首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
To determine the possibility of plasmid transfer occurring between strains of Bacillus thuringiensis in infected lepidopterous larvae, Galleria mellonella and Spodoptera littoralis were infected with two or more strains of B. thuringiensis and the resulting bacteria from the dead insects were examined for plasmid transfer. Transfer rates of plasmids coding for crystal production and tetracycline resistance were high, reaching levels similar to those obtained in laboratory broth cultures. Transfer was higher in G. mellonella than S. littoralis, probably due to the greater ability of B. thuringiensis to colonize the larvae. In broth cultures, B. thuringiensis was also able to transfer plasmids into sporeforming bacteria present in soil samples. The results suggest that plasmid transfer between strains of B. thuringiensis occurs in nature, resulting in the production of new combinations of delta-endotoxins within populations of the bacteria.  相似文献   

2.
The Bacillus thuringiensis var. thuringiensis strain 3A produces a proteinaceous parasporal crystal toxic to larvae of a variety of lepidopteran pests including Spodoptera littoralis (Egyptian cotton leaf worm), Heliothis zeae, H. virescens and Boarmia selenaria. By cloning of individual plasmids of B. thuringiensis in Escherichia coli, we localized a gene coding for the delta-endotoxin on the B. thuringiensis plasmid of about 17 kb designated pTN4. Following partial digestion of the B. thuringiensis plasmid pTN4 and cloning into the E. coli pACYC184 plasmid three clones were isolated in which toxin production was detected. One of these hybrid plasmids pTNG43 carried a 1.7-kb insert that hybridized to the 14-kb BamHI DNA fragments of B. thuringiensis var. thuringiensis strains 3A and berliner 1715. This BamHI DNA fragment of strain berliner 1715 has been shown to contain the gene that codes for the toxic protein of the crystal (Klier et al., 1982). No homologous sequences have been found between pTNG33 and the DNA of B. thuringiensis var. entomocidus strain 24, which exhibited insecticidal activity against S. littoralis similar to that of strain 3A.  相似文献   

3.
To facilitate the analysis of genetic determinants carried by large resident plasmids of Bacillus anthracis, a mating system was developed which promotes plasmid transfer among strains of B. anthracis, B. cereus, and B. thuringiensis. Transfer of the selectable tetracycline resistance plasmid pBC16 and other plasmids from B. thuringiensis to B. anthracis and B. cereus recipients occurred during mixed incubation in broth. Two plasmids, pXO11 and pXO12, found in B. thuringiensis were responsible for plasmid mobilization. B. anthracis and B. cereus transcipients inheriting either pXO11 or pXO12 were, in turn, effective donors. Transcipients harboring pXO12 were more efficient donors than those harboring pXO11; transfer frequencies ranged from 10(-4) to 10(-1) and from 10(-8) to 10(-5), respectively. Cell-to-cell contact was necessary for plasmid transfer, and the addition of DNase had no effect. The high frequencies of transfer, along with the fact that cell-free filtrates of donor cultures were ineffective, suggested that transfer was not phage mediated. B. anthracis and B. cereus transcipients which inherited pXO12 also acquired the ability to produce parasporal crystals (Cry+) resembling those produced by B. thuringiensis, indicating that pXO12 carries a gene(s) involved in crystal formation. Transcipients which inherited pXO11 were Cry-. This mating system provides an efficient method for interspecies transfer of a large range of Bacillus plasmids by a conjugation-like process.  相似文献   

4.
Plasmid transfer between strains of Bacillus thuringiensis subsp. israelensis was studied under a range of environmentally relevant laboratory conditions in vitro, in river water, and in mosquito larvae. Mobilization of pBC16 was detected in vitro at a range of temperatures, pH values, and available water conditions, and the maximum transfer ratio was 10(-3) transconjugant per recipient under optimal conditions. Transfer of conjugative plasmid pXO16::Tn5401 was also detected under this range of conditions. However, a maximum transfer ratio of 1.0 transconjugant per recipient was attained, and every recipient became a transconjugant. In river water, transfer of pBC16 was not detected, probably as a result of the low transfer frequency for this plasmid and the formation of spores by the introduced donor and recipient strains. In contrast, transfer of plasmid pXO16::Tn5401 was detected in water, but at a lower transfer ratio (ca. 10(-2) transconjugant per donor). The number of transconjugants increased over the first 7 days, probably as a result of new transfer events between cells, since growth of both donor and recipient cells in water was not detected. Mobilization of pBC16 was not detected in killed mosquito larvae, but transfer of plasmid pXO16::Tn5401 was evident, with a maximum rate of 10(-3) transconjugant per donor. The reduced transfer rate in insects compared to broth cultures may be accounted for by competition from the background bacterial population present in the mosquito gut and diet or by the maintenance of a large population of B. thuringiensis spores in the insects.  相似文献   

5.
The transfer of plasmids by mating from four Bacillus thuringiensis subspecies to Bacillus anthracis and Bacillus cereus recipients was monitored by selecting transcipients which acquired plasmid pBC16 (Tcr). Transcipients also inherited a specific large plasmid from each B. thuringiensis donor at a high frequency along with a random array of smaller plasmids. The large plasmids (ca. 50 to 120 megadaltons), pXO13, pXO14, pXO15, and pXO16, originating from B. thuringiensis subsp. morrisoni, B. thuringiensis subsp. toumanoffi, B. thuringiensis subsp. alesti, and B. thuringiensis subsp. israelensis, respectively, were demonstrated to be responsible for plasmid mobilization. Transcipients containing any of the above plasmids had donor capability, while B. thuringiensis strains cured of each of them were not fertile, indicating that the plasmids confer conjugation functions. Confirmation that pXO13, pXO14, and pXO16 were self-transmissible was obtained by the isolation of fertile B. anthracis and B. cereus transcipients that contained only pBC16 and one of these plasmids. pXO14 was efficient in mobilizing the toxin and capsule plasmids, pXO1 and pXO2, respectively, from B. anthracis transcipients to plasmid-cured B. anthracis or B. cereus recipients. DNA-DNA hybridization experiments suggested that DNA homology exists among pXO13, pXO14, and the B. thuringiensis subsp. thuringiensis conjugative plasmids pXO11 and pXO12. Matings performed between strains which each contained the same conjugative plasmid demonstrated reduced efficiency of pBC16 transfer. However, in many instances when donor and recipient strains contained different conjugative plasmids, the efficiency of pBC16 transfer appeared to be enhanced.  相似文献   

6.
The cry toxin encoding plasmid pHT73 was transferred from Bacillus thuringiensis subspecies kurstaki KT0 to six B. cereus group strains in three lepidopteran (Spodoptera exigua, Plutella xyllostella and Helicoverpa armigera) larvae by conjugation. The conjugation kinetics of the plasmid was precisely studied during the larval infection using a new protocol. The infections were performed with both vegetative and sporulated strains. However, larval death only occurred when infections were made with spore and toxin preparations. Likewise, spore germinations of both donor and recipient strains were only observed in killed larvae, 44–56 h post-infection. Accordingly, kinetics showed that gene transfer between B. thuringiensis strain KT0 and other B. cereus strains only took place in dead larvae among vegetatively growing bacteria. The conjugational transfer ratios varied among different strain combinations and different larvae. The highest transfer ratio reached 5.83 × 10−6 CFU/donor between the KT0 and the AW05R recipient in Helicoverpa armigera, and all transconjugants gained the ability to produce the insecticidal crystal. These results indicated that horizontal gene transfer among B. cereus group strains might play a key role for the acquisition of extra plasmids and evolution of these strains in toxin susceptible insect larvae.  相似文献   

7.
The cry genes of Bacillus thuringiensis encode a diverse group of crystal-forming proteins that exhibit insecticidal activity, particularly against the larvae of lepidopteran, coleopteran, and dipteran insects. The efficacy of B. thuringiensis-based biopesticides may be improved through the genetic manipulation of these genes. A gene transfer system has been developed for the introduction and maintenance of cloned insecticidal cry genes on small plasmids in B. thuringiensis. This vector system combines a B. thuringiensis plasmid replicon and an indigenous site-specific recombination system that allows for the selective removal of ancillary or foreign DNA from the recombinant bacterium after introduction of the Cry-encoding plasmid. The site-specific recombination system is useful for engineering strains with unique combinations of cry genes, resulting in new active ingredients with improved insecticidal properties.  相似文献   

8.
Bacillus thuringiensis harbors genes encoding Cry proteins found in chromosomes or plasmids of different sizes (4-150 Mb). Although the smaller plasmids are more abundant in B. thuringiensis, their specific function is unknown. As for the megaplasmids, their main recognized function is to harbor cry genes, although the sequencing of some of these plasmids indicates the occurrence of other important genes. This work used a new protocol for practical and rapid extraction of plasmid DNA in order to characterize the plasmid patterns of Brazilian strains belonging to Embrapa Milho e Sorgo research center B. thuringiensis bank. We tried to further assess the relationship of plasmid patterns with strains belonging to the same serovars and strains causing 100% and no mortality to Spodoptera frugiperda (J.E. Smith) larvae. It was possible to characterize 59 strains based on the migration of bands in agarose gel. Strains belonging to the same serovars showed different plasmid sizes (from 1,636 bp to 23,200 bp), with the exception of two strains belonging to serovar galleriae. The strain T09 Bt tolworthi showed a plasmid migration pattern identical to strains belonging to serovar galleriae. Plasmid patterns differed for 46 strains, confirming that this is a useful tool to discriminate specific strains. However, it was not possible to associate the plasmid pattern or the occurrence of particular plasmids with the pathogenicity of a given species towards S. frugiperda larvae.  相似文献   

9.
Studies on the interaction of the insect pathogenic bacterium, Xenorhabdus nematophilus (Enterobacteriaceae), with its nematode and insect hosts would be greatly assisted if a luminescent phenotype were generated that would allow the detection of viable bacteria in vivo without the necessity for disruption of the cellular interactions. The plasmid, pMGM221, containing the luminescence gene (luxCDABE) of Vibrio harveyi was introduced into different strains (DD136 and 19061) and phases (one and two) of X. nematophilus by triparental mating. For reproducible and efficient conjugation, it was necessary to use older cultures (96-160 h) in the stationary phase of X. nematophilus for mating with relatively small differences (<2-fold) in transconjugant yield for the different strains and phases of X. nematophilus. All transconjugants emitted high levels of light with optimum bioluminescence at 27 degrees C in Luria broth at pH 8.0 containing 20 g/L NaCl; pH, osmolarity, and temperature conditions were similar to those encountered by the bacteria in the hemolymph of the larvae of Galleria mellonella. Plasmids were detected in the transconjugants after 6 months of subculturing the bacteria without antibiotic selection. Aside from light emission, luminescent transconjugants had the same physiological properties as the nonluminescent parental strains, including identical rates of growth, production of exoenzymes, removal from and subsequent emergence into the insect's hemolymph, bacterial-induced hemocyte damage, suppression of prophenoloxidase activation, and the ability to kill G. mellonella larvae. Light-emitting larvae could readily be detected by eye in a dark room, and all bacteria reisolated from dead larvae were luminescent. These properties validate the use of luminescent X. nematophilus not only as a means of following bacterial host interactions, but also as a potential agent to follow the infection and death of the insect population.  相似文献   

10.
Plasmid transfer between Bacillus thuringiensis subsp. kurstaki HD1 and B. thuringiensis subsp. tenebrionis donor strains and a streptomycin-resistant B. thuringiensis subsp. kurstaki recipient was studied under environmentally relevant laboratory conditions in vitro, in soil, and in insects. Plasmid transfer was detected in vitro at temperatures of 5 to 37 degrees C, at pH 5.9 to 9.0, and at water activities of 0.965 to 0.995, and the highest transfer ratios (up to 10(-1) transconjugant/donor) were detected within 4 h. In contrast, no plasmid transfer was detected in nonsterile soil, and rapid formation of spores by the introduced strains probably contributed most to the lack of plasmid transfer observed. When a B. thuringiensis subsp. kurstaki strain was used as the donor strain, plasmid transfer was detected in killed susceptible lepidopteran insect (Lacanobia oleracea) larvae but not in the nonsusceptible coleopteran insect Phaedon chocleriae. When a B. thuringiensis subsp. tenerbrionis strain was used as the donor strain, no plasmid transfer was detected in either of these insects even when they were killed. These results show that in larger susceptible lepidopteran insects there is a greater opportunity for growth of B. thuringiensis strains, and this finding, combined with decreased competition due to a low initial background bacterial population, can provide suitable conditions for efficient plasmid transfer in the environment.  相似文献   

11.
Mobilizations of pBC16 and pAND006, containing the replicon of the Bacillus thuringiensis subsp. israelensis plasmid pTX14-3, between strains of B. thuringiensis subsp. israelensis were examined. Transconjugants appeared after a few minutes and reached a maximum frequency after approximately 2 h. Plasmid pBC16 was mobilized at a frequency approximately 200 times that of pAND006. However, pAND006 was consistently transferred, suggesting that the replicon of pTX14-3 is sufficient to sustain mobilization in B. thuringiensis subsp. israelensis. A specific protease-sensitive coaggregation between strains of B. thuringiensis subsp. israelensis was found to be unambiguously correlated with plasmid transfer. Two aggregation phenotypes, Agr+ and Agr-, were identified in this subspecies. Aggregation disappeared when the optical density of the mating mixture at 600 nm exceeded approximately 1, and it did not reappear upon dilution. Aggregation was shown to involve interactions of cells with opposite aggregation phenotypes, and evidence of a proteinaceous molecule on the surface of the Agr- that is cells involved in aggregation formation is presented. Matings and selection for the presence of two antibiotic resistance plasmids followed by identification of the host cell revealed that mobilization was unidirectional, from the Agr+ cell to the Agr- cell. The aggregation phenotype was found to be transferred with high frequency (approximately 100%) in broth matings, and the appearance of Agr- isolates from Agr+ strains suggested that the loci involved in aggregation formation are located on a plasmid. No excreted aggregation-inducing signals were detected in the supernatant or culture filtrate of either the donor, the recipient, or the mating mixture.  相似文献   

12.
As part of an effort to develop systems for genetic analysis of strains of Bacillus pumilus which are being used as a microbial hay preservative, we introduced the conjugative Enterococcus faecalis transposon Tn916 into B. pumilus ATCC 1 and two naturally occurring hay isolates of B. pumilus. B. pumilus transconjugants resistant to tetracycline were detected at a frequency of approximately 6.5 x 10(-7) per recipient after filter mating with E. faecalis CG110. Southern hybridization confirmed the insertion of Tn916 into several different sites in the B. pumilus chromosome. Transfer of Tn916 also was observed between strains of B. pumilus in filter matings, and one donor strain transferred tetracycline resistance to recipients in broth matings at high frequency (up to 3.4 x 10(-5) per recipient). Transfer from this donor strain in broth matings was DNase-resistant and was not mediated by culture filtrates. Transconjugants from these broth matings contained derivatives of a cryptic plasmid (pMGD302, approx 60 kb) from the donor strain with Tn916 inserted at various sites. The plasmids containing Tn916 insertions transferred to a B. pumilus recipient strain at frequencies of approx 5 x 10(-6) per recipient. This evidence suggests that pMGD302 can transfer by a process resembling conjugation between strains of B. pumilus.  相似文献   

13.
Regulation of protoxin synthesis in Bacillus thuringiensis.   总被引:10,自引:3,他引:7       下载免费PDF全文
A derivative of Bacillus thuringiensis subsp. kurstaki (HD-1) formed parasporal inclusions at 25 degrees C, but not at 32 degrees C. This strain differed from the parent only in the loss of a 110-megadalton (Md) plasmid, but plasmid and chromosomal copies of protoxin genes were present in both strains. On the basis of temperature shift experiments, the sensitive period appeared to be during midexponential growth, long before the time of protoxin synthesis at 3 to 4 h after the end of exponential growth. The conditional phenotype could be transferred by cell mating to naturally acrystalliferous Bacillus cereus. In all such cases, a 29-Md protoxin -encoding plasmid was transferred, but this plasmid alone was barely sufficient for protoxin synthesis. Protoxin production increased to detectable levels, but well below those of the parental donor strain, by simultaneous transfer of a 44-Md protoxin -encoding plasmid. Transfer of a 5-Md plasmid with the two larger protoxin -coding plasmids resulted in a protoxin synthesis level approaching that of the donor strain. A role for some of the cryptic plasmids of kurstaki in parasporal body formation was implied. In contrast, a closely related B. thuringiensis strain, HD73 , produced crystals at both 25 and 32 degrees C even when the capacity was transferred on a 50-Md plasmid to B. cereus. The amount of protoxin produced in these B. cereus transcipients , however, was somewhat less than that produced in the parental strain HD73 , implying that catabolic differences, gene dosage, or the presence of a chromosomal gene (or a combination of these) may be necessary for maximum production. A regulatory component of the 29-Md plasmid appeared to be trans-acting and dominant since B. cereus transcipients containing the 29-Md plasmid from kurstaki and the 50-Md plasmid from HD73 produced more protoxin at 25 degrees C than at 30 degrees C. Similar results were obtained when protoxin synthetic capacity was transferred from B. thuringiensis subsp. israelensis to the conditional B. thuringiensis subsp. kurstaki strain.  相似文献   

14.
A low frequency of chromosomal gene transfer from Bacillus thuringiensis to Bacillus cereus was detected by cell mating, with a tryptophan marker being the most frequently transferred gene among four that were tested. The process was resistant to DNase and was not mediated by cell filtrates. Among several B. thuringiensis subspecies tested, transfer was best with a derivative of B. thuringiensis subsp. kurstaki HD1, which lost several plasmids. All of the B. cereus recombinants contained at least one plasmid from the donor B. thuringiensis; frequently, it was a plasmid that encoded a protoxin gene. In matings with B. thuringiensis subsp. kurstaki HD1, a 29-megadalton plasmid that contained a ca. 2.5-kilobase region of homology with the chromosome was always transferred. No detectable transfer of chromosomal genes was found in B. thuringiensis subsp. kurstaki HD1 strains lacking this plasmid, suggesting that there may be chromosome mobilization.  相似文献   

15.
Transfer of chromosomal genes and plasmids in Bacillus thuringiensis   总被引:1,自引:0,他引:1  
A low frequency of chromosomal gene transfer from Bacillus thuringiensis to Bacillus cereus was detected by cell mating, with a tryptophan marker being the most frequently transferred gene among four that were tested. The process was resistant to DNase and was not mediated by cell filtrates. Among several B. thuringiensis subspecies tested, transfer was best with a derivative of B. thuringiensis subsp. kurstaki HD1, which lost several plasmids. All of the B. cereus recombinants contained at least one plasmid from the donor B. thuringiensis; frequently, it was a plasmid that encoded a protoxin gene. In matings with B. thuringiensis subsp. kurstaki HD1, a 29-megadalton plasmid that contained a ca. 2.5-kilobase region of homology with the chromosome was always transferred. No detectable transfer of chromosomal genes was found in B. thuringiensis subsp. kurstaki HD1 strains lacking this plasmid, suggesting that there may be chromosome mobilization.  相似文献   

16.
Survival and conjugation of Bacillus thuringiensis in a soil microcosm   总被引:3,自引:0,他引:3  
The survival and conjugation ability of sporogenic and asporogenic Bacillus thuringiensis strains were investigated in broth, in non-amended sterile clay soil monoculture and in mixed soil culture. The 75 kb pHT73 plasmid carrying an erythromycin resistance determinant and a cry1Ac gene was transferred in mating broth and soil microcosm. Survival of strains was assessed in soil monoculture and in mixed soil culture for up to 20 days. Sporogenic strains rapidly formed viable spores which were maintained until the end of the experiment. The asporogenic strains were no longer recovered after 8 days of incubation. This study shows that the environmental impact of asporogenic B. thuringiensis strains is lower than that of sporogenic B. thuringiensis strains. Thus, the use of asporogenic strains may significantly reduce any potential risk (gene transfer, soil and plant contamination) due to the dissemination of B. thuringiensis-based biopesticides in the environment.  相似文献   

17.
Integrative plasmids were constructed to enable integration of foreign DNA into the chromosome of Bacillus sphaericus 2297 by in vivo recombination. Integration of the aphA3 kanamycin resistance gene by a two-step procedure demonstrated that this strategy was applicable with antibiotic resistance selection. Hybridization experiments evidenced two copies of the operon encoding the binary toxin from B. sphaericus in the recipient strain. The Bacillus thuringiensis subsp. israelensis cry11Aal gene (referred to as cry11A), encoding a delta-endotoxin with toxicity against Culex, Aedes, and Anopheles larvae, was integrated either by a single crossover event [strain 2297 (::pHT5601), harboring the entire recombinant plasmid] or by two successive crossover events [strain 2297 (::cry11A)]. The level of the Cry11A production in B. sphaericus was high; two crystalline inclusions were produced in strain 2297 (::pHT5601). Synthesis of the Cry11A toxin conferred toxicity to the recombinant strains against Aedes aegypti larvae, for which the parental strain was not toxic. Interestingly, the level of larvicidal activity of strain 2297 (::pHT5601) against Anopheles stephensi was as high as that of B. thuringiensis subsp. israelensis and suggested synergy between the B. thuringiensis and B. sphaericus toxins. The toxicities of parental and recombinant B. sphaericus strains against Culex quinquefasciatus were similar, but the recombinant strains killed the larvae more rapidly. The production of the Cry11A toxin in B. sphaericus also partially restored toxicity for C. quinquefasciatus larvae from a population resistant to B. sphaericus 1593. In vivo recombination therefore appears to be a promising approach to the creation of new B. sphaericus strains for vector control.  相似文献   

18.
In sporulating cultures of Bacillus thuringiensis subsp. yunnanensis HD977, two cell types are observed: cells forming only spores and cells forming only crystals. Curing analysis suggested that the crystal proteins are plasmid encoded. Through plasmid transfer experiments, it was established that a 103-MDa plasmid is involved in the crystal production. Conjugal transfer of this plasmid to Cry- recipient cells of Bacillus thuringiensis subsp. kurstaki HD73-26 conferred the ability to produce crystals exclusively on asporogenous cells of the recipient, indicating that the 103-MDa plasmid mediates the unique regulation of Cry protein production. When the dipteran-specific cryIVB gene was introduced into wild-type (Cry+) and Cry- backgrounds of B. thuringiensis subsp. yunnanensis by phage CP51ts45-mediated transduction, similar to all other B. thuringiensis strains, irregular crystals of CryIVB protein were produced by spore-forming cells in both backgrounds. However, the synthesis of the bipyramidal inclusions of B. thuringiensis subsp. yunnanensis was still limited only to asporogenous cells of the transductant. Thus, it appears that the unique property of exclusive crystal formation in asporogenous cells of B. thuringiensis subsp. yunnanensis is associated with the crystal protein gene(s) per se or its cis acting elements. As the crystals in B. thuringiensis subsp. yunnanensis were formed only in asporogenous cells, attempts were made to find out whether crystal formation had any inhibitory effect on sporulation. It was observed that both Cry+ and Cry- strains of B. thuringiensis subsp. yunnanensis (HD977 and HD977-1, respectively) exhibited comparable sporulation efficiencies. In addition, the Cry- B. thuringiensis subsp. kurstaki host (HD73-26) and its Cry+ transconjugant (HD73-26-16), expressing the B. thuringiensis subsp. yunnanensis crystal protein, were also comparable in their sporulation efficiencies, indicating that production of the crystal proteins of B. thuringiensis subsp. yunnanensis does not affect the process of sporulation.  相似文献   

19.
Eighteen different strains of Edwardsiella ictaluri isolated from infected channel catfish (Ictalurus punctatus) were screened to determine whether plasmid DNA was present. Two plasmids of 5,700 and 4,900 base pairs were identified. Restriction enzyme analysis showed that each of the strains harbored these same two plasmids. Restriction maps of the separated plasmids indicated that these plasmids were not closely related to each other. A rapid screening technique was developed that would allow the presence of these plasmids from either broth cultures or single colonies of E. ictaluri to be determined within 2 to 3 h by agarose gel electrophoresis. These results suggest that plasmid fingerprinting of E. ictaluri should become a useful tool in the presumptive identification of this bacterium from infected channel catfish.  相似文献   

20.
Bacillus thuringiensis screening programs based on the official potency bioassay using third-instar larvae and on a neonate bioassay were developed for Heliothis armigera, Earias insulana, and Spondoptera littoralis. In these bioassays, the diets were standardized to be suitable, with minor modifications, for feeding of the three lepidopterans. The bioassay protocol was based on determination of the LC50 of the microbial standard HD-1-S-80 in the insects susceptible to B. thuringiensis var. kurstaki strains. This was followed by preliminary screening of B. thuringiensis strains at the LC50 of the B. thuringiensis standard. The B. thuringiensis strains causing 100% mortality at this LC50 in the larvae were selected for potency determinations. The neonate bioassay was suitable for accurate determinations of potencies also in S. littoralis--a representative of insects weakly susceptible to the HD-1 standard. The role of the official and the neonate bioassays in developing microbial control programs is discussed.  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