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 843 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
The Clostridium perfringens ϵ-toxin is responsible for a severe, often lethal intoxication. In this study, we characterized dominant-negative inhibitors of the ϵ-toxin. Site-specific mutations were introduced into the gene encoding ϵ-toxin, and recombinant proteins were expressed in Escherichia coli. Paired cysteine substitutions were introduced at locations predicted to form a disulfide bond. One cysteine in each mutant was introduced into the membrane insertion domain of the toxin; the second cysteine was introduced into the protein backbone. Mutant proteins with cysteine substitutions at amino acid positions I51/A114 and at V56/F118 lacked detectable cytotoxic activity in a MDCK cell assay. Cytotoxic activity could be reconstituted in both mutant proteins by incubation with dithiothreitol, indicating that the lack of cytotoxic activity was attributable to the formation of a disulfide bond. Fluorescent labeling of the cysteines also indicated that the introduced cysteines participated in a disulfide bond. When equimolar mixtures of wild-type ϵ-toxin and mutant proteins were added to MDCK cells, the I51C/A114C and V56C/F118C mutant proteins each inhibited the activity of wild-type ϵ-toxin. Further analysis of the inhibitory activity of the I51C/A114C and V56C/F118C mutant proteins indicated that these proteins inhibit the ability of the active toxin to form stable oligomeric complexes in the context of MDCK cells. These results provide further insight into the properties of dominant-negative inhibitors of oligomeric pore-forming toxins and provide the basis for developing new therapeutics for treating intoxication by ϵ-toxin.The Clostridium perfringens ϵ-toxin is one of the most potent bacterial toxins (1, 2). The ϵ-toxin can lead to a fatal enterotoxemia characterized by widespread vascular permeability and edema in the heart, lungs, brain, and kidneys (36). The disease most frequently affects livestock animals, though the toxin may also affect humans (79). Because of its extreme potency and the possibility of intoxicating humans, the C. perfringens ϵ-toxin is considered a select agent by the United States Department of Health and Human Services. A vaccine currently is approved for veterinary use, though multiple immunizations are required to provide long-term immunity (1013). There also is an antitoxin approved for veterinary use. However, in the event that an animal exhibits symptoms of intoxication by ϵ-toxin, it is typically too late for the current antitoxin to be effective, and use of the antitoxin is typically limited to prophylactic treatment of unvaccinated animals within a herd (14). There is no treatment currently approved for use in humans. Thus, alternative countermeasures are needed that inhibit the activity of the toxin.One alternative method of countering the cytotoxic activity of bacterial toxins is through dominant-negative inhibitors. Dominant-negative inhibitors are non-cytotoxic mutant forms of active toxins that are able to inhibit the activity of wild-type toxin when the two proteins are mixed together. Such dominant-negative inhibitors have been described for a diverse set of toxins, including Helicobacter pylori VacA (1519), Bacillus anthracis anthrax toxin protective antigen (2025), Bacillus thuringiensis Cry1Ab (26), and Escherichia coli ClyA cytotoxin (27). Like VacA, protective antigen, Cry1Ab, and ClyA, the ϵ-toxin assembles into oligomeric complexes containing multiple toxin monomers (2830). In the case of VacA and protective antigen, the most extensively studied examples of toxins inhibited by dominant-negative mutants, the number of mutations that inactivate the toxins is substantially greater than the number of mutations that lead to a dominant-negative phenotype (16, 17, 24, 31, 32). Although many of the mutations leading to dominant-negative toxins are located within regions of the toxins that are believed to form the membrane insertion domain, some mutations that inactivate the toxins (but are not dominant-negative) also map within the predicted membrane insertion domains (24, 32). Thus, a deeper understanding of the nature of the dominant-negative phenotype is needed.In this study, we sought to generate dominant-negative mutants of the ϵ-toxin. We hypothesized that mutations within the membrane insertion domain of ϵ-toxin, particularly mutations that are expected to restrict movement of this domain, would lead to dominant-negative inhibitors. We expressed wild-type and site-specific mutants of the ϵ-toxin as recombinant proteins in E. coli. The recombinant proteins were purified, and cytotoxicity was assessed using an established cell culture assay. Using this approach, we identified mutant proteins that inhibited the activity of wild-type ϵ-toxin in vitro and determined the mechanism of inhibition.  相似文献   

6.
The mqsR gene has been shown to be positively regulated by the quorum-sensing autoinducer AI-2, which in turn activates a two-component system, the qseB-qseC operon. This operon plays an important role in biofilm formation in Escherichia coli. However, its cellular function has remained unknown. Here, we found that 1 base downstream of mqsR there is a gene, ygiT, that is co-transcribed with mqsR. Induction of mqsR caused cell growth arrest, whereas ygiT co-induction recovered cell growth. We demonstrate that MqsR (98 amino acid residues), which has no homology to the well characterized mRNA interferase MazF, is a potent inhibitor of protein synthesis that functions by degrading cellular mRNAs. In vivo and in vitro primer extension experiments showed that MqsR is an mRNA interferase specifically cleaving mRNAs at GCU. The mRNA interferase activity of purified MqsR was inhibited by purified YgiT (131 residues). MqsR forms a stable 2:1 complex with YgiT, and the complex likely functions as a repressor for the mqsR-ygiT operon by specifically binding to two different palindromic sequences present in the 5′-untranslated region of this operon.It has been reported that quorum sensing is involved in biofilm formation (14). mqsR expression was found to be induced by 8-fold in biofilms (5) and also by the quorum-sensing signal autoinducer AI-2, which is a species-nonspecific signaling molecule produced by both Gram-negative and Gram-positive bacteria, including Escherichia coli (6). It has been reported that induction of mqsR activates a two-component system, the qseB-qseC operon, which is known to play an important role in biofilm formation (6). Thus, it has been proposed that MqsR (98 amino acid residues) is a regulator of biofilm formation because it activates qseB, which controls the flhDC expression required for motility and biofilm formation in E. coli (6). However, the cellular function of MqsR has remained unknown.Interestingly, all free-living bacteria examined to date contain a number of suicide or toxin genes in their genomes (7, 8). Many of these toxins are co-transcribed with their cognate antitoxins in an operon (termed toxin-antitoxin (TA)2 operon) and form a stable complex in the cell, so their toxicity is subdued under normal growth conditions (911). However, the stability of antitoxins is substantially lower than that of their cognate toxins, so any stress causing cellular damage or growth inhibition that induces proteases alters the balance between toxin and antitoxin, leading to toxin release in the cell.To date, 16 (12) TA systems have been reported on the E. coli genome, including relB-relE (13, 14), chpBI-chpBK (15), mazEF (1618), yefM-yoeB (19, 20), dinJ-yafQ (21, 22), hipBA and hicAB (23, 24), prlF-yhaV (25), and ybaJ-hha (26). Interestingly, all of these TA operons appear to use similar modes of regulation: the formation of complexes between antitoxins and their cognate toxins to neutralize toxin activity and the ability of TA complexes to autoregulate their expression. The cellular targets of some toxins have been identified. CcdB directly interacts with gyrase A and blocks DNA replication (27, 28). RelE, which by itself has no endoribonuclease activity, appears to act as a ribosome-associating factor that promotes mRNA cleavage at the ribosome A-site (13, 29, 30). PemK (31), ChpBK (15), and MazF (32) are unique among toxins because they target cellular mRNAs for degradation by functioning as sequence-specific endoribonucleases to effectively inhibit protein synthesis and thereby cell growth.MazF, ChpBK, and PemK have been characterized as sequence-specific endoribonucleases that cleave mRNA at the ACA, ACY (Y is U, A, or G), and UAH (H is C, A, or U) sequences, respectively. They are completely different from other known endoribonucleases such as RNases E, A, and T1, as these toxins function as protein synthesis inhibitors by interfering with the function of cellular mRNAs. It is well known that small RNAs, such as mRNA-interfering cRNA (33), microRNA (34), and small interfering RNA (35), interfere with the function of specific RNAs. These small RNAs bind to specific mRNAs to inhibit their expression. Ribozymes also act on their target RNAs specifically and interfere with their function (36). Therefore, MazF, ChpBK, and PemK homologs form a novel endoribonuclease family that exhibits a new mRNA-interfering mechanism by cleaving mRNAs at specific sequences. Thus, they have been termed “mRNA interferases” (2).During our search for TA systems on the E. coli genome, we found that the mqsR gene is co-transcribed with a downstream gene, ygiT. These two genes appear to function as a TA system, as their size is small (98 residues for MqsR and 131 residues for YgiT) and their respective open reading frames are separated by 1 bp. In this study, we demonstrate that MqsR-YgiT is a new E. coli TA system consisting of a toxin, MqsR, and an antitoxin, YgiT. Moreover, we identify MqsR as a novel mRNA interferase that does not exhibit homology to MazF. This toxin cleaves RNA at GCU sequences in vivo and in vitro. The implication of this finding as to how this mRNA interferase is involved in cell physiology and biofilm formation will be discussed.  相似文献   

7.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

8.
9.
10.
11.
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   

12.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

13.
14.
15.
16.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

17.
Src family kinases (SFKs) signal in response to E-cadherin to support cadherin adhesion and the integrity of cell-cell contacts (McLachlan, R. W., Kraemer, A., Helwani, F. M., Kovacs, E. M., and Yap, A. S. (2007) Mol. Biol. Cell 18, 3214–3223). We now identify the actin-regulatory protein, cortactin, as a target of E-cadherin-activated SFK signaling. Tyr-phosphorylated cortactin was found at cell-cell contacts in established epithelial monolayers, and cortactin became acutely tyrosine-phosphorylated when E-cadherin adhesion was engaged. In all circumstances, cortactin tyrosine phosphorylation was blocked by inhibiting SFK signaling. Importantly, Tyr-phosphorylated cortactin was necessary to preserve the integrity of cadherin contacts and the perijunctional actin cytoskeleton. Moreover, expression of a phosphomimetic cortactin mutant could prevent SFK blockade from disrupting cadherin organization, thereby placing cortactin functionally downstream of SFK signaling at cadherin adhesions. We conclude that SFK and cortactin constitute an important signaling pathway that functionally links E-cadherin adhesion and the actin cytoskeleton.Functional cooperation between cadherin adhesion receptors and the actin cytoskeleton is commonly believed to play a key role in the morphogenesis of cell-cell interactions (1, 2). This functional interplay, and the biochemical mechanisms that underpin it, are much more complex than previously realized. Increasingly it is apparent that a range of distinct actin regulators can be recruited to cadherin adhesions depending on the biological context of cell-cell interactions (2). It is likely that the choice of actin regulator(s) recruited determines the dynamics and organization of the actin cytoskeleton at those contacts, with morphogenetic implications for the formation, modeling, and turnover of cell-cell interactions. Identifying the actin regulators that influence cell-cell interactions and how they cooperate with adhesion receptors are important open issues.Adhesion-activated cell signaling provides a useful paradigm to analyze how classical cadherins regulate the actin cytoskeleton (2, 3). Over the past several years, a range of signal transduction pathways have been identified that are stimulated upon productive engagement of cadherins, such as E-, C-, and N-cadherin (reviewed in Ref. 3). Among these signals are Rho family GTPases, lipid kinases, and protein-tyrosine kinases. Of the latter, we recently identified Src family kinase (SFK)5 activity as a component of E-cadherin signaling (4). SFK was stimulated in an E-cadherin-dependent fashion when cells assembled contacts with one another. Indeed, binding to recombinant cadherin ligands was sufficient to activate SFK, implying that the cadherin itself can serve as a receptor to transduce an adhesive signal to SFK. Furthermore, inhibiting SFK signaling perturbed cadherin adhesion and the integrity of cell-cell contacts. This suggested a model where adhesive ligation of E-cadherin stimulated an SFK signaling cascade to ultimately support cell-cell interactions. An important challenge now is to identify targets of cadherin-activated SFK signaling that contribute to cadherin biology.In this work, we tested whether the actin-binding protein, cortactin, might be just such a target. A multidomain scaffolding protein, cortactin regulates the actin cytoskeleton by interacting with a range of other actin-regulatory proteins (5, 6). It is best understood to participate in actin filament assembly (6) by promoting Arp2/3-mediated actin nucleation and also by stabilizing nascent Arp2/3-generated actin filament branches (7). Cortactin exerts many of these effects through direct interactions with actin filaments and Arp2/3 (8) as well as indirectly by associating with proteins such as N-WASP and WIP, which can themselves activate Arp2/3. Consistent with this, cortactin is often found at sites in the cortex where Arp2/3 drives membrane protrusion, such as lamellipodia and invadopodia (9).Cortactin is also found at cadherin-based cell-cell contacts where a biochemical complex with E-cadherin or N-cadherin has been detected by co-immunoprecipitation analysis (1012). Moreover, ligation of the cadherin with recombinant ligands could induce formation of a complex with cortactin and also recruited cortactin to the cortex at the sites of adhesion (10, 12). Cortactin is found with Arp2/3 at newly forming E-cadherin adhesive contacts, and disruption of cortactin by RNAi or dominant-negative mutants perturbs efficient assembly of cadherin-based contacts (10). At N-cadherin adhesions, cortactin promotes adhesive strengthening and its surface expression (12). Overall, these reports identified a role for cortactin in modulating cadherin biology, likely through regulation of the cadherin-based actin cytoskeleton.Of note, cortactin was first identified as a substrate for v-Src (13) and as a target of fibroblast growth factor-stimulated SFK signaling (14). Tyrosine phosphorylation of cortactin is implicated in cellular events that are accompanied by extensive remodeling of the actin cytoskeleton, such as cell migration and invasion (6, 15, 16). Building on our earlier experience in epithelial cells (4, 10), we now report that E-cadherin ligation induces the tyrosine phosphorylation of cortactin through an SFK-dependent signaling pathway. Furthermore, we demonstrate that phosphorylation at the key Tyr-421, Tyr-466, and Tyr-482 residues is necessary to maintain the integrity of established cell-cell contacts and their perijunctional actin cytoskeleton.  相似文献   

18.
19.
20.
A complete understanding of the biological functions of large signaling peptides (>4 kDa) requires comprehensive characterization of their amino acid sequences and post-translational modifications, which presents significant analytical challenges. In the past decade, there has been great success with mass spectrometry-based de novo sequencing of small neuropeptides. However, these approaches are less applicable to larger neuropeptides because of the inefficient fragmentation of peptides larger than 4 kDa and their lower endogenous abundance. The conventional proteomics approach focuses on large-scale determination of protein identities via database searching, lacking the ability for in-depth elucidation of individual amino acid residues. Here, we present a multifaceted MS approach for identification and characterization of large crustacean hyperglycemic hormone (CHH)-family neuropeptides, a class of peptide hormones that play central roles in the regulation of many important physiological processes of crustaceans. Six crustacean CHH-family neuropeptides (8–9.5 kDa), including two novel peptides with extensive disulfide linkages and PTMs, were fully sequenced without reference to genomic databases. High-definition de novo sequencing was achieved by a combination of bottom-up, off-line top-down, and on-line top-down tandem MS methods. Statistical evaluation indicated that these methods provided complementary information for sequence interpretation and increased the local identification confidence of each amino acid. Further investigations by MALDI imaging MS mapped the spatial distribution and colocalization patterns of various CHH-family neuropeptides in the neuroendocrine organs, revealing that two CHH-subfamilies are involved in distinct signaling pathways.Neuropeptides and hormones comprise a diverse class of signaling molecules involved in numerous essential physiological processes, including analgesia, reward, food intake, learning and memory (1). Disorders of the neurosecretory and neuroendocrine systems influence many pathological processes. For example, obesity results from failure of energy homeostasis in association with endocrine alterations (2, 3). Previous work from our lab used crustaceans as model organisms found that multiple neuropeptides were implicated in control of food intake, including RFamides, tachykinin related peptides, RYamides, and pyrokinins (46).Crustacean hyperglycemic hormone (CHH)1 family neuropeptides play a central role in energy homeostasis of crustaceans (717). Hyperglycemic response of the CHHs was first reported after injection of crude eyestalk extract in crustaceans. Based on their preprohormone organization, the CHH family can be grouped into two sub-families: subfamily-I containing CHH, and subfamily-II containing molt-inhibiting hormone (MIH) and mandibular organ-inhibiting hormone (MOIH). The preprohormones of the subfamily-I have a CHH precursor related peptide (CPRP) that is cleaved off during processing; and preprohormones of the subfamily-II lack the CPRP (9). Uncovering their physiological functions will provide new insights into neuroendocrine regulation of energy homeostasis.Characterization of CHH-family neuropeptides is challenging. They are comprised of more than 70 amino acids and often contain multiple post-translational modifications (PTMs) and complex disulfide bridge connections (7). In addition, physiological concentrations of these peptide hormones are typically below picomolar level, and most crustacean species do not have available genome and proteome databases to assist MS-based sequencing.MS-based neuropeptidomics provides a powerful tool for rapid discovery and analysis of a large number of endogenous peptides from the brain and the central nervous system. Our group and others have greatly expanded the peptidomes of many model organisms (3, 1833). For example, we have discovered more than 200 neuropeptides with several neuropeptide families consisting of as many as 20–40 members in a simple crustacean model system (5, 6, 2531, 34). However, a majority of these neuropeptides are small peptides with 5–15 amino acid residues long, leaving a gap of identifying larger signaling peptides from organisms without sequenced genome. The observed lack of larger size peptide hormones can be attributed to the lack of effective de novo sequencing strategies for neuropeptides larger than 4 kDa, which are inherently more difficult to fragment using conventional techniques (3437). Although classical proteomics studies examine larger proteins, these tools are limited to identification based on database searching with one or more peptides matching without complete amino acid sequence coverage (36, 38).Large populations of neuropeptides from 4–10 kDa exist in the nervous systems of both vertebrates and invertebrates (9, 39, 40). Understanding their functional roles requires sufficient molecular knowledge and a unique analytical approach. Therefore, developing effective and reliable methods for de novo sequencing of large neuropeptides at the individual amino acid residue level is an urgent gap to fill in neurobiology. In this study, we present a multifaceted MS strategy aimed at high-definition de novo sequencing and comprehensive characterization of the CHH-family neuropeptides in crustacean central nervous system. The high-definition de novo sequencing was achieved by a combination of three methods: (1) enzymatic digestion and LC-tandem mass spectrometry (MS/MS) bottom-up analysis to generate detailed sequences of proteolytic peptides; (2) off-line LC fractionation and subsequent top-down MS/MS to obtain high-quality fragmentation maps of intact peptides; and (3) on-line LC coupled to top-down MS/MS to allow rapid sequence analysis of low abundance peptides. Combining the three methods overcomes the limitations of each, and thus offers complementary and high-confidence determination of amino acid residues. We report the complete sequence analysis of six CHH-family neuropeptides including the discovery of two novel peptides. With the accurate molecular information, MALDI imaging and ion mobility MS were conducted for the first time to explore their anatomical distribution and biochemical properties.  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