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 0 毫秒
1.
An extensive collection of Bacillus thuringiensis isolates from around the world were phenotypically profiled using standard biochemical tests. Six phenotypic traits occurred in 20–86% of the isolates and were useful in distinguishing isolates: production of urease (U; 20.5% of isolates), hydrolysis of esculin (E; 32.3% of isolates), acid production from salicin (A; 37.4% of isolates), acid production from sucrose (S; 34.0% of isolates), production of phospholipase C or lecithinase (L; 79.7% of isolates), and hydrolysis of starch (T; 85.8% of isolates). With the exception of acid production from salicin and hydrolysis of esculin, which were associated, the traits assorted independently. Of the 64 possible combinations of these six phenotypic characteristics, 15 combinations accounted for ca. 80% of all isolates, with the most common phenotype being TL (23.6% of isolates). Surprisingly, while the biochemical traits generally assorted independently, certain phenotypic traits associated with the parasporal crystal were correlated with certain combinations of biochemical traits. Crystals that remained attached to spores (which tended to be non-toxic to insects) were highly correlated with the phenotypes that included both L and S. Among the 15 most abundant phenotypes characterizing B. thuringiensis strains, amorphous crystals were associated with TLE, TL, T, and Ø (the absence of positive tested biochemical traits). Amorphous crystal types displayed a distinct bias toward toxicity to dipteran insects. Although all common phenotypes included B. thuringiensis isolates producing bipyramidal crystals toxic to lepidopteran insects, those with the highest abundance of these toxic crystals displayed phenotypes TLU, TLUA, TLUAE, and TLAE.  相似文献   

2.
Published data on insecticidal activity of crystal proteins from Bacillus thuringiensis are incorporated into the Bt toxin specificity relational database. To date, 125 of the 174 holotype known toxins have been tested in ∼1700 bioassays against 163 test species; 49 toxins have not been tested at all; 59 were tested against 71 Lepidoptera species in 1182 bioassays; 53 toxins were tested against 23 Diptera species in 233 bioassays; and 47 were tested against 39 Coleoptera species in 190 bioassays. Activity spectra of the tested toxins were summarized for each order. Comparisons of LC50 values are confounded by high variability of the estimates, mostly due to within-species variation in susceptibility, and errors associated with estimation of toxin protein content. Limited analyses suggest that crystal protein toxicity is not affected by quarternary toxin rank or host used for gene expression, but that pre-ingestion treatment by solubilization or enzymatic processing has a large effect. There is an increasing number of toxin families with cross-order activity, as 15 of the 87 families (secondary rank) that are pesticidal are active against more than one order. Cross-order activity does not threaten environmental safety of B. thuringiensis-based pest control because toxins tend to be much less toxic to taxa outside the family’s primary specificity range.  相似文献   

3.
Transgenic rice to control stem borer damage is under development in China. To assess the potential of Bacillus thuringiensis (Bt) transgenes in stem borer control, the toxicity of five Bt protoxins (Cry1Aa, Cry1Ab, Cry1Ac, Cry1Ba and Cry1Ca) against two rice stem borers, Sesamia inferens (pink stem borer) and Chilo suppressalis (striped stem borer), was evaluated in the laboratory by feeding neonate larvae on artificial diets containing Bt protoxins. The results indicated that Cry1Ca exhibited the highest level of toxicity to both stem borers, with an LC50 of 0.24 and 0.30 μg/g for C. suppressalis and S. inferens, respectively. However, S. inferens was 4-fold lower in susceptibility to Cry1Aa, and 6- and 47-fold less susceptible to Cry1Ab and Cry1Ba, respectively, compared to C. suppressalis. To evaluate interactions among Bt protoxins in stem borer larvae, toxicity assays were performed with mixtures of Cry1Aa/Cry1Ab, Cry1Aa/Cry1Ca, Cry1Ac/Cry1Ca, Cry1Ac/Cry1Ba, Cry1Ab/Cry1Ac, Cry1Ab/Cry1Ba, and Cry1Ab/Cry1Ca at 1:1 (w/w) ratios. All protoxin mixtures demonstrated significant synergistic toxicity activity against C. suppressalis, with values of 1.6- to 11-fold higher toxicity than the theoretical additive effect. Surprisingly, all but one of the Bt protoxin mixtures were antagonistic in toxicity to S. inferens. In mortality-time response experiments, S. inferens demonstrated increased tolerance to Cry1Ab and Cry1Ac compared to C. suppressalis when treated with low or high protoxin concentrations. The data indicate the utility of Cry1Ca protoxin and a Cry1Ac/Cry1Ca mixture to control both stem borer populations.  相似文献   

4.
Rapid detection of Bacillus spores is a challenging task in food and defense industries. In situ labeling of spores would be advantageous for detection by automated systems based on single-cell analysis. Determination of antibiotic-resistance genes in bacterial spores using in situ labeling has never been developed. Most of the in situ detection schemes employ techniques such as fluorescence in situ hybridization (FISH) that target the naturally amplified ribosomal RNA (rRNA). However, the majority of antibiotic-resistance genes has a plasmidic or chromosomal origin and is present in low copy numbers in the cell. The main challenge in the development of low-target in situ detection in spores is the permeabilization procedure and the signal amplification required for detection. This study presents permeabilization and in situ signal amplification protocols, using Bacillus cereus spores as a model, in order to detect antibiotic-resistance genes. The permeabilization protocol was designed based on the different layers of the Bacillus spore. Catalyzed reporter deposition (CARD)–FISH and in situ polymerase chain reaction (PCR) were used as signal amplification techniques. B. cereus was transformed with the high copy number pC194 and low copy number pMTL500Eres plasmids in order to induce resistance to chloramphenicol and erythromycin, respectively. In addition, a rifampicin-resistant B. cereus strain, conferred by a single-nucleotide polymorphism (SNP) in the chromosome, was used. Using CARD–FISH, only the high copy number plasmid pC194 was detected. On the other hand, in situ PCR gave positive results in all tested instances. This study demonstrated that it was feasible to detect antibiotic-resistance genes in Bacillus spores using in situ techniques. In addition, in situ PCR has been shown to be more sensitive and more applicable than CARD–FISH in detecting low copy targets.  相似文献   

5.
Bacillus cereus sphingomyelinase (Bc-SMase) induces hemolysis of sheep erythrocytes which contain large amounts of sphingomyelin. We investigated the mechanism of this hemolysis in comparison to that induced by Clostridium perfringens alpha-toxin. Pertussis toxin, a Gi-specific inhibitor, N-oleoylethernolamine, a ceramidase inhibitor, and dihydrosphingosine, a sphingosine kinase inhibitor, did not inhibit the hemolysis by Bc-SMase, but did inhibit that by alpha-toxin. Bc-SMase broadly bound to whole membranes, and alpha-toxin specifically bound to the detergent-resistant membrane fractions, lipid rafts. The level of ceramide production induced by Bc-SMase in sheep erythrocytes was 6- to 15-fold that induced by alpha-toxin, when the extent of the hemolysis by Bc-SMase was the same as that by the toxin. However, the level of ceramide production induced by Bc-SMase in SM-liposomes was equal to that triggered by the toxin, when the carboxyl fluorescein-release from liposomes induced by Bc-SMase was the same as that induced by alpha-toxin. Confocal laser microscopy showed that treatment of the cells with Bc-SMase resulted in the formation of ceramide-rich domains. A photobleaching analysis suggested that treatment of the cells with Bc-SMase leads to a reduction in membrane fluidity. These results show that Bc-SMase-induced hemolysis of sheep erythrocytes is related to the formation of interface between ceramide-rich domains and ceramide-poor domains through production of ceramide from SM.  相似文献   

6.
A field population (SZ) of Plutella xylostella, collected from the cabbage field in Shenzhen, Guangdong Province of China in 2002, showed 2.3-fold resistance to Cry1Aa, 110-fold to Cry1Ab, 30-fold to Cry1Ac, 2.1-fold to Cry1F, 5.3-fold to Cry2Aa and 6-fold resistance to Bacillus thuringiensis var. kurstaki (Btk) compared with a susceptible strain (ROTH). The SZBT strain was derived from the SZ population through 20 generations of selection with activated Cry1Ac in the laboratory. While the SZBT strain developed 1200-fold resistance to Cry1Ac after selection, resistance to Cry1Aa, Cry1Ab, Cry1F, and Btk increased to 31-, 1900-,>33- and 17-fold compared with the ROTH strain. However, little or no cross-resistance was detected to Cry1B, Cry1C and Cry2Aa in the SZBT strain. Genetic cross analyses between the SZBT and ROTH strains revealed that Cry1Ac-resistance in the SZBT strain was controlled by a single, autosomal, incompletely recessive gene. Binding studies with 125I-labeled Cry1Ac showed that the brush border membrane vesicles (BBMVs) of midguts from the resistant SZBT insects had lost binding to Cry1Ac. Allelic complementation tests demonstrated that the major Bt resistance locus in the SZBT strain was same as that in the Cry1Ac-R strain which has “mode 1” resistance to Bt. An F1 screen of 120 single-pair families between the SZBT strain and three field populations collected in 2008 was carried out. Based on this approach, the estimated frequencies of Cry1Ac-resistance alleles were 0.156 in the Yuxi population from Yunnan province, and 0.375 and 0.472 respectively in the Guangzhou and Huizhou populations from Guangdong province.  相似文献   

7.
Adenosine phosphorylase, a purine nucleoside phosphorylase endowed with high specificity for adenine nucleosides, was purified 117-fold from vegetative forms of Bacillus cereus. The purification procedure included ammonium sulphate fractionation, pH 4 treatment, ion exchange chromatography on DEAE-Sephacel, gel filtration on Sephacryl S-300 HR and affinity chromatography on N6-adenosyl agarose. The enzyme shows a good stability to both temperature and pH. It appears to be a homohexamer of 164 ± 5 kDa. Kinetic characterization confirmed the specificity of this phosphorylase for 6-aminopurine nucleosides. Adenosine was the preferred substrate for nucleoside phosphorolysis (kcat/Km 2.1 × 106 s− 1 M− 1), followed by 2′-deoxyadenosine (kcat/Km 4.2 × 105 s− 1 M− 1). Apparently, the low specificity of adenosine phosphorylase towards 6-oxopurine nucleosides is due to a slow catalytic rate rather than to poor substrate binding.  相似文献   

8.
Bacillus thuringiensis is an insect pathogen used worldwide as a bioinsecticide. It belongs to the Bacillus cereus sensu lato group as well as Bacillus anthracis and B. cereus. Plasmids from this group of organisms have been implicated in pathogenicity as they carry the genes responsible for different types of diseases that affect mammals and insects. Some plasmids, like pAW63 and pBT9727, encode a functional conjugation machinery allowing them to be transferred to a recipient cell. They also share extensive homology with the non-functional conjugation apparatus of pXO2 from B. anthracis. In this study we report the complete sequence of three plasmids from an environmental B. thuringiensis isolate from Argentina, obtained by a shotgun sequencing method. We obtained the complete nucleotide sequence of plasmids pFR12 (12 095 bp), pFR12.5 (12 459 bp) and pFR55 (55 712 bp) from B. thuringiensis INTA-FR7-4. pFR12 and pFR12.5 were classified as cryptic as they do not code for any obvious functions besides replication and mobilization. Both small plasmids were classified as RCR plasmids due to similarities with the replicases they encode. Plasmid pFR55 showed a structural organization similar to that observed for plasmids pAW63, pBT9727 and pXO2. pFR55 also shares a tra region with these plasmids, containing genes related to T4SS and conjugation. A comparison between pFR55 and conjugative plasmids led to the postulation that pFR55 is a conjugative plasmid. Genes related to replication functions in pFR55 are different to those described for plasmids with known complete sequences. pFR55 is the first completely sequenced plasmid with a replication machinery related to that of ori44. The analysis of the complete sequence of plasmids from an environmental isolate of B. thuringiensis permitted the identification of a near complete conjugation apparatus in pFR55, resembling those of plasmids pAW63, pBT9727 and pXO2. The availability of this sequence is a step forward in the study of the molecular basis of the conjugative process in Gram positive bacteria, particularly due to the similarity with known conjugation systems. It is also a contribution to the expansion of the non-pathogenic B. cereus plasmid gene pool.  相似文献   

9.
We examined interaction between Bacillus thuringiensis subsp. kurstaki HD-1 (Foray 48B) and larval midgut bacteria in two lepidopteran hosts, Lymantria dispar and Choristoneura fumiferana. The pathogen multiplied in either moribund (C. fumiferana) or dead (L. dispar) larvae, regardless of the presence of midgut bacteria. Inoculation of L. dispar resulted in a pronounced proliferation of enteric bacteria, which did not contribute to larval death because B. thuringiensis was able to kill larvae in absence of midgut bacteria. Sterile, aureomycin- or ampicillin-treated larvae were killed in a dose-dependent manner but there was no mortality among larvae treated with the antibiotic cocktail used by [Broderick et al., 2006] and [Broderick et al., 2009]. These results do not support an obligate role of midgut bacteria in insecticidal activity of HD-1. The outcome of experiments on the role of midgut bacteria may be more dependent on which bacterial species are dominant at the time of experimentation than on host species per se. The L. dispar cohorts used in our study had a microflora, that was dominated by Enterococcus and Staphylococcus and lacked Enterobacter. Another factor that can confound experimental results is the disk-feeding method for inoculation, which biases mortality estimates towards the least susceptible portion of the test population.  相似文献   

10.
The characterization of the strains containing Coleopteran-specific and also putative novel cry genes in Iranian native Bacillus thuringiensis collection is presented. Characterization was based on PCR analysis using 31 general and specific primers for cry1B, cry1I, cry3A, cry3B, cry3C, cry7A, cry8A, cry8B, cry8C, cry14, cry18, cry26, cry28, cry34 and cry35 genes, protein band patterns as well as their insecticidal activity on Xanthogaleruca luteola Mull. larvae. Forty six isolates (65.7%) contained minimum one Coleopteran-active cry gene. Based on universal primers, strains containing cry18 and cry26 genes were the most abundant and represent 27.1% and 24% of the isolates, respectively, whereas cry14, cry3, cry28, cry34, cry35, cry7, cry8 genes were less abundant, found in 14.2, 12.5, 10, 7, 7 and 5.6% of the strains, respectively. Based on specific primers, isolates containing cry1I were the most abundant (48.5%). Two strains containing Coleopteran-active cry genes showed higher activity against X. luteola larvae than B. thuringiensis subsp. morrisoni pathovar tenebrionis. Thirty isolates, when assayed for cry1C, cry5, cry6, cry8b, cry9, cry10, cry11, cry18, cry24 and cry35 genes, showed unexpected size bands. Cloning and sequencing of the amplicons allowed both the identification of known cry genes and the detection of putative novel cry1C sequences.  相似文献   

11.
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   

12.
Bacillus thuringiensis (Bt) Cry8D insecticidal proteins are unique among Cry8 family proteins in terms of its insecticidal activity against adult Scarab beetles, such as Japanese beetle (Popillia japonica Newman). From the sequence homology with other Bt Cry proteins especially those active against beetles, such as Cry3Aa whose 3D structure is available, the structure of the Cry8D protein has been predicted to be a typical three-domain Cry protein type. In addition, the activation process of Cry8D in gut juice of susceptible insects is presumed to be similar to that of Cry3A (Yamaguchi et al., 2008). In this study, the activation process of Cry8Da in insect gut juice was closely examined. Japanese beetle gut juice proteases digested the 130 kDa Cry8Da protein to produce a 64 kDa protein. This 64 kDa protein was active against both adult and larval Japanese beetle and considered to be an activated toxin. N-terminal sequencing of this 64 kDa protein revealed that the Cry8Da leader sequence consisting of 63 amino acid residues from M1 to F63 was removed. As in the case of Cry3Aa, the proteases further digested the 64 kDa protein to two 8 kDa and 54 kDa fragments. N-terminal amino acid analysis of these smaller fragments indicated that the proteases digested the loop between Alpha Helix (Alpha for short) 3 and Alpha 4. This means that the 8 kDa fragment consists of Alpha 1-3 of Domain I and that the 54 kDa fragment contains the remaining Domain I and full Domain II and Domain III. Size exclusion chromatography and anion exchange chromatography could not separate these 64, 54 and 8 kDa proteins suggesting that the 54 kDa and 8 kDa fragments are still forming the toxin complex equivalent to the 64 kDa protein by size and ionic charge. The sequencing and chromatography results suggest that the gut juice proteases merely nicked the loop between Alpha 3 and Alpha 4. This nicking process appeared to be essential for receptor binding of the Cry8Da toxin. BBMV binding assay revealed that the Cry8Da toxin bound to BBMV preparations from both adult and larval Japanese beetle only after the loop was nicked. Only the 54 kDa fragment bound to the BBMV preparations but not the 64 kDa protein. Ligand blot showed that the protease activated Cry8Da toxin, presumably the 54 kDa fragment, bound to specific BBMV proteins, one or more of those would be receptor(s). The sizes and binding affinities of these Cry8Da-bound proteins of Japanese beetle BBMV differed between larvae and adults.  相似文献   

13.
After binding to specific receptors, Cry toxins form pores in the midgut apical membrane of susceptible insects. The receptors could form part of the pore structure or simply catalyze pore formation and consequently be recycled. To discriminate between these possibilities, the kinetics of pore formation in brush border membrane vesicles isolated from Manduca sexta was studied with an osmotic swelling assay. Pore formation, as deduced from changes in membrane permeability induced by Cry1Ac during a 60-min incubation period, was strongly dose-dependent, but rapidly reached a maximum as toxin concentration was increased. Following exposure of the vesicles to the toxin, the osmotic swelling rate reached a maximum shortly after a delay period. Under these conditions, at relatively high toxin concentrations, the maximal osmotic swelling rate increased linearly with toxin concentration. When vesicles were incubated for a short time with the toxin and then rapidly cooled to prevent the formation of new pores before and during the osmotic swelling experiment, a plateau in the rate of pore formation was observed as toxin concentration was increased. Taken together, these results suggest that the receptors do not act as simple catalysts of pore formation, but remain associated with the pores once they are formed.  相似文献   

14.
A cadherin-like protein has been identified as a putative receptor for Bacillus thuringiensis (Bt) Cry1Ac toxin in Helicoverpa armigera and plays a key role in Bt insecticidal action. In this study, we produced a fragment from this H. armigera Cry1Ac toxin-binding cadherin that included the predicted toxin-binding region. Binding of Cry1Ac toxin to this cadherin fragment facilitated the formation of a 250-kDa toxin oligomer. The cadherin fragment was evaluated for its effect on Cry1Ac toxin-binding and toxicity by ligand blotting, binding assays, and bioassays. The results of ligand blotting and binding assays revealed that the binding of Cry1Ac to H. armigera midgut epithelial cells was reduced under denaturing or native conditions in vitro. Bioassay results indicated that toxicities from Cry1Ac protoxin or activated toxin were reduced in vivo by the H. armigera cadherin fragment. The addition of the cadherin fragment had no effect on Cry2Ab toxicity.  相似文献   

15.
The precise mechanisms underlying Bacillus thuringiensis-mediated killing of pest insects are not clear. In some cases, death may be due to septicaemia caused by Bt and/or gut bacteria gaining access to the insect haemocoel. Since insects protect themselves from microbes using an array of cellular and humoral immune defences, we aimed to determine if a recombinant immunosuppressive wasp venom protein (rVPr1) could increase the susceptibility of two pest Lepidoptera (Lacanobia oleracea and Mamestra brassicae) to Bt. Bio-assays indicated that injection of 6 μl of rVPr1 into the haemocoel of both larvae caused similar levels of mortality (less than 38%). On the other hand, the LD30-40 of Bt for M. brassicae larvae was approximately 20 times higher than that for L. oleracea larvae. Furthermore, in bio-assays where larvae were injected with rVPr1, then fed Bt, a significant reduction in survival of larvae for both species occurred compared to each treatment on its own (P < 0.001); and for L. oleracea larvae, this effect was more than additive. The results are discussed within the context of insect immunity and protection against Bt.  相似文献   

16.
The toxicity and pore-forming ability of the Bacillus thuringiensis Cry9Ca insecticidal toxin, its single-site mutants, R164A and R164K, and the 55-kDa fragment resulting from its proteolytic cleavage at residue 164 were investigated using Manduca sexta neonate larvae and fifth-instar larval midgut brush border membrane vesicles, respectively. Neither the mutations nor the proteolytic cleavage altered Cry9Ca toxicity. Compared with Cry1Ac, Cry9Ca and its mutants formed large poorly selective pores in the vesicles. Pore formation was highly dependent on pH, however, especially for wild-type Cry9Ca and both mutants. Increasing pH from 6.5 to 10.5 resulted in an irregular step-wise decrease in membrane permeabilization that was not related to a change in the ionic selectivity of the pores. Pore formation was much slower with Cry9Ca and its derivatives, including the 55-kDa fragment, than with Cry1Ac and its rate was not influenced by the presence of protease inhibitors or a reducing agent.  相似文献   

17.
To understand the low toxicity of Cry toxins in planthoppers, proteolytic activation of Cry1Ab in Nilaparvata lugens was studied. The proteolytic processing of Cry1Ab protoxin by N. lugens midgut proteases was similar to that by trypsin activated Cry1Ab. The Cry1Ab processed with N. lugens midgut proteases was highly insecticidal against Plutella xylostella. However, Cry1Ab activated either by trypsin or the gut proteases of the brown planthopper showed low toxicity in N. lugens. Binding analysis showed that activated Cry1Ab bound to brush border membrane vesicles (BBMV) from N. lugens at a significantly lower level than to BBMV from P. xylostella.  相似文献   

18.
Microbiological studies of a spent nuclear fuel pool in Argentina were performed to evaluate the risk of microbiological induced corrosion and determine the cultivable bacterial population. Based on standard methods and sequencing of the 16sRNA gene, eighteen microorganisms were identified. Bacillus cereus RE 10 was the predominant organism isolated, and was selected to investigate the biofilm formation process and the corrosion effect on aluminum alloy AA 6061 and on pure aluminum (Al 99.999%). To simulate the environmental conditions, the experiments were performed using a highly diluted medium. After 20 days of exposure, major pits covered with deposits were found on AA 6061 samples exposed to B. cereus RE 10 but not on Al 99.999%. There was a close correlation between biofilm patches, corrosion deposits, pitting and Al-(Fe or Ti)-Si inclusions. We postulate that this correlation is a consequence of a light local alkalinization around the inclusions that produces changes in the expression pattern of B. cereus RE 10 and allows bacterial survival using other substrates. Under these conditions, the generated biofilm induces a crevice corrosion effect around the intermetallic inclusions of the alloy. Our results will be useful for further studies related to the microbial impact on nuclear safety in nuclear waste storage facilities in Argentina.  相似文献   

19.
The pore-forming ability of the Bacillus thuringiensis toxin Cry9Ca, its two single-site mutants R164A and R164K, and the 55-kDa fragment resulting from its proteolytic cleavage at R164 was evaluated under a variety of experimental conditions using an electrophysiological assay. All four toxin preparations depolarized the apical membrane of freshly isolated third-instar Manduca sexta midguts bathing in a solution containing 122 mM KCl at pH 10.5, but the 55-kDa fragment was considerably more active than Cry9Ca and its mutants. The activity of the latter toxins was greatly enhanced, however, when the experiments were conducted in the presence of fifth-instar M. sexta midgut juice. This effect was also observed after midgut juice proteins had been denatured by heating at 95 °C or after inorganic ions and small molecules had been removed from the midgut juice by extensive dialysis. A similar stimulation of toxin activity was also observed when the experiments were carried out in the presence of the lipids extracted from an equivalent volume of midgut juice. Depolarization of the cell membrane was also greatly enhanced, in the absence of midgut juice, by the addition of a cocktail of water-soluble protease inhibitors. These results indicate that, depending on the cleavage site and on the experimental conditions used, further proteolysis of the activated Cry9Ca toxin can either stimulate or be detrimental to its activity and that M. sexta midgut juice probably contains protease inhibitors that could play a major role in the activity of B. thuringiensis toxins in the insect midgut.  相似文献   

20.
Bioassays of insecticidal proteins from Bacillus thuringiensis subsp. israelensis with larvae of the malaria vector mosquito Anophelesalbimanus showed that the cytolytic protein Cyt1Aa was not toxic alone, but it increased the toxicity of the crystalline proteins Cry4Ba and Cry11Aa. Synergism also occurred between Cry4Ba and Cry11Aa toxins. Whereas many previous analyses of synergism have been based on a series of toxin concentrations leading to comparisons between expected and observed values for the concentration killing 50% of insects tested (LC50), we describe and apply a method here that enables testing for synergism based on single concentrations of toxins.  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