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.
2.
3.
DnaA is a replication initiator protein that is conserved among bacteria. It plays a central role in the initiation of DNA replication. In order to monitor its behavior in living Escherichia coli cells, a nonessential portion of the protein was replaced by a fluorescent protein. Such a strain grew normally, and flow cytometry data suggested that the chimeric protein has no substantial loss of the initiator activity. The initiator was distributed all over the nucleoid. Furthermore, a majority of the cells exhibited certain distinct foci that emitted bright fluorescence. These foci colocalized with the replication origin (oriC) region and were brightest during the period spanning the initiation event. In cells that had undergone the initiation, the foci were enriched in less intense ones. In addition, a significant portion of the oriC regions at this cell cycle stage had no colocalized DnaA-enhanced yellow fluorescent protein (EYFP) focus point. It was difficult to distinguish the initiator titration locus (datA) from the oriC region. However, involvement of datA in the initiation control was suggested from the observation that, in ΔdatA cells, DnaA-EYFP maximally colocalized with the oriC region earlier in the cell cycle than it did in wild-type cells and oriC concentration was increased.Initiation of DNA replication is highly regulated to coordinate with cell proliferation. It begins with a series of events in which the replication machinery is assembled at the replication origin of the chromosomal DNA (15, 26, 28, 38). Central to this process are the initiator proteins that bind to the origin of replication and eventually lead to the unwinding of the origin and to helicase loading on the unwound region. Previous biochemical studies and recent structural studies of the bacterial initiator protein DnaA have proposed the molecular mechanism of the action of ATP-DnaA in forming a large oligomeric complex to remodel the unique origin, oriC, and trigger duplex melting (12, 26). However, it is still not clear how the timing of initiation is controlled so that it takes place at a fixed time in the cell cycle. It has been reported that a basal level of DnaA molecules is bound by high-affinity DnaA binding sites (DnaA boxes R1, R2, and R4) at oriC throughout the cell cycle (9, 37). It is also suggested that noncanonical ATP-DnaA binding sites within oriC are occupied at elevated levels of the initiator molecules prior to the initiation event (18, 25). Thus, regulation of the activity and availability of DnaA is an important factor for the initiation control.At least three schemes are known to prevent untimely initiations in Escherichia coli. First, oriC is subject to sequestration, a process that prevents reinitiation, possibly by blocking ATP-DnaA from binding to newly replicated oriC (8, 24). E. coli oriC contains 11 GATC sites that are normally methylated on both strands by Dam methyltransferase. Immediately after passage of the replication fork, GATC sites are in a hemimethylated state, with the newly synthesized strands remaining unmethylated. SeqA binds specifically to such sites and, at oriC, protects these regions from reinitiation for about one-third of the cell cycle (6, 39). Second, in a process termed regulatory inactivation of DNA (RIDA), ATP-DnaA molecules are converted to an inactive ADP-bound form after initiation by the combined action of a β subunit of DNA polymerase III holoenzyme and Hda (16, 17). Newly synthesized DnaA molecules are able to bind ATP for the next initiation event, since its cellular concentration is much higher than that of ADP. ATP-DnaA is also regenerated from the inactive ADP-DnaA later in the cell cycle (21). Finally, the chromosomal segment datA serves to reduce the level of free DnaA protein by titrating a large number of DnaA molecules after replication of the site close to oriC (20).Cytological studies would be very useful for developing our understanding of the regulation mechanisms associated with the initiation step. In the present study, we tagged E. coli DnaA with a fluorescent protein in order to monitor its behavior in live cells. Microscopic observation revealed that DnaA is distributed all over the nucleoid. Remarkably, the majority of cells bore distinct foci that emitted brighter fluorescence against a weak fluorescent background on the nucleoid. We analyzed the behavior of these foci during the cell cycle with respect to oriC and datA.  相似文献   

4.
5.
6.
7.
8.
9.
Cellular and biochemical studies support a role for all five human RecQ helicases in DNA replication; however, their specific functions during this process are unclear. Here we investigate the in vivo association of the five human RecQ helicases with three well-characterized human replication origins. We show that only RECQ1 (also called RECQL or RECQL1) and RECQ4 (also called RECQL4) associate with replication origins in a cell cycle-regulated fashion in unperturbed cells. RECQ4 is recruited to origins at late G1, after ORC and MCM complex assembly, while RECQ1 and additional RECQ4 are loaded at origins at the onset of S phase, when licensed origins begin firing. Both proteins are lost from origins after DNA replication initiation, indicating either disassembly or tracking with the newly formed replisome. Nascent-origin DNA synthesis and the frequency of origin firing are reduced after RECQ1 depletion and, to a greater extent, after RECQ4 depletion. Depletion of RECQ1, though not that of RECQ4, also suppresses replication fork rates in otherwise unperturbed cells. These results indicate that RECQ1 and RECQ4 are integral components of the human replication complex and play distinct roles in DNA replication initiation and replication fork progression in vivo.The RecQ helicases are a family of DNA-unwinding enzymes essential for the maintenance of genome integrity in all kingdoms of life. Five RecQ enzymes have been found in human cells: RECQ1 (also called RECQL or RECQL1), BLM (RECQ2 or RECQL3), WRN (RECQ3 or RECQL2), RECQ4 (RECQL4), and RECQ5 (RECQL5) (3, 7). Here we refer to these helicases as RECQ1, RECQ4, and RECQ5, without the “L” that is present in the official gene names. Mutations in the BLM, WRN, and RECQ4 genes are linked to Bloom syndrome (BS), Werner syndrome (WS), and the subset of Rothmund-Thomson syndrome (RTS) patients at high risk of developing osteosarcomas, respectively (19, 31, 71). RECQ4 mutations have also been associated with RAPADILINO and Baller-Gerold syndrome (56, 61). Although these disorders are all associated with inherent genomic instability and cancer predisposition, they show distinct clinical features, suggesting that BLM, WRN, and RECQ4 are involved in different aspects of DNA metabolism. However, the molecular events underlying the pathogenesis of BS, WS, and RTS remain obscure. Mutations in the remaining two human RecQ helicase genes, RECQ1 and RECQ5, have not as yet been identified as causes of either genomic instability or heritable cancer predisposition disorders.Several lines of evidence suggest that RecQ helicases play an important role in DNA replication control (3, 10). In particular, RecQ helicases are thought to facilitate replication by preserving the integrity of stalled replication forks and by remodeling or repairing damaged or collapsed forks to allow the resumption of replication. Consistent with these ideas, several investigators have shown that primary fibroblasts from BS, WS, and RTS patients and RecQ5-deficient mouse embryonic fibroblasts all show differential hypersensitivity to agents that perturb DNA replication (12, 14, 26, 29). Moreover, BLM and WRN are recruited to DNA replication forks after replicative stress, and DNA fiber track analyses have shown that both BLM and WRN are required for normal fork progression after DNA damage or replication arrest (11-13, 47, 54). In particular, BLM in conjunction with DNA topoisomerase III and two other accessory proteins, RMI-1 and RMI-2, has been shown to catalyze the resolution of double-Holliday-junction recombination intermediates to generate noncrossover products. This dissolution reaction could play an important role in the error-free recombinational repair of damaged or stalled forks during S phase (57, 67). WRN also appears to promote error-free repair by contributing to the resolution of gene conversion events to generate noncrossover products (46). In line with the above observations, WRN and BLM can be found associated with replication foci or other DNA damage response proteins in damaged cells. In contrast, in unperturbed cells, a majority of each protein is found in the nucleolus (WRN) or associated with PML bodies (BLM) (5, 37, 62).RECQ4 has also been implicated in DNA replication. Recent studies have shown that hypomorphic mutants of the Drosophila melanogaster homolog of human RECQ4, DmRECQ4, have reduced DNA replication-dependent chorion gene amplification (65). These findings are thus consistent with a postulated role for Xenopus laevis RECQ4 (XRECQ4) in the initiation of DNA replication (39, 48). The N terminus of XRECQ4 bears homology to the N termini of the yeast proteins Sld2 (Saccharomyces cerevisiae [budding yeast]) and DRC1 (Schizosaccharomyces pombe [fission yeast]), which play a central role, in association with budding yeast Dpb11 and the fission yeast homolog Cut5/Rad4, in the establishment of DNA replication forks (38, 41, 63). Consistently, the N terminus of XRECQ4 has been shown to interact with the X. laevis variant of Cut5, and XRECQ4 depletion severely perturbs DNA replication initiation in X. laevis egg extracts (39, 48). The notion that the function of XRECQ4 is evolutionarily conserved in mammals is supported by the observations that the human protein can complement its Xenopus counterpart in cell-free assays for replication initiation and that depletion of human RECQ4 inhibits cellular proliferation and DNA synthesis (39, 48). Moreover, deletion of the N-terminal region of mouse RECQ4 has been shown to be an embryonic lethal mutation (27). These observations suggest that vertebrate RECQ4 might be a functional homolog of Sld2/DRC11, although its precise function during replication initiation and progression is not known. Recent results, published while this work was in progress, indicate that human RECQ4 interacts with the MCM replicative complex during replication initiation and that this interaction is regulated by CDK phosphorylation of RECQ4 (69). These findings, together with our results below, provide clues to the mechanism regulating RECQ4 interaction with the replication machinery.RECQ1 is the most abundant of the human RecQ helicases and was the first of the human RecQ proteins to be discovered on the basis of its potent ATPase activity (50). Despite this, little is known about the cellular functions of RECQ1, and no human disease associations have been identified to date. Recent studies have shown that RECQ1 is involved in the maintenance of genome integrity and that RECQ1 depletion affects cellular proliferation (51). Moreover, biochemical studies have shown that RECQ1 and BLM display distinct substrate specificities, indicating that these helicases are likely to perform nonoverlapping functions (43). These results suggest an important—though as yet mechanistically ill-defined—role for RECQ1 in cell cycle progression and/or DNA repair (52).In order to better delineate the role of human RecQ helicases in DNA replication, we investigated the in vivo interactions of all five human RecQ enzymes with three well-characterized human DNA replication origins in quantitative chromatin immunoprecipitation (ChIP) assays. We also determined how nascent-origin-dependent DNA synthesis, chromatin binding of replication proteins, origin firing frequency, and replication fork rates were altered by depleting specific human RecQ helicase proteins. We found that only two of the five human RecQ helicases, RECQ1 and RECQ4, specifically interact with origins in unperturbed cells. Our results provide new mechanistic insight into the distinct roles of human RECQ1 and RECQ4 in DNA replication initiation and in replication fork progression.  相似文献   

10.
11.
Mrc1 plays a role in mediating the DNA replication checkpoint. We surveyed replication elongation proteins that interact directly with Mrc1 and identified a replicative helicase, Mcm6, as a specific Mrc1-binding protein. The central portion of Mrc1, containing a conserved coiled-coil region, was found to be essential for interaction with the 168-amino-acid C-terminal region of Mcm6, and introduction of two amino acid substitutions in this C-terminal region abolished the interaction with Mrc1 in vivo. An mcm6 mutant bearing these substitutions showed a severe defect in DNA replication checkpoint activation in response to stress caused by methyl methanesulfonate. Interestingly, the mutant did not show any defect in DNA replication checkpoint activation in response to hydroxyurea treatment. The phenotype of the mcm6 mutant was suppressed when the mutant protein was physically fused with Mrc1. These results strongly suggest for the first time that an Mcm helicase acts as a checkpoint sensor for methyl methanesulfonate-induced DNA damage through direct binding to the replication checkpoint mediator Mrc1.Progression of the DNA replication machinery along chromosomes is a complex process. Replication forks pause occasionally when they encounter genomic regions that are difficult to replicate, such as highly transcribed regions, tRNA genes, and regions with specialized chromatin structure, like centromeric and heterochromatic regions (17). Replication forks also stall when treated with chemicals like methyl methanesulfonate (MMS), which causes DNA damage, or hydroxyurea (HU), which limits the cellular concentration of the deoxynucleoside triphosphate pool (17). Because de novo assembly and programming of the replisome do not occur after the onset of S phase (18), DNA replication forks must be protected from replicative stresses. The DNA replication checkpoint constitutes a surveillance mechanism for S-phase progression that safeguards replication forks from various replicative stresses (22, 38, 40), and malfunction of this checkpoint leads to chromosome instability and cancer development in higher organisms (4, 9).The Saccharomyces cerevisiae DNA replication checkpoint mediator Mrc1 is functionally conserved and is involved directly in DNA replication as a component of the replisome (1, 8, 16, 19, 29, 30). Mrc1, together with Tof1 and Csm3, is required for forming a replication pausing complex when the fork is exposed to replicative stress by HU (16). The pausing complex subsequently triggers events leading to DNA replication checkpoint activation and hence stable replicative arrest. A sensor kinase complex, Mec1-Ddc2 (ATR-ATRIP homolog of higher eukaryotes), is then recruited to the complex (14, 16). Mec1-Ddc2-mediated phosphorylation of Mrc1 activates the pausing complex, and phosphorylated Mrc1 likely recruits Rad53 (a putative homolog of CHK2 of higher eukaryotes), which is then activated via phosphorylation by Mec1-Ddc2 (1, 16, 20, 30). Activated Rad53 subsequently elicits a stress responses, i.e., stabilization of replication forks, induction of repair genes, and suppression of late-firing origins (24). It remains unclear, however, whether DNA replication checkpoint activation is induced in response to DNA damage by MMS, a reagent commonly used to study the DNA replication stress response. Several lines of evidence have suggested that MMS-induced damage is also sensed directly by the replication machinery (38, 40).Although biochemical and genetic interaction data have placed Mrc1 at the center of the replication checkpoint signal transduction cascade, its molecular function remains largely unknown. The proteins Mrc1, Tof1, and Csm3 associate with the Mcm complex (8, 27), a heterohexameric DNA helicase consisting of Mcm2 to Mcm7 proteins which unwinds the parental DNA duplex to allow replisome progression (3, 12, 18, 31, 32, 35). The Mcm complex associates with a specific set of regulatory proteins at forks to form replisome progression complexes (8). In addition to Mcm, Tof1, Csm3, and Mrc1, replisome progression complexes include factors such as Cdc45 and the GINS complex that are also required for fork progression (13, 26, 31, 32, 39). Claspin, a putative Xenopus laevis homolog of Mrc1, is also reported to associate with Cdc45, DNA polymerase ɛ (Polɛ), replication protein A, and two of the replication factor C complexes in aphidicolin-treated Xenopus egg extracts (19). Recently, Mrc1 was reported to interact directly with Polɛ (23).The aim of this study was to provide mechanistic insight into Mrc1 function in the DNA replication checkpoint. For this purpose, it was essential to identify, among all the essential proteins in the replication machinery, a specific protein that interacts with Mrc1 and to examine the role of this interaction in the DNA replication checkpoint. We found that Mrc1 interacts with Mcm6 directly and specifically. When the interaction between Mrc1 and Mcm6 was impaired, cells no longer activated the DNA replication checkpoint in response to MMS-induced replicative stress. Interestingly and unexpectedly, this interaction was not required for DNA replication checkpoint activation in response to HU-induced replicative stress. Our results provide the first mechanistic evidence that cells use separate mechanisms to transmit replicative stresses caused by MMS and HU for DNA replication checkpoint activation.  相似文献   

12.
13.
The β clamp is an essential replication sliding clamp required for processive DNA synthesis. The β clamp is also critical for several additional aspects of DNA metabolism, including DNA mismatch repair (MMR). The dnaN5 allele of Bacillus subtilis encodes a mutant form of β clamp containing the G73R substitution. Cells with the dnaN5 allele are temperature sensitive for growth due to a defect in DNA replication at 49°C, and they show an increase in mutation frequency caused by a partial defect in MMR at permissive temperatures. We selected for intragenic suppressors of dnaN5 that rescued viability at 49°C to determine if the DNA replication defect could be separated from the MMR defect. We isolated three intragenic suppressors of dnaN5 that restored growth at the nonpermissive temperature while maintaining an increase in mutation frequency. All three dnaN alleles encoded the G73R substitution along with one of three novel missense mutations. The missense mutations isolated were S22P, S181G, and E346K. Of these, S181G and E346K are located near the hydrophobic cleft of the β clamp, a common site occupied by proteins that bind the β clamp. Using several methods, we show that the increase in mutation frequency resulting from each dnaN allele is linked to a defect in MMR. Moreover, we found that S181G and E346K allowed growth at elevated temperatures and did not have an appreciable effect on mutation frequency when separated from G73R. Thus, we found that specific residue changes in the B. subtilis β clamp separate the role of the β clamp in DNA replication from its role in MMR.Replication sliding clamps are essential cellular proteins imparting a spectacular degree of processivity to DNA polymerases during genome replication (24, 39-41). Encoded by the dnaN gene, the β clamp is a highly conserved bacterial sliding clamp found in virtually all eubacterial species (reviewed in reference 7). The β clamp is a head-to-tail, ring-shaped homodimer that encircles double-stranded DNA (1, 39). In eukaryotes and archaea, the analog of the β clamp is proliferating cell nuclear antigen (PCNA) (15, 28, 40, 41). Eukaryotic PCNA is a ring-shaped homotrimer that also acts to encircle DNA, increasing the processivity of the replicative DNA polymerases (40, 41). Although the primary structures of the β clamp and PCNA are not conserved, the tertiary structures of these proteins are very similar, demonstrating structural conservation among bacterial, archaeal, and eukaryotic replication sliding clamps (28, 39-41; reviewed in reference 6).The function of the β clamp is not limited to its well-defined role in genome replication. The Escherichia coli β clamp binds Hda, which also binds the replication initiation protein DnaA, regulating the active form of DnaA complexed with ATP (19, 37, 43). This allows the β clamp to regulate replication initiation through the amount of available DnaA-ATP. In Bacillus subtilis, the β clamp binds YabA, a negative regulator of DNA replication initiation (12, 29, 52). It has also been suggested that the B. subtilis β clamp sequesters DnaA from the replication origin during the cell cycle through the binding of DnaA to YabA and the binding of YabA to the β clamp (70). Thus, it is hypothesized that in E. coli and B. subtilis, the β clamp influences the frequency of replication initiation through interactions with Hda and YabA, respectively.The E. coli and B. subtilis β clamp has an important role in translesion DNA synthesis during the replicative bypass of noncoding bases by specialized DNA polymerases belonging to the Y family (20, 33). The roles of the E. coli β clamp in translesion synthesis are well established (5, 8, 30, 31). Binding sites on the E. coli β clamp that accommodate translesion polymerases pol IV (DinB) and pol V (UmuD2′C) have been identified, and the consequence of disrupting their association with the β clamp has illustrated the critical importance of the β clamp to the activity of both of these polymerases (4, 5, 8, 26, 30, 31, 48, 49).In addition to the involvement of the β clamp in replication initiation, DNA replication, and translesion synthesis, the E. coli and B. subtilis β clamp also functions in DNA mismatch repair (MMR) (45, 46, 64). The MMR pathway recognizes and repairs DNA polymerase errors, contributing to the overall fidelity of the DNA replication pathway (reviewed in references 42 and 60). In both E. coli and B. subtilis, deletion of the genes mutS and mutL increases the spontaneous mutation frequency several hundredfold (13, 25, 63). In E. coli, MutS recognizes and binds mismatches, while MutL functions as a “matchmaker,” coordinating the actions of other proteins in the MMR pathway, allowing the removal of the mismatch and resynthesis of the resulting gap (reviewed in references 42 and 60). MutS and MutL of E. coli and B. subtilis physically interact with the β clamp (45, 46, 51, 64). Interaction between the B. subtilis β clamp and MutS is important for efficient MMR and organization of MutS-green fluorescent protein (GFP) into foci in response to replication errors, while the function of MutL binding to the β clamp is unknown (64).These studies show that the β clamp is critical for several aspects of DNA metabolism in E. coli and B. subtilis. In E. coli, many dnaN alleles have been examined and used to define the mechanistic roles of the β clamp in vivo (5, 18, 24, 30, 31, 48, 49, 73). A limitation in studying the mechanistic roles of the B. subtilis β clamp is that only two dnaN alleles (β clamp) are available, dnaN5 and dnaN34 (36) (www.bgsc.org/), and both of these alleles do not support growth at temperatures above 49°C, suggesting that they may cause similar defects (36) (www.bgsc.org/). Of these two dnaN alleles, only dnaN5 has been investigated in any detail (36, 53, 64). The mutant β clamp encoded by dnaN5 contains a G73R substitution [dnaN5(G73R)] in a surface-exposed residue located on the outside rim of the β clamp (53, 64). Our previous studies with this allele showed that dnaN5(G73R) confers an increase in mutation frequency at 30°C and 37°C (64). Further characterization of dnaN5(G73R) showed that the increased mutation frequency is caused by a partial defect in MMR (64). Additionally, dnaN5(G73R)-containing cells have a reduced ability to support MutS-GFP focus formation in response to mismatches (64). These results support the hypothesis that G73R in the β clamp causes a defect in DNA replication at 49°C (36) and impaired MMR manifested by a defect in establishing the assembly of MutS-GFP foci in response to replication errors (64).To understand the roles of the B. subtilis β clamp in MMR and DNA replication, we examined the dnaN5 and dnaN34 alleles. We found that the nucleotide sequences of dnaN5 and dnaN34 and the phenotypes they produce were identical, both producing the G73R missense mutation. We analyzed in vivo β clampG73R protein levels and found that the β clampG73R protein accumulated to wild-type levels at elevated temperatures. To identify amino acid residues that would restore DNA replication at elevated temperatures, we isolated three intragenic suppressors of dnaN5(G73R) that conferred growth of B. subtilis cells at 49°C. Epistasis analysis and determination of the mutation spectrum showed that each dnaN allele isolated in this study caused an MMR-dependent increase in mutation frequency. Additionally, we found that the β clamp binding protein YabA can reduce the efficiency of MMR in vivo when yabA expression is induced. Thus, we have identified residues in the β clamp that are critical for DNA replication and MMR in B. subtilis. We also found that a β clamp binding protein, YabA, can reduce the efficiency of MMR in vivo.  相似文献   

14.
15.
The activation of the human polyomavirus BK causes polyomavirus-associated nephropathy in immunocompromised humans. Studies of the virus have been restricted since the virus DNA replication is species specific. Cell-based and cell-free DNA replication systems, including the BK virus (BKV) monopolymerase DNA replication system using purified proteins, reproduce the species specificity (28). Therefore, the major host proteins comprising this assay, DNA polymerase α-primase (Pol-prim) and replication protein A (RPA), were intensively studied here. We demonstrate that Pol-prim plays a major role in the species specificity of BKV DNA replication. Both large subunits p180 and p68 of the enzyme complex have central functions in modulating the host specificity. Recently, an inhibitory activity of BKV DNA replication was described (C. Mahon, B. Liang, I. Tikhanovich, J. R. Abend, M. J. Imperiale, H. P. Nasheuer, and W. R. Folk, J. Virol. 83:5708-5717, 2009), but neither mouse Pol-prim nor mouse RPA diminishes cell-free BKV DNA replication. However, the inhibition of BKV DNA replication in mouse extracts depends on sequences flanking the core origin. In the presence of human Pol-prim, the inhibitory effect of mouse cell factors is abolished with plasmid DNAs containing the murine polyomavirus early promoter region, whereas the late enhancer region and the core origin are supplied from BKV. Thus, BKV replication is regulated by both Pol-prim, as a core origin species-specific factor, and inhibitory activities, as origin-flanking sequence-dependent factor(s).BK virus (BKV) is a human polyomavirus that was first isolated in the 1970s (15). Up to 90% of adults have serologic evidence of exposure to BKV, but in most humans the virus remains latent (25, 26). Almost all disease accompanied by BKV reactivation has been found in immunocompromised patients (22). In recent years, BKV has been associated with nephropathy (polyomavirus-associated nephropathy, or PVAN) in up to 10% of renal transplant patients. Once established, the disease results in allograft loss in 45 to 70% of the patients (18). Importantly, BKV preferentially replicates in human cells and less well in cells of other primates, and the virus is highly tumorigenic in rodents (21, 41, 44). This fact and the lack of sustainable viral replication in rodents or other convenient, experimental animal models have been an enormous setback to the study of PVAN.As with other members of the Polyomaviridae family, BKV virions are nonenveloped icosahedral particles with a diameter of 45 nm that contain a circular double-stranded DNA genome of 5.3 kb (1). In BKV and in other polyomaviruses, three genomic areas have been distinguished: (i) a noncoding control region including the origin of viral DNA replication, (ii) the early genes encoding large and small T antigens (TAgs), and (iii) the late genes which code for the capsid proteins VP-1, VP-2, and VP-3 and the agnoprotein (22).BKV DNA replication is similar to that of all other members of the Polyomaviridae family and requires only one viral protein, the multifunctional large TAg, whereas all other replication factors are supplied by the host (13, 14, 28, 39, 47). As the first step, TAg binds to the core origin, which contains the early palindrome, an AT-rich sequence, and the TAg binding site II, which consists of two pairs of G(G/A)GGC pentanucleotides. In the presence of ATP, TAg forms a double hexamer and partially melts the early palindrome (EP) and untwists the AT-rich sequence of the BKV core origin (5, 6, 14). Then the TAg double hexamers bidirectionally unwind the viral replication origin, which requires ATP hydrolysis. In the following process the two hexamers remain associated with each other, with the separated single-stranded DNA (ssDNA) threading through the hexameric channels (14). The viral core origin is sufficient to constitute a functional replication origin, but the presence of auxiliary domains increases its activity 5- to 100-fold in vivo (16, 30). After the viral TAg unwinds the core origin and its flanking sequences, replication protein A (RPA), the main eukaryotic ssDNA-binding protein, covers the resulting stretches of ssDNA, whereas topoisomerase I releases the resulting torsional stress and enhances initiation of DNA replication (5, 7, 43). Then, DNA polymerase α-primase (Pol-prim) is loaded onto this TAg-RPA-topoisomerase 1-DNA complex, yielding a functional initiation complex. In the following step, Pol-prim synthesizes short RNA primers at the origin, and these RNA primers are elongated by the DNA polymerase function of the enzyme complex (9, 35, 47). After a polymerase switch from Pol-prim to DNA polymerase δ (Pol δ) with the help of RPA, replication factor C (RFC), and proliferating cell nuclear antigen (PCNA), processive DNA synthesis is completed by Pol δ in association with PCNA, the sliding clamp, on the leading strand (38, 51, 54, 59). Lagging-strand synthesis is discontinuous, and multiple initiation events catalyzed by Pol-prim must take place. Again, after the elongation of the RNA primers by Pol-prim, DNA synthesis is switched to Pol δ, which then synthesizes the complete Okazaki fragments. The maturation of these Okazaki fragments requires the collaboration of RNase H, PCNA, flap endonuclease 1 (Fen-1), Pol δ, and DNA ligase I to establish a continuous strand also on the lagging strand (9, 19, 20, 51, 55).TAg functions in infected cells rely heavily on specific associations with host proteins; for example, TAg interacts with RPA, Pol-prim, and topoisomerase I to replicate viral DNA. Selective interactions with the host p180 and p48 subunits of Pol-prim were shown to be responsible for species-specific replication of simian virus 40 (SV40) and murine polyomavirus (mPyV) DNAs, respectively (8, 47, 50). The subunits of Pol-prim are highly conserved since 88, 80, 89, and 90% of the amino acids are identical between human and murine p180, p68, p58, and p48, respectively. Biochemical studies have shown that TAg interacts independently with all four subunits of Pol-prim (8, 12, 57). Moreover, the p180, p58, and p48 subunits of Pol-prim also physically bind to RPA (7, 11, 57). RPA and TAg binding sites in the Pol-prim complex are essential for SV40 DNA replication in vitro since the presence of an excess of these purified binding peptides diminishes viral DNA replication in vitro (52, 53). Interestingly, species specificity requires the viral origin of DNA replication, whereas physical protein-protein interactions of purified protein complexes are not host specific in the absence of viral origin DNA (29, 42).Consistent with other polyomaviruses, analyses of BKV TAg-dependent DNA replication recently revealed that BKV DNA cannot be replicated in murine cells and that cell extracts are able to mimic this behavior (28). Furthermore, a BKV DNA replication system with the purified human proteins Pol-prim, RPA, topoisomerase I, and BKV TAg was inhibited by murine extracts, whereas SV40 DNA replication was not. Further investigations revealed that the presence of inhibitory activities (IAs) in extracts from murine cells blocks BKV DNA replication at an early step of TAg-mediated unwinding of the BKV origin of replication. Detailed analyses using the BKV monopolymerase DNA replication system, which we report here, show that Pol-prim functions as a species-specific factor associated with core origin functions. In addition, we reveal that the inhibitory activities in murine extracts, which are associated with origin-flanking sequence-dependent factor(s), regulate BKV DNA replication in murine cell extracts in a Pol-prim-independent manner.  相似文献   

16.
We successfully substituted Escherichia coli''s origin of replication oriC with the origin region of Vibrio cholerae chromosome I (oriCIVc). Replication from oriCIVc initiated at a similar or slightly reduced cell mass compared to that of normal E. coli oriC. With respect to sequestration-dependent synchrony of initiation and stimulation of initiation by the loss of Hda activity, replication initiation from oriC and oriCIVc were similar. Since Hda is involved in the conversion of DnaAATP (DnaA bound to ATP) to DnaAADP (DnaA bound to ADP), this indicates that DnaA associated with ATP is limiting for V. cholerae chromosome I replication, which similar to what is observed for E. coli. No hda homologue has been identified in V. cholerae yet. In V. cholerae, dam is essential for viability, whereas in E. coli, dam mutants are viable. Replacement of E. coli oriC with oriCIVc allowed us to specifically address the role of the Dam methyltransferase and SeqA in replication initiation from oriCIVc. We show that when E. coli''s origin of replication is substituted by oriCIVc, dam, but not seqA, becomes important for growth, arguing that Dam methylation exerts a critical function at the origin of replication itself. We propose that Dam methylation promotes DnaA-assisted successful duplex opening and replisome assembly at oriCIVc in E. coli. In this model, methylation at oriCIVc would ease DNA melting. This is supported by the fact that the requirement for dam can be alleviated by increasing negative supercoiling of the chromosome through oversupply of the DNA gyrase or loss of SeqA activity.The genomes of Vibrio cholerae and several related Vibrio spp. are distributed between two circular chromosomes. Characterization of the origins of replication of V. cholerae chromosomes I and II (oriCIVc and oriCIIVc, respectively) has shown that oriCIVc is similar to the origin of replication of the Escherichia coli chromosome, oriC, whereas oriCIIVc is completely different (20). Like oriC, oriCIVc has five R-type DnaA boxes (53) as well as boxes conforming to the I and τ types (52, 61), and the DnaA protein is the rate-limiting factor in the initiation of replication in both cases (18). In E. coli, DnaA associates with both ATP and ADP, and the ATP-bound form is absolutely required for initiation to take place (reviewed in reference 60). When reaching a critical level, DnaAATP (DnaA bound to ATP) protein is proposed to form a helical filament, anchored at one or more R-boxes (54, 69), in which origin DNA wraps around the outside of the DnaA core (21) or where the DnaA wraps around oriC (61). In both cases, the topology of the DnaA-oriC nucleoprotein complex leads to formation of compensatory negative supercoiling that facilitates unwinding of the adjacent AT-rich region resulting in initiation. In both models, DnaAATP is absolutely required for initiation, and in agreement with this, DnaAATP was found to be the rate-limiting factor for initiation in vivo (69).The V. cholerae oriCIVc also resembles oriC in having many potential sites for methylation by DNA adenine methyltransferase (Dam), although the number and position of the GATC sites differ slightly (see Fig. Fig.1).1). The role of Dam in initiation of chromosome replication has been studied mainly in E. coli. After initiation of DNA replication has occurred on a fully methylated oriC, the newly replicated hemimethylated origins are sequestered from the Dam methyltransferase and from reinitiation for approximately one-third of a doubling time. During this time interval, the activity and amount of DnaA available for initiation are reduced to prevent immediate reinitiation (reviewed in references 57 and 83). The sequestration is carried out by the SeqA protein that binds hemimethylated oriC GATC sequences with high affinity (48). In the absence of Dam methylation or SeqA, the same origin can be reinitiated in the same cell cycle, and initiations become asynchronous (9, 48).Open in a separate windowFIG. 1.Alignment of the E. coli minimal oriC with the corresponding region from V. cholerae chromosome I. The AT-rich sequence and the three 13-mer repeats L, M, and R found in E. coli (5) are indicated above the alignment. The 6-mer (A/T)GATCT boxes (80) are underlined. Other DnaA binding sites, i.e., R-boxes (53), I-boxes (52), and τ-boxes (61), are shown as boxed regions. Dam methylation sites (GATC) are shaded gray. The experimentally defined binding sites for integration host factor (IHF) (22) and factor for inversion stimulation (FIS) (65) in E. coli are indicated, and bases that match the consensus sequence are in boldface type. The single base difference between oriCIVc and oriCIVc* (see Materials and Methods) in the minimal origin region is shown below the two sequences. A gap introduced to maximize alignment of the two sequences is indicated by a dash in the sequence. Nucleotides that are identical in the two sequences are indicated by an asterisk below the two sequences.Genes encoding a Dam homologue and a SeqA homologue are present on Vibrio genomes, but there appear to be some differences between the functions of the proteins in E. coli and V. cholerae. dam has been found to be an essential gene in V. cholerae (33, 15), which is not the case in E. coli (48, 51). Conflicting data exist concerning the essentiality of seqA in V. cholerae (15, 72). The roles of Dam and SeqA in oriCIVc replication have been studied using minichromosomes, i.e., plasmids replicating exclusively from a cloned copy of oriCIVc (20). oriCIVc-based minichromosomes can replicate in wild-type E. coli cells but were unable to replicate in dam, seqA, and seqA dam mutants (20). The extrachromosomal existence of minichromosomes is dependent on their ability to initiate replication in synchrony with the chromosomal origin (46, 75). In E. coli cells mutated in dam or seqA, incompatibility exists between the oriC carried on minichromosomes and that of the chromosome due to origin competition (13), and when minichromosomes are maintained under selective pressure, they integrate into the origin region of the host chromosome (46, 75). Minichromosomes based on oriCIVc may also compete with the E. coli oriC for initiations in dam or seqA mutant cells. However, due to limited sequence identity, they may not be able to integrate into the E. coli chromosome. This could provide an explanation for the failure to introduce oriCIVc minichromosomes into dam and seqA mutant cells (20). Both dam and seqA genes could therefore be required for viability of V. cholerae for reasons not related to chromosome replication. In addition to its role in DNA replication, roles for Dam methylation in gene regulation and DNA repair have also been demonstrated in a number of bacteria (for reviews, see references 11, 45, 47, and 50). For V. cholerae as well as for Salmonella spp. and Yersinia pseudotuberculosis, Dam plays a role in virulence possibly through regulation of virulence gene expression (33). Less is known about the functions of seqA apart from its role in E. coli replication, but it has been suggested that SeqA functions as a nucleoid-organizing protein (for a review, see reference 83), and the E. coli chromosome has been demonstrated to have increased supercoiling in a seqA strain (85).Here we describe the first in vivo evidence that Dam plays an important role in the initiation of replication by facilitating the replication initiation at oriCIVc in E. coli. In addition, we show that SeqA does not carry an essential role in the initiation of replication.  相似文献   

17.
The Asf1 and Rad6 pathways have been implicated in a number of common processes such as suppression of gross chromosomal rearrangements (GCRs), DNA repair, modification of chromatin, and proper checkpoint functions. We examined the relationship between Asf1 and different gene products implicated in postreplication repair (PRR) pathways in the suppression of GCRs, checkpoint function, sensitivity to hydroxyurea (HU) and methyl methanesulfonate (MMS), and ubiquitination of proliferating cell nuclear antigen (PCNA). We found that defects in Rad6 PRR pathway and Siz1/Srs2 homologous recombination suppression (HRS) pathway genes suppressed the increased GCR rates seen in asf1 mutants, which was independent of translesion bypass polymerases but showed an increased dependency on Dun1. Combining an asf1 deletion with different PRR mutations resulted in a synergistic increase in sensitivity to chronic HU and MMS treatment; however, these double mutants were not checkpoint defective, since they were capable of recovering from acute treatment with HU. Interestingly, we found that Asf1 and Rad6 cooperate in ubiquitination of PCNA, indicating that Rad6 and Asf1 function in parallel pathways that ubiquitinate PCNA. Our results show that ASF1 probably contributes to the maintenance of genome stability through multiple mechanisms, some of which involve the PRR and HRS pathways.DNA replication must be highly coordinated with chromatin assembly and cell division for correct propagation of genetic information and cell survival. Errors arising during DNA replication are corrected through the functions of numerous pathways including checkpoints and a diversity of DNA repair mechanisms (32, 33, 35). However, in the absence of these critical cellular responses, replication errors can lead to the accumulation of mutations and gross chromosomal rearrangements (GCRs) as well as chromosome loss, a condition generally termed genomic instability (33). Genome instability is a hallmark of many cancers as well as other human diseases (24). There are many mechanisms by which GCRs can arise, and over the last few years numerous genes and pathways have been implicated in playing a role in the suppression of GCRs in Saccharomyces cerevisiae and in some cases in the etiology of cancer (27, 28, 33, 39-47, 51, 53, 56, 58, 60), including S. cerevisiae ASF1, which encodes the main subunit of the replication coupling assembly factor (37, 62).Asf1 is involved in the deposition of histones H3 and H4 onto newly synthesized DNA during DNA replication and repair (62), and correspondingly, asf1 mutants are sensitive to chronic treatment with DNA-damaging agents (2, 30, 62). However, asf1 mutants do not appear to be repair defective and can recover from acute treatment with at least some DNA-damaging agents (2, 8, 30, 31, 54), properties similar to those described for rad9 mutants (68). In the absence of Asf1, both the DNA damage and replication checkpoints become activated during normal cell growth, and in the absence of checkpoint execution, there is a further increase in checkpoint activation in asf1 mutants (30, 46, 54). It has been suggested that asf1 mutants are defective for checkpoint shutoff and that this might account for the increased steady-state levels of checkpoint activation seen in asf1 mutants (8); however, another study has shown that asf1 mutants are not defective for checkpoint shutoff and that in fact Asf1 and the chromatin assembly factor I (CAF-I) complex act redundantly or cooperate in checkpoint shutoff (31). Furthermore, Asf1 might be involved in proper activation of the Rad53 checkpoint protein, as Asf1 physically interacts with Rad53 and this interaction is abrogated in response to exogenous DNA damage (15, 26); however, the physiological relevance of this interaction is unclear. Asf1 is also required for K56 acetylation of histone H3 by Rtt109, and both rtt109 mutants and histone H3 variants that cannot be acetylated (38) share many of the properties of asf1 mutants, suggesting that at least some of the requirement for Asf1 in response to DNA damage is mediated through Rtt109 (11, 14, 22, 61). Subsequent studies of checkpoint activation in asf1 mutants have led to the hypothesis that replication coupling assembly factor defects result in destabilization of replication forks which are then recognized by the replication checkpoint and stabilized, suggesting that the destabilized replication forks account for both the increased GCRs and increased checkpoint activation seen in asf1 mutants (30). This hypothesis is supported by other recent studies implicating Asf1 in the processing of stalled replication forks (16, 57). This role appears to be independent of CAF-I, which can cooperate with Asf1 in chromatin assembly (63). Asf1 has also been shown to function in disassembly of chromatin, suggesting other possibilities for the mechanism of action of Asf1 at the replication fork (1, 2, 34). Thus, while Asf1 is thought to be involved in progression of the replication fork, both the mechanism of action and the factors that cooperate with Asf1 in this process remain obscure.Stalled replication forks, particularly those that stall at sites of DNA damage, can be processed by homologous recombination (HR) (6) or by a mechanism known as postreplication repair (PRR) (reviewed in reference 67). There are two PRR pathways, an error-prone pathway involving translesion synthesis (TLS) by lower-fidelity polymerases and an error-free pathway thought to involve template switching (TS) (67). In S. cerevisiae, the PRR pathways are under the control of the RAD6 epistasis group (64). The error-prone pathway depends on monoubiquitination of proliferating cell nuclear antigen (PCNA) on K164 by Rad6 (an E2 ubiquitin-conjugating enzyme) by Rad18 (E3 ubiquitin ligase) (23). This results in replacement of the replicative DNA polymerase with nonessential TLS DNA polymerases, such as REV3/REV7-encoded DNA polymerase ζ (polζ) and RAD30-encoded DNA polη, which can bypass different types of replication-blocking damage (67). The error-free pathway is controlled by Rad5 (E3) and a complex consisting of Ubc13 and Mms2 (E2 and E2 variant, respectively), which add a K63-linked polyubiquitin chain to monoubiquitinated PCNA, leading to TS to the undamaged nascent sister chromatid (4, 25, 65). Furthermore, in addition to modification with ubiquitin, K164 of PCNA can also be sumoylated by Siz1, resulting in subsequent recruitment of the Srs2 helicase and inhibition of deleterious Rad51-dependent recombination events (50, 52, 55), although it is currently unclear if these are competing PCNA modifications or if both can exist on different subunits in the same PCNA trimer. A separate branch of the Rad6 pathway involving the E3 ligase Bre1 monoubiquitinates the histone H2B (29, 69) as well as Swd2 (66), which stimulates Set1-dependent methylation of K4 and Dot1-dependent methylation of K79 of histone H3 (48, 49, 66). Subsequently, K79-methylated H3 recruits Rad9 and activates the Rad53 checkpoint (19, 70). Activation of Rad53 is also bolstered by Rad6-Rad18-dependent ubiquitination of Rad17, which is part of the 9-1-1 complex that functions upstream in the checkpoint pathway (17). Finally, Rad6 complexes with the E3 Ubr1, which mediates protein degradation by the N-end rule pathway (13).Due to the role of the PRR pathways at stalled replication forks and a recent study implicating the Rad6 pathway in the suppression of GCRs (39), we examined the relationship between these ubiquitination and sumoylation pathways and the Asf1 pathway in order to gain additional insights into the function of Asf1 during DNA replication and repair. Our findings suggest that Asf1 has multiple functions that prevent replication damage or act in the cellular responses to replication damage and that these functions are modified by and interact with the PRR pathways. The TLS PRR pathway does not appear to be involved, and both a Dun1-dependent replication checkpoint and HR are important for preventing the deleterious effects of PRR and Asf1 pathway defects. We hypothesize that this newly observed cooperation between Asf1 and the PRR pathways may be required for resolving stalled replication forks, leading to suppression of GCRs and successful DNA replication.  相似文献   

18.
Porcine circovirus type 1 (PCV1), originally isolated as a contaminant of PK-15 cells, is nonpathogenic, whereas porcine circovirus type 2 (PCV2) causes an economically important disease in pigs. To determine the factors affecting virus replication, we constructed chimeric viruses by swapping open reading frame 1 (ORF1) (rep) or the origin of replication (Ori) between PCV1 and PCV2 and compared the replication efficiencies of the chimeric viruses in PK-15 cells. The results showed that the replication factors of PCV1 and PCV2 are fully exchangeable and, most importantly, that both the Ori and rep of PCV1 enhance the virus replication efficiencies of the chimeric viruses with the PCV2 backbone.Porcine circovirus (PCV) is a single-stranded DNA virus in the family Circoviridae (34). Type 1 PCV (PCV1) was discovered in 1974 as a contaminant of porcine kidney cell line PK-15 and is nonpathogenic in pigs (31-33). Type 2 PCV (PCV2) was discovered in piglets with postweaning multisystemic wasting syndrome (PMWS) in the mid-1990s and causes porcine circovirus-associated disease (PCVAD) (1, 9, 10, 25). PCV1 and PCV2 have similar genomic organizations, with two major ambisense open reading frames (ORFs) (16). ORF1 (rep) encodes two viral replication-associated proteins, Rep and Rep′, by differential splicing (4, 6, 21, 22). The Rep and Rep′ proteins bind to specific sequences within the origin of replication (Ori) located in the intergenic region, and both are responsible for viral replication (5, 7, 8, 21, 23, 28, 29). ORF2 (cap) encodes the immunogenic capsid protein (Cap) (26). PCV1 and PCV2 share approximately 80%, 82%, and 62% nucleotide sequence identity in the Ori, rep, and cap, respectively (19).In vitro studies using a reporter gene-based assay system showed that the replication factors of PCV1 and PCV2 are functionally interchangeable (2-6, 22), although this finding has not yet been validated in a live infectious-virus system. We have previously shown that chimeras of PCV in which cap has been exchanged between PCV1 and PCV2 are infectious both in vitro and in vivo (15), and an inactivated vaccine based on the PCV1-PCV2 cap (PCV1-cap2) chimera is used in the vaccination program against PCVAD (13, 15, 18, 27).PCV1 replicates more efficiently than PCV2 in PK-15 cells (14, 15); thus, we hypothesized that the Ori or rep is directly responsible for the differences in replication efficiencies. The objectives of this study were to demonstrate that the Ori and rep are interchangeable between PCV1 and PCV2 in a live-virus system and to determine the effects of swapped heterologous replication factors on virus replication efficiency in vitro.  相似文献   

19.
20.
Nitrofurazone is reduced by cellular nitroreductases to form N2-deoxyguanine (N2-dG) adducts that are associated with mutagenesis and lethality. Much attention recently has been given to the role that the highly conserved polymerase IV (Pol IV) family of polymerases plays in tolerating adducts induced by nitrofurazone and other N2-dG-generating agents, yet little is known about how nitrofurazone-induced DNA damage is processed by the cell. In this study, we characterized the genetic repair pathways that contribute to survival and mutagenesis in Escherichia coli cultures grown in the presence of nitrofurazone. We find that nucleotide excision repair is a primary mechanism for processing damage induced by nitrofurazone. The contribution of translesion synthesis to survival was minor compared to that of nucleotide excision repair and depended upon Pol IV. In addition, survival also depended on both the RecF and RecBCD pathways. We also found that nitrofurazone acts as a direct inhibitor of DNA replication at higher concentrations. We show that the direct inhibition of replication by nitrofurazone occurs independently of DNA damage and is reversible once the nitrofurazone is removed. Previous studies that reported nucleotide excision repair mutants that were fully resistant to nitrofurazone used high concentrations of the drug (200 μM) and short exposure times. We demonstrate here that these conditions inhibit replication but are insufficient in duration to induce significant levels of DNA damage.Replication in the presence of DNA damage is thought to produce most of the mutagenesis, genomic rearrangements, and lethality that occur in all cells. UV-induced photoproducts, X-ray-induced strand breaks, psoralen- or cis-platin-interstrand cross-links, oxidized bases from reactive oxygen species, and base depurination are just a few of the structurally distinct challenges that the replication machinery must overcome. It seems likely that the mechanisms that process these lesions will vary depending on the nature of the impediment.While a number of the lesions described above are known to block replication, the events associated with UV-induced damage have been the most extensively characterized. UV irradiation causes the formation of cyclobutane pyrimidine dimers and 6-4 photoproducts in DNA that block the progression of the replication fork (16, 29, 30, 37). Following the arrest of replication at UV-induced damage, RecA and several RecF pathway proteins are required to process the replication fork such that the blocking lesion is removed or bypassed (2, 5, 6, 8-10). Cells lacking either RecA or any of several RecF pathway proteins are hypersensitive to UV-induced damage and fail to recover replication following disruption by the lesions (2, 6, 10). RecBCD is an exonuclease/helicase complex that is involved in repairing double-strand breaks (38). It also is required for resistance to UV-induced damage, although it is not required to process or restore disrupted replication forks, and the substrates it acts upon after UV irradiation currently remain unclear (3, 10, 19).Survival and the ability to resume DNA synthesis following UV-induced damage depend predominantly on the removal of the lesions by nucleotide excision repair (5, 7, 36). Cells deficient in nucleotide excision repair are unable to remove UV-induced DNA lesions and exhibit elevated levels of mutagenesis, strand exchanges, rearrangements, and cell lethality (16, 33, 34). In cases where replication fork processing or lesion repair is prevented, the recovery of replication and survival become entirely dependent on translesion synthesis by DNA polymerase V (Pol V) (6). However, in repair-proficient cells, the contribution of translesion synthesis to recovery and survival is minor and is detected only following UV doses that exceed the repair capacity of the cell (5, 6).Less is known about how replication recovers from other forms of DNA damage. We chose to characterize nitrofurazone, because a number of studies suggested that N2-deoxyguanine (N2-dG) adducts induced by this and other agents would be processed differently than UV-induced lesions. Nitrofurazone is a topical antibacterial agent that historically has been used for treating burns and skin grafts in patients and animals (14, 15, 32). Nitrofurazone toxicity is known to require activation by cellular nitroreductases (25, 42). However, the mechanism and targets of its antimicrobial properties have yet to be fully elucidated. In addition to its antimicrobial properties, the reduced nitrofurazone metabolites also target DNA and have been shown to induce free radical damage, strand breaks, and N2-dG adducts (26, 40, 42, 45), and they are mutagenic and carcinogenic in rodent models (1, 15, 24, 39).Whereas nucleotide excision repair is the predominant mechanism required for survival after UV-induced damage, a number of studies suggest that translesion synthesis plays a larger role in survival after nitrofurazone-induced DNA damage. dinB mutants lacking Pol IV were shown to be hypersensitive to nitrofurazone compared to cells that constitutively express the polymerase (17). Biochemically, Pol IV and a number of Pol IV homologs from other organisms have been shown to efficiently replicate over a range of N2-dG adducts in vitro (17, 35, 44). In addition, several studies have reported that uvrA mutants, which are defective in nucleotide excision repair, do not exhibit any hypersensitivity to nitrofurazone or other agents that induce similar adducts in vivo (12, 21, 27). Early studies also observed a direct correlation between nitrofurazone-induced mutations and lethality, suggesting that mutagenic lesions persist in the DNA to cause toxicity (21, 23, 27, 43). Consistent with these observations, nitrofuran-induced lesions were found to be poor substrates for nucleotide excision repair in vitro (46).Taken together, these observations suggest to us that the cellular response to nitrofurazone will be distinct from its response to UV irradiation. However, no study has examined the relative contributions that nucleotide excision repair, translesion synthesis, or recombination has in recovering from nitrofurazone-induced damage. In this study, we characterized the mechanism by which nitrofurazone inhibits DNA replication and identified the genes that contribute to the recovery, survival, and mutagenesis of Escherichia coli treated with nitrofurazone. In contrast to previous studies, we found that survival following nitrofurazone-induced damage depends predominantly on nucleotide excision repair. Similarly to UV-induced DNA damage, both the RecF and RecBC pathways contribute to survival following nitrofurazone-induced DNA damage. The contribution of translesion polymerases to survival was minor and was mediated by Pol IV. In addition, we found that nitrofurazone can act to inhibit DNA replication directly when used at higher concentrations. The direct inhibition of replication is reversible and occurs independently of DNA damage, suggesting that DNA is not the primary target of its antimicrobial properties.  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