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.
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   

2.
The Chinese hamster dihydrofolate reductase (DHFR) origin of replication consists of a broad zone of potential initiation sites scattered throughout a 55-kb intergenic spacer, with at least three sites being preferred (ori-beta, ori-beta', and ori-gamma). We previously showed that deletion of the most active site or region (ori-beta) has no demonstrable effect on initiation in the remainder of the intergenic spacer nor on the time of replication of the DHFR locus as a whole. In the present study, we have now deleted ori-beta', both ori-beta and ori-beta', an 11-kb region just downstream from the DHFR gene, or the central approximately 40-kb core of the spacer. The latter two deletions together encompass >95% of the initiation sites that are normally used in this locus. Two-dimensional gel analysis shows that initiation still occurs in the early S phase in the remainder of the intergenic spacer in each of these deletion variants. Even removal of the 40-kb core fails to elicit a significant effect on the time of replication of the DHFR locus in the S period; indeed, in the truncated spacer that remains, the efficiency of initiation actually appears to increase relative to the corresponding region in the wild-type locus. Thus, if replicators control the positions of nascent strand start sites in this complex origin, either (i) there must be a very large number of redundant elements in the spacer, each of which regulates initiation only in its immediate environment, or (ii) they must lie outside the central core in which the vast majority of nascent strand starts occur.  相似文献   

3.
Previous two-dimensional gel replicon-mapping studies on the amplified dihydrofolate reductase (DHFR) domain in CHOC 400 cells suggested that replication can initiate at any of a large number of sites scattered throughout a 55-kb region lying between two convergently transcribed genes. It could be argued that this unusual distributive initiation mode is unique to amplified chromosomal loci. In this paper, we report the first application of the two-dimensional gel techniques to the analysis of a single-copy locus in mammalian cells. Results obtained with both synchronized and exponentially growing CHO cells suggest that (i) initiation can also occur at any of a large number of sites distributed throughout the intergenic region in the nonamplified DHFR locus, (ii) initiation is confined to the first 2 to 2.5 h of the S period, and (iii) initiation occurs only in a fraction of the DHFR loci in each cell cycle.  相似文献   

4.
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   

5.
The human papillomavirus (HPV) E1 and E2 proteins bind cooperatively to the viral origin of replication (ori), forming an E1-E2-ori complex that is essential for initiation of DNA replication. All other replication proteins, including DNA polymerase α-primase (polα-primase), are derived from the host cell. We have carried out a detailed analysis of the interactions of HPV type 16 (HPV-16) E1 with E2, ori, and the four polα-primase subunits. Deletion analysis showed that a C-terminal region of E1 (amino acids [aa] 432 to 583 or 617) is required for E2 binding. HPV-16 E1 was unable to bind the ori in the absence of E2, but the same C-terminal domain of E1 was sufficient to tether E1 to the ori via E2. Of the polα-primase subunits, only p68 bound E1, and binding was competitive with E2. The E1 region required (aa 397 to 583) was the same as that required for E2 binding but additionally contained 34 N-terminal residues. In confirmation of these differences, we found that a monoclonal antibody, mapping adjacent to the N-terminal junction of the p68-binding region, blocked E1-p68 but not E1-E2 binding. Sequence alignments and secondary-structure prediction for HPV-16 E1 and other superfamily 3 (SF3) viral helicases closely parallel the mapping data in suggesting that aa 439 to 623 constitute a discrete helicase domain. Assuming a common nucleoside triphosphate-binding fold, we have generated a structural model of this domain based on the X-ray structures of the hepatitis C virus and Bacillus stearothermophilus (SF2) helicases. The modelling closely matches the deletion analysis in suggesting that this region of E1 is indeed a structural domain, and our results suggest that it is multifunctional and critical to several stages of HPV DNA replication.  相似文献   

6.
Single-stranded DNA binding proteins (SSBs) are vital in all organisms. SSBs of Escherichia coli (EcoSSB) and Mycobacterium tuberculosis (MtuSSB) are homotetrameric. The N-terminal domains (NTD) of these SSBs (responsible for their tetramerization and DNA binding) are structurally well defined. However, their C-terminal domains (CTD) possess undefined structures. EcoSSB NTD consists of β1-β1′-β2-β3-α-β4-β451-β452-β5 secondary structure elements. MtuSSB NTD includes an additional β-strand (β6) forming a novel hook-like structure. Recently, we observed that MtuSSB complemented an E. coli Δssb strain. However, a chimeric SSB (mβ4-β5), wherein only the terminal part of NTD (β4-β5 region possessing L45 loop) of EcoSSB was substituted with that from MtuSSB, failed to function in E. coli in spite of its normal DNA binding and oligomerization properties. Here, we designed new chimeras by transplanting selected regions of MtuSSB into EcoSSB to understand the functional significance of the various secondary structure elements within SSB. All chimeric SSBs formed homotetramers and showed normal DNA binding. The mβ4-β6 construct obtained by substitution of the region downstream of β5 in mβ4-β5 SSB with the corresponding region (β6) of MtuSSB complemented the E. coli strain indicating a functional interaction between the L45 loop and the β6 strand of MtuSSB.  相似文献   

7.
Mammalian replication origins appear paradoxical. While some studies conclude that initiation occurs bidirectionally from specific loci, others conclude that initiation occurs at many sites distributed throughout large DNA regions. To clarify this issue, the relative number of early replication bubbles was determined at 26 sites in a 110-kb locus containing the dihydrofolate reductase (DHFR)-encoding gene in CHO cells; 19 sites were located within an 11-kb sequence containing ori-β. The ratio of ~0.8-kb nascent DNA strands to nonreplicated DNA at each site was quantified by competitive PCR. Nascent DNA was defined either as DNA that was labeled by incorporation of bromodeoxyuridine in vivo or as RNA-primed DNA that was resistant to λ-exonuclease. Two primary initiation sites were identified within the 12-kb region, where two-dimensional gel electrophoresis previously detected a high frequency of replication bubbles. A sharp peak of nascent DNA occurred at the ori-β origin of bidirectional replication where initiation events were 12 times more frequent than at distal sequences. A second peak occurred 5 kb downstream at a previously unrecognized origin (ori-β′). Thus, the DHFR gene initiation zone contains at least three primary initiation sites (ori-β, ori-β′, and ori-γ), suggesting that initiation zones in mammals, like those in fission yeast, consist of multiple replication origins.  相似文献   

8.
The streptococcal plasmid pMV158 replicates by the rolling-circle mechanism. One feature of this replication mechanism is the generation of single-stranded DNA intermediates which are converted to double-stranded molecules. Lagging-strand synthesis initiates from the plasmid single-stranded origin, sso. We have used the pMV158-derivative plasmid pLS1 (containing the ssoA type of lagging-strand origin) and a set of pLS1 derivatives with mutations in two conserved regions of the ssoA (the recombination site B [RSB] and a conserved 6-nucleotide sequence [CS-6]) to identify sequences important for plasmid lagging-strand replication in Streptococcus pneumoniae. Cells containing plasmids with mutations in the RSB accumulated 30-fold more single-stranded DNA than cells containing plasmids with mutations in the CS-6 sequence. Specificity of lagging-strand synthesis was tested by the development of a new in vitro replication system with pneumococcal cell extracts. Four major initiation sites of lagging-strand DNA synthesis were observed. The specificity of initiation was maintained in plasmids with mutations in the CS-6 region. Mutations in the RSB region, on the other hand, resulted in the loss of specific initiation of lagging-strand synthesis and also severely reduced the efficiency of replication.  相似文献   

9.
Pattern recognition receptors are known to participate in the activation of Prophenoloxidase system. In this study, a 1,3-β-D-glucan recognition protein was detected for the first time in Antheraea pernyi larvae (Ap-βGRP). Ap-βGRP was purified to 99.9% homogeneity from the hemolymph using traditional chromatographic methods. Ap-βGRP specifically bind 1,3-β-D-glucan and yeast, but not E. coli or M. luteus. The 1,3-β-D-glucan dependent phenoloxidase (PO) activity of the hemolymph inhibited by anti-Ap-βGRP antibody could be recovered by addition of purified Ap-βGRP. These results demonstrate that Ap-βGRP acts as a biosensor of 1,3-β-Dglucan to trigger the Prophenoloxidase system. A trace mount of 1,3-β-D-glucan or Ap-βGRP alone was unable to trigger the proPO system, but they both did. Ap-βGRP was specifically degraded following the activation of proPO with 1,3-β-Dglucan. These results indicate the variation in the amount of Ap-βGRP after specific immune challenge in A. pernyi hemolymph is an important regulation mechanism to immune response. [BMB Reports 2013; 46(5): 264-269]  相似文献   

10.
Three-dimensional structures of β2-microglobulin (β2m) from chicken and various mammals have been described previously, but aside from genomic sequences, very little is known about the three-dimensional structures of β2m in species other than warm-blooded vertebrates. Here, we present the first three-dimensional structure of β2m from bony fish grass carp (Ctid-β2m), resolved at 2.1 Å. The key structural differences between this new structure and previously published structures are two new hydrogen bonds at positions Ile37 and Glu38 in strand C and Lys66 in strand E, and a hydrophobic pocket around the center of the protein found in Ctid-β2m. Importantly, Ctid-β2m has a short D strand and a long loop between stands C and D, rather than the flexible region found in other β2m structures that serves as a putative binding region for the major histocompatibility complex heavy chain. Comparing the Ctid-β2m structure with those of bovine and human β2ms, the Cα root mean square deviation of the latter are 1.3 Å and 1.8 Å, respectively. Compared with the constant domains of Lamprey T cell receptor-like receptor (Lamp-TCRLC) and Amphioxus V and C domain-bearing protein (Amphi-VCPC), Ctid-β2m exhibits very different topology. The three-dimensional structures of domains predicted from Amphi-VCPC/Lamp-TCRLC are distinctly lacking in strand A of β2ms. There are 18 amino acids at the N terminus of Amphi-VCPC that may have evolved into strand A of β2ms. A mutation in the BC loops of Amphi-VCPC may have led to the novel topology found in β2m. Based on these results, Ctid-β2m may well reflect evolutionary characteristics of ancestral C set molecules.  相似文献   

11.
T cell receptor (TCR) β gene assembly by V(D)J recombination proceeds via successive Dβ-to-Jβ and Vβ-to-DJβ rearrangements. This two-step process is enforced by a constraint, termed beyond (B)12/23, which prohibits direct Vβ-to-Jβ rearrangements. However the B12/23 restriction does not explain the order of TCRβ assembly for which the regulation remains an unresolved issue. The initiation of V(D)J recombination consists of the introduction of single-strand DNA nicks at recombination signal sequences (RSSs) containing a 12 base-pairs spacer. An RSS containing a 23 base-pairs spacer is then captured to form a 12/23 RSSs synapse leading to coupled DNA cleavage. Herein, we probed RSS nicks at the TCRβ locus and found that nicks were only detectable at Dβ-associated RSSs. This pattern implies that Dβ 12RSS and, unexpectedly, Dβ 23RSS initiate V(D)J recombination and capture their respective Vβ or Jβ RSS partner. Using both in vitro and in vivo assays, we further demonstrate that the Dβ1 23RSS impedes cleavage at the adjacent Dβ1 12RSS and consequently Vβ-to-Dβ1 rearrangement first requires the Dβ1 23RSS excision. Altogether, our results provide the molecular explanation to the B12/23 constraint and also uncover a ‘Dβ1 23RSS-mediated’ restriction operating beyond chromatin accessibility, which directs Dβ1 ordered rearrangements.  相似文献   

12.
Mini-chromosome maintenance (MCM) proteins were originally identified in yeast, and homologues have been identified in several other eukaryotic organisms, including mammals. These findings suggest that the mechanisms by which eukaryotic cells initiate and regulate DNA replication have been conserved throughout evolution. However, it is clear that many mammalian origins are much more complex than those of yeast. An example is the Chinese hamster dihydrofolate reductase (DHFR) origin, which resides in the spacer between the DHFR and 2BE2121 genes. This origin consists of a broad zone of potential sites scattered throughout the 55-kb spacer, with several subregions (e.g. ori-beta, ori-beta', and ori-gamma) being preferred. We show here that antibodies to human MCMs 2-7 recognize counterparts in extracts prepared from hamster cells; furthermore, co-immunoprecipitation data demonstrate the presence of an MCM2-3-5 subcomplex as observed in other species. To determine whether MCM proteins play a role in initiation and/or elongation in Chinese hamster cells, we have examined in vivo protein-DNA interactions between the MCMs and chromatin in the DHFR locus using a chromatin immunoprecipitation (ChIP) approach. In synchronized cultures, MCM complexes associate preferentially with DNA in the intergenic initiation zone early in S-phase during the time that replication initiates. However, significant amounts of MCMs were also detected over the two genes, in agreement with recent observations that the MCM complex co-purifies with RNA polymerase II. As cells progress through S-phase, the MCMs redistribute throughout the DHFR domain, suggesting a dynamic interaction with DNA. In asynchronous cultures, in which replication forks should be found at any position in the genome, MCM proteins were distributed relatively evenly throughout the DHFR locus. Altogether, these data are consistent with studies in yeast showing that MCM subunits localize to origins during initiation and then migrate outward with the replication forks. This constitutes the first evidence that mammalian MCM complexes perform a critical role during the initiation and elongation phases of replication at the DHFR origin in hamster cells.  相似文献   

13.
Bulone V  Girard V  Fèvre M 《Plant physiology》1990,94(4):1748-1755
Enriched 1,3-β-glucan and 1,4-β-glucan synthase fractions from the fungus Saprolegnia were isolated by rate zonal centrifugation on glycerol gradient. Purification was improved by entrapment of the enzymes in their reaction product, i.e. microfibrillar glucans. 1,3-β-Glucan synthases were separated from 1,4-β-glucan synthases following resuspension of entrapped enzymes. Sodium dodecylsulfate-polyacrylamide gel electrophoresis indicated that 1,3-β-glucan and 1,4-β-glucan synthases may have a different polypeptide composition because they were enriched for different protein subunits (34, 48, and 50 kD for the 1,3-β-glucan synthase and 60 kD for the 1,4-β-glucan synthase).  相似文献   

14.
N2,3-Ethenoguanine (N2,3-ϵG) is one of the exocyclic DNA adducts produced by endogenous processes (e.g. lipid peroxidation) and exposure to bioactivated vinyl monomers such as vinyl chloride, which is a known human carcinogen. Existing studies exploring the miscoding potential of this lesion are quite indirect because of the lability of the glycosidic bond. We utilized a 2′-fluoro isostere approach to stabilize this lesion and synthesized oligonucleotides containing 2′-fluoro-N2,3-ϵ-2′-deoxyarabinoguanosine to investigate the miscoding potential of N2,3-ϵG by Y-family human DNA polymerases (pols). In primer extension assays, pol η and pol κ replicated through N2,3-ϵG, whereas pol ι and REV1 yielded only 1-base incorporation. Steady-state kinetics revealed that dCTP incorporation is preferred opposite N2,3-ϵG with relative efficiencies in the order of pol κ > REV1 > pol η ≈ pol ι, and dTTP misincorporation is the major miscoding event by all four Y-family human DNA pols. Pol ι had the highest dTTP misincorporation frequency (0.71) followed by pol η (0.63). REV1 misincorporated dTTP and dGTP with much lower frequencies. Crystal structures of pol ι with N2,3-ϵG paired to dCTP and dTTP revealed Hoogsteen-like base pairing mechanisms. Two hydrogen bonds were observed in the N2,3-ϵG:dCTP base pair, whereas only one appears to be present in the case of the N2,3-ϵG:dTTP pair. Base pairing mechanisms derived from the crystal structures explain the slightly favored dCTP insertion for pol ι in steady-state kinetic analysis. Taken together, these results provide a basis for the mutagenic potential of N2,3-ϵG.  相似文献   

15.
Etheno DNA adducts are a prevalent type of DNA damage caused by vinyl chloride (VC) exposure and oxidative stress. Etheno adducts are mutagenic and may contribute to the initiation of several pathologies; thus, elucidating the pathways by which they induce cellular transformation is critical. Although N2,3-ethenoguanine (N2,3-εG) is the most abundant etheno adduct, its biological consequences have not been well characterized in cells due to its labile glycosidic bond. Here, a stabilized 2′-fluoro-2′-deoxyribose analog of N2,3-εG was used to quantify directly its genotoxicity and mutagenicity. A multiplex method involving next-generation sequencing enabled a large-scale in vivo analysis, in which both N2,3-εG and its isomer 1,N2-ethenoguanine (1,N2-εG) were evaluated in various repair and replication backgrounds. We found that N2,3-εG potently induces G to A transitions, the same mutation previously observed in VC-associated tumors. By contrast, 1,N2-εG induces various substitutions and frameshifts. We also found that N2,3-εG is the only etheno lesion that cannot be repaired by AlkB, which partially explains its persistence. Both εG lesions are strong replication blocks and DinB, a translesion polymerase, facilitates the mutagenic bypass of both lesions. Collectively, our results indicate that N2,3-εG is a biologically important lesion and may have a functional role in VC-induced or inflammation-driven carcinogenesis.  相似文献   

16.
The translation initiation factor aIF2 of the crenarchaeon Sulfolobus solfataricus (Sso) recruits initiator tRNA to the ribosome and stabilizes mRNAs by binding via the γ-subunit to their 5′-triphosphate end. It has been hypothesized that the latter occurs predominantly during unfavorable growth conditions, and that aIF2 or aIF2-γ is released on relief of nutrient stress to enable in particular anew translation of leaderless mRNAs. As leaderless mRNAs are prevalent in Sso and aIF2-γ bound to the 5′-end of a leaderless RNA inhibited ribosome binding in vitro, we aimed at elucidating the mechanism underlying aIF2/aIF2-γ recycling from mRNAs. We have identified a protein termed Trf (translation recovery factor) that co-purified with trimeric aIF2 during outgrowth of cells from prolonged stationary phase. Subsequent in vitro studies revealed that Trf triggers the release of trimeric aIF2 from RNA, and that Trf directly interacts with the aIF2-γ subunit. The importance of Trf is further underscored by an impaired protein synthesis during outgrowth from stationary phase in a Sso trf deletion mutant.  相似文献   

17.
Genomic DNA in higher eucaryotic cells is organized into a series of loops, each of which may be affixed at its base to the nuclear matrix via a specific matrix attachment region (MAR). In this report, we describe the distribution of MARs within the amplified dihydrofolate reductase (DHFR) domain (amplicon) in the methotrexate-resistant CHO cell line CHOC 400. In one experimental protocol, matrix-attached and loop DNA fractions were prepared from matrix-halo structures by restriction digestion and were analyzed for the distribution of amplicon sequences between the two fractions. A second, in vitro method involved the specific binding to the matrix of cloned DNA fragments from the amplicon. Both methods of analysis detected a MAR in the replication initiation locus that we have previously defined in the DHFR amplicon, as well as in the 5'-flanking region of the DHFR gene. The first of these methods also suggests the presence of a MAR in a region mapping approximately 120 kilobases upstream from the DHFR gene. Each of these MARs was detected regardless of whether the matrix-halo structures were prepared by the high-salt or the lithium 3,5-diiodosalicylate extraction protocols, arguing against their artifactual association with the proteinaceous scaffolding of the nucleus during isolation procedures. However, the in vitro binding assay did not detect the MAR located 120 kilobases upstream from the DHFR gene but did detect specific matrix attachment of a sequence near the junction between amplicons. The results of these experiments suggest that (i) MARs can occur next to different functional elements in the genome, with the result that a DNA loop formed between two MARs can be smaller than a replicon; and (ii) different methods of analysis detect a somewhat different spectrum of matrix-attached DNA fragments.  相似文献   

18.
In previous studies, we used two complementary two-dimensional gel electrophoretic methods to examine replication intermediates in the 240-kb amplified dihydrofolate reductase (DHFR) domain of methotrexate-resistant CHOC 400 cells (J. P. Vaughn, P. A. Dijkwel, and J. L. Hamlin, Cell 61:1075-1087, 1990). Surprisingly, in both asynchronous and early-S-phase cultures, initiation bubbles were detected in several contiguous fragments from a previously defined 28-kb initiation locus. However, because of the low levels of bubblelike structures observed on gels, it has been suggested that these structures might represent artifacts, possibly unrelated to replication per se. In this study, we have achieved much more synchronous entry into S phase by using a novel inhibitor and have isolated replication intermediates by a new procedure that largely eliminates branch migration and shear. Under these conditions, we find that (i) the relative number of bubblelike structures detected in fragments from the initiation locus is markedly increased, (ii) bubbles are detected at multiple sites scattered throughout the region lying between the DHFR and 2BE2121 genes, and (iii) bubbles appear and disappear in this region with the kinetics expected of an early-firing origin. These data strengthen the proposal that in vivo, initiation can occur at any of a large number of sites scattered throughout a broad zone in the DHFR domain.  相似文献   

19.
The nature of mammalian origins of DNA replication remains controversial and this is primarily because two-dimensional gel replicon mapping techniques have identified broad zones of replication initiation whereas several other techniques, such as quantitative PCR, have disclosed more discrete sites of initiation at the same chromosomal loci. In this report we analyze the replication of an amplified genomic region encompassing the 3'-end of the GNAI3 gene, the entire GNAT2 gene and the intergenic region between them in exponentially growing Chinese hamster fibroblasts. These cells express GNAI3 but not GNAT2 . The replication pattern was first analyzed by two-dimensional neutral-alkaline gel electrophoresis. Surprisingly, the results revealed a small preferential zone of replication initiation, of at most 1.7 kb, located in a limited part of the GNAI3 - GNAT2 intergenic region. Mapping of this initiation zone was then confirmed by quantitative PCR. The agreement between the two techniques exploited here strengthens the hypothesis that preferred sites of replication initiation do exist in mammalian genomes.  相似文献   

20.
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   

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