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 93 毫秒
1.
2.
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   

3.
The human immunodeficiency virus type 1 structural polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles on cellular membranes. Previous studies demonstrated the importance of the capsid C-terminal domain (CA-CTD), nucleocapsid (NC), and membrane association in Gag-Gag interactions, but the relationships between these factors remain unclear. In this study, we systematically altered the CA-CTD, NC, and the ability to bind membrane to determine the relative contributions of, and interplay between, these factors. To directly measure Gag-Gag interactions, we utilized chimeric Gag-fluorescent protein fusion constructs and a fluorescence resonance energy transfer (FRET) stoichiometry method. We found that the CA-CTD is essential for Gag-Gag interactions at the plasma membrane, as the disruption of the CA-CTD has severe impacts on FRET. Data from experiments in which wild-type (WT) and CA-CTD mutant Gag molecules are coexpressed support the idea that the CA-CTD dimerization interface consists of two reciprocal interactions. Mutations in NC have less-severe impacts on FRET between normally myristoylated Gag proteins than do CA-CTD mutations. Notably, when nonmyristoylated Gag interacts with WT Gag, NC is essential for FRET despite the presence of the CA-CTD. In contrast, constitutively enhanced membrane binding eliminates the need for NC to produce a WT level of FRET. These results from cell-based experiments suggest a model in which both membrane binding and NC-RNA interactions serve similar scaffolding functions so that one can functionally compensate for a defect in the other.The human immunodeficiency virus type 1 (HIV-1) structural precursor polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles (VLPs). Gag is composed of four major structural domains, matrix (MA), capsid (CA), nucleocapsid (NC), and p6, as well as two spacer peptides, SP1 and SP2 (3, 30, 94). Following particle assembly and release, cleavage by HIV-1 protease separates these domains. However, these domains must work together in the context of the full-length Gag polyprotein to drive particle assembly.Previous studies have mapped two major functional domains involved in the early steps of assembly: first, Gag associates with cellular membranes via basic residues and N-terminal myristoylation of the MA domain (10, 17, 20, 35, 39, 87, 91, 106); second, the Gag-Gag interaction domains that span the CA C-terminal domain (CA-CTD) and NC domain promote Gag multimerization (3, 11, 14, 16, 18, 23, 27, 29, 30, 33, 36, 46, 64, 88, 94, 102, 103). Structural and genetic studies have identified two residues (W184 and M185) within a dimerization interface in the CA-CTD that are critical to CA-CA interactions (33, 51, 74, 96). Analytical ultracentrifugation of heterodimers formed between wild-type (WT) Gag and Gag mutants with changes at these residues suggests that the dimerization interface consists of two reciprocal interactions, one of which can be disrupted to form a “half-interface” (22).In addition to the CA-CTD, NC contributes to assembly via 15 basic residues (8, 9, 11, 14, 18, 23, 25, 28, 34, 40, 43, 54, 57, 58, 74, 79, 88, 97, 104, 105), although some researchers have suggested that NC instead contributes to the stability of mature virions after assembly (75, 98, 99). It is thought that the contribution of NC to assembly is due to its ability to bind RNA, since the addition of RNA promotes the formation of particles in vitro (14-16, 37, 46), and RNase treatment disrupts Gag-Gag interactions (11) and immature viral cores (67). However, RNA is not necessary per se, since dimerization motifs can substitute for NC (1, 4, 19, 49, 105). This suggests a model in which RNA serves a structural role, such as a scaffold, to promote Gag-Gag interactions through NC. Based on in vitro studies, it has been suggested that this RNA scaffolding interaction facilitates the low-order Gag multimerization mediated by CA-CTD dimerization (4, 37, 49, 62, 63, 85). Despite a wealth of biochemical data, the relative contributions of the CA-CTD and NC to Gag multimerization leading to assembly are yet to be determined in cells.Mutations in Gag interaction domains alter membrane binding in addition to affecting Gag multimerization. In particular, mutations or truncations of CA reduce membrane binding (21, 74, 82), and others previously reported that mutations or truncations of NC affect membrane binding (13, 78, 89, 107). These findings are consistent with a myristoyl switch model of membrane binding in which Gag can switch between high- and low-membrane-affinity states (38, 71, 76, 83, 86, 87, 92, 95, 107). Many have proposed, and some have provided direct evidence (95), that Gag multimerization mediated by CA or NC interactions promotes the exposure of the myristoyl moiety to facilitate membrane associations.Gag membrane binding and multimerization appear to be interrelated steps of virus assembly, since membrane binding also facilitates Gag multimerization. Unlike betaretroviruses that fully assemble prior to membrane targeting and envelopment (type B/D), lentiviruses, such as HIV, assemble only on cellular membranes at normal Gag expression levels (type C), although non-membrane-bound Gag complexes exist (45, 58, 60, 61, 65). Consistent with this finding, mutations that reduce Gag membrane associations cause a defect in Gag multimerization (59, 74). Therefore, in addition to their primary effects on Gag-Gag interactions, mutations in Gag interaction domains cause a defect in membrane binding, which, in turn, causes a secondary multimerization defect. To determine the relative contributions of the CA-CTD and the NC domain to Gag-Gag interactions at the plasma membrane, it is essential to eliminate secondary effects due to a modulation of membrane binding.Except for studies using a His-tag-mediated membrane binding system (5, 46), biochemical studies of C-type Gag multimerization typically lack membranes. Therefore, these studies do not fully represent particle assembly, which occurs on biological membranes in cells. Furthermore, many biochemical and structural approaches are limited to isolated domains or truncated Gag constructs. Thus, some of these studies are perhaps more relevant to the behavior of protease-cleaved Gag in mature virions. With few exceptions (47, 74), cell-based studies of Gag multimerization have typically been limited to measuring how well mutant Gag is incorporated into VLPs when coexpressed or not with WT Gag. Since VLP production is a complex multistep process, effects of mutations on other steps in the process can confound this indirect measure. For example, NC contributes to VLP production by both promoting multimerization and interacting with the host factor ALIX to promote VLP release (26, 80). To directly assay Gag multimerization in cells, several groups (24, 45, 52, 56) developed microscopy assays based on fluorescence resonance energy transfer (FRET). These assays measure the transfer of energy between donor and acceptor fluorescent molecules that are brought within ∼5 nm by the association of the proteins to which they are attached (41, 48, 90). However, these microscopy-based Gag FRET assays have not been used to fully elucidate several fundamental aspects of HIV-1 Gag multimerization at the plasma membrane of cells, such as the relative contributions of the CA-CTD and NC and the effect of membrane binding on Gag-Gag interactions. In this study, we used a FRET stoichiometry method based on calibrated spectral analysis of fluorescence microscopy images (41). This algorithm determines the fractions of both donor and acceptor fluorescent protein-tagged Gag molecules participating in FRET. For cells expressing Gag molecules tagged with donor (cyan fluorescent protein [CFP]) and acceptor (yellow fluorescent protein [YFP]) molecules, this method measures the apparent FRET efficiency, which is proportional to the mole fraction of Gag constructs in complex. By measuring apparent FRET efficiencies, quantitative estimates of the mole fractions of interacting proteins can be obtained.Using this FRET-based assay, we aim to answer two questions: (i) what are the relative contributions of CA-CTD and NC domains to Gag multimerization when secondary effects via membrane binding are held constant, and (ii) what is the effect of modulating membrane binding on the ability of Gag mutants to interact with WT Gag?Our data demonstrate that the CA-CTD dimerization interface is essential for Gag multimerization at the plasma membrane, as fully disrupting the CA-CTD interaction abolishes FRET, whereas a modest level of FRET is still detected in the absence of NC. We also present evidence that the CA-CTD dimerization interface consists of two reciprocal interactions, allowing the formation of a half-interface that can still contribute to Gag multimerization. Notably, when Gag derivatives with an intact CA-CTD were coexpressed with WT Gag, either membrane binding ability or NC was required for the Gag mutants to interact with WT Gag, suggesting functional compensation between these factors.  相似文献   

4.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

5.
6.
Human immunodeficiency virus type 1 (HIV-1)-associated dementia (HAD) is a severe neurological disease that affects a subset of HIV-1-infected individuals. Increased compartmentalization has been reported between blood and cerebrospinal fluid (CSF) HIV-1 populations in subjects with HAD, but it is still not known when compartmentalization arises during the course of infection. To assess HIV-1 genetic compartmentalization early during infection, we compared HIV-1 populations in the peripheral blood and CSF in 11 primary infection subjects, with analysis of longitudinal samples over the first 18 months for a subset of subjects. We used heteroduplex tracking assays targeting the variable regions of env and single-genome amplification and sequence analysis of the full-length env gene to identify CSF-compartmentalized variants and to examine viral genotypes within the compartmentalized populations. For most subjects, HIV-1 populations were equilibrated between the blood and CSF compartments. However, compartmentalized HIV-1 populations were detected in the CSF of three primary infection subjects, and longitudinal analysis of one subject revealed that compartmentalization during primary HIV-1 infection was resolved. Clonal amplification of specific HIV-1 variants was identified in the CSF population of one primary infection subject. Our data show that compartmentalization can occur in the central nervous system (CNS) of subjects in primary HIV-1 infection in part through persistence of the putative transmitted parental variant or via viral genetic adaptation to the CNS environment. The presence of distinct HIV-1 populations in the CSF indicates that independent HIV-1 replication can occur in the CNS, even early after HIV-1 transmission.Human immunodeficiency virus type 1 (HIV-1) infection of the central nervous system (CNS) can lead to neurological disease in a subset of HIV-infected individuals and may include the development of HIV-1-associated dementia (HAD) (2, 18). HAD is characterized by severe neurological dysfunction, and affected individuals generally have impaired cognitive and motor functions. HIV-1 enters the CNS during primary infection, most likely via the migration of infected monocytes and lymphocytes across the blood-brain barrier (33, 37, 42). The main cell types in the CNS that HIV-1 can productively infect are the perivascular macrophages and microglial cells, which express low receptor densities of CD4, CCR5, and CXCR4 (7, 18, 60, 63). Previous studies have also reported that neurotropic HIV-1 variants are generally macrophage tropic (19, 20, 32, 45, 52, 61). Although cells in the CNS may be infected with HIV-1 during the course of disease, it is still unclear whether productive HIV-1 replication occurs in the CNS early during infection.Genetically compartmentalized HIV-1 variants have been detected in the brains of HAD subjects at autopsy (13, 14, 43, 48, 52) and in the cerebrospinal fluid (CSF) of HAD subjects sampled over the course of infection (26, 46, 51, 59). Extensive compartmentalization between the periphery and the CNS has been reported in subjects with HAD; however, it is not yet known when compartmentalization occurs during the course of HIV-1 infection. Primary HIV-1 infection refers to the acute and early phases of infection, during which peak plasma viremia often occurs and a viral “set point” may be reached (8, 34), within the first year after HIV exposure (64). Studies examining compartmentalization between the blood plasma and CSF during primary infection have been limited, and extensive compartmentalization has not been detected in primary infection subjects (26, 50).In this study, we examined HIV-1 genetic compartmentalization between the peripheral blood and CSF during primary HIV-1 infection. Cross-sectional and longitudinal blood plasma and CSF samples were analyzed for viral compartmentalization using the heteroduplex tracking assay (HTA) and single genome amplification (SGA). We used the HTA to differentiate between HIV-1 variants in the CSF that were either compartmentalized to the CSF or equilibrated with the peripheral blood. Previous studies have used the HTA to separate HIV-1 genetic variants in different anatomical compartments (10, 24, 27, 51) and to follow HIV-1 evolutionary variants over the course of infection (9, 25, 31, 41, 49, 50). We also conducted SGA on a subset of subjects to further examine viral genetic compartmentalization during primary infection. Here we report the detection of compartmentalized and clonally amplified HIV-1 variants in the CSF of subjects in the primary stage of HIV-1 infection. Our results suggest that minor to extensive HIV-1 genetic compartmentalization can occur between the periphery and the CNS during primary HIV-1 infection and that viral compartmentalization, as measured in the CSF, is transient in some subjects.  相似文献   

7.
HIV-1 possesses an exquisite ability to infect cells independently from their cycling status by undergoing an active phase of nuclear import through the nuclear pore. This property has been ascribed to the presence of karyophilic elements present in viral nucleoprotein complexes, such as the matrix protein (MA); Vpr; the integrase (IN); and a cis-acting structure present in the newly synthesized DNA, the DNA flap. However, their role in nuclear import remains controversial at best. In the present study, we carried out a comprehensive analysis of the role of these elements in nuclear import in a comparison between several primary cell types, including stimulated lymphocytes, macrophages, and dendritic cells. We show that despite the fact that none of these elements is absolutely required for nuclear import, disruption of the central polypurine tract-central termination sequence (cPPT-CTS) clearly affects the kinetics of viral DNA entry into the nucleus. This effect is independent of the cell cycle status of the target cells and is observed in cycling as well as in nondividing primary cells, suggesting that nuclear import of viral DNA may occur similarly under both conditions. Nonetheless, this study indicates that other components are utilized along with the cPPT-CTS for an efficient entry of viral DNA into the nucleus.Lentiviruses display an exquisite ability to infect dividing and nondividing cells alike that is unequalled among Retroviridae. This property is thought to be due to the particular behavior or composition of the viral nucleoprotein complexes (NPCs) that are liberated into the cytoplasm of target cells upon virus-to-cell membrane fusion and that allow lentiviruses to traverse an intact nuclear membrane (17, 28, 29, 39, 52, 55, 67, 79). In the case of the human immunodeficiency type I virus (HIV-1), several studies over the years identified viral components of such structures with intrinsic karyophilic properties and thus perfect candidates for mediation of the passage of viral DNA (vDNA) through the nuclear pore: the matrix protein (MA); Vpr; the integrase (IN); and a three-stranded DNA flap, a structure present in neo-synthesized viral DNA, specified by the central polypurine tract-central termination sequence (cPPT-CTS). It is clear that these elements may mediate nuclear import directly or via the recruitment of the host''s proteins, and indeed, several cellular proteins have been found to influence HIV-1 infection during nuclear import, like the karyopherin α2 Rch1 (38); importin 7 (3, 30, 93); the transportin SR-2 (13, 20); or the nucleoporins Nup98 (27), Nup358/RANBP2, and Nup153 (13, 56).More recently, the capsid protein (CA), the main structural component of viral nucleoprotein complexes at least upon their cytoplasmic entry, has also been suggested to be involved in nuclear import or in postnuclear entry steps (14, 25, 74, 90, 92). Whether this is due to a role for CA in the shaping of viral nucleoprotein complexes or to a direct interaction between CA and proteins involved in nuclear import remains at present unknown.Despite a large number of reports, no single viral or cellular element has been described as absolutely necessary or sufficient to mediate lentiviral nuclear import, and important controversies as to the experimental evidences linking these elements to this step exist. For example, MA was among the first viral protein of HIV-1 described to be involved in nuclear import, and 2 transferable nuclear localization signals (NLSs) have been described to occur at its N and C termini (40). However, despite the fact that early studies indicated that the mutation of these NLSs perturbed HIV-1 nuclear import and infection specifically in nondividing cells, such as macrophages (86), these findings failed to be confirmed in more-recent studies (23, 33, 34, 57, 65, 75).Similarly, Vpr has been implicated by several studies of the nuclear import of HIV-1 DNA (1, 10, 21, 43, 45, 47, 64, 69, 72, 73, 85). Vpr does not possess classical NLSs, yet it displays a transferable nucleophilic activity when fused to heterologous proteins (49-51, 53, 77, 81) and has been shown to line onto the nuclear envelope (32, 36, 47, 51, 58), where it can truly facilitate the passage of the viral genome into the nucleus. However, the role of Vpr in this step remains controversial, as in some instances Vpr is not even required for viral replication in nondividing cells (1, 59).Conflicting results concerning the role of IN during HIV-1 nuclear import also exist. Indeed, several transferable NLSs have been described to occur in the catalytic core and the C-terminal DNA binding domains of IN, but for some of these, initial reports of nuclear entry defects (2, 9, 22, 46, 71) were later shown to result from defects at steps other than nuclear import (60, 62, 70, 83). These reports do not exclude a role for the remaining NLSs in IN during nuclear import, and they do not exclude the possibility that IN may mediate this step by associating with components of the cellular nuclear import machinery, such as importin alpha and beta (41), importin 7 (3, 30, 93, 98), and, more recently, transportin-SR2 (20).The central DNA flap, a structure present in lentiviruses and in at least 1 yeast retroelement (44), but not in other orthoretroviruses, has also been involved in the nuclear import of viral DNA (4, 6, 7, 31, 78, 84, 95, 96), and more recently, it has been proposed to provide a signal for viral nucleoprotein complexes uncoating in the proximity of the nuclear pore, with the consequence of providing a signal for import (8). However, various studies showed an absence or weakness of nuclear entry defects in viruses devoid of the DNA flap (24, 26, 44, 61).Overall, the importance of viral factors in HIV-1 nuclear import is still unclear. The discrepancies concerning the role of MA, IN, Vpr, and cPPT-CTS in HIV-1 nuclear import could in part be explained by their possible redundancy. To date, only one comprehensive study analyzed the role of these four viral potentially karyophilic elements together (91). This study showed that an HIV-1 chimera where these elements were either deleted or replaced by their murine leukemia virus (MLV) counterparts was, in spite of an important infectivity defect, still able to infect cycling and cell cycle-arrested cell lines to similar efficiencies. If this result indicated that the examined viral elements of HIV-1 were dispensable for the cell cycle independence of HIV, as infections proceeded equally in cycling and arrested cells, they did not prove that they were not required in nuclear import, because chimeras displayed a severe infectivity defect that precluded their comparison with the wild type (WT).Nuclear import and cell cycle independence may not be as simply linked as previously thought. On the one hand, there has been no formal demonstration that the passage through the nuclear pore, and thus nuclear import, is restricted to nondividing cells, and for what we know, this passage may be an obligatory step in HIV infection in all cells, irrespective of their cycling status. In support of this possibility, certain mutations in viral elements of HIV affect nuclear import in dividing as well as in nondividing cells (4, 6, 7, 31, 84, 95). On the other hand, cell cycle-independent infection may be a complex phenomenon that is made possible not only by the ability of viral DNA to traverse the nuclear membrane but also by its ability to cope with pre- and postnuclear entry events, as suggested by the phenotypes of certain CA mutants (74, 92).Given that the cellular environment plays an important role during the early steps of viral infection, we chose to analyze the role of the four karyophilic viral elements of HIV-1 during infection either alone or combined in a wide comparison between cells highly susceptible to infection and more-restrictive primary cell targets of HIV-1 in vivo, such as primary blood lymphocytes (PBLs), monocyte-derived macrophages (MDM), and dendritic cells (DCs).In this study, we show that an HIV-1-derived virus in which the 2 NLSs of MA are mutated and the IN, Vpr, and cPPT-CTS elements are removed displays no detectable nuclear import defect in HeLa cells independently of their cycling status. However, this mutant virus is partially impaired for nuclear entry in primary cells and more specifically in DCs and PBLs. We found that this partial defect is specified by the cPPT-CTS, while the 3 remaining elements seem to play no role in nuclear import. Thus, our study indicates that the central DNA flap specifies the most important role among the viral elements involved thus far in nuclear import. However, it also clearly indicates that the role played by the central DNA flap is not absolute and that its importance varies depending on the cell type, independently from the dividing status of the cell.  相似文献   

8.
9.
Binding to the primary receptor CD4 induces conformational changes in the human immunodeficiency virus type 1 (HIV-1) gp120 envelope glycoprotein that allow binding to the coreceptor (CCR5 or CXCR4) and ultimately trigger viral membrane-cell membrane fusion mediated by the gp41 transmembrane envelope glycoprotein. Here we report the derivation of an HIV-1 gp120 variant, H66N, that confers envelope glycoprotein resistance to temperature extremes. The H66N change decreases the spontaneous sampling of the CD4-bound conformation by the HIV-1 envelope glycoproteins, thus diminishing CD4-independent infection. The H66N change also stabilizes the HIV-1 envelope glycoprotein complex once the CD4-bound state is achieved, decreasing the probability of CD4-induced inactivation and revealing the enhancing effects of soluble CD4 binding on HIV-1 infection. In the CD4-bound conformation, the highly conserved histidine 66 is located between the receptor-binding and gp41-interactive surfaces of gp120. Thus, a single amino acid change in this strategically positioned gp120 inner domain residue influences the propensity of the HIV-1 envelope glycoproteins to negotiate conformational transitions to and from the CD4-bound state.Human immunodeficiency virus type 1 (HIV-1), the cause of AIDS (6, 29, 66), infects target cells by direct fusion of the viral and target cell membranes. The viral fusion complex is composed of gp120 and gp41 envelope glycoproteins, which are organized into trimeric spikes on the surface of the virus (10, 51, 89). Membrane fusion is initiated by direct binding of gp120 to the CD4 receptor on target cells (17, 41, 53). CD4 binding creates a second binding site on gp120 for the chemokine receptors CCR5 and CXCR4, which serve as coreceptors (3, 12, 19, 23, 25). Coreceptor binding is thought to lead to further conformational changes in the HIV-1 envelope glycoproteins that facilitate the fusion of viral and cell membranes. The formation of an energetically stable six-helix bundle by the gp41 ectodomain contributes to the membrane fusion event (9, 10, 79, 89, 90).The energy required for viral membrane-cell membrane fusion derives from the sequential transitions that the HIV-1 envelope glycoproteins undergo, from the high-energy unliganded state to the low-energy six-helix bundle. The graded transitions down this energetic slope are initially triggered by CD4 binding (17). The interaction of HIV-1 gp120 with CD4 is accompanied by an unusually large change in entropy, which is thought to indicate the introduction of order into the conformationally flexible unliganded gp120 glycoprotein (61). In the CD4-bound state, gp120 is capable of binding CCR5 with high affinity; moreover, CD4 binding alters the quaternary structure of the envelope glycoprotein complex, resulting in the exposure of gp41 ectodomain segments (27, 45, 77, 92). The stability of the intermediate state induced by CD4 binding depends upon several variables, including the virus (HIV-1 versus HIV-2/simian immunodeficiency virus [SIV]), the temperature, and the nature of the CD4 ligand (CD4 on a target cell membrane versus soluble forms of CD4 [sCD4]) (30, 73). For HIV-1 exposed to sCD4, if CCR5 binding occurs within a given period of time, progression along the entry pathway continues. If CCR5 binding is impeded or delayed, the CD4-bound envelope glycoprotein complex decays into inactive states (30). In extreme cases, the binding of sCD4 to the HIV-1 envelope glycoproteins induces the shedding of gp120 from the envelope glycoprotein trimer (31, 56, 58). Thus, sCD4 generally inhibits HIV-1 infection by triggering inactivation events, in addition to competing with CD4 anchored in the target cell membrane (63).HIV-1 isolates vary in sensitivity to sCD4, due in some cases to a low affinity of the envelope glycoprotein trimer for CD4 and in other cases to differences in propensity to undergo inactivating conformational transitions following CD4 binding (30). HIV-1 isolates that have been passaged extensively in T-cell lines (the tissue culture laboratory-adapted [TCLA] isolates) exhibit lower requirements for CD4 than primary HIV-1 isolates (16, 63, 82). TCLA viruses bind sCD4 efficiently and are generally sensitive to neutralization compared with primary HIV-1 isolates. Differences in sCD4 sensitivity between primary and TCLA HIV-1 strains have been mapped to the major variable loops (V1/V2 and V3) of the gp120 glycoprotein (34, 42, 62, 81). Sensitivity to sCD4 has been shown to be independent of envelope glycoprotein spike density or the intrinsic stability of the envelope glycoprotein complex (30, 35).In general, HIV-1 isolates are more sensitive to sCD4 neutralization than HIV-2 or SIV isolates (4, 14, 73). The relative resistance of SIV to sCD4 neutralization can in some cases be explained by a reduced affinity of the envelope glycoprotein trimer for sCD4 (57); however, at least some SIV isolates exhibit sCD4-induced activation of entry into CD4-negative, CCR5-expressing target cells that lasts for several hours after exposure to sCD4 (73). Thus, for some primate immunodeficiency virus envelope glycoproteins, activated intermediates in the CD4-bound conformation can be quite stable.The HIV-1 envelope glycoprotein elements important for receptor binding, subunit interaction, and membrane fusion are well conserved among different viral strains (71, 91). Thus, these elements represent potential targets for inhibitors of HIV-1 entry. Understanding the structure and longevity of the envelope glycoprotein intermediates along the virus entry pathway is relevant to attempts at inhibition. For example, peptides that target the heptad repeat 1 region of gp41 exhibit major differences in potency against HIV-1 strains related to efficiency of chemokine receptor binding (20, 21), which is thought to promote the conformational transition to the next step in the virus entry cascade. The determinants of the duration of exposure of targetable HIV-1 envelope glycoprotein elements during the entry process are undefined.To study envelope glycoprotein determinants of the movement among the distinct conformational states along the HIV-1 entry pathway, we attempted to generate HIV-1 variants that exhibit improved stability. Historically, labile viral elements have been stabilized by selecting virus to replicate under conditions, such as high temperature, that typically weaken protein-protein interactions (38, 39, 76, 102). Thus, we subjected HIV-1 to repeated incubations at temperatures between 42°C and 56°C, followed by expansion and analysis of the remaining replication-competent virus fraction. In this manner, we identified an envelope glycoprotein variant, H66N, in which histidine 66 in the gp120 N-terminal segment was altered to asparagine. The resistance of HIV-1 bearing the H66N envelope glycoproteins to changes in temperature has been reported elsewhere (37). Here, we examine the effect of the H66N change on the ability of the HIV-1 envelope glycoproteins to negotiate conformational transitions, either spontaneously or in the presence of sCD4. The H66N phenotype was studied in the context of both CD4-dependent and CD4-independent HIV-1 variants.  相似文献   

10.
To facilitate the release of infectious progeny virions, human immunodeficiency virus type 1 (HIV-1) exploits the Endosomal Sorting Complex Required for Transport (ESCRT) pathway by engaging Tsg101 and ALIX through late assembly (L) domains in the C-terminal p6 domain of Gag. However, the L domains in p6 are known to be dispensable for efficient particle production by certain HIV-1 Gag constructs that have the nucleocapsid (NC) domain replaced by a foreign dimerization domain to substitute for the assembly function of NC. We now show that one such L domain-independent HIV-1 Gag construct (termed ZWT) that has NC-p1-p6 replaced by a leucine zipper domain is resistant to dominant-negative inhibitors of the ESCRT pathway that block HIV-1 particle production. However, ZWT became dependent on the presence of an L domain when NC-p1-p6 was restored to its C terminus. Furthermore, when the NC domain was replaced by a leucine zipper, the p1-p6 region, but not p6 alone, conferred sensitivity to inhibition of the ESCRT pathway. In an authentic HIV-1 Gag context, the effect of an inhibitor of the ESCRT pathway on particle production could be alleviated by deleting a portion of the NC domain together with p1. Together, these results indicate that the ESCRT pathway dependence of HIV-1 budding is determined, at least in part, by the NC-p1 region of Gag.Human immunodeficiency virus type 1 (HIV-1) and other retroviruses hijack the cellular Endosomal Sorting Complex Required for Transport (ESCRT) pathway to promote the detachment of virions from the cell surface and from each other (3, 21, 42, 44, 47). The ESCRT pathway was initially identified based on its requirement for the sorting of ubiquitinated cargo into multivesicular bodies (MVB) (50, 51). During MVB biogenesis, the ESCRT pathway drives the membrane deformation and fission events required for the inward vesiculation of the limiting membrane of this organelle (26, 29, 50, 51). More recently, it emerged that the ESCRT pathway is also essential for the normal abscission of daughter cells during the final stage of cell division (10, 43). Most of the components of the ESCRT pathway are involved in the formation of four heteromeric protein complexes termed ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III. Additional components include ALIX, which interacts both with ESCRT-I and ESCRT-III, and the AAA ATPase Vps4, which mediates the disassembly of ESCRT-III (29, 42).The deformation and scission of endocytic membranes is thought to be mediated by ESCRT-III, which, together with Vps4, constitutes the most conserved element of the pathway (23, 26, 42). Indeed, it was recently shown that purified yeast ESCRT-III induces membrane deformation (52), and in another study three subunits of yeast ESCRT-III were sufficient to promote the formation of intralumenal vesicles in an in vitro assay (61). In mammals, ESCRT-III is formed by the charged MVB proteins (CHMPs), which are structurally related and tightly regulated through autoinhibition (2, 33, 46, 53, 62). The removal of an inhibitory C-terminal domain induces polymerization and association with endosomal membranes and converts CHMPs into potent inhibitors of retroviral budding (34, 46, 53, 60, 62). Alternatively, CHMPs can be converted into strong inhibitors of the ESCRT pathway and of HIV-1 budding through the addition of a bulky tag such as green fluorescent protein (GFP) or red fluorescent protein (RFP) (27, 36, 39, 54). Retroviral budding in general is also strongly inhibited by catalytically inactive Vps4 (22, 41, 55), or upon Vsp4B depletion (31), confirming the crucial role of ESCRT-III.Retroviruses engage the ESCRT pathway through the activity of so-called late assembly (L) domains in Gag. In the case of HIV-1, the primary L domain maps to a conserved PTAP motif in the C-terminal p6 domain of Gag (24, 28) and interacts with the ESCRT-I component Tsg101 (15, 22, 40, 58). HIV-1 p6 also harbors an auxiliary L domain of the LYPxnL type, which interacts with the V domain of ALIX (20, 35, 39, 54, 59, 63). Interestingly, Tsg101 binding site mutants of HIV-1 can be fully rescued through the overexpression of ALIX, and this rescue depends on the ALIX binding site in p6 (20, 56). In contrast, the overexpression of a specific splice variant of the ubiquitin ligase Nedd4-2 has been shown to rescue the release and infectivity of HIV-1 mutants lacking all known L domains in p6 (12, 57). Nedd4 family ubiquitin ligases had previously been implicated in the function of PPxY-type L domains, which also depend on an intact ESCRT pathway for function (4, 32, 38). However, HIV-1 Gag lacks PPxY motifs, and the WW domains of Nedd4-2, which mediate its interaction with PPxY motifs, are dispensable for the rescue of HIV-1 L domain mutants (57).ALIX also interacts with the nucleocapsid (NC) region of HIV-1 Gag (18, 49), which is located upstream of p6 and the p1 spacer peptide. ALIX binds HIV-1 NC via its Bro1 domain, and the capacity to interact with NC and to stimulate the release of a minimal HIV-1 Gag construct is shared among widely divergent Bro1 domain proteins (48). Based on these findings and the observation that certain mutations in NC cause a phenotype that resembles that of L domain mutants, it has been proposed that NC cooperates with p6 to recruit the machinery required for normal HIV-1 budding (18, 49).NC also plays a role in Gag polyprotein multimerization, and this function of NC depends on its RNA-binding activity (5-8). It has been proposed that the role of the NC-nucleic acid interaction during assembly is to promote the formation of Gag dimers (37), and HIV-1 assembly in the absence of NC can indeed be efficiently rescued by leucine zipper dimerization domains (65). Surprisingly, in this setting the L domains in p6 also became dispensable, since particle production remained efficient even when the entire NC-p1-p6 region of HIV-1 Gag was replaced by a leucine zipper (1, 65). These findings raised the possibility that the reliance of wild-type (WT) HIV-1 Gag on a functional ESCRT pathway is, at least in part, specified by NC-p1-p6. However, it also remained possible that the chimeric Gag constructs engaged the ESCRT pathway in an alternative manner.In the present report, we provide evidence supporting the first of those two possibilities. Particle production became independent of ESCRT when the entire NC-p1-p6 region was replaced by a leucine zipper, and reversion to ESCRT dependence was shown to occur as a result of restoration of p1-p6 but not of p6 alone. Furthermore, although the deletion of p1 alone had little effect in an authentic HIV-1 Gag context, the additional removal of a portion of NC improved particle production in the presence of an inhibitor of the ESCRT pathway. Together, these data imply that the NC-p1 region plays an important role in the ESCRT-dependence of HIV-1 particle production.  相似文献   

11.
12.
The human immunodeficiency virus type 1 (HIV-1) matrix (MA) protein targets HIV-1 precursor Gag (PrGag) proteins to assembly sites at plasma membrane (PM) sites that are enriched in cholesterol and phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2]. MA is myristoylated, which enhances membrane binding, and specifically binds PI(4,5)P2 through headgroup and 2′ acyl chain contacts. MA also binds nucleic acids, although the significance of this association with regard to the viral life cycle is unclear. We have devised a novel MA binding assay and used it to examine MA interactions with membranes and nucleic acids. Our results indicate that cholesterol increases the selectivity of MA for PI(4,5)P2-containing membranes, that PI(4,5)P2 binding tolerates 2′ acyl chain variation, and that the MA myristate enhances membrane binding efficiency but not selectivity. We also observed that soluble PI(4,5)P2 analogues do not compete effectively with PI(4,5)P2-containing liposomes for MA binding but surprisingly do increase nonspecific binding to liposomes. Finally, we have demonstrated that PI(4,5)P2-containing liposomes successfully outcompete nucleic acids for MA binding, whereas other liposomes do not. These results support a model in which RNA binding protects MA from associating with inappropriate cellular membranes prior to PrGag delivery to PM assembly sites.The matrix (MA) domain of the human immunodeficiency virus type 1 (HIV-1) precursor Gag (PrGag) protein serves several functions in the viral replication cycle. One essential function is to target PrGag proteins to their assembly sites at the plasma membranes (PMs) of infected cells (4, 5, 11, 16, 25, 29, 30, 33, 35, 39, 43-45, 47, 50, 54, 56, 57). A second function is the recruitment of the viral surface/transmembrane (SU/TM; also referred to as gp120/gp41) envelope (Env) protein complex into virions (14, 15, 18, 19, 27, 51-53). In addition to these activities, numerous reports have attributed nucleic acid binding properties to retroviral MAs (24, 38, 47), and with some viruses MA appears to serve in an encapsidation capacity (24). While no encapsidation role has been assigned for HIV-1 MA, experiments have shown that MA can substitute for the HIV-1 nucleocapsid (NC) protein assembly function (38) under some circumstances, presumably by virtue of its facility to concentrate PrGag proteins by binding them to RNAs (38).A number of structural studies have been conducted on HIV-1 MA (1, 22, 41, 42, 49). The protein is N terminally myristoylated and composed of six α helices, capped by a three-strand β sheet (7, 22, 41, 42, 49). The protein trimerizes in solution and in crystals (22, 28, 49) and recently has been shown to organize as hexamers of trimers on lipid membranes (1). The membrane binding face of HIV-1 MA is basic, fostering its ability to associate with negatively charged phospholipid headgroups (1, 22, 30, 41, 42, 49). The importance of such an interaction has been underscored in molecular genetic experiments which demonstrated that depletion of PM phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2] reduced the assembly efficiency of HIV-1 (9, 36). Consistent with these observations, HIV-1 MA preferentially binds to soluble PI(4,5)P2 mimics through contacts with the headgroup and 2′ acyl chain, and binding promotes exposure of the MA myristate group and protein oligomerization (17, 21, 40-43, 46). However, PI(4,5)P2 is not the only lipid to demonstrate an association with HIV-1. In particular, HIV-1 appears to assemble at cholesterol-rich PM sites, cholesterol is highly enriched in HIV-1 virions, and cholesterol depletion reduces viral infectivity (2, 6, 8, 20, 23, 26, 31, 34, 37). The HIV-1 lipidome shows additional differences from the PM lipids of infected cells (2, 5, 8), suggesting that other lipids could affect PrGag-membrane binding or virus assembly site selection.To gain a better understanding of the functions and interactions of HIV-1 MA, we have examined the liposome and nucleic acid binding properties of purified myristoylated MA. Using liposome flotation assays and a novel liposome bead binding assay, we have demonstrated that the PI(4,5)P2 binding specificity of MA is enhanced by cholesterol, that protein myristoylation increases membrane binding efficiency but not specificity, and that 2′ acyl chain variation is compatible with PI(4,5)P2 binding. We also examined whether soluble PI(4,5)P2 mimics could compete with liposomes for MA binding. Surprisingly, we found that soluble mimics not only failed to compete with PI(4,5)P2 liposomes but also increased MA binding to membranes that do not contain acidic phospholipids. Finally, we have observed that while MA does bind nucleic acids, nucleic acid binding is outcompeted by PI(4,5)P2-containing liposomes. Our results suggest models for PrGag-membrane and RNA association and the HIV-1 assembly pathway.  相似文献   

13.
14.
15.
16.
During untreated human immunodeficiency virus type 1 (HIV-1) infection, virus-specific CD8+ T cells partially control HIV replication in peripheral lymphoid tissues, but host mechanisms of HIV control in the central nervous system (CNS) are incompletely understood. We characterized HIV-specific CD8+ T cells in cerebrospinal fluid (CSF) and peripheral blood among seven HIV-positive antiretroviral therapy-naïve subjects. All had grossly normal brain magnetic resonance imaging and spectroscopy and normal neuropsychometric testing. Frequencies of epitope-specific CD8+ T cells by direct tetramer staining were on average 2.4-fold higher in CSF than in blood (P = 0.0004), while HIV RNA concentrations were lower. Cells from CSF were readily expanded ex vivo and responded to a broader range of HIV-specific human leukocyte antigen class I restricted optimal peptides than did expanded cells from blood. HIV-specific CD8+ T cells, in contrast to total CD8+ T cells, in CSF and blood were at comparable maturation states, as assessed by CD45RO and CCR7 staining. The strong relationship between higher T-cell frequencies and lower levels of viral antigen in CSF could be the result of increased migration to and/or preferential expansion of HIV-specific T cells within the CNS. This suggests an important role for HIV-specific CD8+ T cells in control of intrathecal viral replication.Human immunodeficiency virus type 1 (HIV-1) invades the central nervous system (CNS) early during primary infection (21, 30, 35), and proviral DNA persists in the brain throughout the course of HIV-1 disease (7, 25, 29, 47, 77, 83). Limited data from human and nonhuman primate studies suggest that little or no viral replication occurs in the brain during chronic, asymptomatic infection, based on the absence of demonstrable viral RNA or proteins (8, 85). In contrast, cognitive impairment affects approximately 40% of patients who progress to advanced AIDS without highly active antiretroviral therapy (21, 30, 35, 65). During HIV-associated dementia, there is active HIV-1 replication in the brain (23, 52, 61, 81), and viral sequence differences between cerebrospinal fluid (CSF) and peripheral tissues suggest distinct anatomic compartments of replication (18, 19, 22, 53, 75, 76, 78). Host mechanisms that control viral replication in the CNS during chronic, asymptomatic HIV-1 infection are incompletely understood.Anti-HIV CD8+ T cells are present in blood and peripheral tissues throughout the course of chronic HIV-1 infection (2, 14). Multiple lines of evidence support a critical role for these cells in controlling HIV-1 replication. During acute HIV-1 infection, the appearance of CD8+ T-cell responses correlates temporally with a decline in viremia (11, 43), and a greater proliferative capacity of peripheral blood HIV-specific CD8+ T cells correlates with better control of viremia (36, 54). In addition, the presence of certain major histocompatibility complex class I human leukocyte antigen (HLA) alleles, notably HLA-B*57, predicts slower progression to AIDS and death during chronic, untreated HIV-1 infection (55, 62). Finally, in the simian immunodeficiency virus (SIV) model, macaques depleted of CD8+ T cells experience increased viremia and rapid disease progression (39, 51, 67).Little is known regarding the role of intrathecal anti-HIV CD8+ T cells in HIV neuropathogenesis. Nonhuman primate studies have identified SIV-specific CD8+ T cells in the CNS early after infection (16, 80). Increased infiltration of SIV antigen-specific CD8+ T cells and cytotoxic T lymphocytes has been detected only in CSF of slow progressors without neurological symptoms (72). In chronically infected macaques with little or no SIV replication in the brain, the frequency of HIV-specific T cells was higher in CSF than in peripheral blood but did not correlate with the level of plasma viremia or CD4+ T-cell counts (56). Although intrathecal anti-HIV CD8+ T cells may help control viral replication, a detrimental role in the neuropathogenesis of HIV-1 has also been postulated (38). Immune responses contribute to neuropathogenesis in models of other infectious diseases, and during other viral infections cytotoxic T lymphocytes can worsen disease through direct cytotoxicity or release of inflammatory cytokines such as gamma interferon (IFN-γ) (3, 17, 31, 37, 42, 44, 71).We tested the hypothesis that quantitative and/or qualitative differences in HIV-specific CD8+ T-cell responses are present in CSF compared to blood during chronic, untreated HIV-1 infection. We characterized HIV-specific CD8+ T-cell responses in CSF among seven antiretroviral therapy-naïve adults with chronic HIV-1 infection, relatively high peripheral blood CD4+ T-cell counts, and low plasma HIV-1 RNA concentrations. We show that among these HIV-positive individuals with no neurological symptoms and with little or no HIV-1 RNA in CSF, frequencies of HIV-specific T cells are significantly higher in CSF than in blood. These CSF cells are at a state of differentiation similar to that of T cells in blood and are functionally competent for expansion and IFN-γ production. The higher frequency of functional HIV-specific CD8+ T cells in CSF, in the context of low or undetectable virus in CSF, suggests that these cells play a role in the control of intrathecal viral replication.  相似文献   

17.
Tetherin (CD317/BST-2), an interferon-induced membrane protein, restricts the release of nascent retroviral particles from infected cell surfaces. While human immunodeficiency virus type 1 (HIV-1) encodes the accessory gene vpu to overcome the action of tetherin, the lineage of primate lentiviruses that gave rise to HIV-2 does not. It has been previously reported that the HIV-2 envelope glycoprotein has a Vpu-like function in promoting virus release. Here we demonstrate that the HIV-2 Rod envelope glycoprotein (HIV-2 Rod Env) is a tetherin antagonist. Expression of HIV-2 Rod Env, but not that of HIV-1 or the closely related simian immunodeficiency virus (SIV) SIVmac1A11, counteracts tetherin-mediated restriction of Vpu-defective HIV-1 in a cell-type-specific manner. This correlates with the ability of the HIV-2 Rod Env to mediate cell surface downregulation of tetherin. Antagonism requires an endocytic motif conserved across HIV/SIV lineages in the gp41 cytoplasmic tail, but specificity for tetherin is governed by extracellular determinants in the mature Env protein. Coimmunoprecipitation studies suggest an interaction between HIV-2 Rod Env and tetherin, but unlike studies with Vpu, we found no evidence of tetherin degradation. In the presence of HIV-2 Rod Env, tetherin localization is restricted to the trans-Golgi network, suggesting Env-mediated effects on tetherin trafficking sequester it from virus assembly sites on the plasma membrane. Finally, we recapitulated these observations in HIV-2-infected CD4+ T-cell lines, demonstrating that tetherin antagonism and sequestration occur at physiological levels of Env expression during virus replication.Various stages of the replication cycle of primate lentiviruses can be targeted by host antiviral restriction factors (reviewed in reference 49). In addition to the well-characterized antiviral effects of members of the APOBEC3 family of cytidine deaminases, particularly APOBEC3G and -3F, and species-specific variants of tripartite motif family 5α, the release of nascent retroviral particles has recently been shown to be a target for a novel restriction factor, tetherin (CD317/bone marrow stromal cell antigen 2 [BST-2]) (31, 46). Tetherin is an interferon-inducible gene that was originally shown to impart a restriction on the release of mutants of human immunodeficiency virus type 1 (HIV-1) that lack a vpu gene (31, 46). In tetherin-positive cells, mature Vpu-defective HIV-1 particles are retained on the cell surface, linked to the plasma membrane (PM) and each other via protease-sensitive tethers, and can be subsequently endocytosed and accumulate in late endosomes (30, 31). Tetherin is not HIV specific and restricts the release of virus-like particles derived from all retroviruses tested (18), as well as those of filoviruses and arenaviruses (18, 19, 39).Tetherin is a small (181-amino-acid) type II membrane protein with an unusual topology that exists mainly as a disulfide-linked dimer (34). It consists of an N-terminal cytoplasmic tail, a transmembrane anchor, an extracellular domain that includes three cysteine residues important for dimerization, a putative coiled-coil, and finally a glycophosphatidyinosityl-linked lipid anchor (22) that is essential for restriction (31). Tetherin localizes to retroviral assembly sites on the PM (18, 31), and this unusual structure is highly suggestive that tetherin restricts virion release by incorporation into the viral membrane and cross-linking virions to cells. Such a mechanism would make tetherin a powerful antiviral effector that can target an obligate part of most, if not all, enveloped virus assembly strategies. Moreover, since tetherin restriction has no specific requirement for virus protein sequences, to avoid its action, mammalian viruses have evolved to encode several distinct countermeasures that specifically inhibit tetherin''s antiviral function.The Vpu accessory protein antagonizes tetherin-mediated restriction of HIV-1 (31, 46). In the presence of Vpu, tetherin is downregulated from the cell surface (2, 46) and is targeted for degradation (10, 13, 14), although whether these processes are required for antagonism of tetherin function is unclear (27). HIV-1 Vpu displays a distinct species specificity in that it is unable to target tetherin orthologues from rhesus macaques or African green monkeys (14, 25). This differential sensitivity maps to the tetherin transmembrane domain, particularly residues that are predicted to have been under high positive selection pressure during primate evolution (14, 16, 25). This suggests that tetherin evolution may have been driven in part by viral countermeasures like Vpu. Vpu, however, is only encoded by HIV-1 and its direct simian immunodeficiency virus (SIV) lineage precursors. The majority of SIVs, including the SIVsm, the progenitor of both HIV-2 and SIVmac, do not encode a Vpu protein (21). In some of these SIVs, tetherin antagonism was recently shown to map to the nef gene (16, 51). SIV Nef proteins, however, are generally ineffective against human tetherin because they target a (G/D)DIWK motif that was deleted from the human tetherin cytoplasmic tail sometime after the divergence of humans and chimpanzees (51). This raises the question of how HIV-2 is able to overcome human tetherin, as recent data show chronically HIV-2-infected CEM T cells have reduced tetherin levels on their surface (10).Interestingly, it has long been known that the envelope glycoprotein of certain HIV-2 isolates can stimulate the release of Vpu-defective HIV-1 virions from cells we now know to be tetherin positive (5, 6, 43). HIV and SIV Envs form trimeric spikes of dimers of the surface subunit (SU-gp105 in HIV-2/SIVmac and gp120 in HIV-1) that bind CD4 and the chemokine coreceptor and gp41 (the transmembrane [TM] subunit that facilitates fusion with and entry into the target cell). Envelope precursors (gp140 or gp160) are synthesized in the endoplasmic reticulum, where they become glycosylated and are exported to the surface via the secretory pathway (8). During transit through the Golgi apparatus and possibly in endosomal compartments, the immature precursors are cleaved by furin-like proteases to form mature spikes (15, 29). Multiple endocytosis motifs in the gp41 cytoplasmic tail lead to only minor quantities of Env being exposed at the cell surface at any given time (7, 40). Recent data demonstrated that the conserved GYxxθ motif, a binding site for the clathrin adaptor protein AP-2 (3), in the membrane-proximal region of HIV-2 gp41 is required to promote Vpu-defective HIV-1 release from HeLa cells (1, 32). Based on experiments with HIV-1/HIV-2 chimeric envelopes, an additional requirement in the extracellular component was suggested (1). In this study we set out to examine the Vpu-like activity of HIV-2 envelope in light of the discovery of tetherin. We demonstrate that the HIV-2 Env is a tetherin antagonist, and we provide mechanistic insight into the basis of this antagonism.  相似文献   

18.
19.
Direct cell-to-cell spread of human immunodeficiency virus type 1 (HIV-1) between T cells at the virological synapse (VS) is an efficient mechanism of viral dissemination. Tetherin (BST-2/CD317) is an interferon-induced, antiretroviral restriction factor that inhibits nascent cell-free particle release. The HIV-1 Vpu protein antagonizes tetherin activity; however, whether tetherin also restricts cell-cell spread is unclear. We performed quantitative cell-to-cell transfer analysis of wild-type (WT) or Vpu-defective HIV-1 in Jurkat and primary CD4+ T cells, both of which express endogenous levels of tetherin. We found that Vpu-defective HIV-1 appeared to disseminate more efficiently by cell-to-cell contact between Jurkat cells under conditions where tetherin restricted cell-free virion release. In T cells infected with Vpu-defective HIV-1, tetherin was enriched at the VS, and VS formation was increased compared to the WT, correlating with an accumulation of virus envelope proteins on the cell surface. Increasing tetherin expression with type I interferon had only minor effects on cell-to-cell transmission. Furthermore, small interfering RNA (siRNA)-mediated depletion of tetherin decreased VS formation and cell-to-cell transmission of both Vpu-defective and WT HIV-1. Taken together, these data demonstrate that tetherin does not restrict VS-mediated T cell-to-T cell transfer of Vpu-defective HIV-1 and suggest that under some circumstances tetherin might promote cell-to-cell transfer, either by mediating the accumulation of virions on the cell surface or by regulating integrity of the VS. If so, inhibition of tetherin activity by Vpu may balance requirements for efficient cell-free virion production and cell-to-cell transfer of HIV-1 in the face of antiviral immune responses.Human immunodeficiency virus type 1 can disseminate between and within hosts by cell-free infection or by direct cell-cell spread. Cell-cell spread of HIV-1 between CD4+ T cells is an efficient means of viral dissemination (65) and has been estimated to be several orders of magnitude more rapid than cell-free virus infection (6, 8, 41, 64, 74). Cell-cell transmission of HIV-1 takes place at the virological synapse (VS), a multimolecular structure that forms at the interface between an HIV-1-infected T cell and an uninfected target T cell during intercellular contact (27). Related structures that facilitate cell-cell spread of HIV-1 between dendritic cells and T cells (42) and between macrophages and T cells (16, 17) and for cell-cell spread of the related retrovirus human T-cell leukemia virus type 1 (HTLV-1) (24) have also been described. Moreover, more long-range cell-cell transfer can occur via cellular projections, including filopodia (71) and membrane nanotubes (75). The VS is initiated by binding of the HIV-1 envelope glycoprotein (Env), which is expressed on the surfaces of infected T cells, to HIV-1 entry receptors (CD4 and either CXCR4 or CCR5) present on the target cell membrane (6, 22, 27, 41, 61, 73). Interactions between LFA-1 and ICAM-1 and ICAM-3 further stabilize the conjugate interface and, together with Env receptor binding, help trigger the recruitment of viral proteins, CD4/coreceptor, and integrins to the contact site (27, 28, 61). The enrichment of viral and cellular proteins at the VS is an active process, dependent on cytoskeletal remodeling, and in the infected T cell both the actin and tubulin network regulate polarization of HIV-1 proteins at the cell-cell interface, thus directing HIV-1 assembly and egress toward the engaged target cell (27, 29). Virus is transferred by budding into the synaptic cleft, and virions subsequently attach to the target cell membrane to mediate entry, either by fusion at the plasma membrane or possibly following endocytic uptake (2, 22). In this way, the VS promotes more rapid infection kinetics and may enhance HIV-1 pathogenesis in vivo.Cells have evolved a number of barriers to resist invading microorganisms. One mechanism that appears to be particularly important in counteracting HIV-1 infection is a group of interferon-inducible, innate restriction factors that includes TRIM5α, APOBEC3G, and tetherin (38, 49, 69, 79). Tetherin (BST-2/CD317) is a host protein expressed by many cell types, including CD4+ T cells, that acts at a late stage of the HIV-1 life cycle to trap (or “tether”) mature virions at the plasma membranes of virus-producing cells, thereby inhibiting cell-free virus release (49, 56, 81). This antiviral activity of tetherin is not restricted to HIV-1, and tetherin can also inhibit the release of other enveloped viruses from infected cells (31, 40, 54, 62). What the cellular function of tetherin is besides its antiviral activity is unclear, but because expression is upregulated following alpha/beta interferon (IFN-α/β) treatment (1) and tetherin can restrict a range of enveloped viruses, tetherin has been postulated to be a broad-acting mediator of the innate immune defense against enveloped viruses.To circumvent restriction of particle release, HIV-1 encodes the 16-kDa accessory protein Vpu, which antagonizes tetherin and restores normal virus budding (47, 78). The molecular mechanisms by which Vpu does this are not entirely clear, but evidence suggests that Vpu may exert its antagonistic function by downregulating tetherin from the cell surface, trapping it in the trans-Golgi network (10) and targeting it for degradation by the proteasome (12, 39, 81) or lysosome (9, 25, 44); however, degradation of tetherin may be dispensable for Vpu activity (13), and in HIV-1-infected T cells, surface downregulation of tetherin has been reported to be minor (45), suggesting that global removal of tetherin from the plasma membrane may not be necessary to antagonize its function.Tetherin-mediated restriction of HIV-1 and antagonism by Vpu have been the focus of much research, and inhibition of cell-free virus infection has been well documented (33, 47-49, 77, 81, 82). In contrast, less studied is the impact of tetherin on direct cell-cell dissemination. For example, it is not clear if tetherin-mediated restriction inhibits T cell-T cell spread as efficiently as cell-free release or whether tetherin affects VS formation. To address these questions, we analyzed Vpu+ and Vpu viruses for their ability to spread directly between Jurkat T cells and primary CD4+ T cells in the presence or absence of endogenous tetherin. Our data suggest that tetherin does not restrict HIV-1 in the context of cell-to-cell transmission of virus between T cells expressing endogenous tetherin. Interestingly, we also that observed that Vpu-defective virus may disseminate more efficiently by cell-cell spread at the VS. We postulate that cell-cell spread may favor viral pathogenesis by allowing HIV-1 to disseminate in the presence of tetherin during an interferon-producing innate response.  相似文献   

20.
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   

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