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 0 毫秒
1.
We isolated overlapping recombinant cosmids that represent the equivalent of two complete dihydrofolate reductase amplicon types from the methotrexate-resistant CHO cell line CHOC400. The type I amplicons are 260 kilobases long, are arranged in head-to-tail fashion, and represent 10 to 15% of the amplicons in the CHOC400 genome. The type II amplicons are 220 kilobases long, are arranged in head-to-head and tail-to-tail configurations, and constituted the majority of the remaining amplicons in CHOC400 cells. The type II amplicon sequences are represented entirely within the type I unit. These are the first complete amplicons to be cloned from a mammalian cell line.  相似文献   

2.
In previous studies, we utilized a neutral/neutral two-dimensional (2-D) gel replicon mapping method to analyze the pattern of DNA synthesis in the amplified dihydrofolate reductase (DHFR) domain of CHOC 400 cells. Replication forks appeared to initiate at any of a large number of sites scattered throughout the 55 kb region lysing between the DHFR and 2BE2121 genes, and subsequently to move outward through the two genes. In the present study, we have analyzed this locus in detail by a complementary, neutral/alkaline 2-D gel technique that determines the direction in which replication forks move through a region of interest. In the early S period, forks are observed to travel in both directions through the intergenic region, but only outward through the DHFR gene. Surprisingly, however, replication forks also move in both directions through the 2BE2121 gene. Furthermore, in early S phase, small numbers of replication bubbles can be detected in the 2BE2121 gene on neutral/neutral 2-D gels. In contrast, replication bubbles have never been detected in the DHFR gene. Thus, replication initiates not only in the intergenic region, but also at a lower frequency in the 2BE2121 gene. We further show that only a small fraction of DHFR amplicons sustains an active initiation event, with the rest being replicated passively by forks from distant amplicons. These findings are discussed in light of other experimental approaches that suggest the presence of a much more narrowly circumscribed initiation zone within the intergenic region.  相似文献   

3.
We have examined the timing of replication of the amplified dihydrofolate reductase genes in the methotrexate-resistant Chinese hamster ovary cell line CHOC 400 using two synchronization procedures. DNA replicated in the presence of 5-bromodeoxyuridine was collected from cells of various times during the DNA synthesis phase and the extent of replication for defined sequences was determined by Southern blotting analysis of CsCl density gradient fractions. We report that under these conditions the DHFR gene replicates throughout the course of S phase in a mode similar to the bulk of the replicated genomic DNA. This contrasts with previous data that shows the non-amplified DHFR gene replicates during the first quarter of S phase. Therefore, we conclude that gene amplification alters the replication timing of the DHFR gene in CHOC 400 cells.  相似文献   

4.
Two-dimensional (2-D) gel analysis of replication intermediates in the Chinese hamster dihydrofolate reductase domain has suggested that nascent chains can initiate at any of a large number of sites scattered throughout a ~50 kb “initiation locus” (although the level of initiation detected at any given site within this region was relatively low). This result contrasts markedly with data from anin vitro strand switching assay suggesting that >80% of initiations occur within a single 500 bp fragment lying within the initiation locus. In an effort to reconcile these two disparate views of the initiation reaction, we have questioned the validity of our 2-D gel data in several ways. We show here that: 1) the number of replication bubbles detected in the DHFR locus in the early S period is markedly increased when the cells are released from a synchronizing agent that inhibits initiationper se, rather than from aphidicolin, which is a chain elongation inhibitor; 2) initiation in the DHFR domain occurs only during the first 90 min of the S period, as would be expected of an early-firing origin; 3) a pulse of3H-thymidine moves through the structures observed on 2-D gels with the kinetics expected ofbonafide replication intermediates; and 4) preparations of replication intermediates that are subsequently analyzed on 2-D gels appear, by electron microscopy, to represent the typical theta structures and single-forked molecules expected of bidirectional origins of replication; no unusual structures (e.g., microbubbles) were seen.  相似文献   

5.
New sublines of BFFR1 and BFFR3 cells were obtained as a result of prolonged cultivation of Chinese hamster cells of Blld-ii-FAF 28 line (clone 431) in the presence of increasing concentrations of methotrexate (MTX). The lines obtained were resistant to 200 and 300 mcM of MTX, respectively. Amplification of the gene for dihydrofolate reductase (DHFR), similar to normal DHFR gene in restriction patterns, was proved by blot-hybridization of the resistant cells' DNA with 32P-labeled plasmid DHFR-26. Correlation is shown between the extent of gene amplification and resistance of the cell lines. In situ hybridization of the metaphase chromosomes of resistant cells with 3H-DHFR-26 results in preferential binding of the label with the regions of marker chromosomes 2 and 5, containing long, so called differential staining regions which are known to be the places of localization of amplified genes.  相似文献   

6.
Overlapping recombinant lambda 1059 phages carrying regions of the dhfr locus from the amplified Chinese hamster ovary (CHO) cell clone MK42 have been isolated. In addition, dhfr cDNAs from this cell line have been cloned into plasmid pBR322. Restriction analysis of these recombinant molecules has led to a map of the Chinese hamster dhfr gene. This gene has a minimum size of 26 kb and contains six exons as defined by hybridization to a combination of mouse and CHO cDNA probes. The latter probes reveal 3' exonic sequences that are not present in mouse cDNA. The CHO dhfr gene thus extends about 700 bp further 3' than in the mouse, consistent with the larger size of the hamster mRNA. At least five intervening sequences are present, of approximate sizes: 0.3, 2.5, 8.6, 2.6 and 9.4 kb. Four sequences from highly repeated families are situated in introns within the dhfr gene. The overall structure of this gene is strikingly similar to that of the mouse. Evolutionary conservation of interrupted gene structure among mammals thus extends to genes that code for household enzymes as well as specialized or structural proteins.  相似文献   

7.
T H Leu  B Anachkova  J L Hamlin 《Genomics》1990,7(3):428-433
We have previously demonstrated that one of the replication initiation loci in the dihydrofolate reductase (DHFR) domain of Chinese hamster cells contains a repeated sequence that is enriched in the early-replicating fraction of the Chinese hamster genome. Here we present the sequence of the initiation locus, identify the relevant repeated element, and show that, while this element is enriched in early-replicating DNA, its synthesis is not confined to early S.  相似文献   

8.
The formation and removal of UV-induced pyrimidine dimers were measured in restriction fragments near and within the essential dihydrofolate reductase (DHFR) gene in Chinese hamster ovary cells in order to map the genomic fine structure of DNA repair. Dimer frequencies were determined at 0, 8, and 24 h after irradiating the cells with 20 J/m2 UV light (254 nm). Within 8 h, the cells had removed more than 40% of the dimers from sequences near the 5' end of the gene, somewhat fewer from the 3' end, but only 2% from the 3' flanking region and 10% from a region upstream from the gene. The corresponding extent of repair in the genome as a whole is 5-10% in the 8-h period. Isoschizomeric restriction enzyme analysis was used to detect the level of methylation in the fragments in which repair was measured. We found that the only hypomethylated sites in and around the DHFR gene were in the fragment near its 5' end, which displayed maximal DNA repair efficiency. The size of the region of preferential DNA repair at the DHFR locus appears to be in the range of 50-80 kilobases, and this finding is discussed in relation to genomic domains and the structure of mammalian chromatin.  相似文献   

9.
We isolated and characterized three spontaneous mutants of Chinese hamster ovary cells that were deficient in dihydrofolate reductase activity. All three mutants contained no detectable enzyme activity and produced dihydrofolate reductase mRNA species that were shorter than those of the wild type by about 120 bases. Six exons are normally represented in this mRNA; exon 5 was missing in all three mutant mRNAs. Nuclease S1 analysis of the three mutants indicated that during the processing of the mutant RNA, exon 4 was spliced to exon 6. The three mutant genes were cloned, and the regions around exons 4 and 5 were sequenced. In one mutant, the GT dinucleotide at the 5' end of intron 5 had changed to CT. In a second mutant, the first base in exon 5 had changed from G to T. In a revertant of this mutant, this base was further mutated to A, a return to a purine. Approximately 25% of the mRNA molecules in the revertant were spliced correctly to produce an enzyme with one presumed amino acid change. In the third mutant, the AG at the 3' end of intron 4 had changed to AA. A mutation that partially reversed the mutant phenotype had changed the dinucleotide at the 5' end of intron 4 from GT to AT. The splicing pattern in this revertant was consistent with the use of cryptic donor and acceptor splice sites close to the original sites to produce an mRNA with three base changes and a protein with two amino acid changes. These mutations argue against a scanning model for the selection of splice site pairs and suggest that only a single splice site need be inactivated to bring about efficient exon skipping (a regulatory mechanism for some genes). The fact that all three mutants analyzed exhibited exon 5 splicing mutations indicates that these splice sites are hot spots for spontaneous mutation.  相似文献   

10.
During stepwise increases in the methotrexate concentration in culture medium, we selected Chinese hamster ovary cells that contained elevated dihydrofolate reductase levels which were proportional to the number of dihydrofolate reductase gene copies (i.e., gene amplification). We studied the dihydrofolate reductase levels in individual cells that underwent the initial steps of methotrexate resistance by using the fluorescence-activated cell sorter technique. Such cells constituted a heterogeneous population with differing dihydrofolate reductase levels, and they characteristically lost the elevated enzyme levels when they were grown in the absence of methotrexate. The progeny of individual cells with high enzyme levels behaved differently and could lose all or variable numbers of the amplified genes.  相似文献   

11.
Stable isolates of Chinese hamster ovary cells that are highly resistant to methotrexate have been selected in a multistep selection process. Quantitative immunoprecipitations have indicated that these isolates synthesize dihydrofolate reductase at an elevated rate over its synthesis in sensitive cells. Restriction enzyme and Southern blot analyses with a murine reductase cDNA probe indicate that the highly resistant isolates contain amplifications of the dihydrofolate reductase gene number. Depending upon the parenteral line used to select these resistant cells, they overproduce either a wild-type enzyme or a structurally altered enzyme. Karyotype analysis shows that some of these isolates contain chromosomes with homogeneously staining regions whereas others do not contain such chromosomes.  相似文献   

12.
We have previously described methotrexate-resistant Chinese hamster ovary cells which appear to contain normal levls of a structurally altered dihydrofolate reductase (EC 1.5.1.3) (Flintoff, W.F., Davidson, S.V., and Siminovitch, L. (1976) Somatic Cell Genet.2,245-261). By selecting for increased resistance form these class I cells, class III resistant cells were isolated which appeared to possess an increased activity of the altered enzyme. In the report, we describe the purification and several properties of the reductase from wild-type cells, two independently selected class I cells, and class III resistant cell. The reductases from wild-type and resistant cells had similar specific activities using folate and dihydrofolate as substrates, and similar molecular weights as determined by sodium dodecyl sulfate gel electrophoresis. The mutant enzymes, however, were about six- to eight-fold more resistant to inhibition by methotrexate than the wild-type enzyme, suggesting a decreased affinity of the mutant reductases to methotrexate-binding. Small differences between various enzymes were also seen in other physicochemical properties such as pH optima and Km values for folate, and in their heat stabilities, which suggest that different structural alterations may lead to the same mutant phenotype. As expected from earlier studies with crude extracts, class III cells did produce a higher (about 10-fold) yield of the reductase than the class I or wild-type cells.  相似文献   

13.
14.
Previous radiolabeling and two-dimensional (2-D) gel studies of the dihydrofolate reductase (DHFR) domain of Chinese hamster cells have suggested that replication can initiate at any one of a very large number of inefficient sites scattered throughout the 55-kb intergenic spacer region, with two broad subregions (ori-beta and ori-gamma) preferred. However, high-resolution analysis by a PCR-based nascent strand abundance assay of the 12-kb subregion encompassing ori-beta has suggested the presence of a relatively small number of fixed, highly efficient initiation sites distributed at infrequent intervals that correspond to genetic replicators. To attempt to reconcile these observations, two different approaches were taken in the present study. In the first, neutral-neutral 2-D gel analysis was used to examine replication intermediates in 31 adjacent and overlapping restriction fragments in the spacer, ranging in size from 1.0 to 18 kb. Thirty of 31 fragments displayed the complete bubble arcs characteristic of centered origins. Taking into account overlapping fragments, these data suggest a minimum of 14 individual start sites in the spacer. In the second approach, a quantitative early labeled fragment hybridization assay was performed in which radioactive origin-containing DNA 300 to 1,000 nucleotides in length was synthesized in the first few minutes of the S period and used to probe 15 clones distributed throughout the intergenic spacer but separated on average by more than 1,000 bp. This small nascent DNA fraction hybridized to 14 of the 15 clones, ranging from just above background to a maximum at the ori-beta locus. The only silent region detected was a small fragment lying just upstream from a centered matrix attachment region--the same region that was also negative for initiation by 2-D gel analysis. Results of both approaches suggest a minimum of approximately 20 initiation sites in the spacer (two of them being ori-beta and ori-gamma), with ori-beta accounting for a maximum of approximately 20% of initiations occurring in the spacer. We believe that the results of all experimental approaches applied to this locus so far can be fitted to a model in which the DHFR origin consists of a 55-kb intergenic zone of potential sites that are used with very different efficiencies and which are separated in many cases by a few kilobases or less.  相似文献   

15.
Utilizing an in vivo labeling method on synchronized cultures, we have previously defined a 28-kilobase (kb) replication initiation locus in the amplified dihydrofolate reductase domain of a methotrexate-resistant Chinese hamster ovary cell line (CHOC 400) (N. H. Heintz and J. L. Hamlin, Proc. Natl. Acad. Sci. USA 79:4083-4087, 1982; N. H. Heintz and J. L. Hamlin, Biochemistry 22:3552-3557, 1983; N. H. Heintz, J. D. Milbrandt, K. S. Greisen, and J. L. Hamlin, Nature [London] 302:439-441, 1983). To locate the origin of replication in this 243-kb amplicon with more precision, we used an in-gel renaturation procedure (I. Roninson, Nucleic Acids Res. 11:5413-5431, 1983) to examine the labeling pattern of restriction fragments from the amplicon in the early S phase. This method eliminates background labeling from single-copy sequences and allows quantitation of the relative radioactivity in individual fragments. We used this procedure to follow the movement of replication forks through the amplicons, to roughly localize the initiation locus, and to estimate the rate of fork travel. We also used a slight modification of this method (termed hybridization enhancement) to illuminate the labeling pattern of smaller restriction fragments derived solely from the initiation locus itself, thereby increasing resolution. Our preliminary results suggest that there are actually two distinct initiation sites in the amplicon that are separated by approximately 22 kb.  相似文献   

16.
17.
We have investigated different parameters characterizing carcinogen-mediated enhancement of methotrexate resistance in Chinese hamster ovary (CHO) cells and in simian virus 40-transformed Chinese hamster embryo (C060) cells. We show that this enhancement reflects dihydrofolate reductase (dhfr) gene amplification. The carcinogens used in this work are alkylating agents and UV irradiation. Both types of carcinogens induce a transient enhancement of methotrexate resistance which increases gradually from the time of treatment to 72 to 96 h later and decreases thereafter. Increasing doses of carcinogens decrease cell survival and increase the enhancement of methotrexate resistance. Enhancement was observed when cells were treated at different stages in the cell cycle, and it was maximal when cells were treated during the early S phase. These studies of carcinogen-mediated dhfr gene amplification coupled with our earlier studies on viral DNA amplification in simian virus 40-transformed cells demonstrate that the same parameters characterize the amplification of both genes. Possible cellular mechanisms responsible for the carcinogen-mediated gene amplification phenomenon are discussed.  相似文献   

18.
Genome-scale mapping of pre-replication complex proteins has not been reported in mammalian cells. Poor enrichment of these proteins at specific sites may be due to dispersed binding, poor epitope availability or cell cycle stage-specific binding. Here, we have mapped sites of biotin-tagged ORC and MCM protein binding in G1-synchronized populations of Chinese hamster cells harboring amplified copies of the dihydrofolate reductase (DHFR) locus, using avidin-affinity purification of biotinylated chromatin followed by high-density microarray analysis across the DHFR locus. We have identified several sites of significant enrichment for both complexes distributed throughout the previously identified initiation zone. Analysis of the frequency of initiations across stretched DNA fibers from the DHFR locus confirmed a broad zone of de-localized initiation activity surrounding the sites of ORC and MCM enrichment. Mapping positions of mononucleosomal DNA empirically and computing nucleosome-positioning information in silico revealed that ORC and MCM map to regions of low measured and predicted nucleosome occupancy. Our results demonstrate that specific sites of ORC and MCM enrichment can be detected within a mammalian initiation zone, and suggest that initiation zones may be regions of generally low nucleosome occupancy where flexible nucleosome positioning permits flexible pre-RC assembly sites.  相似文献   

19.
20.
Expression of human tissue-type plasminogen activator (t-PA) at high levels has been achieved in Chinese hamster ovary (CHO) cells by cotransfection and subsequent coamplification of the transfected sequences. Expression vectors containing the t-PA cDNA gene and dihydrofolate reductase (DHFR) cDNA gene were cotransfected into CHO DHFR-deficient cells. Transformants expressing DHFR were selected by growth in media lacking nucleosides and contained low numbers of t-PA genes and DHFR genes. Stepwise selection of the DHFR+ transformants in increasing concentrations of methotrexate generated cells which had amplified both DHFR genes and t-PA genes over 100-fold. These cell lines expressed elevated levels of enzymatically active t-PA. To optimize both t-PA sequence amplification and t-PA expression, various modifications of the original procedure were used. These included alterations to the DHFR expression vector, optimization of the molar ratio of t-PA to DHFR sequences in the cotransfection, and modification of the methotrexate resistance selection procedure. The structure of the amplified DNA, its chromosomal location, and its stability during growth in the absence of methotrexate are reported.  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