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 15 毫秒
1.
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   

2.
The interactions of the herpes simplex virus processivity factor UL42 with the catalytic subunit of the viral polymerase (Pol) and DNA are critical for viral DNA replication. Previous studies, including one showing that substitution of glutamine residue 282 with arginine (Q282R) results in an increase of DNA binding in vitro, have indicated that the positively charged back surface of UL42 interacts with DNA. To investigate the biological consequences of increased DNA binding by UL42 mutations, we constructed two additional UL42 mutants, including one with a double substitution of alanine for aspartic acid residues (D270A/D271A) and a triple mutant with the D270A/D271A and Q282R substitutions. These UL42 mutants exhibited increased and prolonged DNA binding without an effect on binding to a peptide corresponding to the C terminus of Pol. Plasmids expressing any of the three UL42 mutants with an increased positive charge on the back surface of UL42 were qualitatively competent for complementation of growth and DNA replication of a UL42 null mutant on Vero cells. We then engineered viruses expressing these mutant proteins. The UL42 mutants were more resistant to detergent extraction than wild-type UL42, suggesting that they are more tightly associated with DNA in infected cells. All three UL42 mutants formed smaller plaques on Vero cells and replicated to reduced yields compared with results for a control virus expressing wild-type UL42. Moreover, mutants with double and triple mutations, which contain D270A/D271A mutations, exhibited increased mutation frequencies, and mutants containing the Q282R mutation exhibited elevated ratios of virion DNA copies per PFU. These results suggest that herpes simplex virus has evolved so that UL42 neither binds DNA too tightly nor too weakly to optimize virus production and replication fidelity.Processivity factors of DNA polymerases promote long-chain DNA synthesis by preventing dissociation of the DNA polymerase from the primer/template. Processivity factors also can influence DNA replication fidelity, as indicated by numerous in vivo and in vitro studies (1-3, 5, 6, 11, 12, 18, 28, 36). A major class of processivity factors known as “sliding clamps” includes proliferating cell nuclear antigen (PCNA) of eukaryotic cells (23) and gp45 of T4 bacteriophage (27). Sliding clamps are homodimers or homotrimers that encircle DNA and interact with the catalytic subunits (Pols) of their cognate DNA polymerases to promote processive DNA synthesis.A second class of processivity factors includes those encoded by herpesviruses and is exemplified by herpes simplex virus (HSV) UL42. UL42 forms a heterodimer with the HSV Pol. Both subunits are essential for production of infectious virus and for viral DNA replication (20, 26). UL42 can stimulate long-chain DNA synthesis by Pol, and template challenge experiments established that this stimulation is due to increased processivity (15). In addition to its interaction with Pol, which is mediated by the C terminus of Pol, UL42 also binds DNA directly with high affinity (14, 15, 30, 37). This mode of DNA binding differs from that of sliding clamps, which do not form high-affinity direct interactions with DNA (13) but must be loaded onto DNA with the aid of ATP-dependent clamp loaders for their normal functioning (16). Nevertheless, the structure of UL42 is very similar to a monomer of the sliding clamp PCNA (39). Like other processivity factors, UL42 also plays a role in maintaining DNA replication fidelity both in vivo and in vitro (5, 18).The “back face” (opposite face to the side that binds Pol) of a UL42 molecule contains several positively charged residues. By titrating the effects of cations on UL42 DNA binding, it was determined that charge-charge interactions are involved in the interaction (22). Substitutions of alanine for any of four arginine residues on the back face of UL42 resulted in substantial reductions in DNA binding without affecting the binding to peptide corresponding to the C terminus of Pol in vitro (31), while substitutions of lysine for arginine had little or no effect on DNA binding affinity (22). A UL42 mutant (Q282R) containing a substitution of arginine for a negatively charged glutamine residue on the back face of UL42 exhibited a fourfold increase in DNA binding without altering the interaction with the Pol C-terminal peptide in vitro (22). Therefore, the positively charged surface of UL42 is important for the interaction between UL42 and DNA. A question raised by these studies is whether UL42 could bind DNA so tightly as to affect HSV replication.Mutant viruses engineered to encode individual arginine-to-alanine substitution mutations in UL42 exhibit several phenotypes, including a delayed onset of viral DNA replication, reduced virus yields, and reduced fidelity of DNA replication (18). Recombinant viruses expressing UL42 with multiple substitutions of alanine for arginine residues exhibit even greater effects on viral DNA replication and virus yields (19). Thus, reducing DNA binding by UL42 deleteriously affects viral growth and DNA replication fidelity. However, these studies did not address whether increasing DNA binding by UL42 would have any effects on viral DNA replication, replication fidelity, or virus production.In this study we engineered two new UL42 mutant proteins (with the D270A/D271A or Q282R/D270A/D271A mutations) that contain less negative charge on the back face and examined the effects of these substitutions on DNA and Pol peptide binding. In addition, recombinant viruses were constructed to examine the effect of these multiple substitutions and the single Q282R substitution on virus production, DNA replication, and the fidelity of DNA replication.  相似文献   

3.
4.
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   

5.
6.
7.
8.
Many microorganisms encode proteins that interact with molecules involved in host immunity; however, few of these molecules have been proven to promote immune evasion in vivo. Herpes simplex virus type 1 (HSV-1) glycoprotein C (gC) binds complement component C3 and inhibits complement-mediated virus neutralization and lysis of infected cells in vitro. To investigate the importance of the interaction between gC and C3 in vivo, we studied the virulence of a gC-null strain in complement-intact and C3-deficient animals. Using a vaginal infection model in complement-intact guinea pigs, we showed that gC-null virus grows to lower titers and produces less severe vaginitis than wild-type or gC rescued virus, indicating a role for gC in virulence. To determine the importance of complement, studies were performed with C3-deficient guinea pigs; the results demonstrated significant increases in vaginal titers of gC-null virus, while wild-type and gC rescued viruses showed nonsignificant changes in titers. Similar findings were observed for mice where gC null virus produced significantly less disease than gC rescued virus at the skin inoculation site. Proof that C3 is important was provided by studies of C3 knockout mice, where disease scores of gC-null virus were significantly higher than in complement-intact mice. The results indicate that gC-null virus is approximately 100-fold (2 log10) less virulent that wild-type virus in animals and that gC-C3 interactions are involved in pathogenesis.  相似文献   

9.
本文利用PCR技术建立一种对HSV直接基因分型的方法。在HSV-Ⅰ、Ⅱ两型的DNA多聚酶基因上设计了一条两型共同的上游引物(HDP-B)和两条型特异的下游引物(HDP-1、HDP-2)。三条引物共同组成一个扩增反应体系,在HSV-Ⅰ产生543bp条带,HSV-Ⅱ产生372bp条带,据此在基因水平上对HSV进行分型。5株不同来源的HSV(2株Ⅰ型,3株Ⅱ型)分型结果与病毒分离及血清学方法完全一致。该反应体系与其它来源的DNA不产生特异反应,敏感性可达1fg。应用该法对151份临床可疑HSV感染的标本进行检测并分型,结果与免疫学方法完全一致。  相似文献   

10.
The catalytic subunit of herpes simplex virus 1 DNA polymerase (HSV-1 Pol) has been extensively studied; however, its full complement of functional domains has yet to be characterized. A crystal structure has revealed a previously uncharacterized pre-NH2-terminal domain (residues 1 to 140) within HSV-1 Pol. Due to the conservation of the pre-NH2-terminal domain within the herpesvirus Pol family and its location in the crystal structure, we hypothesized that this domain provides an important function during viral replication in the infected cell distinct from 5′-3′ polymerase activity. We identified three pre-NH2-terminal Pol mutants that exhibited 5′-3′ polymerase activity indistinguishable from that of wild-type Pol in vitro: deletion mutants PolΔN43 and PolΔN52 that lack the extreme N-terminal 42 and 51 residues, respectively, and mutant PolA6, in which a conserved motif at residues 44 to 49 was replaced with alanines. We constructed the corresponding pol mutant viruses and found that the polΔN43 mutant displayed replication kinetics similar to those of wild-type virus, while polΔN52 and polA6 mutant virus infection resulted in an 8-fold defect in viral yield compared to that achieved with wild type and their respective rescued derivative viruses. Additionally, both polΔN52 and polA6 viruses exhibited defects in viral DNA synthesis that correlated with the observed reduction in viral yield. These results strongly indicate that the conserved motif within the pre-NH2-terminal domain is important for viral DNA synthesis and production of infectious virus and indicate a functional role for this domain.  相似文献   

11.
12.
13.
14.
Herpes simplex virus (HSV) encodes seven proteins necessary for viral DNA synthesis—UL9 (origin-binding protein), ICP8 (single-strand DNA [ssDNA]-binding protein), UL30/UL42 (polymerase), and UL5/UL8/UL52 (helicase/primase). It is our intention to provide an up-to-date analysis of our understanding of the structures of these replication proteins and how they function during HSV replication. The potential roles of host repair and recombination proteins will also be discussed.The Herpesviridae are a large family of double-stranded DNA viruses responsible for many human and veterinary diseases. Although members of this family differ in tissue tropism and many aspects of their interactions with their hosts, the mechanisms by which they replicate their DNA during productive (“lytic”) infection are largely conserved. The molecular mechanisms involved in herpesvirus DNA replication and its regulation are of interest as they provide important models for the study of eukaryotic DNA replication. Many of the replication proteins encoded by herpesviruses represent functional analogs of the eukaryotic DNA replication machinery, with informative similarities and differences. In addition, viral enzymes involved in DNA replication have provided a rich store of useful targets for antiviral therapy. This work will focus primarily on DNA replication of herpes simplex virus 1 and 2 (HSV-1 and HSV-2), but will refer, on occasion, to findings from other herpesviruses. Because this work is intended to update the work on HSV DNA replication written for the previous edition of DNA Replication and Human Disease (Weller and Coen 2006), it draws primarily from work published during the last 6 years. Topics covered in detail in the previous work by the authors (Weller and Coen 2006) and reviewed elsewhere will be summarized only briefly (see Coen 2009; Weller 2010; Livingston and Kyratsous 2011; Ward and Weller 2011; Weitzman and Weller 2011).  相似文献   

15.
16.
DNA聚合酶高保真机理的新发现及其在SNP分析中的应用   总被引:3,自引:0,他引:3  
高保真DNA聚合酶在遗传与进化等生命活动中具有十分重要的生理与病理意义。高保真聚合酶除具有广为人知的校正功能外,最近的实验进一步表明, 由不能及时校正或难于纠正的错配碱基引发的“关”闭DNA聚合反应的效应, 同样保证了DNA聚合反应终产物的纯度。高保真聚合酶这一“关”闭DNA聚合反应的能力, 促成了其与耐外切酶消化的3´末端碱基特异性引物共同构成一个SNP敏感性纳米级复合分子“开/关”,高保真聚合酶分子中相距三纳米的聚合中心和3´→5´外切酶酶解中心则既合作又独立地起到了复合分子开关中“开”和“关”的效能:对于配对的引物,则直接在该酶的聚合中心进行聚合反应,即“开”的效应;而对于3´末端错配的引物,则从该酶的聚合中心转移至3´→5´外切酶的酶解中心,由于引物修饰了的3´末端耐外切酶的特点,继而出现了一种长时间无酶解产物的酶解过程,最后因酶的聚合中心空转而“关”闭DNA聚合反应,即“关”的效应。这一新的复合分子“开/关”在很大程度上满足了后基因时代对SNP分析的要求。该SNP分子开关的应用, 使基因诊断提高到单碱基水平。同时, 利用该方法通过SNP对基因组扫描, 在单基因遗传病病因研究及法医学鉴定上具有很强的理论和实用价值。  相似文献   

17.
单纯疱疹病毒1型(Herpes simplex virus type 1, HSV-1) UL42作为病毒编码的DNA聚合酶辅助亚基之一,是一种多功能蛋白,其在催化和调节病毒在细胞核内的有效复制发挥了重要的作用。已知UL42能提高DNA聚合酶催化亚基UL30的持续合成能力,激活病毒DNA聚合酶活性;介导DNA聚合酶的入核;与DNA模板链结合,提高病毒复制的保真度,以及含有抑制DNA聚合酶活性的肽段,提示其在病毒复制过程中也可能具有负调控作用。近期亦有报道显示,UL42能够阻断肿瘤坏死因子α(tumor necrosis factor-α, TNF-α)激活的核转录因子(nuclear factor kappa-B,NF-κB)信号通路以及干扰素调控因子3(interferon regulatory factor 3, IRF-3)的功能,提示其在病毒逃逸宿主天然免疫反应中发挥了一定的功能,但具体的作用机制尚不明确。本文对目前国内外HSV-1 UL42的结构特点、主要功能、作用机制及其在抗病毒药物研发中的研究进展进行综述,为后续揭示病毒致病机制和抗病毒药物的研发提供参考。  相似文献   

18.
In addition to adenoviruses, which are capable of completely helping adenovirus-associated virus (AAV) multiplication, only herpesviruses are known to provide any AAV helper activity, but this activity has been thought to be partial (i.e., AAV DNA, RNA, and protein syntheses are induced, but infectious particles are not assembled). In this study, however, we show that herpes simplex virus type 1 (HSV-1) and type 2 (HSV-2) are in fact complete AAV helpers and that AAV type 2 (AAV2) infectivity yields can approach those obtained when coinfections are carried out with a helper adenovirus. AAV helper activity was demonstrated in KB cells with two HSV-1 strains (11124 and 17MP) and an HSV-2 strain (HG52). Each herpesvirus supported AAV2 multiplication with comparable efficiency. AAV2 multiplication was similarly efficient in HSV-1 coinfections of HeLa cells, whereas lower yields were obtained in HEp-2 and primary human embryonic kidney cells. HSV-1 also supported AAV1 multiplication in HeLa cells but, at corresponding multiplicities of infection, AAV1 grew less efficiently than AAV2. Comparisons of the time courses of AAV2 DNA, RNA, and protein syntheses after coinfection with either adenovirus type 5 or HSV-1 revealed that, in each case, the onset of synthesis and attainment of maximal synthesis rate occurred earlier in coinfections with HSV-1. These findings demonstrate the linkage of AAV macromolecular synthesis to an event(s) in the helper virus cycle. Aside from this temporal association, helper-related differences in AAV macromolecular synthesis were not apparent.  相似文献   

19.
We recently reported that the herpes simplex virus 1 (HSV-1) Us3 protein kinase phosphorylates threonine at position 887 (Thr-887) in the cytoplasmic tail of envelope glycoprotein B (gB) (A. Kato, J. Arii, I. Shiratori, H. Akashi, H. Arase, and Y. Kawaguchi, J. Virol. 83:250-261, 2009; T. Wisner, C. C. Wright, A. Kato, Y. Kawaguchi, F. Mou, J. D. Baines, R. J. Roller and D. C. Johnson, J. Virol. 83:3115-3126, 2009). In the studies reported here, we examined the effect(s) of this phosphorylation on viral replication and pathogenesis in vivo and present data showing that replacement of gB Thr-887 by alanine significantly reduced viral replication in the mouse cornea and development of herpes stroma keratitis and periocular skin disease in mice. The same effects have been reported for mice infected with a recombinant HSV-1 carrying a kinase-inactive mutant of Us3. These observations suggested that Us3 phosphorylation of gB Thr-887 played a critical role in viral replication in vivo and in HSV-1 pathogenesis. In addition, we generated a monoclonal antibody that specifically reacted with phosphorylated gB Thr-887 and used this antibody to show that Us3 phosphorylation of gB Thr-887 regulated subcellular localization of gB, particularly on the cell surface of infected cells.The herpes simplex virus 1 (HSV-1) Us3 gene encodes a serine/threonine protein kinase with an amino acid sequence that is conserved in the subfamily Alphaherpesvirinae (9, 20, 29). The Us3 kinase phosphorylation target site has been reported to be similar to that of protein kinase A (PKA), a cellular cyclic AMP-dependent protein kinase (3, 12). Us3 catalytic activity plays important roles in viral replication and pathogenesis in vivo, based on studies showing that recombinant Us3 null mutant viruses and recombinant viruses encoding catalytically inactive Us3 have significantly reduced virulence, pathogenicity, and replication in mouse models (21, 34). In contrast, Us3 is not essential for growth in tissue culture cells (29). Thus, recombinant Us3 mutants grow as well as wild-type virus in Vero cells and have modestly impaired growth in a specific cell line such as HEp-2 cells (32, 33). The catalytic activity of Us3 is, in part, regulated by autophosphorylation of its serine at position 147 (Ser-147), and regulation of Us3 activity by autophosphorylation of Ser-147 appears to play a critical role in HSV-1 replication in vivo and in HSV-1 pathogenesis (34). Numerous studies have elucidated the potential downstream effects of Us3, including blocking apoptosis (18, 26-28), promoting nuclear egress of progeny nucleocapsids through the nuclear membrane (24, 32, 33), redistributing and phosphorylating nuclear membrane-associated viral nuclear egress factors UL31 and UL34 (13, 24, 30, 31) and cellular proteins including lamin A/C and emerin (16, 22, 23), controlling infected cell morphology (12, 27), and downregulating cell surface expression of viral envelope glycoprotein B (gB) (11).Two substrates that mediate some of the Us3 functions described above have been identified. First, it has been shown that Us3 phosphorylates Thr-887 in the cytoplasmic tail of gB, which appears to downregulate cell surface expression of gB (11). This conclusion is based on the observation that a T887A mutation in gB (gB-T887A) markedly upregulated cell surface expression of gB in infected cells: this upregulation was also observed with a recombinant virus encoding a Us3 kinase-inactive mutant, whereas a phosphomimetic substitution for gB Thr-887 restored wild-type cell surface expression of gB (11). Us3 phosphorylation of gB Thr-887 has also been proposed to be involved in regulation of fusion of the nascent progeny virion envelope with the cell''s outer nuclear membrane, based on the observation that virions accumulated aberrantly in the perinuclear space in cells infected with a mutant virus carrying the gB-T887A substitution mutation and lacking the capacity to produce gH (42). Second, it has been shown that Us3 may phosphorylate some or all of the six serines in the UL31 N-terminal region (24). Such phosphorylation might regulate proper localization of UL31 and UL34 at the nuclear membrane, nuclear egress of nucleocapsids, and viral growth in cell cultures since the Us3 kinase-inactive mutant phenotype for nuclear egress (i.e., mislocalization of UL31 and UL34 at the nuclear membrane, aberrant accumulation of virions within herniations of the nuclear membrane, and decreased viral growth in cell cultures) is also produced by replacement of the six serines in the UL31 N-terminal region with alanines while phosphomimetic substitutions of the six serines restored the wild-type phenotype (24).Thus, the molecular mechanisms for some of the downstream effects of Us3 phosphorylation have been gradually elucidated. However, it remains to be shown whether the Us3 functions reported to date are in fact involved in viral replication and pathogenicity in vivo. In the present study, we focused on Us3 phosphorylation of gB Thr-887 and examined the effect(s) of this phosphorylation on viral replication and pathogenesis in vivo. These studies have shown that replacement of gB Thr-887 by alanine significantly reduced viral replication in the mouse cornea and development of herpes stroma keratitis (HSK) and periocular skin disease in mice, as reported for infection of mice with a recombinant virus carrying a Us3 kinase-inactive mutant (34). These observations suggested that Us3 phosphorylation of gB Thr-887 played a critical role in viral replication in vivo and in HSV-1 pathogenesis. In addition, we generated a monoclonal antibody that specifically recognized phosphorylated gB Thr-887 and used this antibody to directly study the functional consequences of Us3 phosphorylation of gB Thr-887 in infected cells. We also present data showing that Us3 phosphorylation of gB Thr-887 regulated subcellular localization of gB, particularly gB localization on the cell surface of infected cells.  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