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.
2.
Highly active antiretroviral therapy (HAART) can reduce human immunodeficiency virus type 1 (HIV-1) viremia to clinically undetectable levels. Despite this dramatic reduction, some virus is present in the blood. In addition, a long-lived latent reservoir for HIV-1 exists in resting memory CD4+ T cells. This reservoir is believed to be a source of the residual viremia and is the focus of eradication efforts. Here, we use two measures of population structure—analysis of molecular variance and the Slatkin-Maddison test—to demonstrate that the residual viremia is genetically distinct from proviruses in resting CD4+ T cells but that proviruses in resting and activated CD4+ T cells belong to a single population. Residual viremia is genetically distinct from proviruses in activated CD4+ T cells, monocytes, and unfractionated peripheral blood mononuclear cells. The finding that some of the residual viremia in patients on HAART stems from an unidentified cellular source other than CD4+ T cells has implications for eradication efforts.Successful treatment of human immunodeficiency virus type 1 (HIV-1) infection with highly active antiretroviral therapy (HAART) reduces free virus in the blood to levels undetectable by the most sensitive clinical assays (18, 36). However, HIV-1 persists as a latent provirus in resting, memory CD4+ T lymphocytes (6, 9, 12, 16, 48) and perhaps in other cell types (45, 52). The latent reservoir in resting CD4+ T cells represents a barrier to eradication because of its long half-life (15, 37, 40-42) and because specifically targeting and purging this reservoir is inherently difficult (8, 25, 27).In addition to the latent reservoir in resting CD4+ T cells, patients on HAART also have a low amount of free virus in the plasma, typically at levels below the limit of detection of current clinical assays (13, 19, 35, 37). Because free virus has a short half-life (20, 47), residual viremia is indicative of active virus production. The continued presence of free virus in the plasma of patients on HAART indicates either ongoing replication (10, 13, 17, 19), release of virus after reactivation of latently infected CD4+ T cells (22, 24, 31, 50), release from other cellular reservoirs (7, 45, 52), or some combination of these mechanisms. Finding the cellular source of residual viremia is important because it will identify the cells that are still capable of producing virus in patients on HAART, cells that must be targeted in any eradication effort.Detailed analysis of this residual viremia has been hindered by technical challenges involved in working with very low concentrations of virus (13, 19, 35). Recently, new insights into the nature of residual viremia have been obtained through intensive patient sampling and enhanced ultrasensitive sequencing methods (1). In a subset of patients, most of the residual viremia consisted of a small number of viral clones (1, 46) produced by a cell type severely underrepresented in the peripheral circulation (1). These unique viral clones, termed predominant plasma clones (PPCs), persist unchanged for extended periods of time (1). The persistence of PPCs indicates that in some patients there may be another major cellular source of residual viremia (1). However, PPCs were observed in a small group of patients who started HAART with very low CD4 counts, and it has been unclear whether the PPC phenomenon extends beyond this group of patients. More importantly, it has been unclear whether the residual viremia generally consists of distinct virus populations produced by different cell types.Since the HIV-1 infection in most patients is initially established by a single viral clone (23, 51), with subsequent diversification (29), the presence of genetically distinct populations of virus in a single individual can reflect entry of viruses into compartments where replication occurs with limited subsequent intercompartmental mixing (32). Sophisticated genetic tests can detect such population structure in a sample of viral sequences (4, 39, 49). Using two complementary tests of population structure (14, 43), we analyzed viral sequences from multiple sources within individual patients in order to determine whether a source other than circulating resting CD4+ T cells contributes to residual viremia and viral persistence. Our results have important clinical implications for understanding HIV-1 persistence and treatment failure and for improving eradication strategies, which are currently focusing only on the latent CD4+ T-cell reservoir.  相似文献   

3.
Hantaviruses infect endothelial cells and cause 2 vascular permeability-based diseases. Pathogenic hantaviruses enhance the permeability of endothelial cells in response to vascular endothelial growth factor (VEGF). However, the mechanism by which hantaviruses hyperpermeabilize endothelial cells has not been defined. The paracellular permeability of endothelial cells is uniquely determined by the homophilic assembly of vascular endothelial cadherin (VE-cadherin) within adherens junctions, which is regulated by VEGF receptor-2 (VEGFR2) responses. Here, we investigated VEGFR2 phosphorylation and the internalization of VE-cadherin within endothelial cells infected by pathogenic Andes virus (ANDV) and Hantaan virus (HTNV) and nonpathogenic Tula virus (TULV) hantaviruses. We found that VEGF addition to ANDV- and HTNV-infected endothelial cells results in the hyperphosphorylation of VEGFR2, while TULV infection failed to increase VEGFR2 phosphorylation. Concomitant with the VEGFR2 hyperphosphorylation, VE-cadherin was internalized to intracellular vesicles within ANDV- or HTNV-, but not TULV-, infected endothelial cells. Addition of angiopoietin-1 (Ang-1) or sphingosine-1-phosphate (S1P) to ANDV- or HTNV-infected cells blocked VE-cadherin internalization in response to VEGF. These findings are consistent with the ability of Ang-1 and S1P to inhibit hantavirus-induced endothelial cell permeability. Our results suggest that pathogenic hantaviruses disrupt fluid barrier properties of endothelial cell adherens junctions by enhancing VEGFR2-VE-cadherin pathway responses which increase paracellular permeability. These results provide a pathway-specific mechanism for the enhanced permeability of hantavirus-infected endothelial cells and suggest that stabilizing VE-cadherin within adherens junctions is a primary target for regulating endothelial cell permeability during pathogenic hantavirus infection.Hantaviruses cause 2 human diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (50). HPS and HFRS are multifactorial in nature and cause thrombocytopenia, immune and endothelial cell responses, and hypoxia, which contribute to disease (7, 11, 31, 42, 62). Although these syndromes sound quite different, they share common components which involve the ability of hantaviruses to infect endothelial cells and induce capillary permeability. Edema, which results from capillary leakage of fluid into tissues and organs, is a common finding in both HPS and HFRS patients (4, 7, 11, 31, 42, 62). In fact, both diseases can present with renal or pulmonary sequelae, and the renal or pulmonary focus of hantavirus diseases is likely to result from hantavirus infection of endothelial cells within vast glomerular and pulmonary capillary beds (4, 7, 11, 31, 42, 62). All hantaviruses predominantly infect endothelial cells which line capillaries (31, 42, 44, 61, 62), and endothelial cells have a primary role in maintaining fluid barrier functions of the vasculature (1, 12, 55). Although hantaviruses do not lyse endothelial cells (44, 61), this primary cellular target underlies hantavirus-induced changes in capillary integrity. As a result, understanding altered endothelial cell responses following hantavirus infection is fundamental to defining the mechanism of permeability induced by pathogenic hantaviruses (1, 12, 55).Pathogenic, but not nonpathogenic, hantaviruses use β3 integrins on the surface of endothelial cells and platelets for attachment (19, 21, 23, 39, 46), and β3 integrins play prominent roles in regulating vascular integrity (3, 6, 8, 24, 48). Pathogenic hantaviruses bind to basal, inactive conformations of β3 integrins (35, 46, 53) and days after infection inhibit β3 integrin-directed endothelial cell migration (20, 46). This may be the result of cell-associated virus (19, 20, 22) which keeps β3 in an inactive state but could also occur through additional regulatory processes that have yet to be defined. Interestingly, the nonpathogenic hantaviruses Prospect Hill virus (PHV) and Tula virus (TULV) fail to alter β3 integrin functions, and their entry is consistent with the use of discrete α5β1 integrins (21, 23, 36).On endothelial cells, αvβ3 integrins normally regulate permeabilizing effects of vascular endothelial growth factor receptor-2 (VEGFR2) (3, 24, 48, 51). VEGF was initially identified as an edema-causing vascular permeability factor (VPF) that is 50,000 times more potent than histamine in directing fluid across capillaries (12, 14). VEGF is responsible for disassembling adherens junctions between endothelial cells to permit cellular movement, wound repair, and angiogenesis (8, 10, 12, 13, 17, 26, 57). Extracellular domains of β3 integrins and VEGFR2 reportedly form a coprecipitable complex (3), and knocking out β3 causes capillary permeability that is augmented by VEGF addition (24, 47, 48). Pathogenic hantaviruses inhibit β3 integrin functions days after infection and similarly enhance the permeability of endothelial cells in response to VEGF (22).Adherens junctions form the primary fluid barrier of endothelial cells, and VEGFR2 responses control adherens junction disassembly (10, 17, 34, 57, 63). Vascular endothelial cadherin (VE-cadherin) is an endothelial cell-specific adherens junction protein and the primary determinant of paracellular permeability within the vascular endothelium (30, 33, 34). Activation of VEGFR2, another endothelial cell-specific protein, triggers signaling responses resulting in VE-cadherin disassembly and endocytosis, which increases the permeability of endothelial cell junctions (10, 12, 17, 34). VEGF is induced by hypoxic conditions and released by endothelial cells, platelets, and immune cells (2, 15, 38, 52). VEGF acts locally on endothelial cells through the autocrine or paracrine activation of VEGFR2, and the disassembly of endothelial cell adherens junctions increases the availability of nutrients to tissues and facilitates leukocyte trafficking and diapedesis (10, 12, 17, 55). The importance of endothelial cell barrier integrity is often in conflict with requirements for endothelial cells to move in order to permit angiogenesis and repair or cell and fluid egress, and as a result, VEGF-induced VE-cadherin responses are tightly controlled (10, 17, 18, 32, 33, 59). This limits capillary permeability while dynamically responding to a variety of endothelial cell-specific factors and conditions. However, if unregulated, this process can result in localized capillary permeability and edema (2, 9, 10, 12, 14, 17, 29, 60).Interestingly, tissue edema and hypoxia are common findings in both HPS and HFRS patients (11, 31, 62), and the ability of pathogenic hantaviruses to infect human endothelial cells provides a means for hantaviruses to directly alter normal VEGF-VE-cadherin regulation. In fact, the permeability of endothelial cells infected by pathogenic Andes virus (ANDV) or Hantaan virus (HTNV) is dramatically enhanced in response to VEGF addition (22). This response is absent from endothelial cells comparably infected with the nonpathogenic TULV and suggests that enhanced VEGF-induced endothelial cell permeability is a common underlying response of both HPS- and HFRS-causing hantaviruses (22). In these studies, we comparatively investigate responses of human endothelial cells infected with pathogenic ANDV and HTNV, as well as nonpathogenic TULV.  相似文献   

4.
5.
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   

6.
7.
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   

8.
9.
10.
11.
12.
13.
14.
BST-2/tetherin is an interferon-inducible protein that restricts the release of enveloped viruses from the surface of infected cells by physically linking viral and cellular membranes. It is present at both the cell surface and in a perinuclear region, and viral anti-tetherin factors including HIV-1 Vpu and HIV-2 Env have been shown to decrease the cell surface population. To map the domains of human tetherin necessary for both virus restriction and sensitivity to viral anti-tetherin factors, we constructed a series of tetherin derivatives and assayed their activity. We found that the cytoplasmic tail (CT) and transmembrane (TM) domains of tetherin alone produced its characteristic cellular distribution, while the ectodomain of the protein, which includes a glycosylphosphatidylinositol (GPI) anchor, was sufficient to restrict virus release when presented by the CT/TM regions of a different type II membrane protein. To counteract tetherin restriction and remove it from the cell surface, HIV-1 Vpu required the specific sequence present in the TM domain of human tetherin. In contrast, the HIV-2 Env required only the ectodomain of the protein and was sensitive to a point mutation in this region. Strikingly, the anti-tetherin factor, Ebola virus GP, was able to overcome restriction conferred by both tetherin and a series of functional tetherin derivatives, including a wholly artificial tetherin molecule. Moreover, GP overcame restriction without significantly removing tetherin from the cell surface. These findings suggest that Ebola virus GP uses a novel mechanism to circumvent tetherin restriction.Pathogenic viruses often have evolved mechanisms to neutralize host defenses that act at the cellular level to interfere with the virus life cycle. Such cellular restriction factors have been most extensively characterized for HIV-1 (38) and include the interferon-inducible membrane protein BST-2/HM1.24/CD317/tetherin (28, 40). If unchecked, tetherin blocks the release of newly formed HIV-1 particles from cells by physically tethering them at the cell surface (7, 28, 32, 40). In addition, tetherin has been shown to act against a broad range of enveloped viral particles, including retroviruses, filoviruses, arenaviruses, and herpesviruses (17, 18, 23, 35). In turn, certain viruses that are targeted by tetherin appear to have evolved counteracting activities, and anti-tetherin factors so far identified include HIV-1 Vpu; HIV-2 Env; simian immunodeficiency virus (SIV) Nef, Vpu, and Env proteins; Ebola virus GP; and Kaposi''s sarcoma-associated herpesvirus (KSHV) K5 (11, 16, 18, 20, 23, 28, 36, 40, 44, 45).Tetherin is a homodimeric type II integral membrane protein containing an N-terminal cytoplasmic tail (CT), a single-pass transmembrane domain (TM), an ectodomain-containing predicted coiled-coil regions, two glycoslyation sites, three conserved cysteines, and a C-terminal glycosylphosphatidylinositol (GPI) anchor (2, 19, 31). This unusual topology, with two independent membrane anchors, has led to the suggestion that the retention of virions at the cell surface arises from tetherin''s ability to be inserted simultaneously in both host and viral membranes (28, 32, 41) or, alternatively, that dimers or higher-order complexes of tetherin conferred by the ectodomain mediate this effect (39). Interestingly, an artificial tetherin containing the same structural features as the native protein but constructed from unrelated sequences was able to restrict both HIV-1 and Ebola virus particles (32). This suggests that the viral lipid envelope is the target of tetherin and provides an explanation for tetherin''s broad activity against diverse enveloped viruses.A fraction of tetherin is present at the plasma membrane of cells (9, 14), and it has been proposed that viral anti-tetherin factors function by removing this cell surface fraction (40). This now has been shown to occur in the presence of HIV-1 Vpu (5, 7, 15, 26, 34, 40, 44), HIV-2 Env (5, 20), SIV Env (11), SIV Nef (15), and KSHV K5 (3, 23). In addition, certain anti-tetherin factors also may promote the degradation of tetherin, as has been observed for both HIV-1 Vpu (3, 5, 7, 10, 22, 26, 27) and KSHV K5 (3, 23), although Vpu also appears able to block tetherin restriction in the absence of degradation (8), and no effects on tetherin steady-state levels have been observed in the presence of either the HIV-2 or SIVtan Env (11, 20). Simply keeping tetherin away from the cell surface, or targeting it for degradation, may not be the only mechanism used by anti-tetherin factors, since it also has been reported that Vpu does not affect the levels of surface tetherin or its total cellular levels in certain T-cell lines (27).The interactions between tetherin and viral anti-tetherin factors show evidence of species specificity, suggesting ongoing evolution between viruses and their hosts. HIV-1 Vpu is active against human and chimpanzee tetherin but not other primate tetherins (10, 25, 34, 36, 44, 45), while SIV Nef proteins are active against primate but not human tetherins (16, 36, 44, 45). This suggests that, unlike tetherin restriction, the action of the anti-tetherin factors may involve specific sequence interactions. Indeed, the TM domain has been recognized as a target for HIV-1 Vpu (10, 15, 16, 25, 34), while a single point mutation introduced into the extracellular domain of human tetherin can block its antagonism by the SIVtan Env (11).In the present study, we investigated the roles of the different domains of tetherin in both promoting virus restriction and conferring susceptibility to the anti-tetherin factors encoded by HIV-1, HIV-2, and Ebola virus. We confirmed that tetherin restriction can be conferred by proteins that retain the two distinct membrane anchors, while signals for the cellular localization of the protein reside in the CT/TM domains of the protein. We found that the Vpu protein targets the TM domain of tetherin, while the HIV-2 Env targets the ectodomain of the protein. In contrast, the Ebola virus GP appears to use a non-sequence-specific mechanism to counteract tetherin restriction, since even an artificial tetherin could be successfully overcome by GP expression. Interestingly, Ebola virus GP counteracted tetherin restriction without removing the protein from the cell surface, suggesting that it is possible to overcome this restriction by mechanisms other than blocking tetherin''s cell surface expression.  相似文献   

15.
Cell migration is critical for normal development and for pathological processes including cancer cell metastasis. Dynamic remodeling of focal adhesions and the actin cytoskeleton are crucial determinants of cell motility. The Rho family and the mitogen-activated protein kinase (MAPK) module consisting of MEK-extracellular signal-regulated kinase (ERK) are important regulators of these processes, but mechanisms for the integration of these signals during spreading and motility are incompletely understood. Here we show that ERK activity is required for fibronectin-stimulated Rho-GTP loading, Rho-kinase function, and the maturation of focal adhesions in spreading cells. We identify p190A RhoGAP as a major target for ERK signaling in adhesion assembly and identify roles for ERK phosphorylation of the C terminus in p190A localization and activity. These observations reveal a novel role for ERK signaling in adhesion assembly in addition to its established role in adhesion disassembly.Cell migration is a highly coordinated process essential for physiological and pathological processes (69). Signaling through Rho family GTPases (e.g., Rac, Cdc42, and Rho) is crucial for cell migration. Activated Rac and Cdc42 are involved in the production of a dominant lamellipodium and filopodia, respectively, whereas Rho-stimulated contractile forces are required for tail retraction and to maintain adhesion to the matrix (57, 58, 68). Rac- and Cdc42-dependent membrane protrusions are driven by the actin cytoskeleton and the formation of peripheral focal complexes; Rho activation stabilizes protrusions by stimulating the formation of mature focal adhesions and stress fibers. Active Rho influences cytoskeletal dynamics through effectors including the Rho kinases (ROCKs) (2, 3).Rho activity is stimulated by GEFs that promote GTP binding and attenuated by GTPase-activating proteins (GAPs) that enhance Rho''s intrinsic GTPase activity. However, due to the large number of RhoGEFs and RhoGAPs expressed in mammalian cells, the molecular mechanisms responsible for regulation of Rho activity in time and space are incompletely understood. p190A RhoGAP (hereafter p190A) is implicated in adhesion and migration signaling. p190A contains an N-terminal GTPase domain, a large middle domain juxtaposed to the C-terminal GAP domain, and a short C-terminal tail (74). The C-terminal tail of ∼50 amino acids is divergent between p190A and the closely related family member p190B (14) and thus may specify the unique functional roles for p190A and p190B revealed in gene knockout studies (10, 11, 41, 77, 78). p190A activity is dynamically regulated in response to external cues during cell adhesion and migration (5, 6, 59). Arthur et al. (5) reported that p190A activity is required for the transient decrease in RhoGTP levels seen in fibroblasts adhering to fibronectin. p190A activity is positively regulated by tyrosine phosphorylation (4, 5, 8, 17, 31, 39, 40, 42): phosphorylation at Y1105 promotes its association with p120RasGAP and subsequent recruitment to membranes or cytoskeleton (8, 17, 27, 31, 71, 75, 84). However, Y1105 phosphorylation is alone insufficient to activate p190A GAP activity (39). While the functions of p190A can be irreversibly terminated by ubiquitinylation in a cell-cycle-dependent manner (80), less is known about reversible mechanisms that negatively regulate p190A GAP activity during adhesion and motility.The integration of Rho family GTPase and extracellular signal-regulated kinase (ERK) signaling is important for cell motility (48, 50, 63, 76, 79). Several studies have demonstrated a requirement for ERK signaling in the disassembly of focal adhesions in migrating cells, in part through the activation of calpain proteases (36, 37) that can downregulate focal adhesion kinase (FAK) signaling (15), locally suppress Rho activity (52), and sever cytoskeletal linkers to focal adhesions (7, 33). Inhibition of ERK signaling increases focal adhesion size and retards disassembly of focal adhesions in adherent cells (57, 64, 85, 86). It is also recognized that ERK modulates Rho-dependent cellular processes, including membrane protrusion and migration (18, 25, 64, 86). Interestingly, ERK activated in response to acute fibronectin stimulation localizes not only to mature focal adhesions, but also to peripheral focal complexes (32, 76). Since these complexes can either mature or be turned over (12), ERK may play a distinct role in focal adhesion assembly. ERK is proposed to promote focal adhesion formation by activating myosin light chain kinase (MLCK) (21, 32, 50).Here we find that ERK activity is required for Rho activation and focal adhesion formation during adhesion to fibronectin and that p190A is an essential target of ERK signaling in this context. Inspection of the p190A C terminus reveals a number of consensus ERK sites and indeed p190A is phosphorylated by recombinant ERK only on its C terminus in vitro, and on the same C-terminal peptide in vivo. Mutation of the C-terminal ERK phosphorylation sites to alanine increases the biochemical and biological activity of p190A. Finally, inhibition of MEK or mutation of the C-terminal phosphorylation sites enhances retention of p190A in peripheral membranes during spreading on fibronectin. Our data support the conclusion that ERK phosphorylation inhibits p190A allowing increases in RhoGTP and cytoskeletal changes necessary for focal adhesion formation.  相似文献   

16.
Kinetic analyses of bacterial growth, carbohydrate consumption, and metabolite production of five butyrate-producing clostridial cluster XIVa colon bacteria grown on acetate plus fructose, oligofructose, inulin, or lactate were performed. A gas chromatography method was set up to assess H2 and CO2 production online and to ensure complete coverage of all metabolites produced. Method accuracy was confirmed through the calculation of electron and carbon recoveries. Fermentations with Anaerostipes caccae DSM 14662T, Roseburia faecis DSM 16840T, Roseburia hominis DSM 16839T, and Roseburia intestinalis DSM 14610T revealed similar patterns of metabolite production with butyrate, CO2, and H2 as the main metabolites. R. faecis DSM 16840T and R. intestinalis DSM 14610T were able to degrade oligofructose, displaying a nonpreferential breakdown mechanism. Lactate consumption was only observed with A. caccae DSM 14662T. Roseburia inulinivorans DSM 16841T was the only strain included in the present study that was able to grow on fructose, oligofructose, and inulin. The metabolites produced were lactate, butyrate, and CO2, without H2 production, indicating an energy metabolism distinct from that of other Roseburia species. Oligofructose degradation was nonpreferential. In a coculture of R. inulinivorans DSM 16841T with the highly competitive strain Bifidobacterium longum subsp. longum LMG 11047 on inulin, hardly any production of butyrate and CO2 was detected, indicating a lack of competitiveness of the butyrate producer. Complete recovery of metabolites during fermentations of clostridial cluster XIVa butyrate-producing colon bacteria allowed stoichiometric balancing of the metabolic pathway for butyrate production, including H2 formation.The implementation of 16S rRNA gene-based analytical techniques in the ongoing exploration of the microbial diversity of the human colon ecosystem has both broadened and sharpened the prevailing image of its population (17, 24, 32). While a rather conservative perception of the composition of the colon microbiota has dominated gut research for several decades (36), recent studies have revealed the importance of previously largely neglected bacterial groups and have reduced historically numerically overestimated subpopulations to their actual (marginal) size (8, 22, 52). The human colon has been shown to be a remarkably selective environment, which is reflected by a rather shallow microbial diversity (32). Species belonging to the bacterial divisions Firmicutes, Bacteroidetes, Proteobacteria, and Actinobacteria make up more than 98% of the bacterial population of the human colon (2, 17, 24). However, this superficial uniformity only covers an overwhelming diversity at the lower taxonomic levels; the human colon has been estimated to harbor between 500 and 1,000 species, representing over 7,000 strains, with up to 80% of them considered uncultivable using presently available methodologies (14, 28, 53).Assessing identity and abundance of the major microbial groups composing the colon microbiota is a first and indispensable step toward a better understanding of the ecosystem of the large intestine (48). However, defining a complex ecosystem such as the human colon requires more than the construction of a catalog of its members (32). A major challenge of gastrointestinal microbiology lies in linking phylogenetic subgroups with particular ecological habitats and niches (7, 8, 23). The latter requires further development of highly discriminating 16S rRNA gene-targeted probes to monitor spatial bacterial distribution, combined with renewed efforts toward species isolation through the application of innovative cultivation methods and media, and extensive metabolic characterization of representative strains (19, 35, 48).Recently, a global ecological approach, combining efforts in probe development (1, 27), species isolation (3), and metabolic characterization (4, 11, 15, 20), has led to the identification of a functional group of microorganisms, composed of species belonging to the clostridial clusters IV and XIVa, that are responsible for colon butyrate production. As butyrate is regarded as a key metabolite for the maintenance of colon health, this functional subunit of the colon microbiota could have a major influence on human well-being and might be considered as a target for prebiotic dietary interventions (25, 35, 45). Some recently described lactate- and/or acetate-converting colon butyrate producers have been reported to be able to degrade prebiotic inulin-type fructans, although the kinetics of their respective breakdown mechanisms have hardly been investigated (10, 20). The enhancement of colon butyrate production observed after consumption of oligofructose or inulin (6, 31, 40)—the so-called butyrogenic effect—as well as the limited stimulatory effect of these prebiotics on the clostridial cluster IV and XIVa colon populations (16, 30) have been attributed to cross-feeding with bifidobacteria, which are still considered the primary fructan degraders (5, 38). Anaerostipes caccae as well as Roseburia spp. have been shown to be able to (co)metabolize end products of bifidobacterial fructan fermentation (lactate and/or acetate) or to grow on short oligosaccharides and monosaccharides released by Bifidobacterium spp. during fructan degradation (4, 20).Recently, many clostridial cluster IV and XIVa butyrate producers characterized in detail have been shown to produce gases, mainly CO2 and H2 (12, 15, 20, 46). Consequently, they might be responsible for an enhancement of gas production as a result of fructan fermentation, through either cross-feeding or direct degradation of inulin-type fructans (15, 16). Indeed, inulin-type fructan consumption has been reported to cause some gastrointestinal discomfort related to gas production—essentially, flatulence and bloating (43)—while bifidobacteria, the main beneficiaries of dietary fructan intake, do not produce gases (19, 49). Although CO2 and H2 production by colon butyrate producers could have implications for human intestinal well-being, (in vitro) production has not been satisfactorily monitored up to now, probably due to limited availability of a performant apparatus for (online) gas analysis (15, 20). Moreover, the currently proposed pathway for colon butyrate production does not provide a conclusive quantitative link between bacterial (co)substrate metabolism and H2 formation (11).This study investigated the kinetics of inulin-type fructan degradation by representatives of the genera Anaerostipes and Roseburia. A method based on online gas chromatography (GC) was developed to assess gas production qualitatively and quantitatively in a continuously sparged fermentation vessel for complete coverage of metabolite production. The competitiveness of inulin-degrading butyrate producers was investigated through coculture fermentations with Bifidobacterium longum subsp. longum LMG 11047, a strain representing a highly competitive cluster of bifidobacteria that share both high fructose consumption and oligofructose degradation rates and are able to perform partial breakdown of inulin (18, 20). A stoichiometrically balanced pathway for butyrate production, including H2 production, is proposed.  相似文献   

17.
18.
19.
Vesicle-associated-membrane protein 8 (VAMP8) is highly expressed in the kidney, but the exact physiological and molecular functions executed by this v-SNARE protein in nephrons remain elusive. Here, we show that the depletion of VAMP8 in mice resulted in hydronephrosis. Furthermore, the level of the vasopressin-responsive water channel aquaporin 2 (AQP2) was increased by three- to fivefold in VAMP8-null mice. Forskolin and [desamino-Cys1, D-Arg8]-vasopressin (DDAVP)-induced AQP2 exocytosis was impaired in VAMP8-null collecting duct cells. VAMP8 was revealed to colocalize with AQP2 on intracellular vesicles and to interact with the plasma membrane t-SNARE proteins syntaxin4 and syntaxin3, suggesting that VAMP8 mediates the regulated fusion of AQP2-positive vesicles with the plasma membrane.Aquaporins are water channels that can facilitate the movement of water through membranes. They play a crucial role in maintaining body water homeostasis (22, 23). Of the known aquaporins, aquaporin 2 (AQP2) is the major, if not the only, vasopressin-responsive aquaporin. Under physiological conditions, AQP2 exists in a dynamic equilibrium between the plasma membrane and intracellular vesicles. The binding of vasopressin to the G-protein-coupled V2 receptor induces a series of coordinated signaling events in collecting duct cells that eventually lead to elevated levels of surface AQP2 and, consequently, increased water reabsorption. The signaling events include, but may not be limited to, the activation of adenylate cyclase, which elevates the cyclic AMP (cAMP) level, the phosphorylation of AQP2 at S256 by protein kinase A (PKA), the reorganization of the cytoskeleton, and a transient increase in free intracellular Ca2+. In addition to rapidly deploying AQP2 to the plasma membrane from intracellular vesicles, vasopressin can execute a long-term effect on AQP2 by upregulating its gene expression level (7, 22).The regulated exocytosis of AQP2 ultimately requires the fusion of AQP2-containing vesicles with the plasma membrane, a process known to be driven by SNARE proteins (12, 14). Several SNARE proteins have been identified in kidney collecting duct cells. These include syntaxin2, syntaxin3, and syntaxin4 on the plasma membrane (15), VAMP2 and VAMP3 on intracellular vesicles (5, 20), and SNAP23 on both intracellular vesicles and the plasma membrane (13). Although proposed long ago, and there are some functional studies with cultured cells (10, 24), the roles of these proteins in AQP2 exocytosis have yet to be clarified and confirmed in animal models.It has been known for more than a decade that the tissue that expresses the most VAMP8 is the kidney (1), but the exact role of VAMP8 in this organ remains elusive. Although initially identified as the endosomal SNARE protein endobrevin (1, 3, 37), VAMP8 recently has been shown to play a major role in the regulated secretory pathway of the exocrine system (8, 34, 35), platelets (26), basophilic cells (16), and mast cells (25, 28, 33). In this study, we show that VAMP8 is present on intracellular AQP2-containing vesicles and appears to play an important role in AQP2 exocytosis.  相似文献   

20.
The DosS (DevS) and DosT histidine kinases form a two-component system together with the DosR (DevR) response regulator in Mycobacterium tuberculosis. DosS and DosT, which have high sequence similarity to each other over the length of their amino acid sequences, contain two GAF domains (GAF-A and GAF-B) in their N-terminal sensory domains. Complementation tests in conjunction with phylogenetic analysis showed that DevS of Mycobacterium smegmatis is more closely related to DosT than DosS. We also demonstrated in vivo that DosS and DosT of M. tuberculosis play a differential role in hypoxic adaptation. DosT responds to a decrease in oxygen tension more sensitively and strongly than DosS, which might be attributable to their different autooxidation rates. The different responsiveness of DosS and DosT to hypoxia is due to the difference in their GAF-A domains accommodating the hemes. Multiple alignment analysis of the GAF-A domains of mycobacterial DosS (DosT) homologs and subsequent site-directed mutagenesis revealed that just one substitution of E87, D90, H97, L118, or T169 of DosS with the corresponding residue of DosT is sufficient to convert DosS to DosT with regard to the responsiveness to changes in oxygen tension.Oxygen sensing is important for facultative anaerobes to adapt to changes in metabolic necessities during the transition between aerobic and anaerobic conditions. Although Mycobacterium tuberculosis (MTB) is an obligate aerobe, a gradual depletion of O2 from its culture is known to lead to a drastic change in gene expression (8, 21, 24, 28, 34, 37, 39). Approximately 48 genes of M. tuberculosis were reported to be induced under early hypoxic conditions, which is mediated by the DosSR (DevSR) two-component system (16, 24, 34). The induction of the DosR regulon is important for survival of M. tuberculosis under hypoxic conditions and for it to enter the nonreplicating dormant state (2, 19). The DosSR two-component system consists of the DosS histidine kinase (HK) and its cognate DosR response regulator (RR) (24, 26, 29). The DosT HK, which shares high sequence similarity to DosS over the length of their primary structures, was also found to cross talk with DosR (26, 30). The N-terminal domains of DosS and DosT contain two tandem GAF domains (GAF-A and GAF-B from their N termini), and the three-dimensional structure of the GAF-A and GAF-B domains was determined (5, 25). A b-type heme is embedded in the GAF-A domain, composed of one five-stranded antiparallel β-sheet and four α-helices (5, 14, 25, 32). The heme is positioned nearly perpendicular to the β-sheet, and H149 and H147 of the polypeptides serve as the proximal axial ligands for DosS and DosT, respectively (5, 25). The ligand-binding state at the distal axial position of heme and the redox state of the heme iron modulate the autokinase activity of DosS and DosT. The O2-bound (oxyferrous) and ferric forms of the HKs are inactive, whereas the unliganded ferrous (deoxyferrous) form as well as NO- and CO-bound forms are active (17, 36). The heme iron of DosT is stable against autooxidation of Fe2+ to Fe3+ in the presence of O2, indicating that its conversion between deoxyferrous and oxyferrous forms is the mechanism by which DosT recognizes O2 (17). However, the autooxidation property of oxyferrous DosS remains controversial. Kumar et al. (17) and Cho et al. (5) reported that DosS undergoes autooxidation on exposure to O2, while other research groups demonstrated that the oxyferrous form of DosS is stable against autooxidation (13, 14, 36). Recently, different roles of DosS and DosT in O2 sensing by M. tuberculosis were suggested. DosT plays a more important role in the early phase of hypoxic conditions than DosS when the growth of M. tuberculosis is transferred from aerobic to hypoxic conditions (11).Mycobacterium smegmatis possesses a single DevS HK that phosphorylates the DevR RR (20). The DevSR two-component system is also implemented in hypoxic adaptation of this bacterium (20). Like DosT of M. tuberculosis, the autokinase activity of M. smegmatis DevS was shown to be controlled by the ligand-binding state of its heme (18). Regarding the autooxidation property, DevS of M. smegmatis was suggested to be similar to DosT rather than DosS; i.e., the heme iron in DevS is resistant to autooxidation from an oxyferrous to a ferric state in the presence of O2 (18).In this paper we report several lines of evidence for the functional difference between DosS and DosT in the hypoxic adaptation of mycobacteria and discuss the implications of these findings.  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