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 62 毫秒
1.
2.
Tegument is a unique structure of herpesvirus, which surrounds the capsid and interacts with the envelope. Morphogenesis of gammaherpesvirus is poorly understood due to lack of efficient lytic replication for Epstein-Barr virus and Kaposi''s sarcoma-associated herpesvirus/human herpesvirus 8, which are etiologically associated with several types of human malignancies. Murine gammaherpesvirus 68 (MHV-68) is genetically related to the human gammaherpesviruses and presents an excellent model for studying de novo lytic replication of gammaherpesviruses. MHV-68 open reading frame 33 (ORF33) is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. However, the specific role of ORF33 in gammaherpesvirus replication has not yet been characterized. We describe here that ORF33 is a true late gene and encodes a tegument protein. By constructing an ORF33-null MHV-68 mutant, we demonstrated that ORF33 is not required for viral DNA replication, early and late gene expression, viral DNA packaging or capsid assembly but is required for virion morphogenesis and egress. Although the ORF33-null virus was deficient in release of infectious virions, partially tegumented capsids produced by the ORF33-null mutant accumulated in the cytoplasm, containing conserved capsid proteins, ORF52 tegument protein, but virtually no ORF45 tegument protein and the 65-kDa glycoprotein B. Finally, we found that the defect of ORF33-null MHV-68 could be rescued by providing ORF33 in trans or in an ORF33-null revertant virus. Taken together, our results indicate that ORF33 is a tegument protein required for viral lytic replication and functions in virion morphogenesis and egress.Gammaherpesviruses are associated with tumorigenesis. Like other herpesviruses, they are characterized as having two distinct stages in their life cycle: lytic replication and latency (15, 16, 18, 21, 54). Latency provides the viruses with advantages to escape host immune surveillance and to establish lifelong persistent infection and contributes to transformation and development of malignancies. However, it is through lytic replication that viruses propagate and transmit among hosts to maintain viral reservoirs. Both viral latency and lytic replication play important roles in tumorigenesis. The gammaherpesvirus subfamily includes Epstein-Barr virus (EBV), Kaposi''s sarcoma-associated herpesvirus (KSHV)/human herpesvirus 8 and murine gammaherpesvirus 68 (MHV-68), among others. EBV is associated with Burkitt''s lymphoma, nasopharyngeal carcinoma, Hodgkin''s disease, and lymphoproliferative diseases in immunodeficient patients (28). KSHV is etiologically linked with Kaposi''s sarcoma, primary effusion lymphoma, and multicentric Castleman''s disease (11-13, 22, 52). Neither in vivo nor in vitro studies of EBV and KSHV are convenient due to their propensity to establish latency in cell culture and their limited host ranges.MHV-68 is genetically related to these two human gammaherpesviruses, especially to KSHV, based on the alignment of their genomic sequences and other biological properties (55). As a natural pathogen of wild rodents, MHV-68 also infects laboratory mice (6, 40, 46) and replicates to a high titer in a variety of fibroblast and epithelial cell lines. These advantages make MHV-68 an excellent model for studying the lytic replication of gammaherpesviruses in vitro and certain aspects of virus-host interactions in vivo. In addition, the MHV-68 genome has been cloned as a bacterial artificial chromosome (BAC) that can propagate in Escherichia coli (1, 2, 36, 51), making it convenient to study the function of each open reading frame (ORF) by genetic methods. Exploring the functions of MHV-68 ORFs will likely shed light on the functions of their homologues in human gammaherpesviruses.Gammaherpesviral particles have a characteristic multilayered architecture. An infectious virion contains a double-stranded DNA genome, an icosahedral capsid shell, a thick, proteinaceous tegument compartment, and a lipid bilayer envelope spiked with glycoproteins (14, 30, 47, 49). As a unique structure of herpesviruses, the tegument plays important roles in multiple aspects of the viral life cycle, including virion assembly and egress (38, 48, 53), translocation of nucleocapsids into the nucleus, transactivation of viral immediate-early genes, and modulation of host cell gene expression, innate immunity, and signal transduction (9, 10, 23, 60). Some components of MHV-68 tegument have been identified by a mass spectrometric study (8), and the functions of some tegument proteins have been revealed, such as ORF45, ORF52, and ORF75c (7, 24, 29).MHV-68 ORF33 is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. Its homologues include human herpes simplex virus type 1 (HSV-1) UL16, human herpes simplex virus type 2 (HSV-2) UL16, human cytomegalovirus (HCMV) UL94, EBV BGLF2, KSHV ORF33, and rhesus monkey rhadinovirus (RRV) ORF33. HSV-1 UL16 has been identified as a tegument protein and may function in viral DNA packaging, virion assembly, budding, and egress (5, 32, 35, 41, 44). HCMV UL94 is a virion associated protein and might function in virion assembly and budding (31, 57). EBV BGLF2, KSHV ORF33, and RRV ORF33 are also virion-associated proteins, but their functions are not clear (26, 43, 59). The mass spectrometric study of MHV-68 did not identify ORF33 as a virion component (8), although ORF33 is found to be essential for viral lytic replication by transposon mutagenesis of the MHV-68 genome cloned as a BAC (51). However, insertion of the 1.2-kbp Mu transposon in that study may influence the expression of ORFs approximate to ORF33. Consequently, the role ORF33 plays in viral replication needs to be confirmed, preferably through site-directed mutagenesis. Whether ORF33 is a tegument protein and the exact viral replication stage in which it functions also need to be investigated.We determined that MHV-68 ORF33 encodes a tegument protein and is expressed with true late kinetics. To explore the function of ORF33 in viral lytic phase, we used site-directed mutagenesis and generated an ORF33-null mutant, taking advantage of the MHV-68 BAC system. We showed that the ORF33-null mutant is capable of viral DNA replication, early and late gene expression, capsid assembly, and DNA packaging, but incapable of virion release. The defect of ORF33-null mutant can be rescued in trans by an ORF33 expression plasmid.  相似文献   

3.
4.
5.
6.
Murine gammaherpesvirus 68 (MHV-68) is closely related to Epstein-Barr virus and Kaposi''s sarcoma-associated herpesvirus (KSHV) and provides a small-animal model to study the pathogenesis of gammaherpesvirus (γHV) infections. According to the colinear organization of the γHV genomes, the M10 locus is situated at a position equivalent to the K12 locus of KSHV, which codes for proteins of the kaposin family. The M10 locus of MHV-68 has been predicted to code for three overlapping open reading frames (M10a, M10b, and M10c [M10a-c]) with unknown function. In addition, the M10 locus contains a lytic origin of replication (oriLyt). To elucidate the function of the M10 locus during lytic and latent infections, we investigated, both in vitro and in vivo, the following four recombinant viruses which were generated using MHV-68 cloned as a bacterial artificial chromosome: (i) a mutant virus with a deletion which affects both the coding region for M10a-c and the oriLyt; (ii) a revertant virus in which both the M10a-c coding region and the oriLyt were reverted to those of the wild type; (iii) a virus with an ectopic insertion of the oriLyt, which restores the function of the oriLyt but not the M10a-c coding region; and (iv) a mutant virus with a deletion in the oriLyt only. While the mutants were slightly attenuated with regard to lytic replication in cell culture, they showed severe growth defects in vivo. Both lytic replication and latency amplification were strongly reduced. In contrast, both the revertant virus and the virus with the ectopic oriLyt insertion grew very similarly to the parental wild-type virus both in vitro and in vivo. Thus, we provide genetic evidence that mutation of the oriLyt, and not of putative protein coding sequences within the M10a-c region, is responsible for the observed phenotype. We conclude that the oriLyt in the M10 locus plays an important role during infection of mice with MHV-68.Diseases caused by gammaherpesviruses continue to be a challenge for human health. The prototypic gamma-1 herpesvirus Epstein-Barr virus (EBV) is associated with lymphomas and nasopharyngeal carcinoma (22). Human herpesvirus 8 (also called Kaposi''s sarcoma-associated herpesvirus [KSHV]), a gamma-2 herpesvirus, is associated with lymphoproliferative disorders and Kaposi''s sarcoma (24). In vivo studies of gammaherpesvirus pathogenesis have been limited to clinical investigation of the infection because of the restricted host range of these viruses. The murine gammaherpesvirus 68 (MHV-68) is also a member of the gammaherpesvirus subfamily and is closely related to KSHV and EBV. Since there exist no good animal models for KSHV and EBV, MHV-68 serves as a small-animal model to investigate gammaherpesvirus pathogenesis (6, 9, 10, 13, 21, 25, 26, 30). MHV-68 is a natural pathogen of wild rodents (7) and is capable of infecting laboratory mice. The nucleotide sequence of MHV-68 is similar to that of EBV and even more closely related to that of KSHV (29). MHV-68 contains genes which are homologous to cellular genes or to genes of other gammaherpesviruses. In addition, it contains virus-specific genes. Many of the latency- and transformation-associated proteins of the gammaherpesviruses, for example, EBNA and LMP of EBV, appear to be encoded by virus-specific genes, yet it has been suggested that pathogenesis-associated genes of gammaherpesviruses may be contained in similarly positioned genome regions (29). The virus-specific genes of MHV-68 were originally designated M1 to M14 (29). The M10 locus has been predicted to code for three overlapping open reading frames (M10a, M10b, and M10c [M10a-c]) (29). While several MHV-68-specific genes have been shown to code for proteins with important functions, the function of M10 is still unknown. A more recent report even considered M10a-c rather unlikely to code for proteins (21). Importantly, the M10 locus also contains a lytic origin of replication (oriLyt) (3, 8). According to the colinear organization of the gammaherpesvirus genomes, the M10 locus is situated at a position equivalent to that of the K12 locus of KSHV. K12 encodes proteins of the kaposin family. Kaposin proteins are involved in cellular transformation and in stabilization of cytokine mRNAs (16-18,20). Of note, the K12 locus also contains an oriLyt (5).Here, we investigated the function of the M10 locus during lytic and latent infections by studying mutant viruses with deletions in the M10 loci, either affecting both the coding region for M10a-c and the oriLyt or the oriLyt only. While the mutants were slightly attenuated with regard to lytic replication in cell culture, they showed severe growth defects in vivo. Both lytic replication and latency amplification were strongly reduced in mice infected with the mutant viruses. In contrast, a revertant virus in which both the M10a-c coding region and the oriLyt were reverted to those of the wild type and a virus with an ectopic insertion of the oriLyt which restores the function of the oriLyt but not the M10a-c coding region grew very similarly to the parental wild-type virus both in vitro and in vivo. Thus, we provide genetic evidence that mutation of the oriLyt, and not of putative protein coding sequences within the M10a-c region, is responsible for the observed phenotype.  相似文献   

7.
8.
9.
10.
11.
12.
13.
14.
We previously showed that agonistic antibodies to CD40 could substitute for CD4 T-cell help and prevent reactivation of murine gammaherpesvirus 68 (MHV-68) in the lungs of major histocompatibility complex (MHC) class II−/− (CII−/−) mice, which are CD4 T cell deficient. Although CD8 T cells were required for this effect, no change in their activity was detected in vitro. A key question was whether anti-CD40 treatment (or CD4 T-cell help) changed the function of CD8 T cells or another cell type in vivo. To address this question, in the present study, we showed that adoptive transfer of CD8 T cells from virus-infected wild-type mice or anti-CD40-treated CII−/− mice caused a significant reduction in lung viral titers, in contrast to those from control CII−/− mice. Anti-CD40 treatment also greatly prolonged survival of infected CII−/− mice. This confirms that costimulatory signals cause a change in CD8 T cells enabling them to maintain effective long-term control of MHV-68. We investigated the nature of this change and found that expression of the inhibitory receptor PD-1 was significantly increased on CD8 T cells in the lungs of MHV-68-infected CII−/−, CD40−/−, or CD80/86−/− mice, compared with that in wild-type or CD28/CTLA4−/− mice, correlating with the level of viral reactivation. Furthermore, blocking PD-1-PD-L1 interactions significantly reduced viral reactivation in CD4 T-cell-deficient mice. In contrast, the absence of another inhibitory receptor, NKG2A, had no effect. These data suggest that CD4 T-cell help programs a change in CD8 T-cell function mediated by altered PD-1 expression, which enables effective long-term control of MHV-68.Murine gammaherpesvirus 68 (MHV-68) is a naturally occurring rodent pathogen which is closely related to Epstein-Barr virus (EBV) and Kaposi''s sarcoma-associated herpesvirus (KSHV) (17, 64). Intranasal administration of MHV-68 to mice results in acute productive infection of lung epithelial cells and a latent infection in various cell types, including B lymphocytes, dendritic cells, epithelial cells, and macrophages (18, 19, 52, 53, 61, 65). The virus induces an inflammatory infiltrate in the lungs, lymph node enlargement, splenomegaly, and mononucleosis comprising increased numbers of activated CD8 T cells in the blood (53, 58). It has also been reported to induce lymphoproliferative disease/lymphoma in immunocompromised mice (30, 55, 60). Thus, the pathogenesis resembles that of EBV in humans, although structurally, the virus is more closely related to KSHV.Infectious MHV-68 is cleared from the lungs by a T-cell-dependent mechanism 10 to 15 days after infection (18, 53, 56). In wild-type mice, the lungs remain clear of replicating virus thereafter. Although CD4 T cells are not essential for primary clearance of replicating virus, they are required for effective long-term control (11). Thus, major histocompatibility complex (MHC) class II−/− mice that lack CD4 T cells or mice rendered CD4 deficient by antibody treatment initially clear infectious virus from the lungs. However, infectious virus reactivates in the lungs 10 to 15 days later and gradually increases in titer (11, 43). The infected CD4-deficient mice eventually die, apparently from long-term lung damage due to continuing lytic viral replication (11). MHC class II−/− mice do not produce antibody to T-dependent antigens (10). Cytotoxic T-lymphocyte (CTL) epitopes have been identified in open reading frame (ORF) 6 (p56, H-2Db-restricted), and ORF 61 (p79, H-2Kb-restricted) gene products, which appear to encode early lytic-phase proteins (32, 49). The epitopes are presented during two distinct phases during MHV-68 infection, which changes the pattern of CTL dominance (32, 51). However, there is no significant difference in the numbers of CD8 T cells specific for each epitope in wild-type mice and CD4 T-cell-deficient mice (4, 50). In addition, CTL activity measured in vitro does not differ substantially in the lungs of wild-type mice or CD4 T-cell-deficient mice (4, 11, 50). Furthermore, postexposure vaccination with the p56 epitope failed to prevent viral reactivation in class II−/− mice, despite dramatically expanding the number of CD8 T cells specific for the peptide (5). In contrast, vaccination of wild-type mice against these epitopes reduced lytic viral titers in the lung dramatically on subsequent challenge with MHV-68. B-cell-deficient mice clear MHV-68 with the kinetics of wild-type mice and do not show viral reactivation in the lungs (13, 61), suggesting that antibody is not essential for control of the virus. Depletion of CD4 T cells during the latent phase of infection in B-cell-deficient mice does not induce viral reactivation, whereas depletion of both CD4 and CD8 T-cell subsets provokes viral reactivation in the lungs (52). Short-term depletion of both CD4 and CD8 T-cell subsets during the latent phase of infection in wild-type mice does not lead to viral reactivation probably due to the presence of neutralizing antibody (11). Taken together, these results suggest that CD4 and CD8 T cells and B cells play overlapping roles in preventing or controlling reactivation of MHV-68 during the latent phase of infection. However, the B-cell- and CD8 T-cell-mediated control mechanisms do not develop in the absence of CD4 T cells.We, and others, have previously shown that the costimulatory molecule CD28 is not required for long-term control of MHV-68 (28, 29). However, interestingly, mice lacking both of the ligands for CD28, CD80 and CD86, show viral reactivation in the lung (21, 35). Our previously published data showed that agonistic antibodies to CD40 could substitute for CD4 T-cell function in the long-term control of MHV-68 (46). CD8 T-cell receptor-positive (TCR+) cells were required for this effect, while antibody production was not restored (45, 46). MHV-68-infected CD40L−/− mice (7) and CD40−/− mice (29) also showed viral reactivation in the lungs. However, no change in CD8 CTL activity was detected in in vitro assays following anti-CD40 treatment (46). A key question was whether anti-CD40 treatment (or CD4 T-cell help) caused a direct change in CD8 T-cell function or whether both CD8 T cells and an independent anti-CD40-sensitive step were required for viral control. To address this question, we used adoptive transfer of CD8 T cells from MHV-68-infected wild-type mice, anti-CD40-treated mice, or control MHC class II−/− mice to MHV-68-infected class II−/− recipients. We also investigated whether anti-CD40 treatment prolonged survival in addition to reducing lung viral titers. The heterodimeric molecule CD94/NKG2A has been implicated in negatively regulating the CD8 T-cell response to polyomavirus (38) and herpes simplex virus (HSV) (54), while the inhibitory receptor PD-1 (programmed death 1) has been implicated in T-cell exhaustion following infection with several other persistent viruses (2, 15, 20, 22, 26, 36, 39-41, 57, 67). In the present study, we investigated the effect of signaling via various costimulatory molecules on the expression of NKG2A and PD-1 and how these molecules influenced viral control.  相似文献   

15.
16.
Intranasal mouse hepatitis virus type 1 (MHV-1) infection of mice induces lung pathology similar to that observed in severe acute respiratory syndrome (SARS) patients. However, the severity of MHV-1-induced pulmonary disease varies among mouse strains, and it has been suggested that differences in the host immune response might account for this variation. It has also been suggested that immunopathology may represent an important clinical feature of SARS. Little is known about the host immune response to MHV-1 and how it might contribute to some of the pathological changes detected in infected mice. In this study we show that an intact type I interferon system and the adaptive immune responses are required for controlling MHV-1 replication and preventing morbidity and mortality in resistant C57BL/6J mice after infection. The NK cell response also helps minimize the severity of illness following MHV-1 infection of C57BL/6J mice. In A/J and C3H/HeJ mice, which are highly susceptible to MHV-1-induced disease, we demonstrate that both CD4 and CD8 T cells contribute to morbidity during primary infection, and memory responses can enhance morbidity and mortality during subsequent reexposure to MHV-1. However, morbidity in A/J and C3H/HeJ mice can be minimized by treating them with immune serum prior to MHV-1 infection. Overall, our findings highlight the role of the host immune response in contributing to the pathogenesis of coronavirus-induced respiratory disease.Severe acute respiratory syndrome (SARS) is caused by a zoonotic coronaviral infection that reached epidemic proportions beginning in late 2002 (37, 52, 55, 76, 84, 86). The etiologic agent, SARS-coronavirus (CoV), is a novel group 2 CoV that emerged in the human population exposed to infected animals that were present in wet markets in various provinces of southern China (16, 22, 35, 45, 57, 61). Although the outbreak was quickly contained by the application of aggressive public health measures, it highlighted the deadly potential of this novel pathogen as more than 8,000 people in more than 25 countries were affected, and almost 800 infected individuals died (37, 76, 84, 86). Although there have not been additional outbreaks of this disease in the general population since 2003, due to the continued presence of related viruses in bats and other animals and to cultural practices prevalent in the local population in southern China, the reemergence of this pathogen in the human population may occur in the future (40).Currently, there are no rigorously tested efficacious prophylactic or therapeutic agents targeting this pathogen. Given the lethal potential of this virus, it is imperative to develop specific antiviral therapies that can be rapidly and universally applied. One of the serious drawbacks in the field is the paucity of appropriate animal models that faithfully reproduce the clinical features of SARS (52, 60). Although a mouse-adapted strain of this virus is available, studies with this strain need to be performed in biosafety level 3 facilities (48, 59). Logistical issues associated with such requirements hamper the rapidity and ease with which one can perform a comprehensive and detailed systemic examination of the dynamics of host-pathogen interactions. Recently, it was reported that intranasal infection of certain strains of mice with a related group 2 respiratory CoV, mouse hepatitis virus type 1 (MHV-1), induced pulmonary disease that was very similar to that observed in human subjects infected with SARS-CoV (11). In addition to the phylogenetic proximity of MHV-1 and SARS-CoV, they also share similarities in genome organization and in mechanisms of replication (63, 68). Hence, it is likely that the pathophysiology observed in MHV-1-infected mice mimics important pathological features associated with SARS-CoV infection in humans. A dysregulated immune response characterized by aberrant cytokine production is postulated to contribute to clinical disease in patients with SARS (8, 26, 55, 58, 72, 75, 82, 83). MHV-1 infection of susceptible strains of mice is also associated with an altered cytokine profile, and published reports suggest that the host immune response to the virus is an important contributor to the pathology observed in susceptible strains of mice (11). Examination of the immune response to a pathogen is critical for the purpose of designing rational and effective vaccination approaches. In addition, it also helps identify potentially deleterious effects of the immune response that can subsequently be manipulated to the advantage of the host, thereby maximizing recovery and minimizing morbidity.In the present study we have carried out a comprehensive analysis of the immune response to MHV-1 following intranasal infection of both resistant and susceptible strains of inbred mice. Our observations in alpha/beta interferon (type I IFN) receptor-knockout (IFN-αβR-KO) mice and NK cell-depleted mice shed light on the protective role of these components of the innate immune response in resistant C57BL/6J (B6) mice. And our examination of the adaptive immune responses to MHV-1 shows that they function as a double-edged sword, mediating protection in resistant strains and contributing to pathology in susceptible strains of mice.  相似文献   

17.
18.
Open reading frame 45 (ORF45) of Kaposi''s sarcoma-associated herpesvirus 8 (KSHV) is an immediate-early phosphorylated tegument protein and has been shown to play important roles at both early and late stages of viral infection. Homologues of ORF45 exist only in gammaherpesviruses, and their homology is limited. These homologues differ in their protein lengths and subcellular localizations. We and others have reported that KSHV ORF45 is localized predominantly in the cytoplasm, whereas its homologue in murine herpesvirus 68 is localized exclusively in the nucleus. We observed that ORF45s of rhesus rhadinovirus and herpesvirus saimiri are found exclusively in the nucleus. As a first step toward understanding the mechanism underlying the distinct intracellular distribution of KSHV ORF45, we identified the signals that control its subcellular localization. We found that KSHV ORF45 accumulated rapidly in the nucleus in the presence of leptomycin B, an inhibitor of CRM1 (exportin 1)-dependent nuclear export, suggesting that it could shuttle between the nucleus and cytoplasm. Mutational analysis revealed that KSHV ORF45 contains a CRM1-dependent, leucine-rich-like nuclear export signal and an adjacent nuclear localization signal. Replacement of the key residues with alanines in these motifs of ORF45 disrupts its shuttling between the cytoplasm and nucleus. The resulting ORF45 mutants have restricted subcellular localizations, being found exclusively either in the cytoplasm or in the nucleus. Recombinant viruses were reconstituted by introduction of these mutations into KSHV bacterial artificial chromosome BAC36. The resultant viruses have distinct phenotypes. A mutant virus in which ORF45 is restricted to the cytoplasm behaves as an ORF45-null mutant and produces 5- to 10-fold fewer progeny viruses than the wild type. In contrast, mutants in which the ORF45 protein is mostly restricted to the nucleus produce numbers of progeny viruses similar to those produced by the wild type. These data suggest that the subcellular localization signals of ORF45 have important functional roles in KSHV lytic replication.Kaposi''s sarcoma-associated herpesvirus (KSHV) is a DNA tumor virus and the causative agent of several human cancers, including Kaposi''s sarcoma (KS), primary effusion lymphoma, and multicentric Castleman''s disease (3, 6). Like all herpesviruses, KSHV has two alternative life cycles, a latent and a lytic cycle. During latency, only a few viral genes are expressed, and no progeny viruses are produced. Under appropriate conditions, latent viral genomes are activated, initiate lytic replication, and express a full panel of viral genes, in a process that leads to viral assembly, release of progeny virus particles, and de novo infection of naïve cells (3, 6). KSHV establishes latent infection in the majority of infected cells in cases of KS, primary effusion lymphoma, and multicentric Castleman''s disease, but lytic replications occur in a small fraction. The recurrent and periodic lytic cycles of KSHV are believed to play critical roles in viral pathogenesis (6, 7).Open reading frame 45 (ORF45) is a KSHV-encoded gene product that plays a critical role in the viral lytic cycle. It is an immediate-early protein and is also present in viral particles as tegument protein (26, 27, 30). Disruption of ORF45 has no significant effect on overall viral lytic gene expression or DNA replication in BAC36-reconstituted 293T cells induced with both tetradecanoyl phorbol acetate (TPA) and sodium butyrate together, but the ORF45-null mutant produces 5- to 10-fold fewer progeny viruses than the wild type and the mutant virus has dramatically reduced infectivity, suggesting that ORF45 plays important roles at both early and late stages of viral infection (29). In addition to its roles as a tegument component, which are possibly involved in viral ingress and egress processes, KSHV ORF45 interacts with cellular proteins and modulates the cellular environment. At least two such functions have been described. First, KSHV ORF45 inhibits activation of interferon regulatory factor 7 (IRF-7) and therefore antagonizes the host innate antiviral response (28). Second, KSHV ORF45 interacts with p90 ribosomal kinase 1 and 2 (RSK1/RSK2) and modulates the extracellular signal-regulated kinase/RSK signaling pathway, which is known to play essential roles in KSHV reactivation and lytic replication (12). All of these data suggest that KSHV ORF45 is a multifunctional protein.ORF45 is unique to the gammaherpesviruses; it has no homologue in the alpha- or betaherpesviruses. ORF45 homologues have been identified as virion protein components in other gammaherpesviruses, such as Epstein-Barr virus (EBV), rhesus rhadinovirus (RRV), and murine herpesvirus 68 (MHV-68), suggesting that certain tegument functions of ORF45 are conserved (2, 11, 18). ORF45 homologues differ in protein length. KSHV ORF45 is the longest, at 407 amino acids (aa); RRV, EBV, MHV-68, and herpesvirus saimiri (HVS) have proteins of 353, 217, 206, and 257 aa, respectively. The limited homologies lie mostly at the amino- and carboxyl-terminal ends. The middle portion of KSHV ORF45 diverges from those of its homologues. The homologues differ in subcellular localization. We and others have reported previously that KSHV ORF45 is found predominantly in the cytoplasm (1, 21, 28, 30), whereas ORF45 of MHV-68 is found exclusively in the nucleus (9). Recently, we found KSHV ORF45 also present in the nuclei of BCBL-1 cells in what resembled viral replication compartments, suggesting that ORF45 could shuttle into the nucleus (12).Nucleocytoplasmic trafficking of proteins across the nuclear membrane occurs through nuclear pore complexes. Small molecules of up to approximately 9 nm in diameter, corresponding to a globular protein of approximately 40 to 60 kDa, can in principle enter or leave the nucleus by diffusion through nuclear pores (15, 17, 24). Large molecules are transported with the aid of a related family of transport factors, importins and exportins, which recognize nuclear localization sequence (NLS)-containing or nuclear export sequence (NES)-containing proteins (15, 17, 23). CRM1 (exportin 1) has been identified as a common export receptor that recognizes human immunodeficiency virus Rev-like leucine-rich NES sequences and is responsible for the export of such NES-containing proteins (4, 5, 19, 22). CRM1-dependent nuclear export is specifically inhibited by a pharmacological compound, leptomycin B (LMB), that interacts with CRM1 and thus blocks such NES-mediated protein export (4).To understand the mechanism underlying the distinct intracellular distribution of KSHV ORF45, we attempted to locate the signals that control its subcellular localization. In the research reported here, we identified a leucine-rich NES and an adjacent basic NLS in KSHV ORF45. We demonstrated that the regulated intracellular trafficking of ORF45, especially its translocation into the nucleus, is important for KSHV lytic replication.  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