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.
The catalytic subunit of herpes simplex virus DNA polymerase (Pol), a member of the B family polymerases, possesses both polymerase and exonuclease activities. We previously demonstrated that a recombinant virus (YD12) containing a double mutation within conserved exonuclease motif III of the Pol was highly mutagenic and rapidly evolved to contain an additional leucine-to-phenylalanine mutation at residue 774 (L774F), which is located within the finger subdomain of the polymerase domain. We further demonstrated that the recombinant L774F virus replicated DNA with increased fidelity and that the L774F mutant Pol exhibited altered enzyme kinetics and impaired polymerase activity to extension from mismatched primer termini. In this study, we demonstrated that addition of the L774F mutation to the YD12 Pol did not restore the exonuclease deficiency. However, the polymerase activity of the YD12 Pol to extension from mismatched primer termini and on the nucleotide incorporation pattern was altered upon addition of the L774F mutation. The L774F mutation-containing YD12 Pol also supported the growth of viral progeny and replicated DNA more efficiently and more accurately than did the YD12 Pol. Together, these studies demonstrate that a herpes simplex virus Pol mutant with a highly mutagenic ability can rapidly acquire additional mutations, which may be selected for their survival and outgrowth. Furthermore, the studies demonstrate that the polymerase activity of HSV-1 Pol on primer extension is influenced by sequence context and that herpes simplex virus type 1 Pol may dissociate more frequently at G·C sites during the polymerization reaction. The implications of the findings are discussed.Herpes simplex virus (HSV) DNA polymerase consists of the catalytic subunit of the polymerase (Pol) and the processivity factor UL42. The Pol subunit contains three well-defined activities: polymerization (replication), exonuclease proofreading (editing), and UL42 binding (5, 6, 28). The UL42 binding activity is mediated by amino acid residues located at the C terminus (5, 6). Although the UL42 binding residues are unique to certain alphaherpesvirus DNA polymerases, the sequences comprising the polymerase and exonuclease domains are conserved among the B family (or the α-like) polymerases (2-4). The exonuclease domain of the HSV type 1 (HSV-1) Pol contains conserved exonuclease I (Exo I), II, and III motifs, whereas the polymerase domain contains seven conserved regions (I to VII); conserved region IV overlaps with the Exo II motif. The Exo III motif is located within the δ region C, which is highly conserved among the B family polymerases (Fig. (Fig.1A).1A). These conserved regions are located within the palm, the thumb, and the finger subdomains, which comprise the structural components of the polymerase domain. The crystal structure of the HSV-1 Pol subunit revealed three grooves that form the putative polymerase, exonuclease, and DNA binding sites. The putative exonuclease site is defined as a groove formed between the exonuclease domain and the tip of the thumb subdomain. The palm and thumb subdomains form a groove proposed to be the putative duplex DNA binding site for both the editing and the polymerization complexes (23). Thus, the polymerase and exonuclease domains of HSV-1 are structurally and functionally interconnected (1, 7, 16, 21, 23, 27, 28), although they are organized into two different domains.Open in a separate windowFIG. 1.(A) Schematic diagram of the conserved regions and motifs within HSV-1 Pol. The relative locations of the conserved regions of HSV-1 Pol are shown at the top; regions I to VII and δ region C are represented by open boxes. The conserved exonuclease motifs I, II, and III are indicated with closed boxes. The functional and structural domains (determined by crystal structure analysis [23]) of the HSV-1 Pol are shown below. The N-terminal domain (N domain) is composed of two regions separated by the 3′-to-5′ exonuclease domain (23). (B) Schematic diagram of wild-type and mutant YDL Pol. The BamHI fragment of the wild-type pol from the plasmid pHC629 is shown at the top. The relative location of conserved region VI and the Exo III motif are shown below, with corresponding wild-type and mutant (YDL) amino acid sequences. B, BamHI; M, MstI; N, NotI.The high fidelity of DNA replication is achieved by three different mechanisms: nucleotide discrimination during the polymerization reaction, editing immediately after the polymerization reaction, and postreplication repair. HSV-1 mutant Pol containing mutations within the conserved regions of the polymerase domain can result in altered enzyme kinetics and DNA replication fidelity (8, 9, 11, 12, 18, 26). Similarly, mutation of conserved Exo domain residues can lead to the loss of exonuclease activity and to altered nucleotide selection and incorporation kinetics as well as the mutator phenotype (1, 10, 13, 14, 21, 25). Our previous studies demonstrated that a mutant Pol (YD12) containing a tyrosine-to-histidine substitution at residue 577 and an aspartic acid-to-alanine substitution at residue 581 (Y577H/D581A) is exonuclease deficient (exo) and that recombinant virus expressing the mutant Pol exhibits a mutator phenotype in vivo (14). However, this recombinant virus rapidly evolved to contain an additional leucine-to-phenylalanine substitution at residue 774 (L774F), which is located within conserved region VI of the polymerase domain (18). Interestingly, a recombinant virus containing the L774F Pol mutation exhibits increased fidelity of DNA replication (18). Our recent study also demonstrated that the mutant L774F Pol exhibits altered enzyme kinetics (26). These results led to the hypothesis that the emerged L774F mutation in the context of the YD12 Pol mutant may also affect enzyme activity, DNA replication, and fidelity.  相似文献   

3.
The UL33 protein of herpes simplex virus type 1 (HSV-1) is thought to be a component of the terminase complex that mediates the cleavage and packaging of viral DNA. In this study we describe the generation and characterization of a series of 15 UL33 mutants containing insertions of five amino acids located randomly throughout the 130-residue protein. Of these mutants, seven were unable to complement the growth of the UL33-null virus dlUL33 in transient assays and also failed to support the cleavage and packaging of replicated amplicon DNA into capsids. The insertions in these mutants were clustered between residues 51 and 74 and between 104 and 116, within the most highly conserved regions of the protein. The ability of the mutants to interact with the UL28 component of the terminase was assessed in immunoprecipitation and immunofluorescence assays. All four mutants with insertions between amino acids 51 and 74 were impaired in this interaction, whereas two of the three mutants in the second region (with insertions at positions 111 and 116) were not affected. These data indicate that the ability of UL33 to interact with UL28 is probably necessary, but not sufficient, to support viral growth and DNA packaging.During the packaging of the double-stranded DNA genome of herpes simplex virus type 1 (HSV-1), the cleavage of replicated concatemeric viral DNA into single-genome lengths is tightly coupled to its insertion into preassembled spherical procapsids. Upon genome insertion, the internal scaffold protein of the procapsid is lost, and the capsid shell angularizes. Genetic analysis has revealed that successful packaging requires a cis-acting DNA sequence (the a sequence) together with seven proteins, encoded by the UL6, UL15, UL17, UL25, UL28, UL32, and UL33 genes (6, 10). By analogy with double-stranded bacteriophage, the encapsidation of HSV-1 DNA is thought to be mediated by a heteromultimeric terminase enzyme. It is envisaged that the terminase is involved in the recognition of packaging signals present in the concatemers and the association with procapsids via an interaction with the capsid portal protein. Terminase initiates packaging by cleaving at an a sequence present between adjacent genomes within concatemers and subsequently provides energy for genome insertion through the hydrolysis of ATP. Packaging is terminated by a second cleavage event at the next similarly orientated a sequence, resulting in the encapsidation of a unit-length genome.An accumulating body of evidence suggests that the HSV-1 terminase is comprised of the UL15, UL28, and UL33 gene products. Viruses lacking a functional version of any of these three proteins are unable to initiate DNA packaging, and uncleaved concatemers and abortive B-capsids (angularized forms containing scaffold but no DNA) accumulate in the nuclei of infected cells (2, 4, 5, 11, 25, 27, 30, 36, 38). Protein sequence comparisons revealed a distant relationship between UL15 and the large subunit of bacteriophage T4 terminase, gp17, including the presence of Walker A and B box motifs characteristic of ATP binding proteins (13). Subsequent experiments demonstrated that point mutations affecting several of the most highly conserved residues abolished the ability of the resulting mutant viruses to cleave and package viral DNA (26, 39). The UL28 component has been reported to interact with the viral DNA packaging signal (3), a property shared with the homologous protein of human cytomegalovirus (CMV), UL56 (9). Furthermore, both UL15 and UL28 are able to interact with UL6 (33, 37), which form a dodecameric portal complex through which DNA is inserted into the capsid (22, 23, 31). Within the terminase complex, strong interactions have previously been reported between UL15 and UL28 and between UL28 and UL33 (1, 7, 17, 19, 34). Evidence also suggests that UL15 and UL33 may be able to interact directly, albeit more weakly than UL28 and UL33 (7, 15). Temperature-sensitive (ts) lesions in UL33 or UL15 reduced both the interaction of the thermolabile protein with the other members of the terminase complex and viral growth at the nonpermissive temperature (36). Recent evidence suggests that the terminase complex assembles in the cytoplasm and is imported into the nucleus via a mechanism involving a nuclear localization signal within UL15 (35). UL15 is also necessary for the localization of the terminase to nuclear sites of DNA replication and packaging (15). At present, the enzymatic activities necessary for DNA packaging have not been demonstrated for either the complex or individual subunits of the HSV-1 terminase.This study concerns the UL33 protein, which, at 130 residues, is the smallest subunit of the presumptive terminase (7, 27). No specific role in terminase activity has yet been ascribed to UL33, but several possibilities have been proposed including (i) ensuring correct folding or assembly of the complex, (ii) regulating the functions of the other subunits, (iii) performing an essential enzymatic role per se, and (iv) ensuring correct localization of the terminase to sites of DNA packaging (7). However, recent immunofluorescence studies using mutants with defects in the individual terminase subunits suggest that UL33 is unlikely to be involved in this last function (15).In order to further investigate the role of UL33 in the cleavage-packaging process, we utilized transposon-mediated mutagenesis to introduce insertions of five codons throughout the UL33 ORF. We report the generation and characterization of 15 mutants in terms of their ability to support viral growth and DNA packaging and to interact with the terminase component UL28.  相似文献   

4.
5.
6.
7.
The processing of lagging-strand intermediates has not been demonstrated in vitro for herpes simplex virus type 1 (HSV-1). Human flap endonuclease-1 (Fen-1) was examined for its ability to produce ligatable products with model lagging-strand intermediates in the presence of the wild-type or exonuclease-deficient (exo) HSV-1 DNA polymerase (pol). Primer/templates were composed of a minicircle single-stranded DNA template annealed to primers that contained 5′ DNA flaps or 5′ annealed DNA or RNA sequences. Gapped DNA primer/templates were extended but not significantly strand displaced by the wild-type HSV-1 pol, although significant strand displacement was observed with exo HSV-1 pol. Nevertheless, the incubation of primer/templates containing 5′ flaps with either wild-type or exo HSV-1 pol and Fen-1 led to the efficient production of nicks that could be sealed with DNA ligase I. Both polymerases stimulated the nick translation activity of Fen-1 on DNA- or RNA-containing primer/templates, indicating that the activities were coordinated. Further evidence for Fen-1 involvement in HSV-1 DNA synthesis is suggested by the ability of a transiently expressed green fluorescent protein fusion with Fen-1 to accumulate in viral DNA replication compartments in infected cells and by the ability of endogenous Fen-1 to coimmunoprecipitate with an essential viral DNA replication protein in HSV-1-infected cells.Herpes simplex virus type 1 (HSV-1), the prototypic member of the family of Herpesviridae and that of the alphaherpesviridae subfamily, has served as the model for understanding the replication of herpesvirus genomes during lytic virus replication (29). The 152-kbp genome of herpes simplex virus type 1 (HSV-1) possesses approximately 85 genes, 7 of which have been shown to be necessary and sufficient for viral DNA replication within host cells (reviewed in references 5 and 38). These seven genes encode a DNA polymerase (pol) and its processivity factor (UL42), a heterotrimeric complex containing a DNA helicase (UL5), primase (UL52), and noncatalytic accessory protein (UL8), a single-stranded DNA binding protein (infected cell protein 8 [ICP-8]), and an origin binding protein with DNA helicase activity (UL9). There is strong evidence in support of the circularization of the linear virion DNA shortly after entry, and DNA replication then is thought to initiate at one or more of the three redundant origins of replication (29, 38). At least in the earliest stages of viral DNA replication, UL9 protein is required, presumably to bind to and unwind the DNA and to attract the other DNA replication proteins (29, 38). The electron microscopic examination of pulse-labeled replicating HSV-1 DNA indicates the presence of lariats, eye-forms, and D-forms (21), which is consistent with bidirectional theta-like replication from origins. To date, however, no biochemical assay has demonstrated origin-dependent DNA replication in vitro. However, in the absence of UL9, the other six HSV DNA replication proteins can support initiation and replication from a circular single-stranded DNA (ssDNA) template in an origin-independent fashion (15, 26), resembling the rolling-circle mode of replication thought to occur during the later stages of viral replication.Although nicks and small gaps have been observed in isolated replicating and virion DNA (38), the evidence for bidirectional duplex synthesis, the rapid rate of viral DNA replication, and the absence of long stretches of ssDNA in replicating and mature DNA isolated from HSV-1-infected cells suggest that leading- and lagging-strand synthesis are closely coordinated in vivo. Falkenberg et al. (15) used a minicircle DNA template with a strand bias and the six essential HSV-1 DNA replication proteins needed for rolling circle replication to demonstrate lagging-strand synthesis in vitro. However, replication from the parental strand template (leading-strand synthesis) was more efficient than synthesis from the complementary-strand template (lagging-strand synthesis). These results suggest the possibility that one or more host functions required for efficient lagging-strand synthesis or for its close coordination with leading-strand synthesis is missing in such in vitro systems.Although leading- and lagging-strand syntheses share many of the same requirements for bulk DNA synthesis, lagging-strand synthesis is a more complex process. Because the direction of polymerization of lagging-strand intermediates is opposite the direction of replication fork movement, lagging-strand synthesis requires that priming and extension occur many times to produce discontinuous segments called Okazaki fragments (reviewed in reference 25). Okazaki fragments need to be processed to remove the RNA primer, to fill in the area previously occupied by the RNA, and to seal the remaining nick between fragments, all of which must occur efficiently, accurately, and completely. Failure to do so would result in the accumulation of DNA breaks, multiple mutations, delayed DNA replication, and/or cell death (16, 61).In eukaryotes, what is currently known regarding the process of lagging-strand synthesis is based on genetic and biochemical studies with Saccharomyces cerevisiae and on in vitro reconstitution studies to define the mammalian enzymes required for simian virus 40 (SV40) T-antigen-dependent DNA replication (17, 37, 44, 57, 58). These studies have revealed that the extension of a newly synthesized Okazaki fragment DNA with pol δ causes the strand displacement of the preceding fragment to produce a 5′ flap (25). Results suggest that flap endonuclease 1 (Fen-1) is the activity responsible for the removal of the bulk of the 5′ flaps generated (1, 44, 48), although dna2 protein may facilitate the removal of longer flaps coated with the ssDNA binding protein complex (2, 44). In addition, the overexpression of exonuclease I can partially compensate for the loss of Fen-1 function in yeast (24, 51). For the proper processing of lagging-strand intermediates, the entire 5′ flap and all of the RNA primer need to be removed, and the gap must be filled to achieve a ligatable nick. DNA ligase I has been shown to be the enzyme involved in sealing Okazaki fragments in yeast and in humans (3, 31, 50, 56, 57). DNA ligase I requires a nick in which there is a 5′ phosphate on one end and a 3′ hydroxyl linked to a deoxyribose sugar entity on the other, and it works poorly in the presence of mismatches (54). The close coordination of Fen-1 and DNA ligase I activities for Okazaki fragment processing is facilitated by the interactions of these proteins with proliferating cell nuclear antigen (PCNA), the processivity factor for pol δ and ɛ (6, 30, 32, 46, 52, 53).HSV-1 does not appear to encode a protein with DNA ligase activity or one that can specifically cleave 5′ flaps, although it does encode a 5′-to-3′ exonuclease activity (UL12 [10, 20]) and a 3′-to-5′ exonuclease activity that is part of the HSV-1 pol catalytic subunit (27). As for most eukaryotes, RNA primers are essential for HSV-1 DNA synthesis, as demonstrated by the presence of oligoribonucleotides in replicating DNA in vivo (4), by the well-characterized ability of the UL52 protein in complex with the UL5 helicase activity to synthesize oligoribonucleotide primers on ssDNA in vitro (11, 13), and by the requirement of the conserved catalytic residues in the UL52 primase in vitro and in HSV-1-infected cells (14, 26). It is the strand displacement activity of pol δ that produces the 5′ flaps that are key to the removal of RNA primers during Okazaki fragment processing (6, 25). However, we previously demonstrated that wild-type HSV-1 DNA polymerase possesses poor strand displacement activity (62), in contrast to mammalian DNA pol δ (25). Thus, it is not apparent how RNA primers would be removed when encountered by HSV-1 pol during HSV-1 lagging-strand synthesis or how such intermediates would be processed.We wished to test the hypothesis that the nick translation activity of mammalian Fen-1 could function in collaboration with HSV-1 pol to facilitate the proper removal of RNA primers and/or short flaps to produce the ligatable products required for Okazaki fragment processing. In this report, we have examined the ability of wild-type and exonuclease-deficient (exo) HSV-1 pol, which differ in their respective strand displacement activities, to extend model lagging-strand substrates in the presence or absence of mammalian Fen-1. Our results demonstrate that both wild-type and exo HSV-1 pol can cooperate with and enhance Fen-1 activity to achieve a ligatable nick in vitro. Moreover, colocalization and coimmunoprecipitation studies reveal a physical association of Fen-1 with HSV-1 DNA replication proteins, supporting a model for the involvement of Fen-1 in HSV-1 DNA replication.  相似文献   

8.
9.
10.
Kaposi''s sarcoma-associated herpesvirus is an emerging pathogen whose mechanism of replication is poorly understood. PF-8, the presumed processivity factor of Kaposi''s sarcoma-associated herpesvirus DNA polymerase, acts in combination with the catalytic subunit, Pol-8, to synthesize viral DNA. We have solved the crystal structure of residues 1 to 304 of PF-8 at a resolution of 2.8 Å. This structure reveals that each monomer of PF-8 shares a fold common to processivity factors. Like human cytomegalovirus UL44, PF-8 forms a head-to-head dimer in the form of a C clamp, with its concave face containing a number of basic residues that are predicted to be important for DNA binding. However, there are several differences with related proteins, especially in loops that extend from each monomer into the center of the C clamp and in the loops that connect the two subdomains of each protein, which may be important for determining PF-8''s mode of binding to DNA and to Pol-8. Using the crystal structures of PF-8, the herpes simplex virus catalytic subunit, and RB69 bacteriophage DNA polymerase in complex with DNA and initial experiments testing the effects of inhibition of PF-8-stimulated DNA synthesis by peptides derived from Pol-8, we suggest a model for how PF-8 might form a ternary complex with Pol-8 and DNA. The structure and the model suggest interesting similarities and differences in how PF-8 functions relative to structurally similar proteins.Most if not all organisms with DNA genomes have mechanisms to ensure processive DNA synthesis. In bacteria, archaea, and eukaryotes, DNA polymerase subunits include a catalytic subunit and a processivity factor, often referred to as a “sliding clamp.” In these organisms, a clamp loader protein is required to assemble the processivity factor onto the DNA (27, 37). The bacterial sliding (beta) clamp is made up of homodimers of a subunit that comprises three structurally similar subdomains (26), whereas archaeal and eukaryotic proliferating cell nuclear antigen (PCNA) is composed of homotrimers that comprise two structurally similar subdomains (27, 37). For both of these clamps, the monomers assemble head-to-tail to form a closed homodimeric or homotrimeric ring, respectively, around the DNA. In these organisms, a clamp loader protein is required to efficiently load the clamp onto DNA, using an ATP-dependent process. Once loaded on DNA, the processivity factor is capable of binding directly to the DNA polymerase, conferring extended strand synthesis without falling off of the template (50).Herpesviruses encode their own DNA polymerases. However, unlike bacteria, archaea, and eukaryotes, herpesviruses do not encode clamp loaders to assemble their processivity factors onto the DNA. Yet, the accessory subunits of the herpesvirus DNA polymerases still associate with DNA with nanomolar affinity to enable long-chain DNA synthesis (9, 16, 23, 25, 29, 35, 44, 46, 53, 56). Human herpesviruses are divided into three classes, namely, the alpha-, beta-, and gammaherpesviruses, based on homologies found in their genomic organization as well as in protein sequences and function (45). Crystal structures have been determined for the processivity factor UL42 from the alphaherpesvirus herpes simplex virus type 1 (HSV-1) and for UL44 from the betaherpesvirus human cytomegalovirus (HCMV) (2, 3, 58). Despite having little if any sequence homology with processivity factors outside of their herpesvirus subfamily, these structures all share the “processivity fold” originally seen in the structure of the bacterial beta clamp (26). Interestingly, some of these processivity factors have a different quaternary structure. PCNA forms a head-to-tail trimeric ring (18, 27), HSV-1 UL42 is a monomer (10, 14, 16, 46, 58) equivalent to one-third of the PCNA complex, and HCMV UL44 is a head-to-head dimer in the form of a C-shaped clamp (2, 3, 9).Both HSV-1 UL42 and HCMV UL44 have a basic face that has been shown to be important for interacting with DNA (25, 35). In the case of dimeric HCMV UL44, the basic surface of each monomer faces inward, toward the center of the C clamp, and includes a basic loop, called the “gap loop,” that is thought to wrap around DNA (24). Recently the crystal structure of the bacterial beta clamp was determined in complex with DNA (15). In that structure, DNA was found to be located in the central pore of the clamp. Amino acid residues that interacted with DNA were in positions structurally homologous to those found on the positively charged faces of UL42 and UL44.UL42 and UL44 each also has a surface, facing away from the DNA binding face, that is important for interacting with the catalytic subunit of the viral DNA polymerase. Indeed, both of these proteins have been crystallized in complex with C-terminal peptides from their respective catalytic subunits, HSV-1 UL30 and HCMV UL54 (2, 58). Together with biochemical and mutational analyses, these crystal structures indicated that, although the details of the interaction are different, the catalytic subunit of the polymerase binds to a region including and in close proximity to a long loop that connects the N- and C-terminal subdomains, called the interdomain connector loop (32-34). The corresponding region of PCNA is also important for polymerase attachment and mediates the interactions of PCNA with many other cellular proteins (40). Both UL54 and UL30 were shown to attach to their respective subunits, UL44 and UL42, by way of their extreme C termini. The C-terminal residues responsible for this interaction correspond to amino acids that are not detectably conserved, either by sequence or by structure, among herpesvirus catalytic subunits. The HSV-1 UL30-UL42 interaction involves a groove to one side of the UL42 connector loop, with hydrophilic interactions being critical (58). The HCMV UL54-UL44 interaction involves a crevice near the UL44 connector loop, and hydrophobic interactions are crucial (2, 32, 33). Moreover, the HCMV UL44 crevice is on the opposite side of the connector loop with respect to the HSV-1 UL42 groove.Kaposi''s sarcoma-associated herpesvirus (KSHV), a gammaherpesvirus, encodes a viral DNA polymerase catalytic subunit, Pol-8, and an accessory subunit, PF-8 (4, 7, 8, 29, 48, 57). PF-8 can bind to Pol-8 directly and specifically (8, 29) and is required for long-chain DNA synthesis in vitro (29). Similarly to UL44, PF-8 forms dimers in solution and when bound to DNA (9). Although it is likely that UL44 and PF-8 are the processivity factors for HCMV and KSHV, respectively, rigorous experiments demonstrating this have not been performed. However, for the sake of brevity and clarity, we will refer to these proteins as processivity factors.Here we present the crystal structure of PF-8 and show that PF-8 forms a head-to-head homodimer akin to UL44 but lacking the long gap loops which are thought to wrap around DNA. This suggests that PF-8 binds DNA differently than does UL44 or UL42. Because Pol-8 appears to lack a long, flexible C-terminal tail with a length comparable to those of other herpesvirus Pols, we expect the mode of binding of the catalytic subunit to be different as well. Based on structural data, information from homologs, and initial biochemical results, we were able to identify possible sites for interactions with DNA and Pol-8 and to propose a model for the simultaneous interaction of all three components of the complex. Further, the availability of crystal structures for all three herpesvirus classes provides new insights into comparative structure, function, and evolution.  相似文献   

11.
Nitrofurazone is reduced by cellular nitroreductases to form N2-deoxyguanine (N2-dG) adducts that are associated with mutagenesis and lethality. Much attention recently has been given to the role that the highly conserved polymerase IV (Pol IV) family of polymerases plays in tolerating adducts induced by nitrofurazone and other N2-dG-generating agents, yet little is known about how nitrofurazone-induced DNA damage is processed by the cell. In this study, we characterized the genetic repair pathways that contribute to survival and mutagenesis in Escherichia coli cultures grown in the presence of nitrofurazone. We find that nucleotide excision repair is a primary mechanism for processing damage induced by nitrofurazone. The contribution of translesion synthesis to survival was minor compared to that of nucleotide excision repair and depended upon Pol IV. In addition, survival also depended on both the RecF and RecBCD pathways. We also found that nitrofurazone acts as a direct inhibitor of DNA replication at higher concentrations. We show that the direct inhibition of replication by nitrofurazone occurs independently of DNA damage and is reversible once the nitrofurazone is removed. Previous studies that reported nucleotide excision repair mutants that were fully resistant to nitrofurazone used high concentrations of the drug (200 μM) and short exposure times. We demonstrate here that these conditions inhibit replication but are insufficient in duration to induce significant levels of DNA damage.Replication in the presence of DNA damage is thought to produce most of the mutagenesis, genomic rearrangements, and lethality that occur in all cells. UV-induced photoproducts, X-ray-induced strand breaks, psoralen- or cis-platin-interstrand cross-links, oxidized bases from reactive oxygen species, and base depurination are just a few of the structurally distinct challenges that the replication machinery must overcome. It seems likely that the mechanisms that process these lesions will vary depending on the nature of the impediment.While a number of the lesions described above are known to block replication, the events associated with UV-induced damage have been the most extensively characterized. UV irradiation causes the formation of cyclobutane pyrimidine dimers and 6-4 photoproducts in DNA that block the progression of the replication fork (16, 29, 30, 37). Following the arrest of replication at UV-induced damage, RecA and several RecF pathway proteins are required to process the replication fork such that the blocking lesion is removed or bypassed (2, 5, 6, 8-10). Cells lacking either RecA or any of several RecF pathway proteins are hypersensitive to UV-induced damage and fail to recover replication following disruption by the lesions (2, 6, 10). RecBCD is an exonuclease/helicase complex that is involved in repairing double-strand breaks (38). It also is required for resistance to UV-induced damage, although it is not required to process or restore disrupted replication forks, and the substrates it acts upon after UV irradiation currently remain unclear (3, 10, 19).Survival and the ability to resume DNA synthesis following UV-induced damage depend predominantly on the removal of the lesions by nucleotide excision repair (5, 7, 36). Cells deficient in nucleotide excision repair are unable to remove UV-induced DNA lesions and exhibit elevated levels of mutagenesis, strand exchanges, rearrangements, and cell lethality (16, 33, 34). In cases where replication fork processing or lesion repair is prevented, the recovery of replication and survival become entirely dependent on translesion synthesis by DNA polymerase V (Pol V) (6). However, in repair-proficient cells, the contribution of translesion synthesis to recovery and survival is minor and is detected only following UV doses that exceed the repair capacity of the cell (5, 6).Less is known about how replication recovers from other forms of DNA damage. We chose to characterize nitrofurazone, because a number of studies suggested that N2-deoxyguanine (N2-dG) adducts induced by this and other agents would be processed differently than UV-induced lesions. Nitrofurazone is a topical antibacterial agent that historically has been used for treating burns and skin grafts in patients and animals (14, 15, 32). Nitrofurazone toxicity is known to require activation by cellular nitroreductases (25, 42). However, the mechanism and targets of its antimicrobial properties have yet to be fully elucidated. In addition to its antimicrobial properties, the reduced nitrofurazone metabolites also target DNA and have been shown to induce free radical damage, strand breaks, and N2-dG adducts (26, 40, 42, 45), and they are mutagenic and carcinogenic in rodent models (1, 15, 24, 39).Whereas nucleotide excision repair is the predominant mechanism required for survival after UV-induced damage, a number of studies suggest that translesion synthesis plays a larger role in survival after nitrofurazone-induced DNA damage. dinB mutants lacking Pol IV were shown to be hypersensitive to nitrofurazone compared to cells that constitutively express the polymerase (17). Biochemically, Pol IV and a number of Pol IV homologs from other organisms have been shown to efficiently replicate over a range of N2-dG adducts in vitro (17, 35, 44). In addition, several studies have reported that uvrA mutants, which are defective in nucleotide excision repair, do not exhibit any hypersensitivity to nitrofurazone or other agents that induce similar adducts in vivo (12, 21, 27). Early studies also observed a direct correlation between nitrofurazone-induced mutations and lethality, suggesting that mutagenic lesions persist in the DNA to cause toxicity (21, 23, 27, 43). Consistent with these observations, nitrofuran-induced lesions were found to be poor substrates for nucleotide excision repair in vitro (46).Taken together, these observations suggest to us that the cellular response to nitrofurazone will be distinct from its response to UV irradiation. However, no study has examined the relative contributions that nucleotide excision repair, translesion synthesis, or recombination has in recovering from nitrofurazone-induced damage. In this study, we characterized the mechanism by which nitrofurazone inhibits DNA replication and identified the genes that contribute to the recovery, survival, and mutagenesis of Escherichia coli treated with nitrofurazone. In contrast to previous studies, we found that survival following nitrofurazone-induced damage depends predominantly on nucleotide excision repair. Similarly to UV-induced DNA damage, both the RecF and RecBC pathways contribute to survival following nitrofurazone-induced DNA damage. The contribution of translesion polymerases to survival was minor and was mediated by Pol IV. In addition, we found that nitrofurazone can act to inhibit DNA replication directly when used at higher concentrations. The direct inhibition of replication is reversible and occurs independently of DNA damage, suggesting that DNA is not the primary target of its antimicrobial properties.  相似文献   

12.
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   

13.
The herpes simplex virus type 1 (HSV-1) gene UL12 encodes a conserved alkaline DNase with orthologues in all herpesviruses. The HSV-1 UL12 gene gives rise to two separately promoted 3′ coterminal mRNAs which encode distinct but related proteins: full-length UL12 and UL12.5, an amino-terminally truncated form that initiates at UL12 codon 127. Full-length UL12 localizes to the nucleus where it promotes the generation of mature viral genomes from larger precursors. In contrast, UL12.5 is predominantly mitochondrial and acts to trigger degradation of the mitochondrial genome early during infection. We examined the basis for these very different subcellular localization patterns. We confirmed an earlier report that the amino-terminal region of full-length UL12 is required for nuclear localization and provide evidence that multiple nuclear localization determinants are present in this region. In addition, we demonstrate that mitochondrial localization of UL12.5 relies largely on sequences located between UL12 residues 185 and 245 (UL12.5 residues 59 to 119). This region contains a sequence that resembles a typical mitochondrial matrix localization signal, and mutations that reduce the positive charge of this element severely impaired mitochondrial localization. Consistent with matrix localization, UL12.5 displayed a detergent extraction profile indistinguishable from that of the matrix protein cyclophilin D. Mitochondrial DNA depletion required the exonuclease activity of UL12.5, consistent with the idea that UL12.5 located within the matrix acts directly to destroy the mitochondrial genome. These results clarify how two highly related viral proteins are targeted to different subcellular locations with distinct functional consequences.All members of the Herpesviridae encode a conserved alkaline DNase that displays limited homology to bacteriophage λ red α (2, 24), an exonuclease that acts in conjunction with the synaptase red β to catalyze homologous recombination between DNA molecules (23). The most thoroughly characterized member of the herpesvirus alkaline nuclease family is encoded by the herpes simplex virus type 1 (HSV-1) gene UL12 (7, 9, 22). HSV-1 UL12 has both endo- and exonuclease activity (15-17, 36) and binds the viral single-stranded DNA binding protein ICP8 (37, 39) to form a recombinase that displays in vitro strand exchange activity similar to that for red α/β (27). UL12 localizes to the nucleus (26) where it plays an important, but as-of-yet ill-defined, role in promoting the production of mature packaged unit-length linear progeny viral DNA molecules (12, 20), perhaps via a recombination mechanism (27, 28). The importance of UL12 is documented by the observation that UL12 null mutants display a ca. 1,000-fold reduction in the production of infectious progeny virions (41).HSV-1 also produces an amino-terminally truncated UL12-related protein termed UL12.5, which is specified by a separately promoted mRNA that initiates within UL12 coding sequences (7, 9, 21). UL12.5 is translated in the same reading frame as UL12 but initiates at UL12 codon 127 and therefore lacks the first 126 amino acid residues of the full-length protein. UL12.5 retains the nuclease and ICP8 binding activities of UL12 (4, 14, 26) but does not accumulate to high levels in the nucleus (26) and is unable to efficiently substitute for UL12 in promoting viral genome maturation (14, 21). We recently showed that UL12.5 localizes predominantly to mitochondria, where it triggers massive degradation of the host mitochondrial genome early during HSV infection (31). Mammalian mitochondrial DNA (mt DNA) is a 16.5-kb double-stranded circle located within the mitochondrial matrix that encodes 13 proteins involved in oxidative phosphorylation and the RNA components of the mitochondrial translational apparatus (reviewed in reference 10). Inherited mutations that inactivate or deplete mt DNA impair oxidative phosphorylation, leading to a wide range of pathological conditions, including neuropathy and myopathy (reviewed in references 8 and 40). Thus, although the contribution of mt DNA depletion to the biology of HSV infection has yet to be determined, it likely has a major negative impact on host cell functions.UL12 and UL12.5 provide a striking example of a pair of highly related proteins that share a common biochemical activity yet differ markedly in subcellular location and biological function. The basis for their distinct subcellular localization patterns is of considerable interest, as the only difference in the primary sequences is that UL12.5 lacks the first 126 residues of UL12. Reuven et al. (26) demonstrated that this UL12-specific region contains one or more signals able to target enhanced green fluorescent protein (eGFP) to the nucleus. However, the determinants of the mitochondrial localization of UL12.5 have not been previously examined. Most proteins that are imported into the mitochondrial matrix bear a matrix targeting sequence that is located at or close to the amino terminus (reviewed in references 25 and 38). We speculated that UL12.5 bears such an amino-terminal matrix targeting sequence and that the function of this element is masked in the full-length UL12 protein by the UL12-specific amino-terminal extension, which contains the nuclear localization signal(s) (NLS). Our results broadly support this hypothesis and indicate that the mitochondrial localization sequence of UL12.5 is located ca. 60 residues from its N terminus.  相似文献   

14.
15.
A conserved family of herpesvirus protein kinases plays a crucial role in herpesvirus DNA replication and virion production. However, despite the fact that these kinases are potential therapeutic targets, no systematic studies have been performed to identify their substrates. We generated an Epstein-Barr virus (EBV) protein array to evaluate the targets of the EBV protein kinase BGLF4. Multiple proteins involved in EBV lytic DNA replication and virion assembly were identified as previously unrecognized substrates for BGLF4, illustrating the broad role played by this protein kinase. Approximately half of the BGLF4 targets were also in vitro substrates for the cellular kinase CDK1/cyclin B. Unexpectedly, EBNA1 was identified as a substrate and binding partner of BGLF4. EBNA1 is essential for replication and maintenance of the episomal EBV genome during latency. BGLF4 did not prevent EBNA1 binding to sites in the EBV latency origin of replication, oriP. Rather, we found that BGLF4 was recruited by EBNA1 to oriP in cells transfected with an oriP vector and BGLF4 and in lytically induced EBV-positive Akata cells. In cells transfected with an oriP vector, the presence of BGLF4 led to more rapid loss of the episomal DNA, and this was dependent on BGLF4 kinase activity. Similarly, expression of doxycycline-inducible BGLF4 in Akata cells led to a reduction in episomal EBV genomes. We propose that BGLF4 contributes to effective EBV lytic cycle progression, not only through phosphorylation of EBV lytic DNA replication and virion proteins, but also by interfering with the EBNA1 replication function.Herpesviruses encode two families of serine/threonine protein kinases, one of which, the BGLF4 (Epstein-Barr virus [EBV])/UL97 (human cytomegalovirus)/UL13 (herpes simplex virus)/ORF36 (Kaposi''s sarcoma-associated herpesvirus)/ORF47 (varicella-zoster virus) family, is the sole protein kinase encoded by beta and gamma herpesviruses. The protein kinases phosphorylate both viral and host proteins (16, 21, 42) and are necessary for efficient virus lytic replication. Consequently, these kinases have been of interest as potential targets for antiviral drug development (37), and the compound 1263W94 (maribavir), which inhibits the cytomegalovirus UL97 protein (3), has been used in phase I clinical trials (27, 31, 47).EBV infection is prevalent worldwide, and primary infection in adolescence or early adulthood is associated in 30 to 40% of cases with infectious mononucleosis. EBV efficiently infects B cells in the lymphoid tissues of the Waldeyer ring (43). EBV infection of B cells is biased toward establishment of latency with limited viral-gene expression (49). During latent infection, EBV genomes are maintained as extrachromosomal episomes. Replication of episomal genomes utilizes the latency origin of replication, oriP. The only EBV-encoded protein required is the origin binding protein EBNA1. All other essential replication factors are provided by the cell. Expression of the EBV replicative cycle and production of progeny virus take place in terminally differentiated plasma B cells (11, 29), and epithelial cells may also contribute to the cycle of virus replication and spread that is an important component of both persistent infection of the individual and transmission of virus from one individual to the next (4, 22). Lytic DNA replication initiates at separate origins, oriLyt. EBV encodes a set of six core lytic replication proteins, along with ancillary proteins, such as thymidine kinase (TK), that are involved in nucleotide metabolism (13, 44).Several substrates have been described for the EBV BGLF4 protein kinase, including the core lytic EBV replication protein BMRF1, the polymerase processivity factor (8, 17). BGLF4 has also been found to locate to sites of lytic viral replication (46), to be required for efficient lytic DNA replication and release of nucleocapsids from the nucleus (18), and to contribute to the compaction of cell chromatin seen in cells undergoing lytic replication (32). Protein chip technology provides a new tool for global analysis of activities for biologically important enzymes, such as ubiquitin ligases, DNA repair enzymes, and kinases (7, 19, 36, 38, 52). Using an EBV protein array for unbiased screening, we identified multiple new BGLF4 substrates involved in lytic DNA replication, capsid assembly, and DNA packaging. Unexpectedly, we also identified EBNA1 as a substrate and binding partner for BGLF4. The data suggest that the contribution of BGLF4 to the EBV lytic cycle extends beyond the previously recognized contributions to lytic DNA replication and virion production and includes facilitating the switch from latent to lytic DNA replication by downregulating the EBNA1 replication function.  相似文献   

16.
17.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号