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 578 毫秒
1.
Six genes, including UL32, have been implicated in the cleavage and packaging of herpesvirus DNA into preassembled capsids. We have isolated a UL32 insertion mutant which is capable of near-wild-type levels of viral DNA synthesis; however, the mutant virus is unable to cleave and package viral DNA, consistent with the phenotype of a previously isolated temperature-sensitive herpes simplex virus type 1 mutant, tsN20 (P. A. Schaffer, G. M. Aron, N. Biswal, and M. Benyesh-Melnick, Virology 52:57–71, 1973). A polyclonal antibody which recognizes UL32 was previously used by Chang et al. (Y. E. Chang, A. P. Poon, and B. Roizman, J. Virol. 70:3938–3946, 1996) to demonstrate that UL32 accumulates predominantly in the cytoplasm of infected cells. In this report, a functional epitope-tagged version of UL32 showed that while UL32 is predominantly cytoplasmic, some nuclear staining which colocalizes with the major DNA binding protein (ICP8, UL29) in replication compartments can be detected. We have also used a monoclonal antibody (5C) specific for the hexon form of major capsid protein VP5 to study the distribution of capsids during infection. In cells infected with wild-type KOS (6 and 8 h postinfection), 5C staining patterns indicate that capsids are present in nuclei within replication compartments. These results suggest that cleavage and packaging occur in replication compartments at least at 6 and 8 h postinfection. Cells infected with the UL32 mutant exhibit a hexon staining pattern which is more diffusely distributed throughout the nucleus and which is not restricted to replication compartments. We propose that UL32 may play a role in “bringing” preassembled capsids to the sites of DNA packaging and that the failure to localize to replication compartments may explain the cleavage/packaging defect exhibited by this mutant. These results suggest that the UL32 protein is required at a step distinct from those at which other cleavage and packaging proteins are required and may be involved in the correct localization of capsids within infected cells.During infection of cells with herpes simplex virus type 1 (HSV-1), the large concatemeric products of DNA replication are cleaved to unit length and packaged into preassembled capsids. Capsids are icosahedral structures composed of 150 hexons and 12 pentons. Three types of capsids (A, B, and C) can be isolated from infected cells by velocity centrifugation (20). C capsids contain the viral DNA genome; B capsids contain the scaffolding protein; and A capsids contain neither DNA nor the scaffolding protein. Pulse chase experiments with another alphaherpesvirus, equine herpesvirus 1, indicate that at least some B capsids can package DNA and mature into infectious virions, while A capsids cannot (46). By analogy with the bacteriophages, these results suggest that B capsids represent procapsids which are intermediates in the packaging process. However, a new intermediate in the assembly process has recently been identified (41, 62). These newly identified capsid forms observed in in vitro assembly extracts have the same protein content as B capsids but are more spherical; these capsids are unstable and adopt the more angular form characteristic of B capsids after prolonged incubation in vitro. These results suggest that the unstable spherical forms may represent the true procapsid intermediate (41, 62).In many bacteriophages, the procapsid contains at least three essential components: an icosahedrally arranged protein shell, an internal scaffold, and a dodecameric ring called the portal vertex through or around which the phage DNA is taken up (8, 11, 18). For HSV-1, the outer shell is composed of four proteins: the major capsid protein, VP5; a small protein bound to hexons, VP26; and a triplex structure made up of heterotrimers of VP19C and VP23 (reviewed in reference 56). VP24, VP21, and VP22a are found in the interior of the capsid and are encoded by overlapping genes UL26 and UL26.5; VP21 and VP22a are present in B but not A or C capsids and are considered to make up the internal scaffold (reviewed in reference 56). Although bacteriophages contain a portal vertex, no such structure has been observed in HSV-1 capsids. Whether the herpesviruses have a unique portal vertex through which viral DNA is taken up is unclear; it is possible that this type of unique vertex is only needed in viruses which have a tail. Capsids indistinguishable from those isolated from HSV-1-infected cells have been observed in extracts from insect cells infected with recombinant baculoviruses bearing HSV-1 capsid genes (42, 60). Therefore, it is clear that these proteins are sufficient for capsid assembly in vitro; however, it is not known whether capsids formed in vitro are competent for DNA uptake. It is possible that minor components of capsids play important roles in genome encapsidation.In addition to the capsid proteins, at least six genes are essential for the encapsidation of viral DNA: the UL6, UL15, UL25, UL28, UL32, and UL33 genes. Temperature-sensitive (ts) strains with mutations in these genes have similar phenotypes, in that viral DNA can be replicated but not cleaved and packaged (1, 2, 4, 6, 48, 51, 54, 55, 66). Strains with null mutations in the UL6, UL15, UL25, UL28, and UL33 genes have been isolated and characterized, thereby confirming the roles of these genes in cleavage and packaging (5, 27, 37, 45, 59, 68). Despite the identification of these required genes, the mechanism by which viral DNA is cleaved and packaged is not understood, nor has the role of any of the gene products been determined. The UL6 and UL25 proteins have been detected in A, B, and C capsids as well as in virions (3, 28, 37, 44); however, the precise role of these two proteins in capsids remains to be determined.A ts UL32 mutant, tsN20, defective in cleavage and packaging, has been reported previously (51). Because mutants with lesions resulting in temperature sensitivity are often prone to problems associated with incomplete penetrance at the nonpermissive temperature, we isolated a UL32 insertion mutant, hr64. Characterization of hr64 confirms that UL32 is essential for cleavage and packaging. Previous studies demonstrated that UL32 localizes to the cytoplasm of infected cells (13). We have used a functional epitope-tagged version of UL32 to confirm that in infected cells, this protein is mainly cytoplasmic, although some nuclear staining was observed.HSV-1 DNA replication occurs in globular nuclear domains termed “replication compartments” initially identified by ICP8 (UL29) staining patterns in an immunofluorescence assay (49). All seven replication proteins have now been localized within replication compartments (10, 24, 2931, 43) as has regulatory protein ICP4 (26, 50). Ward et al. have recently reported that at late times after infection (18 h), capsids accumulate in the nucleus in regions distinct from replication compartments (64). These authors suggest that these regions represent assembly stations in which DNA is packaged. We report herein, however, that at 6 and 8 h postinfection, capsids colocalize with ICP8 in replication compartments. This suggests that at these early times, cleavage and packaging occur within replication compartments. Furthermore, we report that in cells infected with the UL32 mutant virus, capsids are distributed throughout the nucleus, accumulating in regions outside the replication compartments. This suggests that UL32 may play a role in the efficient localization of capsids in infected cells.  相似文献   

2.
The C terminus of the herpes simplex virus type 1 origin-binding protein, UL9ct, interacts directly with the viral single-stranded DNA-binding protein ICP8. We show that a 60-amino acid C-terminal deletion mutant of ICP8 (ICP8ΔC) also binds very strongly to UL9ct. Using small angle x-ray scattering, the low resolution solution structures of UL9ct alone, in complex with ICP8ΔC, and in complex with a 15-mer double-stranded DNA containing Box I of the origin of replication are described. Size exclusion chromatography, analytical ultracentrifugation, and electrophoretic mobility shift assays, backed up by isothermal titration calorimetry measurements, are used to show that the stoichiometry of the UL9ct-dsDNA15-mer complex is 2:1 at micromolar protein concentrations. The reaction occurs in two steps with initial binding of UL9ct to DNA (Kd ∼ 6 nm) followed by a second binding event (Kd ∼ 0.8 nm). It is also shown that the stoichiometry of the ternary UL9ct-ICP8ΔC-dsDNA15-mer complex is 2:1:1, at the concentrations used in the different assays. Electron microscopy indicates that the complex assembled on the extended origin, oriS, rather than Box I alone, is much larger. The results are consistent with a simple model whereby a conformational switch of the UL9 DNA-binding domain upon binding to Box I allows the recruitment of a UL9-ICP8 complex by interaction between the UL9 DNA-binding domains.The initiation of DNA replication for most double-stranded DNA (dsDNA)6 viral genomes begins with the recognition of the origin by specific origin-binding proteins. The herpes simplex virus type 1 (HSV-1) genome encodes seven proteins required for origin-dependent DNA replication. These are the DNA polymerase (UL30) and its accessory protein (UL42), a heterotrimeric helicase-primase complex (UL5, UL8, and UL52), the single-stranded DNA-binding protein (ICP8 or UL29), and the origin-binding protein (UL9) (reviewed in Ref. 1). HSV-1 contains three functional origins, oriL and two copies of oriS. OriS, which is about 80 bp in length, consists of three UL9 recognition sites, in Boxes I, II, and III, which are arranged in two overlapping palindromes (2). Box I and Box III are part of an evolutionarily conserved palindrome that forms a stable hairpin in single-stranded DNA, which may be important in the origin rearrangement (3) during initiation of replication. Box I and II are separated by an AT-rich spacer sequence, which varies in length and nucleotide composition between the different members of the α-herpesvirus subfamily (2, 46).UL9 is a homodimer in solution, and EM studies, with UL9 bound to oriS, indicate the existence of a dimer or pair of dimers assembled on oriS (7). Several reports indicate that UL9 can physically interact not only with ICP8 (8) but also with other members of the HSV-1 replication complex, including UL8 (9) and UL42 (10). Thus UL9 functions as a docking protein to recruit these essential replication proteins to the viral origins. ICP8 stimulates the helicase activity of UL9 (11, 12) and binds to its C-terminal 27-aa residues (13). In the presence of ICP8, UL9 will open dsDNA containing Box I, leading to a conformational change in the origin, thus facilitating unwinding (1416). As stated above, the changes in DNA conformation in the complete oriS may be more complex (3). Recently, it has been suggested that single-stranded oriS folds into a unique and evolutionarily conserved conformation, oriS*, which is stably bound by UL9. oriS* contains a hairpin formed by complementary base pairing between Box I and Box III in oriS (17). UL9, in the presence of the single-stranded DNA-binding protein ICP8, can convert an 80-bp double-stranded minimal oriS fragment to oriS* and form a UL9-oriS* complex. The formation of a UL9-oriS* complex requires ATP hydrolysis (18). Therefore, the UL9-oriS* complex may serve as an assembly site for the herpesvirus replisome. Macao et al. (3) proposed a model in which full-length UL9 would be required to adopt a different conformation when binding to oriS or oriS*. The implication is that UL9 partially unwinds and introduces a hairpin into the origin of replication and that the formation of oriS* is aided, in some way, by ICP8 and requires ATP hydrolysis. Macao et al. (3) suggest that the length of the single-stranded tail of the probe DNA determines the stoichiometry of the UL9-DNA complex. oriS may bind two molecules of UL9, whereas oriS* may only bind one because the hairpin formation prevents the second interaction.Photo-cross-linking studies have shown that, although the UL9 protein binds Box I as a dimer, only one of the two monomers contacts Box I, suggesting that the C terminus of UL9 undergoes a conformational change upon binding to Box I (19). The results reported here are consistent with this observation. To date there is no three-dimensional structural information available on the full-length UL9 or either of the functionally characterized (helicase and DNA binding) domains. The ability to adopt different conformations and a tendency to proteolytic degradation may be responsible for this. It has been shown that UL9 binds with very high specificity to the Box I through its DNA-binding domain, consisting of the C-terminal 317 aa (UL9ct) (20, 21). Although the importance of the binding between UL9ct and oriS for the viral life cycle is well established, the mechanism behind this interaction still remains unclear. Even though UL9ct exists as a monomer in solution, uncertainty remains as to whether one or two molecules bind to a single Box I recognition sequence. Some reports have suggested that one UL9ct molecule binds to a single copy of the sequence (2224), whereas others have proposed that UL9ct forms a dimer when bound to DNA (25, 26). This apparent difference may well result from the different protein concentrations used in different assays/experiments, which in turn highlights the difficulty of translating in vitro equilibrium experiments into cellular nonequilibrium situations.A few years ago, the crystal structure of a 60-residue C-terminal deletion mutant of ICP8 (ICP8ΔC) was determined to 3 Å resolution (Protein Data Bank code 1URJ (27)). The structure of ICP8ΔC consists of a large N-terminal domain (aa 9–1038) and a smaller entirely helical C-terminal domain (aa 1049–1120) connected to the N-terminal domain by a disordered linker (aa 1038–1049) spanning around 18 Å in the crystal structure. ICP8 preferentially binds ssDNA over dsDNA in a nonsequence-specific and cooperative manner (28). ICP8 is a zinc metalloprotein containing one zinc atom per molecule, which is coordinated by three cysteines (Cys-499, Cys-502, and Cys-510) and a histidine (His-512) (27).In this study, we show that the 60-amino acid C-terminal deletion of ICP8 (ICP8ΔC) binds strongly to UL9ct. We present three low resolution structures in solution using small angle x-ray scattering as follows: that of the UL9ct alone, in complex with ICP8ΔC, and in complex with a 15-mer dsDNA (dsDNA15-mer) containing the Box I sequence. Using these data and a variety of biophysical techniques, we demonstrate that the stoichiometries of the UL9ct-dsDNA15-mer and UL9ct-ICP8ΔC-dsDNA15-mer complexes are 2:1 and 2:1:1, respectively, at the micromolar protein concentrations used in this study. Using EM we visualize the assembly of the ICP8ΔC-UL9ct complex on oriS and estimate the size of the complex.  相似文献   

3.
With the use of a high-throughput biochemical DNA helicase assay as a screen, T157602, a 2-amino thiazole compound, was identified as a specific inhibitor of herpes simplex virus (HSV) DNA replication. T157602 inhibited reversibly the helicase activity of the HSV UL5-UL8-UL52 (UL5/8/52) helicase-primase complex with an IC50 (concentration of compound that yields 50% inhibition) of 5 μM. T157602 inhibited specifically the UL5/8/52 helicase and not several other helicases. The primase activity of the UL5/8/52 complex was also inhibited by T157602 (IC50 = 20 μM). T157602 inhibited HSV growth in a one-step viral growth assay (IC90 = 3 μM), and plaque formation was completely prevented at concentrations of 25 to 50 μM T157602. Vero, human foreskin fibroblast (HFF), and Jurkat cells could be propagated in the presence of T157602 at concentrations exceeding 100 μM with no obvious cytotoxic effects, indicating that the window between antiviral activity and cellular toxicity is at least 33-fold. Seven independently derived T157602-resistant mutant viruses (four HSV type 2 and three HSV type 1) carried single base pair mutations in the UL5 that resulted in single amino acid changes in the UL5 protein. Marker rescue experiments demonstrated that the UL5 gene from T157602-resistant viruses conferred resistance to T157602-sensitive wild-type viruses. Recombinant UL5/8/52 helicase-primase complex purified from baculoviruses expressing mutant UL5 protein showed complete resistance to T157602 in the in vitro helicase assay. T157602 and its analogs represent a novel class of specific and reversible anti-HSV agents eliciting their inhibitory effects on HSV replication by interacting with the UL5 helicase.Herpes simplex virus types 1 and 2 (HSV-1 and HSV-2) each comprise at least 77 genes whose expression is tightly regulated (42). These genes are assigned to four kinetic classes, designated as α, β, γ1, and γ2 on the basis of the timing of and requirements for their expression (46). The five α genes, α0, α4, α22, α27, and α47, are expressed first in the absence of viral protein synthesis and are responsible for the regulated expression of the other viral genes. The β genes require functional α gene products for their expression and encode proteins and enzymes that are directly involved in DNA synthesis and nucleotide metabolism. The γ genes form the last set of viral genes to be expressed, with the γ2 class having viral DNA replication as a strict requirement for their expression.The HSV genome contains three origins of replication (44, 45, 47, 48, 50, 54) and encodes seven viral proteins that are essential for DNA replication (34, 59). These include an origin binding protein (OBP) encoded by open reading frame (ORF) UL9 (14, 15, 17, 35), a DNA binding protein encoded by UL29 (40, 53, 54), a DNA polymerase encoded by ORF UL30 and its accessory factor encoded by UL42 (1, 4, 8, 18, 19, 21, 24, 37), and a heterotrimeric complex consisting of proteins encoded by ORFs UL5, UL8, and UL52, which include both 5′-to-3′ helicase activity and primase activity (1012). Although extensively studied, the roles of the individual subunits of the helicase-primase complex and their specific interactions with each other have not been completely defined. However, several lines of evidence suggest that the UL5 gene encodes the helicase activity of the complex. Examination of the amino acid sequence of the UL5 protein revealed that it contains six conserved motifs that are found in many DNA and RNA helicases, two of these motifs defining an ATP binding site (20, 25, 32, 52, 61). Site-specific mutagenesis of amino acids within each of the six motifs revealed that all six are critical for the function of the UL5 protein as a helicase in transient replication assays (60, 61).The observation that recombinant UL5, UL52, and UL8 proteins could be purified from baculovirus-infected insect cells as a complex that displays DNA-dependent ATPase, helicase, and primase activities that are identical to those produced during a herpesvirus infection allowed functional and biochemical analyses of the individual components of the complex (10, 13, 38). Although the UL5 protein alone contained the defining helicase amino acid sequence motifs, the UL5 protein does not display helicase activity in vitro in the absence of the UL52 protein. Purified UL5 protein has less than 1% of the ATPase activity of the complex UL5-UL8-UL52 (UL5/8/52) complex (2, 43). In addition, studies with recombinant herpesviruses carrying mutations in the UL5 gene that abolish helicase activity revealed that the UL5 protein could still form specific interactions with UL8 and UL52 proteins (60). These results indicate that the functional domains of UL5 protein required for helicase activity are separate from those involved in protein-protein interactions and that UL5 and UL52 must interact to yield efficient helicase activity. Further mutagenesis studies with the UL52 protein identified mutations that abolish the primase activity of the complex, while the helicase and ATPase activities are unaffected, suggesting that the UL52 protein is responsible for the primase activity of the complex (27). The third component of the helicase-primase complex, the UL8 protein, interacts with other viral replication proteins, including the OBP, the single-stranded DNA binding protein, and the viral DNA polymerase (30, 33). It has been postulated that the interaction of the UL8 protein with the OBP (encoded by the UL9 gene) may function to recruit helicase-primase complexes to initiation complexes at viral origins (30). The UL8 protein is also required for stimulation of primer synthesis by the UL52 protein and for stimulation of the helicase activity of the helicase-primase complex which is crucial to allow efficient unwinding of long stretches of duplex DNA (16, 43, 49). Additionally, UL8 appears to be required for efficient nuclear entry of the helicase-primase complex (1, 3, 31).As the UL5, UL8, and UL52 gene products are essential for HSV replication and have not been exploited previously for antiviral drug discovery, they represent attractive targets for the development of novel anti-HSV agents. Current anti-HSV drugs include vidarabine (adenine arabinoside; Ara-A), foscarnet (phosphonoformic acid; PFA), and a wide variety of nucleoside analogs, the most clinically successful being acyclovir (ACV) and its analogs valacyclovir and famciclovir. ACV is phosphorylated by viral thymidine kinase (TK) to its monophosphate form, an event that occurs to a much lesser extent in uninfected cells. Subsequent phosphorylation events by cellular enzymes convert the ACV monophosphate to its triphosphate form. The ACV triphosphate derivative directly inhibits the DNA polymerase by competing as a substrate with dGTP. Because the ACV triphosphate lacks the 3′ hydroxyl group required to elongate the DNA chain, DNA replication is terminated. The triphosphorylated form of ACV is a much better substrate for the viral DNA polymerase than it is for the cellular DNA polymerase; thus, very little ACV triphosphate is incorporated into cellular DNA. Although ACV has proven to be safe and successful at reducing the duration, severity, and in some cases recurrence of HSV infections, eradication of the infection symptoms is far from complete and latent virus can reactivate frequently (5558). In addition, primarily as a result of poor patient compliance with inconvenient ACV dosage regimens, virulent HSV strains resistant to ACV that contain mutations in either the viral TK or DNA polymerase gene have arisen (6, 7, 9, 26, 39). More potent and efficacious drugs that target other essential components of the virus replicative cycle would be invaluable as therapeutic agents to treat HSV and ACV-resistant HSV infections.To identify novel inhibitors of the HSV helicase-primase enzyme, we developed a high-throughput in vitro helicase assay and screened >190,000 samples. Using this biochemical approach, we identified T157602, a 2-amino thiazole, as a specific inhibitor of HSV replication. By generating and analyzing T157602-resistant viruses, we further demonstrate genetically that the molecular target of T157602 is the UL5 component of the HSV helicase-primase complex.  相似文献   

4.
2,5,6-Trichloro-1-β-d-ribofuranosyl benzimidazole (TCRB) is a potent and selective inhibitor of human cytomegalovirus (HCMV) replication. TCRB acts via a novel mechanism involving inhibition of viral DNA processing and packaging. Resistance to the 2-bromo analog (BDCRB) has been mapped to the UL89 open reading frame (ORF), and this gene product was proposed as the viral target of the benzimidazole nucleosides. In this study, we report the independent isolation of virus that is 20- to 30-fold resistant to TCRB (isolate C4) and the characterization of the virus. The six ORFs known to be essential for viral DNA cleavage and packaging (UL51, UL52, UL56, UL77, UL89, and UL104) were sequenced from wild-type HCMV, strain Towne, and from isolate C4. Mutations were identified in UL89 (D344E) and in UL56 (Q204R). The mutation in UL89 was identical to that previously reported for virus resistant to BDCRB, but the mutation in UL56 is novel. Marker transfer analysis demonstrated that each of these mutations individually caused ∼10-fold resistance to the benzimidazoles and that the combination of both mutations caused ∼30-fold resistance. The rate and extent of replication of the mutants was the same as for wild-type virus, but the viruses were less sensitive to inhibition of DNA cleavage by TCRB. Mapping of resistance to UL56 supports and extends recent work showing that UL56 codes for a packaging motif binding protein which also has specific nuclease activity (E. Bogner et al., J. Virol. 72:2259–2264, 1998). Resistance which maps to two different genes suggests that their putative proteins interact and/or that either or both have a benzimidazole ribonucleoside binding site. The results also suggest that the gene products of UL89 and UL56 may be antiviral drug targets.Human cytomegalovirus (HCMV) can cause significant morbidity and mortality in immunocompromised populations (3). It is a common opportunistic disease in patients with AIDS and is often a factor in their death (38). HCMV infection has been implicated in increased risk of organ rejection following heart (28) and kidney transplants (8) and in restenosis of diseased arteries following angioplasty (41, 63). It is also a leading cause of birth defects (16).Current therapies for HCMV infection include ganciclovir (GCV) (22), cidofovir (30), and foscarnet (20). Each of these drugs has several limitations to its use: none are orally bioavailable, all have dose-limiting toxicity, and resistance has developed to each (26). Because all three of these drugs inhibit viral replication through an interaction with the virally encoded DNA polymerase (25, 31, 37), the possibility of cross-resistance exists. Thus, additional drugs with unique mechanisms of action are needed for the treatment of HCMV infections.In 1995, we reported that 2-bromo-5,6-dichloro-1-(β-d-ribofuranosyl)benzimidazole (BDCRB; Fig. Fig.1)1) and the 2-chloro analog [2,5,6-trichloro-1-(β-d-ribofuranosyl)benzimidazole TCRB] are potent and selective inhibitors of HCMV replication (55). These compounds have a novel mechanism of action, which unlike the current therapies for HCMV infection, does not involve inhibition of DNA synthesis. The benzimidazole ribonucleosides prevent the cleavage of high-molecular-weight viral DNA concatemers to monomeric genomic lengths (57). Resistance to BDCRB has been mapped to the HCMV UL89 open reading frame (ORF), which, by analogy to gene gp17 from bacteriophage T4, may be a terminase (23, 57). Consequently, we have proposed that the benzimidazole ribonucleosides inhibit the product of this gene and that the UL89 gene product is involved in the viral DNA concatemer cleavage process (57). Open in a separate windowFIG. 1Structure of benzimidazole ribonucleosides. TCRB, R = Cl; BDCRB, R = Br.HCMV replication proceeds in a manner which is conserved among herpesviruses. The virally encoded DNA polymerase produces large, complex head-to-tail concatemers (10, 29, 33) which must be cleaved into genomic-length pieces before insertion into preformed capsids (59). With herpes simplex virus type 1 (HSV-1), temperature-sensitive mutants which are unable to cleave and package the concatemeric DNA have been derived (1, 2, 4, 45, 49, 50, 61). By this process, six HSV-1 genes have been found to be involved in concatemer cleavage and packaging. They are UL6, UL15, UL25, UL28, UL32, and UL33. In addition, recent studies in Homa’s laboratory have established that the product of UL25 is required for viral DNA encapsidation but not cleavage (39). Homologs of these genes exist in HCMV and are UL104, UL89, UL77, UL56, UL52, and UL51, respectively (18).In our continuing investigation of the mode of action of benzimidazole nucleosides, we report herein the independent isolation of HCMV strains resistant to TCRB, characterization of these strains, and identification of the mutations responsible for the development of resistance. The results demonstrate that the mechanism of action of the benzimidazole ribonucleosides is more complex than previously proposed and that a second gene product implicated in DNA cleavage and packaging is involved.  相似文献   

5.
6.
7.
8.
Coordinated execution of DNA replication, checkpoint activation, and postreplicative chromatid cohesion is intimately related to the replication fork machinery. Human AND-1/chromosome transmission fidelity 4 is localized adjacent to replication foci and is required for efficient DNA synthesis. In S phase, AND-1 is phosphorylated in response to replication arrest in a manner dependent on checkpoint kinase, ataxia telangiectasia-mutated, ataxia telangiectasia-mutated and Rad3-related protein, and Cdc7 kinase but not on Chk1. Depletion of AND-1 increases DNA damage, delays progression of S phase, leads to accumulation of late S and/or G2 phase cells, and induces cell death in cancer cells. It also elevated UV-radioresistant DNA synthesis and caused premature recovery of replication after hydroxyurea arrest, indicating that lack of AND-1 compromises checkpoint activation. This may be partly due to the decreased levels of Chk1 protein in AND-1-depleted cells. Furthermore, AND-1 interacts with cohesin proteins Smc1, Smc3, and Rad21/Scc1, consistent with proposed roles of yeast counterparts of AND-1 in sister chromatid cohesion. Depletion of AND-1 leads to significant inhibition of homologous recombination repair of an I-SceI-driven double strand break. Based on these data, we propose that AND-1 coordinates multiple cellular events in S phase and G2 phase, such as DNA replication, checkpoint activation, sister chromatid cohesion, and DNA damage repair, thus playing a pivotal role in maintenance of genome integrity.Replication fork is not only the site of DNA synthesis but also the center for coordinated execution of various chromosome transactions. The preparation for replication forks starts in the G1 phase, when the prereplicative complex composed of origin recognition and minichromosome maintenance assembles on the chromosome. At the G1-S boundary, Cdc45, GINS complex, and other factors join the prereplicative complex to generate a complex capable of initiating DNA replication. A series of phosphorylation events mediated by cyclin-dependent kinase and Cdc7 kinase play crucial roles in this process and facilitate the generation of active replication forks (16). Purification of the putative replisome complex in yeast indicated the presence of the checkpoint mediator Mrc1 and fork protection complex proteins Tof1 and Csm3 in the replication fork machinery (7), consistent with a previous report on the genome-wide analyses with chromatin immunoprecipitation analyses on chip (microarray) (8). Mcm10 is another factor present in the isolated complex, required for loading of replication protein A (RPA)2 and primase-DNA polymerase α onto the replisome complex (7, 9, 10).Replication fork machinery can cope with various stresses, including shortage of the cellular nucleotide pool and replication fork blockages that interfere with its progression. Stalled replication forks activate checkpoint pathways, leading to cell cycle arrest, DNA repair, restart of DNA replication, or cell death in some cases (1114). Single-stranded DNAs coated with RPA at the stalled replication forks are recognized by the ATR-ATR-interacting protein kinase complex and Rad17 for loading of the Rad9-Rad1-Hus1 checkpoint clamp (1416). Factors present in the replisome complex are also known to be required for checkpoint activation. Claspin, Tim, and Tipin functionally and physically associate with sensor and effector kinases and serve as mediator/adaptors (1723). Mcm7, a component of the replicative DNA helicase in eukaryotes, was reported to associate with the checkpoint clamp loader Rad17 (24) and to have a distinct function in checkpoint (24, 25). We recently reported that Cdc7 kinase, known to be required for DNA replication initiation, plays a role in activation of DNA replication checkpoint possibly through regulating Claspin phosphorylation (26). Thus, it appears that DNA replication and checkpoint activation functionally and physically interact with each other.Another crucial cellular event for maintenance of genome stability is sister chromatid cohesion. The cohesin complex, a conserved apparatus required for sister chromatid cohesion, contains Smc1, Smc3, and Rad21/Scc1/Mcd1 proteins. The assembled cohesin complexes are loaded onto chromatin prior to DNA replication in G1 phase and link the sister chromosomes during S and G2 phase until mitosis when they separate (27, 28). The mitotic cohesion defects are not rescued by supplementing cohesin in G2 phase, and it has been suggested that establishment of sister chromatid cohesion is coupled with DNA replication (29, 30). Indeed, yeast mutants in some replisome components show defect in sister chromosome cohesion or undergo chromosome loss (3133). Cdc7 kinase is also required for efficient mitotic chromosome cohesion (34, 35).Human AND-1 is the putative homolog of budding yeast CTF4/Pob1/CHL15 and fission yeast Mcl1/Slr3. The budding yeast counterpart was identified as a replisome component described above (7), which travels along with the replication fork (29). CTF4 is nonessential for viability, but its interactions with primase, Rad2 (FEN1 family of nuclease), and Dna2 have implicated CTF4 in lagging strand synthesis and/or Okazaki fragment processing (3639). Yeast CTF4 and Mcl1 are involved in chromosome cohesion (33, 40, 41) and genetically interact with a cohesin, Mcd1/Rad21 (40, 42). Recently, it was reported that human AND-1 protein interacts with human primase-DNA polymerase α and Mcm10 and is required for DNA synthesis (43).Here we confirm that human AND-1 protein is required for DNA replication and efficient progression of S phase, and we further show that it facilitates replication checkpoint. Depletion of AND-1 causes accumulation of DNA damage and cell cycle arrest at late S to G2 phase, ultimately leading to cell death. Furthermore, we also show that human AND-1 physically interacts with cohesin proteins Smc1, Smc3, Rad21/Scc1, suggesting a possibility that AND-1 may physically and functionally link replisome and cohesin complexes in vivo. Recent studies indicate that sister chromatid cohesion is required for recombinational DNA repair (4447). Thus, we examined the requirement of AND-1 for repair of artificially induced double-stranded DNA breaks and showed that AND-1 depletion leads to significant reduction of the double strand break repair. Possible roles of AND-1 in coordination of various chromosome transactions at a replication fork and in maintenance of genome integrity during S phase will be discussed.  相似文献   

9.
10.
Mcm10 is an essential eukaryotic protein required for the initiation and elongation phases of chromosomal replication. Specifically, Mcm10 is required for the association of several replication proteins, including DNA polymerase α (pol α), with chromatin. We showed previously that the internal (ID) and C-terminal (CTD) domains of Mcm10 physically interact with both single-stranded (ss) DNA and the catalytic p180 subunit of pol α. However, the mechanism by which Mcm10 interacts with pol α on and off DNA is unclear. As a first step toward understanding the structural details for these critical intermolecular interactions, x-ray crystallography and NMR spectroscopy were used to map the binary interfaces between Mcm10-ID, ssDNA, and p180. The crystal structure of an Mcm10-ID·ssDNA complex confirmed and extended our previous evidence that ssDNA binds within the oligonucleotide/oligosaccharide binding-fold cleft of Mcm10-ID. We show using NMR chemical shift perturbation and fluorescence spectroscopy that p180 also binds to the OB-fold and that ssDNA and p180 compete for binding to this motif. In addition, we map a minimal Mcm10 binding site on p180 to a small region within the p180 N-terminal domain (residues 286–310). These findings, together with data for DNA and p180 binding to an Mcm10 construct that contains both the ID and CTD, provide the first mechanistic insight into how Mcm10 might use a handoff mechanism to load and stabilize pol α within the replication fork.To maintain their genomic integrity, cells must ensure complete and accurate DNA replication once per cell cycle. Consequently, DNA replication is a highly regulated and orchestrated series of molecular events. Multiprotein complexes assembled at origins of replication lead to assembly of additional proteins that unwind chromosomal DNA and synthesize nascent strands. The first event is the formation of a pre-replicative complex, which is composed of the origin recognition complex, Cdc6, Cdt1, and Mcm2–7 (for review, see Ref. 1). Initiation of replication at the onset of S-phase involves the activity of cyclin- and Dbf4-dependent kinases concurrent with recruitment of key factors to the origin. Among these, Mcm10 (2, 3) is recruited in early S-phase and is required for loading of Cdc45 (4). Mcm2–7, Cdc45, and the GINS complex form the replicative helicase (58). Origin unwinding is followed by loading of RPA,3 And-1/Ctf4, and pol α onto ssDNA (912). In addition, recruitment of Sld2, Sld3, and Dpb11/TopBP1 are essential for replication initiation (13, 14), and association of topoisomerase I, proliferating cellular nuclear antigen (PCNA), replication factor C, and the replicative DNA polymerases δ and ϵ completes the replisome (for review, see Ref. 15).Mcm10 is exclusive to eukaryotes and is essential to both initiation and elongation phases of chromosomal DNA replication (6, 8, 16). Mutations in Mcm10 in yeast result in stalled replication, cell cycle arrest, and cell death (2, 3, 1719). These defects can be explained by the number of genetic and physical interactions between Mcm10 and many essential replication proteins, including origin recognition complex, Mcm2–7, and PCNA (3, 12, 2024). In addition, Mcm10 has been shown to stimulate the phosphorylation of Mcm2–7 by Dbf4-dependent kinase in vitro (25). Thus, Mcm10 is an integral component of the replication machinery.Importantly, Mcm10 physically interacts with and stabilizes pol α and helps to maintain its association with chromatin (16, 26, 27). This is a critical interaction during replication because pol α is the only enzyme in eukaryotic cells that is capable of initiating DNA synthesis de novo. Indeed, Mcm10 stimulates the polymerase activity of pol α in vitro (28), and interestingly, the fission yeast Mcm10, but not Xenopus Mcm10, has been shown to exhibit primase activity (29, 30). Mcm10 is composed of three domains, the N-terminal (NTD), internal (ID), and C-terminal (CTD) domains (29). The NTD is presumably an oligomerization domain, whereas the ID and CTD both interact with DNA and pol α (29). The CTD is not found in yeast, whereas the ID is highly conserved among all eukaryotes. The crystal structure of Mcm10-ID showed that this domain is composed of an oligonucleotide/oligosaccharide binding (OB)-fold and a zinc finger motif, which form a unified DNA binding platform (31). An Hsp10-like motif important for the interaction with pol α has been identified in the sequence of Saccharomyces cerevisiae Mcm10-ID (16, 26).DNA pol α-primase is composed of four subunits: p180, p68, p58, and p48. The p180 subunit possesses the catalytic DNA polymerase activity, and disruption of this gene is lethal (32, 33). p58 and p48 form the DNA-dependent RNA polymerase (primase) activity (34, 35), whereas the p68 subunit has no known catalytic activity but serves a regulatory role (36, 37). Pol α plays an essential role in lagging strand synthesis by first creating short (7–12 nucleotide) RNA primers followed by DNA extension. At the critical length of ∼30 nucleotides, replication factor C binds to the nascent strand to displace pol α and loads PCNA with pols δ and ϵ (for review, see Ref. 38).The interaction between Mcm10 and pol α has led to the suggestion that Mcm10 may help recruit the polymerase to the emerging replisome. However, the molecular details of this interaction and the mechanism by which Mcm10 may recruit and stabilize the pol α complex on DNA has not been investigated. Presented here is the high resolution structure of the conserved Mcm10-ID bound to ssDNA together with NMR chemical shift perturbation competition data for pol α binding in the presence of ssDNA. Collectively, these data demonstrate a shared binding site for DNA and pol α in the OB-fold cleft of Mcm10-ID, with a preference for ssDNA over pol α. In addition, we have mapped the Mcm10-ID binding site on pol α to a 24-residue segment of the N-terminal domain of p180. Based on these results, we propose Mcm10 helps to recruit pol α to origins of replication by a molecular hand-off mechanism.  相似文献   

11.
12.
13.
14.
The pseudorabies virus (PrV) gene homologous to herpes simplex virus type 1 (HSV-1) UL53, which encodes HSV-1 glycoprotein K (gK), has recently been sequenced (J. Baumeister, B. G. Klupp, and T. C. Mettenleiter, J. Virol. 69:5560–5567, 1995). To identify the corresponding protein, a rabbit antiserum was raised against a 40-kDa glutathione S-transferase–gK fusion protein expressed in Escherichia coli. In Western blot analysis, this serum detected a 32-kDa polypeptide in PrV-infected cell lysates as well as a 36-kDa protein in purified virion preparations, demonstrating that PrV gK is a structural component of virions. After treatment of purified virions with endoglycosidase H, a 34-kDa protein was detected, while after incubation with N-glycosidase F, a 32-kDa protein was specifically recognized. This finding indicates that virion gK is modified by N-linked glycans of complex as well as high-mannose type. For functional analysis, the UL53 open reading frame was interrupted after codon 164 by insertion of a gG-lacZ expression cassette into the wild-type PrV genome (PrV-gKβ) or by insertion of the bovine herpesvirus 1 gB gene into a PrV gB genome (PrV-gKgB). Infectious mutant virus progeny was obtained only on complementing gK-expressing cells, suggesting that gK has an important function in the replication cycle. After infection of Vero cells with either gK mutant, only single infected cells or small foci of infected cells were visible. In addition, virus yield was reduced approximately 30-fold, and penetration kinetics showed a delay in entry which could be compensated for by phenotypic gK complementation. Interestingly, the plating efficiency of PrV-gKβ was similar to that of wild-type PrV on complementing and noncomplementing cells, pointing to an essential function of gK in virus egress but not entry. Ultrastructurally, virus assembly and morphogenesis of PrV gK mutants in noncomplementing cells were similar to wild-type virus. However, late in infection, numerous nucleocapsids were found directly underneath the plasma membrane in stages typical for the entry process, a phenomenon not observed after wild-type virus infection and also not visible after infection of gK-complementing cells. Thus, we postulate that presence of gK is important to inhibit immediate reinfection.Herpesvirions are complex structures consisting of a nucleoprotein core, capsid, tegument, and envelope. They comprise at least 30 structural proteins (35). Pseudorabies virus (PrV), a member of the Alphaherpesvirinae, is an economically important animal pathogen, causing Aujeszky’s disease in swine. It is also highly pathogenic for most other mammals except higher primates, including humans (28, 45), and a wide range of cultured cells from different species support productive virus replication, reflecting the wide in vivo host range. Envelope glycoproteins play major roles in the early and late interactions between virion and host cell. They are required for virus entry and participate in release of free virions and viral spread by direct cell-to-cell transmission (27, 37). For PrV, 10 glycoproteins, designated gB, gC, gD, gE, gG, gH, gI, gL, gM, and gN, have been characterized (20, 27); these glycoproteins are involved in the attachment of virion to host cell (gC and gD), fusion of viral envelope and cellular cytoplasmic membrane (gB, gD, gH, and gL), spread from infected to noninfected cells (gB, gE, gH, gI, gL, and gM), and egress (gC, gE, and gI) (27, 37). Homologs of these glycoproteins are also present in other alphaherpesviruses (37). The gene coding for a potential 11th PrV glycoprotein, gK, has been described recently (3), but the protein and its function have not been identified.The product of the homologous UL53 open reading frame (ORF) of herpes simplex virus type 1 (HSV-1) is gK (13, 32). gK was detected in nuclear membranes and in membranes of the endoplasmic reticulum but was not observed in the plasma membrane (14). Also, it did not appear to be present in purified virion preparations (15). The latter result was surprising since earlier studies identified several mutations in HSV-1 gK resulting in syncytium-inducing phenotypes (7, 14), which indicates participation of gK in membrane fusion events during HSV-1 infection. Moreover, HSV-1 mutants in gK exhibited a delayed entry into noncomplementing cells, which is difficult to reconcile with absence of gK from virions (31). Mutants deficient for gK expression have been isolated and investigated by different groups (16, 17). Mutant F-gKβ carries a lacZ gene insertion in the HSV-1 strain F gK gene, which interrupts the ORF after codon 112 (16). In mutant ΔgK, derived from HSV-1 KOS, almost all of the UL53 gene was deleted (17). Both mutants formed small plaques on Vero cells, and virus yield was reduced to an extent which varied with the different confluencies of the infected cells, cell types, and mutants used for infection. However, both HSV-1 gK mutants showed a defect in efficient translocation of virions from the cytoplasm to the extracellular space, and only a few enveloped virions were present in the extracellular space after infection of Vero cells (16, 17). The authors therefore suggested that HSV-1 gK plays a role in virion transport during egress.Different routes of final envelopment and egress of alphaherpesvirions are discussed. It has been suggested that HSV-1 nucleocapsids acquire their envelope at the inner nuclear membrane and are transported as enveloped particles through the endoplasmic reticulum to the Golgi stacks, where glycoproteins are modified in situ during transport (5, 6, 19, 39), although other potential egress pathways cannot be excluded (4). In contrast, maturation of varicella-zoster virus and PrV involves primary envelopment at the nuclear membrane, followed by release of nucleocapsids into the cytoplasm and secondary envelopment in the trans-Golgi area (10, 12, 43). Final egress of virions appears to occur via transport vesicles containing one or more virus particles by fusion of vesicle and cell membrane. The possibility of different routes of virion egress is supported by studies of other proteins involved in egress, e.g., the UL20 proteins of HSV-1 and PrV and the PrV UL3.5 protein, which lacks a homolog in the HSV-1 genome (1, 8, 9). In UL20-negative HSV-1, virions accumulated in the perinuclear cisterna of Vero cells (1), while PrV UL20 virions accumulated and were retained in cytoplasmic vesicles (9). PrV UL3.5 is important for budding of nucleocapsids into Golgi-derived vesicles during secondary envelopment (8). Thus, there appear to be profound differences in the egress pathways. Since HSV-1 gK was also implicated in egress, we were interested in identifying the PrV homolog and analyzing its function.  相似文献   

15.
16.
17.
Upon DNA replication stress, stalled DNA replication forks serve as a platform to recruit many signaling proteins, leading to the activation of the DNA replication checkpoint. Activation of Rad53, a key effector kinase in the budding yeast Saccharomyces cerevisiae, is essential for stabilizing DNA replication forks during replication stress. Using an activity-based assay for Rad53, we found that Mrc1, a replication fork-associated protein, cooperates with Mec1 to activate Rad53 directly. Reconstitution of Rad53 activation using purified Mec1 and Mrc1 showed that the addition of Mrc1 stimulated a more than 70-fold increase in the ability of Mec1 to activate Rad53. Instead of increasing the catalytic activity of Mec1, Mrc1 was found to facilitate the phosphorylation of Rad53 by Mec1 via promotion of a stronger enzyme-substrate interaction between them. Further, the conserved C-terminal domain of Mrc1 was found to be required for Rad53 activation. These results thus provide insights into the role of the adaptor protein Mrc1 in activating Rad53 in the DNA replication checkpoint.Faithful replication of the genome is important for the survival of all organisms. During DNA replication, replication stress can arise from a variety of situations, including intrinsic errors made by DNA polymerases, difficulties in replicating repeated DNA sequences, and failures to repair damaged DNA caused by either endogenous oxidative agents or exogenous mutagens such as UV light and DNA-damaging chemicals (13). In eukaryotes, there is an evolutionarily conserved DNA replication checkpoint that becomes activated in response to DNA replication stress. It helps to stabilize DNA replication forks, block late replication origin firing, and delay mitosis and ultimately helps recovery from stalled replication forks after DNA repair (47). Defects in the DNA replication checkpoint could result in elevated genomic instabilities, cancer development, or cell death (8, 9).Aside from replicating the genome, the DNA replication forks also provide a platform to assemble many signaling proteins that function in the DNA replication checkpoint. In the budding yeast Saccharomyces cerevisiae, Mec1, an ortholog of human ATR,2 is a phosphoinositide 3-kinase-like kinase (PIKK) involved in sensing stalled DNA replication forks. Mec1 forms a protein complex with Ddc2 (ortholog of human ATRIP). The Mec1-Ddc2 complex is recruited to stalled replication forks through replication protein A (RPA)-coated single-stranded DNA (10, 11). The Mec3-Rad17-Ddc1 complex, a proliferating cell nuclear antigen (PCNA)-like checkpoint clamp and ortholog of the human 9-1-1 complex, was shown to be loaded onto the single- and double-stranded DNA junction of the stalled replication forks by the clamp loader Rad24-RFC complex (12). Once loaded, the Mec3-Rad17-Ddc1 complex stimulates Mec1 kinase activity (13). Dbp11 and its homolog TopBP1 in vertebrates are known components of the replication machinery (14). In addition to regulating the initiation of DNA replication, they were found to play a role in the DNA replication checkpoint (1517). They interact with the 9-1-1 complex and directly stimulate Mec1/ATR activity in vitro (1820). Thus, the assembly of multiple protein complexes at stalled DNA replication forks appears to facilitate activation of the DNA replication checkpoint (13, 18).Mrc1 (for mediator of replication checkpoint) was originally identified to be important for cells to respond to hydroxyurea in S. cerevisiae and Schizosaccharomyces pombe (21, 22). Mrc1 is a component of the DNA replisome and travels with the replication forks along chromosome during DNA synthesis (2325). Deletion of MRC1 causes defects in DNA replication, indicating its role in the normal progression of DNA replication (23). Interestingly, when DNA replication is blocked by hydroxyurea, Mrc1 undergoes Mec1- and Rad3 (S. pombe ortholog of Mec1)-dependent phosphorylation (21, 22). In S. cerevisiae, mutations of Mrc1 at the (S/T)Q sites, which are consensus phosphorylation sites of the Mec1/ATR family kinases, abolishes hydroxyurea-induced Mrc1 phosphorylation in vivo, suggesting a direct phosphorylation of Mrc1 by Mec1 (21, 22).Rad53 and Cds1, homologs of human Chk2, are the major effector kinases in the DNA replication checkpoints in S. cerevisiae and S. pombe, respectively. Activation of Rad53 is a hallmark of DNA replication checkpoint activation and is important for the maintenance of DNA replication forks in response to DNA replication stress (5, 6). Thus, it is important to understand how Rad53 activity is controlled. Interestingly, mutation of all the (S/T)Q sites of Mrc1 not only abolishes the phosphorylation of Mrc1 by Mec1 but also compromises hydroxyurea-induced Rad53 activation in S. cerevisiae (21). Similarly, mutation of the TQ sites of Mrc1 in S. pombe was shown to abolish the binding between Cds1 and Mrc1 as well as Cds1 activation (22). Further, mutation of specific TQ sites of Mrc1 in S. pombe abolishes its binding to Cds1 in vitro and the activation of Cds1 in vivo (26). Thus, Mec1/Rad3-dependent phosphorylation of Mrc1 is responsible for Mrc1 binding to Rad53/Cds1, which is essential for Rad53/Cds1 activation.An intriguing property of the Chk2 family kinases is their ability to undergo autophosphorylation and activation in the absence of other proteins in vitro (27, 28). First, autophosphorylation of a conserved threonine residue in the activation loop of Chk2 family kinase was found to be an essential part of their activation processes (26, 2931). Second, a direct and trans-phosphorylation of the N-terminal TQ sites of the Chk2 family kinases by the Mec1/ATR family kinases is also important for their activation in vivo. Analogous to the requirement of N-terminal TQ site phosphorylation of Chk2 by ATR in human (32), the activation of Rad53/Cds1 in vivo requires phosphorylation of TQ sites in their N termini by Mec1/Rad3 (33, 34).Considering that Mec1, Mrc1, and many other proteins are recruited at stalled DNA replication forks and have been shown to be involved in DNA replication checkpoint activation, a key question remains unresolved: what is the minimal system that is capable of activating Rad53 directly? Given the direct physical interaction between Mrc1 and Rad53 and the requirement of Mrc1 and Mec1 in vivo, it is likely that they both play a role in Rad53 activation. Furthermore, what is the molecular mechanism of Rad53 activation by its upstream activators? To address these questions, a faithful reconstitution of the activation of Rad53 using purified proteins is necessary. In this study, we developed an activity-based assay consisting of the Dun1 kinase, a downstream substrate of Rad53, and Sml1, as a substrate of Dun1, to quantitatively measure the activity of Rad53. Using this coupled kinase assay from Rad53 to Dun1 and then to Sml1, we screened for Mrc1 and its associated factors to see whether they could directly activate Rad53 in vitro. Our results showed that Mec1 and Mrc1 collaborate to constitute a minimal system in direct activation of Rad53.  相似文献   

18.
The herpes simplex virus replicon consists of cis-acting sequences, oriS and oriL, and the origin binding protein (OBP) encoded by the UL9 gene. Here we identify essential structural features in the initiator protein OBP and the replicator sequence oriS, and we relate the appearance of these motifs to the evolutionary history of the alphaherpesvirus replicon. Our results reveal two conserved sequence elements in herpes simplex virus type 1, OBP; the RVKNL motif, common to and specific for all alphaherpesviruses, is required for DNA binding, and the WP XXXGAXXFXX L motif, found in a subset of alphaherpesviruses, is required for specific binding to the single strand DNA-binding protein ICP8. A 121-amino acid minimal DNA binding domain containing conserved residues is not soluble and does not bind DNA. Additional sequences present 220 amino acids upstream from the RVKNL motif are needed for solubility and function. We also examine the binding sites for OBP in origins of DNA replication and how they are arranged. NMR and DNA melting experiments demonstrate that origin sequences derived from many, but not all, alphaherpesviruses can adopt stable boxI/boxIII hairpin conformations. Our results reveal a stepwise evolutionary history of the herpes simplex virus replicon and suggest that replicon divergence contributed to the formation of major branches of the herpesvirus family.Herpesviruses have been found in animal species ranging from molluscs to man. According to the International Committee on Taxonomy of Viruses, the order of Herpesvirales consists of three families as follows: Alloherpesviridae, Herpesviridae, and Malacoherpesviridae (1). The subfamilies Alphaherpesvirinae, Betaherpesvirinae, and Gammaherpesvirinae are found within the family of Herpesviridae. The events leading to establishment of a new virus species are poorly understood, but in the case of herpesviruses it is commonly assumed that viruses co-evolve with their hosts (2). Herpesviruses have thus become well adapted to their hosts and may reside in a latent state in the host for a lifetime with little or no overt signs of infection. Upon reactivation, the infectious virus will be released. The viruses remain faithful to their hosts, and infections across species borders are rare but may under specific circumstances give rise to fatal disease.Replication of herpes simplex virus type 1 (HSV-1),2 requires a cis-acting DNA sequence, the replicator, termed oriS or oriL, an initiator protein, OBP or UL9 protein, and a replisome composed of DNA polymerase, helicase-primase, and a single strand DNA-binding protein referred to as ICP8 or UL29 protein (3). OBP assisted by ICP8 can in an ATP-dependent reaction unwind double-stranded oriS (4, 5). The resulting single-stranded DNA adopts a hairpin conformation, which is stably bound by OBP (6, 7). The herpesvirus replisome once assembled on DNA is capable of synthesizing leading and lagging strands processively in a coordinated fashion (8). DNA replication is likely to start on circular molecules produced by the action of DNA ligase IV/XRCC4 and proceed in a θ-type manner (9, 10). Later, a rolling circle mode of replication dominates giving rise to characteristic head-to-tail concatemers. The initiator protein OBP appears to be strictly required only during the first few hours of the infectious cycle (11, 12).The HSV-1 origin binding protein was first isolated using an assay monitoring specific binding to HSV-1 oriS (13). A minimal DNA-binding site was identified using footprinting techniques as well as binding studies with double-stranded oligonucleotides (14). For alphaherpesviruses the high affinity binding site is always TTCGCAC, with a minor exception for CHV1 also referred to as monkey B virus (3 The C-terminal 317 amino acids of alphaherpesvirus OBP can be isolated as a soluble protein, which remains capable of high affinity binding to the sequence TTCGCAC (15). The C-terminal domain of HSV-1 OBP, here referred to as ΔOBP, binds as a monomer to the major groove in the DNA and makes contacts with base pairs as well as the deoxyribose-phosphate backbone (16, 17). ΔOBP binds DNA specifically with an estimated Kd of 0.3 nm, a value that is highly influenced by the composition of the assay buffer (16). At high protein concentrations ΔOBP form aggregates, which, still in a sequence-specific manner, binds DNA (16). A number of studies have attempted to define amino acids involved in DNA binding (1821). In addition, sequence comparisons between alphaherpesviruses and roseoloviruses have helped to identify amino acids in OBP potentially involved in DNA binding as well as corresponding recognition sequences in origins of DNA replication (2225). However, a comprehensive and quantitative study of evolutionarily conserved amino acids required for DNA binding is still lacking.

TABLE 1

Functionally significant sequence motifs for herpes simplex virus replicon evolutionOrigins of DNA replication have been identified from sequence analysis. For roseoloviruses the sequences for two slightly different binding sites for OBP have been listed (31). The number and orientation of OBP-binding sites in relation to an AT-rich spacer sequence are schematically presented by symbols > and <. Note that since a virus often has more than one origin of replication they may exist as variants. This is indicated by symbols within parentheses. The conserved amino acids within the ICP8-binding motif are shown in boldface type. Sources and nomenclature for DNA sequences have been presented in Footnote 3.Open in a separate windowThe single strand DNA-binding protein, ICP8 encoded by the UL29 gene, is involved in initiation of DNA replication, and it also participates in the elongation phase at the replication fork (4, 5, 26). ICP8 forms a specific complex with OBP (26, 27). Studies with deletion mutants have demonstrated that important sequences are found close to the C terminus of OBP, but amino acids directly participating in the high affinity interaction have not been identified. The interaction is biologically significant, because deletion of the extreme C terminus enhances the helicase activity of OBP but reduces origin-dependent DNA synthesis (26).The HSV-1 oriS contains three copies of the binding site for OBP; two binding sites, box I and box II, are high affinity sites, and the third binding site, box III, has a very low affinity for OBP (14) (Fig. 1). All sites are required for efficient replication, and in a competitive situation there is a strong selection for the most efficient replicator sequence (28). Box III and box I are arranged in a palindrome, which becomes rearranged upon activation to form an alternative conformation, most likely a hairpin (4, 6, 7). Point mutations that prevent formation of a hairpin reduce replication, and compensatory mutations restore complementary base pairing as well as the ability to replicate (7).Open in a separate windowFIGURE 1.Schematic presentation of the herpes simplex virus replicon. Upper part, HSV-1 oriS. The linear genome contains three homologous replication origins, two copies of oriS and one copy oriL, and encodes seven replication proteins. Middle part, HSV-1 OBP. OBP or UL9 protein is a superfamily II DNA helicase as well as a sequence-specific DNA-binding protein. Here the helicase domain is represented by two connected ellipsoids, and the C-terminal DNA binding is drawn as a circle. The OBP-binding sites in oriS are shown. The OBP dimer binds two double-stranded DNA box I oligonucleotides but only one hairpin with a single-stranded tail (48). The figure is intended to demonstrate conformational changes affecting the DNA binding domain, referred to as ΔOBP, during activation of oriS. Lower part, DNA binding domain ΔOBP. A schematic presentation of the following three motifs discussed in this publication: the F553 XX KYL motif required for proper folding of the DNA binding domain, the R756VKNL motif necessary for DNA binding, and the W839PXXXGAXXFXXL motif involved in binding to ICP8.To learn more about the mechanism of virus evolution, we have examined the evolutionary history of some functionally significant features of the HSV-1 replicon and related them to a sequence-based evolutionary tree. The results indicate that replicon divergence, characterized by the stepwise appearance of the DNA-binding RVKNL motif, the WPXXXGAXXFXXL ICP8-binding motif, and the boxIII–boxI palindrome, may have played important roles in establishing major branches of the alphaherpesvirus tree.  相似文献   

19.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

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