a digested form (∼62 kDa) in the larval midgut 18 and 36 h after conidial ingestion. The median lethal concentration (LC50) of BbV28 against the second-instar larvae feeding on cabbage leaves sprayed with conidial suspensions was 26.2-fold lower than that of the wild-type strain on day 3 and 1.1-fold lower on day 7. The same sprays applied to both larvae and leaves for their feeding reduced the LC50 of the transformant 17.2- and 1.3-fold on days 3 and 7, respectively. Median lethal times (LT50s) of BbV28 were shortened by 23 to 35%, declining with conidial concentrations. The larvae infected by ingestion of BbV28 conidia showed typical symptoms of Vip3A action, i.e., shrinkage and palsy. However, neither LC50 nor LT50 trends differed between BbV28 and its parental strain if the infection occurred through the cuticle only. Our findings indicate that fungal conidia can be used as vectors for spreading the highly insecticidal Vip3A protein for control of foliage feeders such as S. litura.Entomopathogenic fungi, such as Beauveria bassiana and Metarhizium anisopliae, infect hosts through the cuticle and are classic biocontrol agents used against insect pests with either chewing or sucking mouthparts (3, 10, 25, 31). However, their action on target pests is quite slow due to a latent period of several days. This slow action is an obstacle to commercial development of mycoinsecticides (11, 29).Fungal infection starts from the adhesion of conidia to the host cuticle, followed by germination and cuticle penetration into the hemocoel, where fungal cells are propagated by budding until the mycosis-affected host dies from nutrition depletion (10). Conidia ingested by foliage feeders, such as caterpillars, rarely cause substantial infection before excretion due to the lack of intestine-specific virulence factors (29). For this reason, fungal agents are often not very effective against leaf-ingesting insects, which are still able to cause considerable damages before dying of mycosis. Thus, efforts have been increasing to enhance fungal virulence by genetic manipulation. Fungal infection of aphids was accelerated by overexpressing a silkworm chitinase or a hybrid chitinase in B. bassiana (6, 8). Integration of a scorpion neurotoxin in M. anisopliae for specific expression in insect hemolymph after cuticle penetration resulted in 22-, 9-, and 16-fold increases of fungal toxicity to the tobacco hornworm Manduca sexta, the yellow fever mosquito Aedes aegypti, and the coffee berry borer Hypothenemus hampei, respectively (21, 33). Similarly, B. bassiana expressing the scorpion neurotoxin showed a 15-fold increase of insecticidal activity against Masson''s pine caterpillar (Dendrolimus punctatus), and its action on the larvae of D. punctatus and Galleria mellonella was accelerated 24 and 40% (17), respectively. These studies have shed light upon the feasibility of enhancing fungal biocontrol potential by gene transformation. However, none of the genes transformed into the fungal pathogens was an intestine-specific virulence factor to enhance fungal infection per os, although foliage feeders may ingest a large number of conidia sprayed on crop leaves.Vegetative insecticidal proteins (Vips) are a novel class of toxins secreted by Bacillus thuringiensis at the stages of vegetative and stationary growth and show insecticidal activities only in insect intestines (26). Among these, Vip3A has been proven to kill a broad spectrum of lepidopteran insects (2, 5, 18, 19, 26) and may overcome pest resistance to Bacillus thuringiensis endotoxins (Bt endotoxins) (4, 5). As an intestine-specific virulence factor, Vip3A lyses midgut epithelium cells of insects after ingestion (38), forming pores on the cell membrane (12, 13). Not surprisingly, the Vip-encoding genes are considered excellent candidates for new generation of transgenic crops (34). Transgenic plants expressing Vip3A not only exhibit high efficacy against the cotton bollworm Helicoverpa armigera (16, 32) but also are safe to vertebrates (1, 22). However, none of the Vip toxins has been developed commercially like Bt endotoxin formulations, perhaps largely due to their limited yield in cell cultures and an instability of their noncrystal structure. Thus, a new approach is needed to develop Vip-vectoring products for pest control.Fungal conidia can readily be mass produced on solid substrates, such as small grains (36), and usually are formulated as active ingredients of mycoinsecticides (3), making them potential vectors to carry the Vip toxins onto crop foliages by field spray. This study sought to express Vip3Aa1 toxin (a Vip3A member) in a wild-type strain of B. bassiana. Our goal was to enhance the insecticidal activity of transgenic conidia by per os infection in addition to normal cuticle infection. A genetically stable transformant expressing the toxin was generated and assayed for its oral and cuticle infectivity to the larvae of the oriental leafworm moth Spodoptera litura (Noctuidae) in parallel with the wild-type strain.  相似文献   

15.
Porcine circovirus type 1 (PCV1), originally isolated as a contaminant of PK-15 cells, is nonpathogenic, whereas porcine circovirus type 2 (PCV2) causes an economically important disease in pigs. To determine the factors affecting virus replication, we constructed chimeric viruses by swapping open reading frame 1 (ORF1) (rep) or the origin of replication (Ori) between PCV1 and PCV2 and compared the replication efficiencies of the chimeric viruses in PK-15 cells. The results showed that the replication factors of PCV1 and PCV2 are fully exchangeable and, most importantly, that both the Ori and rep of PCV1 enhance the virus replication efficiencies of the chimeric viruses with the PCV2 backbone.Porcine circovirus (PCV) is a single-stranded DNA virus in the family Circoviridae (34). Type 1 PCV (PCV1) was discovered in 1974 as a contaminant of porcine kidney cell line PK-15 and is nonpathogenic in pigs (31-33). Type 2 PCV (PCV2) was discovered in piglets with postweaning multisystemic wasting syndrome (PMWS) in the mid-1990s and causes porcine circovirus-associated disease (PCVAD) (1, 9, 10, 25). PCV1 and PCV2 have similar genomic organizations, with two major ambisense open reading frames (ORFs) (16). ORF1 (rep) encodes two viral replication-associated proteins, Rep and Rep′, by differential splicing (4, 6, 21, 22). The Rep and Rep′ proteins bind to specific sequences within the origin of replication (Ori) located in the intergenic region, and both are responsible for viral replication (5, 7, 8, 21, 23, 28, 29). ORF2 (cap) encodes the immunogenic capsid protein (Cap) (26). PCV1 and PCV2 share approximately 80%, 82%, and 62% nucleotide sequence identity in the Ori, rep, and cap, respectively (19).In vitro studies using a reporter gene-based assay system showed that the replication factors of PCV1 and PCV2 are functionally interchangeable (2-6, 22), although this finding has not yet been validated in a live infectious-virus system. We have previously shown that chimeras of PCV in which cap has been exchanged between PCV1 and PCV2 are infectious both in vitro and in vivo (15), and an inactivated vaccine based on the PCV1-PCV2 cap (PCV1-cap2) chimera is used in the vaccination program against PCVAD (13, 15, 18, 27).PCV1 replicates more efficiently than PCV2 in PK-15 cells (14, 15); thus, we hypothesized that the Ori or rep is directly responsible for the differences in replication efficiencies. The objectives of this study were to demonstrate that the Ori and rep are interchangeable between PCV1 and PCV2 in a live-virus system and to determine the effects of swapped heterologous replication factors on virus replication efficiency in vitro.  相似文献   

16.
17.
KSR1 is a mitogen-activated protein (MAP) kinase scaffold that enhances the activation of the MAP kinase extracellular signal-regulated kinase (ERK). The function of KSR1 in NK cell function is not known. Here we show that KSR1 is required for efficient NK-mediated cytolysis and polarization of cytolytic granules. Single-cell analysis showed that ERK is activated in an all-or-none fashion in both wild-type and KSR1-deficient cells. In the absence of KSR1, however, the efficiency of ERK activation is attenuated. Imaging studies showed that KSR1 is recruited to the immunological synapse during T-cell activation and that membrane recruitment of KSR1 is required for recruitment of active ERK to the synapse.Kinase suppressor of Ras was originally identified in Drosophila melanogaster (53) and Caenorhabditis elegans (19, 32, 52) as a positive regulator of the extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase signaling pathway. It is thought to function as a MAP kinase scaffold because it can bind to Raf, MEK, and ERK (18, 19, 27, 28, 44, 59). While the exact function of KSR is unknown, preassembling the three components of the ERK MAP kinase cascade could function to enhance the efficiency of ERK activation, potentially regulate the subcellular location of ERK activation, and promote access to specific subcellular substrates (16, 45, 46).While only one isoform of KSR is expressed in Drosophila (53), two KSR isoforms have been identified in C. elegans (19, 32, 52) and most higher organisms. They are referred to as KSR1 and KSR2 (32, 43). While KSR1 mRNA and protein are detectable in a wide variety of cells and tissues, including brain, thymus, and muscle (10, 11, 29), little is known about the expression pattern of KSR2.We previously reported the phenotype of KSR1-deficient mice (30). These mice are born at Mendelian ratios and develop without any obvious defects. Using gel filtration, we showed that KSR1 promotes the formation of large signaling complexes containing KSR1, Raf, MEK, and ERK (30). Using both primary T cells stimulated with antibodies to the T-cell receptor as well as fibroblasts stimulated with growth factors, we showed that KSR1-deficient cells exhibit an attenuation of ERK activation with defects in cell proliferation.Here we explored the role of KSR1 in NK cell-mediated cytolysis. The killing of a target cell by a cytolytic T cell or NK cell is a complicated process that involves cell polarization with microtubule-dependent movement of cytolytic granules to an area that is proximal to the contact surface or immunological synapse (7, 33, 34, 48-50, 54). A variety of different signaling molecules are also involved, including calcium (23), phosphatidylinositol-3,4,5-triphosphate (13, 17), and activation of the ERK MAP kinase (6, 42, 56). Recently, the recruitment of activated ERK to the immunological synapse (IS) has been shown to be a feature of successful killing of a target by cytotoxic T lymphocytes (58).How active ERK is recruited to the synapse is not known. Since KSR1 is known to be recruited to the plasma membrane by Ras activation (24), and since the immunological synapse is one of the major sites of Ras activation (26, 41), it seemed plausible to test the hypothesis that KSR1 recruitment to the plasma membrane functions to recruit ERK to the immunological synapse and facilitate its activation. We found that KSR1 was recruited to the immunological synapse and that KSR1 appeared to be required for the localization of active ERK at the contact site. As KSR1-deficient cells exhibit a defect in killing, this suggests that KSR1 recruitment to the synapse may be important in the cytolytic killing of target cells.  相似文献   

18.
19.
Hepatitis E virus genotype 1 strain Sar55 replicated in subcloned Caco-2 intestinal cells and Huh7 hepatoma cells that had been transfected with in vitro transcribed viral genomes, and hepatitis E virions were released into the culture medium of both cell lines. Virus egress from cells depended on open reading frame 3 (ORF3) protein, and a proline-rich sequence in ORF3 was important for egress from cultured cells and for infection of macaques. Both intracellular ORF3 protein accumulation and virus release occurred at the apical membrane of polarized Caco-2 cells. ORF3 protein and lipids were intimately associated with virus particles produced in either cell line; ORF2 epitopes were masked in these particles and could not be immunoprecipitated with anti-ORF2.Hepatitis E virus (HEV) remains enigmatic in spite of recent advances (see references 7 and 16 for reviews). HEV is a major cause of acute hepatitis in numerous developing countries, but hepatitis E is infrequently detected in industrialized countries even though seroprevalence rates of anti-HEV as high as 20% in these countries have been reported. Although hepatitis E normally is a self-limited acute disease, recent studies have identified it as an emerging cause of chronic hepatitis in immunocompromised patients. Whereas contaminated drinking water is the source of most infections in developing countries, the sources in industrialized countries are not fully evaluated, but many, if not most, infections appear linked to eating undercooked meat, especially pork. These differences in epidemiology may reflect the fact that most infections in developing countries are caused by genotypes 1 and 2 while those in industrialized countries are mainly due to genotypes 3 and 4.HEV was initially classified as a calicivirus, but subsequent sequence analysis suggested that it was more closely related to the enveloped rubella virus. However, although HEV may be associated with lipids under some conditions (22), HEV virions do not possess an envelope. Four genotypes of HEV that infect humans have been identified (4). Genotypes 1 and 2 infect primates exclusively, whereas genotypes 3 and 4 are zoonotic and commonly also infect swine and rarely other nonprimates. Recent identification of a strain infecting farmed rabbits in China suggests that other reservoirs may exist (32).The capsid protein encoded by open reading frame 2 (ORF2) is able to form infectious virus particles, but these particles remain cell associated. The crystal structure of a truncated recombinant protein has been solved, but the size of the protein in mature virions is unknown (11, 15, 28, 31). The virus is not cytopathic, and it is unclear how it gets out of cells.The 7.2-kb genome of HEV is a capped mRNA that contains three ORFs that encode proteins involved in replication (ORF1), a capsid protein (ORF2), and a small protein of only 113 to 114 amino acids (ORF3). All but the 5′ terminus of ORF3 is overlapped by ORF2, and both proteins are translated from the same bicistronic subgenomic RNA (10). When overexpressed in cell culture, ORF2 is glycosylated, and ORF3 is phosphorylated (26); this phosphorylated ORF3 protein binds to nonglycosylated ORF2 protein in cell culture, but phosphorylation is not required for infection of macaques (9). The virus has been exceedingly difficult to propagate in cell culture, but recently Okamoto and colleagues reported the successful adaptation of both a genotype 3 and a genotype 4 strain to efficient growth in cultures of PLC/PRF/5 hepatoma or A549 lung cells (23, 24).The tiny ORF3 protein is particularly intriguing because it has a significant impact on virus propagation through mechanisms that have yet to be defined. Data from experiments performed with overexpressed ORF3 protein have suggested that, among other things, ORF3 may interact with cellular proteins, including signaling proteins containing Src homology 3 domains (14), bikunin (27), hemopexin (21), and microtubule proteins (13), and it may function to modulate the acute-phase disease response (3), protect cells from mitochondrial depolarization (18), and enhance expression of glycolytic pathway enzymes (17). Yet within transfected hepatoma cells in culture, virions of an ORF3 null mutant of genotype 1 were assembled in the absence of ORF3 protein and were infectious for naïve hepatoma cells (6) although this same ORF3 null mutant was unable to mount a detectable infection in rhesus monkeys (8). Also, swine transfected with genotype 3 mutant genomes encoding a truncated ORF3 protein did not get infected, indicating that an intact ORF3 protein is needed for infectivity in vivo (12). This lack of infectivity in vivo is possibly explained by the recent demonstration that the ORF3 protein of genotype 3 virus is important for export of virions out of cultured cells in vitro (30); however, this dependence on ORF3 for virion egress has not been confirmed in vivo or for strains of the other three genotypes.The four major genotypes of human HEV appear to segregate naturally into two distinct groups. One group contains genotype 1 and 2 strains that lack a zoonotic component and are spread mainly via contaminated water; in contrast, the second group contains genotype 3 and 4 strains which are able to cross species boundaries and are zoonotic since humans have been infected as a result of eating undercooked meat (16, 25). The molecular basis for the two groupings is unknown, and much more extensive comparative analyses are required to determine which variables are epidemiologically relevant. Here, for lack of an efficient cell culture system for genotype 1 or 2 strains, we have utilized an infectious cDNA clone of a genotype 1 strain in order to explore the role of the ORF3 protein in this group.  相似文献   

20.
Norovirus GII/4 is a leading cause of acute viral gastroenteritis in humans. We examined here how the GII/4 virus evolves to generate and sustain new epidemics in humans, using 199 near-full-length GII/4 genome sequences and 11 genome segment clones from human stool specimens collected at 19 sites in Japan between May 2006 and February 2009. Phylogenetic studies demonstrated outbreaks of 7 monophyletic GII/4 subtypes, among which a single subtype, termed 2006b, had continually predominated. Phylogenetic-tree, bootscanning-plot, and informative-site analyses revealed that 4 of the 7 GII/4 subtypes were mosaics of recently prevalent GII/4 subtypes and 1 was made up of the GII/4 and GII/12 genotypes. Notably, single putative recombination breakpoints with the highest statistical significance were constantly located around the border of open reading frame 1 (ORF1) and ORF2 (P ≤ 0.000001), suggesting outgrowth of specific recombinant viruses in the outbreaks. The GII/4 subtypes had many unique amino acids at the time of their outbreaks, especially in the N-term, 3A-like, and capsid proteins. Unique amino acids in the capsids were preferentially positioned on the outer surface loops of the protruding P2 domain and more abundant in the dominant subtypes. These findings suggest that intersubtype genome recombination at the ORF1/2 boundary region is a common mechanism that realizes independent and concurrent changes on the virion surface and in viral replication proteins for the persistence of norovirus GII/4 in human populations.Norovirus (NoV) is a nonenveloped RNA virus that belongs to the family Caliciviridae and can cause acute gastroenteritis in humans. The NoV genome is a single-stranded, positive-sense, polyadenylated RNA that encodes three open reading frames, ORF1, ORF2, and ORF3 (68). ORF1 encodes a long polypeptide (∼200 kDa) that is cleaved in the cells by the viral proteinase (3Cpro) into six proteins (4). These proteins function in NoV replication in host cells (19). ORF2 encodes a viral capsid protein, VP1. The capsid gene evolved at a rate of 4.3 × 10−3 nucleotide substitutions/site/year (7), which is comparable to the substitution rates of the envelope and capsid genes of human immunodeficiency virus (30). The capsid protein of NoV consists of a shell (S) and two protruding (P) domains: P1 and P2 (47). The S domain is relatively conserved within the same genetic lineages of NoVs (38) and is responsible for the assembly of VP1 (6). The P1 subdomain is also relatively conserved (38) and has a role in enhancing the stability of virus particles (6). The P2 domain is positioned at the most exposed surface of the virus particle (47) and forms binding clefts for putative infection receptors, such as human histo-blood group antigens (HBGA) (8, 13, 14, 60). The P2 domain also contains epitopes for neutralizing antibodies (27, 33) and is consistently highly variable even within the same genetic lineage of NoVs (38). ORF3 encodes a VP2 protein that is suggested to be a minor structural component of virus particles (18) and to be responsible for the expression and stabilization of VP1 (5).Thus far, the NoVs found in nature are classified into five genogroups (GI to GV) and multiple genotypes on the basis of the phylogeny of capsid sequences (71). Among them, genogroup II genotype 4 (GII/4), which was present in humans in the mid-1970s (7), is now the leading cause of NoV-associated acute gastroenteritis in humans (54). The GII/4 is further subclassifiable into phylogenetically distinct subtypes (32, 38, 53). Notably, the emergence and spread of a new GII/4 subtype with multiple amino acid substitutions on the capsid surface are often associated with greater magnitudes of NoV epidemics (53, 54). In 2006 and 2007, a GII/4 subtype, termed 2006b, prevailed globally over preexisting GII/4 subtypes in association with increased numbers of nonbacterial acute gastroenteritis cases in many countries, including Japan (32, 38, 53). The 2006b subtype has multiple unique amino acid substitutions that occur most preferentially in the protruding subdomain of the capsid, the P2 subdomain (32, 38, 53). Together with information on human population immunity against NoV GII/4 subtypes (12, 32), it has been postulated that the accumulation of P2 mutations gives rise to antigenic drift and plays a key role in new epidemics of NoV GII/4 in humans (32, 38, 53).Genetic recombination is common in RNA viruses (67). In NoV, recombination was first suggested by the phylogenetic analysis of an NoV genome segment clone: a discordant branching order was noted with the trees of the 3Dpol and capsid coding regions (21). Subsequently, many studies have reported the phylogenetic discordance using sequences from various epidemic sites in different study periods (1, 10, 11, 16, 17, 22, 25, 40, 41, 44-46, 49, 51, 57, 63, 64, 66). These results suggest that genome recombination frequently occurs among distinct lineages of NoV variants in vivo. However, the studies were done primarily with direct sequencing data of the short genome portion, and information on the cloned genome segment or full-length genome sequences is very limited (21, 25). Therefore, we lack an overview of the structural and temporal dynamics of viral genomes during NoV epidemics, and it remains unclear whether NoV mosaicism plays a role in these events.To clarify these issues, we collected 199 near-full-length genome sequences of GII/4 from NoV outbreaks over three recent years in Japan, divided them into monophyletic subtypes, analyzed the temporal and geographical distribution of the subtypes, collected phylogenetic evidence for the viral genome mosaicism of the subtypes, identified putative recombination breakpoints in the genomes, and isolated mosaic genome segments from the stool specimens. We also performed computer-assisted sequence and structural analyses with the identified subtypes to address the relationship between the numbers of P2 domain mutations at the times of the outbreaks and the magnitudes of the epidemics. The obtained data suggest that intersubtype genome recombination at the ORF1/2 boundary region is common in the new GII/4 outbreaks and promotes the effective acquisition of mutation sets of heterogeneous capsid surface and viral replication proteins.  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