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 46 毫秒
1.
Kasugamycin (KSM), a unique aminoglycoside antibiotic, has been used in agriculture for many years to control not only rice blast caused by the fungus Magnaporthe grisea but also rice bacterial grain and seedling rot or rice bacterial brown stripe caused by Burkholderia glumae or Acidovorax avenae subsp. avenae, respectively. Since both bacterial pathogens are seed-borne and cause serious injury to rice seedlings, the emergence of KSM-resistant B. glumae and A. avenae isolates highlights the urgent need to understand the mechanism of resistance to KSM. Here, we identified a novel gene, aac(2')-IIa, encoding a KSM 2'-N-acetyltransferase from both KSM-resistant pathogens but not from KSM-sensitive bacteria. AAC(2')-IIa inactivates KSM, although it reveals no cross-resistance to other aminoglycosides. The aac(2')-IIa gene from B. glumae strain 5091 was identified within the IncP genomic island inserted into the bacterial chromosome, indicating the acquisition of this gene by horizontal gene transfer. Although excision activity of the IncP island and conjugational gene transfer was not detected under the conditions tested, circular intermediates containing the aac(2')-IIa gene were detected. These results indicate that the aac(2')-IIa gene had been integrated into the IncP island of a donor bacterial species. Molecular detection of the aac(2')-IIa gene could distinguish whether isolates are resistant or susceptible to KSM. This may contribute to the production of uninfected rice seeds and lead to the effective control of these pathogens by KSM.  相似文献   

2.
Aims: This study was carried out to find the prevalence of various plasmid‐mediated quinolone‐resistant (PMQR) determinants among the quinolone‐resistant clinical isolates of Shigella sp. from paediatric patients in Andaman & Nicobar Islands. Methods and Results: A total of 106 quinolone‐resistant Shigella isolates obtained from paediatric patients during hospital‐based surveillance from January 2003 to June 2010 were screened for the presence of various PMQR determinants. Of 106 isolates, 8 (7·5%) showed the presence of aac (6′)‐Ib‐cr and 3 (2·8%) harboured the qnrB genes with 2 (1·9%) of these isolates showing the presence of both. All the 9 isolates had uniform mutations in gyrA (S83L) and in parC (S80I). Conclusions: The prevalence of fluoroquinolone‐acetylating aminoglycoside acetyltransferase {aac (6′)‐Ib‐cr} gene is higher than qnrB gene among the clinical Shigella isolates. These PMQR determinants were detected in the Shigella isolates obtained from 2008–2010, indicating that it happens in a stepwise manner following the multiple mutations in quinolone resistance‐determining regions increase or extend resistance to quinolones or fluoroquinolones. Significance and Impact of Study: The prevalence of these genes are of grave concern as it may be horizontally transferred to other human pathogenic bacteria and can lead to therapeutic failure as a consequence of antimicrobial resistance, not only for the islands but also for the entire south‐east region. The results obtained should encourage further studies on the implications of the presence, distribution, association and variation of these determinants in our quest for understanding PMQR.  相似文献   

3.
Temporal and spatial regulation of genes mediated by tissue‐specific promoters and conditional gene expression systems provide a powerful tool to study gene function in health, disease, and during development. Although transgenic mice expressing the Cre recombinase in the gastric epithelium have been reported, there is a lack of models that allow inducible and reversible gene modification in the stomach. Here, we exploited the gastrointestinal epithelium‐specific expression pattern of the three trefoil factor (Tff) genes and bacterial artificial chromosome transgenesis to generate a novel mouse strain that expresses the CreERT2 recombinase and the reverse tetracycline transactivator (rtTA). The Tg(Tff1‐CreERT2;Tff2‐rtTA;Tff3‐Luc) strain confers tamoxifen‐inducible irreversible somatic recombination and allows simultaneous doxycycline‐dependent reversible gene activation in the gastric epithelium of developing and adult mice. This strain also confers luciferase activity to the intestinal epithelium to enable in vivo bioluminescence imaging. Using fluorescent reporters as conditional alleles, we show Tff1‐CreERT2 and Tff2‐rtTA transgene activity in a partially overlapping subset of long‐term regenerating gastric stem/progenitor cells. Therefore, the Tg(Tff1‐CreERT2;Tff2‐rtTA;Tff3‐Luc) strain can confer intermittent transgene expression to gastric epithelial cells that have undergone previous gene modification, and may be suitable to genetically model therapeutic intervention during development, tumorigenesis, and other genetically tractable diseases. Birth Defects Research (Part A) 106:626–635, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

4.
5.
In this study, a homogenous polysaccharide (FSP), with an average molecular weight of 9.08 × 104 Da, was isolated from Forsythia suspense and its antibacterial potential against Enterobacter cloacae producing SHV‐12 ESBL was investigated. Growth kinetics, in vitro competition and biofilm formation experiments demonstrated that SHV‐12 ESBL contributed to a fitness benefit to E cloacae strain. The antibacterial activity of FSP (2.5, 5.0 and 10.0 μg/mL) was tested against E cloacae bearing SHV‐12 ESBL gene using bacterial sensitivity, agar bioassay and agar well diffusion assays. It was found that the addition of FSP demonstrated potent antibacterial activities against this bacterial as showed by the decrease of bacterial growth and the increase of the inhibition zone diameter. Furthermore, SHV‐12 ESBL gene expression was decreased in E cloacae strain following different FSP treatment in a concentration‐dependent manner. In conclusion, these data showed that FSP exhibited potent good antibacterial activity against E cloacae producing SHV‐12 ESBL via inhibition of SHV‐12 ESBL gene expression, which may promote the development of novel natural antibacterial agents to treat infections caused by this drug‐resistant bacterial pathogen.  相似文献   

6.
Using DNA fingerprinting by pulse‐field gel electrophoresis and repetitive extragenic pallindromic (REP)‐polymerase chain reaction (PCR), two distinct groups were confirmed among 64 Acidovorax avenae subsp. citrulli strains collected from a range of cucurbitaceous hosts in the USA, China, Taiwan, Thailand, Canada, Australia, Brazil and Israel. Eighty‐two percent of the group I strains were recovered from non‐watermelon hosts and the subspecies type strain was the only member of this group that utilized l ‐leucine as a sole carbon source. On the contrary, 94% of the group II strains were recovered from watermelon and 96% of them utilized l ‐leucine. Two‐week‐old watermelon cv. Crimson sweet, cantaloupe cv. Athena, pumpkin cv. Lumina and squash cv. Early yellow crookneck seedlings were susceptible to A. avenae subsp. citrulli strains representing each group with the exception of the subspecies type strain. Overall, seedlings of watermelon cv. Crimson Sweet were most susceptible to A. avenae subsp. citrulli infection followed by cantaloupe, pumpkin and squash. Group II strains were more aggressive watermelon than on other hosts. On the contrary, group I strains were moderately aggressive on all cucurbit hosts tested.  相似文献   

7.
Despite the recent unprecedented development of efficient dopant‐free hole transporting materials (HTMs) for high‐performance perovskite solar cells (PSCs) on small‐area devices (≤0.1 cm2), low‐cost dopant‐free HTMs for large‐area PSCs (≥1 cm2) with high power conversion efficiencies (PCEs) have rarely been reported. Herein, two novel HTMs, 3,3′,6,6′ (or 2,2′,7,7′)‐tetrakis(N,N′‐di‐p‐methoxyphenylamine)‐N,N′‐bicarbazole (3,6 BCz‐OMeTAD or 2,7 BCz‐OMeTAD), are synthesized via an extremely simple route from very cheap raw materials. Owing to their excellent film‐forming abilities and matching energy levels, 3,6 BCz‐OMeTAD and 2,7 BCz‐OMeTAD can be successfully employed as a perfect ultrathin (≈30 nm) hole transporting layer in large‐area PSCs up to 1 cm2. The 3,6 BCz‐OMeTAD and 2,7 BCz‐OMeTAD based large‐area PSCs show highest PCEs up to 17.0% and 17.6%, respectively. More importantly, high performance large‐area PSCs based on 2,7 BCz‐OMeTAD retain 90% of the initial efficiency after 2000 h storage in an ambient environment without encapsulation.  相似文献   

8.
9.
The rdpA gene of strains Delftia acidovorans MC1, Rhodoferax sp. P230, and Sphingobium herbicidovorans MH proved to be identical. However, when RdpA [(R)‐2‐(2,4‐dichlorophenoxy)propionate/α‐ketoglutarate dioxygenase] was investigated after purification from the various strains, significant differences in the kinetics and some chemical properties of the enzymes were observed. The preference for substrates ranged in the order (R)‐2‐(2,4‐dichlorophenoxy)propionate (2,4‐DP) > (R)‐2‐(4‐chloro‐2‐methylphenoxy)propionate (MCPP) >> 2,4‐dichlorophenoxyacetate (2,4‐D) ~ 4‐chloro‐2‐methylphenoxyacetate (MCPA), but detailed kinetic investigations revealed significant strain‐dependent differences in the kcat and KM values. While the KM values of RdpA from the various strains were low and their range rather narrow with 2,4‐DP (19–60 μM) and MCPP (35–64 μM), larger differences were observed with phenoxyacetates which were distinctly higher and spanned a wider range with 2,4‐D (237–935 μM) and MCPA (164–510 μM). The lowest KM values with 2,4‐D and MCPA were found for RdpA originating from strain P230. Investigation of the enzymes from the various sources by 2D gel electrophoresis revealed up to three monomeric enzyme forms which differed in the pI value. The 2D‐patterns were similar with RdpA from strains MC1 and MH, and after heterologous expression of the enzyme in Escherichia coli, but differed significantly from that of strain P230. The presence of enzyme forms and their different composition coincided apparently with the differences observed in the kinetic properties of RdpA in the various strains. The effects are discussed in terms of posttranslational modification of RdpA which appears to be different in extent and kind in the various strains.  相似文献   

10.
We analysed interactions in the system of two Barley Yellow Dwarf Virus (BYDV) strains (MAV and PAV), and wheat (cv. Tinos) as host plant for the virus, and the cereal aphid Sitobion avenae (F.) as vector, in particular whether or not infection by the virus might alter host plant suitability in favour of vector development. By measuring the amino acid and sugar content in the phloem sap of infected and non‐infected wheat plants we found a significant reduction in the concentration of the total amount of amino acids on BYDV‐infected plants. Qualitative and quantitative analysis of honeydew and honeydew excretion indicated a lower efficiency of phloem sap utilisation by S. avenae on infected plants. In addition, S. avenae excreted less honeydew on infected plants. Both BYDV strains significantly affected aphid development by a reduction in the intrinsic rate of natural increase. Hence, infection by the virus reduced the host suitability in terms of aphid population growth potential on BYDV‐infected plants. However, more alate morphs developed on virus‐infected plants. These findings are discussed in relation to the population dynamics of S. avenae, and, as a consequence, the spread of BYDV.  相似文献   

11.
Phenotypic plasticity contributes to the adaptative evolution of populations exposed to new or altered environments. Feeding plasticity is a component of phenotypic plasticity not usually considered in insect strains adapted to insecticide‐altered environments, but which may either accentuate or mitigate insecticide resistance. This is a concern in the pyrethroid‐resistant strains of the maize weevil Sitophilus zeamais Motsch. (Col., Curculionidae), and the reason for this study. A pyrethroid‐susceptible and two pyrethroid‐resistant strains of maize weevil were subjected to free‐choice and no‐choice tests with maize grains sprayed with increasing doses of the pyrethroid, deltamethrin. The insects from the pyrethroid‐resistant strains exhibited higher feeding avoidance with increased deltamethrin doses than insects from the susceptible strain when subjected to free‐choice tests. The strains of maize weevil physiologically resistant to pyrethroids were also behaviourally resistant to deltamethrin – an additional management concern. The resistant strains avoid deltamethrin‐sprayed grains and are less nutritionally affected by this compound, with divergent responses from the susceptible strain with increased doses of deltamethrin. Furthermore, the higher relative growth rate and consequently higher efficiency of food conversion observed in the insecticide‐resistant strains were significant even without insecticide exposure, indicating that these traits are stimulus‐independent and may persist even without further insecticide selection, potentially limiting the options available for their management.  相似文献   

12.
SHV‐12 is the most widespread resistance determinant of Enterobacter cloacae in Taiwan; however, blaSHV‐12 has rarely been mobilized. Six multidrug‐resistant E. cloacae isolates were collected. After conjugal transfer, plasmid profiling and analysis of incompatibility groups was performed to characterize the genetic context of blaSHV‐12‐containing fragments. The presence of mobile genetic elements was demonstrated by PCR, cloning, sequencing and bioinformatics analyses. Four different β‐lactamase genes (blaTEM‐1, blaSHV‐12, blaCTX‐M‐3 and/or blaCTX‐M‐14) were observed in the conjugative plasmids belonging to the IncHI2 (n = 4), IncI1 or IncP incompatibility groups. The IS26‐blaSHV‐12‐IS26 locus was located in five different genetic environments. A novel structural organization of a class 1 integron with the aac(6')‐IIc cassette truncated by IS26 was identified in one isolate. Thus, blaSHV‐12 was obtained from different plasmids through IS26‐mediated homologous recombination. IS26 plays a vital role in the distribution of mobile resistance elements between different plasmids found in multidrug‐resistant E. cloacae isolates.  相似文献   

13.
A synthesis method of novel N‐cycloalkylcarbonyl‐N′‐arylthioureas was developed. It consists of sequential addition of equimolecular amounts of ammonium isothiocyanate and substituted anilines to cycloalkylcarbonyl chlorides. The identity and purity of products were confirmed by LC/MS spectra, their structure by elemental analysis, IR and 1H‐NMR spectra. Preliminary antimicrobial screening for standard microorganisms and molecular docking allowed to select several structures for antifungal and genetic toxicity studies. Conducted in vitro screening of 9 compounds for antifungal potential against 11 phytopathogenic fungi and three Phytophthora strains revealed that two N‐(arylcarbamothioyl) cyclopropanecarboxamides at a concentration of 50 μg/ml exhibited activities comparable to the standard antifungal agent ‘Cyproconazole’. Analysis of mutagenicity of novel thioureas using the Salmonella reverse mutagenicity assay (‘Ames Test’) showed a low gene‐toxicity profile.  相似文献   

14.
Vibrio cholerae non‐O1, non‐O139 (VC_NAG) organisms are universally present in the aquatic environment and regarded as non‐pathogenic bacteria. However, considering that they do occasionally induce gastroenteritis, a study of their virulence and antibiotic resistance genes is important. The presence of enteropathogenic genes, including ctxA, VC_NAG‐specific heat‐stable toxin gene (st), hemolysin (hly), and zona occludens toxin (zot) was determined by PCR in 100 VC_NAG strains isolated in southern Vietnam in 2010–2013 from 94 environmental and six human origins. These 100 VC_NAG strains were also tested phenotypically and genotypically for the presence of the New Delhi metallo‐β‐lactamase (NDM‐1). Of the 100 VC_NAG strains tested, six were positive for ctxA; five from the environment and one of human origin. The st gene was detected in 17 isolates, 15 and two of which were of environmental and human origins, respectively. Gene hly was detected in 19 VC_NAG strains examined, two of which were isolated from humans and 17 from environments. The zot gene was not detected in any of the strains tested. Three VC_NAG strains of environmental origin were confirmed to produce NDM‐1 and the blaNDM‐1 gene was detected in those strains by PCR. Of note, one of the three NDM‐1‐producing VC_NAG strains was confirmed to carry ctxA, st and hly genes concurrently. This is the first report of isolation of NDM‐1‐producing VC_NAG strains in Vietnam.  相似文献   

15.
Six geographic isolates of Heterodera avenae, including two isolates each from Sweden, Australia, and the United States, were compared on the basis of 2-D PAGE protein patterns and the complete DNA sequence for the two internal transcribed ribosomal DNA spacers (rDNA ITS1 and ITS2) and the 5.8S rRNA gene. The protein pattern data and rDNA ITS sequence data both indicated that the Swedish Gotland strain of H. avenae differed markedly from the rest of the isolates. Protein patterns for the Australia isolates differed more from a Swedish strict H. avenae isolate and isolates from Oregon and Idaho, than the two U.S. isolates and the Swedish strict H. avenae isolate differed from each other. Except for the Gotland strain isolate, the rDNA ITS sequences were highly conserved among all of the H. avenae isolates, just as we earlier found them to be conserved among species of the schachtii group of Heterodera.  相似文献   

16.
The aim of this work was to investigate the ability of Acidovorax avenae ssp. citrulli, the causal agent of bacterial fruit blotch of cucurbits (BFB), to colonize female watermelon blossoms, and to explore the relationship between blossom inoculum dosage and seed infestation. Under greenhouse conditions A. avenae ssp. citrulli colonized stigmas and styles of female watermelon blossoms reaching populations of ≈107 to 108 colony‐forming units (CFU) per blossom for 96 h after inoculation. Acidovorax avenae ssp. citrulli growth on stigmas was slower than that of Pseudomonas syringae Cit7, a non‐pathogenic, foliar epiphyte of tomato. While pollination reduced growth of A. avenae ssp. citrulli, but P. syringae Cit7 was unaffected. Both bacteria colonized style tissues but bacterial growth in the style was significantly less than the stigma. Blossom inoculation with ≈1 × 103A. avenae ssp. citrulli CFU/blossom led to 36–55% infested seedlots within symptomless fruits. On average 14% of the seedlings produced from these seedlots displayed BFB symptoms. There was a strong positive correlation between A. avenae ssp. citrulli inoculum concentration applied to blossoms and the percentage of infested seedlots, as determined by the seedling grow‐out assay (R2 = 0.94). However, this relationship was weaker when seedlot infestation was determined by a polymerase chain reaction‐based assay (R2 = 0.34). There was also a strong positive linear relationship between A. avenae ssp. citrulli blossom inoculum dose and the mean percentage of BFB‐infected seedlings (R2 = 0.99) produced in seedling grow‐out assays. These data support the hypothesis that blossom colonization might be involved in seed infestation under field conditions.  相似文献   

17.
18.
19.
20.
The genetic diversity and phylogenetic position of 10 strains of picocyanobacteria from the Arabian Sea were examined using partial sequences from three loci: 16S rDNA, RNA polymerase rpoC1, and two elements of the phycoerythrin (PE) locus, cpeA and cpeB which encode for the α and β subunit of PE. Nine of the strains showed nearly identical spectral phenotypes based on the in vivo excitation spectrum for PE fluorescence emission and appear to be strains synthesizing a phycourobilin (PUB)–lacking PE. These strains include one, Synechococcus sp. G2.1, already known to be closely related to filamentous cyanobacteria and not to the commonly studied 5.1 subcluster of marine Synechococcus. The 10th strain was a PE‐lacking strain that was of interest because it was isolated from open‐ocean conditions where picocyanobacteria with this phenotype are relatively uncommon. Phylogenetic analysis of the concatenated 16S rDNA and rpoC1 data sets showed that none of the previously described strains were members of the 5.1 subcluster of marine Synechococcus, nor were they closely related to strain G2.1. Instead, they form a well‐supported and previously undescribed clade of cyanobacteria that is sister to Cyanobium. Thus, these strains represent the first PE‐containing Cyanobium from oceanic waters, and the lineage they define includes a strain with a PE‐lacking phenotype from the same environment. Analysis of the PE sequence data showed the PE apoprotein has evolved independently in the G2.1 lineage and the Cyanobium‐like lineage represented by the study strains. It also revealed a hypervariable region of the β‐subunit not described previously; variation in this region shows a pattern among a wide range of PE‐containing organisms congruent with the phylogenetic relationships inferred from other genes. This suggests that the PUB‐lacking spectral phenotype is more likely to have evolved in distantly related phylogenetic lineages by either divergent or convergent evolution than by lateral gene transfer. Both the conserved PE gene sequences and the inferred amino acid sequences for the hypervariable region show considerable divergence among Prochlorococcus PEs, red algal PEs, PUB‐containing PEs from the marine Synechococcus 5.1 subcluster, PEs from the Cyanobium‐like strains, and PEs from other cyanobacteria (including strain G2.1). Thus, it appears that the hypervariable region of the PE gene can be used as a taxon‐specific marker.  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