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.
Type I interferons (IFNs) play a critical role in the host defense against viruses. Lymphocytic choriomeningitis virus (LCMV) infection induces robust type I IFN production in its natural host, the mouse. However, the mechanisms underlying the induction of type I IFNs in response to LCMV infection have not yet been clearly defined. In the present study, we demonstrate that IRF7 is required for both the early phase (day 1 postinfection) and the late phase (day 2 postinfection) of the type I IFN response to LCMV, and melanoma differentiation-associated gene 5 (MDA5)/mitochondrial antiviral signaling protein (MAVS) signaling is crucial for the late phase of the type I IFN response to LCMV. We further demonstrate that LCMV genomic RNA itself (without other LCMV components) is able to induce type I IFN responses in various cell types by activation of the RNA helicases retinoic acid-inducible gene I (RIG-I) and MDA5. We also show that expression of the LCMV nucleoprotein (NP) inhibits the type I IFN response induced by LCMV RNA and other RIG-I/MDA5 ligands. These virus-host interactions may play important roles in the pathogeneses of LCMV and other human arenavirus diseases.Type I interferons (IFNs), namely, alpha interferon (IFN-α) and IFN-β, are not only essential for host innate defense against viral pathogens but also critically modulate the development of virus-specific adaptive immune responses (6, 8, 28, 30, 36, 50, 61). The importance of type I IFNs in host defense has been demonstrated by studying mice deficient in the type I IFN receptor, which are highly susceptible to most viral pathogens (2, 47, 62).Recent studies have suggested that the production of type I IFNs is controlled by different innate pattern recognition receptors (PRRs) (19, 32, 55, 60). There are three major classes of PRRs, including Toll-like receptors (TLRs) (3, 40), retinoic acid-inducible gene I (RIG-I)-like receptors (RLRs) (25, 48, 51), and nucleotide oligomerization domain (NOD)-like receptors (9, 22). TLRs are a group of transmembrane proteins expressed on either cell surfaces or endosomal compartments. RLRs localize in the cytosol. Both TLRs and RLRs are involved in detecting viral pathogens and controlling the production of type I IFNs (52, 60). In particular, the endosome-localized TLRs (TLR3, TLR7/8, and TLR9) play important roles in detecting virus-derived double-stranded RNA (dsRNA), single-stranded RNA (ssRNA), and DNA-containing unmethylated CpG motifs, respectively. In contrast, RIG-I detects virus-derived ssRNA with 5′-triphosphates (5′-PPPs) or short dsRNA (<1 kb), whereas melanoma differentiation-associated gene 5 (MDA5) is responsible for recognizing virus-derived long dsRNA as well as a synthetic mimic of viral dsRNA poly(I):poly(C) [poly(I·C)] (24, 60). Recognition of viral pathogen-associated molecular patterns (PAMPs) ultimately leads to the activation and nuclear translocation of interferon regulatory factors (IRFs) and nuclear factor κB (NF-κB), which, in turn, switches on a cascade of genes controlling the production of both type I IFNs and other proinflammatory cytokines (10, 11, 60).Lymphocytic choriomeningitis virus (LCMV) infection in its natural host, the mouse, is an excellent system to study the impact of virus-host interactions on viral pathogenesis and to address important issues related to human viral diseases (1, 45, 49, 67). LCMV infection induces type I IFNs as well as other proinflammatory chemokines and cytokines (6, 41). Our previous studies have demonstrated that TLR2, TLR6, and CD14 are involved in LCMV-induced proinflammatory chemokines and cytokines (66). The mechanism by which LCMV induces type I IFN responses, however, has not been clearly defined (7, 8, 31, 44). The role of the helicase family members RIG-I and MDA5 in virus-induced type I IFN responses has been recently established. RIG-I has been found to be critical in controlling the production of type I IFN in response to a number of RNA viruses, including influenza virus, rabies virus, Hantaan virus, vesicular stomatitis virus (VSV), Sendai virus (SeV), etc. In contrast, MDA5 is required for responses to picornaviruses (15, 25, 63).In the present study, we demonstrated that LCMV genomic RNA strongly activates type I IFNs through a RIG-I/MDA5-dependent signaling pathway. Our present study further demonstrated that the LCMV nucleoprotein (NP) blocks LCMV RNA- and other viral ligand-induced type I IFN responses.  相似文献   

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

4.
Intracranial (i.c.) infection of mice with lymphocytic choriomeningitis virus (LCMV) results in anorexic weight loss, mediated by T cells and gamma interferon (IFN-γ). Here, we assessed the role of CD4+ T cells and IFN-γ on immune cell recruitment and proinflammatory cytokine/chemokine production in the central nervous system (CNS) after i.c. LCMV infection. We found that T-cell-depleted mice had decreased recruitment of hematopoietic cells to the CNS and diminished levels of IFN-γ, CCL2 (MCP-1), CCL3 (MIP-1α), and CCL5 (RANTES) in the cerebrospinal fluid (CSF). Mice deficient in IFN-γ had decreased CSF levels of CCL3, CCL5, and CXCL10 (IP-10), and decreased activation of both resident CNS and infiltrating antigen-presenting cells (APCs). The effects of IFN-γ signaling on macrophage lineage cells was assessed using transgenic mice, called “macrophages insensitive to interferon gamma” (MIIG) mice, that express a dominant-negative IFN-γ receptor under the control of the CD68 promoter. MIIG mice had decreased levels of CCL2, CCL3, CCL5, and CXCL10 compared to controls despite having normal numbers of LCMV-specific CD4+ T cells in the CNS. MIIG mice also had decreased recruitment of infiltrating macrophages and decreased activation of both resident CNS and infiltrating APCs. Finally, MIIG mice were significantly protected from LCMV-induced anorexia and weight loss. Thus, these data suggest that CD4+ T-cell production of IFN-γ promotes signaling in macrophage lineage cells, which control (i) the production of proinflammatory cytokines and chemokines, (ii) the recruitment of macrophages to the CNS, (iii) the activation of resident CNS and infiltrating APC populations, and (iv) anorexic weight loss.Immune cell recruitment to and infiltration of the central nervous system (CNS) is central to the pathology of a variety of inflammatory neurological diseases, including infectious meningoencephalitis, multiple sclerosis, and cerebral ischemia (59, 60). Chemokines have been shown to be highly upregulated in both human diseases and animal models of neuroinflammation and are thought to be important mediators of immune cell entry into the CNS (59, 60). For example, during experimental autoimmune encephalomyelitis (EAE) and multiple sclerosis (MS), the chemokines CCL2 (monocyte chemoattractant protein 1 [MCP-1α]), CCL3 (macrophage inflammatory protein 1α [MIP-1α]), CCL5 (regulated upon activation, T-cell expressed and secreted [RANTES]), and CXCL10 (gamma interferon [IFN-γ]-inducible protein 10 [IP-10]) are produced by either resident CNS cells or infiltrating cells (27) and serve to amplify the ongoing inflammatory response (25, 28). However, in some EAE studies, neither CCL3 nor CXCL10 were required for disease (72, 73). During CNS viral infection, CXCL10 and CCL5 are highly produced in several models (2, 41, 48, 82). In addition, mice deficient in CCR5, which binds (among others) CCL3 and CCL5, do not display impaired CNS inflammation after certain viral infections (13). Thus, the role of chemokines in CNS inflammation is likely complex and dissimilar between autoimmune and viral infection models.IFN-γ is present in the CNS during autoimmunity and infection (7, 54, 69). Several studies suggest that IFN-γ can be a potent inducer of CNS chemokine expression. Adenoviral expression of IFN-γ in the CNS strongly induced CCL5 and CXCL10 mRNA and protein, and this induction was dependent on the presence of the IFN-γ receptor (50). In EAE and Toxoplasma infection, mice deficient in IFN-γ or the IFN-γ receptor demonstrated reduced expression of several chemokines, including CCL2, CCL3, CCL5, and CXCL10 (26, 69). However, given the near-ubiquitous expression of the IFN-γ receptor (44), the mechanisms by which IFN-γ regulates CNS chemokine production remain to be elucidated.We studied neuroinflammation and immune-mediated disease using a well-studied mouse model of infection with lymphocytic choriomeningitis virus (LCMV). Intracranial (i.c.) injection of mice with LCMV results in seizures and death 6 to 8 days after inoculation. The onset of symptoms is associated with a massive influx of mononuclear cells into the cerebrospinal fluid (CSF), meninges, choroid plexus, and ependymal membranes (6, 8, 18), as well as the presence of proinflammatory cytokines (7, 38). The immune response is critical for disease, since infection of irradiated or T-cell-depleted mice leads to persistent infection with very high levels of virus in multiple tissues without the development of lethal meningitis (18, 34, 64). i.c. LCMV infection of β2-microglobulin-deficient mice (β2m−/− mice) also results in meningitis and production of proinflammatory cytokines and chemokines; however, meningitis occurs with a later onset and lower severity compared to wild-type mice (17, 24, 53, 57). Interestingly, i.c. LCMV infection of these mice also causes severe anorexia and weight loss (33, 38, 46, 52, 57) that is mediated by major histocompatibility complex (MHC) class II-restricted, CD4+ T cells (17, 46, 53, 57). Anorexia and weight loss are also observed in wild-type mice, but they succumb to lethal meningitis shortly thereafter (33), making study of this particular aspect of disease difficult. LCMV-induced weight loss, similar to what we have observed in β2m−/− mice also occurs in perforin-deficient mice, which possess CD8+ T cells (37). Although some reports have observed weight loss after peripheral LCMV infection (11, 45), we note that these studies used high doses of the clone 13 strain of LCMV, in contrast to our studies which have used the Armstrong strain of LCMV and orders of magnitude less virus (33, 38, 46, 52, 57). Although we cannot exclude a contribution of peripheral cells to weight loss in our i.c. Armstrong infection model, we previously showed that this weight loss does not occur with peripheral infection with LCMV Armstrong (33, 38), indicating that interactions between the CNS and the immune system are contribute substantially to disease.During LCMV infection, there is biphasic production of IFN-γ: a small, early peak of IFN-γ (most likely produced by NK or NKT cells), followed by T-cell-mediated production of IFN-γ (23, 75). Further, both CD4+ T cells and CD8+ T cells produce large amounts of IFN-γ after LCMV infection and T-cell production of IFN-γ is critical for LCMV-induced weight loss (35). Chemokines, especially CXCL10, CCL5, and CCL2, and their receptors, are upregulated in the brain after i.c. LCMV infection (2, 13). Brain chemokine mRNA expression after i.c. LCMV infection is reduced in IFN-γ-deficient mice and relatively absent in athymic mice (2). However, the mechanism(s) by which T cells and IFN-γ mediate the effects on CNS chemokine expression, cellular infiltration into the CNS, and LCMV-induced anorexic weight loss remain unclear.In the present study, we focused on two major questions. The first question concerned the role of IFN-γ on immune cell recruitment to and chemokine/cytokine production within the CNS? We found that macrophages and myeloid dendritic cells (DCs) require IFN-γ for their accumulation within the CNS. Second, since macrophages and myeloid DCs are the predominant cellular infiltrate, we sought to determine whether IFN-γ signaling on these cells was direct with regard to their recruitment and to chemokine/cytokine production. We found that IFN-γ signaling in macrophage lineage cells contributes significantly to their recruitment, to chemokine production in the CNS, and to anorexic weight loss. Together, these data suggest that much of the proinflammatory effects of IFN-γ in the CNS are mediated by the effects of IFN-γ on CD68-bearing cells.  相似文献   

5.
6.
7.
8.
9.
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   

10.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

11.
Prion strain interference can influence the emergence of a dominant strain from a mixture; however, the mechanisms underlying prion strain interference are poorly understood. In our model of strain interference, inoculation of the sciatic nerve with the drowsy (DY) strain of the transmissible mink encephalopathy (TME) agent prior to superinfection with the hyper (HY) strain of TME can completely block HY TME from causing disease. We show here that the deposition of PrPSc, in the absence of neuronal loss or spongiform change, in the central nervous system corresponds with the ability of DY TME to block HY TME infection. This suggests that DY TME agent-induced damage is not responsible for strain interference but rather prions compete for a cellular resource. We show that protein misfolding cyclic amplification (PMCA) of DY and HY TME maintains the strain-specific properties of PrPSc and replicates infectious agent and that DY TME can interfere, or completely block, the emergence of HY TME. DY PrPSc does not convert all of the available PrPC to PrPSc in PMCA, suggesting the mechanism of prion strain interference is due to the sequestering of PrPC and/or other cellular components required for prion conversion. The emergence of HY TME in PMCA was controlled by the initial ratio of the TME agents. A higher ratio of DY to HY TME agent is required for complete blockage of HY TME in PMCA compared to several previous in vivo studies, suggesting that HY TME persists in animals coinfected with the two strains. This was confirmed by PMCA detection of HY PrPSc in animals where DY TME had completely blocked HY TME from causing disease.Prions are infectious agents of animals, including humans, which are comprised of PrPSc, a misfolded isoform of the noninfectious host encoded protein PrPC (17, 24, 50, 63). Prion diseases of humans are unique neurodegenerative disorders in that they can have either a sporadic, familial, or infectious etiology. Prions cause disease in economically important domestic and wild animal species such as bovine spongiform encephalopathy in cattle and chronic wasting disease in wild and captive cervids (20, 62). Prion diseases can be zoonotic as illustrated by the transmission of bovine spongiform encephalopathy to humans that resulted in the emergence of variant Creutzfeldt-Jacob disease (14, 19, 22, 23, 46, 61, 68). Prion diseases are inevitably fatal and there are currently no effective treatments (21).Prion strains are defined by a characteristic set of features that breed true upon experimental passage (33, 34). Strain-specific differences have been identified in incubation period, clinical signs, agent distribution, overdominance, host range, neuropathology, and biochemical properties of PrPSc (5, 10, 11, 13, 28, 34, 42, 44). Strain-specific conformations of PrPSc are hypothesized to encode prion strain diversity; however, it is not understood how these differences result in the distinct strain properties (11, 19, 40, 47, 59, 66).Prion strain interference may be involved in the emergence of a dominant strain from a mixture as could occur during prion adaptation to a new host species or during prion evolution (4, 36, 43, 48, 56). In the natural prion diseases, there are examples where an individual host may be infected with more than one prion strain (15, 25, 55, 57, 58). Experimentally, coinfection or superinfection of prion strains can result in interference where a blocking, long incubation period strain extends the incubation period or completely blocks a superinfecting, short incubation period strain from causing disease (26, 27). Prion interference has been described in experimental studies of mice and hamsters infected with a wide variety of prion strains and routes of inoculation, suggesting it may be a common property of prion disease (3, 27, 52, 53, 60).It has been proposed that prion strains compete for a shared “replication site”; however, mechanistic details are not known, and it is unclear whether the blocking strain destroys or occupies the replication sites required for the superinfecting strain (28). The transport to and relative onset of replication of interfering strains in a common population of neurons is an important factor that can determine which strain will emerge (8). In the present study, we sought to determine whether the blocking strain disables transport and spread of the superinfecting strain or whether prion interference is due to competition for a cellular resource.  相似文献   

12.
Persistent viral infections are a major health concern worldwide. During persistent infection, overwhelming viral replication and the rapid loss of antiviral T-cell function can prevent immune-mediated clearance of the infection, and therapies to reanimate the immune response and purge persistent viruses have been largely unsuccessful. Adoptive immunotherapy using memory T cells is a highly successful therapeutic approach to eradicate a persistent viral infection. Understanding precisely how therapeutically administered memory T cells achieve clearance should improve our ability to terminate states of viral persistence in humans. Mice persistently infected from birth with lymphocytic choriomeningitis virus are tolerant to the pathogen at the T-cell level and thus provide an excellent model to evaluate immunotherapeutic regimens. Previously, we demonstrated that adoptively transferred memory T cells require recipient dendritic cells to effectively purge an established persistent viral infection. However, the mechanisms that reactivate and sustain memory T-cell responses during clearance of such an infection remain unclear. Here we establish that therapeutic memory T cells require CD80 and CD86 costimulatory signals to efficiently clear an established persistent viral infection in vivo. Early blockade of costimulatory pathways with CTLA-4-Fc decreased the secondary expansion of virus-specific CD8+ and CD4+ memory T cells as well as their ability to produce antiviral cytokines and purge the persistent infection. Late costimulation blockade also reduced virus-specific T-cell numbers, illustrating that sustained interactions with costimulatory molecules is required for efficient T-cell expansion. These findings indicate that antiviral memory T cells require costimulation to efficiently clear a persistent viral infection and that costimulatory pathways can be targeted to modulate the magnitude of an adoptive immunotherapeutic regimen.Persistent viruses, such as human immunodeficiency virus, hepatitis B virus, and hepatitis C virus, cause major health problems worldwide and are extraordinarily difficult to clear following the establishment of persistence. Given the challenges associated with clearing persistent infections, it is important to develop and mechanistically understand therapeutic strategies that successfully achieve viral eradication without inducing permanent damage in the host. Studies using the lymphocytic choriomeningitis virus (LCMV) model system have convincingly demonstrated that a systemic persistent viral infection can be completely purged from a murine host by using a therapeutic approach referred to as adoptive immunotherapy (1, 15, 22, 29, 30). Remarkably, total body control of multiple persistent viral infections in both the mouse (1, 15, 22, 29, 30) and humans (8, 14, 24, 26, 31) can be achieved using adoptive immunotherapy. When mice are persistently infected at birth or in utero with LCMV (referred to as carrier mice), the virus establishes systemic persistence (6). Adult LCMV carrier mice are tolerant to the virus at the T-cell level and thus are unable to eradicate the pathogen (23), which provides an excellent model to study immunotherapeutic regimens. Immunocytotherapy relies on the adoptive transfer of virus-specific memory CD8 and CD4 T cells from LCMV-immune donor mice into recipient carrier mice (1, 15, 22, 29, 30). Following the therapeutic administration of memory cells, LCMV is purged from most peripheral tissues of carrier mice in 14 days, whereas more than 100 days are required to clear virus from the central nervous system (CNS) and kidneys (1, 15, 22). Furthermore, successful viral clearance requires antiviral “memory” but not “effector” T cells (11). Thus, in addition to its proven therapeutic relevance, this model also provides a paradigm to understand factors that regulate memory T cells following secondary exposure to pathogens in vivo.The mechanisms leading to activation of naïve T cells have been well described and involve recognition of major histocompatibility complex (MHC) peptide through the T-cell receptor (TCR) as well as costimulation (e.g., CD80 and CD86 interactions) (4, 25, 27). On the other hand, the factors that govern the activation and secondary expansion of memory CD8+ and CD4+ T cells are less clearly defined, particularly in an in vivo therapeutic setting. When memory T cells reencounter cognate antigen, they respond rapidly by producing cytokines and dividing. Previous studies indicated that there was no role for dendritic cells or costimulation (4, 27) in the reactivation of memory T cells; however, three recent studies have shown that dendritic cells (DCs) stimulate memory T-cell activity upon antigen rechallenge (2, 33) and during adoptive immunotherapy (15). Because MHC class I antigen (MHC-I) is expressed on nearly all cell types but costimulatory molecules are not, these three studies strongly suggested that DCs were influencing memory T cells with costimulatory pathways thought only to be required during priming. Indeed, when the issue was reexamined, it was revealed that memory CD8+ and CD4+ T cells require CD28-CD80/CD86 costimulation to be fully reactivated upon secondary exposure to antigen (3, 7, 21).Because therapeutically administered memory T cells require effective interactions with the host hematopoietic system (10), in particular dendritic cells (15), to achieve successful viral clearance, we set out to address several unanswered questions. First, is costimulation required for the immunotherapeutic clearance of an established persistent viral infection? This is a particularly important question because the requirements imposed on therapeutically administered memory T cells, which encounter immediate and overwhelmingly high levels of virus, heightened antigenic stimulation, and a unique inflammatory milieu, are likely to be different than those faced by endogenous memory T cells following pathogen rechallenge in an otherwise-quiescent environment. The second question we set out to address in this study was whether costimulation blockade could modulate the activities of an immunotherapeutic regimen consisting of memory T cells. This question is of great importance in a clinical setting where pathogen-specific memory T cells can induce severe tissue pathology through the release of effector molecules (12). Thus, it is critical to have a strategy to limit the magnitude of an undesirable response without impeding viral clearance.  相似文献   

13.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

14.
We previously reported that CD4C/human immunodeficiency virus (HIV)Nef transgenic (Tg) mice, expressing Nef in CD4+ T cells and cells of the macrophage/dendritic cell (DC) lineage, develop a severe AIDS-like disease, characterized by depletion of CD4+ T cells, as well as lung, heart, and kidney diseases. In order to determine the contribution of distinct populations of hematopoietic cells to the development of this AIDS-like disease, five additional Tg strains expressing Nef through restricted cell-specific regulatory elements were generated. These Tg strains express Nef in CD4+ T cells, DCs, and macrophages (CD4E/HIVNef); in CD4+ T cells and DCs (mCD4/HIVNef and CD4F/HIVNef); in macrophages and DCs (CD68/HIVNef); or mainly in DCs (CD11c/HIVNef). None of these Tg strains developed significant lung and kidney diseases, suggesting the existence of as-yet-unidentified Nef-expressing cell subset(s) that are responsible for inducing organ disease in CD4C/HIVNef Tg mice. Mice from all five strains developed persistent oral carriage of Candida albicans, suggesting an impaired immune function. Only strains expressing Nef in CD4+ T cells showed CD4+ T-cell depletion, activation, and apoptosis. These results demonstrate that expression of Nef in CD4+ T cells is the primary determinant of their depletion. Therefore, the pattern of Nef expression in specific cell population(s) largely determines the nature of the resulting pathological changes.The major cell targets and reservoirs for human immunodeficiency virus type 1 (HIV-1)/simian immunodeficiency virus (SIV) infection in vivo are CD4+ T lymphocytes and antigen-presenting cells (macrophages and dendritic cells [DC]) (21, 24, 51). The cell specificity of these viruses is largely dependent on the expression of CD4 and of its coreceptors, CCR5 and CXCR-4, at the cell surface (29, 66). Infection of these immune cells leads to the severe disease, AIDS, showing widespread manifestations, including progressive immunodeficiency, immune activation, CD4+ T-cell depletion, wasting, dementia, nephropathy, heart and lung diseases, and susceptibility to opportunistic pathogens, such as Candida albicans (1, 27, 31, 37, 41, 82, 93, 109). It is reasonable to assume that the various pathological changes in AIDS result from the expression of one or many HIV-1/SIV proteins in these immune target cells. However, assigning the contribution of each infected cell subset to each phenotype has been remarkably difficult, despite evidence that AIDS T-cell phenotypes can present very differently depending on the strains of infecting HIV-1 or SIV or on the cells targeted by the virus (4, 39, 49, 52, 72). For example, the T-cell-tropic X4 HIV strains have long been associated with late events and severe CD4+ T-cell depletion (22, 85, 96). However, there are a number of target cell subsets expressing CD4 and CXCR-4, and identifying which one is responsible for this enhanced virulence has not been achieved in vivo. Similarly, the replication of SIV in specific regions of the thymus (cortical versus medullary areas), has been associated with very different outcomes but, unfortunately, the critical target cells of the viruses were not identified either in these studies (60, 80). The task is even more complex, because HIV-1 or SIV can infect several cell subsets within a single cell population. In the thymus, double (CD4 CD8)-negative (DN) or triple (CD3 CD4 CD8)-negative (TN) T cells, as well as double-positive (CD4+ CD8+) (DP) T cells, are infectible by HIV-1 in vitro (9, 28, 74, 84, 98, 99, 110) and in SCID-hu mice (2, 5, 91, 94). In peripheral organs, gut memory CCR5+ CD4+ T cells are primarily infected with R5 SIV, SHIV, or HIV, while circulating CD4+ T cells can be infected by X4 viruses (13, 42, 49, 69, 70, 100, 101, 104). Moreover, some detrimental effects on CD4+ T cells have been postulated to originate from HIV-1/SIV gene expression in bystander cells, such as macrophages or DC, suggesting that other infected target cells may contribute to the loss of CD4+ T cells (6, 7, 32, 36, 64, 90).Similarly, the infected cell population(s) required and sufficient to induce the organ diseases associated with HIV-1/SIV expression (brain, heart, and kidney) have not yet all been identified. For lung or kidney disease, HIV-specific cytotoxic CD8+ T cells (1, 75) or infected podocytes (50, 95), respectively, have been implicated. Activated macrophages have been postulated to play an important role in heart disease (108) and in AIDS dementia (35), although other target cells could be infected by macrophage-tropic viruses and may contribute significantly to the decrease of central nervous system functions (11, 86, 97), as previously pointed out (25).Therefore, because of the widespread nature of HIV-1 infection and the difficulty in extrapolating tropism of HIV-1/SIV in vitro to their cell targeting in vivo (8, 10, 71), alternative approaches are needed to establish the contribution of individual infected cell populations to the multiorgan phenotypes observed in AIDS. To this end, we developed a transgenic (Tg) mouse model of AIDS using a nonreplicating HIV-1 genome expressed through the regulatory sequences of the human CD4 gene (CD4C), in the same murine cells as those targeted by HIV-1 in humans, namely, in immature and mature CD4+ T cells, as well as in cells of the macrophage/DC lineages (47, 48, 77; unpublished data). These CD4C/HIV Tg mice develop a multitude of pathologies closely mimicking those of AIDS patients. These include a gradual destruction of the immune system, characterized among other things by thymic and lymphoid organ atrophy, depletion of mature and immature CD4+ T lymphocytes, activation of CD4+ and CD8+ T cells, susceptibility to mucosal candidiasis, HIV-associated nephropathy, and pulmonary and cardiac complications (26, 43, 44, 57, 76, 77, 79, 106). We demonstrated that Nef is the major determinant of the HIV-1 pathogenicity in CD4C/HIV Tg mice (44). The similarities of the AIDS-like phenotypes of these Tg mice to those in human AIDS strongly suggest that such a Tg mouse approach can be used to investigate the contribution of distinct HIV-1-expressing cell populations to their development.In the present study, we constructed and characterized five additional mouse Tg strains expressing Nef, through distinct regulatory elements, in cell populations more restricted than in CD4C/HIV Tg mice. The aim of this effort was to assess whether, and to what extent, the targeting of Nef in distinct immune cell populations affects disease development and progression.  相似文献   

15.
16.
Cytotoxicity and proliferation capacity are key functions of antiviral CD8 T cells. In the present study, we investigated a series of markers to define these functions in virus-specific CD8 T cells. We provide evidence that there is a lack of coexpression of perforin and CD127 in human CD8 T cells. CD127 expression on virus-specific CD8 T cells correlated positively with proliferation capacity and negatively with perforin expression and cytotoxicity. Influenza virus-, cytomegalovirus-, and Epstein-Barr virus/human immunodeficiency virus type 1-specific CD8 T cells were predominantly composed of CD127+ perforin/CD127 perforin+, and CD127/perforin CD8 T cells, respectively. CD127/perforin and CD127/perforin+ cells expressed significantly more PD-1 and CD57, respectively. Consistently, intracellular cytokine (gamma interferon, tumor necrosis factor alpha, and interleukin-2 [IL-2]) responses combined to perforin detection confirmed that virus-specific CD8 T cells were mostly composed of either perforin+/IL-2 or perforin/IL-2+ cells. In addition, perforin expression and IL-2 secretion were negatively correlated in virus-specific CD8 T cells (P < 0.01). As previously shown for perforin, changes in antigen exposure modulated also CD127 expression. Based on the above results, proliferating (CD127+/IL-2-secreting) and cytotoxic (perforin+) CD8 T cells were contained within phenotypically distinct T-cell populations at different stages of activation or differentiation and showed different levels of exhaustion and senescence. Furthermore, the composition of proliferating and cytotoxic CD8 T cells for a given antiviral CD8 T-cell population appeared to be influenced by antigen exposure. These results advance our understanding of the relationship between cytotoxicity, proliferation capacity, the levels of senescence and exhaustion, and antigen exposure of antiviral memory CD8 T cells.Cytotoxic CD8 T cells are a fundamental component of the immune response against viral infections and mediate an important role in immunosurveillance (7, 10, 55), and the induction of vigorous CD8 T-cell responses after vaccination is thought to be a key component of protective immunity (37, 41, 49, 50, 58, 60, 69). Cytotoxic CD8 T cells exert their antiviral and antitumor activity primarily through the secretion of cytotoxic granules containing perforin (pore-forming protein) and several granule-associated proteases, including granzymes (Grms) (5, 15, 20, 44). Several studies have recently advanced the characterization of the mechanism of granule-dependent cytotoxic activity and performed a comprehensive investigation of the content of cytotoxic granules in human virus-specific CD8 T cells (2, 19, 29, 44, 53).Heterogeneous profiles of cytotoxic granules have been identified in different virus-specific memory CD8 T cells and associated with distinct differentiation stages of memory CD8 T cells (2, 19, 29, 44). Furthermore, we have observed a hierarchy among the cytotoxic granules in setting the efficiency of cytotoxic activity and demonstrated that perforin (and to a lesser extent GrmB) but not GrmA or GrmK were associated with cytotoxic activity (29). Recently, a novel mechanism of perforin-dependent granule-independent CTL cytotoxicity has also been demonstrated (45).Major advances in the characterization of antigen (Ag)-specific CD4 and CD8 T cells have been made recently and have aimed at identifying functional profiles that may correlate with protective CD8 T-cell responses (1, 3, 4, 12, 13, 24, 28, 36-38, 40, 41, 49, 50, 56-58, 60, 64, 68). In particular, the functional characterization of antigen-specific T cells was mainly performed on the basis of (i) the pattern of cytokines secreted (i.e., gamma interferon [IFN-γ], tumor necrosis factor alpha [TNF-α], interleukin-2 [IL-2], or macrophage inflammatory protein 1β [MIP-1β]), (ii) the proliferation capacity, and (iii) the cytotoxic capacity (13, 28, 59). Of note, degranulation activity (i.e., CD107a mobilization following specific stimulation) has been used as a surrogate marker of cytotoxic activity (11, 13).The term “polyfunctional” has been used to define T-cell immune responses that, in addition to typical effector functions such as secretion of IFN-γ, TNF-α, or MIP-1β and cytotoxic activity (measured by the degranulation capacity), comprise distinct T-cell populations able to secrete IL-2 and retain proliferation capacity (13, 28, 49, 50). Some evidence indicates that a hallmark of protective immune responses is the presence of polyfunctional T-cell responses (59). Furthermore, the ability to secrete IL-2 was shown to be linked to proliferation capacity, and both factors have been associated with protective antiviral immunity (13, 28, 49, 50). Although a lack of correlation between degranulation activity and GrmB expression was reported in mice (65), the relationship between degranulation activity and perforin expression has never been comprehensively investigated in mice and in humans.The private α chain of the IL-7 receptor (IL-7Rα, also called CD127) has been suggested to selectively identify CD8 T cells that will become long-lived memory cells (6, 34, 36). Moreover, it was shown in mice (34, 36) and humans (14, 48, 63) that the CD127high memory-precursor CD8 T cells produced IL-2 in contrast to CD127low effector CD8 T cells. Of interest, CD127 expression has also been shown to correlate with Ag-specific proliferation capacity in mice (34, 36). A similar correlation was observed in humans, although only for polyclonal stimulations (48). With the exception of studies performed in HIV-1 infection, where an association between CD127 expression and HIV-1 viremia has been shown (21, 22, 42, 48, 54), very limited information is available on the CD127 expression in human virus-specific CD8 T cells other that HIV-1.Although cytotoxic activity and proliferation capacity are key components of the antiviral cellular immune response, the relationship between these functions has been only investigated in nonprogressive HIV-1 infection (46), where these two functions were shown to be related. However, it still remains to be determined whether these functions are mediated by the same or by different T-cell populations.In the present study, we performed a comprehensive characterization of virus-specific CD8 T-cell responses against HIV-1, cytomegalovirus (CMV), Epstein Barr virus (EBV), and influenza virus (Flu) in order to (i) analyze the degree of concordance between degranulation activity and perforin/Grm expression; (ii) identify the relevance of CD127 in identifying virus-specific CD8 T cells endowed with proliferation capacity; (iii) delineate the relationship between proliferation capacity, cytotoxic activity, activation/differentiation stage, and level of exhaustion of CD8 T cells; and (iv) determine the influence of antigen exposure in shaping the functional composition of virus-specific CD8 T cells.Our data indicate that cytotoxic (as defined by perforin expression) and proliferating (as defined by CD127 expression or IL-2 secretion) virus-specific CD8 T cells are contained within distinct CD8 T-cell populations. Furthermore, the proportion of proliferating and cytotoxic T cells within a given virus-specific CD8 T-cell population appears to be influenced by antigen exposure. These results advance our understanding of the relationship between cytotoxicity, proliferative capacity, differentiation stage, and Ag exposure of memory CD8 T cells.  相似文献   

17.
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   

18.
The structural precursor polyprotein, Gag, encoded by all retroviruses, including the human immunodeficiency virus type 1 (HIV-1), is necessary and sufficient for the assembly and release of particles that morphologically resemble immature virus particles. Previous studies have shown that the addition of Ca2+ to cells expressing Gag enhances virus particle production. However, no specific cellular factor has been implicated as mediator of Ca2+ provision. The inositol (1,4,5)-triphosphate receptor (IP3R) gates intracellular Ca2+ stores. Following activation by binding of its ligand, IP3, it releases Ca2+ from the stores. We demonstrate here that IP3R function is required for efficient release of HIV-1 virus particles. Depletion of IP3R by small interfering RNA, sequestration of its activating ligand by expression of a mutated fragment of IP3R that binds IP3 with very high affinity, or blocking formation of the ligand by inhibiting phospholipase C-mediated hydrolysis of the precursor, phosphatidylinositol-4,5-biphosphate, inhibited Gag particle release. These disruptions, as well as interference with ligand-receptor interaction using antibody targeted to the ligand-binding site on IP3R, blocked plasma membrane accumulation of Gag. These findings identify IP3R as a new determinant in HIV-1 trafficking during Gag assembly and introduce IP3R-regulated Ca2+ signaling as a potential novel cofactor in viral particle release.Assembly of the human immunodeficiency virus (HIV) is determined by a single gene that encodes a structural polyprotein precursor, Gag (71), and may occur at the plasma membrane or within late endosomes/multivesicular bodies (LE/MVB) (7, 48, 58; reviewed in reference 9). Irrespective of where assembly occurs, the assembled particle is released from the plasma membrane of the host cell. Release of Gag as virus-like particles (VLPs) requires the C-terminal p6 region of the protein (18, 19), which contains binding sites for Alix (60, 68) and Tsg101 (17, 37, 38, 41, 67, 68). Efficient release of virus particles requires Gag interaction with Alix and Tsg101. Alix and Tsg101 normally function to sort cargo proteins to LE/MVB for lysosomal degradation (5, 15, 29, 52). Previous studies have shown that addition of ionomycin, a calcium ionophore, and CaCl2 to the culture medium of cells expressing Gag or virus enhances particle production (20, 48). This is an intriguing observation, given the well-documented positive role for Ca2+ in exocytotic events (33, 56). It is unclear which cellular factors might regulate calcium availability for the virus release process.Local and global elevations in the cytosolic Ca2+ level are achieved by ion release from intracellular stores and by influx from the extracellular milieu (reviewed in reference 3). The major intracellular Ca2+ store is the endoplasmic reticulum (ER); stores also exist in MVB and the nucleus. Ca2+ release is regulated by transmembrane channels on the Ca2+ store membrane that are formed by tetramers of inositol (1,4,5)-triphosphate receptor (IP3R) proteins (reviewed in references 39, 47, and 66). The bulk of IP3R channels mediate release of Ca2+ from the ER, the emptying of which signals Ca2+ influx (39, 51, 57, 66). The few IP3R channels on the plasma membrane have been shown to be functional as well (13). Through proteomic analysis, we identified IP3R as a cellular protein that was enriched in a previously described membrane fraction (18) which, in subsequent membrane floatation analyses, reproducibly cofractionated with Gag and was enriched in the membrane fraction only when Gag was expressed. That IP3R is a major regulator of cytosolic calcium concentration (Ca2+) is well documented (39, 47, 66). An IP3R-mediated rise in cytosolic Ca2+ requires activation of the receptor by a ligand, inositol (1,4,5)-triphosphate (IP3), which is produced when phospholipase C (PLC) hydrolyzes phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2] at the plasma membrane (16, 25, 54). Paradoxically, PI(4,5)P2 binds to the matrix (MA) domain in Gag (8, 55, 59), and the interaction targets Gag to PI(4,5)P2-enriched regions on the plasma membrane; these events are required for virus release (45). We hypothesized that PI(4,5)P2 binding might serve to target Gag to plasma membrane sites of localized Ca2+ elevation resulting from PLC-mediated PI(4,5)P2 hydrolysis and IP3R activation. This idea prompted us to investigate the role of IP3R in Gag function.Here, we show that HIV-1 Gag requires steady-state levels of IP3R for its efficient release. Three isoforms of IP3R, types 1, 2, and 3, are encoded in three independent genes (39, 47). Types 1 and 3 are expressed in a variety of cells and have been studied most extensively (22, 39, 47, 73). Depletion of the major isoforms in HeLa or COS-1 cells by small interfering RNA (siRNA) inhibited viral particle release. Moreover, we show that sequestration of the IP3R activating ligand or blocking ligand formation also inhibited Gag particle release. The above perturbations, as well as interfering with receptor expression or activation, led to reduced Gag accumulation at the cell periphery. The results support the conclusion that IP3R activation is required for efficient HIV-1 viral particle release.  相似文献   

19.
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   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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