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.
A major obstacle to gene transduction by viral vectors is inactivation by human complement in vivo. One way to overcome this is to incorporate complement regulatory proteins, such as CD55/decay accelerating factor (DAF), into viral particles. Lentivirus vectors pseudotyped with the baculovirus envelope protein GP64 have been shown to acquire more potent resistance to serum inactivation and longer transgene expression than those pseudotyped with the vesicular stomatitis virus (VSV) envelope protein G. However, the molecular mechanisms underlying resistance to serum inactivation in pseudotype particles bearing the GP64 have not been precisely elucidated. In this study, we generated pseudotype and recombinant VSVs bearing the GP64. Recombinant VSVs generated in human cell lines exhibited the incorporation of human DAF in viral particles and were resistant to serum inactivation, whereas those generated in insect cells exhibited no incorporation of human DAF and were sensitive to complement inactivation. The GP64 and human DAF were detected on the detergent-resistant membrane and were coprecipitated by immunoprecipitation analysis. A pseudotype VSV bearing GP64 produced in human DAF knockdown cells reduced resistance to serum inactivation. In contrast, recombinant baculoviruses generated in insect cells expressing human DAF or carrying the human DAF gene exhibited resistance to complement inactivation. These results suggest that the incorporation of human DAF into viral particles by interacting with baculovirus GP64 is involved in the acquisition of resistance to serum inactivation.Gene therapy is a potential treatment option for genetic diseases, malignant diseases, and other acquired diseases. To this end, safe and efficient gene transfer into specific target cells is a central requirement, and a variety of nonviral and viral vector systems have been developed (6, 44). Recombinant viruses can be used for efficient gene transfer. Retroviruses, adeno-associated viruses, and lentiviruses are able to integrate foreign genes into host genomes and are suitable for gene therapeutics by virtue of their permanent expression of the therapeutic genes, whereas adenoviruses, herpesviruses, and baculoviruses can transiently express foreign genes (6, 12, 44). Pseudotype particles bearing other viral envelope proteins have been developed to improve transduction efficiency and the safety of viral vectors, including retrovirus (4, 7), lentivirus (25), vesicular stomatitis virus (VSV) (29), and baculovirus (17, 42). Pseudotype retroviruses and lentiviruses bearing the baculovirus envelope protein GP64 of Autographa californica nucleopolyhedrosis virus (AcNPV) have been shown to exhibit efficient gene transduction into a wide variety of cells with a lower cytotoxicity compared to those bearing the VSV envelope protein G (VSVG), which is commonly used for pseudotyping (18, 32, 35, 36).However, a drawback of gene transduction by viral vectors is that human sera inactivate the vectors (11, 40). Complement is a major element of the innate immune response and serves to link innate and adaptive immunity (8). Complement activation can occur via classical, lectin, and alternative pathways (2, 8). All pathways invoke several responses, such as virus opsonization, virolysis, anaphylatoxin, and chemotaxin production, as well as others (2, 8). VSV and baculovirus are inactivated by human sera via the classical pathway (1, 11). Because complement activation also induces potential damage to host cells, the complement system is tightly regulated by the complement regulatory proteins (CRPs), including CD55/decay-accelerating factor (DAF), CD46/membrane cofactor protein (MCP), and CD59 (2, 8, 15). DAF and CD46 inhibit activation of C3/C5-converting enzymes, which regulate the activation of classical and alternative pathways, whereas CD59 regulates the assembly of the membrane attack complex (2, 8, 15).Viral vectors can be manipulated to confer resistance to the complement inactivation. Human immunodeficiency virus (HIV) is known to develop resistance to human complement through the incorporation of DAF, CD46, and CD59 to the viral particles (22, 30, 31, 38). Moloney murine leukemia virus vectors produced in HT1080 cells are resistant to complement inactivation (5). Baculovirus and lentivirus vectors bearing DAF or the fusion protein between the functional domains of human DAF and the GP64 were resistant to complement inactivation (9, 13). It has been shown that lentivirus vectors pseudotyped with the GP64 are more resistant to inactivation in the sera of mice and rats (14, 32) and are capable of executing longer expression of the transgenes in nasal epithelia compared to those pseudotyped with the VSVG (35, 36). However, the precise mechanisms underlying the resistance to complement inactivation by pseudotyping of the GP64 is not known.To clarify the molecular mechanisms underlying the resistance of the viral vectors pseudotyped with the GP64 to the complement inactivation, we produced pseudotype and recombinant VSVs bearing the GP64. The recombinant VSVs carrying the gp64 gene generated in human cells but not in insect cells exhibited incorporation of human DAF on the viral particles and were resistant to the complement inactivation. Furthermore, production of the gp64 pseudotype VSV in the DAF knockdown human cells impaired serum resistance, whereas production of the gp64 recombinant VSV in the CHO cell lines stably expressing human DAF and the recombinant baculoviruses in the insect cells stably expressing human DAF or encoding the DAF gene in the genome conferred resistance to the complement inactivation. These results suggest that DAF incorporation into viral particles bearing baculovirus GP64 confers resistance to serum inactivation.  相似文献   

2.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

3.
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   

4.
5.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

6.
7.
The spatial organization of metastable paramyxovirus fusion (F) and attachment glycoprotein hetero-oligomers is largely unknown. To further elucidate the organization of functional fusion complexes of measles virus (MeV), an archetype of the paramyxovirus family, we subjected central predictions of alternative docking models to experimental testing using three distinct approaches. Carbohydrate shielding through engineered N-glycans indicates close proximity of a membrane-distal, but not membrane-proximal, section of the MeV attachment (H) protein stalk domain to F. Directed mutagenesis of this section identified residues 111, 114, and 118 as modulators of avidity of glycoprotein interactions and determinants of F triggering. Stalk-length variation through deletion or insertion of HR elements at positions flanking this section demonstrates that the location of the stalk segment containing these residues cannot be altered in functional fusion complexes. In contrast, increasing the distance between the H head domains harboring the receptor binding sites and this section through insertion of structurally rigid α-helical domains with a pitch of up to approximately 75 Å downstream of stalk position 118 partially maintains functionality in transient expression assays and supports efficient growth of recombinant virions. In aggregate, these findings argue against specific protein-protein contacts between the H head and F head domains but instead support a docking model that is characterized by short-range contacts between the prefusion F head and the attachment protein stalk, possibly involving H residues 111, 114, and 118, and extension of the head domain of the attachment protein above prefusion F.Paramyxoviruses infect cells through fusion of the viral envelope with target cell membranes. For all members of the Paramyxovirinae subfamily, this involves the concerted action of two envelope glycoproteins, the fusion (F) and attachment (H, HN, or G, depending on the Paramyxovirinae genus) proteins. Both proteins feature short lumenal tails, a single transmembrane domain, and large ectodomains. The F protein, in type I orientation, forms homotrimers, while homodimers or homotetramers have been suggested as functional units for attachment proteins of different Paramyxovirinae subfamily members (7, 14, 28, 41, 49, 50, 66). For entry, upon receptor binding, the attachment protein is considered to initiate a series of conformational rearrangements in the metastable prefusion F protein (15, 77), which ultimately brings together transmembrane domains and fusion peptides and, thus, donor and target membranes (3, 32, 45, 53, 80).Multiple studies have demonstrated that specific interactions between compatible F and attachment proteins of paramyxovirinae are imperative for the formation of functional fusion complexes (6, 29, 36, 42, 43, 56, 75). However, the molecular nature of these interactions and the spatial organization of functional glycoprotein hetero-oligomers remain largely unknown. Individual ectodomain and partial ectodomain crystal structures have been obtained for different paramyxovirus F (13, 76, 77) and attachment (8, 14, 17, 28, 35, 79) proteins, respectively. For F, a stabilized human parainfluenza virus type 5 (HPIV5) ectodomain that is believed to represent a prefusion conformation folds into a globular head structure that is attached to the transmembrane domains through a helical stalk consisting of the membrane-proximal heptad repeat B (HR-B) domains (77). For the attachment protein, a globular head that harbors the receptor binding sites is considered to be connected to the transmembrane region through extended stalk domains (34, 78). Crystal structures of isolated head domains have been solved for several paramyxovirus attachment proteins, including measles virus (MeV) H, and reveal the six-blade propeller fold typical of sialidase structures (8, 14, 17, 28, 79). However, morbilliviruses recognize proteinaceous receptors (for MeV, the regulator of complement activation [CD46] and/or signaling lymphocytic activation molecule [SLAM], depending on the virus strain) (21, 40, 46, 51, 64, 65). X-ray data do not extend to the stalk domains, but circular dichroism analysis (78) and structure predictions (36, 78) support an α-helical coiled-coil configuration of the stalk.The nature of individual residues that engage in specific intermolecular interactions between glycoproteins of paramyxovirinae prior to refolding has been studied most extensively for the attachment protein. The stalk domains of several paramyxovirus HN proteins have been implicated in mediating specificity for their homotypic F proteins (18, 20, 43, 63, 70, 72). We have found that this extends to MeV and canine distemper virus H and, thus, to paramyxovirinae recognizing proteinaceous receptors (36), supporting the general hypothesis that F-interacting residues may reside in the stalk region of the attachment protein (30, 78).Considerably less information concerning the nature of F microdomains that mediate attachment protein specificity is available. Among the few exceptions are peptides derived from Newcastle disease virus (NDV) and Sendai virus F HR-B domains, which interact with soluble variants of the respective HN proteins in vitro (25, 67). Multiple domains have been suggested to mediate specificity of HPIV2 F for its HN (69). However, a conclusive N-glycan shielding study (43) and structural information (77, 78) argue against direct contacts between NDV F HR-B domains and HN in native glycoprotein complexes. Thus, the role of individual HPIV2 F residues in HN binding is unclear (25, 43).Building on the observation that MeV H is able to engage in productive heterotypic interactions with F proteins derived from some but not all isolates of closely related canine distemper virus, we have recently identified residues in morbillivirus F (MeV F residue 121) and H (H stalk residues 110 to 114) that interdependently contribute to physical MeV glycoprotein interaction and F triggering for fusion (36). While these residues could mediate reciprocal glycoprotein specificity through long-range effects, molecular modeling of the MeV H stalk in an α-helical conformation has posited F residue 121 at the same level above the viral envelope as H residues 110 to 114, making direct contacts structurally conceivable (36). This spatial organization of functional fusion complexes furthermore provides a comprehensive explanation for previous demonstrations of a specific role for attachment protein stalk domains of paramyxovirinae in functional and physical interactions with F (18, 43, 63, 70, 72). However, this “staggered-head” model mandates positioning the globular head of the attachment protein above the prefusion F trimer (36), as opposed to a suggested “parallel-head” alignment of the glycoproteins (31, 47). The latter is mostly based on transmission electron microscopy micrographs of viral particles apparently showing glycoprotein spikes of equal length (33). Unfortunately, these images lack the resolution for an identification of the molecular nature of the spikes (attachment or F protein) or the distinguishing between densely packaged H and F head domains of different heights and laterally aligned head domains. Indeed, a recent single-particle reconstruction based on cryo-electron microscopy images of HPIV5 particles revealed that defined spikes correspond to F in a postfusion conformation, which was interpreted as a product of possible premature F refolding (38). These two-dimensional images of heavy-metal-stained particles did not reveal F spikes in a prefusion conformation. Rather, a dense surface layer was considered to correspond to prefusion glycoprotein hetero-oligomers (38). In addition to further-advanced image reconstructions, biochemical assessment of alternative docking modes is imperative for the elucidation of the organization of functional fusion complexes of paramyxovirinae.In this study, we subjected central predictions of the hypothetical alignment models to experimental analysis. By employing carbohydrate shielding, directed mutagenesis, and variation of the length of the H stalk domain, we examined the proximity of different regions of the H stalk to F, probed a role of individual residues around the previously identified H stalk section from positions 110 to 114 in the formation of functional fusion complexes, tested the effect of varying the length of the H stalk membrane proximal and membrane distal to this section, and explored the general possibility of whether specific contacts between the prefusion F and H head domains are required for F triggering. Experimental data were interpreted in the light of a working model of MeV glycoprotein hetero-oligomers prior to receptor binding.  相似文献   

8.
9.
10.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

11.
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   

12.
13.
14.
15.
The quest to create a human immunodeficiency virus type 1 (HIV-1) vaccine capable of eliciting broadly neutralizing antibodies against Env has been challenging. Among other problems, one difficulty in creating a potent immunogen resides in the substantial overall sequence variability of the HIV envelope protein. The membrane-proximal region (MPER) of gp41 is a particularly conserved tryptophan-rich region spanning residues 659 to 683, which is recognized by three broadly neutralizing monoclonal antibodies (bnMAbs), 2F5, Z13, and 4E10. In this study, we first describe the variability of residues in the gp41 MPER and report on the invariant nature of 15 out of 25 amino acids comprising this region. Subsequently, we evaluate the ability of the bnMAb 2F5 to recognize 31 varying sequences of the gp41 MPER at a molecular level. In 19 cases, resulting crystal structures show the various MPER peptides bound to the 2F5 Fab′. A variety of amino acid substitutions outside the 664DKW666 core epitope are tolerated. However, changes at the 664DKW666 motif itself are restricted to those residues that preserve the aspartate''s negative charge, the hydrophobic alkyl-π stacking arrangement between the β-turn lysine and tryptophan, and the positive charge of the former. We also characterize a possible molecular mechanism of 2F5 escape by sequence variability at position 667, which is often observed in HIV-1 clade C isolates. Based on our results, we propose a somewhat more flexible molecular model of epitope recognition by bnMAb 2F5, which could guide future attempts at designing small-molecule MPER-like vaccines capable of eliciting 2F5-like antibodies.Eliciting broadly neutralizing antibodies (bnAbs) against primary isolates of human immunodeficiency virus type I (HIV-1) has been identified as a major milestone to attain in the quest for a vaccine in the fight against AIDS (12, 28). These antibodies would need to interact with HIV-1 envelope glycoproteins gp41 and/or gp120 (Env), target conserved regions and functional conformations of gp41/gp120 trimeric complexes, and prevent new HIV-1 fusion events with target cells (21, 57, 70, 71). Although a humoral response generating neutralizing antibodies against HIV-1 can be detected in HIV-1-positive individuals, the titers are often very low, and virus control is seldom achieved by these neutralizing antibodies (22, 51, 52, 66, 67). The difficulty in eliciting a broad and potent neutralizing antibody response against HIV-1 is thought to reside in the high degree of genetic diversity of the virus, in the heterogeneity of Env on the surface of HIV-1, and in the masking of functional regions by conformational covering, by an extensive glycan shield, or by the ability of some conserved domains to partition to the viral membrane (24, 25, 29, 30, 38, 39, 56, 68, 69). So far, vaccine trials using as immunogens mimics of Env in different conformations have primarily elicited antibodies with only limited neutralization potency across different HIV-1 clades although recent work has demonstrated more encouraging results (4, 12, 61).The use of conserved regions on gp41 and gp120 Env as targets for vaccine design has been mostly characterized by the very few anti-HIV-1 broadly neutralizing monoclonal antibodies (bnMAbs) that recognize them: the CD4 binding-site on gp120 (bnMAb b12), a CD4-induced gp120 coreceptor binding site (bnMAbs 17b and X5), a mannose cluster on the outer face of gp120 (bnMAb 2G12), and the membrane proximal external region (MPER) of gp41 (bnMAbs 2F5, Z13 and 4E10) (13, 29, 44, 58, 73). The gp41 MPER region is a particularly conserved part of Env that spans residues 659 to 683 (HXB2 numbering) (37, 75). Substitution and deletion studies have linked this unusually tryptophan-rich region to the fusion process of HIV-1, possibly involving a series of conformational changes (5, 37, 41, 49, 54, 74). Additionally, the gp41 MPER has been implicated in gp41 oligomerization, membrane leakage ability facilitating pore formation, and binding to the galactosyl ceramide receptor on epithelial cells for initial mucosal infection mediated by transcytosis (2, 3, 40, 53, 63, 64, 72). This wide array of roles for the gp41 MPER will put considerable pressure on sequence conservation, and any change will certainly lead to a high cost in viral fitness.Monoclonal antibody 2F5 is a broadly neutralizing monoclonal anti-HIV-1 antibody isolated from a panel of sera from naturally infected asymptomatic individuals. It reacts with a core gp41 MPER epitope spanning residues 662 to 668 with the linear sequence ELDKWAS (6, 11, 42, 62, 75). 2F5 immunoglobulin G binding studies and screening of phage display libraries demonstrated that the DKW core is essential for 2F5 recognition and binding (15, 36, 50). Crystal structures of 2F5 with peptides representing its core gp41 epitope reveal a β-turn conformation involving the central DKW residues, flanked by an extended conformation and a canonical α-helical turn for residues located at the N terminus and C terminus of the core, respectively (9, 27, 45, 47). In addition to binding to its primary epitope, evidence is accumulating that 2F5 also undergoes secondary interactions: multiple reports have demonstrated affinity of 2F5 for membrane components, possibly through its partly hydrophobic flexible elongated complementarity-determining region (CDR) H3 loop, and it has also been suggested that 2F5 might interact in a secondary manner with other regions of gp41 (1, 10, 23, 32, 33, 55). Altogether, even though the characteristics of 2F5 interaction with its linear MPER consensus epitope have been described extensively, a number of questions persist about the exact mechanism of 2F5 neutralization at a molecular level.One such ambiguous area of the neutralization mechanism of 2F5 is investigated in this study. Indeed, compared to bnMAb 4E10, 2F5 is the more potent neutralizing antibody although its breadth across different HIV-1 isolates is more limited (6, 35). In an attempt to shed light on the exact molecular requirements for 2F5 recognition of its primary gp41 MPER epitope, we performed structural studies of 2F5 Fab′ with a variety of peptides. The remarkable breadth of possible 2F5 interactions reveals a somewhat surprising promiscuity of the 2F5 binding site. Furthermore, we link our structural observations with the natural variation observed within the gp41 MPER and discuss possible routes of 2F5 escape from a molecular standpoint. Finally, our discovery of 2F5''s ability to tolerate a rather broad spectrum of amino acids in its binding, a spectrum that even includes nonnatural amino acids, opens the door to new ways to design small-molecule immunogens potentially capable of eliciting 2F5-like neutralizing antibodies.  相似文献   

16.
17.
18.
The herpes simplex virus (HSV) virion host shutoff protein (vhs) encoded by gene UL41 is an mRNA-specific RNase that triggers accelerated degradation of host and viral mRNAs in infected cells. We report here that vhs is also able to modulate reporter gene expression without greatly altering the levels of the target mRNA in transient-transfection assays conducted in HeLa cells. We monitored the effects of vhs on a panel of bicistronic reporter constructs bearing a variety of internal ribosome entry sites (IRESs) located between two test cistrons. As expected, vhs inhibited the expression of the 5′ cistrons of all of these constructs; however, the response of the 3′ cistron varied with the IRES: expression driven from the wild-type EMCV IRES was strongly suppressed, while expression controlled by a mutant EMCV IRES and the cellular ApaF1, BiP, and DAP5 IRES elements was strongly activated. In addition, several HSV type 1 (HSV-1) 5′ untranslated region (5′ UTR) sequences also served as positive vhs response elements in this assay. IRES activation was also observed in 293 and HepG2 cells, but no such response was observed in Vero cells. Mutational analysis has yet to uncouple the ability of vhs to activate 3′ cistron expression from its shutoff activity. Remarkably, repression of 5′ cistron expression could be observed under conditions where the levels of the reporter RNA were not correspondingly reduced. These data provide strong evidence that vhs can modulate gene expression at the level of translation and that it is able to activate cap-independent translation through specific cis-acting elements.The virion host shutoff protein (vhs) encoded by herpes simplex virus (HSV) gene UL41 is an endoribonuclease that is packaged into the tegument of mature HSV virions. Once delivered into the cytoplasm of newly infected cells, vhs triggers shutoff of host protein synthesis, disruption of preexisting polysomes, and degradation of host mRNAs (reviewed in reference 62). The vhs-dependent shutoff system destabilizes many cellular and viral mRNAs (36, 46, 67). The rapid decline in host mRNA levels presumably helps viral mRNAs gain access to the cellular translational apparatus. In addition, the relatively short half-lives of viral mRNAs contribute to the sharp transitions between the successive phases of viral protein synthesis by tightly coupling changes in the rates of synthesis of viral mRNAs to altered mRNA levels (46). These effects enhance virus replication and may account for the modest reduction in virus yield displayed by vhs mutants in cultured Vero cells (55, 61).vhs also plays a critical role in HSV pathogenesis: vhs mutants are severely impaired for replication in the corneas and central nervous systems of mice and cannot efficiently establish or reactivate from latency (63, 65, 66). Mounting evidence indicates that this attenuation stems at least in part from an impaired ability to disarm elements of the innate and adaptive host immune responses (reviewed in reference 62). For example, vhs suppresses certain innate cellular antiviral responses, including production of proinflammatory cytokines and chemokines (68); dampens the type I interferon system (11, 45, 49, 78); and blocks activation of dendritic cells (58). Moreover, vhs mutants display enhanced virulence in knockout mice lacking type I interferon (IFN) receptors (37, 45) or Stat1 (48) and are hypersensitive to the antiviral effects of IFN in some cells in tissue culture (11, 49, 68). Thus, vhs is arguably a bona fide virulence factor.vhs present in extracts of HSV virions or purified from bacteria has nonspecific RNase activity capable of degrading all RNA substrates (15, 70, 71, 79). However, vhs is highly selective in vivo, targeting mRNAs and sparing other cytoplasmic RNAs (36, 46). In vivo and in mammalian whole-cell extracts, vhs-induced decay of at least some mRNAs initiates near regions of translation initiation and proceeds in an overall 5′-to-3′ direction (12, 13, 29, 52). Moreover, vhs binds to the translation initiation factors eIF4H, eIF4B, and eIF4A II, all components of the cap recognition factor eIF4F (10, 16, 17). Thus, it has been proposed that vhs selectively targets actively translated mRNAs through interactions with eIF4F components (17). Consistent with this hypothesis, recent data document that eIF4H is required for vhs activity in vivo (59).A previous report from this laboratory documented that the internal ribosome entry sites (IRESs) of the picornaviruses poliovirus and encephalomyocarditis virus (EMCV) strongly target vhs-induced RNA cleavage events to sequences immediately 3′ to the IRES in an in vitro translation system derived from rabbit reticulocyte lysates (RRL) (13). IRES elements are highly structured RNA sequences that are able to direct cap-independent translational initiation (reviewed in references 21, 25, 30, and 64). In the case of the poliovirus and EMCV elements, this is achieved by directly recruiting the eIF4F scaffolding protein eIF4G, thus bypassing the requirement for the cap-binding eIF4F subunit, eIF4E (reviewed in reference 30). Based on these data, we suggested that vhs is strongly targeted to the picornavirus IRES elements via interactions with eIF4 factors.A growing number of cellular mRNAs have been proposed to bear IRES elements in their 5′ untranslated regions (5′ UTRs). These include many that are involved in cellular stress responses, apoptosis, and cell cycle progression (24, 64, 74). Given the striking ability of picornavirus IRES elements to target vhs RNase activity in vitro, we asked whether viral and cellular IRES elements are able to modify the susceptibility of mRNAs to vhs in vivo. During the course of preliminary experiments designed to test this hypothesis, we unexpectedly discovered that vhs is able to strongly activate gene expression controlled by some cellular IRES elements and HSV 5′ UTR sequences in in vivo bicistronic reporter assays. These observations are the subject of the present report.  相似文献   

19.
We identify Kette, a key regulator of actin polymerization, as a substrate for Drosophila protein tyrosine phosphatase PTP61F, as well as for dAbl tyrosine kinase. We further show that dAbl is a direct substrate for PTP61F. Therefore, Kette phosphotyrosine levels are regulated both directly and indirectly by PTP61F. Kette and PTP61F genetically interact in the regulation of F-actin organization in pupal eye discs, suggesting that tyrosine phosphorylation is essential for the proper regulation of Kette-mediated actin dynamics. This hypothesis was confirmed by demonstrating the loss of Kette-mediated F-actin organization and lamella formation in S2 cells in a Kette Y482F mutant in which the dAbl phosphorylation site was eliminated. Our results establish for the first time that PTP61F and dAbl ensure proper actin organization through the coordinated and reversible tyrosine phosphorylation of Kette.The actin cytoskeleton is regulated as a function of development, cell motility, intracellular transport, and the cell cycle by the polymerization of G-actin to F-actin (34). Correct regulation of actin cytoskeletal dynamics is essential to numerous differentiating and cellular processes in the nervous system (9) and musculature (42), among others. Actin polymerization is regulated by a number of proteins, among which is human NCK-associated protein 1 (NAP-1 [3, 4, 45]). It and its Drosophila orthologue, Kette (Hem in FlyBase), are critical components in both SCAR/WAVE and WASP complexes, which play essential roles in transducing Rac1 signals to initiate Arp2/3-dependent actin polymerization (6, 25, 40, 48). Murine NAP-1 interacts with NCK, an SH2-SH3 adaptor protein (4), and is essential for proper neuronal differentiation in the cortex (53). Neuronal differentiation and neural tube defects are observed in NAP-1 mutant mice, apparently due to reduced localization of WAVE1 to the cell membrane (53).In Drosophila, loss of kette activity specifically results in the accumulation of cytosolic F-actin (6). Kette protein associates with F-actin in the cytosol, but also at focal contact sites, where it apparently antagonizes SCAR/WAVE function and activates WASP-dependent actin polymerization (6). Despite its role in repressing SCAR/WAVE function, Kette serves to protect the complex from proteosome-mediated degradation and is critical to its intracellular localization (25). At the level of the organism, kette alleles affect axonal growth and pathfinding due to aberrant actin cytoskeleton formation, for example, altering crossing of the embryonic ventral midline by VUM neuron axons, as well as generating aberrant axonal projections in both motor and sensory neurons (21). Like mammalian NAP-1, Drosophila kette also interacts with the fly NCK orthologue, dreadlocks (dock) (21). Other evidence for the conserved interaction of Kette with signaling cascades is provided by the observation that kette mutant phenotypes are partially rescuable by overexpression of the small G protein Rac1 (21). The interaction of kette with dock suggests the possibility of tyrosine phosphorylation in the regulation of Kette activity, but no evidence supporting this hypothesis has been reported.Signaling by tyrosine phosphorylation in various metazoans controls numerous processes involved in cellular differentiation and proliferation. Many of the components regulating tyrosine phosphorylation have been identified and characterized using genetic, biochemical, molecular, and genomic sequence analyses (31). However, in contrast to the very well-characterized regulation of cellular processes by kinase-mediated tyrosine phosphorylation (15, 52), their regulation by dephosphorylation by protein tyrosine phosphatases (PTPs) has generally lagged behind. Although the functions of several receptor PTPs have been clearly defined as playing essential roles in axon guidance in both Drosophila (12, 23, 41, 47, 50) and mammals (44, 49), our understanding of nontransmembrane PTPs (NT-PTPs) is more limited. Only three of the eight putative Drosophila NT-PTPs have been characterized genetically. Corkscrew (Csw) acts as a downstream effector of various receptor protein tyrosine kinases (PTKs) and is essential for R7 photoreceptor development (35). PTP-enhancer of Ras1 has been characterized as an essential regulator antagonizing signaling mediated by Ras1, possibly through tyrosine dephosphorylation of mitogen-activated protein kinase (24, 36). More recently, it has been shown that PTP-meg participates in the establishment and maintenance of axon projections in the Drosophila brain (51). Other than these, the functions of Drosophila NT-PTPs remain largely unknown.PTP61F was originally identified as an NT-PTP that contains one phosphatase domain in the N-terminal region and five proline-rich motifs in the C-terminal tail (29). It is the Drosophila orthologue of mammalian PTP1B and T-cell PTP (TC-PTP) (1), which have been implicated in the regulation of signaling by both insulin (39) and JAK/STAT (33). Two PTP61F isoforms due to alternative splicing possess unique sequences at the C terminus, which determine either internal membrane-association (PTP61Fm) or nuclear localization (PTP61Fn) (29). To date, limited data suggest that PTP61F may participate in the downregulation of JAK/STAT signaling (2, 32), although the underlying mechanism remains unexplored. While PTP61F may recognize the adaptor proteins DOCK (10) and Abi (20) as potential substrates, the signaling pathways involving these interactions have not been clearly defined. In this study, we demonstrate for the first time that the regulation of Kette, and hence the localization and polymerization of the actin cytoskeleton, is achieved by reversible tyrosine phosphorylation under the control of both PTP61F and the PTK dAbl.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号