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 390 毫秒
1.
O-Methyltransferase I catalyzes both the conversion of demethylsterigmatocystin to sterigmatocystin and the conversion of dihydrodemethylsterigmatocystin to dihydrosterigmatocystin during aflatoxin biosynthesis. In this study, both genomic cloning and cDNA cloning of the gene encoding O-methyltransferase I were accomplished by using PCR strategies, such as conventional PCR based on the N-terminal amino acid sequence of the purified enzyme, 5' and 3' rapid amplification of cDNA ends PCR, and thermal asymmetric interlaced PCR (TAIL-PCR), and genes were sequenced by using Aspergillus parasiticus NIAH-26. A comparison of the genomic sequences with the cDNA of the dmtA region revealed that the coding region is interrupted by three short introns. The cDNA of the dmtA gene is 1,373 bp long and encodes a 386-amino-acid protein with a deduced molecular weight of 43,023, which is consistent with the molecular weight of the protein determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The C-terminal half of the deduced protein exhibits 76.3% identity with the coding region of the Aspergillus nidulans StcP protein, whereas the N-terminal half of dmtA exhibits 73.0% identity with the 5' flanking region of the stcP gene, suggesting that translation of the stcP gene may start at a site upstream from methionine that is different from the site that has been suggested previously. Also, an examination of the 5' and 3' flanking regions of the dmtA gene in which TAIL-PCR was used demonstrated that the dmtA gene is located in the aflatoxin biosynthesis cluster between (and in the same orientation as) the omtA and ord-2 genes. Northern blotting revealed that expression of the dmtA gene is influenced by both medium composition and culture temperature and that the pattern correlates with the patterns observed for other genes in the aflatoxin gene cluster. Furthermore, Southern blotting and PCR analyses of the dmtA gene showed that a dmtA homolog is present in Aspergillus oryzae SYS-2.  相似文献   

2.
3.
Localized sets of random point mutations generated by PCR amplification can be transferred efficiently to the chromosome of Acinetobacter ADP1 (also known as strain BD413) by natural transformation. The technique does not require cloning of PCR fragments in plasmids: PCR-amplified DNA fragments are internalized by cells and directly incorporated into their genomes by homologous recombination. Previously such procedures for random mutagenesis could be applied only to Acinetobacter genes affording the selection of mutant phenotypes. Here we describe the construction of a vector and recipient that allow for mutagenesis, recovery, and expression of heterologous genes that may lack a positive selection. The plasmid carries an Acinetobacter chromosomal segment interrupted by a multiple cloning site next to a kanamycin resistance marker. The insertion of heterologous DNA into the multiple cloning site prepares the insert as a target for PCR mutagenesis. PCR amplifies the kanamycin resistance marker and a flanking region of Acinetobacter DNA along with the insert of heterologous DNA. Nucleotide sequence identity between the flanking regions and corresponding chromosomal segments in an engineered Acinetobacter recipient allows homologous recombination of the PCR-amplified DNA fragments into a specific chromosomal docking site from which they can be expressed. The recipient strain contains only a portion of the kanamycin resistance gene, so donor DNA containing both this gene and the mutagenized insert can be selected by demanding growth of recombinants in the presence of kanamycin. The effectiveness of the technique was demonstrated with the relatively GC-rich Pseudomonas putida xylE gene. After only one round of PCR amplification (35 cycles), donor DNA produced transformants of which up to 30% carried a defective xylE gene after growth at 37°C. Of recombinant clones that failed to express xylE at 37°C, about 10% expressed the gene when grown at 22°C. The techniques described here could be adapted to prepare colonies with an altered function in any gene for which either a selection or a suitable phenotypic screen exists.  相似文献   

4.
NotI linking clones contain sequences flanking NotI recognition sites and were previously shown to be tightly associated with CpG islands and genes. To directly assess the value of NotI clones in genome research, high density grids with 50 000 NotI linking clones originating from six representative NotI linking libraries were constructed. Altogether, these libraries contained nearly 100 times the total number of NotI sites in the human genome. A total of 3437 sequences flanking NotI sites were generated. Analysis of 3265 unique sequences demonstrated that 51% of the clones displayed significant protein similarity to SWISSPROT and TREMBL database proteins based on MSPcrunch filtering with stringent parameters. Of the 3265 sequences, 1868 (57.2%) were new sequences, not present in the EMBL and EST databases (similarity  90%). Among these new sequences, 795 (24.3%) showed similarity to known proteins and 712 (21.8%) displayed an identity of >75% at the nucleotide level to sequences from EMBL or EST databases. The remaining 361 (11.1%) sequences were completely new, i.e. <75% identical. The work also showed tight, specific association of NotI sites with the first exon and suggest that the so-called 3′ ESTs can actually be generated from 5′-ends of genes that contain NotI sites in their first exon.  相似文献   

5.
G-DNA is a four-stranded DNA structure with diverse putative biological roles. We have previously purified and cloned a novel G-DNA-binding protein TGP1 from the ciliate Tetrahymena thermophila. Here we report the molecular cloning of TGP3, an additional G-DNA-binding protein from the same organism. The TGP3 cDNA encodes a 365 amino acid protein that is homologous to TGP1 (34% identity and 44% similarity). The proteins share a sequence pattern that contains two novel repetitive and homologous motifs flanking an extensively hydrophilic and basic region. A nuclear fractionation experiment showed that TGP1 and TGP3 activities are localized predominantly in the nuclear fraction. To further investigate the biological roles of the proteins in vivo, we have generated separate macronuclear gene knockout (KO) strains (TGP1KO and TGP3KO) for each of the two genes. Southern blot analysis demonstrated that the macronuclear copies of each gene were completely disrupted. Mobility shift assays showed that the corresponding G-DNA-binding activity for each protein was abolished in the KO strains. Growth analysis showed that both KO strains grew at near wild-type rates, indicating that neither of the genes is essential for cell growth. Nevertheless, nuclear staining analysis revealed that both TGP1KO and TGP3KO cells have an increased occurrence (more than 2-fold) of extra micronuclei, implying faulty control of micronuclear division in the KO cells.  相似文献   

6.
7.
The nor-1 gene is involved in aflatoxin biosynthesis in Aspergillus parasiticus and was predicted to encode a norsolorinic acid ketoreductase. Recombinant Nor-1 expressed in Escherichia coli converted the 1′ keto group of norsolorinic acid to the 1′ hydroxyl group of averantin in crude E. coli cell extracts in the presence of NADPH. The results confirm that Nor-1 functions as a ketoreductase in vitro.  相似文献   

8.
Escherichia coli mRNA translation is facilitated by sequences upstream and downstream of the initiation codon, called Shine–Dalgarno (SD) and downstream box (DB) sequences, respectively. In E.coli enhancing the complementarity between the DB sequences and the 16S rRNA penultimate stem resulted in increased protein accumulation without a significant affect on mRNA stability. The objective of this study was to test whether enhancing the complementarity of plastid mRNAs downstream of the AUG (downstream sequence or DS) with the 16S rRNA penultimate stem (anti-DS or ADS region) enhances protein accumulation. The test system was the tobacco plastid rRNA operon promoter fused with the E.coli phage T7 gene 10 (T7g10) 5′-untranslated region (5′-UTR) and DB region. Translation efficiency was tested by measuring neomycin phosphotransferase (NPTII) accumulation in tobacco chloroplasts. We report here that the phage T7g10 5′-UTR and DB region promotes accumulation of NPTII up to ~16% of total soluble leaf protein (TSP). Enhanced mRNA stability and an improved NPTII yield (~23% of TSP) was obtained from a construct in which the T7g10 5′-UTR was linked with the NPTII coding region via a NheI site. However, replacing the T7g10 DB region with the plastid DS sequence reduced NPTII and mRNA levels to 0.16 and 28%, respectively. Reduced NPTII accumulation is in part due to accelerated mRNA turnover.  相似文献   

9.
Certain genes exhibit notable diversity in their expression patterns both within and between species. One such gene is the vasopressin receptor 1a gene (Avpr1a), which exhibits striking differences in neural expression patterns that are responsible for mediating differences in vasopressin-mediated social behaviors. The genomic mechanisms that contribute to these remarkable differences in expression are not well understood. Previous work has suggested that both the proximal 5′ flanking region and a polymorphic microsatellite element within that region of the vole Avpr1a gene are associated with variation in V1a receptor (V1aR) distribution and behavior, but neither has been causally linked. Using homologous recombination in mice, we reveal the modest contribution of proximal 5′ flanking sequences to species differences in V1aR distribution, and confirm that variation in V1aR distribution impacts stress-coping in the forced swim test. We also demonstrate that the vole Avpr1a microsatellite structure contributes to Avpr1a expression in the amygdala, thalamus, and hippocampus, mirroring a subset of the inter- and intra-species differences observed in central V1aR patterns in voles. This is the first direct evidence that polymorphic microsatellite elements near behaviorally relevant genes can contribute to diversity in brain gene expression profiles, providing a mechanism for generating behavioral diversity both at the individual and species level. However, our results suggest that many features of species-specific expression patterns are mediated by elements outside of the immediate 5′ flanking region of the gene.  相似文献   

10.
In the aflatoxin biosynthetic pathway, 5′-oxoaverantin (OAVN) cyclase, the cytosolic enzyme, catalyzes the reaction from OAVN to (2′S,5′S)-averufin (AVR) (E. Sakuno, K. Yabe, and H. Nakajima, Appl. Environ. Microbiol. 69:6418-6426, 2003). Interestingly, the N-terminal 25-amino-acid sequence of OAVN cyclase completely matched an internal sequence of the versiconal (VHOH) cyclase that was deduced from its gene (vbs). The purified OAVN cyclase also catalyzed the reaction from VHOH to versicolorin B (VB). In a competition experiment using the cytosol fraction of Aspergillus parasiticus, a high concentration of VHOH inhibited the enzyme reaction from OAVN to AVR, and instead VB was newly formed. The recombinant Vbs protein, which was expressed in Pichia pastoris, showed OAVN cyclase activity, as well as VHOH cyclase activity. A mutant of A. parasiticus SYS-4 (= NRRL 2999) with vbs deleted accumulated large amounts of OAVN, 5′-hydroxyaverantin, averantin, AVR, and averufanin in the mycelium. These results indicated that the cyclase encoded by the vbs gene is also involved in the reaction from OAVN to AVR in aflatoxin biosynthesis. Small amounts of VHOH, VB, and aflatoxins also accumulated in the same mutant, and this accumulation may have been due to an unknown enzyme(s) not involved in aflatoxin biosynthesis. This is the first report of one enzyme catalyzing two different reactions in a pathway of secondary metabolism.  相似文献   

11.
Homologous recombination in yeast can be exploited to reliably generate libraries of >107 transformants from a pool of PCR products and a linearized plasmid vector. Homology in the PCR insertion products drives shuffling of these genes in vivo by yeast homologous recombination. Two scFvs that share 89.8% homology were shuffled in vivo by homologous recombination, and chimeric genes were generated regardless of whether or not one of the scFv PCR products lacked 5′ homology to the cut vector and the second scFv PCR product lacked 3′ homology to the cut vector, or both PCR products had both 5′ and 3′ homology to the cut vector. A majority of the chimeras had single crossovers; however, double and triple crossovers were isolated. Crossover points were evenly distributed in the hybrids created and homology of as little as two nucleotides was able to produce a chimeric clone. The numbers of clones isolated with a given number of crossovers was approximated well by a Poisson distribution. Transformation efficiencies for the chimeric libraries were of the order of 104–105 transformants per microgram of insert, which is the same order of magnitude as when a single PCR product is inserted alone into the display vector by homologous recombination. This method eliminates ligation and Escherichia coli transformation steps of previous methods for generating yeast-displayed libraries, requires fewer PCR cycles than in vitro DNA shuffling and, unlike site-specific recombination methods, allows for recombination anywhere that homology exists between the genes to be recombined. This simple technique should prove useful for protein engineering in general and antibody engineering, specifically in yeast.  相似文献   

12.
Different strains of Streptococcus suis serotypes 1 and 2 isolated from pigs either contained a restriction-modification (R-M) system or lacked it. The R-M system was an isoschizomer of Streptococcus pneumoniae DpnII, which recognizes nucleotide sequence 5′-GATC-3′. The nucleotide sequencing of the genes encoding the R-M system in S. suis DAT1, designated SsuDAT1I, showed that the SsuDAT1I gene region contained two methyltransferase genes, designated ssuMA and ssuMB, as does the DpnII system. The deduced amino acid sequences of M.SsuMA and M.SsuMB showed 70 and 90% identity to M.DpnII and M.DpnA, respectively. However, the SsuDAT1I system contained two isoschizomeric restriction endonuclease genes, designated ssuRA and ssuRB. The deduced amino acid sequence of R.SsuRA was 49% identical to that of R.DpnII, and R.SsuRB was 72% identical to R.LlaDCHI of Lactococcus lactis subsp. cremoris DCH-4. The four SsuDAT1I genes overlapped and were bounded by purine biosynthetic gene clusters in the following gene order: purF-purM-purN-purH-ssuMA-ssuMB-ssuRA-ssuRB-purD-purE. The G+C content of the SsuDAT1I gene region (34.1%) was lower than that of the pur region (48.9%), suggesting horizontal transfer of the SsuDAT1I system. No transposable element or long-repeat sequence was found in the flanking regions. The SsuDAT1I genes were functional by themselves, as they were individually expressed in Escherichia coli. Comparison of the sequences between strains with and without the R-M system showed that only the region from 53 bp upstream of ssuMA to 5 bp downstream of ssuRB was inserted in the intergenic sequence between purH and purD and that the insertion target site was not the recognition site of SsuDAT1I. No notable substitutions or insertions could be found, and the structures were conserved among all the strains. These results suggest that the SsuDAT1I system could have been integrated into the S. suis chromosome by an illegitimate recombination mechanism.  相似文献   

13.
The pathway oxoaverantin (OAVN) → averufin (AVR) → hydroxyversicolorone (HVN) → versiconal hemiacetal acetate (VHA) is involved in aflatoxin biosynthesis, and the cypX and moxY genes, which are present in the aflatoxin gene cluster, have been previously suggested to be involved in this pathway. To clarify the function of these two genes in more detail, we disrupted the genes in aflatoxigenic Aspergillus parasiticus NRRL 2999. The cypX-deleted mutant lost aflatoxin productivity and accumulated AVR in the mycelia. Although this mutant converted HVN, versicolorone (VONE), VHA, and versiconol acetate (VOAc) to aflatoxins in feeding experiments, it could not produce aflatoxins from either OAVN or AVR. The moxY-deleted mutant also lost aflatoxin productivity, whereas it newly accumulated HVN and VONE. In feeding experiments, this mutant converted either VHA or VOAc to aflatoxins but did not convert OAVN, AVR, HVN, or VONE to aflatoxins. These results demonstrated that cypX encodes AVR monooxygenase, catalyzing the reaction from AVR to HVN, and moxY encodes HVN monooxygenase, catalyzing a Baeyer-Villiger reaction from HVN to VHA as well as from VONE to VOAc. In this work, we devised a simple and rapid method to extract DNA from many fungi for PCR analyses in which cell disruption with a shaker and phenol extraction were combined.  相似文献   

14.
During aflatoxin biosynthesis, 5′-hydroxyaverantin (HAVN) is converted to averufin (AVR). Although we had previously suggested that this occurs in one enzymatic step, we demonstrate here that this conversion is composed of two enzymatic steps by showing that the two enzyme activities in the cytosol fraction of Aspergillus parasiticus were clearly separated by Mono Q column chromatography. An enzyme, HAVN dehydrogenase, catalyzes the first reaction from HAVN to a novel intermediate, another new enzyme catalyzes the next reaction from the intermediate to AVR, and the intermediate is a novel substance, 5′-oxoaverantin (OAVN), which was determined by physicochemical methods. We also purified both of the enzymes, HAVN dehydrogenase and OAVN cyclase, from the cytosol fraction of A. parasiticus by using ammonium sulfate fractionation and successive chromatographic steps. The HAVN dehydrogenase is a homodimer composed of 28-kDa subunits, and it requires NAD, but not NADP, as a cofactor for its activity. Matrix-assisted laser desorption ionization-time of flight mass spectrometry analysis of tryptic peptides of the purified HAVN dehydrogenase revealed that this enzyme coincides with a protein deduced from the adhA gene in the aflatoxin gene cluster of A. parasiticus. Also, the OAVN cyclase enzyme is a homodimer composed of 79-kDa subunits which does not require any cofactor for its activity. Further characterizations of both enzymes were performed.  相似文献   

15.
Soil isolates of Aspergillus flavus from a transect extending from eastern New Mexico through Georgia to eastern Virginia were examined for production of aflatoxin B1 and cyclopiazonic acid in a liquid medium. Peanut fields from major peanut-growing regions (western Texas; central Texas; Georgia and Alabama; and Virginia and North Carolina) were sampled, and fields with other crops were sampled in regions where peanuts are not commonly grown. The A. flavus isolates were identified as members of either the L strain (n = 774), which produces sclerotia that are >400 μm in diameter, or the S strain (n = 309), which produces numerous small sclerotia that are <400 μm in diameter. The S-strain isolates generally produced high levels of aflatoxin B1, whereas the L-strain isolates were more variable in aflatoxin production; variation in cyclopiazonic acid production also was greater in the L strain than in the S strain. There was a positive correlation between aflatoxin B1 production and cyclopiazonic acid production in both strains, although 12% of the L-strain isolates produced only cyclopiazonic acid. Significant differences in production of aflatoxin B1 and cyclopiazonic acid by the L-strain isolates were detected among regions. In the western half of Texas and the peanut-growing region of Georgia and Alabama, 62 to 94% of the isolates produced >10 μg of aflatoxin B1 per ml. The percentages of isolates producing >10 μg of aflatoxin B1 per ml ranged from 0 to 52% in the remaining regions of the transect; other isolates were often nonaflatoxigenic. A total of 53 of the 126 L-strain isolates that did not produce aflatoxin B1 or cyclopiazonic acid were placed in 17 vegetative compatibility groups. Several of these groups contained isolates from widely separated regions of the transect.  相似文献   

16.
Aflatoxin, a mycotoxin synthesized by Aspergillus spp., is among the most potent naturally occurring carcinogens known. Little is known about the subcellular organization of aflatoxin synthesis. Previously, we used transmission electron microscopy and immunogold labeling to demonstrate that the late aflatoxin enzyme OmtA localizes primarily to vacuoles in fungal cells on the substrate surface of colonies. In the present work, we monitored subcellular localization of Ver-1 in real time in living cells. Aspergillus parasiticus strain CS10-N2 was transformed with plasmid constructs that express enhanced green fluorescent protein (EGFP) fused to Ver-1. Analysis of transformants demonstrated that EGFP fused to Ver-1 at either the N or C terminus functionally complemented nonfunctional Ver-1 in recipient cells. Western blot analysis detected predominantly full-length Ver-1 fusion proteins in transformants. Confocal laser scanning microscopy demonstrated that Ver-1 fusion proteins localized in the cytoplasm and in the lumen of up to 80% of the vacuoles in hyphae grown for 48 h on solid media. Control EGFP (no Ver-1) expressed in transformants was observed in only 13% of the vacuoles at this time. These data support a model in which middle and late aflatoxin enzymes are synthesized in the cytoplasm and transported to vacuoles, where they participate in aflatoxin synthesis.  相似文献   

17.
18.
19.
Cyclopiazonic acid (CPA), an indole-tetramic acid mycotoxin, is produced by many species of Aspergillus and Penicillium. In addition to CPA Aspergillus flavus produces polyketide-derived carcinogenic aflatoxins. Aflatoxin biosynthesis genes form a gene cluster in a subtelomeric region. Isolates of A. flavus lacking aflatoxin production due to the loss of the entire aflatoxin gene cluster and portions of the subtelomeric region are often unable to produce CPA, which suggests a physical link of genes involved in CPA biosynthesis to the aflatoxin gene cluster. Examining the subtelomeric region in A. flavus isolates of different chemotypes revealed a region possibly associated with CPA production. Disruption of three of the four genes present in this region predicted to encode a monoamine oxidase, a dimethylallyl tryptophan synthase, and a hybrid polyketide non-ribosomal peptide synthase abolished CPA production in an aflatoxigenic A. flavus strain. Therefore, some of the CPA biosynthesis genes are organized in a mini-gene cluster that is next to the aflatoxin gene cluster in A. flavus.  相似文献   

20.
A new ketoreductase useful for asymmetric synthesis of chiral alcohols was identified in the cyanobacterium Synechococcus sp. strain PCC 7942. Mass spectrometry of trypsin-digested peptides identified the protein as 3-ketoacyl-[acyl-carrier-protein] reductase (KR) (EC 1.1.1.100). The gene, referred to as fabG, was cloned, functionally expressed in Escherichia coli, and subsequently purified to homogeneity. The enzyme displayed a temperature optimum at 44°C and a broad pH optimum between pH 7 and pH 9. The NADPH-dependent KR was able to asymmetrically reduce a variety of prochiral ketones with good to excellent enantioselectivities (>99.8%). The KR showed particular high specific activity for asymmetric reduction of ethyl 4-chloroacetoacetate (38.29 ± 2.15 U mg−1) and 2′,3′,4′,5′,6′-pentafluoroacetophenone (8.57 ± 0.49 U mg−1) to the corresponding (S)-alcohols. In comparison with an established industrial enzyme like the alcohol dehydrogenase from Lactobacillus brevis, the KR showed seven-times-higher activity toward 2′,3′,4′,5′,6′-pentafluoroacetophenone, with a remarkably higher enantiomeric excess (>99.8% [S] versus 43.3% [S]).  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