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.
The early steps of the hepatitis B virus (HBV) life cycle are still poorly understood. Indeed, neither the virus receptor at the cell surface nor the mechanism by which nucleocapsids are delivered to the cytosol of infected cells has been identified. Extensive mutagenesis studies in pre-S1, pre-S2, and most of the S domain of envelope proteins revealed the presence of two regions essential for HBV infectivity: the 77 first residues of the pre-S1 domain and a conformational motif in the antigenic loop of the S domain. In addition, at the N-terminal extremity of the S domain, a putative fusion peptide, partially overlapping the first transmembrane (TM1) domain and preceded by a PEST sequence likely containing several proteolytic cleavage sites, was identified. Since no mutational analysis of these two motifs potentially implicated in the fusion process was performed, we decided to investigate the ability of viruses bearing contiguous deletions or substitutions in the putative fusion peptide and PEST sequence to infect HepaRG cells. By introducing the mutations either in the L and M proteins or in the S protein, we demonstrated the following: (i) that in the TM1 domain of the L protein, three hydrophobic clusters of four residues were necessary for infectivity; (ii) that the same clusters were critical for S protein expression; and, finally, (iii) that the PEST sequence was dispensable for both assembly and infection processes.The hepatitis B virus (HBV) is the main human pathogen responsible for severe hepatic diseases like cirrhosis and hepatocellular carcinoma. Even though infection can be prevented by immunization with an efficient vaccine, about 2 billion people have been infected worldwide, resulting in 350 million chronic carriers that are prone to develop liver diseases (56). Current treatments consist either of the use of interferon α, which modulates antiviral defenses and controls infection in 30 to 40% of cases, or of the use of viral polymerase inhibitors that allow a stronger response to treatment but require long-term utilization and frequently lead to the outcome of resistant viruses (34, 55). A better understanding of the virus life cycle, and particularly of the mechanism by which the virus enters the cell, could provide background for therapeutics that inhibit the early steps of infection, as recently illustrated with the HBV pre-S1-derived entry inhibitor (25, 45).HBV belongs to the Hepadnaviridae family whose members infect different species. All viruses of this family share common properties. The capsid containing a partially double-stranded circular DNA genome is surrounded by a lipid envelope, in which two (in avihepadnaviruses infecting birds) or three (in orthohepadnaviruses infecting mammals) envelope proteins are embedded. A single open reading frame bearing several translation initiation sites encodes these surface proteins. Thus, the HBV envelope contains three proteins: S, M, and L that share the same C-terminal extremity corresponding to the small S protein that is crucial for virus assembly (7, 8, 46) and infectivity (1, 31, 53). These proteins are synthesized in the endoplasmic reticulum (ER), assembled, and secreted as particles through the Golgi apparatus (15, 42). The current model for the transmembrane structure of the S domain implies the luminal exposition of both N- and C-terminal extremities and the presence of four transmembrane (TM) domains: the TM1 and TM2 domains, both necessary for cotranslational protein integration into the ER membrane, and the TM3 and TM4 domains, located in the C-terminal third of the S domain (for a review, see reference 6). Among the four predicted TM domains, only the TM2 domain has a defined position between amino acids 80 and 98 of the S domain. The exact localization of the TM1 domain is still unclear, probably because of the relatively low hydrophobicity of its sequence, which contains polar residues and two prolines. The M protein corresponds to the S protein extended by an N-terminal domain of 55 amino acids called pre-S2. Its presence is dispensable for both assembly and infectivity (20, 21, 37). Finally, the L protein corresponds to the M protein extended by an N-terminal domain of 108 amino acids called pre-S1 (genotype D). The pre-S1 and pre-S2 domains of the L protein can be present either at the inner face of viral particles (on the cytoplasmic side of the ER), playing a crucial role in virus assembly (5, 8, 10, 11, 46), or on the outer face (on the luminal side of the ER), available for the interaction with target cells and necessary for viral infectivity (4, 14, 36). The pre-S translocation is independent from the M and S proteins and is driven by the L protein TM2 domain (33). Finally, HBV surface proteins are not only incorporated into virion envelopes but also spontaneously bud from ER-Golgi intermediate compartment membranes (30, 43) to form empty subviral particles (SVPs) that are released from the cell by secretion (8, 40).One approach to decipher viral entry is to interfere with the function of envelope proteins. Thus, by a mutagenesis approach, two envelope protein domains crucial for HBV infectivity have already been identified: (i) the 77 first amino acids of the pre-S1 domain (4, 36) including the myristic acid at the N-terminal extremity (9, 27) and (ii) possibly a cysteine motif in the luminal loop of the S domain (1, 31). In addition, a putative fusion peptide has been identified at the N-terminal extremity of the S domain due to its sequence homology with other viral fusion peptides (50). This sequence, either N-terminal in the S protein or internal in the L and M proteins, is conserved among the Hepadnaviridae family and shares common structural and functional properties with other fusion peptides (49, 50). Finally, a PEST sequence likely containing several proteolytic cleavage sites has been identified in the L and M proteins upstream of the TM1 domain (39). A cleavage within this sequence could activate the fusion peptide.In this study, we investigated whether the putative fusion peptide and the PEST sequence were necessary for the infection process. For this purpose, we constructed a set of mutant viruses bearing contiguous deletions in these regions and determined their infectivity using an in vitro infection model based on HepaRG cells (28). The introduction of mutations either in the L and M proteins or in only the S protein allowed us to demonstrate that, in the TM1 domain of L protein, three hydrophobic clusters not essential for viral assembly were crucial for HBV infectivity while their presence in the S protein was critical for envelope protein expression. In addition, we showed that the PEST sequence was clearly dispensable for both assembly and infection processes.  相似文献   

2.
3.
4.
The hepatitis B virus (HBV) particles bear a receptor-binding site located in the pre-S1 domain of the large HBV envelope protein. Using the hepatitis delta virus (HDV) as a surrogate of HBV, a second infectivity determinant was recently identified in the envelope proteins antigenic loop (AGL), and its activity was shown to depend upon cysteine residues that are essential for the structure of the HBV immunodominant “a” determinant. Here, an alanine-scanning mutagenesis approach was used to precisely map the AGL infectivity determinant to a set of conserved residues, which are predicted to cluster together with cysteines in the AGL disulfide bridges network. Several substitutions suppressed both infectivity and the “a” determinant, whereas others were infectivity deficient with only a partial impact on antigenicity. Interestingly, G145R, a substitution often arising under immune pressure selection and detrimental to the “a” determinant, had no effect on infectivity. Altogether, these findings indicate that the AGL infectivity determinant is closely related to, yet separable from, the “a” determinant. Finally, a selection of HDV entry-deficient mutations were introduced at the surface of HBV virions and shown to also abrogate infection in the HBV model. Therefore, a function can at last be assigned to the orphan “a” determinant, the first-discovered marker of HBV infection. The characterization of the AGL functions at viral entry may lead to novel approaches in the development of antivirals against HBV.Hepatitis B virus (HBV) causes acute and chronic infections in humans; such infections are often associated with severe liver diseases, including cirrhosis and hepatocellular carcinoma (10). To date, it is estimated that approximately 350 millions individuals worldwide suffer from chronic infection despite the availability of an effective vaccine for more than 25 years. Remarkably, the development of a vaccine soon after the HBV discovery was, at least in part, the consequence of a very peculiar feature that is unique to members of the Hepadnaviridae family: viral envelope proteins are produced in quantities far exceeding the amounts required for assembly of HBV virions (6) and, owing to their capacity for autoassembly, the vast majority are secreted as subviral particles. Besides the practical consequences in the original vaccine development, in nature, the phenomenon of HBV envelope protein overexpression has provided a helper function to the hepatitis delta virus (HDV) (29). The HBV envelope proteins assist in packaging the HDV ribonucleoprotein (RNP) in case of HBV-HDV coinfection, thereby ensuring spreading of the satellite HDV. As a result, the coats of HBV and HDV particles are similar, consisting of cell-derived lipids and the HBV envelope proteins—large, middle, and small—bearing the HBV surface antigen (HBsAg) and referred to as L-HBsAg, M-HBsAg, and S-HBsAg, respectively (4, 14).The HBsAg includes an immunodominant determinant common to all HBV strains, referred to as “a,” and several mutually exclusive subtype-specific determinants referred to as “d”/“y” and “w”/“r” (21). The “a” determinant is defined by a specific conformation of the antigenic loop (AGL) polypeptide present at the surface of subviral, HBV, or HDV particles. The AGL itself resides between the transmembrane domain II (residues 80 to 100) and the hydrophobic carboxyl terminus (residues 165 to 226) of the envelope proteins S domain (see Fig. Fig.1).1). It is the “a” determinant that elicits the most effective neutralizing antibody response upon vaccination or infection (32). Surprisingly, a function in the HBV life cycle had never been assigned to the “a” determinant, the first identified HBV marker, until the recent demonstration of its involvement in HDV entry (2, 15). More precisely, it was shown that the AGL cysteine residues were critical for both the structure of the “a” determinant and HDV infectivity (2).Open in a separate windowFIG. 1.Schematic representation of the HBV envelope protein AGL. (A) The topology of the L-, M-, and S-HBsAg proteins (L, M, and S, respectively) is represented. The determinants of viral entry, pre-S1 and AGL, are indicated in red. The M-HBsAg protein, represented in gray, is dispensable for infectivity. Open boxes represent transmembrane regions in the S domain. (B) Alignment of the AGL amino acids sequences (positions 101 to 172 in the S domain) of HBV (genotype D, ayw3 phenotype), WMHBV, and WHV. The GenBank sequence numbers of the isolates are as follows: J02203 (HBV), AY226578 (WMHBV), and NC_004107 (WHV). HBV amino acid residues important for infectivity (the present study) are indicated in blue. A hyphen denotes amino acid identity with the HBV sequence.It is now well established that both HBV and HDV entry rely on the pre-S1 domain of L-HBsAg as the primary infectivity determinant that is likely to promote attachment to a specific receptor at the surface of human hepatocytes (11). The AGL determinant could thus fulfill complementary functions for attachment, uptake, or particle disassembly after entry (2, 15).In the present study, the AGL infectivity determinant was mapped and confirmed to be closely related to the “a” determinant. Moreover, its essential function at viral entry was demonstrated in the HBV model.  相似文献   

5.
Hepatits B virus (HBV)-specific T cells play a key role both in the control of HBV replication and in the pathogenesis of liver disease. Human immunodeficiency virus type 1 (HIV-1) coinfection and the presence or absence of HBV e (precore) antigen (HBeAg) significantly alter the natural history of chronic HBV infection. We examined the HBV-specific T-cell responses in treatment-naïve HBeAg-positive and HBeAg-negative HIV-1-HBV-coinfected (n = 24) and HBV-monoinfected (n = 39) Asian patients. Peripheral blood was stimulated with an overlapping peptide library for the whole HBV genome, and tumor necrosis factor alpha and gamma interferon cytokine expression in CD8+ T cells was measured by intracellular cytokine staining and flow cytometry. There was no difference in the overall magnitude of the HBV-specific T-cell responses, but the quality of the response was significantly impaired in HIV-1-HBV-coinfected patients compared with monoinfected patients. In coinfected patients, HBV-specific T cells rarely produced more than one cytokine and responded to fewer HBV proteins than in monoinfected patients. Overall, the frequency and quality of the HBV-specific T-cell responses increased with a higher CD4+ T-cell count (P = 0.018 and 0.032, respectively). There was no relationship between circulating HBV-specific T cells and liver damage as measured by activity and fibrosis scores, and the HBV-specific T-cell responses were not significantly different in patients with either HBeAg-positive or HBeAg-negative disease. The quality of the HBV-specific T-cell response is impaired in the setting of HIV-1-HBV coinfection and is related to the CD4+ T-cell count.There are 40 million people worldwide infected with human immunodeficiency virus type 1 (HIV-1), and 6 to 15% of HIV-1-infected patients are also chronically infected with hepatitis B virus (HBV) (13, 20, 35, 38, 40-42, 47, 50, 61, 69). The highest rates of coinfection with HIV-1 and HBV are in Asia and Africa, where HBV is endemic (33, 68). Following the introduction of highly active antiretroviral therapy (HAART), liver disease is now the major cause of non-AIDS-related deaths in HIV-1-infected patients (12, 13, 38, 59, 65).Coinfection of HBV with HIV-1 alters the natural history of HBV infection. Individuals with HIV-1-HBV coinfection seroconvert from HBV e (precore) antigen (HBeAg) to HBV e antibody less frequently and have higher HBV DNA levels but lower levels of alanine aminotransferase (ALT) and milder necroinflammatory activity on histology than those infected with HBV alone (18, 26, 49). Progression to cirrhosis, however, seems to be more rapid and more common, and liver-related mortality is higher, in HIV-1-HBV coinfection than with either infection alone (47, 59). HBeAg is an accessory protein of HBV and is not required for viral replication or infection; however, chronic HBV infection typically is divided into two distinct phases: HBeAg positive and HBeAg negative (reviewed in reference 15). Most natural history studies of HIV-1-HBV coinfection to date have primarily focused on HBeAg-positive patients from non-Asian countries (23, 44, 46).We previously developed an overlapping peptide library for the HBV genome to detect HBV-specific CD4+ and CD8+ T-cell responses to all HBV gene products from multiple HBV genotypes (17). In a small cross-sectional study of patients recruited in Australia, we found that in coinfected patients, HBV-specific CD4+ T-cell responses, as measured by gamma interferon (IFN-γ) production, were diminished compared to those seen in HBV-monoinfected patients (17). However, patients had varying lengths of exposure to anti-HBV-active HAART at the time of analysis. In this study, therefore, we aimed to characterize the HBV-specific T-cell response in untreated HBeAg-positive and HBeAg-negative HIV-1-HBV-coinfected patients and to determine the relationship between the HBV-specific immune response, HBeAg status, and liver disease.  相似文献   

6.
7.
8.
Hepatitis B and C viruses (HBV and HCV, respectively) are different and distinct viruses, but there are striking similarities in their disease potential. Infection by either virus can cause chronic hepatitis, liver cirrhosis, and ultimately, liver cancer, despite the fact that no pathogenetic mechanisms are known which are shared by the two viruses. Our recent studies have suggested that replication of either of these viruses upregulates a cellular protein called serine protease inhibitor Kazal (SPIK). Furthermore, the data have shown that cells containing HBV and HCV are more resistant to serine protease-dependent apoptotic death. Since our previous studies have shown that SPIK is an inhibitor of serine protease-dependent apoptosis, it is hypothesized that the upregulation of SPIK caused by HBV and HCV replication leads to cell resistance to apoptosis. The evasion of apoptotic death by infected cells results in persistent viral replication and constant liver inflammation, which leads to gradual accumulation of genetic changes and eventual development of cancer. These findings suggest a possibility by which HBV and HCV, two very different viruses, can share a common mechanism in provoking liver disease and cancer.Hepatitis B virus (HBV) and hepatitis C virus (HCV) infections are serious worldwide health problems, with more than 500 million people believed to be chronically infected with at least one of these viruses (36). HBV is a DNA virus belonging to the Hepadnaviridae family (21), while HCV is an RNA virus belonging to the Flaviviridae family (7). Despite the fact that they are two very different viruses, they share a common pathology in the ability to cause chronic hepatitis, liver cirrhosis, and ultimately, hepatocellular carcinoma (HCC) (34). It remains unclear why these two viruses, which are fundamentally so different, can both lead to similar disease states and the development of HCC.Numerous studies suggest that in chronic viral hepatitis, the host''s immune system is unable to clear infected cells (34). The persistent viral replication further stimulates liver inflammation, and prolonged inflammation and viral persistence result in a gradual accumulation of genetic changes which can subsequently lead to transformation and development of HCC (3, 13). It is possible that part of this failure of the host to clear infected cells results from an inability to induce apoptosis in these cells. For example, persistent HBV/HCV infection suppresses cytotoxic-T-lymphocyte (CTL)-induced apoptosis (3, 4). Apoptosis, or programmed cell death, plays a critical role in embryonic development, immune system function, and the overall maintenance of tissue homeostasis in multicellular organisms. It is also important in the host''s control of viral infection (4). The execution of the apoptotic program has traditionally been considered the result of the activation of a family of proteases known as caspases. Caspase-dependent cell apoptosis (CDCA) usually initiates by activating caspases 8 and 10 through proteolysis of their proenzymes, which further activates the executioner caspases, such as caspase 3 and caspase 7, resulting in the degradation of chromosomal DNA and cell death (28, 29). Recent evidence, however, has suggested that apoptotic cell death can also be promoted and triggered by serine proteases in a caspase-independent manner (5, 6, 39). Serine protease-dependent cell apoptosis (SPDCA) differs from CDCA in that serine proteases, not caspases, are critical to the apoptotic process (1, 6, 39). Interestingly, certain viral infections have been shown to induce SPDCA (27, 39).Failure of the immune-mediated removal of malignant cells through apoptosis may be due to the upregulation of apoptosis inhibitors in these cells (12, 18). We recently demonstrated that SPDCA can be inhibited by a small, 79-amino-acid protein called serine protease inhibitor Kazal (SPIK) (22). SPIK, which is also known as SPINK1, TATI (tumor-associated trypsin inhibitor), and PSTI (pancreas secretory trypsin inhibitor) (8, 24, 38), was first discovered in the pancreas as an inhibitor of autoactivation of trypsinogen (9). The expression of SPIK in normal tissue is limited or inactivated outside the pancreas, but expression of SPIK is elevated in numerous cancers, such as colorectal tumors, renal cell carcinoma, gastric carcinoma, and intrahepatic cholangiocarcinoma (ICC) (16, 19, 24, 31, 40, 41). It remains unknown, however, what role SPIK may play in cancer formation and development. Additionally, overexpression of SPIK was also found in HBV/HCV-infected human livers (32), and an even higher level of expression of SPIK was found in HBV/HCV-associated HCC tissue (19, 31). This implies that SPIK may be closely associated with hepatitis virus infection and development of HCC.Here we show direct evidence that HBV/HCV replication does in fact upregulate expression of the apoptosis inhibitor SPIK, resulting in resistance to SPDCA, which could ultimately lead to the development of chronic hepatitis and liver cancer.  相似文献   

9.
Alpha interferon (IFN-α) is an approved medication for chronic hepatitis B. Gamma interferon (IFN-γ) is a key mediator of host antiviral immunity against hepatitis B virus (HBV) infection in vivo. However, the molecular mechanism by which these antiviral cytokines suppress HBV replication remains elusive. Using an immortalized murine hepatocyte (AML12)-derived cell line supporting tetracycline-inducible HBV replication, we show in this report that both IFN-α and IFN-γ efficiently reduce the amount of intracellular HBV nucleocapsids. Furthermore, we provide evidence suggesting that the IFN-induced cellular antiviral response is able to distinguish and selectively accelerate the decay of HBV replication-competent nucleocapsids but not empty capsids in a proteasome-dependent manner. Our findings thus reveal a novel antiviral mechanism of IFNs and provide a basis for a better understanding of HBV pathobiology.Hepatitis B virus (HBV) is a noncytopathic hepatotropic DNA virus which belongs to the family Hepadnaviridae (11, 44). Despite the fact that most adulthood HBV infections are transient, approximately 5 to 10% of infected adults and more than 90% of infected neonates fail to clear the virus and develop a lifelong persistent infection, which may progress to chronic hepatitis, cirrhosis, and primary hepatocellular carcinoma (4, 33, 34). It has been shown by several research groups that resolution of HBV and other animal hepadnavirus infection in vivo depends on both killing of infected hepatocytes by viral antigen-specific cytotoxic T lymphocytes and noncytolytic suppression of viral replication, which is most likely mediated by inflammatory cytokines, such as gamma interferon (IFN-γ) and tumor necrosis factor α (TNF-α) (10, 12, 15, 20, 26, 27, 48). Moreover, together with five nucleoside or nucleotide analogs that inhibit HBV DNA polymerase, alpha IFN (IFN-α) and pegylated IFN-α are currently available antiviral medications for the management of chronic hepatitis B. Compared to the viral DNA polymerase inhibitors, the advantages of IFN-α therapy include a lack of drug resistance, a finite and defined treatment course, and an increased likelihood for hepatitis B virus surface antigen (HBsAg) clearance (8, 39). However, only approximately 30% of treated patients achieve a sustained virological response to a standard 48-month pegylated IFN-α therapy (6, 32). Thus far, the antiviral mechanism of IFN-α and IFN-γ and the parameters determining the success or failure of IFN-α therapy in chronic hepatitis B remain elusive. Elucidation of the mechanism by which the cytokines suppress HBV replication represents an important step toward understanding the pathobiology of HBV infection and the molecular basis of IFN-α therapy of chronic hepatitis B.Considering the mechanism by which IFNs noncytolytically control HBV infection in vivo, it is possible that the cytokines either induce an antiviral response in hepatocytes to directly limit HBV replication or modulate the host antiviral immune response to indirectly inhibit the virus infection. However, due to the fact that IFN-α and -γ do not inhibit or only modestly inhibit HBV replication in human hepatoma-derived cell lines (5, 22, 23, 30), the direct antiviral effects of the cytokines and their antiviral mechanism against HBV have been studied with either an immortalized hepatocyte cell line derived from HBV transgenic mice or duck hepatitis B virus (DHBV) infection of primary duck hepatocytes (37, 53). While these studies revealed that IFN treatment significantly reduced the amount of encapsidated viral pregenomic RNA (pgRNA) in both mouse and duck hepatocytes, further mechanistic analyses suggested that IFN-α inhibited the formation of pgRNA-containing nucleocapsids in murine hepatocytes (52) but shortened the half-life of encapsidated pgRNA in DHBV-replicating chicken hepatoma cells (21). Moreover, the fate of viral DNA replication intermediates or nucleocapsids in the IFN-treated hepatocytes was not investigated in the previous studies.To further define the target(s) of IFN-α and -γ in the HBV life cycle and to create a robust cell culture system for the identification of IFN-stimulated genes (ISGs) that mediate the antiviral response of the cytokines (25), we established an immortalized murine hepatocyte (AML-12)-derived stable cell line that supported a high level of HBV replication in a tetracycline-inducible manner. Consistent with previous reports, we show that both IFN-α and IFN-γ potently inhibited HBV replication in murine hepatocytes (37, 40). With the help of small molecules that inhibit HBV capsid assembly (Bay-4109) (7, 47) and prevent the incorporation of pgRNA into nucleocapsids (AT-61) (9, 29), we obtained evidence suggesting that the IFN-induced cellular antiviral response is able to distinguish and selectively accelerate the decay of HBV replication-competent nucleocapsids but not empty capsids in a proteasome-dependent manner. Our findings provide a basis for further studies toward better understanding of IFN′s antiviral mechanism, which might ultimately lead to the development of strategies to improve the efficacy of IFN therapy of chronic hepatitis B.  相似文献   

10.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

11.
12.
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.
15.
16.
17.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

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

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