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.
The parvovirus adeno-associated virus (AAV) contains a small single-stranded DNA genome with inverted terminal repeats that form hairpin structures. In order to propagate, AAV relies on the cellular replication machinery together with functions supplied by coinfecting helper viruses such as adenovirus (Ad). Here, we examined the host cell response to AAV replication in the context of Ad or Ad helper proteins. We show that AAV and Ad coinfection activates a DNA damage response (DDR) that is distinct from that seen during Ad or AAV infection alone. The DDR was also triggered when AAV replicated in the presence of minimal Ad helper proteins. We detected autophosphorylation of the kinases ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) and signaling to downstream targets SMC1, Chk1, Chk2, H2AX, and XRCC4 and multiple sites on RPA32. The Mre11 complex was not required for activation of the DDR to AAV infection. Additionally, we found that DNA-PKcs was the primary mediator of damage signaling in response to AAV replication. Immunofluorescence revealed that some activated damage proteins were found in a pan-nuclear pattern (phosphorylated ATM, SMC1, and H2AX), while others such as DNA-PK components (DNA-PKcs, Ku70, and Ku86) and RPA32 accumulated at AAV replication centers. Although expression of the large viral Rep proteins contributed to some damage signaling, we observed that the full response required replication of the AAV genome. Our results demonstrate that AAV replication in the presence of Ad helper functions elicits a unique damage response controlled by DNA-PK.Replication of viral genomes produces a large amount of extrachromosomal DNA that may be recognized by the cellular DNA damage machinery. This is often accompanied by activation of DNA damage response (DDR) signaling pathways and recruitment of cellular repair proteins to sites of viral replication. Viruses therefore provide good model systems to study the recognition and response to DNA damage (reviewed in reference 48). The Mre11/Rad50/Nbs1 (MRN) complex functions as a sensor of chromosomal DNA double-strand breaks (DSBs) and is involved in activation of damage signaling (reviewed in reference 41). The MRN complex also localizes to DNA DSBs and is found at viral replication compartments during infection with a number of DNA viruses (6, 40, 47, 70, 75, 77, 87, 93). The phosphatidylinositol 3-kinase-like kinases (PIKKs) ataxia telangiectasia mutated (ATM), ATM and Rad3-related kinase (ATR), and the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) are involved in the signal transduction cascades activated by DNA damage (reviewed in references 43, 51, and 71). These kinases respond to distinct types of damage and regulate DSB repair during different phases of the cell cycle (5), either through nonhomologous end-joining (NHEJ) or homologous recombination pathways (reviewed in references 63, 81, and 86). The DNA-PK holoenzyme is composed of DNA-PKcs and two regulatory subunits, the Ku70 and Ku86 heterodimer. DNA-PK functions with XRCC4/DNA ligase IV to repair breaks during NHEJ, and works with Artemis to process DNA hairpin structures during VDJ recombination and during a subset of DNA DSB events (46, 50, 86). While the kinase activity of DNA-PKcs leads to phosphorylation of a large number of substrates in vitro as well as autophosphorylation of specific residues (reviewed in references 16 and 85), it is currently unclear how DNA-PKcs contributes to signaling in cells upon different types of damage.The adeno-associated virus (AAV) genome consists of a molecule of single-stranded DNA with inverted terminal repeats (ITRs) at both ends that form double-hairpin structures due to their palindromic sequences (reviewed in reference 52). The ITRs are important for replication and packaging of the viral genome and for integration into the host genome. Four viral Rep proteins (Rep78, Rep68, Rep52, and Rep40) are also required for replication and packaging of the AAV genome into virions assembled from the Cap proteins. Although the Rep and Cap genes are replaced in recombinant AAV vectors (rAAV) that retain only the ITRs flanking the gene of interest, these vectors can be replicated by providing Rep in trans (reviewed in reference 7). Productive AAV infection requires helper functions supplied by adenovirus (Ad) or other viruses such as herpes simplex virus (HSV) (reviewed in reference 27), together with components of the host cell DNA replication machinery (54, 55, 58). In the presence of helper viruses or minimal helper proteins from Ad or HSV, AAV replicates in the nucleus at centers where the viral DNA and Rep proteins accumulate (35, 76, 84, 89). Cellular and viral proteins involved in AAV replication, including replication protein A (RPA), Ad DNA-binding protein (DBP), and HSV ICP8, localize with Rep proteins at these viral centers (29, 33, 76).A number of published reports suggest associations between AAV and the cellular DNA damage machinery. For example, transduction by rAAV vectors is increased by genotoxic agents and DNA damaging treatments (1, 62, 91) although the mechanisms involved remain unclear. Additionally, the ATM kinase negatively regulates rAAV transduction (64, 92), and we have shown that the MRN complex poses a barrier to both rAAV transduction and wild-type AAV replication (11, 67). UV-inactivated AAV particles also appear to activate a DDR involving ATM and ATR kinases that perturbs cell cycle progression (39, 60, 88). It has been suggested that this response is provoked by the AAV ITRs (60) and that UV-treated particles mimic stalled replication forks in infected cells (39). In addition to AAV genome components, the viral Rep proteins have been observed to exhibit cytotoxicity and induce S-phase arrest (3, 65).The role of cellular repair proteins in AAV genome processing has also been explored by examining the molecular fate of rAAV vectors, which are converted into circular and concatemeric forms that persist episomally (18, 19, 66). Proteins shown to regulate circularization in cell culture include ATM and the MRN complex (14, 64), while in vivo experiments using mouse models have implicated ATM and DNA-PK in this process (14, 20, 72). Additionally, DNA-PKcs and Artemis have recently been shown to cleave the ITR hairpins of rAAV vectors in vivo in a tissue-dependent manner (36). Despite these studies, it is not clear how damage response factors function together and how they impact AAV transduction and replication in human cells.In this study we examined the cellular response to AAV replication in the context of Ad infection or helper proteins. We show that coinfection with AAV and Ad activates a DDR that is distinct from that seen during infection with Ad alone. The ATM and DNA-PKcs damage kinases are activated and signal to downstream substrates, but the response does not require the MRN complex and is primarily mediated by DNA-PKcs. Although expression of the large Rep proteins induced some DDR events, full signaling appeared to require AAV replication and was accompanied by accumulation of DNA-PK at viral replication compartments. Our results demonstrate that AAV replication induces a unique DNA damage signal transduction response and provides a model system for studying DNA-PK.  相似文献   

3.
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   

4.
5.
6.
7.
Adeno-associated virus (AAV) type 2 and 5 proteins Rep52 and Rep40 were polyubiquitinated during AAV-adenovirus type 5 (Ad5) coinfection and during transient transfection in either the presence or absence of Ad5 E4orf6 and E1b-55k. Polyubiquitination of small Rep proteins via lysine 48 (K48) linkages, normally associated with targeting of proteins for proteasomal degradation, was detected only in the presence of E4orf6. The small Rep proteins were ubiquitinated via lysine 63 (K63) following transfection in either the presence or absence of E4orf6 or following coinfection with Ad5. E4orf6/E1b-55k-dependent K48-specific polyubiquitination of small Rep proteins could be inhibited using small interfering RNA (siRNA) to cullin 5.Together, adenovirus type 5 (Ad5) early gene products E1a, E1b-55k, E2a, E4orf6, and virus-associated (VA) RNA can support efficient replication of adeno-associated virus (AAV) (4, 31). E4orf6 and E1b-55k are known to interact with cellular cullin 5 (cul5), elongins B and C, and the ring box protein Rbx1 to form an E3 ubiquitin ligase complex that specifically targets a small population of cellular proteins for degradation by the proteasome (1, 7, 21, 22, 24, 27). This property has been implicated in a number of functions presumed to be required for both Ad and AAV replication (3, 8-10, 17, 23, 24, 34, 35).Previously, only p53, Mre11, DNA ligase IV, and integrin α3 had been shown to be substrates of the Ad5 E3 ubiquitin ligase complex (1, 7, 21, 22, 24, 27); however, we have recently shown (16, 17) that the small Rep proteins and capsid proteins of AAV5 are also degraded in the presence of Ad E4orf6 and E1b-55k in a proteasome-dependent manner. These proteins were restored to levels required during infection by the action of VA RNA (17). The targeting for degradation of AAV5 protein by the E4orf6/E1b-55k E3 ubiquitin ligase complex required functional BC-box motifs in E4orf6 and could be inhibited by depletion of the scaffolding protein cullin 5 using directed small interfering RNA (siRNA) (16). In addition, the degradation of AAV5 protein was partially prevented by overexpression of pUBR7, a plasmid that generates a dominant-negative ubiquitin (16). The role this targeted degradation plays in the life cycle of AAV has not yet been clarified; however, E4orf6 mutants that cannot function in this regard do not support AAV replication as well as wild-type E4orf6 (R. Nayak and D. J. Pintel, unpublished data). Degradation of Mre11 by the Ad5 E3 ligase has also been implicated in allowing efficient Ad5 and AAV replication (24). Ubiquitination of AAV Rep proteins during viral infection, however, has not previously been reported.  相似文献   

8.
9.
10.
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   

11.
12.
13.
14.
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   

15.
16.
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   

17.
18.
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   

19.
20.
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