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 31 毫秒
1.
Antibodies against the extracellular virion (EV or EEV) form of vaccinia virus are an important component of protective immunity in animal models and likely contribute to the protection of immunized humans against poxviruses. Using fully human monoclonal antibodies (MAbs), we now have shown that the protective attributes of the human anti-B5 antibody response to the smallpox vaccine (vaccinia virus) are heavily dependent on effector functions. By switching Fc domains of a single MAb, we have definitively shown that neutralization in vitro—and protection in vivo in a mouse model—by the human anti-B5 immunoglobulin G MAbs is isotype dependent, thereby demonstrating that efficient protection by these antibodies is not simply dependent on binding an appropriate vaccinia virion antigen with high affinity but in fact requires antibody effector function. The complement components C3 and C1q, but not C5, were required for neutralization. We also have demonstrated that human MAbs against B5 can potently direct complement-dependent cytotoxicity of vaccinia virus-infected cells. Each of these results was then extended to the polyclonal human antibody response to the smallpox vaccine. A model is proposed to explain the mechanism of EV neutralization. Altogether these findings enhance our understanding of the central protective activities of smallpox vaccine-elicited antibodies in immunized humans.The smallpox vaccine, live vaccinia virus (VACV), is frequently considered the gold standard of human vaccines and has been enormously effective in preventing smallpox disease. The smallpox vaccine led to the worldwide eradication of the disease via massive vaccination campaigns in the 1960s and 1970s, one of the greatest successes of modern medicine (30). However, despite the efficacy of the smallpox vaccine, the mechanisms of protection remain unclear. Understanding those mechanisms is key for developing immunologically sound vaccinology principles that can be applied to the design of future vaccines for other infectious diseases (3, 101).Clinical studies of fatal human cases of smallpox disease (variola virus infection) have shown that neutralizing antibody titers were either low or absent in patient serum (24, 68). In contrast, neutralizing antibody titers for the VACV intracellular mature virion (MV or IMV) were correlated with protection of vaccinees against smallpox (68). VACV immune globulin (VIG) (human polyclonal antibodies) is a promising treatment against smallpox (47), since it was able to reduce the number of smallpox cases ∼80% among variola-exposed individuals in four case-controlled clinical studies (43, 47, 52, 53, 69). In animal studies, neutralizing antibodies are crucial for protecting primates and mice against pathogenic poxviruses (3, 7, 17, 21, 27, 35, 61, 66, 85).The specificities and the functions of protective antipoxvirus antibodies have been areas of intensive research, and the mechanics of poxvirus neutralization have been debated for years. There are several interesting features and problems associated with the antibody response to variola virus and related poxviruses, including the large size of the viral particles and the various abundances of many distinct surface proteins (18, 75, 91, 93). Furthermore, poxviruses have two distinct virion forms, intracellular MV and extracellular enveloped virions (EV or EEV), each with a unique biology. Most importantly, MV and EV virions share no surface proteins (18, 93), and therefore, there is no single neutralizing antibody that can neutralize both virion forms. As such, an understanding of virion structure is required to develop knowledge regarding the targets of protective antibodies.Neutralizing antibodies confer protection mainly through the recognition of antigens on the surface of a virus. A number of groups have discovered neutralizing antibody targets of poxviruses in animals and humans (3). The relative roles of antibodies against MV and EV in protective immunity still remain somewhat unclear. There are compelling data that antibodies against MV (21, 35, 39, 66, 85, 90, 91) or EV (7, 16, 17, 36, 66, 91) are sufficient for protection, and a combination of antibodies against both targets is most protective (66). It remains controversial whether antibodies to one virion form are more important than those to the other (3, 61, 66). The most abundant viral particles are MV, which accumulate in infected cells and are released as cells die (75). Neutralization of MV is relatively well characterized (3, 8, 21, 35). EV, while less abundant, are critical for viral spread and virulence in vivo (93, 108). Neutralization of EV has remained more enigmatic (3).B5R (also known as B5 or WR187), one of five known EV-specific proteins, is highly conserved among different strains of VACV and in other orthopoxviruses (28, 49). B5 was identified as a protective antigen by Galmiche et al., and the available evidence indicated that the protection was mediated by anti-B5 antibodies (36). Since then, a series of studies have examined B5 as a potential recombinant vaccine antigen or as a target of therapeutic monoclonal antibodies (MAbs) (1, 2, 7, 17, 40, 46, 66, 91, 110). It is known that humans immunized with the smallpox vaccine make antibodies against B5 (5, 22, 62, 82). It is also known that animals receiving the smallpox vaccine generate antibodies against B5 (7, 20, 27, 70). Furthermore, previous neutralization assays have indicated that antibodies generated against B5 are primarily responsible for neutralization of VACV EV (5, 83). Recently Chen at al. generated chimpanzee-human fusion MAbs against B5 and showed that the MAbs can protect mice from lethal challenge with virulent VACV (17). We recently reported, in connection with a study using murine monoclonal antibodies, that neutralization of EV is highly complement dependent and the ability of anti-B5 MAbs to protect in vivo correlated with their ability to neutralize EV in a complement-dependent manner (7).The focus of the study described here was to elucidate the mechanisms of EV neutralization, focusing on the human antibody response to B5. Our overall goal is to understand underlying immunobiological and virological parameters that determine the emergence of protective antiviral immune responses in humans.  相似文献   

2.
The current vaccine against smallpox is an infectious form of vaccinia virus that has significant side effects. Alternative vaccine approaches using recombinant viral proteins are being developed. A target of subunit vaccine strategies is the poxvirus protein A33, a conserved protein in the Chordopoxvirinae subfamily of Poxviridae that is expressed on the outer viral envelope. Here we have determined the structure of the A33 ectodomain of vaccinia virus. The structure revealed C-type lectin-like domains (CTLDs) that occur as dimers in A33 crystals with five different crystal lattices. Comparison of the A33 dimer models shows that the A33 monomers have a degree of flexibility in position within the dimer. Structural comparisons show that the A33 monomer is a close match to the Link module class of CTLDs but that the A33 dimer is most similar to the natural killer (NK)-cell receptor class of CTLDs. Structural data on Link modules and NK-cell receptor-ligand complexes suggest a surface of A33 that could interact with viral or host ligands. The dimer interface is well conserved in all known A33 sequences, indicating an important role for the A33 dimer. The structure indicates how previously described A33 mutations disrupt protein folding and locates the positions of N-linked glycosylations and the epitope of a protective antibody.Poxviruses are large DNA viruses that infect a wide range of hosts. The smallpox virus devastated human populations until its eradication 3 decades ago. Other poxviruses are emerging, such as monkeypox virus, which also infects humans and causes disease (61). The smallpox vaccine is a model of vaccine efficacy, but how the vaccine induces protection is not well understood. Knowledge of how the vaccine produces protection will also likely be important for efforts to produce vaccines that are effective against other pathogens. Though highly successful in the population overall, the smallpox vaccine uses an infectious strain of vaccinia virus and has a significant level of serious side effects (6). One focus of current research is to develop vaccines using recombinant poxvirus proteins that are as protective as the live virus vaccine but produce fewer complications. A more detailed structural characterization of the protein antigens that are important for conferring protection will improve our knowledge of how the smallpox vaccine works and lead to a better understanding of poxvirus biology.There are two morphologically distinct forms of poxviruses: the mature virion (MV) and the enveloped virion (EV) (53). The MV, also known as the intracellular mature virion (IMV), is found inside the infected cell (62, 65). Enveloped virions are formed from MVs that have been wrapped by modified Golgi or endosomal membranes (53). MVs are thought to be responsible for host-to-host proliferation of the virus, while the EVs mediate virus spread within a host (40, 49). EVs that are attached to the cell surface, also termed cell-associated enveloped virions (CEV), are thought to propagate viral infection to neighboring cells (65). EVs that are released from the cell surface, also termed extracellular enveloped virus (EEV), mediate longer-range dissemination in the host (65). The outer membranes of the MV and EV forms each have a distinct assemblage of proteins. Candidate subunit vaccines have been shown to require proteins from both virus forms to be most effective (17, 27, 28).A33 is a type II integral membrane glycoprotein found on the surface of the EV form of the virus and is also expressed on the host cell membrane (62). A33 is a disulfide-bonded homodimer with both N- and O-linked glycosylation (57, 62). Deletion of the A33R gene in vaccinia virus results in a small-plaque phenotype, defects in actin tail formation, and inefficient cell-to-cell spread in cell culture (63). Evidence implicates A33 in the spreading of virus from cell to cell by a mechanism that is antibody resistant (40). Antibodies against A33 in cell culture prevent the formation of comet-shaped viral plaques, which are assayed in an overlay of plaques with liquid and are indicative of cell-to-cell spreading of the EV (2, 17). A33 has been shown to interact through its cytoplasmic and transmembrane regions with the EV proteins A36 (18, 64, 72, 76) and B5 (58, 64). Vaccination with A33 is protective in a number of animal models of poxvirus infection as a component of protein subunit vaccination (16, 17, 19, 77), DNA vaccination (19, 27-29), or a combination of the two methods (24). Despite the inclusion of A33 in vaccination studies, the functions of A33 in the virus are unclear.Members of the C-type lectin-like domain (CTLD) superfamily of proteins are found in organisms ranging from bacteria to humans (13, 78). The first crystal structures of carbohydrate-binding domains with this fold gave rise to the name “lectin” for the family, but many family members do not bind carbohydrates. The classification of a domain as being C type lectin-like is currently based on similarities in protein sequence and fold (74). CTLDs have been shown to bind noncarbohydrate small molecules, lipids, proteins, and other structures, such as ice (13, 78). One group of type II transmembrane proteins that contain CTLDs is comprised of the natural killer (NK)-cell receptors of the innate immune system (78). The NK-cell receptors are composed of dimers of CTLDs, which are the only dimers out of the several hundred C-type lectin-like structures that are known. Protein binding by NK-cell receptors occurs on a surface formed by the “long loop” and nearby residues that are on the opposite side of the dimer from the N and C termini (60). A second group of proteins contains a monomeric CTLD, termed a “Link module” CTLD, that binds glycosaminoglycans but lacks the “long loop” region that is present in almost all other CTLDs (78).To gain a better understanding of the structure and possible functions of A33 and to further its development in vaccines, we determined the X-ray crystal structure of the A33 ectodomain from vaccinia virus. Based on the structure and sequence of A33, the carbohydrate-binding site of the canonical CTLD is not present. The structure revealed A33 to have dimers of CTLDs. Comparison of A33 with other CTLDs, including dimers from NK-cell receptors and monomers from Link modules, indicates that A33 contains an unusual CTLD that likely binds ligands of host or virus origins.  相似文献   

3.
Oncolytic vaccinia viruses have shown compelling results in preclinical cancer models and promising preliminary safety and antitumor activity in early clinical trials. However, to facilitate systemic application it would be useful to improve tumor targeting and antitumor efficacy further. Here we report the generation of vvdd-VEGFR-1-Ig, a targeted and armed oncolytic vaccinia virus. Tumor targeting was achieved by deletion of genes for thymidine kinase and vaccinia virus growth factor, which are necessary for replication in normal but not in cancer cells. Given the high vascularization typical of kidney cancers, we armed the virus with the soluble vascular endothelial growth factor (VEGF) receptor 1 protein for an antiangiogenic effect. Systemic application of high doses of vvdd-VEGFR-1-Ig resulted in cytokine induction in an immunocompromised mouse model. Upon histopathological analysis, splenic extramedullary hematopoiesis was seen in all virus-injected mice and was more pronounced in the vvdd-VEGFR-1-Ig group. Analysis of the innate immune response after intravenous virus injection revealed high transient and dose-dependent cytokine elevations. When medium and low doses were used for intratumoral or intravenous injection, vvdd-VEGFR-1-Ig exhibited a stronger antitumor effect than the unarmed control. Furthermore, expression of VEGFR-1-Ig was confirmed, and a concurrent antiangiogenic effect was seen. In an immunocompetent model, systemic vvdd-VEGFR-1-Ig exhibited superior antitumor efficacy compared to the unarmed control virus. In conclusion, the targeted and armed vvdd-VEGFR-1-Ig has promising anticancer activity in renal cell cancer models. Extramedullary hematopoiesis may be a sensitive indicator of vaccinia virus effects in mice.In 2002 renal cell cancer accounted for more than 200,000 cases and 100,000 deaths worldwide (33). Unfortunately, chemotherapy, radiotherapy, and immunotherapy yield low response rates (9, 17) in this cancer type. Thus, prognosis for patients is poor, especially when the disease is metastatic, as median survival is only 8 months (19). Although recently approved drugs, such as sorafenib, sunitinib, temsirolimus, and bevacizumab, have provided additional tools for treatment of renal cell cancer (7), they are usually not curative, and thus new treatment approaches are needed.Oncolytic vaccinia viruses are promising agents for cancer treatment and have shown compelling results in preclinical tumor models (40, 42, 45). Moreover, good safety and preliminary evidence of antitumor efficacy were seen in phase 1 clinical trials (22, 26, 32). Vaccinia virus has a strong oncolytic effect due to its fast replication cycle (45) and a high innate tropism to cancer tissue (34). Tumor targeting can be further improved by deleting vaccinia virus genes that are necessary for replication in normal cells but not in cancer cells. For example, deletions of either thymidine kinase (TK) or vaccinia virus growth factor (VGF) or both have been shown to reduce pathogenicity compared to wild-type virus (3, 5, 27). To enhance antitumor potency, oncolytic vaccinia viruses can be armed with therapeutic transgenes, such as immunostimulatory factors (26) or suicide genes (14, 16, 35). With regard to kidney cancer, an arming approach with antiangiogenenic molecules seems logical, considering the high vascularization characteristic of renal tumors (20).Vascular endothelial growth factor (VEGF) is a major player in tumor angiogenesis and is highly expressed in renal cell cancers (29). VEGF binds to the fms-like-tyrosine kinase receptor (flt-1 or VEGFR-1) and kinase domain region receptor (KDR or VEGFR-2) with high affinity (13). The soluble vascular endothelial growth factor receptor 1-Ig fusion protein (VEGFR-1-Ig) used in this study is derived from the membrane-bound VEGFR-1 and binds human and murine VEGF without inducing vascular endothelial cell mitogenesis (31). Blocking VEGF with this or closely related molecules has been shown to inhibit tumor growth in several cancer models (18, 21, 25, 39).Although tumor cell selective replication can be enhanced by deletion of TK and/or VGF to reduce pathogenicity (3, 5, 27), high doses of attenuated vaccinia virus may increase serum cytokine concentrations which parallel the onset of toxic events, as seen with other viral vectors (2, 38). The potential “early” toxicity associated with oncolytic vaccinia viruses has not been completely elucidated heretofore (36, 46).Given the high vascularization of renal cell cancers and the pressing need to generate new antitumor agents with increased safety and efficacy, we hypothesized that an oncolytic vaccinia virus targeted by TK and VGF deletions and armed with VEGFR-1-Ig would exhibit enhanced antitumor efficacy due to its antiangiogenic properties in renal cell cancer models compared to a nonarmed control virus, allowing reduction of the treatment dose.  相似文献   

4.
The vaccinia virus (VACV) complement control protein (VCP) is an immunomodulatory protein that is both secreted from and expressed on the surface of infected cells. Surface expression of VCP occurs though an interaction with the viral transmembrane protein A56 and is dependent on a free N-terminal cysteine of VCP. Although A56 and VCP have been shown to interact in infected cells, the mechanism remains unclear. To investigate if A56 is sufficient for surface expression, we transiently expressed VCP and A56 in eukaryotic cell lines and found that they interact on the cell surface in the absence of other viral proteins. Since A56 contains three extracellular cysteines, we hypothesized that one of the cysteines may be unpaired and could therefore form a disulfide bridge with VCP. To test this, we generated a series of A56 mutants in which each cysteine was mutated to a serine, and we found that mutation of cysteine 162 abrogated VCP cell surface expression. We also tested the ability of other poxvirus complement control proteins to bind to VACV A56. While the smallpox homolog of VCP is able to bind VACV A56, the ectromelia virus (ECTV) VCP homolog is only able to bind the ECTV homolog of A56, indicating that these proteins may have coevolved. Surface expression of poxvirus complement control proteins may have important implications in viral pathogenesis, as a virus that does not express cell surface VCP is attenuated in vivo. This suggests that surface expression of VCP may contribute to poxvirus pathogenesis.Poxviruses, including vaccinia virus (VACV), encode large numbers of immunomodulatory proteins that help them establish an infection and combat the host''s immune response (10, 32). One of these is the vaccinia virus complement control protein (VCP), which is both secreted from and expressed on the surface of infected cells (9, 14, 16, 17). VCP acts against the complement system, a series of soluble proteins that is an important early component of the innate immune system and also shapes adaptive immune responses (15, 42, 43). In response to viral infection, complement can opsonize or inactivate virions and can lyse enveloped virus or infected cells (1, 3, 7, 12). Because of these pressures, a number of viruses, including herpes simplex virus, flaviviruses, and poxviruses, encode novel or host-derived regulators of complement, while others, including HIV and poxviruses, incorporate host complement regulatory proteins into virus particles (7, 11, 31, 39). Many orthopoxviruses encode a complement regulator (8, 20, 23, 29), and the most studied of these is VCP. Structurally, VCP is made up of four short consensus repeats (SCR) that are the basic units of mammalian complement regulators (17, 25), and VCP has been shown to interfere with the complement cascade at multiple steps (2, 16, 20-22, 25, 28-30, 33). Additionally, a VCP knockout virus generates smaller lesions in animal models (14, 16). While some host complement control proteins (CCPs) are secreted, many contain transmembrane domains (or a glycophosphatidylinositol anchor) and are thus expressed on the cell surface (42, 43). Thus, when we found that VCP is also expressed on the infected cell surface and protects infected cells from complement-mediated lysis in vitro (9), we believed this to be an important interaction that required further investigation. We previously found that the N-terminal cysteine on VCP was needed for surface expression and that the VACV transmembrane protein A56 was also required (9). The vaccinia virus A56 protein is a type 1 transmembrane glycoprotein that is found on the surface of infected cells and on extracellular virus particles (4, 18, 26, 27, 36). It interacts with another viral protein, K2 (19, 37, 45), which lacks a transmembrane domain and binds to A56 noncovalently (36). The A56/K2 complex prevents syncytium formation between infected cells and superinfection by interacting with the vaccinia virus entry/fusion complex on virions (24, 38, 40, 41). Here we provide evidence that the N-terminal cysteine on VCP forms an intermolecular disulfide bond with cysteine 162 on the ectodomain of A56. We also demonstrate that similar interactions can occur with other poxvirus CCPs, as the smallpox virus and ectromelia virus homologs of VCP also exhibit A56-dependent surface expression.  相似文献   

5.
Poxviruses produce complement regulatory proteins to subvert the host''s immune response. Similar to the human pathogen variola virus, ectromelia virus has a limited host range and provides a mouse model where the virus and the host''s immune response have coevolved. We previously demonstrated that multiple components (C3, C4, and factor B) of the classical and alternative pathways are required to survive ectromelia virus infection. Complement''s role in the innate and adaptive immune responses likely drove the evolution of a virus-encoded virulence factor that regulates complement activation. In this study, we characterized the ectromelia virus inhibitor of complement enzymes (EMICE). Recombinant EMICE regulated complement activation on the surface of CHO cells, and it protected complement-sensitive intracellular mature virions (IMV) from neutralization in vitro. It accomplished this by serving as a cofactor for the inactivation of C3b and C4b and by dissociating the catalytic domain of the classical pathway C3 convertase. Infected murine cells initiated synthesis of EMICE within 4 to 6 h postinoculation. The levels were sufficient in the supernatant to protect the IMV, upon release, from complement-mediated neutralization. EMICE on the surface of infected murine cells also reduced complement activation by the alternative pathway. In contrast, classical pathway activation by high-titer antibody overwhelmed EMICE''s regulatory capacity. These results suggest that EMICE''s role is early during infection when it counteracts the innate immune response. In summary, ectromelia virus produced EMICE within a few hours of an infection, and EMICE in turn decreased complement activation on IMV and infected cells.Poxviruses encode in their large double-stranded DNA genomes many factors that modify the immune system (30, 56). The analysis of these molecules has revealed a delicate balance between viral pathogenesis and the host''s immune response (2, 21, 31, 61). Variola, vaccinia, monkeypox, cowpox, and ectromelia (ECTV) viruses each produce an orthologous complement regulatory protein (poxviral inhibitor of complement enzymes [PICE]) that has structural and functional homology to host proteins (14, 29, 34, 38, 41, 45, 54). The loss of the regulatory protein resulted in smaller local lesions with vaccinia virus lacking the vaccinia virus complement control protein (VCP) (29) and in a greater local inflammatory response in the case of cowpox lacking the inflammation-modulatory protein (IMP; the cowpox virus PICE) (35, 45, 46). Additionally, the complete loss of the monkeypox virus inhibitor of complement enzymes (MOPICE) may account for part of the reduced mortality observed in the West African compared to Congo basin strains of monkeypox virus (12).The complement system consists of proteins on the cell surface and in blood that recognize and destroy invading pathogens and infected host cells (36, 52). Viruses protect themselves from the antiviral effects of complement activation in a variety of ways, including hijacking the host''s complement regulatory proteins or producing their own inhibitors (7, 8, 15, 20, 23). Another effective strategy is to incorporate the host''s complement regulators in the outermost viral membrane, which then protects the virus from complement attack (62). The extracellular enveloped virus (EEV) produced by poxviruses acquires a unique outer membrane derived from the Golgi complex or early endosomes that contain the protective host complement regulators (58, 62). Poxviruses have multiple infectious forms, and the most abundant, intracellular mature virions (IMV), are released when infected cells lyse (58). The IMV lacks the outermost membrane found on EEV and is sensitive to complement-mediated neutralization. The multiple strategies viruses have evolved to evade the complement system underscore its importance to innate and adaptive immunity (15, 36).The most well-characterized PICE is VCP (24-29, 34, 49, 50, 53, 55, 59, 60). Originally described as a secreted complement inhibitor (34), VCP also attaches to the surface of infected cells through an interaction with the viral membrane protein A56 that requires an unpaired N-terminal cysteine (26). This extra cysteine also adds to the potency of the inhibitor by forming function-enhancing dimers (41). VCP and the smallpox virus inhibitor of complement enzymes (SPICE) bind heparin in vitro, and this may facilitate cell surface interactions (24, 38, 50, 59). The coevolution of variola virus with its only natural host, humans, likely explains the enhanced activity against human complement observed with SPICE compared to the other PICEs (54, 64).Our recent work with ECTV, the causative agent of mousepox infection, demonstrated that the classical and alternative pathways of the complement system are required for host survival (48). The mouse-specific pathogen ECTV causes severe disease in most strains and has coevolved with its natural host, analogous to variola virus in humans (9). This close host-virus relationship is particularly important for evaluating the role of the complement system, given the species specificity of many complement proteins, receptors, and regulators (10, 47, 62). Additionally, the availability of complement-deficient mice permits dissection of the complement activation pathways involved. Naïve C57BL/6 mouse serum neutralizes the IMV of ECTV in vitro, predominately through opsonization (48). Maximal neutralization requires natural antibody, classical-pathway activation, and amplification by the alternative pathway. C3 deficiency in the normally resistant C57BL/6 strain results in acute mortality, similar to immunodeficiencies in important elements of the antiviral immune response, including CD8+ T cells (19, 32), natural killer cells (18, 51), and gamma interferon (33). During ECTV infection, the complement system acts in the first few hours and days to delay the spread of infection, resulting in lower levels of viremia and viral burden in tissues (48).This study characterized the PICE produced by ECTV, ectromelia virus inhibitor of complement enzymes (EMICE), and assessed its complement regulatory activity. Recombinant EMICE (rEMICE) decreased activation of both human and mouse complement. Murine cells produced EMICE at 4 to 6 h postinfection prior to the release of the majority of the complement-sensitive IMV from infected cells. rEMICE protected ECTV IMV from complement-mediated neutralization. Further, EMICE produced during natural infection inhibited complement deposition on infected cells by the alternative pathway. ECTV likely produces this abundance of EMICE to protect both the IMV and infected cells.  相似文献   

6.
Little is known about the in vivo kinetics of T-cell responses in smallpox/monkeypox. We showed that macaque Vγ2Vδ2 T cells underwent 3-week-long expansion after smallpox vaccine immunization and displayed simple reexpansion in association with sterile anti-monkeypox virus (anti-MPV) immunity after MPV challenge. Virus-activated Vγ2Vδ2 T cells exhibited gamma interferon-producing effector function after phosphoantigen stimulation. Surprisingly, like αβ T cells, suboptimally primed Vγ2Vδ2 T cells in vaccinia virus/cidofovir-covaccinated macaques mounted major recall-like expansion after MPV challenge. Finally, Vγ2Vδ2 T cells localized in inflamed lung tissues for potential regulation. Our studies provide the first in vivo evidence that viruses, despite their inability to produce exogenous phosphoantigen, can induce expansion, reexpansion, and recall-like expansion of Vγ2Vδ2 T cells and stimulate their antimicrobial cytokine response.Human γδ T cells appear to contribute to both innate and adaptive immune responses (4, 6, 10, 19). Vγ2Vδ2 T cells exist only in primates, and in humans, they constitute 60 to 95% of total blood γδ T cells. The capacity of Vγ2Vδ2 T cells to undergo major clonal expansion in primary infection and to mount rapid recall expansion upon reinfection has been proposed as an adaptive immune response (6), which is consistent with memory phenotypes of Vγ2Vδ2 T cells (7), long-term expansion of memory-like Vδ2 T cells, and in vitro recall expansion of blood γδ T cells in vaccinated or infected humans (1, 15a, 16, 17, 25). It is important to note that the microbial antigen recognized by Vγ2Vδ2 T cells is temporally limited to (E)-4-hydroxy-3-methyl-but-2-enyl pyrophosphate (HMBPP), commonly referred to as phosphoantigen, produced in the newly discovered 2-C-methyl-d-erythritol-4-phosphate pathway of isoprenoid biosynthesis in bacteria, but not viruses (8). Our recent study has demonstrated that HMBPP is presented by a putative molecule on antigen-presenting cell membranes and recognized by Vγ2Vδ2 T-cell receptor (TCR) (24). Since viruses do not produce exogenous HMBPP recognized by Vγ2Vδ2 T cells, in vivo Vγ2Vδ2 T-cell expansion usually occurs only in HMBPP-producing bacterial or protozoal infections.Monkeypox virus (MPV) (an orthopoxvirus) has biological features similar to those of smallpox virus, and MPV infection is clinically similar to smallpox in humans (9, 12, 22). Immune responses of Vγ2Vδ2 T cells during lethal MPV infection have not been studied, although some laboratories have undertaken in vitro studies of γδ T-cell immune responses to vaccinia virus (1, 2, 15). We presume that initial vaccinia virus immunization and subsequent MPV challenge of macaques would provide an ideal in vivo setting in which to determine whether Vγ2Vδ2 T cells can mount innate-like or recall-like responses to orthopoxvirus infections. We made a novel observation indicating expansion, reexpansion, and recall-like expansion of Vγ2Vδ2 T cells with effector function in response to smallpox vaccination and MPV infection in macaques.  相似文献   

7.
Modified vaccinia virus Ankara (MVA) has a highly restricted host range in cell culture and is apathogenic in vivo. MVA was derived from the parental chorioallantois vaccinia virus Ankara (CVA) by more than 570 passages in chicken embryo fibroblast (CEF) cells. During CEF cell passaging, six major deletions comprising 24,668 nucleotides occurred in the CVA genome. We have cloned both the MVA and the parental CVA genome as bacterial artificial chromosomes (BACs) and have sequentially introduced the six major MVA deletions into the cloned CVA genome. Reconstituted mutant CVA viruses containing up to six major MVA deletions showed no detectable replication restriction in 12 of 14 mammalian cell lines tested; the exceptions were rabbit cell lines RK13 and SIRC. In mice, CVA mutants with up to three deletions showed slightly enhanced virulence, suggesting that gene deletion in replicating vaccinia virus (VACV) can result in gain of fitness in vivo. CVA mutants containing five or all six deletions were still pathogenic, with a moderate degree of attenuation. Deletion V was mainly responsible for the attenuated phenotype of these mutants. In conclusion, loss or truncation of all 31 open reading frames in the six major deletions is not sufficient to reproduce the specific MVA phenotype of strong attenuation and highly restricted host range. Mutations in viral genes outside or in association with the six major deletions appear to contribute significantly to this phenotype. Host range restriction and avirulence of MVA are most likely a cooperative effect of gene deletions and mutations involving the major deletions.Modified vaccinia virus Ankara (MVA) was derived from the parental strain chorioallantois vaccinia virus Ankara (CVA) by more than 570 passages in chicken embryo fibroblast (CEF) cells and became severely host cell restricted to avian cells (6, 9, 26). MVA is apathogenic in mammalian hosts, while maintaining excellent immunogenicity (5, 18, 41, 46). Due to the versatility of MVA as a gene expression vector and its immunogenicity, MVA offers an attractive basis for recombinant vector vaccines (17, 48, 50). In addition, the recent appreciation of the possibility of accidental or deliberate release of the smallpox virus renewed interest in an MVA-based smallpox vaccine. MVA was used in the 1970s as a priming vaccine prior to the administration of conventional smallpox vaccine in a two-step program. No significant adverse events were reported after the administration of MVA to more than 120,000 primary vaccinees in Germany (27, 45). Recent clinical studies using a third-generation vaccine, MVA-BN, as a stand-alone smallpox vaccine confirmed its excellent safety profile (56, 57) and underscored its potential as a safe vaccine vector against human infections with various pathogens as well as against cancer.One important reason for the versatility of MVA as a vaccine vector is its particular host range phenotype, which allows the viral gene expression program to proceed until late times in infection, resulting in efficient expression of viral as well as recombinant proteins. The block in viral replication occurs very late during assembly of mature and infectious viral particles (42, 48). With the notable exception of the Syrian hamster cell line BHK-21 and the recently described rat cell line IEC-6, MVA has a very limited ability to productively replicate in mammalian cells (9, 16, 35). The genetic basis of the particular host range restriction of MVA is still not well defined. Comparisons with NYVAC, another replication-deficient vaccinia virus vector (54), showed that significant differences in gene expression programs, apoptosis induction, and immunogenicity exist between NYVAC and MVA (19, 31, 32), although they have very similar host ranges in vitro. Thus, comprehensive knowledge of the genetic factors determining the host ranges of such vectors is necessary to provide a deeper understanding of the basis for the safety and immunogenicity to eventually allow their further optimization.In the course of passaging CVA on CEF cells, the virus acquired six large genomic deletions totaling more than 24 kbp of genomic DNA and deleting or truncating 31 open reading frames (ORFs). In addition to the six major deletions, a multitude of shorter deletions and insertions as well as point mutations have occurred in the MVA genome, resulting in gene fragmentation, truncation, short internal deletions, and amino acid exchanges (29). Some or all of these mutations might also contribute to the attenuated phenotype of MVA. MVA no longer encodes many of the known poxviral immune evasion and virulence factors (2). Of the five classical host range genes present in vaccinia virus (VACV), only C12L/SPI-1 and K1L are deleted or truncated in MVA, whereas C7L, K3L, and E3L are preserved. Deletion of C12L/SPI-1 and K1L contributed to the limited MVA host range, but their reconstitution only partially reversed the MVA host range restriction in selected cell lines (49, 59). Marker rescue experiments using large fragments of the CVA genome suggested that at least two further host range genes apart from C12L/SPI-1, C7L, and K1L might reside in the left part of the VACV genome (59).It is presently unknown how the multiple genetic alterations of MVA determine its limited ability to replicate in most mammalian cells and its lack of pathogenicity in vivo. Since restricted host range and lack of pathogenicity of MVA have commonly been associated with the large deletions in the MVA genome, we aimed at sequentially introducing the six major MVA-like deletions in CVA. To facilitate and accelerate mutagenesis, we have generated bacterial artificial chromosome (BAC) clones of both the MVA and CVA genomes. The use of a counterselectable marker allowed multiple consecutive rounds of mutagenesis of the CVA-BAC (39, 58, 62). The resulting CVA mutants showed that even the introduction of all six major MVA deletions did not create an MVA-like host range phenotype and caused only slight attenuation of the parental CVA virus in a murine intranasal infection model. This result indicates that major host range determinants of MVA are located outside the six large deletions. The host range and virulence phenotype of MVA most probably result from a combined effect of mutation of these unknown factors in conjunction with the six major deletions.  相似文献   

8.
9.
Cidofovir (CDV) is one of the most effective antiorthopoxvirus drugs, and it is widely accepted that viral DNA replication is the main target of its activity. In the present study, we report a detailed analysis of CDV effects on the replicative cycles of distinct vaccinia virus (VACV) strains: Cantagalo virus, VACV-IOC, and VACV-WR. We show that despite the approximately 90% inhibition of production of virus progeny, virus DNA accumulation was reduced only 30%, and late gene expression and genome resolution were unaltered. The level of proteolytic cleavage of the major core proteins was diminished in CDV-treated cells. Electron microscopic analysis of virus-infected cells in the presence of CDV revealed reductions as great as 3.5-fold in the number of mature forms of virus particles, along with a 3.2-fold increase in the number of spherical immature particles. A detailed analysis of purified virions recovered from CDV-treated cells demonstrated the accumulation of unprocessed p4a and p4b and nearly 67% inhibition of DNA encapsidation. However, these effects of CDV on virus morphogenesis resulted from a primary effect on virus DNA synthesis, which led to later defects in genome encapsidation and virus assembly. Analysis of virus DNA by atomic force microscopy revealed that viral cytoplasmic DNA synthesized in the presence of CDV had an altered structure, forming aggregates with increased strand overlapping not observed in the absence of the drug. These aberrant DNA aggregations were not encapsidated into virus particles.Vaccinia virus (VACV) is the prototypical member of the Poxviridae, a family of large DNA-containing viruses. During infection, a cascade of temporally regulated viral gene expression occurs exclusively in the cell cytoplasm, where viral DNA replication takes place. DNA replication is essential for the onset of the intermediate and late steps of viral gene expression (37). VACV morphogenesis is a complex process that starts within the virus factories, or virosomes, in parallel with the late stage of gene expression. Crescent-shaped virus membranes evolve into immature spherical particles (IV) that subsequently progress to form brick-shaped mature virions (MV) (reviewed in reference 8). Virus assembly and maturation are complex processes requiring telomere resolution of newly replicated DNA (20, 36, 40), genome encapsidation (6, 22, 50), and the proteolytic processing of major structural proteins (38, 51). For the past decade, several reports have analyzed in detail these numerous steps of VACV morphogenesis, unraveling the role of distinct virus late proteins in the progression of viral particle formation (reviewed in reference 8).Cantagalo virus (CTGV) is a strain of VACV isolated from pustular lesions on cows in Brazil (10). Similar outbreaks of vaccinia-like viruses have been reported frequently over the past 8 years (14, 39, 48). Interestingly, the majority of these vaccinia viruses circulating in the wild in Brazil bear a striking similarity to the Brazilian vaccine strain used for systematic vaccination during the eradication campaign, which was produced in Rio de Janeiro (19) and called strain IOC (10). This similarity raises the interesting possibility that the circulating vaccinia viruses represent feral derivatives of IOC or of a closely related ancestor. Little is known about the sensitivity of these novel vaccinia viruses to antiviral compounds. In the absence of an active smallpox vaccination campaign, the spread of these vaccinia viruses in the wild, the prevalence of cowpox infections in Europe and elsewhere (39), and the occurrence of complications from smallpox vaccination (52) make the need for effective antipoxvirus treatment a worldwide concern.Cidofovir (CDV), an acyclic pyrimidine phosphonate analogue, has shown a potent antiviral effect on several poxvirus infections (4, 15, 44, 46). Recently, we have reported the efficacy of CDV in inhibiting the replication of the Brazilian VACV strains CTGV and IOC (26). The mechanism of action of CDV on reactions catalyzed by the VACV DNA polymerase has been studied in vitro. CDV is not a chain-terminating analogue but drastically slows chain extension and inhibits the 3′-5′ exonuclease proofreading activity of the enzyme (34). In addition, templates containing CDV cause inhibition of DNA elongation (33). Differences between the effects of CDV on human cytomegalovirus (55) and VACV enzymes have been observed, but overall it has been widely accepted that CDV acts by inhibiting the process of virus DNA replication. Moreover, most CDV-resistant VACV strains contain mutations in the catalytic domain or in the 3′-5′ exonuclease domain of the DNA polymerase (2, 5, 28, 45).Despite the consensus regarding mechanisms of action, the effects of CDV on the stages of the VACV replicative cycle have never been analyzed. We report here that although CDV led to approximately 90% inhibition of VACV progeny production, we observed only 30% inhibition of DNA replication and normal levels of postreplicative virus gene expression. However, the encapsidation of DNA into virus particles and the proteolytic processing of the major core proteins were inhibited in CDV-treated cells, leading to an impairment of virus morphogenesis. These effects on virus assembly are an indirect result of a primary effect of CDV on VACV DNA synthesis. Atomic force microscopy (AFM) analysis revealed that virus DNA isolated from the cytoplasm of CDV-treated cells formed aggregates of highly entangled and intertwined DNA molecules that were not observed in cytoplasmic viral DNA isolated from untreated cells. In addition, these DNA aggregates were not detected in encapsidated virus genomes isolated from particles purified from untreated or CDV-treated cells. Our data suggest that incorporation of CDV into VACV DNA during the replication process may lead to aberrant DNA structures, which are less able to be packaged into virus particles.  相似文献   

10.
11.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

12.
13.
14.
Human immunodeficiency virus type 2 (HIV-2) infection affects about 1 to 2 million individuals, the majority living in West Africa, Europe, and India. As for HIV-1, new strategies for the prevention of HIV-2 infection are needed. Our aim was to produce new vaccine immunogens that elicit the production of broadly reactive HIV-2 neutralizing antibodies (NAbs). Native and truncated envelope proteins from the reference HIV-2ALI isolate were expressed in vaccinia virus or in bacteria. This source isolate was used due to its unique phenotype combining CD4 independence and CCR5 usage. NAbs were not elicited in BALB/c mice by single immunization with a truncated and fully glycosylated envelope gp125 (gp125t) or a recombinant polypeptide comprising the C2, V3, and C3 envelope regions (rpC2-C3). A strong and broad NAb response was, however, elicited in mice primed with gp125t expressed in vaccinia virus and boosted with rpC2-C3. Serum from these animals potently neutralized (median 50% neutralizing titer, 3,200) six of six highly divergent primary HIV-2 isolates. Coreceptor usage and the V3 sequence of NAb-sensitive isolates were similar to that of the vaccinating immunogen (HIV-2ALI). In contrast, NAbs were not reactive on three X4 isolates that displayed major changes in V3 loop sequence and structure. Collectively, our findings demonstrate that broadly reactive HIV-2 NAbs can be elicited by using a vaccinia virus vector-prime/rpC2-C3-boost immunization strategy and suggest a potential relationship between escape to neutralization and cell tropism.Human immunodeficiency virus type 2 (HIV-2) infection affects 1 to 2 million individuals, most of whom live in India, West Africa, and Europe (17). HIV-2 has diversified into eight genetic groups named A to H, of which group A is by far the most prevalent worldwide. Nucleotide sequences of Env can differ up to 21% within a particular group and by over 35% between groups.The mortality rate in HIV-2-infected patients is at least twice that of uninfected individuals (26). Nonetheless, the majority of HIV-2-infected individuals survive as elite controllers (17). In the absence of antiretroviral therapy, the numbers of infected cells (39) and viral loads (36) are much lower among HIV-2-infected individuals than among those who are HIV-1 infected. This may be related to a more effective immune response produced against HIV-2. In fact, most HIV-2-infected individuals have proliferative T-cell responses and strong cytotoxic responses to Env and Gag proteins (17, 31). Moreover, autologous and heterologous neutralizing antibodies (NAbs) are raised in most HIV-2-infected individuals (8, 32, 48, 52), and the virus seems unable to escape from these antibodies (52). As for HIV-1, the antibody specificities that mediate HIV-2 neutralization and control are still elusive. The V3 region in the envelope gp125 has been identified as a neutralizing target by some but not by all investigators (3, 6, 7, 11, 40, 47, 54). Other weakly neutralizing epitopes were identified in the V1, V2, V4, and C5 regions in gp125 and in the COOH-terminal region of the gp41 ectodomain (6, 7, 41). A better understanding of the neutralizing determinants in the HIV-2 Env will provide crucial information regarding the most relevant targets for vaccine design.The development of immunogens that elicit the production of broadly reactive NAbs is considered the number one priority for the HIV-1 vaccine field (4, 42). Most current HIV-1 vaccine candidates intended to elicit such broadly reactive NAbs are based on purified envelope constructs that mimic the structure of the most conserved neutralizing epitopes in the native trimeric Env complex and/or on the expression of wild-type or modified envelope glycoproteins by different types of expression vectors (4, 5, 29, 49, 58). With respect to HIV-2, purified gp125 glycoprotein or synthetic peptides representing selected V3 regions from HIV-2 strain SBL6669 induced autologous and heterologous NAbs in mice or guinea pigs (6, 7, 22). However, immunization of cynomolgus monkeys with a subunit vaccine consisting of gp130 (HIV-2BEN) micelles offered little protection against autologous or heterologous challenge (34). Immunization of rhesus (19, 44, 45) and cynomolgus (1) monkeys with canarypox or attenuated vaccinia virus expressing several HIV-2 SBL6669 proteins, including the envelope glycoproteins, in combination with booster immunizations with gp160, gp125, or V3 synthetic peptides, elicited a weak neutralizing response and partial protection against autologous HIV-2 challenge. Likewise, vaccination of rhesus monkeys with immunogens derived from the historic HIV-2ROD strain failed to generate neutralizing antibodies and to protect against heterologous challenge (55). Finally, baboons inoculated with a DNA vaccine expressing the tat, nef, gag, and env genes of the HIV-2UC2 group B isolate were partially protected against autologous challenge without the production of neutralizing antibodies (33). These studies illustrate the urgent need for new vaccine immunogens and/or vaccination strategies that elicit the production of broadly reactive NAbs against HIV-2. The present study was designed to investigate in the mouse model the immunogenicity and neutralizing response elicited by novel recombinant envelope proteins derived from the reference primary HIV-2ALI isolate, when administered alone or in different prime-boost combinations.  相似文献   

15.
It was shown previously that the highly conserved vaccinia virus A35 gene is an important virulence factor in respiratory infection of mice. We show here that A35 is also required for full virulence by the intraperitoneal route of infection. A virus mutant in which the A35 gene has been removed replicated normally and elicited improved antibody, gamma interferon-secreting cell, and cytotoxic T-lymphocyte responses compared to wild-type virus, suggesting that A35 increases poxvirus virulence by immunomodulation. The enhanced immune response correlated with an improved control of viral titers in target organs after the development of the specific immune response. Finally, the A35 deletion mutant virus also provided protection from lethal challenge (1,000 50% lethal doses) equal to that of the wild-type virus. Together, these data suggest that A35 deletion viruses will make safer and more efficacious vaccines for poxviruses. In addition, the A35 deletion viruses will serve as improved platform vectors for other infectious diseases and cancer and will be superior vaccine choices for postexposure poxvirus vaccination, as they also provide improved kinetics of the immune response.Poxviruses are large, complex viruses with a broad host range and worldwide distribution (30). Members of the family Poxviridae include variola virus, the causative agent of smallpox, which induced a fatality rate of approximately 30% and killed hundreds of millions of people before its eradication in 1980 (28). Currently, the most dangerous extant human-infecting poxvirus is monkeypox virus, which is commonly found in African rodents. Monkeypox virus causes a smallpox-like illness with a 10% fatality rate. A recent study showed that 1.7% of people in the Likouala region in Africa had monkeypox-specific immunoglobulin M (IgM), indicating a significant ongoing infection rate (25). An outbreak of a low-virulence strain of monkeypox virus occurred in the United States in 2003, causing more than 80 human infections and several hospitalizations (6). This outbreak raises concern that monkeypox virus could establish itself in wild-rodent populations in North America (34), thus creating a local zoonotic reservoir for this emerging pathogen. Of further concern are the facts that monkeypox is spreading more efficiently in humans (18, 24, 31) and that the current poxvirus vaccine is not universally protective against monkeypox infection (27). Both variola and monkeypox viruses are considered bioterrorism and biowarfare concerns and are category A select-agent pathogens. There are also other poxvirus infections that sporadically cause human outbreaks, including Cantagalo virus in South America (10, 41) and buffalopox virus in India (23), and the incidence of tanapox virus appears to be increasing (12, 43). Molluscum contagiosum poxvirus accounts for approximately 300,000 doctor visits each year in the United States alone (29). Thus, the study of virulence mechanisms in this group of viruses is important.The eradication of smallpox was accomplished through the use of the related vaccinia virus (VV) as a live-virus vaccine. Despite its phenomenal success, the public vaccination program was discontinued because of the high incidence of complications due to the virulence of wild-type VV. It is estimated that approximately 25% of the population should not receive this vaccine because of immunodeficiency, eczema, pregnancy, or heart disease (14, 21, 45). Safer vaccines are necessary to protect against emerging or released poxviruses. In addition, poxviruses are being used as platform vaccines for other diseases such as human immunodeficiency virus, malaria, and cancer because they induce a robust immune response and accommodate the insertion of large pieces of foreign DNA. It is therefore of great importance to identify poxvirus virulence genes in order to develop safer and more effective poxvirus vaccines. Replication-defective strains, such as Modified Vaccinia Ankara, have been used in an effort to reduce the risks associated with vaccination (8, 47), but the production of these viruses can be challenging, they require higher doses of vaccine, and their protective efficacy against poxvirus infections in humans is unknown. As poxviruses inhibit the activation of antigen-presenting cells and antigen presentation (26, 38), another way to construct a safer vaccine is to develop replication-competent vaccine strains that exclude immunosuppressive genes (3) while retaining protective antigenic epitopes (17, 32). We show herein that the A35R gene is an excellent candidate for removal from vaccine strains.The VV A35 gene is highly conserved in mammalian-tropic poxviruses, and a sequence identity search has revealed that the protein has little similarity to any other poxvirus protein or any nonpoxvirus protein, suggesting that this gene has an important and novel function (39). We have shown that the A35 gene is not required for viral replication in vitro but is required for full virulence in the mouse model (39).We therefore tested the effects of A35R on immune responses during infection in the mouse model and tested its protective efficacy against virulent challenge.  相似文献   

16.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

17.
18.
19.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

20.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  相似文献   

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