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 500 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.
Pore formation in the apical membrane of the midgut epithelial cells of susceptible insects constitutes a key step in the mode of action of Bacillus thuringiensis insecticidal toxins. In order to study the mechanism of toxin insertion into the membrane, at least one residue in each of the pore-forming-domain (domain I) interhelical loops of Cry1Aa was replaced individually by cysteine, an amino acid which is normally absent from the activated Cry1Aa toxin, using site-directed mutagenesis. The toxicity of most mutants to Manduca sexta neonate larvae was comparable to that of Cry1Aa. The ability of each of the activated mutant toxins to permeabilize M. sexta midgut brush border membrane vesicles was examined with an osmotic swelling assay. Following a 1-h preincubation, all mutants except the V150C mutant were able to form pores at pH 7.5, although the W182C mutant had a weaker activity than the other toxins. Increasing the pH to 10.5, a procedure which introduces a negative charge on the thiol group of the cysteine residues, caused a significant reduction in the pore-forming abilities of most mutants without affecting those of Cry1Aa or the I88C, T122C, Y153C, or S252C mutant. The rate of pore formation was significantly lower for the F50C, Q151C, Y153C, W182C, and S252C mutants than for Cry1Aa at pH 7.5. At the higher pH, all mutants formed pores significantly more slowly than Cry1Aa, except the I88C mutant, which formed pores significantly faster, and the T122C mutant. These results indicate that domain I interhelical loop residues play an important role in the conformational changes leading to toxin insertion and pore formation.Once ingested by susceptible insect larvae, the insecticidal crystal proteins of Bacillus thuringiensis are solubilized and converted to their toxic form by midgut proteases. The activated toxins bind to specific receptors on the surface of the luminal membrane of midgut columnar cells, insert into the membrane, and form pores that abolish transmembrane ionic gradients and osmotic balance, leading to the disruption of the epithelium and death of the insect (47, 51). Members of the B. thuringiensis Cry toxin family for which the atomic structure has been reported share a similar three-domain organization in which domain I is composed of a bundle of six amphipathic α-helices surrounding a hydrophobic helix (α5), and domains II and III are formed mostly of β-sheets (7, 8, 18, 26, 37, 38, 43). While domains II and III are thought to be involved in receptor binding and toxin specificity (47), domain I is believed to play a major role in membrane insertion and pore formation (51). Toxin fragments corresponding to domain I of Cry1Ac (62), Cry3Aa (53), and Cry3Ba (61) or to the first five α-helices of Cry4B (48) have been shown to form pores in model membranes. Pore formation in artificial membranes has also been demonstrated with synthetic peptides corresponding to α5 of Cry1Ac (13) and Cry3Aa (19, 21) and to the α4-loop-α5 segment of Cry3Aa (23). Spectroscopic studies have also revealed that while synthetic peptides corresponding to α4 and α5 can coassemble within a lipid bilayer, those corresponding to α2, α3, α6, and α7 adopt a membrane surface orientation (20, 22). In agreement with these findings, α4 was shown to line the lumen of the pores (42). On the other hand, convincing evidence supporting previous suggestions that most of the toxin molecule may become imbedded in the membrane (3, 39, 60) has recently been reported (44, 45).Thus, several models have been proposed for the mechanism of toxin insertion and pore formation (4, 9, 28, 32, 39, 44, 52, 56). Although these models differ in the identities of the toxin segments that are suggested to insert into the membrane, they all imply that the toxin undergoes conformational changes following binding to the membrane surface. Even though such changes imply rotations about the polypeptide backbone in domain I interhelical loops, little attention has been devoted so far to the role of domain I loop residues in pore formation.In the present study, amino acid residues strategically located within each of these loops in Cry1Aa were replaced by a cysteine using site-directed mutagenesis. The resulting mutant toxins were assayed with Manduca sexta midgut brush border membrane vesicles using a light-scattering technique. Mutations mapping within several of these loops altered the functional properties of Cry1Aa, suggesting the involvement of most domain I α-helices in the pore-forming process.  相似文献   

3.
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   

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.
Bacillus thuringiensis subsp. israelensis is the most widely used microbial control agent against mosquitoes and blackflies. Its insecticidal success is based on an arsenal of toxins, such as Cry4A, Cry4B, Cry11A, and Cyt1A, harbored in the parasporal crystal of the bacterium. A fifth toxin, Cry10Aa, is synthesized at very low levels; previous attempts to clone and express Cry10Aa were limited, and no parasporal body was formed. By using a new strategy, the whole Cry10A operon was cloned in the pSTAB vector, where both open reading frames ORF1 and ORF2 (and the gap between the two) were located, under the control of the cyt1A operon and the STAB-SD stabilizer sequence characteristic of this vector. Once the acrystalliferous mutant 4Q7 of B. thuringiensis subsp. israelensis was transformed with this construct, parasporal bodies were observed by phase-contrast microscopy and transmission electron microscopy. Discrete, ca. 0.9-μm amorphous parasporal bodies were observed in the mature sporangia, which were readily purified by gradient centrifugation once autolysis had occurred. Pure parasporal bodies showed two major bands of ca. 68 and 56 kDa on sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis. These bands were further characterized by N-terminal sequencing of tryptic fragments using matrix-assisted laser desorption ionization-time of flight mass spectrometry analysis, which identified both bands as the products of ORF1 and ORF2, respectively. Bioassays against fourth-instar larvae of Aedes aegypti of spore-crystal complex and pure crystals of Cry10Aa gave estimated 50% lethal concentrations of 2,061 ng/ml and 239 ng/ml, respectively. Additionally, synergism was clearly detected between Cry10A and Cyt1A, as the synergistic levels (potentiation rates) were estimated at 13.3 for the mixture of Cyt1A crystals and Cry10Aa spore-crystal complex and 12.6 for the combination of Cyt1A and Cry10Aa pure crystals.The subspecies Bacillus thuringiensis subsp. israelensis (serotype H-14) was discovered by Goldberg and Margalit in 1977 (11). To date, its insecticidal potential has not been overcome by any other bacterium (or any biological control agent) as an effective control measure against mosquito and blackfly larvae (8). Recently, its toxicity spectrum has been expanded to a coleopteran pest, the coffee berry borer (Hypothenemus hampei) (23), indicating that this strain may have potential versatility. Also, the so-called pBtoxis megaplasmid harbored in this strain, containing all the endotoxin-encoding genes found in its parasporal crystal, including cry4A, cry4B, cry10A, cry11A, and cyt1A, was recently sequenced (1). Among many other interesting aspects of this serotype, the occurrence of this mosquitocidal arsenal in one strain and their synergistic interaction make this bacterium scientifically and technologically attractive.The parasporal crystal of B. thuringiensis subsp. israelensis contains large amounts of Cry4A, Cry4B, Cry11A, and Cyt1A toxins (14), and consequently, most of the knowledge about the toxicity of this strain has been focused on these proteins, acting either as a complex (31) or tested separately (6). Although the cry10Aa gene was originally cloned in 1986 (known then as cryIVC) (30), to date, little is known about cry10Aa and the protein it encodes, mostly due to its very low levels of expression (10) in B. thuringiensis subsp. israelensis. Interestingly, cry10Aa is an operon as it includes two open reading frames (ORFs), previously reported as pBt047 and pBt048 (hereafter referred to only as ORF1 and ORF2, respectively), separated by a 48-bp untranslated gap (1). ORF1 contains the complete δ-endotoxin sequence (active toxin), with a coding capacity for a 78-kDa protein. Interestingly, ORF2 shows high identity with the coding sequence of the C-terminal half of Cry4-type proteins, with a coding capacity for a 56-kDa protein. Therefore, it is believed that a putative ancestral cry10Aa gene is similar in size to the cry4-type genes (ca. 4 kbp), but either a small sequence had been inserted in the middle of the coding sequence or site mutations produced end codons (two end codons flank the gap) in this region (1).Previous attempts to clone and express the cry10Aa gene included ORF1 and only part of ORF2 (7, 10, 30). This was a reasonable strategy, as most of the so-called “complete” protoxins are partially digested to become active toxins (δ-endotoxins) (28), and ORF1 included the complete sequence to code the Cry10Aa δ-endotoxin. However, in all these cases, the expression levels were very low, and no parasporal body was formed. Similar results were obtained when the promoter was changed and a stabilizing sequence was added to the construction (13). The low expression levels achieved in these cases led to conclusions that assumed low toxic levels of Cry10Aa when tested against mosquito larvae (30). In spite of the low toxicity of Cry10Aa found against mosquito larvae, a synergistic effect was reported between Cry10Aa and Cry4Ba toxins in Culex (7). Obtaining high levels of expression and crystallization of Cry10Aa are required to properly characterize and understand the toxic spectrum of this protein.In this report, we show the formation of parasporal bodies of Cry10Aa, achieved by cloning the whole Cry10Aa operon under the control of the cyt1A promoter and the STAB-SD sequence. We also show that Cry10Aa is as toxic as most of the other B. thuringiensis subsp. israelensis toxins acting separately, and in synergism with the Cyt1A toxin.  相似文献   

6.
7.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

8.
9.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

10.
The binding affinities and specificities of six truncated S-layer homology domain (SLH) polypeptides of mosquitocidal Bacillus sphaericus strain C3-41 with the purified cell wall sacculi have been assayed. The results indicated that the SLH polypeptide comprised of amino acids 31 to 210 was responsible for anchoring the S-layer subunits to the rigid cell wall layer via a distinct type of secondary cell wall polymer and that a motif of the recombinant SLH polypeptide comprising amino acids 152 to 210 (rSLH152-210) was essential for the stable binding of the S-layer with the bacterial cell walls. The quantitative assays revealed that the KD (equilibrium dissociation constant) values of rSLH152-210 and rSLH31-210 with purified cell wall sacculi were 1.11 × 10−6 M and 1.40 × 10−6 M, respectively. The qualitative assays demonstrated that the SLH domain of strain C3-41 could bind only to the cell walls or the cells treated with 5 M guanidinium hydrochloride of both toxic and nontoxic B. sphaericus strains but not to those from other bacteria, indicating the species-specific binding of the SLH polypeptide of B. sphaericus with bacterial cell walls.Crystalline bacterial cell surface layers (S-layers) cover the cell surfaces of many bacteria and archaea during all stages of growth and division. S-layers are composed of identical protein or glycoprotein subunits, which can assemble into two-dimensional crystalline arrays and exhibit oblique, square, or hexagonal symmetry (27, 28, 30). S-layers play key roles in the interaction between bacterial cells and environment as protective coats, molecular sieves, ion traps, cell adhesion mediators, and attachment structures (4, 21, 26, 29). Many S-layer proteins possess an N-terminal region with highly conserved amino acid sequences, which is called an S-layer homology (SLH) domain. An SLH domain contains one, two, or three repeating SLH motifs (6, 16). Each SLH motif is composed of about 55 amino acids containing 10 to 15 conserved residues (6, 17). It is suggested that the SLH domain of S-layer proteins is responsible for the binding of the S-layer subunits to the rigid cell wall layer (6, 15, 17, 19, 25), while the middle and C-terminal parts include the domains which are involved in the self-assembly process (27). In the case of Bacillaceae, secondary cell wall polymers (SCWP) are responsible for binding with SLH domains (13, 18, 19), but the SLH domains of some other bacteria have an affinity for peptidoglycan (33).Bacillus sphaericus is a gram-positive soil bacterium that represents a strictly aerobic group of mesophilic endospore-forming bacteria. Due to its specific toxicity to target mosquito larvae and the limited environment impact, some strains of this bacterium have been successfully used worldwide in integrated mosquito control programs. Previous studies revealed that some nontoxic strains of B. sphaericus contained S-layer proteins, and the S-layer proteins of B. sphaericus NCTC 9602, JG-A12, P1, and CCM 2177 have been studied in detail elsewhere (3, 7-9, 12, 22).B. sphaericus C3-41, a highly active strain isolated from a mosquito-breeding site in China in 1987, has different levels of toxicity against Culex spp., Anopheles spp., and Aedes spp. This strain belongs to the flagella serotype H5a5b, like strains 2362 and 1593 (32), and it has been developed as a commercial larvicide (JianBao) for mosquito larva control in China during the last decade (31). The genomic analysis of strain C3-41 revealed that an S-layer protein gene (slpC) (GenBank accession no. EF535606) exists on the chromosomal genome and its sequence is identical to the S-layer protein of B. sphaericus 2362 (1, 10), composed of 3,531 bp encoding a protein of 1,176 amino acids with a molecular size of 125 kDa. Although the binding function of S-layers has been identified in some nontoxic B. sphaericus strains (6, 11), it is not well documented in mosquitocidal B. sphaericus strains, and there are few reports on the binding function of each SLH motif and the binding specificity.In this study, the binding affinities and specificities of each SLH motif of S-layer protein from mosquitocidal B. sphaericus C3-41 alone and in combination with the different cell wall preparations have been investigated, and the species-specific binding of SLH polypeptide with bacterial cell walls has been demonstrated.  相似文献   

11.
12.
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   

13.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

14.
15.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  相似文献   

16.
17.
18.
19.
20.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  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