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 31 毫秒
1.
Coveny AM  Dray T  Gloor GB 《Genetics》2002,161(1):249-258
We examined the influence that heterologous sequences of different sizes have on the frequency of double-strand-break repair by gene conversion in Drosophila melanogaster. We induced a double-strand break on one X chromosome in female flies by P-element excision. These flies contained heterologous insertions of various sizes located 238 bp from the break site in cis or in trans to the break, or both. We observed a significant decrease in double-strand-break repair with large heterologous insertions located either in cis or in trans to the break. Reestablishing the homology by including the same heterologous sequence in cis and in trans to the double-strand break restored the frequency of gene conversion to wild-type levels. In one instance, an allelic nonhomologous insertion completely abolished repair by homologous recombination. The results show that the repair of a double-strand break by gene conversion requires chromosome pairing in the local region of the double-strand break.  相似文献   

2.
We have used a differential cloning approach to isolate ribosomal/non-ribosomal frontier sequences from Xenopus laevis. A ribosomal intergenic spacer sequence (IGS) was cloned and shown not to be physically linked with the ribosomal locus. This ribosomal orphon contained the IGS sequences found immediately downstream of the 28S gene and included an array of enhancer repetitions and a non-functional spacer promoter. The orphon sequence was flanked by a member of the novel 'Frt' low copy repetitive element family. Three individual Frt repeats were sequenced and all members of this family were shown to lie clustered at two chromosomal sites, one of which contained the ribosomal orphon. One of the Frt elements contained an insertion of 297 bp that showed extensive homology to sequences within at least three other Xenopus genes. Each homology region was flanked by members of the T2 family of short interspersed repetitive elements, (SINEs), and by its target insertion sequence, suggesting multiple translocation events. The data are discussed in terms of the evolution of the ribosomal gene locus.  相似文献   

3.
Double-strand breaks can initiate meiotic recombination in S. cerevisiae   总被引:29,自引:0,他引:29  
A L Kolodkin  A J Klar  F W Stahl 《Cell》1986,46(5):733-740
We investigated the effects of double-strand breaks on meiotic recombination in yeast. A double-strand break was introduced at the MATa locus by sporulation of a MAT alpha inc/MATa diploid under inducing conditions for the HO-encoded endonuclease; 14% of the resulting tetrads had undergone 4 alpha:0a conversion. Conversion at MAT was associated with co-conversion of a closely linked marker and an increased recombination frequency for flanking markers. We also studied the sporulation products of a diploid heterozygous at the HIS4 locus for an insertion of a 100 bp fragment of MATa containing the HO endonuclease cut site. Under inducing conditions, a significant number of tetrads were formed that had undergone gene conversions in favor of the HIS4+ allele. Although double-strand breaks can initiate meiotic recombination in yeast, the data suggest that they do not normally do so.  相似文献   

4.
Gene targeting has been used to direct mutations into specific chromosomal loci in murine embryonic stem (ES) cells. The altered locus can be studied in vivo with chimeras and, if the mutated cells contribute to the germ line, in their offspring. Although homologous recombination is the basis for the widely used gene targeting techniques, to date, the mechanism of homologous recombination between a vector and the chromosomal target in mammalian cells is essentially unknown. Here we look at the nature of gene targeting in ES cells by comparing an insertion vector with replacement vectors that target hprt. We found that the insertion vector targeted up to ninefold more frequently than a replacement vector with the same length of homologous sequence. We also observed that the majority of clones targeted with replacement vectors did not recombine as predicted. Analysis of the recombinant structures showed that the external heterologous sequences were often incorporated into the target locus. This observation can be explained by either single reciprocal recombination (vector insertion) of a recircularized vector or double reciprocal recombination/gene conversion (gene replacement) of a vector concatemer. Thus, single reciprocal recombination of an insertion vector occurs 92-fold more frequently than double reciprocal recombination of a replacement vector with crossover junctions on both the long and short arms.  相似文献   

5.
Plasmids containing heteroallelic copies of the Saccharomyces cerevisiae HIS3 gene undergo intramolecular gene conversion in mitotically dividing S. cerevisiae cells. We have used this plasmid system to determine the minimum amount of homology required for gene conversion, to examine how conversion tract lengths are affected by limited homology, and to analyze the role of flanking DNA sequences on the pattern of exchange. Plasmids with homologous sequences greater than 2 kilobases have mitotic exchange rates as high as 2 x 10(-3) events per cell per generation. As the homology is reduced, the exchange rate decreases dramatically. A plasmid with 26 base pairs (bp) of homology undergoes gene conversion at a rate of approximately 1 x 10(-10) events per cell per generation. These studies have also shown that an 8-bp insertion mutation 13 bp from a border between homologous and nonhomologous sequences undergoes conversion, but that a similar 8-bp insertion 5 bp from a border does not. Examination of independent conversion events which occurred in plasmids with heteroallelic copies of the HIS3 gene shows that markers within 280 bp of a border between homologous and nonhomologous sequences undergo conversion less frequently than the same markers within a more extensive homologous sequence. Thus, proximity to a border between homologous and nonhomologous sequences shortens the conversion tract length.  相似文献   

6.
We have used DNA-mediated gene transfer to study homologous recombination in cultured mammalian cells. A family of plasmids with insertion and deletion mutations in the coding region of the herpes simplex type 1 thymidine kinase (tk) gene served as substrates for DNA-mediated gene transfer into mouse Ltk- cells by the calcium phosphate technique. Intermolecular recombination events were scored by the number of colonies in hypoxanthine-aminopterin-thymidine selective medium. We used supercoiled plasmids containing tk gene fragments to demonstrate that an overlap of 62 base pairs (bp) of homologous DNA was sufficient for intermolecular recombination. Addition of 598 bp of flanking homology separated from the region of recombination by a double-strand gap, deletion, or insertion of heterologous DNA increased the frequency of recombination by 300-, 20-, or 40-fold, respectively. Linearizing one of the mutant plasmids in a pair before cotransfer by cutting in the area of homology flanking a deletion of 104 bp or an insertion of less than 24 bp increased the frequency of recombination relative to that with uncut plasmids. However, cutting an insertion mutant of greater than or equal to 24 bp in the same manner did not increase the frequency. We show how our data are consistent with models that postulate at least two phases in the recombination process: homologous pairing and heteroduplex formation.  相似文献   

7.
Mutations were targeted to the Hprt locus of mouse embryo-derived stem cells by using 22 different sequence replacement and sequence insertion vectors. The targeting frequency was examined at two sites within the Hprt locus as a function of the extent of homology between the targeting vector and the target locus. The targeting frequency was also compared by using vectors prepared from isogenic and nonisogenic DNA sources. With one exception, all of the vectors showed the same exponential dependence of targeting efficiency on the extent of homology between the targeting vector and the target locus. This was true regardless of whether they were sequence replacement or sequence insertion vectors, whether they were directed toward either of the two different sites within the Hprt locus, or whether they were prepared from isogenic or nonisogenic DNA sources. Vectors prepared from isogenic DNA targeted four to five times more efficiently than did the corresponding vectors prepared from nonisogenic DNA. The single case of unexpectedly low targeting efficiency involved one of the vectors prepared from nonisogenic DNA and could be attributed to an unfavorable distribution of heterology between the Hprt sequences present in the targeting vector and the endogenous Hprt gene.  相似文献   

8.
The effect of the structure of donor DNA molecules on the initiation of recombination for double strand break repair in human nuclear extracts, was investigated here. A unique double strand break was introduced into M13 duplex derivatives by digestion with restriction enzymes. After coincubation of the cleaved DNA in human nuclear extracts, with a plasmid containing M13 sequences spanning the break, double strand break repair was estimated by the plating efficiency in JM109 (RecA1) bacteria. We first confirm that a short heterologous insert (8bp) close to the break on the recipient cleaved M13 DNA inhibits recombination with circular as well as with linear donor molecules. The results indicate that, with these substrates, recombination is initiated at the level of the break, requires uninterrupted homology on both sides of the break, and is associated with a decreasing gradient of gene conversion. When the heterologous insertion is located on the plasmid donor DNA, similar results are obtained with a circular donor DNA. In contrast, with a linear donor molecule, bearing the insert, homology requirements, in the region of the break in M13 DNA, are abolished. This last result suggests that recombination could be initiated at the extremities of the linear donor DNA.  相似文献   

9.
目的研究不同波长光照对草履虫增殖的影响,克隆草履虫感光蛋白基因.方法以双小核草履虫 Paramecium aureli为研究对象,分别置于黄色光(578~592 nm)、蓝色光(446~464 nm)、红色光(620~760 nm)、白光和自然光下,每隔1 h随机抽样法显微观察并计数;RT-PCR克隆草履虫感光蛋白基因.结果不同波长光的照射下,与自然光比较,第1天黄光组草履虫增殖显著上升(P<0.01),蓝光组和红光组草履虫增殖受到不同程度的抑制(P<0.01, P<0.05),白光组无明显差异(P>0.05);第2天,蓝光组、红光组和白光组草履虫增殖仍受到抑制(P<0.01),黄光组作用不显著(P>0.05).黄光组和白光组草履虫总RNA作为模板,克隆出大小约500 bp的rhodopsin-like基因cDNA片段, 5个不同光照组均克隆出大小约195 bp的Long wave sensitive opsin-like 基因cDNA片段.结论黄色光显著地促进草履虫增殖,蓝色光和红色光抑制其增殖;黄光和白光能诱导草履虫rhodopsin-like基因表达;Long wave sensitive opsin-like 基因在草履虫有表达.  相似文献   

10.
Using simple linear fragments of the Chinese hamster adenine phosphoribosyltransferase (APRT) gene as targeting vectors, we have investigated the homology dependence of targeted recombination at the endogenous APRT locus in Chinese hamster ovary (CHO) cells. We have examined the effects of varying either the overall length of targeting sequence homology or the length of 5' or 3' flanking homology on both the frequency of targeted homologous recombination and the types of recombination events that are obtained. We find an exponential (logarithmic) relationship between length of APRT targeting homology and the frequency of targeted recombination at the CHO APRT locus, with the frequency of targeted recombination dependent upon both the overall length of targeting homology and the length of homology flanking each side of the target gene deletion. Although most of the APRT+ recombinants analyzed reflect simple targeted replacement or conversion of the target gene deletion, a significant fraction appear to have arisen by target gene-templated extension and correction of the targeting fragment sequences. APRT fragments with limited targeting homology flanking one side of the target gene deletion yield proportionately fewer target gene conversion events and proportionately more templated extension and vector correction events than do fragments with more substantial flanking homology.  相似文献   

11.
P element-induced gene conversion has been previously used to modify the white gene of Drosophila melanogaster in a directed fashion. The applicability of this approach of gene targeting in Drosophila melanogaster, however, has not been analyzed quantitatively for other genes. We took advantage of the P element-induced forked allele, f(hd), which was used as a target, and we constructed a vector containing a modified forked fragment for converting f(hd). Conversion frequencies were analyzed for this locus as well as for an alternative white allele, w(eh812). Combination of both P element-induced mutant genes allowed the simultaneous analysis of conversion frequencies under identical genetic, developmental, and environmental conditions. This paper demonstrates that gene conversion through P element-induced gap repair can be applied with similar success rates at the forked locus and in the white gene. The average conversion frequency at forked was 0.29%, and that at white was 0.17%. These frequencies indicate that in vivo gene targeting in Drosophila melanogaster should be applicable for other genes in this species at manageable rates. We also confirmed the homolog dependence of reversions at the forked locus, indicating that P elements transpose via a cut-and-paste mechanism. In a different experiment, we attempted conversion with a modified forked allele containing the su(Hw) binding site. Despite an increased sample size, there were no conversion events with this template. One interpretation (under investigation) is that the binding of the su(Hw) product prevents double-strand break repair.  相似文献   

12.
13.
Homologous recombination has been used to introduce site-specific mutations into murine embryonic stem (ES) cells with both insertion and replacement vectors. In this study, we compared the frequency of gene targeting with various lengths of homology and found a dramatic increase in targeting with an increase in homology from 1.3 to 6.8 kb. We examined in detail the relationship between the length of homology and the gene-targeting frequency for replacement vectors and found that a critical length of homology is needed for targeting. Adding greater lengths of homology to this critical length has less of an effect on the targeting frequency. We also analyzed the lengths of homology necessary on both arms of the vector for gene replacement events and found that 472 bp of homology is used as efficiently as 1.2 kb in the formation and resolution of crossover junctions.  相似文献   

14.
The involvement of a double strand break in the initiation of homologous recombination was examined in human nuclear extracts. M13 duplex derivatives, containing inserts in the LacZ' region (producing white plaques), were cleaved by restriction enzymes and coincubated in the extracts with a circular plasmid containing the LacZ' region without insert, and unable to produce plaques. Repair was estimated by the ability to produce plaques after transfection into JM109 (recA1) bacteria. Recombination with the plasmid enhances the number of plaques and also the frequency of M13 producing blue plaques. Heterologous insertions in the region surrounding the break were analyzed for their effects on initiation of recombination. The extent of repair by recombination (number of plaques) was compared with the number of blue plaques among the repaired population. Initiation of recombination is inhibited when heterologous insertions are located at 7bp from the break, on the right side as well as on the left side. A low level of recombination is measurable for 27 bp of homology but the maximum efficiency of recombination occurred with homologies of 165 or 320 bp from the break to the heterologous insertion. At 320 bp, the extent of recombinational repair remained at a plateau level but the frequency of blue plaques progressively decreases. We have also analyzed the effect of different sizes of inserts. With longer inserts, a longer length of homology adjacent to the break is required for optimum recombination. However, the size of the insert does not affect the low level of recombination that occurred with a short homology (27 bp). The results indicate that the process is initiated at or near the break, requires homology on both sides of the break and is followed by an elongation from the double strand break to the distal regions of the DNA. Our data provide some support to the double-strand-break repair model established for meiotic recombination in yeast.  相似文献   

15.
We have examined the effects of placing nonhomologous DNA on the ends of an insertion-type gene targeting vector. The presence of terminal heterologies was found to be compatible with insertion targeting, and the terminal heterologies were efficiently removed. Terminal heterologies reduced the frequency of gene targeting to variable extents. The degree of inhibition of targeting was dependent on the length and the position of the heterology: 2.1kb heterologous sequences were more inhibitory than shorter regions of heterology, and heterology placed on the end of the long (4.8kb) arm of homology was more inhibitory than heterology positioned on the end of the short (0.8kb) arm. When heterology was placed on both arms of the targeting vector the targeting efficiencies were similar to or higher than when heterology was present on the long arm only. These results suggest that terminal sequences are removed simultaneously from both ends of targeting vectors. The removal of terminal sequences probably occurs by exonucleolytic degradation of both strands at each end, and removal of at least one of the strands is intimately coupled with the process of homologous recombination. These findings have implications for the design of gene targeting vectors.  相似文献   

16.
We have cloned a full-length cDNA for testicular 20β-HSD in yellow catfish. The validated 20β-HSD cDNA full-length sequence, 1141 bp in length, contained a 108 bp 5'-untranslated region (UTR), a 202 bp 3'-UTR with an AATAAAA frame, and an 831 bp open reading frame (ORF) which encoded a propeptide of 277 amino acid residues. The enzyme shows the highest structural homology with that of zebrafish, and rainbow trout. Quantitative real-time PCR revealed that 20β-HSD has widespread tissue distribution, with expression being abundant in tissues with high metabolic rates like gonads, liver, intestine, stomach and gill. In vivo experiments showed that expression level was highest at testicular mature stage indicating that 20β-HSD could play an important role in testicular developmental maturation in yellow catfish. During testicular mature stage, 20β-HSD related metabolism was regulated by GnRH and LH. Moreover, structural analysis showed that the predicted 20β-HSD contained 7 functional motifs of SDR superfamily of enzymes, including the putative coenzyme binding domain (Rossmann fold), GlyXXXGlyIleuGly, and the region responsible for nucleophilic attack of the substrate pocket, TyrXXXLys. These motifs are strictly conserved in yellow catfish 20β-HSD. Comprehensive functional analysis revealed that this enzyme has multiple functions, such as xenobiotic metabolism, and steroid conversion. Catfish 20β-HSD contains multiple potential post-translational modification sites. Its subcellular location, theoretical isoelectric point and molecular weight were also investigated. Furthermore, we constructed its phylogenetic tree and secondary structure. All results provided basic information for further studies of its structure, functions and properties.  相似文献   

17.
目的:用缺口修复等技术构建Myostatin(肌肉生长抑制素,MSTN)基因打靶载体,并对大白猪胎儿成纤维体细胞进行转染,获得基因敲除细胞。方法与结果:首先构建用于MSTN基因同源长臂(LA)的抓捕载体,然后在大肠杆菌内利用Red同源重组系统介导的缺口修复,从含大白猪MSTN基因座的细菌人工染色体上亚克隆9.9 kb的LA到抓捕载体上,经过部分序列测定,同源性为100%;通过PCR获得1.4 kb的同源短臂(SA);将LA和SA连入载体pLOXP,构建含有neo和tk正负筛选标记基因的MSTN基因打靶载体pLOXP-MSTN-KO;将线性化的pLOXP-MSTN-KO通过电转染整合到大白猪胎儿成纤维细胞基因组中,利用G418和丙氧鸟苷进行药物筛选,获得抗性细胞克隆890个,通过PCR和DNA测序鉴定获得基因敲除的细胞克隆4个。结论:构建了有效的MSTN基因打靶载体,通过转染获得基因敲除细胞,为利用体细胞核移植制备MSTN基因敲除猪奠定了基础。  相似文献   

18.
Double-strand gap repair in a mammalian gene targeting reaction.   总被引:16,自引:9,他引:16       下载免费PDF全文
To better understand the mechanism of homologous recombination in mammalian cells that facilitates gene targeting, we have analyzed the recombination reaction that inserts a plasmid into a homologous chromosomal locus in mouse embryonic stem cells. A partially deleted HPRT gene was targeted with various plasmids capable of correcting the mutation at this locus, and HPRT+ recombinants were directly selected in HAT medium. The structures of the recombinant loci were then determined by genomic Southern blot hybridizations. We demonstrate that plasmid gaps of 200, 600, and 2,500 bp are efficiently repaired during the integrative recombination reaction. Targeting plasmids that carry a double-strand break or gap in the region of DNA homologous to the target locus produce 33- to 140-fold more hypoxanthine-aminopterin-thymidine-resistant recombinants than did these same plasmids introduced in their uncut (supercoiled) forms. Our data suggest that double-strand gaps and breaks may be enlarged prior to the repair reaction since sequence heterologies carried by the incoming plasmids located close to them are often lost. These results extend the known similarities between mammalian and yeast recombination mechanisms and suggest several features of the insertional (O-type) gene targeting reaction that should be considered when one is designing mammalian gene targeting experiments.  相似文献   

19.
We have cloned and characterized the white gene of Anopheles albimanus. Comparison of the deduced amino acid sequence of this white gene with its homologs from six species of Diptera show that the An. albimanus gene is most similar to the white gene of An. gambiae (92% identity). A spontaneous white-eyed mutant An. albimanus was caused by an approximately 10 kb insertion into a CT dinucleotide repeat region of intron 2 of the white locus. The flanks of this insertion are long (at least 400 bp), nearly perfect inverted terminal repeat sequences. This cloned white gene should be useful as a marker for germ line transformation of An. albimanus. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

20.
Mammalian cells are able to repair chromosomal double-strand breaks (DSBs) both by homologous recombination and by mechanisms that require little or no homology. Although spontaneous homologous recombination is rare, DSBs will stimulate recombination by 2 to 3 orders of magnitude when homology is provided either from exogenous DNA in gene-targeting experiments or from a repeated chromosomal sequence. Using a gene-targeting assay in mouse embryonic stem cells, we now investigate the effect of heterology on recombinational repair of DSBs. Cells were cotransfected with an endonuclease expression plasmid to induce chromosomal DSBs and with substrates containing up to 1.2% heterology from which to repair the DSBs. We find that heterology decreases the efficiency of recombinational repair, with 1.2% sequence divergence resulting in an approximately sixfold reduction in recombination. Gene conversion tract lengths were examined in 80 recombinants. Relatively short gene conversion tracts were observed, with 80% of the recombinants having tracts of 58 bp or less. These results suggest that chromosome ends in mammalian cells are generally protected from extensive degradation prior to recombination. Gene conversion tracts that were long (up to 511 bp) were continuous, i.e., they contained an uninterrupted incorporation of the silent mutations. This continuity suggests that these long tracts arose from extensive degradation of the ends or from formation of heteroduplex DNA which is corrected with a strong bias in the direction of the unbroken strand.  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