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.
2.
The parvovirus adeno-associated virus (AAV) contains a small single-stranded DNA genome with inverted terminal repeats that form hairpin structures. In order to propagate, AAV relies on the cellular replication machinery together with functions supplied by coinfecting helper viruses such as adenovirus (Ad). Here, we examined the host cell response to AAV replication in the context of Ad or Ad helper proteins. We show that AAV and Ad coinfection activates a DNA damage response (DDR) that is distinct from that seen during Ad or AAV infection alone. The DDR was also triggered when AAV replicated in the presence of minimal Ad helper proteins. We detected autophosphorylation of the kinases ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) and signaling to downstream targets SMC1, Chk1, Chk2, H2AX, and XRCC4 and multiple sites on RPA32. The Mre11 complex was not required for activation of the DDR to AAV infection. Additionally, we found that DNA-PKcs was the primary mediator of damage signaling in response to AAV replication. Immunofluorescence revealed that some activated damage proteins were found in a pan-nuclear pattern (phosphorylated ATM, SMC1, and H2AX), while others such as DNA-PK components (DNA-PKcs, Ku70, and Ku86) and RPA32 accumulated at AAV replication centers. Although expression of the large viral Rep proteins contributed to some damage signaling, we observed that the full response required replication of the AAV genome. Our results demonstrate that AAV replication in the presence of Ad helper functions elicits a unique damage response controlled by DNA-PK.Replication of viral genomes produces a large amount of extrachromosomal DNA that may be recognized by the cellular DNA damage machinery. This is often accompanied by activation of DNA damage response (DDR) signaling pathways and recruitment of cellular repair proteins to sites of viral replication. Viruses therefore provide good model systems to study the recognition and response to DNA damage (reviewed in reference 48). The Mre11/Rad50/Nbs1 (MRN) complex functions as a sensor of chromosomal DNA double-strand breaks (DSBs) and is involved in activation of damage signaling (reviewed in reference 41). The MRN complex also localizes to DNA DSBs and is found at viral replication compartments during infection with a number of DNA viruses (6, 40, 47, 70, 75, 77, 87, 93). The phosphatidylinositol 3-kinase-like kinases (PIKKs) ataxia telangiectasia mutated (ATM), ATM and Rad3-related kinase (ATR), and the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) are involved in the signal transduction cascades activated by DNA damage (reviewed in references 43, 51, and 71). These kinases respond to distinct types of damage and regulate DSB repair during different phases of the cell cycle (5), either through nonhomologous end-joining (NHEJ) or homologous recombination pathways (reviewed in references 63, 81, and 86). The DNA-PK holoenzyme is composed of DNA-PKcs and two regulatory subunits, the Ku70 and Ku86 heterodimer. DNA-PK functions with XRCC4/DNA ligase IV to repair breaks during NHEJ, and works with Artemis to process DNA hairpin structures during VDJ recombination and during a subset of DNA DSB events (46, 50, 86). While the kinase activity of DNA-PKcs leads to phosphorylation of a large number of substrates in vitro as well as autophosphorylation of specific residues (reviewed in references 16 and 85), it is currently unclear how DNA-PKcs contributes to signaling in cells upon different types of damage.The adeno-associated virus (AAV) genome consists of a molecule of single-stranded DNA with inverted terminal repeats (ITRs) at both ends that form double-hairpin structures due to their palindromic sequences (reviewed in reference 52). The ITRs are important for replication and packaging of the viral genome and for integration into the host genome. Four viral Rep proteins (Rep78, Rep68, Rep52, and Rep40) are also required for replication and packaging of the AAV genome into virions assembled from the Cap proteins. Although the Rep and Cap genes are replaced in recombinant AAV vectors (rAAV) that retain only the ITRs flanking the gene of interest, these vectors can be replicated by providing Rep in trans (reviewed in reference 7). Productive AAV infection requires helper functions supplied by adenovirus (Ad) or other viruses such as herpes simplex virus (HSV) (reviewed in reference 27), together with components of the host cell DNA replication machinery (54, 55, 58). In the presence of helper viruses or minimal helper proteins from Ad or HSV, AAV replicates in the nucleus at centers where the viral DNA and Rep proteins accumulate (35, 76, 84, 89). Cellular and viral proteins involved in AAV replication, including replication protein A (RPA), Ad DNA-binding protein (DBP), and HSV ICP8, localize with Rep proteins at these viral centers (29, 33, 76).A number of published reports suggest associations between AAV and the cellular DNA damage machinery. For example, transduction by rAAV vectors is increased by genotoxic agents and DNA damaging treatments (1, 62, 91) although the mechanisms involved remain unclear. Additionally, the ATM kinase negatively regulates rAAV transduction (64, 92), and we have shown that the MRN complex poses a barrier to both rAAV transduction and wild-type AAV replication (11, 67). UV-inactivated AAV particles also appear to activate a DDR involving ATM and ATR kinases that perturbs cell cycle progression (39, 60, 88). It has been suggested that this response is provoked by the AAV ITRs (60) and that UV-treated particles mimic stalled replication forks in infected cells (39). In addition to AAV genome components, the viral Rep proteins have been observed to exhibit cytotoxicity and induce S-phase arrest (3, 65).The role of cellular repair proteins in AAV genome processing has also been explored by examining the molecular fate of rAAV vectors, which are converted into circular and concatemeric forms that persist episomally (18, 19, 66). Proteins shown to regulate circularization in cell culture include ATM and the MRN complex (14, 64), while in vivo experiments using mouse models have implicated ATM and DNA-PK in this process (14, 20, 72). Additionally, DNA-PKcs and Artemis have recently been shown to cleave the ITR hairpins of rAAV vectors in vivo in a tissue-dependent manner (36). Despite these studies, it is not clear how damage response factors function together and how they impact AAV transduction and replication in human cells.In this study we examined the cellular response to AAV replication in the context of Ad infection or helper proteins. We show that coinfection with AAV and Ad activates a DDR that is distinct from that seen during infection with Ad alone. The ATM and DNA-PKcs damage kinases are activated and signal to downstream substrates, but the response does not require the MRN complex and is primarily mediated by DNA-PKcs. Although expression of the large Rep proteins induced some DDR events, full signaling appeared to require AAV replication and was accompanied by accumulation of DNA-PK at viral replication compartments. Our results demonstrate that AAV replication induces a unique DNA damage signal transduction response and provides a model system for studying DNA-PK.  相似文献   

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

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.
11.
A recent clinical trial in patients with hemophilia B has suggested that adeno-associated virus (AAV) capsid-specific cytotoxic T lymphocytes (CTLs) eliminated AAV-transduced hepatocytes and resulted in therapeutic failure. AAV capsids elicit a CTL response in animal models; however, these capsid-specific CTLs fail to kill AAV-transduced target cells in mice. To better model the human clinical trial data in mice, we introduced an immunodominant epitope derived from ovalbumin (OVA; SIINFEKL) into the AAV capsid and tested CTL-mediated killing of AAV2-transduced target tissues in vivo. Initially, in vitro experiments demonstrated both classical class I and cross-presentation of the OVA antigen, following endogenous expression or AAV2-OVA vector transduction, respectively. Furthermore, an OVA-specific CTL response was elicited after muscular or systemic injection of the AAV2-OVA vector. Finally, CTL reactivity was enhanced in mice with established SIINFEKL-specific immunity after AAV2-OVA/α1 anti-trypsin (AAT) administration. Most importantly, these OVA-specific CTLs decreased AAT expression in mice treated with AAV2-OVA/AAT vector that followed a time course mimicking uncoating kinetics of AAV2 transduction in OVA-immunized mice. These results demonstrate that AAV capsid-derived antigens elicit CD8+ CTL reactivity, and these CTLs eliminated AAV-transduced target cells in mice. Notably, this model system can be exploited to study the kinetics of capsid presentation from different serotypes of AAV and permit the design of novel strategies to block CTL-mediated killing of AAV-transduced cells.Adeno-associated virus (AAV) is a single-stranded DNA parvovirus. Its replication relies on coinfection of a helper virus such as adenovirus or herpesvirus. In the absence of a helper virus, AAV establishes latency to integrate into the AAVS1 site of host chromosome 19 (11). The genome of AAV is ∼4.7 kb and contains two open reading frames encoding replication proteins and structural capsid proteins (21). The capsid proteins (VP) are composed of VP1, VP2, and VP3. The VP3 protein is the major structural component and constitutes nearly 80% of the virion shell with an overall ratio of 1:1:8 for VP1, VP2, VP3, respectively. While VP2 is thought to be nonessential for AAV transduction (30), the VP1 subunit contains a phospholipase A2 domain required for infectivity (9). Recombinant AAV (rAAV) vectors require only the 145-bp terminal repeats of the AAV genome in cis and all other viral factors supplied in trans for production (18). rAAV vectors have rapidly gained popularity in gene therapy applications and have proven effective in preclinical studies/clinical trials for a number of diseases (20, 31, 33).AAV vectors mount a potent humoral immune response against capsid in animals and human. However, AAV vectors only contain the therapeutic gene flanked by two 145-bp AAV terminal repeats devoid of any AAV genes(23). In addition, AAV initiates long-term stable therapeutic gene expression in animal models (3-5, 17, 31). Based on these observations AAV has been thought to be relatively nonimmunogenic regarding the induction of cytotoxic T lymphocytes (CTLs) specific for capsid proteins. In spite of all of these observations, the recent clinical trial for hemophilia B (F9) gene therapy has otherwise suggested that AAV2 capsid initiates cell-mediated immunity that eliminates the AAV2 encoding F9 (AAV2/F9) vector transduced liver cells (15). Against this backdrop, numerous attempts to replicate aforementioned observations in animal models have been made. Preliminary results from these studies support direct presentation and cross-presentation of the AAV2 capsid in animal models (6, 12, 13, 22, 29). However, capsid-specific CTLs did not eliminate AAV2-transduced target cells in mice (12, 13, 29), inconsistent with observations made in a clinical trial for hemophilia B with AAV2/F9 gene therapy. A potential explanation for this discrepancy is the weak immunogenicity of the AAV2 capsid in mice. Accordingly, we hypothesized that incorporation of a peptide epitope into the AAV2 capsid would increase immunogenicity of the rAAV and therefore could be exploited to mimic events ongoing in humans and study approaches to block capsid-specific CTL reactivity in mice.We chose to introduce the MHC-H2Kb-restricted SIINFEKL peptide derived from ovalbumin (OVA) into AAV2 capsid. Integration of the OVA epitope into AAV capsids elicited a specific CTL response. Most importantly, after administration of genetically engineered AAV2 vectors into OVA peptide-immunized mice, OVA-specific CTL reactivity was further enhanced, thereby limiting transgene expression in vivo. The modified vector described herein is a potentially valuable tool for future studies focused on developing strategies to evade capsid-specific CTL-mediated elimination of AAV-transduced target cells in animal models.  相似文献   

12.
In animal models of infection, glycoprotein E (gE) is required for efficient herpes simplex virus type 1 (HSV-1) spread from the inoculation site to the cell bodies of innervating neurons (retrograde direction). Retrograde spread in vivo is a multistep process, in that HSV-1 first spreads between epithelial cells at the inoculation site, then infects neurites, and finally travels by retrograde axonal transport to the neuron cell body. To better understand the role of gE in retrograde spread, we used a compartmentalized neuron culture system, in which neurons were infected in the presence or absence of epithelial cells. We found that gE-deleted HSV-1 (NS-gEnull) retained retrograde axonal transport activity when added directly to neurites, in contrast to the retrograde spread defect of this virus in animals. To better mimic the in vivo milieu, we overlaid neurites with epithelial cells prior to infection. In this modified system, virus infects epithelial cells and then spreads to neurites, revealing a 100-fold retrograde spread defect for NS-gEnull. We measured the retrograde spread defect of NS-gEnull from a variety of epithelial cell lines and found that the magnitude of the spread defect from epithelial cells to neurons correlated with epithelial cell plaque size defect, indicating that gE plays a similar role in both types of spread. Therefore, gE-mediated spread between epithelial cells and neurites likely explains the retrograde spread defect of gE-deleted HSV-1 in vivo.Herpes simplex virus type 1 (HSV-1) is an alphaherpesvirus that characteristically infects skin and mucosal surfaces before spreading to sensory neurons, where it establishes a lifelong persistent infection. The virus periodically returns to the periphery via sensory axons and causes recurrent lesions as well as asymptomatic shedding. This life cycle requires viral transport along axons in two directions: toward the neuron cell body (retrograde direction) and away from the neuron cell body (anterograde direction).Many studies of alphaherpesvirus neuronal spread have focused on pseudorabies virus (PRV), a virus whose natural host is the pig. Three PRV proteins, glycoprotein E (gE), gI, and Us9, have been shown to mediate anterograde neuronal spread both in animal models of infection and in cultured neurons. However, these three proteins are dispensable for retrograde spread (3, 8, 11, 12, 31, 46). In contrast, numerous animal models of infection have shown that HSV-1 gE is required for retrograde spread from the inoculation site to the cell bodies of innervating neurons (4, 9, 44, 56). In the murine flank model, wild-type (WT) virus replicates in the skin and then infects sensory neurons and spreads in a retrograde direction to the dorsal root ganglia (DRG). In this model, gE-deleted HSV-1 replicates in the skin but is not detected in the DRG (9, 44). This phenotype differs from gE-deleted PRV, which is able to reach the DRG at WT levels (8). Thus, unlike PRV, gE-deleted HSV-1 viruses have a retrograde spread defect in vivo.HSV-1 gE is a 552-amino-acid type I membrane protein found in the virion membrane as well as in the trans-Golgi and plasma membranes of infected cells (1). gE forms a heterodimer with another viral glycoprotein, gI. The gE/gI complex is important for HSV-1 immune evasion through its Fc receptor activity. gE/gI binds to the Fc domain of antibodies directed against other viral proteins, sequestering these antibodies and blocking antibody effector functions (27, 32, 40). Additionally, gE/gI promotes spread between epithelial cells. Viruses lacking either gE or gI form characteristically small plaques in cell culture and small inoculation site lesions in mice (4, 9, 18, 40, 58). In animal models, gE and gI also mediate viral spread in both anterograde and retrograde directions (4, 19, 44, 56).In order to better understand the role of gE in HSV-1 retrograde neuronal spread, we employed a compartmentalized neuron culture system that has been used to study directional neuronal spread of PRV and West Nile virus (12, 14, 45). In the Campenot chamber system, neurites are contained in a compartment that is separate from their corresponding cell bodies. Therefore, spread in an exclusively retrograde direction can be measured by infecting neurites and detecting spread to neuron cell bodies.HSV-1 replication requires retrograde transport of incoming viral genomes to the nucleus. In neurites, fusion between viral and cellular membranes occurs at the plasma membrane (43, 48). Upon membrane fusion, the capsid and a subset of tegument proteins (the inner tegument) dissociate from glycoproteins and outer tegument proteins, which remain at the plasma membrane (28, 38). Unenveloped capsids and the associated inner tegument proteins are then transported in the retrograde direction to the nucleus (7, 48, 49).For both neurons and epithelial cells, retrograde transport is dependent upon microtubules, ATP, the retrograde microtubule motor dynein, and the dynein cofactor dynactin (22, 34, 49, 52). Several viral proteins interact with components of the dynein motor complex (23, 39, 60). However, none of these proteins suggest a completely satisfactory mechanism by which viral retrograde transport occurs, either because they are not components of the complex that is transported to the nucleus (UL34, UL9, VP11/12) or because capsids lacking that protein retain retrograde transport activity (VP26) (2, 17, 21, 28, 37). This implies that additional viral proteins are involved in retrograde trafficking.We sought to better characterize the role of gE in retrograde spread and found that gE is dispensable for retrograde axonal transport; however, it promotes HSV-1 spread from epithelial cells to neurites. This epithelial cell-to-neuron spread defect provides a plausible explanation for the retrograde spread defect of gE-deleted HSV-1 in animal models of infection.  相似文献   

13.
14.
15.
16.
Us3 protein kinases encoded by herpes simplex virus 1 (HSV-1) and 2 (HSV-2) are serine/threonine protein kinases and play critical roles in viral replication and pathogenicity in vivo. In the present study, we investigated differences in the biological properties of HSV-1 and HSV-2 Us3 protein kinases and demonstrated that HSV-2 Us3 did not have some of the HSV-1 Us3 kinase functions, including control of nuclear egress of nucleocapsids, localization of UL31 and UL34, and cell surface expression of viral envelope glycoprotein B. In agreement with the observations that HSV-2 Us3 was less important for these functions, the effect of HSV-2 Us3 kinase activity on virulence in mice following intracerebral inoculation was much lower than that of HSV-1 Us3. Furthermore, we showed that alanine substitution in HSV-2 Us3 at a site (aspartic acid at position 147) corresponding to one that can be autophosphorylated in HSV-1 Us3 abolished HSV-2 Us3 kinase activity. Thus, the regulatory and functional effects of Us3 kinase activity are different between HSV-1 and HSV-2.Us3 protein kinases encoded by herpes simplex virus 1 (HSV-1) and 2 (HSV-2) are serine/threonine protein kinases with amino acid sequences that are conserved in the subfamily Alphaherpesvirinae (6, 24, 36). Based on studies showing that recombinant Us3 mutants of HSV-1 and HSV-2 have significantly impaired viral replication and virulence in mice models, it has been concluded that both HSV-1 and HSV-2 Us3 protein kinases play important roles in viral replication and pathogenicity in vivo (25, 33, 41). In contrast, HSV-1 and HSV-2 Us3 protein kinases are not essential for growth in tissue culture cells (33, 36). Thus, recombinant Us3 mutants grow as well as wild-type viruses in Vero cells, and the mutants exhibit modestly impaired replication in HEp-2 cells (33, 36, 39, 40). The possible functions of Us3 have been extensively studied and gradually elucidated for HSV-1 Us3, but much less is known about HSV-2 Us3. These functions include (i) blocking apoptosis (1, 22, 30, 31, 35); (ii) promoting nuclear egress of progeny nucleocapsids through the nuclear membrane (39, 40, 45); (iii) redistributing and phosphorylating nuclear membrane-associated viral nuclear egress factors UL31 and UL34 (14, 37, 38) and cellular proteins, including lamin A/C and emerin (21, 27, 28); (iv) controlling infected cell morphology (13, 31, 32); and (v) downregulating cell surface expression of viral envelope glycoprotein B (gB) (12).To determine the molecular mechanisms for a viral protein kinase''s effects in infected cells, the kinase''s physiological substrates and its phosphorylation sites must be identified. This can involve studies showing that the altered phenotypes observed in cells infected with a mutant virus lacking the protein kinase activity is also detected in cells infected with a mutant virus in which the substrate''s phosphorylation sites have been modified by mutations. Although more than 15 potential HSV Us3 substrates have been reported, HSV-1 Us3 phosphorylation of only three substrates (Us3 itself, gB, and UL31) has been demonstrated to be linked directly with Us3 functions in infected cells (12, 13, 29, 41) as follows. (i) Us3 has been reported to autophosphorylate serine at position 147 (Ser-147), and this phosphorylation augments Us3''s kinase activity in infected cells (13, 41). Even though only a small fraction of Us3 is autophosphorylated at Ser-147 in infected cells, alanine replacement of Ser-147 in Us3 significantly reduced HSV-1 replication in the mouse cornea and pathogenic manifestations of herpes stroma keratitis and periocular skin disease in mice (41). These results indicated that Us3 kinase activity was, in part, regulated by autophosphorylation of Ser-147, and regulation of Us3 activity by autophosphorylation played a critical role in viral replication in vivo and HSV-1 pathogenesis. (ii) It has been reported that HSV-1 Us3 phosphorylates Thr-887 in the cytoplasmic tail of gB, and this phosphorylation downregulates the cell surface expression of gB (12). Us3 phosphorylation of gB at Thr-887 also has been proposed to be involved in the 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 mutant viruses carrying the amino acid substitution mutation T887A in gB and lacking the capacity to produce gH (45). The Us3 phosphorylation of gB at Thr-887 appeared to be critical for HSV-1 replication and pathogenesis in vivo, based on studies showing that the T887A substitution in the phosphorylation site in gB significantly reduced viral replication in the mouse cornea and pathogenic manifestations of herpes stroma keratitis and periocular skin disease in mice (Takahiko Imai, Ken Sagou, and Yasushi Kawaguchi, unpublished observations). (iii) It has been shown that Us3 phosphorylated some or all of the six serines in the UL31 N-terminal region, and this phosphorylation regulated the proper localization of UL31 and UL34 at the nuclear membrane and nuclear egress of nucleocapsids (29). Thus, the molecular basis of HSV-1 Us3 effects in infected cells have been gradually elucidated.However, the Us3 phosphorylation sites in Us3 itself and in gB are not conserved between HSV serotypes (12, 13). The amino acid residues in HSV-2 Us3 and gB corresponding to HSV-1 Us3 Ser-147 and gB Thr-887 are aspartic acid (Asp-147) and alanine (Ala-887), respectively. These results suggest that some HSV-1 Us3 functions, such as regulation of its own catalytic activity and control of gB expression on the cell surface, are not regulated by HSV-2 Us3 or are regulated in a manner(s) different from HSV-1 Us3. In agreement with this suggestion, there is a marked difference between HSV-1 and HSV-2 virulence in mice following intracerebral infection, with the HSV-1 Us3 null mutant being >104-fold less virulent than the parent wild-type virus (25), while the HSV-2 Us3 null mutant was only ∼10-fold less virulent (33). Although these results were from different reports and the mouse strains used in the studies were different, they indicate that some HSV-1 Us3 functions are different from those of HSV-2 Us3.Therefore, we investigated differences in the biological properties of HSV-1 and HSV-2 Us3 protein kinases. It was of particular interest to examine whether Asp-147 in HSV-2 Us3 is required for its own kinase activity, since it is well established that acidic amino acids such as Asp or glutamic acid sometimes mimic the negative charges produced by phosphorylation (29, 46). In the present study, using a genetic manipulation system of HSV-2 with our newly constructed HSV-2 bacterial artificial chromosome (BAC) clone, we have shown that HSV-2 Us3 exhibited marked differences from HSV-1 Us3 in its catalytic functions, including the regulation of UL31/UL34 localization, nuclear egress of nucleocapsids, cell surface expression of gB, and virulence in mice. We also found that Asp-147 in HSV-2 Us3 was critical for its kinase activity, raising a possibility that the activity of Us3 kinases was regulated differently in HSV-1 and HSV-2.  相似文献   

17.
18.
Although the herpes simplex virus type 1 (HSV-1) genome might be expected to induce a DNA damage response, the ATR kinase is not activated in infected cells. We previously proposed that spatial uncoupling of ATR from its interaction partner, ATRIP, could be the basis for inactivation of the ATR kinase in infected cells; however, we now show that ATR and ATRIP are in fact both recruited to HSV-1 replication compartments and can be coimmunoprecipitated from infected-cell lysates. ATRIP and replication protein A (RPA) are recruited to the earliest detectable prereplicative sites, stage II microfoci. In a normal cellular DNA damage response, ATR/ATRIP are recruited to stretches of RPA-coated single-stranded DNA in an RPA- and kinase-dependent manner, resulting in the phosphorylation of RPA by ATR in damage foci. In contrast, in HSV-1-infected cells, RPA is not phosphorylated, and endogenous phosphorylated RPA is excluded from stage II microfoci; in addition, the recruitment of ATR/ATRIP is independent of RPA and the kinase activity of ATR. Furthermore, we show that ATR/ATRIP play a beneficial role in viral gene expression and virus production. Although ICP0 has been shown to be important for partial inactivation of other cellular DNA repair pathways, we show that ICP0 is not responsible for the inactivation of ATR signaling and, furthermore, that neither ATR nor ATRIP is a target of ICP0 degradation. Thus, ATR and ATRIP may function outside the context of the canonical ATR damage signaling pathway during HSV-1 infection to participate in the viral life cycle.Herpes simplex virus type 1 (HSV-1) is a large linear double-stranded DNA virus that replicates in the nucleus of the host cell. The incoming viral genome contains nicks and gaps (42), and cellular DNA repair machinery might be expected to recognize it as damaged, resulting in the activation of one or more cellular DNA damage pathways. Activation of DNA damage response pathways can result not only in repair of the damaged DNA but also in cell cycle arrest, gene silencing, and apoptosis (9). The later outcomes could result in suppression of viral gene expression and DNA replication and thus have negative consequences for lytic infection. Activation of a cellular DNA damage response during viral infection could, therefore, represent a form of intrinsic antiviral immunity (14, 15). On the other hand, HSV-1 and other DNA viruses which replicate in the nucleus have also been shown to utilize cellular DNA repair machinery to promote productive infection (28). Thus, HSV-1 has apparently evolved to manipulate the host DNA damage response by utilizing some components and inactivating others in an attempt to create an environment conducive to lytic viral infection.The cellular DNA damage response is regulated by the three phosphoinositide 3-kinase-related kinases (PIKKs), DNA-PK (DNA-dependent protein kinase), ATM (ataxia-telangiectasia-mutated), and ATR (ATM and Rad3-related) (1, 9). DNA-PK and ATM respond predominantly to double-strand breaks, and ATR responds to stalled replication forks and long stretches of single-stranded DNA (ssDNA). DNA-PK is required for nonhomologous end joining (NHEJ), while ATM activation promotes homologous recombination. Interestingly, in some cell types, the catalytic subunit of DNA-PK (DNA-PKcs) is proteolytically degraded during infection by the immediate-early (IE) protein ICP0, a viral E3 ubiquitin ligase (25, 37), thereby resulting in the probable inactivation of the NHEJ pathway. ATM kinase activity, on the other hand, is activated during HSV-1 infection once viral DNA replication is initiated (26, 47, 56). Despite phosphorylation of several ATM targets, ATM signaling is also modulated by ICP0, which degrades the ubiquitin ligases RNF8 and RNF168. The function of these ubiquitin ligases is to promote the tethering of ATM pathway proteins at sites of cellular DNA damage (27). Thus, ICP0 functions to partially inactivate portions of both the DNA-PK- and ATM-mediated repair pathways.During a cellular DNA damage response, ATM activation and processing of DNA ends generate ssDNA adjacent to double-stranded DNA (dsDNA), a structure that is known to activate ATR (9, 38). The ssDNA is coated by the cellular ssDNA binding protein, replication protein A (RPA), which then serves to recruit ATR through a direct interaction with ATR-interacting protein (ATRIP) (4, 12, 58). ATR signaling results in the phosphorylation of many substrates, including RPA and Chk1. During HSV-1 infection, the ATR substrates RPA and Chk1 are not phosphorylated (47, 54-56), indicating that ATR signaling may be disabled.A hallmark of HSV-1 infection is the reorganization of the infected-cell nucleus, resulting in the formation of large globular replication compartments as well as the rearrangement of cellular proteins involved in several homeostatic pathways. In addition to cellular DNA repair proteins, HSV-1 infection also causes the reorganization of components of the cellular protein quality control pathways, resulting in the formation of virus-induced chaperone-enriched (VICE) domains, which act to maintain nuclear protein quality control during infection (31). Viral gene expression, DNA replication, and encapsidation of viral genomes occur in replication compartments (24, 39, 41). In this work we revisit the study of proteins recruited to and restricted from replication compartments in an attempt to better understand how HSV-1 manipulates components of the cellular DNA damage response for its own benefit.  相似文献   

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