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 46 毫秒
1.
2.
3.
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   

4.
Although the herpes simplex virus type 1 (HSV-1) genome might be expected to induce a DNA damage response, the ATR kinase is not activated in infected cells. We previously proposed that spatial uncoupling of ATR from its interaction partner, ATRIP, could be the basis for inactivation of the ATR kinase in infected cells; however, we now show that ATR and ATRIP are in fact both recruited to HSV-1 replication compartments and can be coimmunoprecipitated from infected-cell lysates. ATRIP and replication protein A (RPA) are recruited to the earliest detectable prereplicative sites, stage II microfoci. In a normal cellular DNA damage response, ATR/ATRIP are recruited to stretches of RPA-coated single-stranded DNA in an RPA- and kinase-dependent manner, resulting in the phosphorylation of RPA by ATR in damage foci. In contrast, in HSV-1-infected cells, RPA is not phosphorylated, and endogenous phosphorylated RPA is excluded from stage II microfoci; in addition, the recruitment of ATR/ATRIP is independent of RPA and the kinase activity of ATR. Furthermore, we show that ATR/ATRIP play a beneficial role in viral gene expression and virus production. Although ICP0 has been shown to be important for partial inactivation of other cellular DNA repair pathways, we show that ICP0 is not responsible for the inactivation of ATR signaling and, furthermore, that neither ATR nor ATRIP is a target of ICP0 degradation. Thus, ATR and ATRIP may function outside the context of the canonical ATR damage signaling pathway during HSV-1 infection to participate in the viral life cycle.Herpes simplex virus type 1 (HSV-1) is a large linear double-stranded DNA virus that replicates in the nucleus of the host cell. The incoming viral genome contains nicks and gaps (42), and cellular DNA repair machinery might be expected to recognize it as damaged, resulting in the activation of one or more cellular DNA damage pathways. Activation of DNA damage response pathways can result not only in repair of the damaged DNA but also in cell cycle arrest, gene silencing, and apoptosis (9). The later outcomes could result in suppression of viral gene expression and DNA replication and thus have negative consequences for lytic infection. Activation of a cellular DNA damage response during viral infection could, therefore, represent a form of intrinsic antiviral immunity (14, 15). On the other hand, HSV-1 and other DNA viruses which replicate in the nucleus have also been shown to utilize cellular DNA repair machinery to promote productive infection (28). Thus, HSV-1 has apparently evolved to manipulate the host DNA damage response by utilizing some components and inactivating others in an attempt to create an environment conducive to lytic viral infection.The cellular DNA damage response is regulated by the three phosphoinositide 3-kinase-related kinases (PIKKs), DNA-PK (DNA-dependent protein kinase), ATM (ataxia-telangiectasia-mutated), and ATR (ATM and Rad3-related) (1, 9). DNA-PK and ATM respond predominantly to double-strand breaks, and ATR responds to stalled replication forks and long stretches of single-stranded DNA (ssDNA). DNA-PK is required for nonhomologous end joining (NHEJ), while ATM activation promotes homologous recombination. Interestingly, in some cell types, the catalytic subunit of DNA-PK (DNA-PKcs) is proteolytically degraded during infection by the immediate-early (IE) protein ICP0, a viral E3 ubiquitin ligase (25, 37), thereby resulting in the probable inactivation of the NHEJ pathway. ATM kinase activity, on the other hand, is activated during HSV-1 infection once viral DNA replication is initiated (26, 47, 56). Despite phosphorylation of several ATM targets, ATM signaling is also modulated by ICP0, which degrades the ubiquitin ligases RNF8 and RNF168. The function of these ubiquitin ligases is to promote the tethering of ATM pathway proteins at sites of cellular DNA damage (27). Thus, ICP0 functions to partially inactivate portions of both the DNA-PK- and ATM-mediated repair pathways.During a cellular DNA damage response, ATM activation and processing of DNA ends generate ssDNA adjacent to double-stranded DNA (dsDNA), a structure that is known to activate ATR (9, 38). The ssDNA is coated by the cellular ssDNA binding protein, replication protein A (RPA), which then serves to recruit ATR through a direct interaction with ATR-interacting protein (ATRIP) (4, 12, 58). ATR signaling results in the phosphorylation of many substrates, including RPA and Chk1. During HSV-1 infection, the ATR substrates RPA and Chk1 are not phosphorylated (47, 54-56), indicating that ATR signaling may be disabled.A hallmark of HSV-1 infection is the reorganization of the infected-cell nucleus, resulting in the formation of large globular replication compartments as well as the rearrangement of cellular proteins involved in several homeostatic pathways. In addition to cellular DNA repair proteins, HSV-1 infection also causes the reorganization of components of the cellular protein quality control pathways, resulting in the formation of virus-induced chaperone-enriched (VICE) domains, which act to maintain nuclear protein quality control during infection (31). Viral gene expression, DNA replication, and encapsidation of viral genomes occur in replication compartments (24, 39, 41). In this work we revisit the study of proteins recruited to and restricted from replication compartments in an attempt to better understand how HSV-1 manipulates components of the cellular DNA damage response for its own benefit.  相似文献   

5.
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   

6.
7.
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   

8.
Adenovirus infection activates cellular DNA damage response and repair pathways. Viral proteins that are synthesized before viral DNA replication prevent recognition of viral genomes as a substrate for DNA repair by targeting members of the sensor complex composed of Mre11/Rad50/NBS1 for degradation and relocalization, as well as targeting the effector protein DNA ligase IV. Despite inactivation of these cellular sensor and effector proteins, infection results in high levels of histone 2AX phosphorylation, or γH2AX. Although phosphorylated H2AX is a characteristic marker of double-stranded DNA breaks, this modification was widely distributed throughout the nucleus of infected cells and was coincident with the bulk of cellular DNA. H2AX phosphorylation occurred after the onset of viral DNA replication and after the degradation of Mre11. Experiments with inhibitors of the serine-threonine kinases ataxia telangiectasia mutated (ATM), AT- and Rad3-related (ATR), and DNA protein kinase (DNA-PK), the kinases responsible for H2AX phosphorylation, indicate that H2AX may be phosphorylated by ATR during a wild-type adenovirus infection, with some contribution from ATM and DNA-PK. Viral DNA replication appears to be the stimulus for this phosphorylation event, since infection with a nonreplicating virus did not elicit phosphorylation of H2AX. Infected cells also responded to high levels of input viral DNA by localized phosphorylation of H2AX. These results are consistent with a model in which adenovirus-infected cells sense and respond to both incoming viral DNA and viral DNA replication.Cellular DNA damage response pathways protect and preserve the integrity of the genome. These pathways, which are activated in response to various forms of DNA damage, involve a number of proteins that participate in both DNA repair and cell cycle progression (62). The serine-threonine kinases ataxia telangiectasia mutated (ATM), AT- and Rad3-related (ATR), and DNA protein kinase (DNA-PK) are activated in response to distinct types of damage. The ATM pathway is activated primarily by double-stranded DNA breaks (4, 30). DNA-PK acts in conjunction with the DNA ligase IV/XRCC4 complex to mediate the ligation of double-stranded breaks through nonhomologous end joining (34). The ATR pathway can be activated in response to a wide range of genotoxic stresses, such as base or nucleotide excision, double-stranded breaks, or single-stranded breaks. Activation of ATR is generally thought to occur via the recognition of single-stranded tracks of DNA (63). Each of these pathways leads to the phosphorylation and activation of a number of cellular proteins such as the variant histone H2AX, checkpoint kinases 1 and 2 (Chk1 and Chk2), and Nijmegen break syndrome protein 1 (NBS1), among others (62). Signals transmitted by a cascade of phosphorylation events result in cell cycle arrest and the accumulation of repair protein complexes at sites of DNA damage.Upon recognition of a double-stranded DNA break by the cell, H2AX is phosphorylated on an extended C-terminal tail at serine 139 by the phosphatidylinositol 3-kinase (PI3K)-related kinases ATM, ATR, and DNA-PK (9, 41, 44, 58). Considered one of the earliest indications of a double-stranded DNA break, phosphorylated H2AX (γH2AX) acts as a scaffolding protein to which a number of DNA repair factors can dock to facilitate repair of the damaged DNA (36, 42, 53). Areas of phosphorylated H2AX, termed γH2AX foci, are enriched for proteins involved in both homologous recombination and nonhomologous end joining, such as NBS1, BRCA1 (42), and Mdc1 (24, 50).Although adenovirus is able to activate both ATM and ATR pathways (11), adenoviral proteins limit the extent and consequences of signaling through these pathways. The E1B-55K and E4orf6 proteins form an E3 ubiquitin ligase with the cellular proteins Cullin-5, elongins B and C, and Rbx1 (28, 43). This complex targets key cellular proteins involved in cellular response to DNA damage, including p53 (28, 43), Mre11 (51), and DNA ligase IV (3). The E4orf3 gene product targets cellular proteins central to both the cellular DNA damage response and the antiviral response. The E4orf3 protein of species C adenoviruses alters the localization of Mre11/Rad50/NBS1 (MRN) complex members within the nucleus to prevent association with centers of viral DNA replication and to ensure efficient viral DNA replication (17, 18, 52). In addition, these three viral early proteins direct members of the MRN complex (2, 35) and the single-stranded DNA-binding protein 2 (20) to cytoplasmic aggresomes, where these sequestered proteins are effectively inactivated. These viral activities, along with the inactivation of DNA-PK by E4orf3 and E4orf6 gene products (7), appear to prevent recognition of viral genomes by the MRN complex and prevent ligation of these genomes through nonhomologous end joining. In cells infected with a virus with E4 deleted, Mre11 physically binds to viral DNA in an NBS1-dependent manner and may prevent efficient genome replication (37). The overlapping means by which adenovirus disables the MRN complex and prevents DNA damage repair serves to illustrate the importance of this activity for a productive adenovirus infection. However, despite having DNA damage signaling and DNA repair pathways dismantled, adenovirus-infected cells exhibit some characteristic changes associated with DNA damage signaling events, such as the phosphorylation of H2AX (6, 15). Thus, it appears that adenovirus effectively inhibits DNA repair activity but may not fully suppress the early events of DNA damage signaling.The focus of the present study was to elucidate the activation of DNA damage signaling pathways revealed by phosphorylation of the variant histone H2AX during wild-type adenovirus infection and to determine what stage of the virus life cycle leads to this activation. We demonstrate that infected cells respond to viral genome replication with high levels of H2AX phosphorylation throughout the cell nucleus. This phosphorylation event is not localized to viral replication centers and does not appear to be concurrent with cellular double-stranded DNA breaks; rather, H2AX phosphorylation occurs coincident with the bulk of cellular chromatin. H2AX phosphorylation follows viral DNA replication and reaches peak levels after the degradation of the Mre11. In addition, we observed that infected cells can respond to both the presence of incoming viral genomes and genome replication by initiating H2AX phosphorylation.  相似文献   

9.
10.
11.
12.
13.
Controlling the loading of Rad51 onto DNA is important for governing when and how homologous recombination is used. Here we use a combination of genetic assays and indirect immunofluorescence to show that the F-box DNA helicase (Fbh1) functions in direct opposition to the Rad52 orthologue Rad22 to curb Rad51 loading onto DNA in fission yeast. Surprisingly, this activity is unnecessary for limiting spontaneous direct-repeat recombination. Instead it appears to play an important role in preventing recombination when replication forks are blocked and/or broken. When overexpressed, Fbh1 specifically reduces replication fork block-induced recombination, as well as the number of Rad51 nuclear foci that are induced by replicative stress. These abilities are dependent on its DNA helicase/translocase activity, suggesting that Fbh1 exerts its control on recombination by acting as a Rad51 disruptase. In accord with this, overexpression of Fbh1 also suppresses the high levels of recombinant formation and Rad51 accumulation at a site-specific replication fork barrier in a strain lacking the Rad51 disruptase Srs2. Similarly overexpression of Srs2 suppresses replication fork block-induced gene conversion events in an fbh1Δ mutant, although an inability to suppress deletion events suggests that Fbh1 has a distinct functionality, which is not readily substituted by Srs2.Homologous recombination (HR) is often described as a double-edged sword: it can maintain genome stability by promoting DNA repair, while its injudicious action can disturb genome stability by causing gross chromosome rearrangement (GCR) or loss of heterozygosity (LOH). Both GCR and LOH are potential precursors of diseases such as cancer, and consequently there is need to control when and how HR is used.A key step in most HR is the loading of the Rad51 recombinase onto single-stranded DNA (ssDNA), which forms a nucleoprotein filament (nucleofilament) that catalyzes the pairing of homologous DNAs and subsequent strand invasion (32). This is a critical point at which recombination can be regulated through the removal of the Rad51 filament (60). Early removal can prevent strand invasion altogether, freeing the DNA for alternative processing. Later removal may limit unnecessary filament growth, free the 3′-OH of the invading strand to prime DNA synthesis, and ultimately enable ejection of the invading strand, which is important for the repair of double-strand breaks (DSBs) by synthesis-dependent strand annealing (SDSA). SDSA avoids the formation of Holliday junctions that can be resolved into reciprocal exchange products (crossovers), which may result in GCR or LOH if the recombination is ectopic or allelic, respectively.One enzyme that appears to be able to control Rad51 in the aforementioned manner is the yeast superfamily 1 DNA helicase Srs2 (42). In Saccharomyces cerevisiae, Srs2 is recruited to stalled replication forks by the SUMOylation of PCNA, and there it appears to block Rad51-dependent HR in favor of Rad6- and Rad18-dependent postreplication repair (1, 2, 35, 50, 53, 58). In vitro Srs2 can strip Rad51 from ssDNA via its DNA translocase activity (31, 62) and therefore probably controls HR at stalled replication forks by acting as a Rad51 disruptase. In accord with this, chromatin immunoprecipitation analysis has shown that Rad51 is enriched at or near replication forks in an srs2 mutant (50). Srs2 also plays an important role in crossover avoidance during DSB repair, where it is thought to promote SDSA by both disrupting Rad51 nucleofilaments and dissociating displacement (D) loops (20, 27).Srs2 is conserved in the fission yeast Schizosaccharomyces pombe (19, 43, 63) and has a close relative in bacteria called UvrD, which can similarly control HR by disrupting RecA nucleofilaments (61). However, an obvious homologue in mammals has not been detected. Recently, two mammalian members of the RecQ DNA helicase family, BLM and RECQL5, were shown to disrupt Rad51 nucleofilaments in vitro (11, 25), although in the case of BLM, this activity appears to be relatively weak (5, 55). Nevertheless these data have led to speculation that both BLM and RECQL5 might perform a function similar to that of Srs2 in vivo (6). Certainly mutational inactivation of either helicase results in elevated levels of HR and genome instability, with an associated increased rate of cancer (23, 25). However, BLM and RECQL5 are not the only potential Rad51 disruptases in mammals; a relative of Srs2 and UvrD called FBH1 was recently implicated in this role by genetic studies of its orthologue in S. pombe and by its ability to partially compensate for the loss of Srs2 in S. cerevisiae, which, unlike S. pombe, lacks an FBH1 orthologue (15). FBH1 is so named because of an F box near its N terminus—a feature that makes it unique among DNA helicases (28). The F box is important for its interaction with SKP1 and therefore the formation of an E3 ubiquitin ligase SCF (SKP1-Cul1-F-box protein) complex (29). The targets of this complex are currently unknown. In S. pombe, mutations within Fbh1''s F-box block interaction with Skp1 and prevent Fbh1 from localizing to the nucleus and forming damage-induced foci therein (57). Fbh1''s role in constraining Rad51 activity in S. pombe is evidenced by the increase in spontaneous Rad51 foci and accumulation of UV irradiation-induced Rad51-dependent recombination intermediates in an fbh1Δ mutant (47). Moreover, loss of both Fbh1 and Srs2 in S. pombe results in a synergistic reduction in cell viability, and like Srs2, Fbh1 is essential for viability in the absence of the S. pombe RecQ family DNA helicase Rqh1, which processes recombination intermediates (47, 48). In both cases the synthetic interaction is suppressed by deleting rad51, suggesting that Fbh1 works in parallel with Srs2 and Rqh1 to prevent the formation of toxic recombination intermediates. In yeast, Rad51-mediated recombination is dependent on Rad52 (Rad22 in S. pombe), which is believed to promote the nucleation of Rad51 onto DNA that is coated with the ssDNA binding protein replication protein A (RPA) (18, 32). Intriguingly, the genotoxin sensitivity and recombination deficiency of a rad22 mutant are suppressed in a Rad51-dependent manner by deleting fbh1 (48). This suggests that Fbh1 and Rad22 act in opposing ways to modulate the assembly of the Rad51 nucleofilament. Although current data indicate a role for Fbh1 in controlling HR, the only evidence so far that Fbh1 limits recombinant formation is in chicken DT40 cells, for which a modest increase in sister chromatid exchange has been noted when FBH1 is deleted (30).Here we present in vivo evidence suggesting that Fbh1 does indeed act as a Rad51 disruptase, which is dependent on its DNA helicase/translocase activity. We confirm predictions that this activity works in opposition to Rad22 for the loading of Rad51 onto DNA and show that Fbh1''s modulation of Rad51 activity, while not essential for limiting spontaneous direct-repeat recombination, is critical for preventing recombination at blocked replication forks. Finally, we highlight similarities and differences between Fbh1 and Srs2, based on their mutant phenotypes and relative abilities to suppress recombination when overexpressed. Overall our data affirm that Fbh1 is one of the principal modulators of Rad51 activity in fission yeast and therefore may play a similar role in vertebrates.  相似文献   

14.
Sister chromatid recombination (SCR) is a potentially error-free pathway for the repair of DNA lesions associated with replication and is thought to be important for suppressing genomic instability. The mechanisms regulating the initiation and termination of SCR in mammalian cells are poorly understood. Previous work has implicated all the Rad51 paralogs in the initiation of gene conversion and the Rad51C/XRCC3 complex in its termination. Here, we show that hamster cells deficient in the Rad51 paralog XRCC2, a component of the Rad51B/Rad51C/Rad51D/XRCC2 complex, reveal a bias in favor of long-tract gene conversion (LTGC) during SCR. This defect is corrected by expression of wild-type XRCC2 and also by XRCC2 mutants defective in ATP binding and hydrolysis. In contrast, XRCC3-mediated homologous recombination and suppression of LTGC are dependent on ATP binding and hydrolysis. These results reveal an unexpectedly general role for Rad51 paralogs in the control of the termination of gene conversion between sister chromatids.DNA double-strand breaks (DSBs) are potentially dangerous lesions, since their misrepair may cause chromosomal translocations, gene amplifications, loss of heterozygosity (LOH), and other types of genomic instability characteristic of human cancers (7, 9, 21, 40, 76, 79). DSBs are repaired predominantly by nonhomologous end joining or homologous recombination (HR), two evolutionarily conserved DSB repair mechanisms (8, 12, 16, 33, 48, 60, 71). DSBs generated during the S or G2 phase of the cell cycle may be repaired preferentially by HR, using the intact sister chromatid as a template for repair (12, 26, 29, 32, 71). Sister chromatid recombination (SCR) is a potentially error-free pathway for the repair of DSBs, which has led to the proposal that SCR protects against genomic instability, cancer, and aging. Indeed, a number of human cancer predisposition genes are implicated in SCR control (10, 24, 45, 57, 75).HR entails an initial processing of the DSB to generate a free 3′ single-stranded DNA (ssDNA) overhang (25, 48, 56). This is coupled to the loading of Rad51, the eukaryotic homolog of Escherichia coli RecA, which polymerizes to form an ssDNA-Rad51 “presynaptic” nucleoprotein filament. Formation of the presynaptic filament is tightly regulated and requires the concerted action of a large number of gene products (55, 66, 68). Rad51-coated ssDNA engages in a homology search by invading homologous duplex DNA. If sufficient homology exists between the invading and invaded strands, a triple-stranded synapse (D-loop) forms, and the 3′ end of the invading (nascent) strand is extended, using the donor as a template for gene conversion. This recombination intermediate is thought to be channeled into one of the following two major subpathways: classical gap repair or synthesis-dependent strand annealing (SDSA) (48). Gap repair entails the formation of a double Holliday junction, which may resolve into either crossover or noncrossover products. Although this is a major pathway in meiotic recombination, crossing-over is highly suppressed in somatic eukaryotic cells (26, 44, 48). Indeed, the donor DNA molecule is seldom rearranged during somatic HR, suggesting that SDSA is the major pathway for the repair of somatic DSBs (26, 44, 49, 69). SDSA terminates when the nascent strand is displaced from the D-loop and pairs with the second end of the DSB to form a noncrossover product. The mechanisms underlying displacement of the nascent strand are not well understood. However, failure to displace the nascent strand might be expected to result in the production of longer gene conversion tracts during HR (36, 44, 48, 63).Gene conversion triggered in response to a Saccharomyces cerevisiae or mammalian chromosomal DSB generally results in the copying of a short (50- to 300-bp) stretch of information from the donor (short-tract gene conversion [STGC]) (14, 47, 48, 67, 69). A minority of gene conversions in mammalian cells entail more-extensive copying, generating gene conversion tracts that are up to several kilobases in length (long-tract gene conversion [LTGC]) (26, 44, 51, 54, 64). In yeast, very long gene conversions can result from break-induced replication (BIR), a highly processive form of gene conversion in which a bona fide replication fork is thought to be established at the recombination synapse (11, 36, 37, 39, 61, 63). In contrast, SDSA does not require lagging-strand polymerases and appears to be much less processive than a conventional replication fork (37, 42, 78). BIR in yeast has been proposed to play a role in LOH in aging yeast, telomere maintenance, and palindromic gene amplification (5, 41, 52). It is unclear to what extent a BIR-like mechanism operates in mammalian cells, although BIR has been invoked to explain telomere elongation in tumors lacking telomerase (13). It is currently unknown whether LTGC and STGC in somatic mammalian cells are products of mechanistically distinct pathways or whether they represent alternative outcomes of a common SDSA pathway.Vertebrate cells contain five Rad51 paralogs—polypeptides with limited sequence homology to Rad51—Rad51B, Rad51C, Rad51D, XRCC2, and XRCC3 (74). The Rad51 paralogs form the following two major complexes: Rad51B/Rad51C/Rad51D/XRCC2 (BCDX2) and Rad51C/XRCC3 (CX3) (38, 73). Genetic deletion of any one of the rad51 paralogs in the mouse germ line produces early embryonic lethality, and mouse or chicken cells lacking any of the rad51 paralogs reveal hypersensitivity to DNA-damaging agents, reduced frequencies of HR and of sister chromatid exchanges, increased chromatid-type errors, and defective sister chromatid cohesion (18, 72, 73, 82). Collectively, these data implicate the Rad51 paralogs in SCR regulation. The purified Rad51B/Rad51C complex has been shown to assist Rad51-mediated strand exchange (62). XRCC3 null or Rad51C null hamster cells reveal a bias toward production of longer gene conversion tracts, suggesting a role for the CX3 complex in late stages of SDSA (6, 44). Rad51C copurifies with branch migration and Holliday junction resolution activities in mammalian cell extracts (35), and XRCC3, but not XRCC2, facilitates telomere shortening by reciprocal crossing-over in telomeric T loops (77). These data, taken together with the meiotic defects observed in Rad51C hypomorphic mice, suggest a specialized role for CX3, but not for BCDX2, in resolving Holliday junction structures (31, 58).To further address the roles of Rad51 paralogs in late stages of recombination, we have studied the balance between long-tract (>1-kb) and short-tract (<1-kb) SCR in XRCC2 mutant hamster cells. We found that DSB-induced gene conversion in both XRCC2 and XRCC3 mutant cells is biased in favor of LTGC. These defects were suppressed by expression of wild-type (wt) XRCC2 or XRCC3, respectively, although the dependence upon ATP binding and hydrolysis differed between the two Rad51 paralogs. These results indicate that Rad51 paralogs play a more general role in determining the balance between STGC and LTGC than was previously appreciated and suggest roles for both the BCDX2 and CX3 complexes in influencing the termination of gene conversion in mammals.  相似文献   

15.
16.
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   

17.
18.
19.
20.
The initiator protein E1 from human papillomavirus (HPV) is a helicase essential for replication of the viral genome. E1 contains three functional domains: a C-terminal enzymatic domain that has ATPase/helicase activity, a central DNA-binding domain that recognizes specific sequences in the origin of replication, and a N-terminal region necessary for viral DNA replication in vivo but dispensable in vitro. This N-terminal portion of E1 contains a conserved nuclear export signal (NES) whose function in the viral life cycle remains unclear. In this study, we provide evidence that nuclear export of HPV31 E1 is inhibited by cyclin E/A-Cdk2 phosphorylation of two serines residues, S92 and S106, located near and within the E1 NES, respectively. Using E1 mutant proteins that are confined to the nucleus, we determined that nuclear export of E1 is not essential for transient viral DNA replication but is important for the long-term maintenance of the HPV episome in undifferentiated keratinocytes. The findings that E1 nuclear export is not required for viral DNA replication but needed for genome maintenance over multiple cell divisions raised the possibility that continuous nuclear accumulation of E1 is detrimental to cellular growth. In support of this possibility, we observed that nuclear accumulation of E1 dramatically reduces cellular proliferation by delaying cell cycle progression in S phase. On the basis of these results, we propose that nuclear export of E1 is required, at least in part, to limit accumulation of this viral helicase in the nucleus in order to prevent its detrimental effect on cellular proliferation.Human papillomaviruses (HPV) are small double-stranded DNA viruses that infect keratinocytes of the differentiating epithelium of the skin or mucosa (reviewed in references 4 and 63). Of more than 150 different HPV types identified thus far, about 25 infect the anogenital region (9). The low-risk types, such as HPV11 and HPV6, are associated with the development of genital warts, while the high-risk types, such as HPV16, -18, and -31, cause high-grade lesions that can progress to invasive cervical carcinoma (17, 38, 61).The HPV life cycle is coupled with the differentiation program that keratinocytes undergo in the epithelium. After infection of the basal cell layer of the epithelium, the virus establishes and maintains its genome as an extrachromosomal element (episome) in the nucleus of infected cells. While the viral episome is maintained at low levels in basal cells, its amplification to a high copy number is trigged in the upper layers of the epithelium by the action of the viral oncogenes E6 and E7 and the differentiation of the infected keratinocytes (reviewed in reference 21). Replication of the HPV genome relies on the viral proteins E1 and E2 and the host DNA replication machinery. Viral DNA replication is initiated by the binding of E2 to specific sites on the viral origin where it facilitates the recruitment and assembly of E1 into a double hexamer that is required to unwind DNA ahead of the bidirectional replication fork (3, 14, 15, 31, 33, 36, 43-45, 52, 60). In addition to its helicase activity, E1 interacts with several cellular replication factors, including polymerase α-primase, replication protein A (RPA), and topoisomerase I, to replicate the viral episome (5, 6, 19, 32, 35, 39).E1, which belongs to helicase superfamily III (SF3) (22, 26), can be divided into three functional regions. Its C-terminal domain has ATPase and helicase activity and can self-assemble into hexamers. It is also this domain that is contacted by E2 to recruit E1 at the origin (50, 57, 58). The middle portion of E1 encompasses the origin-binding domain (OBD) that binds and dimerizes on specific sequences in the origin (55, 56). We and others previously found that a fragment of E1 containing only the C-terminal enzymatic domain and the OBD is capable of supporting viral DNA replication in vitro but is inactive in vivo (2, 51). This suggested that the N-terminal region of E1 plays an essential regulatory function in vivo. As such, it has been shown for HPV11 E1 that this region contains a cyclin E/A-Cdk2 (cyclin-dependent kinase 2) binding motif (CBM), a bipartite nuclear localization signal (NLS) and an CRM1-dependent nuclear export signal (NES), which together regulate the nucleocytoplasmic shuttling of the protein (10, 30, 34). Specifically, it has been shown that phosphorylation of HPV11 E1 on three serine residues within its N-terminal region inhibits its nuclear export (10, 62). Interestingly, bovine papillomavirus (BPV) E1 was also shown to shuttle between the nucleus and the cytoplasm in a phosphorylation-dependent manner. In this case, however, Cdk2 phosphorylation was found to promote, rather than inhibit, the export of the viral helicase (24). This apparent discrepancy between HPV11 and BPV E1 prompted us to examine the regulation of a third E1 protein, specifically that of the high-risk HPV31.We report here that HPV31 E1 also shuttles between the nucleus and the cytoplasm through its conserved NLS and NES. We determined that nuclear export of HPV31 E1 is dependent on the CRM1 export pathway and is inhibited by Cdk2 phosphorylation of serines 92 and 106. We also found that nuclear export of E1 is not required for transient viral DNA replication and thus investigated its role in viral genome maintenance and amplification in immortalized keratinocytes. In contrast to the wild type (WT), a mutant genome carrying a defective E1 NES was poorly maintained and progressively lost upon cell division, indicating that nuclear export of E1 is required for long-term maintenance of the viral episome. Because nuclear export of E1 is not required for viral DNA replication per se but needed for episomal maintenance over several cell divisions, we investigated the possibility that continuous accumulation of E1 into the nucleus is detrimental to cellular proliferation. In support of this possibility, we found that the accumulation of E1 at high levels in the nucleus impedes cellular proliferation by delaying cell cycle progression in the S phase. In addition, we found that this delay was alleviated when nuclear export of E1 was increased. Altogether, these results suggest that nuclear export of E1 is required, at least in part, to limit accumulation of this viral helicase in the nucleus in order to prevent its detrimental effect on cellular proliferation.  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