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.
Pseudorabies virus (PRV) Us9 is a small, tail-anchored (TA) membrane protein that is essential for axonal sorting of viral structural proteins and is highly conserved among other members of the alphaherpesvirus subfamily. We cloned the Us9 homologs from two human pathogens, varicella-zoster virus (VZV) and herpes simplex virus type 1 (HSV-1), as well as two veterinary pathogens, equine herpesvirus type 1 (EHV-1) and bovine herpesvirus type 1 (BHV-1), and fused them to enhanced green fluorescent protein to examine their subcellular localization and membrane topology. Akin to PRV Us9, all of the Us9 homologs localized to the trans-Golgi network and had a type II membrane topology (typical of TA proteins). Furthermore, we examined whether any of the Us9 homologs could compensate for the loss of PRV Us9 in anterograde, neuron-to-cell spread of infection in a compartmented chamber system. EHV-1 and BHV-1 Us9 were able to fully compensate for the loss of PRV Us9, whereas VZV and HSV-1 Us9 proteins were unable to functionally replace PRV Us9 when they were expressed in a PRV background.Alphaherpesviruses are classified by their variable host range, short reproductive cycle, and ability to establish latency in the peripheral nervous system (PNS) (36, 37). Commonly studied pathogens of this subfamily include herpes simplex virus (HSV) and varicella-zoster virus (VZV), as well as the veterinary pathogens pseudorabies virus (PRV), equine herpesvirus (EHV), and bovine herpesvirus (BHV). Initial infection begins with the virus entering the host mucosal surfaces and spreading between cells of the mucosal epithelium. Invariably, virus enters the PNS through the infection of peripheral nerves that innervate this region. The virus establishes a latent infection in PNS neurons that can be reactivated and that persists for the life of the host (36). In most natural infections, virus replication in the PNS never spreads to the central nervous system (CNS). However, on rare occasions, invasion of the CNS does occur, resulting in devastating encephalitis (46). Trafficking of virus particles from infected epithelial cells into the axon and subsequent transport to neuronal cell bodies is known as retrograde spread of infection. Trafficking of virus particles that are assembled in the neuronal cell body and subsequently sorted into axons for transport to epithelial cells at the initial site of infection (upon reactivation from latency) is known as anterograde spread of infection.Though the natural host of PRV is swine, the virus infects a wide variety of animals, including rodents, cats, dogs, rabbits, cattle, and chicken embryos, but not higher primates (1, 30, 47). In contrast to the well-contained spread of PRV within its natural host, infection of other mammals is usually lethal. Instead of stopping in the PNS, infection continues on to second-order and third-order neurons in the CNS (reviewed in reference 35). This facet of PRV infection makes it a useful tracer of neuronal connections (18). Work in our lab has identified three PRV proteins, Us9 and the gE/gI heterodimer, which are critical for efficient anterograde spread of infection in vivo (i.e., spread from presynaptic to postsynaptic neurons) (6, 45). The molecular mechanism by which these proteins function has been further elucidated in vitro using primary neuronal cultures of superior cervical ganglion (SCG) harvested from embryonic rat pups. PRV Us9 and, to a lesser extent, gE/gI are required for efficient axonal targeting of viral structural proteins, a necessary step for subsequent anterograde, transneuronal spread (10, 11, 27, 28, 42).PRV Us9 is a type II, tail-anchored (TA) membrane protein that is highly enriched in lipid raft microdomains and resides predominantly in or near the trans-Golgi network (TGN) inside infected cells (5-7, 27). It has homologs in most of the alphaherpesviruses, including VZV (16), HSV-1 (22), HSV-2 (17), EHV-1 (21, 40), EHV-4 (41), BHV-1 (25), and BHV-5 (14). Though several studies have examined individually the Us9 proteins encoded by VZV (16), HSV-1 (4, 22, 34, 39), BHV-1 (13), and BHV-5 (14), several gaps in our understanding of Us9 biology remain, namely, whether all of the PRV Us9 homologs are type II membrane proteins, if the proteins localize to similar subcellular compartments within different cell types, and if they can functionally substitute for the loss of PRV Us9 in axonal sorting and anterograde spread of infection. The aim of this study is to examine PRV Us9 in parallel assays with its homologs from VZV, HSV-1, EHV-1, and BHV-1 to identify potential similarities and differences between these highly conserved alphaherpesvirus proteins.  相似文献   

3.
Upon entry, neuroinvasive herpesviruses traffic from axon terminals to the nuclei of neurons resident in peripheral ganglia, where the viral DNA is deposited. A detailed analysis of herpes simplex virus type 1 (HSV-1) transport dynamics in axons following entry is currently lacking. Here, time lapse fluorescence microscopy was used to compare the postentry viral transport of two neurotropic herpesviruses: HSV-1 and pseudorabies virus (PRV). HSV-1 capsid transport dynamics were indistinguishable from those of PRV and did not differ in neurons of human, mouse, or avian origin. Simultaneous tracking of capsids and tegument proteins demonstrated that the composition of actively transporting HSV-1 is remarkably similar to that of PRV. This quantitative assessment of HSV-1 axon transport following entry demonstrates that HSV-1 and PRV share a conserved mechanism for postentry retrograde transport in axons and provides the foundation for further studies of the retrograde transport process.Herpes simplex virus type 1 (HSV-1) and the veterinary herpesvirus pathogen pseudorabies virus (PRV) establish latent infections within the peripheral nervous systems (PNS) of their hosts. Neurotropic herpesviruses gain access to the PNS at nerve endings present in infected skin or mucosal tissue. Upon entry at the nerve terminal, viral particles are transported in axons toward the neuronal cell body to ultimately deposit the viral genome into the nucleus. This process is referred to as retrograde transport and is critical for the establishment of latency. Following reactivation, progeny viral particles travel anterogradely from the ganglia toward the nerve terminals, resulting in reinfection of the dermis or other innervated tissues. Reactivated infection can manifest in various forms, including asymptomatic virus shedding or mild focal lesions (herpes labialis), or less frequently in more-severe disease (herpes keratitis, encephalitis, and in the case of varicella-zoster virus, shingles).All herpesviruses consist of an icosahedral capsid that contains the viral genome surrounded by a layer of proteins known as the tegument, which is contained within a membrane envelope (33). HSV-1 and PRV capsids disassociate from the viral envelope (2, 13, 14, 22, 23, 25, 28, 30, 40) and several tegument proteins (13, 16, 21, 25) upon fusion-mediated entry into cells. However, following entry into epithelial cell lines, the VP1/2 and UL37 tegument proteins are detected in association with cytosolic capsids of PRV by immunogold electron microscopy (16) and colocalize with HSV-1 capsids at the nuclear membrane by immunofluorescence microscopy (8). In primary sensory neurons, VP1/2 and UL37 are observed to be cotransported with PRV capsids during retrograde transport by time lapse fluorescence microscopy (21), and the kinetics of axon transport have been assessed (39).Although HSV-1 and PRV share similarities in their neurotropism in vivo (reviewed in reference 12), studies of axon transport have indicated possible mechanistic differences relevant to the underlying cell biology of neural transmission (reviewed in reference 10). As a result, a live-cell analysis comparing PRV and HSV-1 is needed to determine if axon transport mechanisms are conserved between the two neuroinvasive herpesvirus genera: Simplexvirus (HSV-1) and Varicellovirus (PRV). In this study, the retrograde transport process that delivers capsids to the nuclei of sensory neurons was compared for HSV-1 (strains KOS and F) and PRV (strain Becker).  相似文献   

4.
5.
Immediate-early protein ICP0 of herpes simplex virus type 1 (HSV-1) is an E3 ubiquitin ligase of the RING finger class that is required for efficient lytic infection and reactivation from latency. Other alphaherpesviruses also express ICP0-related RING finger proteins, but these have limited homology outside the core RING domain. Existing evidence indicates that ICP0 family members have similar properties, but there has been no systematic comparison of the biochemical activities and biological functions of these proteins. Here, we describe an inducible cell line system that allows expression of the ICP0-related proteins of bovine herpes virus type 1 (BHV-1), equine herpesvirus type 1 (EHV-1), pseudorabies virus (PRV), and varicella-zoster virus (VZV) and their subsequent functional analysis. We report that the RING domains of all the proteins have E3 ubiquitin ligase activity in vitro. The BHV-1, EHV-1, and PRV proteins complement ICP0-null mutant HSV-1 plaque formation and induce derepression of quiescent HSV-1 genomes to levels similar to those achieved by ICP0 itself. VICP0, the ICP0 expressed by VZV, was found to be extremely unstable, which limited its analysis in this system. We compared the abilities of the ICP0-related proteins to disrupt ND10, to induce degradation of PML and Sp100, to affect key components of the interferon signaling pathway, and to interfere with induction of interferon-stimulated genes. We found that the property that correlated most closely with their biological activities was the ability to preclude the recruitment of cellular ND10 proteins to sites closely associated with incoming HSV-1 genomes and early replication compartments.The members of the alphaherpesvirus subfamily are characterized by their ability to establish life-long latent infections in neuronal tissues after the primary infection. Although certain core genes are conserved in all herpesviruses of all subfamilies, there are also genes that are characteristic of particular subfamilies. Among these are the genes that encode the ICP0-related proteins of the alphaherpesviruses, of which the most widely studied is ICP0 of herpes simplex virus type 1 (HSV-1). The interest in ICP0 stems from its biological roles in stimulating lytic infection and reactivation from latency (for reviews, see references 17, 18 and 33). Members of the ICP0 family of proteins are characterized by the presence of a RING finger domain near their N termini, a zinc-stabilized fold that in many other proteins confers E3 ubiquitin ligase activity (43). This has proved to be true of ICP0 (3), and the available evidence indicates that other members of the ICP0 family have similar biochemical functions (13, 61). Although a number of ICP0-related alphaherpesvirus proteins have been studied in a variety of contexts, notably those expressed by bovine herpesvirus 1 (BHV-1), equine herpes virus 1 (EHV-1), pseudorabies virus (PRV), and varicella-zoster virus (VZV), there has been no systematic comparison of their abilities to complement ICP0 null mutant HSV-1 or to induce derepression of quiescent HSV-1 genomes.This paper describes a comparative study of the ICP0-related proteins expressed by the viruses listed above. In terms of nomenclature, the proteins expressed by BHV-1 and EHV-1 have been named BICP0 and EICP0, so although other names have been used for the PRV and VZV proteins (such as EP0 and orf61, respectively), we have adopted the names PICP0 and VICP0 for this study. Previous work found that, like ICP0 itself, all four proteins activate gene expression in reporter assays in a RING finger-dependent manner (4, 5, 8, 29, 38, 41, 45, 51, 54, 59, 64, 75, 76, 78). VICP0 and EICP0 also complement, at least partially, ICP0 null mutant HSV-1 (15, 48, 53, 54). BHV-1, EHV-1, PRV, and VZV mutants in which the ICP0-related genes have been deleted have been isolated and found to have reduced replication efficiencies, as expected by analogy with ICP0 null mutant HSV-1 (2, 7, 11, 12, 30, 46, 74, 77).A prominent property of ICP0 is its localization to and disruption of cellular nuclear substructures known as ND10 or promyelocytic leukemia (PML) nuclear domains. Interactions between ND10 and BICP0, EICP0, PICP0, and VICP0 have also been observed, with various consequences for ND10 integrity (47, 60, 63). Whereas ICP0 achieves ND10 disruption through induction of the degradation of PML and SUMO-modified forms of Sp100 (21, 60), EICP0 appears less efficient than ICP0 in inducing PML degradation (60) while VICP0 is inactive (47). While it is likely that all the ICP0 family members discussed here have RING finger-mediated E3 ubiquitin ligase activity (61), the only other protein for which this has been confirmed is BICP0 (13).The similarities between these members of the ICP0 family of proteins and their apparent differences prompted us to investigate in more detail the properties of these proteins in order to determine which of their properties correlate most closely with biological functions in complementing ICP0 null mutant HSV-1. In addition, there was no existing evidence on whether the related proteins could, like ICP0, induce derepression of gene expression from quiescent HSV-1 genomes. We have taken two approaches to these issues. The first is the use of an inducible cell line system that has been used to study ICP0 itself (24, 26). Although inducible cell line systems have been described for VICP0 and BICP0 (53, 69), much of the work described in the current study is novel. The second approach is in vitro analysis of the E3 ubiquitin ligase activities of the isolated RING finger domains of the proteins. The major findings of the study are the following: (i) that all the proteins studied are active in E3 ubiquitin ligase assays; (ii) that VICP0 is extremely unstable, compromising comparative functional analysis in this system; (iii) that BICP0, EICP0, and PICP0 complement to various degrees the plaque-forming defect of ICP0 null mutant HSV-1; (iv) that these three proteins also efficiently stimulate derepression of gene expression from quiescent HSV-1 genomes; (v) that none of the ICP0 family members impedes interferon (IFN)-induced expression of IFN-stimulated genes (ISGs) or affects the stability of important components of the IFN signaling system (namely STAT1, STAT2, and IRF3); (vi) that BICP0, EICP0, and PICP0 cause some disruption of ND10 integrity and have various effects on PML and Sp100 abundance; and (vii) that the property of the proteins that correlated most closely with their stimulation of ICP0 null mutant HSV-1 infection and derepression of quiescent genomes is their ability to inhibit the recruitment of PML and other ND10 proteins to sites associated with parental HSV-1 genomes and early replication compartments.  相似文献   

6.
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   

7.
8.
9.
10.
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   

11.
Glycoprotein B (gB), the most conserved protein in the family Herpesviridae, is essential for the fusion of viral and cellular membranes. Information about varicella-zoster virus (VZV) gB is limited, but homology modeling showed that the structure of VZV gB was similar to that of herpes simplex virus (HSV) gB, including the putative fusion loops. In contrast to HSV gB, VZV gB had a furin recognition motif ([R]-X-[KR]-R-|-X, where | indicates the position at which the polypeptide is cleaved) at residues 491 to 494, thought to be required for gB cleavage into two polypeptides. To investigate their contribution, the putative primary fusion loop or the furin recognition motif was mutated in expression constructs and in the context of the VZV genome. Substitutions in the primary loop, W180G and Y185G, plus the deletion mutation Δ491RSRR494 and point mutation 491GSGG494 in the furin recognition motif did not affect gB expression or cellular localization in transfected cells. Infectious VZV was recovered from parental Oka (pOka)-bacterial artificial chromosomes that had either the Δ491RSRR494 or 491GSGG494 mutation but not the point mutations W180G and Y185G, demonstrating that residues in the primary loop of gB were essential but gB cleavage was not required for VZV replication in vitro. Virion morphology, protein localization, plaque size, and replication were unaffected for the pOka-gBΔ491RSRR494 or pOka-gB491GSGG494 virus compared to pOka in vitro. However, deletion of the furin recognition motif caused attenuation of VZV replication in human skin xenografts in vivo. This is the first evidence that cleavage of a herpesvirus fusion protein contributes to viral pathogenesis in vivo, as seen for fusion proteins in other virus families.Varicella-zoster virus (VZV), an alphaherpesvirus, causes chicken pox (varicella) as a primary infection and shingles (zoster) upon reactivation from infected ganglia in humans (reviewed in reference 16). Although not yet investigated in VZV, herpesvirus entry requires fusion of the virus envelope with cell membranes governed by viral glycoprotein B (gB) and gH/gL, which are conserved across the family Herpesviridae (12, 27, 57). gB is the most conserved glycoprotein, with its function as a fusion protein well documented for several of the herpesviruses (10, 19, 38, 48, 51, 52).Open reading frame 31 (ORF31) codes for the 931 amino acids of VZV gB (18, 37). The successive N- and O-linked glycosylation plus the sialation and sulfation of VZV gB yields a mature protein with a molecular mass of approximately 140 kDa (45). Upon maturation, gB is cleaved, presumably by cellular proteases, into two polypeptides of 66 and 68 kDa. Intracellular trafficking of gB was shown to be dependent upon amino acid motifs in the cytoplasmic domain (24-26). In transfection studies, gB was transported to the cellular surface where it was endocytosed and localized to the trans-Golgi, where envelopment of viral particles is thought to occur.The structures of gB in the two human alphaherpesviruses, VZV and herpes simplex virus type 1 (HSV-1), are likely to be very similar as they have 49% amino acid identity (reviewed in reference 16). The ectodomain of HSV-1 gB was shown to form a spike that consisted of trimers with the structural homology to gG of vesicular stomatitis virus (28). Heldwein et al. (28) proposed that HSV-1 gB is a class II fusion protein based on homology to VSV G. The herpesvirus gB monomer was divided into five domains, I to V. Domain I consisted of a continuous amino acid sequence that folded into a pleckstrin homology-like domain, while domain II was comprised of two discontinuous segments, which also had a pleckstrin homology-like domain. A loop region exposed to the exterior of gB connected domain II with domain III. Domain III was comprised of three discontinuous segments and connected to the external loop by a long α helix that ended in a central coiled coil. Domain IV crowned gB and was connected to domain V, which stretched from the top to the bottom of the gB monomer, forming the core of the trimer making contacts with the two other subunits. The structural homology and lack of furin cleavage suggest that the herpesvirus gB and VSV G proteins have undergone convergent evolution.Although not proven experimentally, VZV gB is likely to be cleaved by the subtilisin-like proprotein convertase furin as the glycoprotein has a furin recognition motif [R]-X-[KR]-R-|-X (where | indicates the position at which the polypeptide is cleaved) (29). The [R]-X-[KR]-R-|-X motif is conserved in gBs for all of the herpesvirus families (5, 9, 21, 36, 40, 53, 63, 64). This site has been shown to be dispensable for the replication of human cytomegalovirus (HCMV), bovine herpesvirus type 1 (BHV-1), and pseudorabies virus (PRV) in vitro (32, 49, 58). Furin site mutants for BHV-1 and PRV show an altered phenotype in vitro, but effects were not examined in vivo. HSV-1 gB is not cleaved and lacks the [R]-X-[KR]-R-|-X motif at the canonical site, which is of interest because HSV-1 is genetically the most closely related human herpesvirus to VZV.Domain I of HSV gB showed structural conservation of putative fusion loops similar to those found in domain IV of the VSV G protein (28). Despite the lack of conserved amino acids within these loops, the hydrophobicity of the residues appears to be conserved for the Herpesviridae (4). Substitution of hydrophobic residues in Epstein-Barr virus gB and linker insertion mutagenesis close to the putative fusion loops of HSV-1 gB abrogated fusion based on in vitro transfection studies (4, 22, 34). However, the effect of substitutions in these putative fusion loops on viral replication has not been characterized. Since the development of fusion assays for VZV has proven elusive, the effect of substitutions in the putative fusion loop using viral mutagenesis to make recombinant viruses provides an alternative approach for identifying functional residues in VZV gB.In contrast to HSV-1, VZV is a human-restricted pathogen (reviewed in reference 16). To study the pathogenesis of VZV in vivo, well-established human xenograft models have been developed using SCID mice (6, 7, 13, 14, 41, 44, 54, 65). Lesions formed by VZV in the skin are similar to those seen in human subjects following primary infection (15, 43). The relevance of the model was demonstrated by studies with the varicella vaccine virus (vOka) that exhibited decreased growth in skin xenografts in vivo but does not cause disease in the healthy human host. In contrast, the vaccine virus and its parent strain, parental Oka (pOka), have indistinguishable replication kinetics in vitro (15, 43).The present study was designed to investigate the effects of structure-based targeted mutations in VZV gB on viral replication in cultured cells and in human skin xenografts in the SCIDhu mouse model. This was performed in the context of infectious virus recovered using the self-excisable bacterial artificial chromosome (BAC) containing the genome of a clinical isolate, Oka (62). The roles of the conserved residues W180 (gB-W180G) and Y185 (gB-Y185G) in the putative fusion loop were evaluated using glycine substitution, and the role of the furin recognition motif (491RSRR494) was assessed by a complete deletion of the furin motif (gBΔ491RSRR494) or a substitution of the arginine residues with glycine (gB491GSGG494) to conserve the carbon backbone.  相似文献   

12.
13.
14.
15.
16.
Varicella-zoster virus (VZV) causes varicella and herpes zoster, diseases characterized by distinct cutaneous rashes. Dendritic cells (DC) are essential for inducing antiviral immune responses; however, the contribution of DC subsets to immune control during natural cutaneous VZV infection has not been investigated. Immunostaining showed that compared to normal skin, the proportion of cells expressing DC-SIGN (a dermal DC marker) or DC-LAMP and CD83 (mature DC markers) were not significantly altered in infected skin. In contrast, the frequency of Langerhans cells was significantly decreased in VZV-infected skin, whereas there was an influx of plasmacytoid DC, a potent secretor of type I interferon (IFN). Langerhans cells and plasmacytoid DC in infected skin were closely associated with VZV antigen-positive cells, and some Langerhans cells and plasmacytoid DC were VZV antigen positive. To extend these in vivo observations, both plasmacytoid DC (PDC) isolated from human blood and Langerhans cells derived from MUTZ-3 cells were shown to be permissive to VZV infection. In VZV-infected PDC cultures, significant induction of alpha IFN (IFN-α) did not occur, indicating the VZV inhibits the capacity of PDC to induce expression of this host defense cytokine. This study defines changes in the response of DC which occur during cutaneous VZV infection and implicates infection of DC subtypes in VZV pathogenesis.Varicella-zoster virus (VZV) is a highly species-specific human herpesvirus that causes the diseases varicella (chicken pox) and herpes zoster (shingles). Varicella results from the primary phase of infection and is characterized by a diffuse rash of vesiculopustular lesions that appear in crops and usually resolve within 1 to 2 weeks (7, 26). Primary infection is initiated by inoculation of mucosal sites, such as the upper respiratory tract and the conjunctiva, with infectious virus, usually contained within respiratory droplets (3, 23). Following inoculation, there is a 10- to 21-day incubation period during which VZV is transported to the regional lymph nodes; however, it remains unclear which cell types are responsible for transport of VZV during natural infection (3). It has been hypothesized that dendritic cells (DC) of the respiratory mucosa may be among the first cells to encounter VZV during primary infection and are capable of virus transport to the draining lymph nodes (1, 45). It is postulated that within lymph nodes, VZV undergoes a period of replication, resulting in a primary cell-associated viremia, during which time virus is transported to the reticuloendothelial organs, where it undergoes another period of replication that results in a secondary cell-associated viremia and virus transport to the skin (3, 23). However, VZV has recently been shown to have tropism for human tonsillar CD4+ T lymphocytes (37), and it has been demonstrated that these T lymphocytes express skin homing markers that may allow them to transport VZV directly from the lymph node to the skin during primary viremia (38). Once the virus reaches the skin, it infects cutaneous epithelial cells, resulting in distinctive vesiculopustular lesions.During the course of primary infection, VZV establishes a lifelong latent infection within the sensory ganglia, from which virus may reactivate years later to cause herpes zoster (22, 42, 53). VZV reactivation results in the production of new infectious virus and a characteristic vesiculopustular rash, which differs from that of varicella insofar as the distribution of the lesions is typically unilateral and covers only 1 to 2 dermatomes (8). In both primary and reactivated VZV infection of human skin, VZV antigens are detectable in the epidermis and dermis (2, 30, 46, 47, 49, 52), and although some studies have examined the immune infiltrate present in these lesions, most have focused on T lymphocytes, macrophages, and NK cells (40, 48, 50, 51, 58). The role of DC subsets in VZV infection in human skin has not been previously explored in vivo.Our laboratory provided the first evidence that VZV could productively infect human immature and mature monocyte-derived dendritic cells (MDDC) in vitro (1, 45), and Hu and Cohen (2005) showed that VZV ORF47 was critical for replication of virus in human immature DC but not mature DC (29). However, whether DC become directly infected during natural VZV skin infection and the impact VZV infection may have on DC subsets has yet to be elucidated. The two subsets of DC that are normally present in the skin and which may be involved in the pathogenesis of VZV infection are the Langerhans cells (LC) of the epidermis and dermal DC (DDC) (60). LC are present in an immature state in uninfected skin and in upper respiratory tract epithelium. Upon capture of foreign antigens, LC have the capacity to migrate from the periphery to the lymph nodes, where they seek interaction with T lymphocytes (60). Although the location of cutaneous DC suggests that they are a DC subset likely to be involved in the pathogenesis of VZV infection, other subsets of DC, such as the blood-derived myeloid DC (MDC) and plasmacytoid DC (PDC), are also potentially important in the pathogenesis of VZV infection. Of particular interest are PDC, since these cells are important in innate antiviral immune responses due to their ability to recruit to sites of inflammation and secrete high levels of alpha interferon (IFN-α) (6, 18, 56). PDC also participate in adaptive immune responses through their secretion of cytokines and chemokines that promote activation of effector cells, including NK cells, NKT cells, B lymphocytes, and T lymphocytes, and also through their capacity to present antigen to T lymphocytes (9, 63). Whether PDC and LC can be infected with VZV and their roles during infection have not been previously studied.In this study, we sought to identify and compare the subsets of DC present in human skin lesions following natural VZV infection and to assess DC permissiveness to VZV infection. We utilized immunohistochemical (IHC) and immunofluorescent (IFA) staining to characterize DC subsets within the skin of multiple patients with either varicella or herpes zoster, and identified profound changes in the frequency of LC and PDC as a consequence of cutaneous VZV infection. In addition, some LC and PDC costained with a range of VZV antigens indicative of productive infection. PDC isolated from human blood and LC derived from the MUTZ-3 cells were shown to be permissive to productive VZV infection in vitro. This study defines changes in the type and distribution of DC during natural cutaneous VZV infection and implicates infection of specific DC subsets in VZV pathogenesis.  相似文献   

17.
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