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 46 毫秒
1.
Human immunodeficiency virus type 1 (HIV-1) elite controllers (EC) maintain viremia below the limit of commercial assay detection (<50 RNA copies/ml) in the absence of antiviral therapy, but the mechanisms of control remain unclear. HLA-B57 and the closely related allele B*5801 are particularly associated with enhanced control and recognize the same Gag240-249 TW10 epitope. The typical escape mutation (T242N) within this epitope diminishes viral replication capacity in chronically infected persons; however, little is known about TW10 epitope sequences in residual replicating viruses in B57/B*5801 EC and the extent to which mutations within this epitope may influence steady-state viremia. Here we analyzed TW10 in a total of 50 B57/B*5801-positive subjects (23 EC and 27 viremic subjects). Autologous plasma viral sequences from both EC and viremic subjects frequently harbored the typical cytotoxic T-lymphocyte (CTL)-selected mutation T242N (15/23 sequences [65.2%] versus 23/27 sequences [85.1%], respectively; P = 0.18). However, other unique mutants were identified in HIV controllers, both within and flanking TW10, that were associated with an even greater reduction in viral replication capacity in vitro. In addition, strong CTL responses to many of these unique TW10 variants were detected by gamma interferon-specific enzyme-linked immunospot assay. These data suggest a dual mechanism for durable control of HIV replication, consisting of viral fitness loss resulting from CTL escape mutations together with strong CD8 T-cell immune responses to the arising variant epitopes.A subset of human immunodeficiency virus type 1 (HIV-1)-infected persons who control viremia to below the limit of detection (<50 RNA copies/ml plasma) without antiviral therapy has been termed elite controllers/suppressors (EC) (2, 3, 6, 13, 32). Some of these individuals have been infected in excess of 30 years, indicating prolonged containment of HIV replication, but the mechanisms associated with this extreme viremia control remain elusive (13). Among EC, certain HLA class I alleles are overrepresented, in particular HLA-B57, strongly suggesting that HIV-1-specific cytotoxic T-lymphocyte (CTL) responses restricted by these alleles may be crucial for viremia control (16, 29, 32). However, to date, there has been no clear explanation as to why some subjects can control viremia but others cannot, even when carrying the same allegedly protective HLA alleles. Moreover, the characteristics of virus-specific immune responses as well as the impact of viral escape mutations on in vitro replicative fitness in persons with different disease outcomes remain unclear.Growing numbers of studies suggest that CTL targeting Gag, particularly the p24 capsid protein, play an important role in controlling viremia (7, 15, 22, 26, 32, 33, 38). Indeed, the most protective HLA class I allele, B57, which is present in over 40% of EC (32), restricts four immunodominant CTL epitopes in the p24 capsid protein. Previous studies have failed to find differences in the recognition of Gag epitopes or in gamma interferon (IFN-γ) responses to HIV proteins between B57-positive (B57+) long-term nonprogressors and B57+ progressors (28). Other studies have shown differences in the frequency of polyfunctional CD8+ T cells between B57+ EC and B57+ progressors (5); likewise, differences in the frequency of IFN-γ/interleukin-2-producing CD8+ T cells between controllers and progressors with protective HLA alleles were reported (16). Recently, Bailey et al. reported that plasma viruses in B57+ EC can harbor CTL escape mutations in the Gag protein, and in some cases these autologous variants were recognized by CTL (3). However, since there were no comparisons to progressors, it is unclear whether the viral variants that were detected or the apparent de novo CTL responses to the variant viruses are characteristic features among B57+ persons who maintain persistent control.Of the four immunodominant Gag CTL epitopes restricted by HLA-B57, TW10 (TSTLQEQIGW [Gag residues 240 to 249]) is known to be the earliest target in acute infection (1, 11, 36), therefore likely playing an important role in defining the plasma viral load set point. This epitope is also known to be presented by the closely related B*5801 allele, which is also associated with viral control (21). One of the most frequently detected mutations within this epitope, T242N, is known to occur rapidly and almost universally after acute infection in persons expressing HLA-B57/B*5801 (11, 17, 23). The same mutation has been shown to have a negative impact on viral replication capacity (VRC) by both clinical observation and in vitro experiments (8, 23, 25). Moreover, as plasma viral load increases, compensatory mutations accumulate, restoring VRC to some extent (8). Additional studies, predominantly with children, indicated that some TW10 escape variants may be targeted by specific immune responses (17). Together, these data suggest a hypothesis to explain the diverse disease courses among B57+ subjects, namely, that a combination of fitness cost by CTL escape from the TW10 response, variable accumulation of compensatory mutations, and variable generation of specific CTL responses to the new variant influence plasma viral loads.In this study, we investigated plasma viral sequences and IFN-γ-specific enzyme-linked immunospot (ELISPOT) assay responses to autologous Gag TW10 sequences in HLA-B57/B*5801-positive EC and compared these data to those obtained from persons with detectable viremia. Our results indicate that the TW10 T242N mutation does not differentiate HLA-B57/B*5801 EC from those with viremia and that CTL responses to this variant epitope are frequently detected in both viremic and aviremic subjects. However, some rare variants within and flanking this epitope were observed exclusively in HIV controllers, most of which not only reduced VRC but also were recognized by specific CTL at a high magnitude. These data suggest that the additive effects of both CTL-mediated selection for less fit viral variants and CD8 T-cell responses to the variant viruses contribute to strict viremia control in HLA-B57/B*5801-positive controllers.  相似文献   

2.
A broad Gag-specific CD8+ T-cell response is associated with effective control of adult human immunodeficiency virus (HIV) infection. The association of certain HLA class I molecules, such as HLA-B*57, -B*5801, and -B*8101, with immune control is linked to mutations within Gag epitopes presented by these alleles that allow HIV to evade the immune response but that also reduce viral replicative capacity. Transmission of such viruses containing mutations within Gag epitopes results in lower viral loads in adult recipients. In this study of pediatric infection, we tested the hypothesis that children may tend to progress relatively slowly if either they themselves possess one of the protective HLA-B alleles or the mother possesses one of these alleles, thereby transmitting a low-fitness virus to the child. We analyzed HLA type, CD8+ T-cell responses, and viral sequence changes for 61 mother-child pairs from Durban, South Africa, who were monitored from birth. Slow progression was significantly associated with the mother or child possessing one of the protective HLA-B alleles, and more significantly so when the protective allele was not shared by mother and child (P = 0.007). Slow progressors tended to make CD8+ T-cell responses to Gag epitopes presented by the protective HLA-B alleles, in contrast to progressors expressing the same alleles (P = 0.07; Fisher''s exact test). Mothers expressing the protective alleles were significantly more likely to transmit escape variants within the Gag epitopes presented by those alleles than mothers not expressing those alleles (75% versus 21%; P = 0.001). Reversion of transmitted escape mutations was observed in all slow-progressing children whose mothers possessed protective HLA-B alleles. These data show that HLA class I alleles influence disease progression in pediatric as well as adult infection, both as a result of the CD8+ T-cell responses generated in the child and through the transmission of low-fitness viruses by the mother.Human immunodeficiency virus (HIV)-specific CD8+ T cells play a central role in controlling viral replication (12). It is the specificity of the CD8+ T-cell response, particularly the response to Gag, that is associated with low viral loads in HIV infection (7, 17, 34). Although immune control is undermined by the selection of viral mutations that prevent recognition by the CD8+ T cells, evasion of Gag-specific responses mediated by protective class I HLA-B alleles typically brings a reduction in viral replicative capacity, facilitating subsequent immune control of HIV (2, 20, 21). The same principle has been demonstrated in studies of simian immunodeficiency virus infection (18, 22).Recent studies showed that the class I HLA-B alleles that protect against disease progression present more Gag-specific CD8+ T-cell epitopes and drive the selection of more Gag-specific escape mutations than those alleles that are associated with high viral loads (23). These protective HLA-B alleles not only are beneficial to infected individuals expressing those alleles but also benefit a recipient following transmission, since the transmitted virus carrying multiple Gag escape mutations may have substantially reduced fitness (3, 4, 8). However, there is no benefit to the recipient if he or she shares the same protective allele as the donor because the transmitted virus carries escape mutations in the Gag epitopes that would otherwise be expected to mediate successful immune control in the recipient (8, 11).The sharing of HLA alleles between donor and recipient occurs frequently in mother-to-child transmission (MTCT). The risk of MTCT is related to viral load in the mother, and a high viral load is associated with nonprotective alleles, such as HLA-B*18 and -B*5802. This may contribute in two distinct ways to the more rapid progression observed in pediatric HIV infection (24, 26, 27). First, because infected children share 50% or more of their HLA alleles with the transmitting mother, they are less likely than adults to carry protective HLA alleles (16). Thus, infected children as a group carry fewer protective HLA alleles and more nonprotective HLA alleles. Second, even when the child has a protective allele, such as HLA-B*27, this allele does not offer protection if the maternally transmitted virus carries escape mutations within the key Gag epitopes that are presented by the protective allele (11, 19).However, it is clear that infected children who possess protective alleles, such as HLA-B*27 or HLA-B*57, can achieve durable immune control of HIV infection if the virus transmitted from the mother is not preadapted to those alleles (6, 10). HIV-specific CD8+ T-cell responses are detectable from birth in infected infants (32). Furthermore, as in adult infection (3, 8), HIV-infected children have the potential to benefit from transmission of low-fitness viruses in the situation where the mother possesses protective HLA alleles and the child does not share those protective alleles. MTCT of low-fitness viruses carrying CD8+ T-cell escape mutations was recently documented (28; J. Prado et al., unpublished data).In this study, undertaken in Durban, South Africa, we set out to test the hypothesis that HIV-infected children are less likely to progress rapidly to disease if either the infected child or the transmitting mother possesses a protective HLA allele that is not shared. The HLA alleles most strongly associated with low viral loads and high CD4 counts in a cohort of >1,200 HIV-infected adults in Durban are HLA-B*57 (-B*5702 and -B*5703), HLA-B*5801, and HLA-B*8101 (16; A. Leslie et al., unpublished data). These four alleles all present Gag-specific CD8+ T-cell epitopes, and in each case the escape mutations selected in these epitopes reduce viral replicative capacity (2-4, 8, 21, 23).Analyzing a previously described cohort of 61 HIV-infected children in Durban (24, 26, 32), South Africa, who were all monitored from birth, we first addressed the question of whether possession of any of these four alleles by either mother or child is associated with slower disease progression in the child and then determined whether sharing of protective alleles by mother and child affects the ability of the child to make the Gag-specific CD8+ T-cell responses restricted by the shared allele.  相似文献   

3.
Previous studies have identified a central role for HLA-B alleles in influencing control of HIV infection. An alternative possibility is that a small number of HLA-B alleles may have a very strong impact on HIV disease outcome, dominating the contribution of other HLA alleles. Here, we find that even following the exclusion of subjects expressing any of the HLA-B class I alleles (B*57, B*58, and B*18) identified to have the strongest influence on control, the dominant impact of HLA-B alleles on virus set point and absolute CD4 count variation remains significant. However, we also find that the influence of HLA on HIV control in this C-clade-infected cohort from South Africa extends beyond HLA-B as HLA-Cw type remains a significant predictor of virus and CD4 count following exclusion of the strongest HLA-B associations. Furthermore, there is evidence of interdependent protective effects of the HLA-Cw*0401-B*8101, HLA-Cw*1203-B*3910, and HLA-A*7401-B*5703 haplotypes that cannot be explained solely by linkage to a protective HLA-B allele. Analysis of individuals expressing both protective and detrimental alleles shows that even the strongest HLA alleles appear to have an additive rather than dominant effect on HIV control at the individual level. Finally, weak but significant frequency-dependent effects in this cohort can be detected only by looking at an individual''s combined HLA allele frequencies. Taken together, these data suggest that although individual HLA alleles, particularly HLA-B, can have a strong impact, HIV control overall is likely to be influenced by the additive effect of some or all of the other HLA alleles present.HIV-specific CD8+ T cells play a central role in resolution of primary viremia and the long-term suppression of viral replication (13). Supporting this notion is the observed correlation between possession of particular human leukocyte antigen (HLA) class I alleles and control of HIV, measured both directly by time-to-AIDS (5, 6) and indirectly via clinical markers of disease progression (viral load [VL] and CD4 count) (15, 26, 28). Specific HLA class I alleles have been associated with relatively successful control of viral replication and slow disease progression, most notably, alleles HLA-B*57 and HLA-B*27 (1, 7, 12, 15, 21, 23), and also with relatively ineffective control of viral replication and rapid disease progression [B*35(Px), B*5802, and B*18] (5, 15, 17, 23). In addition, general trends suggesting an HLA class I heterozygote advantage (5) and rare allele advantage (28) and, most recently, a correlation between levels of surface expression linked to certain HLA-Cw alleles (11, 27) and HIV control has also been described.Among the different HLA class I loci, the HIV-specific CD8+ T-cell responses restricted by HLA-B alleles are thought to play the central role in determining disease outcome: the majority of detectable HIV-specific CD8+ T-cell responses are restricted by HLA-B alleles (3, 15, 16), HLA-B-restricted responses typically express a more effective “polyfunctional” phenotype (14), the strongest HLA-associations with either slow or rapid progression are with HLA-B alleles (5, 10, 11, 15), and HLA-B-restricted CD8+ T cells exert the strongest selection pressure on the virus (15, 19, 24). However, whether this apparent association between HIV immune control and HLA-B is a general and causal trend or, rather, is biased by the coincidence that the strongest HLA associations with either extreme of disease control happen, by chance, to involve HLA-B alleles remains uncertain.In order to further investigate the correlation between HLA type and HIV infection control, we here examine a cohort now comprising >1,200 chronically HIV C-clade-infected, treatment-naïve subjects from Durban, South Africa, in an extended analysis following from our previous studies of a smaller cohort (15). We first address the question of whether the dominant role of HLA-B in this population compared to the roles of HLA-A or HLA-C results from the influence of HLA-B alleles in general or is dependent on a few known strong associations, such as that between HLA-B*57 alleles and low viremia. Second, in light of recent data (11, 27), we assess the impact of HLA-C alleles on HIV disease outcome and examine the effect of HLA haplotypes on observed HLA associations with disease control. Third, we investigate the question of whether the impact of certain HLA-B alleles on HIV outcome dominates that of other HLA-B alleles to negate the contribution of the latter or whether the impact of individual HLA alleles can be additive. Finally, we compare the impact of individual HLA alleles on HIV on immune control to the impact of heterozygote and rare allele advantage in this cohort.  相似文献   

4.
The control of human immunodeficiency virus type 1 (HIV-1) associated with particular HLA class I alleles suggests that some CD8+ T-cell responses may be more effective than others at containing HIV-1. Unfortunately, substantial diversities in the breadth, magnitude, and function of these responses have impaired our ability to identify responses most critical to this control. It has been proposed that CD8 responses targeting conserved regions of the virus may be particularly effective, since the development of cytotoxic T-lymphocyte (CTL) escape mutations in these regions may significantly impair viral replication. To address this hypothesis at the population level, we derived near-full-length viral genomes from 98 chronically infected individuals and identified a total of 76 HLA class I-associated mutations across the genome, reflective of CD8 responses capable of selecting for sequence evolution. The majority of HLA-associated mutations were found in p24 Gag, Pol, and Nef. Reversion of HLA-associated mutations in the absence of the selecting HLA allele was also commonly observed, suggesting an impact of most CTL escape mutations on viral replication. Although no correlations were observed between the number or location of HLA-associated mutations and protective HLA alleles, limiting the analysis to mutations selected by acute-phase immunodominant responses revealed a strong positive correlation between mutations at conserved residues and protective HLA alleles. These data suggest that control of HIV-1 may be associated with acute-phase CD8 responses capable of selecting for viral escape mutations in highly conserved regions of the virus, supporting the inclusion of these regions in the design of an effective vaccine.Despite substantial advances in antiretroviral therapies, development of an effective human immunodeficiency virus type 1 (HIV-1) vaccine remains a critical goal (6, 39, 82). Unfortunately, current vaccine efforts have failed to reduce infection rates in humans (9, 75) and have only achieved modest decreases in viral loads in the simian immunodeficiency virus (SIV)/SHIV macaque model (21, 44, 81). A majority of these vaccine approaches have focused on inducing T-cell responses, utilizing large regions of the virus in an attempt to induce a broad array of immune responses (6, 34, 44, 81). While it is well established that CD8+ T-cell responses play a critical role in the containment of HIV-1 (45, 49, 67), supported in part by the strong association of particular HLA class I alleles with control of HIV (20, 33, 42, 61), it remains unclear which particular CD8+ T-cell responses are best able to control the virus and thus should be preferentially targeted by a vaccine. Studies comparing the magnitude, breadth, and function of CD8+ T-cell responses in subjects exhibiting either enhanced or poor control of HIV-1 have yielded few clues as to the specific factors associated with an effective CD8+ T-cell response (2, 28, 64, 67). Various differences in the functional capacity of T-cell responses have been observed in long-term nonprogressors (1, 26, 64), although it is possible that these differences may be reflective of an intact immune response, as opposed to having had directly enhanced immune control. As such, efforts are needed to identify factors or phenotypes associated with protective CD8+ T-cell responses in order to enable vaccines to induce the most effective responses.Recent studies have begun to suggest that the specificity of the CD8+ T-cell response, or the targeting of specific regions of the virus, may be associated with control of HIV-1. Preferential targeting of Gag, a structurally conserved viral protein responsible for multiple functions, has been associated with lower viral loads (25, 43, 56, 60, 77, 85). Furthermore, Kiepiela et al. (43) recently illustrated in a large cohort of 578 clade C-infected subjects that Gag-specific responses were associated with lowered viremia, in contrast to Env-specific responses, which were associated with higher viremia. These data are in line with previous observations that many of the major histocompatibility complex (MHC) class I alleles most strongly associated with control of HIV-1 and SIV, namely, HLA-B57, HLA-B27, and Mamu-A*01, restrict immunodominant CD8+ T-cell responses against the Gag protein (8, 10, 24, 63, 68, 83). However, other alleles associated with slower disease progression, such as HLA-B51 in humans and Mamu-B08 and B-17 in the rhesus macaque, do not immunodominantly target Gag, suggesting that targeting of some other regions of the virus may also be capable of eliciting control (8, 52-54). In addition, recent studies investigating the pattern of HIV-1-specific CD8+ T-cell responses during acute infection reveal that only a small subset of CD8+ T-cell responses restricted by any given HLA allele arise during acute infection and that there exist clear immunodominance patterns to these responses (8, 77, 85). Since control of HIV-1 is likely to be established or lost during the first few weeks of infection, these data suggest that potentially only a few key CD8+ T-cell responses may be needed to adequately establish early control of HIV-1.One of the major factors limiting the effectiveness of CD8+ T-cell responses is the propensity for HIV-1 to evade these responses through sequence evolution or viral escape (3, 13, 66). Even single point mutations within a targeted CD8 epitope can effectively abrogate recognition by either the HLA allele or the T-cell receptor. However, recent studies have begun to highlight that many sequence polymorphisms will revert to more common consensus residues upon transmission of HIV-1 to a new host, including many cytotoxic T-lymphocyte (CTL) escape mutations (4, 30, 33, 48, 50). Notably, the more rapidly reverting mutations have been observed to preferentially occur at conserved residues, indicating that structurally conserved regions of the virus may be particularly refractory to sequence changes (50). In support of these data, many CTL escape mutations have now been observed to directly impair viral replication (15, 23, 55, 74), in particular those known to either revert or require the presence of secondary compensatory mutations (15, 23, 73, 74). Taken together, these data suggest that, whereas CTL escape mutations provide a benefit to the virus to enable the evasion of host immune pressures, some of these mutations may come at a substantial cost to viral replication. These data may also imply that the association between Gag-specific responses and control of HIV-1 may be due to the targeting of highly conserved regions of the virus that are difficult to evade through sequence evolution.The propensity by which HIV-1 escapes CD8+ T-cell responses, and the reproducibility by which mutations arise at precise residues in targeted CD8 epitopes (3, 48), also enables the utilization of sequence data to predict which responses may be most capable of exerting immune selection pressure on the virus. Studies in HIV-1, SIV, and hepatitis C virus (16, 58, 65, 78) are now rapidly identifying immune-driven CTL escape mutations across these highly variable pathogens at the population level by correlating sequence polymorphisms in these viruses with the expression of particular HLA alleles. We provide here an analysis of HLA-associated mutations across the entire HIV-1 genome using a set of sequences derived from clade B chronically infected individuals. Through full-length viral genome coverage, these data provide an unbiased analysis of the location of these mutations and suggest that the control of HIV-1 by particular HLA alleles correlates with their ability to preferentially restrict early CD8+ T-cell responses capable of selecting for viral escape mutations at highly conserved residues of the virus. These data provide support for the inclusion of specific highly conserved regions of HIV-1 into vaccine antigens.  相似文献   

5.
The mechanisms underlying HIV-1 control by protective HLA class I alleles are not fully understood and could involve selection of escape mutations in functionally important Gag epitopes resulting in fitness costs. This study was undertaken to investigate, at the population level, the impact of HLA-mediated immune pressure in Gag on viral fitness and its influence on HIV-1 pathogenesis. Replication capacities of 406 recombinant viruses encoding plasma-derived Gag-protease from patients chronically infected with HIV-1 subtype C were assayed in an HIV-1-inducible green fluorescent protein reporter cell line. Viral replication capacities varied significantly with respect to the specific HLA-B alleles expressed by the patient, and protective HLA-B alleles, most notably HLA-B*81, were associated with lower replication capacities. HLA-associated mutations at low-entropy sites, especially the HLA-B*81-associated 186S mutation in the TL9 epitope, were associated with lower replication capacities. Most mutations linked to alterations in replication capacity in the conserved p24 region decreased replication capacity, while most in the highly variable p17 region increased replication capacity. Replication capacity also correlated positively with baseline viral load and negatively with baseline CD4 count but did not correlate with the subsequent rate of CD4 decline. In conclusion, there is evidence that protective HLA alleles, in particular HLA-B*81, significantly influence Gag-protease function by driving sequence changes in Gag and that conserved regions of Gag should be included in a vaccine aiming to drive HIV-1 toward a less fit state. However, the long-term clinical benefit of immune-driven fitness costs is uncertain given the lack of correlation with longitudinal markers of disease progression.There is broad heterogeneity in the ability of HIV-infected individuals to control virus replication, ranging from elite controllers, who maintain undetectable viral loads without treatment, to rapid progressors, who progress to AIDS within 2 years of infection (9, 22, 32). Many interrelated factors, including host and viral genetic factors involved in antiviral immunity and the viral life cycle, may partially account for the differences in the course of disease progression (10, 11, 30, 41). The complex interplay between host genetic factors and viral factors is exemplified by human leukocyte antigen (HLA) class I-restricted cytotoxic T-lymphocyte (CTL) responses, which exert considerable immune pressure on the virus, resulting in escape mutations that affect the interaction of viral and host proteins, thereby influencing infection outcome.The exact mechanisms by which some HLA class I alleles, such as HLA-B*57 and HLA-B*27, are associated with slower progression to AIDS, while others, such as B*5802 and B*18, are associated with accelerated disease progression (6, 20, 42), are unclear. The magnitude and/or breadth of HLA-restricted CTL responses to the conserved Gag protein has been correlated inversely with disease progression or markers of disease progression in several studies (12, 21, 28, 31, 35, 43, 46), although there are some exceptions (4, 16, 37), while preferential targeting of the highly variable envelope protein (as occurs in HLA-B*5802-positive individuals) correlates with higher viral loads (21, 29). Protective HLA alleles restrict CTL responses that impose a strong selection pressure on a few specific Gag p24 epitopes, resulting in escape mutations (14) for which fitness costs have been demonstrated either through site-directed mutations introduced into a reference strain background (2, 8, 25, 38) or through in vivo reversion of these mutations after transmission to an HLA-mismatched individual (8, 24). Recent evidence suggests that Gag escape mutations with a fitness cost, particularly those in p24, are a significant determinant of disease progression: the transmitted number of HLA-B-associated polymorphisms in Gag was found to significantly impact the viral set point in recipients (although an associated fitness cost was not shown) (7, 15), and in a small number of infants, decreased fitness of the transmitted virus with HLA-B*5703/5801-selected mutations in Gag p24 epitopes resulted in slower disease progression (33, 39). Also, the number of reverting Gag mutations (thought to revert as a consequence of fitness costs) associated with individual HLA-B alleles was strongly correlated with the HLA-linked viral set point in chronically infected patients (26). A recent in vitro study showed that HLA-associated variation in Gag-protease, with resulting reduced replication capacity, may contribute to viral control in HIV-1 subtype B-infected elite controllers (27). Taken together, these studies suggest that CTL responses restricted by favorable HLA alleles select for escape mutations in conserved epitopes, particularly those in Gag, resulting in a fitness cost to HIV and therefore at least partly explaining the slower disease progression in individuals carrying these alleles.To date, many of the studies investigating the fitness cost of Gag escape mutations and their clinical relevance have concentrated on escape mutations associated with protective HLA alleles, have not assessed fitness consequences in the natural sequence background (in the presence of other escape and compensatory mutations), and/or have focused on a limited number of patients. Most importantly, the majority of studies have focused on HIV-1 subtype B. The present study is the first to use a large population-based approach and clinically derived Gag-protease sequences to investigate comprehensively the relationships between immune-driven sequence variation in Gag, viral replication capacity, and markers of disease progression in chronic infection with HIV-1 subtype C, the most predominant subtype in the epidemic. We assayed the replication capacity of recombinant viruses encoding patient Gag-protease in an HIV-1-inducible green fluorescent protein (GFP) reporter cell line and found associations between lower replication capacities, protective HLA alleles, protective HLA-associated mutations, lower baseline viral loads, and higher baseline CD4 counts. However, Gag-protease replication capacity did not correlate with the subsequent rate of CD4 decline.  相似文献   

6.
Human immunodeficiency virus type 1 (HIV-1) controllers maintain viremia at <2,000 RNA copies/ml without antiretroviral therapy. Viruses from controllers with chronic infection were shown to exhibit impaired replication capacities, in part associated with escape mutations from cytotoxic-T-lymphocyte (CTL) responses. In contrast, little is known about viruses during acute/early infection in individuals who subsequently become HIV controllers. Here, we examine the viral replication capacities, HLA types, and virus sequences from 18 HIV-1 controllers identified during primary infection. gag-protease chimeric viruses constructed using the earliest postinfection samples displayed significantly lower replication capacities than isolates from persons who failed to control viremia (P = 0.0003). Protective HLA class I alleles were not enriched in these early HIV controllers, but viral sequencing revealed a significantly higher prevalence of drug resistance mutations associated with impaired viral fitness in controllers than in noncontrollers (6/15 [40.0%] versus 10/80 [12.5%], P = 0.018). Moreover, of two HLA-B57-positive (B57+) controllers identified, both harbored, at the earliest time point tested, signature escape mutations within Gag that likewise impair viral replication capacity. Only five controllers did not express “protective” alleles or harbor viruses with drug resistance mutations; intriguingly, two of them displayed typical B57 signature mutations (T242N), suggesting the acquisition of attenuated viruses from B57+ donors. These data indicate that acute/early stage viruses from persons who become controllers have evidence of reduced replication capacity during the initial stages of infection which is likely associated with transmitted or acquired CTL escape mutations or transmitted drug resistance mutations. These data suggest that viral dynamics during acute infection have a major impact on HIV disease outcome.Human immunodeficiency virus type I (HIV-1)-infected individuals who control viremia spontaneously without antiviral therapy have been termed HIV controllers (3, 18, 21, 48, 52). Unraveling the mechanisms associated with this phenotype should provide important insights regarding HIV pathogenesis and could contribute to vaccine development.Host and viral genetics, as well as host innate and adaptive immune responses, influence the rate of disease progression in HIV-1 infection (reviewed in reference 18). Several studies have reported the correlation between in vitro HIV replication capacity and level of plasma virus loads or disease progression in individuals with chronic infection (6, 13, 35, 45, 50, 55). Studies of HIV-1 elite controllers (EC), who control viremia to below the limit of detection in commercial assays, have revealed the presence of replication-competent viruses in these individuals (7), although these viruses appear to be less fit based on studies of envelope (35) and Gag-protease (45). This fitness defect in the chronic phase of infection is due at least in part to fitness-impairing mutations induced by cytotoxic-T-lymphocyte (CTL) responses restricted by “protective” HLA class I alleles (46).In contrast, little is known about viruses obtained from the acute/early phase of infection in persons who subsequently become HIV-1 controllers, largely due to the difficulty in enrolling such people during the acute/early phase of infection. The characterization of acute/early-phase viruses in individuals who subsequently achieve low set-point virus loads is of paramount importance to our understanding of the mechanisms of HIV-1 control.In the present study, we analyzed acute/early-phase plasma HIV RNA sequences from 18 untreated individuals who were diagnosed during the acute/early phase and subsequently became controllers (<2,000 RNA copies/ml). We compared these to sequences from a group of HIV-1 noncontrollers enrolled similarly during acute/early infection. We also generated chimeric viruses carrying patient-derived gag-protease sequences from acute/early-phase infection and compared the viral replication capacities of the chimeric viruses from controllers and from noncontrollers.We observed that the chimeric viruses derived from controllers have significantly reduced replicative capacities compared to those from noncontrollers. Moreover, we observed that at least 80% of these individuals who go on to become controllers featured transmission of attenuated drug-resistant viruses, transmission of HLA-B57-restricted CTL escape variants to HLA-mismatched recipients, selection of attenuated CTL escape variants in HLA-B57-positive (B57+) recipients, or combinations of these factors. Taken together, these results indicate that the initial viral dynamics have a major influence on the subsequent course of disease.  相似文献   

7.
8.
9.
Human immunodeficiency virus type 1 (HIV-1) envelope protein (Env) is subject to both neutralizing antibody (NAb) and CD8 T-cell (cytotoxic T-lymphocyte [CTL]) immune pressure. We studied the reversion of the Env CTL escape mutant virus to the wild type and the relationship between the reversion of CTL mutations with N-linked glycosylation site (NLGS)-driven NAb escape in pigtailed macaques. Env CTL mutations either did not revert to the wild type or only transiently reverted 5 to 7 weeks after infection. The CTL escape mutant reversion was coincident, for the same viral clones, with the loss of NLGS mutations. At one site studied, both CTL and NLGS mutations were needed to confer NAb escape. We conclude that CTL and NAb escape within Env can be tightly linked, suggesting opportunities to induce effective multicomponent anti-Env immunity.CD8 T-cell responses against human immunodeficiency virus (HIV) have long been observed to select for viral variants that avoid cytotoxic T-lymphocyte (CTL) recognition (2, 5, 15, 18, 27). These immune escape mutations may, however, result in reduced replication competence (“fitness cost”) (11, 20, 26). CTL escape variants have been shown to revert to the wild type (WT) upon passage to major histocompatibility complex-mismatched hosts, both in macaques with simian immunodeficiency virus (SIV) or chimeric SIV/HIV (SHIV) infection (11, 12) and in humans with HIV type 1 (HIV-1) infection (1, 19).Most analyses of CTL escape and reversion have studied Gag CTL epitopes known to facilitate control of viremia (7, 14, 21, 30). Fewer analyses have studied Env-specific CTL epitopes. Recent sequencing studies suggest the potential for mutations within predicted HIV-1 Env-specific CTL epitopes to undergo reversion to the WT (16, 23). Env-specific CTL responses may, however, have less impact on viral control of both HIV-1 and SIV/SHIV than do Gag CTL responses (17, 24, 25), presumably reflecting either less-potent inhibition of viral replication or minimal fitness cost of escape (9).Serial viral escape from antibody pressure also occurs in both macaques and humans (3, 13, 28). Env is extensively glycosylated, and this “evolving glycan shield” can sterically block antibody binding without mutation at the antibody-binding site (8, 16, 31). Mutations at glycosylation sites, as well as other mutations, are associated with escape from neutralizing antibody (NAb) responses (4, 13, 29). Mutations in the amino acid sequences of N-linked glycosylation sites (NLGS) can alter the packing of the glycan cloud that surrounds the virion, by a loss, gain, or shift of an NLGS (32), thus facilitating NAb escape.Env is the only viral protein targeted by both CTL and NAb responses. The serial viral escape from both Env-specific CTL and NAb responses could have implications for viral fitness and the reversion of multiple mutations upon transmission to naïve hosts.We previously identified three common HIV-1 Env-specific CD8 T cell epitopes, RY8788-795, SP9110-118, and NL9671-679, and their immune escape patterns in pigtail macaques (Macaca nemestrina) infected with SHIVmn229 (25). SHIVmn229 is a chimeric virus constructed from an SIVmac239 backbone and an HIV-1HXB2 env fragment that was passaged through macaques to become pathogenic (11). This earlier work provided an opportunity for detailed studies of how viruses with Env-specific CTL escape mutations, as well as mutations in adjacent NLGS, evolve when transmitted to naïve pigtail macaques.  相似文献   

10.
11.
We previously showed that agonistic antibodies to CD40 could substitute for CD4 T-cell help and prevent reactivation of murine gammaherpesvirus 68 (MHV-68) in the lungs of major histocompatibility complex (MHC) class II−/− (CII−/−) mice, which are CD4 T cell deficient. Although CD8 T cells were required for this effect, no change in their activity was detected in vitro. A key question was whether anti-CD40 treatment (or CD4 T-cell help) changed the function of CD8 T cells or another cell type in vivo. To address this question, in the present study, we showed that adoptive transfer of CD8 T cells from virus-infected wild-type mice or anti-CD40-treated CII−/− mice caused a significant reduction in lung viral titers, in contrast to those from control CII−/− mice. Anti-CD40 treatment also greatly prolonged survival of infected CII−/− mice. This confirms that costimulatory signals cause a change in CD8 T cells enabling them to maintain effective long-term control of MHV-68. We investigated the nature of this change and found that expression of the inhibitory receptor PD-1 was significantly increased on CD8 T cells in the lungs of MHV-68-infected CII−/−, CD40−/−, or CD80/86−/− mice, compared with that in wild-type or CD28/CTLA4−/− mice, correlating with the level of viral reactivation. Furthermore, blocking PD-1-PD-L1 interactions significantly reduced viral reactivation in CD4 T-cell-deficient mice. In contrast, the absence of another inhibitory receptor, NKG2A, had no effect. These data suggest that CD4 T-cell help programs a change in CD8 T-cell function mediated by altered PD-1 expression, which enables effective long-term control of MHV-68.Murine gammaherpesvirus 68 (MHV-68) is a naturally occurring rodent pathogen which is closely related to Epstein-Barr virus (EBV) and Kaposi''s sarcoma-associated herpesvirus (KSHV) (17, 64). Intranasal administration of MHV-68 to mice results in acute productive infection of lung epithelial cells and a latent infection in various cell types, including B lymphocytes, dendritic cells, epithelial cells, and macrophages (18, 19, 52, 53, 61, 65). The virus induces an inflammatory infiltrate in the lungs, lymph node enlargement, splenomegaly, and mononucleosis comprising increased numbers of activated CD8 T cells in the blood (53, 58). It has also been reported to induce lymphoproliferative disease/lymphoma in immunocompromised mice (30, 55, 60). Thus, the pathogenesis resembles that of EBV in humans, although structurally, the virus is more closely related to KSHV.Infectious MHV-68 is cleared from the lungs by a T-cell-dependent mechanism 10 to 15 days after infection (18, 53, 56). In wild-type mice, the lungs remain clear of replicating virus thereafter. Although CD4 T cells are not essential for primary clearance of replicating virus, they are required for effective long-term control (11). Thus, major histocompatibility complex (MHC) class II−/− mice that lack CD4 T cells or mice rendered CD4 deficient by antibody treatment initially clear infectious virus from the lungs. However, infectious virus reactivates in the lungs 10 to 15 days later and gradually increases in titer (11, 43). The infected CD4-deficient mice eventually die, apparently from long-term lung damage due to continuing lytic viral replication (11). MHC class II−/− mice do not produce antibody to T-dependent antigens (10). Cytotoxic T-lymphocyte (CTL) epitopes have been identified in open reading frame (ORF) 6 (p56, H-2Db-restricted), and ORF 61 (p79, H-2Kb-restricted) gene products, which appear to encode early lytic-phase proteins (32, 49). The epitopes are presented during two distinct phases during MHV-68 infection, which changes the pattern of CTL dominance (32, 51). However, there is no significant difference in the numbers of CD8 T cells specific for each epitope in wild-type mice and CD4 T-cell-deficient mice (4, 50). In addition, CTL activity measured in vitro does not differ substantially in the lungs of wild-type mice or CD4 T-cell-deficient mice (4, 11, 50). Furthermore, postexposure vaccination with the p56 epitope failed to prevent viral reactivation in class II−/− mice, despite dramatically expanding the number of CD8 T cells specific for the peptide (5). In contrast, vaccination of wild-type mice against these epitopes reduced lytic viral titers in the lung dramatically on subsequent challenge with MHV-68. B-cell-deficient mice clear MHV-68 with the kinetics of wild-type mice and do not show viral reactivation in the lungs (13, 61), suggesting that antibody is not essential for control of the virus. Depletion of CD4 T cells during the latent phase of infection in B-cell-deficient mice does not induce viral reactivation, whereas depletion of both CD4 and CD8 T-cell subsets provokes viral reactivation in the lungs (52). Short-term depletion of both CD4 and CD8 T-cell subsets during the latent phase of infection in wild-type mice does not lead to viral reactivation probably due to the presence of neutralizing antibody (11). Taken together, these results suggest that CD4 and CD8 T cells and B cells play overlapping roles in preventing or controlling reactivation of MHV-68 during the latent phase of infection. However, the B-cell- and CD8 T-cell-mediated control mechanisms do not develop in the absence of CD4 T cells.We, and others, have previously shown that the costimulatory molecule CD28 is not required for long-term control of MHV-68 (28, 29). However, interestingly, mice lacking both of the ligands for CD28, CD80 and CD86, show viral reactivation in the lung (21, 35). Our previously published data showed that agonistic antibodies to CD40 could substitute for CD4 T-cell function in the long-term control of MHV-68 (46). CD8 T-cell receptor-positive (TCR+) cells were required for this effect, while antibody production was not restored (45, 46). MHV-68-infected CD40L−/− mice (7) and CD40−/− mice (29) also showed viral reactivation in the lungs. However, no change in CD8 CTL activity was detected in in vitro assays following anti-CD40 treatment (46). A key question was whether anti-CD40 treatment (or CD4 T-cell help) caused a direct change in CD8 T-cell function or whether both CD8 T cells and an independent anti-CD40-sensitive step were required for viral control. To address this question, we used adoptive transfer of CD8 T cells from MHV-68-infected wild-type mice, anti-CD40-treated mice, or control MHC class II−/− mice to MHV-68-infected class II−/− recipients. We also investigated whether anti-CD40 treatment prolonged survival in addition to reducing lung viral titers. The heterodimeric molecule CD94/NKG2A has been implicated in negatively regulating the CD8 T-cell response to polyomavirus (38) and herpes simplex virus (HSV) (54), while the inhibitory receptor PD-1 (programmed death 1) has been implicated in T-cell exhaustion following infection with several other persistent viruses (2, 15, 20, 22, 26, 36, 39-41, 57, 67). In the present study, we investigated the effect of signaling via various costimulatory molecules on the expression of NKG2A and PD-1 and how these molecules influenced viral control.  相似文献   

12.
Mutations that allow escape from CD8 T-cell responses are common in HIV-1 and may attenuate pathogenesis by reducing viral fitness. While this has been demonstrated for individual cases, a systematic investigation of the consequence of HLA class I-mediated selection on HIV-1 in vitro replication capacity (RC) has not been undertaken. We examined this question by generating recombinant viruses expressing plasma HIV-1 RNA-derived Gag-Protease sequences from 66 acute/early and 803 chronic untreated subtype B-infected individuals in an NL4-3 background and measuring their RCs using a green fluorescent protein (GFP) reporter CD4 T-cell assay. In acute/early infection, viruses derived from individuals expressing the protective alleles HLA-B*57, -B*5801, and/or -B*13 displayed significantly lower RCs than did viruses from individuals lacking these alleles (P < 0.05). Furthermore, acute/early RC inversely correlated with the presence of HLA-B-associated Gag polymorphisms (R = −0.27; P = 0.03), suggesting a cumulative effect of primary escape mutations on fitness during the first months of infection. At the chronic stage of infection, no strong correlations were observed between RC and protective HLA-B alleles or with the presence of HLA-B-associated polymorphisms restricted by protective alleles despite increased statistical power to detect these associations. However, RC correlated positively with the presence of known compensatory mutations in chronic viruses from B*57-expressing individuals harboring the Gag T242N mutation (n = 50; R = 0.36; P = 0.01), suggesting that the rescue of fitness defects occurred through mutations at secondary sites. Additional mutations in Gag that may modulate the impact of the T242N mutation on RC were identified. A modest inverse correlation was observed between RC and CD4 cell count in chronic infection (R = −0.17; P < 0.0001), suggesting that Gag-Protease RC could increase over the disease course. Notably, this association was stronger for individuals who expressed B*57, B*58, or B*13 (R = −0.27; P = 0.004). Taken together, these data indicate that certain protective HLA alleles contribute to early defects in HIV-1 fitness through the selection of detrimental mutations in Gag; however, these effects wane as compensatory mutations accumulate in chronic infection. The long-term control of HIV-1 in some persons who express protective alleles suggests that early fitness hits may provide lasting benefits.The host immune response elicited by CD8+ cytotoxic T lymphocytes (CTLs) is a major contributor to viral control following human immunodeficiency virus type 1 (HIV-1) infection (6, 39), but antiviral pressure exerted by CTLs is diminished by the selection of escape mutations in targeted regions throughout the viral proteome (7, 18, 29, 35, 41, 45, 57). A comprehensive identification of HLA-associated viral polymorphisms has recently been achieved through population-based analyses of HIV-1 sequences and HLA class I types from different cohorts worldwide (3, 8, 13-15, 34, 43, 50, 56, 63). However, despite improved characterization of the sites and pathways of immune escape, effective ways to incorporate these findings into immunogen design remain an area of debate. A better understanding of the impact of escape mutations on viral fitness may provide novel directions for HIV-1 vaccines that are designed to attenuate pathogenesis.The development of innovative vaccine strategies that can overcome the extreme diversity of HIV is a key priority (4). One proposed approach is to target the most conserved T-cell epitopes, which presumably cannot escape from CTL pressure easily due to structural or functional constraints on the viral protein (55). Complementary approaches include the design of polyvalent and/or mosaic immunogens that incorporate commonly observed viral diversity (4, 38) or the specific targeting of vulnerable regions of the viral proteome that do escape but only at a substantial cost to viral replication capacity (RC) (1, 40). A chief target of such vaccine approaches is the major HIV-1 structural protein Gag, which is known to be highly immunogenic and to elicit CTL responses that correlate with the natural control of infection (22, 36, 66). Indeed, several lines of evidence support a relationship between the selection of CTL escape mutations and reduced HIV-1 fitness. These include the reversion of escape mutations following transmission to an HLA-mismatched recipient who cannot target the epitope (19, 24, 41) as well as reduced plasma viral load (pVL) set point following the transmission of certain escape variants from donors who expressed protective HLA alleles (17, 27). Notably, these in vivo observations have been made most often for variations within Gag that are attributed to CTL responses restricted by the protective alleles HLA-B*57 and -B*5801 (17, 19, 27, 41). Most recently, reduced in vitro RCs of clinical isolates and/or engineered strains encoding single or multiple escape mutations in Gag selected in the context of certain protective HLA alleles, including B*57, B*5801, B*27, and B*13, have been demonstrated (9, 10, 42, 53, 59, 62). Despite these efforts, the goal of a T-cell vaccine that targets highly conserved and attenuation-inducing sites is hampered by a lack of knowledge concerning the contribution of most escape mutations to HIV-1 fitness as well as a poor understanding of the relative influence of HLA on the viral RC at different stages of infection.The mutability of HIV-1 permits the generation of progeny viruses encoding compensatory mutations that restore normal protein function and/or viral fitness. Detailed studies have demonstrated that the in vitro RC of escape variants in human and primate immunodeficiency viruses can be enhanced by the addition of secondary mutations outside the targeted epitope (10, 20, 52, 59, 65). Thus, vaccine strategies aimed at attenuating HIV-1 must also consider, among other factors, the frequency, time course, and extent to which compensation might overcome attenuation mediated by CTL-induced escape. Despite its anticipated utility for HIV-1 vaccine design, systematic studies to examine the consequences of naturally occurring CTL escape and compensatory mutations on viral RC have not been undertaken.We have described previously an in vitro recombinant viral assay to examine the impact of Gag-Protease mutations on HIV-1 RC (47, 49). Gag and protease have been included in each virus to minimize the impact of sequence polymorphisms at Gag cleavage sites, which coevolve with changes in protease (5, 37). Using this approach, we have demonstrated that viruses derived from HIV-1 controllers replicated significantly less well than those derived from noncontrollers and that these differences were detectable at both the acute/early (49) and chronic (47) stages. Escape mutations in Gag associated with the protective HLA-B*57 allele, as well as putative compensatory mutations outside known CTL epitopes, contributed to this difference in RC (47). However, substantial variability was observed for viruses from controllers and noncontrollers, indicating that additional factors were likely to be involved. Benefits of this assay include its relatively high-throughput capacity as well as the fact that clinically derived HIV-1 sequences are used in their entirety. Thus, it is possible to examine a large number of “real-world” Gag-Protease sequences, to define an RC value for each one, and to identify sequences within the population of recombinant strains that are responsible for RC differences.Here, we use this recombinant virus approach to examine the contribution of HLA-associated immune pressure on Gag-Protease RC during acute/early (n = 66) and chronic (n = 803) infections in the context of naturally occurring HIV-1 subtype B isolates from untreated individuals. In a recent report (64), we employed this system to examine the Gag-Protease RC in a similar cohort of chronic HIV-1 subtype C-infected individuals. The results of these studies provide important insights into the roles of immune pressure and fitness constraints on HIV-1 evolution that may contribute to the rational design of an effective vaccine.  相似文献   

13.
The prominent role of antiviral cytotoxic CD8+ T-lymphocytes (CD8-TL) in containing the acute viremia of human and simian immunodeficiency viruses (HIV-1 and SIV) has rationalized the development of T-cell-based vaccines. However, the presence of escape mutations in the acute stage of infection has raised a concern that accelerated escape from vaccine-induced CD8-TL responses might undermine vaccine efficacy. We reanalyzed previously published data of 101,822 viral genomes of three CD8-TL epitopes, Nef103-111RM9 (RM9), Tat28-35SL8 (SL8), and Gag181-189CM9 (CM9), sampled by ultradeep pyrosequencing from eight macaques. Multiple epitope variants appeared during the resolution of acute viremia, followed by the predominance of a single mutant epitope. By fitting a mathematical model, we estimated the first acute escape rate as 0.36 day−1 within escape-prone epitopes, RM9 and SL8, and the chronic escape rate as 0.014 day−1 within the CM9 epitope. Our estimate of SIV acute escape rates was found to be comparable to very early HIV-1 escape rates. The timing of the first escape was more highly correlated with the timing of the peak CD8-TL response than with the magnitude of the CD8-TL response. The transmitted epitope decayed more than 400 times faster during the acute viral decline stage than predicted by a neutral evolution model. However, the founder epitope persisted as a minor population even at the viral set point; in contrast, the majority of acute escape epitopes were completely cleared. Our results suggest that a reservoir of SIV infection is preferentially formed by virus with the transmitted epitope.A critical role of CD8+ T-lymphocytes (CD8-TL) in controlling the peak of acute viral replication has been demonstrated both in HIV-1 (10, 31, 57) and experimental SIV infections (51). HIV-1-infected patients with strong HIV-1-specific CD8-TL responses early after the onset of the acute retroviral syndrome showed more effective control of primary viremia than patients with low or undetectable virus-specific CD8-TL activity (10). Delayed HIV-1-specific CD8-TL responses within an acutely infected individual was found to be one factor contributing to the patient''s persistent viremia, symptoms, and low CD4+ T-cell counts (31). A close temporal association between the magnitude of immunodominant B57-restrcited HIV-1-specific CD8 T cells and viral load was observed (57). In nonhuman primate models, the effect of CD8+ T cells on acute viral containment has been more directly probed by administering an anti-CD8 antibody to transiently deplete CD8+ lymphocytes from the peripheral blood (51). The resolution of peak viremia was much slower in the CD8+ lymphocyte-depleted rhesus macaques than in the untreated control animals (51).CT8-TL responses provide selective pressure within human leukocyte antigen (HLA)-restricted regions of the viral genome, which can select for escape variants. Understanding the kinetics of viral escape has important implications for the development of T-cell-based vaccines. Recently, in acutely infected HIV-1 subjects, single-genome amplification (SGA) and sequencing have shown that while only random mutations were observed prior to peak viremia (50), CD8-TL escape mutations were prominent as early as 20 to 30 days after the acute peak of viremia (24), well before the establishment of the viral set point. Indeed, it was observed that the emergence of viral escape mutants occurred coincidently with the expansion of the epitope-specific CD8-TL population in the acutely infected host, and that it resulted in amino acid substitutions in the transmitted/founder virus that diminished recognition by CT8-TL specific for the original (transmitted) epitope (24).Quantitatively, the average rate of CD8-TL escape mutation within 20 days of HIV-1 infection since the first screening has been estimated as 0.33 day−1 (24). This early escape rate is substantially greater than the chronic escape rate, which has been estimated as 0.04 day−1 (6). However, these prior estimates (6, 24) have been based on Sanger sequencing data from a limited number of virus clones. The availability of ultradeep pyrosequencing methods provides the opportunity to revisit these estimates using much richer data sets, which can detect mutations with a frequency of as little as 1% (8). The quantification of the rate of CD8-TL escape in SIV and HIV-1 is important, since it can serve as a surrogate measure of the magnitude and effectiveness of the host CD8-TL response. Mathematical models have been developed to quantify the process of viral CD8-TL escape (6, 7, 23), which framed the escape phenomenon as a synergetic outcome of the differences of wild and mutant epitopes in terms of susceptibility to cytotoxic T-lymphocyte (CTL) killing versus their intrinsic viral fitness.The goal of the present study was to quantify escape dynamics within three well-defined CD8-TL epitopes by rigorously analyzing both previously published and newly generated ultradeep pyrosequencing data from a set of eight SIV-infected macaques (8). Bimber and colleagues (8) previously demonstrated multifarious patterns of CTL escape in these SIV-infected macaques, and a recently published analysis of the same data set by Hughes et al. revealed that the persistence of low levels of inoculum sequence and its consistent loss kinetics enable the reliable inference of the wild-type sequence when only samples from later in infection are available for study (26). Here, we used the same extensive sequence data set, in combination with newly generated data, to quantify viral escape dynamics for three well-defined CD8-TL epitopes relative to the transmitted (wild-type) epitope sequence. By fitting a mathematical model of CD8-TL escape (6) to the experimentally determined CT8-TL escape kinetics, we compared the rate of the first CD8-TL escape of the escape-prone epitopes, Nef103-111RM9 and Tat28-35SL8, to that of the escape-resistant epitope, Gag181-189CM9. For this purpose, we define the time to first CD8-TL escape as the time when the first CD8-TL escape mutant comprises 50% of the combined population of the transmitted (wild) sequence and the first escape mutant clone. This definition is different from the timing of the first emergence of amino acid variants within an epitope. Our definition can be used when individual clones are obtained either by single-genome amplification (42, 49) or pyrosequencing (32, 48).In this study, by employing a rich data set from ultradeep pyrosequencing, we tested the hypothesis that the transmitted epitope contributes to the formation of a reservoir of infection. Our results suggest that this is indeed the case, and they also suggest that viral reversion (13, 21, 34, 37) is complicated in some cases by the unexpected persistence of wild-type, transmitted virus strains long after initial infection.  相似文献   

14.
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   

15.
Binding to the primary receptor CD4 induces conformational changes in the human immunodeficiency virus type 1 (HIV-1) gp120 envelope glycoprotein that allow binding to the coreceptor (CCR5 or CXCR4) and ultimately trigger viral membrane-cell membrane fusion mediated by the gp41 transmembrane envelope glycoprotein. Here we report the derivation of an HIV-1 gp120 variant, H66N, that confers envelope glycoprotein resistance to temperature extremes. The H66N change decreases the spontaneous sampling of the CD4-bound conformation by the HIV-1 envelope glycoproteins, thus diminishing CD4-independent infection. The H66N change also stabilizes the HIV-1 envelope glycoprotein complex once the CD4-bound state is achieved, decreasing the probability of CD4-induced inactivation and revealing the enhancing effects of soluble CD4 binding on HIV-1 infection. In the CD4-bound conformation, the highly conserved histidine 66 is located between the receptor-binding and gp41-interactive surfaces of gp120. Thus, a single amino acid change in this strategically positioned gp120 inner domain residue influences the propensity of the HIV-1 envelope glycoproteins to negotiate conformational transitions to and from the CD4-bound state.Human immunodeficiency virus type 1 (HIV-1), the cause of AIDS (6, 29, 66), infects target cells by direct fusion of the viral and target cell membranes. The viral fusion complex is composed of gp120 and gp41 envelope glycoproteins, which are organized into trimeric spikes on the surface of the virus (10, 51, 89). Membrane fusion is initiated by direct binding of gp120 to the CD4 receptor on target cells (17, 41, 53). CD4 binding creates a second binding site on gp120 for the chemokine receptors CCR5 and CXCR4, which serve as coreceptors (3, 12, 19, 23, 25). Coreceptor binding is thought to lead to further conformational changes in the HIV-1 envelope glycoproteins that facilitate the fusion of viral and cell membranes. The formation of an energetically stable six-helix bundle by the gp41 ectodomain contributes to the membrane fusion event (9, 10, 79, 89, 90).The energy required for viral membrane-cell membrane fusion derives from the sequential transitions that the HIV-1 envelope glycoproteins undergo, from the high-energy unliganded state to the low-energy six-helix bundle. The graded transitions down this energetic slope are initially triggered by CD4 binding (17). The interaction of HIV-1 gp120 with CD4 is accompanied by an unusually large change in entropy, which is thought to indicate the introduction of order into the conformationally flexible unliganded gp120 glycoprotein (61). In the CD4-bound state, gp120 is capable of binding CCR5 with high affinity; moreover, CD4 binding alters the quaternary structure of the envelope glycoprotein complex, resulting in the exposure of gp41 ectodomain segments (27, 45, 77, 92). The stability of the intermediate state induced by CD4 binding depends upon several variables, including the virus (HIV-1 versus HIV-2/simian immunodeficiency virus [SIV]), the temperature, and the nature of the CD4 ligand (CD4 on a target cell membrane versus soluble forms of CD4 [sCD4]) (30, 73). For HIV-1 exposed to sCD4, if CCR5 binding occurs within a given period of time, progression along the entry pathway continues. If CCR5 binding is impeded or delayed, the CD4-bound envelope glycoprotein complex decays into inactive states (30). In extreme cases, the binding of sCD4 to the HIV-1 envelope glycoproteins induces the shedding of gp120 from the envelope glycoprotein trimer (31, 56, 58). Thus, sCD4 generally inhibits HIV-1 infection by triggering inactivation events, in addition to competing with CD4 anchored in the target cell membrane (63).HIV-1 isolates vary in sensitivity to sCD4, due in some cases to a low affinity of the envelope glycoprotein trimer for CD4 and in other cases to differences in propensity to undergo inactivating conformational transitions following CD4 binding (30). HIV-1 isolates that have been passaged extensively in T-cell lines (the tissue culture laboratory-adapted [TCLA] isolates) exhibit lower requirements for CD4 than primary HIV-1 isolates (16, 63, 82). TCLA viruses bind sCD4 efficiently and are generally sensitive to neutralization compared with primary HIV-1 isolates. Differences in sCD4 sensitivity between primary and TCLA HIV-1 strains have been mapped to the major variable loops (V1/V2 and V3) of the gp120 glycoprotein (34, 42, 62, 81). Sensitivity to sCD4 has been shown to be independent of envelope glycoprotein spike density or the intrinsic stability of the envelope glycoprotein complex (30, 35).In general, HIV-1 isolates are more sensitive to sCD4 neutralization than HIV-2 or SIV isolates (4, 14, 73). The relative resistance of SIV to sCD4 neutralization can in some cases be explained by a reduced affinity of the envelope glycoprotein trimer for sCD4 (57); however, at least some SIV isolates exhibit sCD4-induced activation of entry into CD4-negative, CCR5-expressing target cells that lasts for several hours after exposure to sCD4 (73). Thus, for some primate immunodeficiency virus envelope glycoproteins, activated intermediates in the CD4-bound conformation can be quite stable.The HIV-1 envelope glycoprotein elements important for receptor binding, subunit interaction, and membrane fusion are well conserved among different viral strains (71, 91). Thus, these elements represent potential targets for inhibitors of HIV-1 entry. Understanding the structure and longevity of the envelope glycoprotein intermediates along the virus entry pathway is relevant to attempts at inhibition. For example, peptides that target the heptad repeat 1 region of gp41 exhibit major differences in potency against HIV-1 strains related to efficiency of chemokine receptor binding (20, 21), which is thought to promote the conformational transition to the next step in the virus entry cascade. The determinants of the duration of exposure of targetable HIV-1 envelope glycoprotein elements during the entry process are undefined.To study envelope glycoprotein determinants of the movement among the distinct conformational states along the HIV-1 entry pathway, we attempted to generate HIV-1 variants that exhibit improved stability. Historically, labile viral elements have been stabilized by selecting virus to replicate under conditions, such as high temperature, that typically weaken protein-protein interactions (38, 39, 76, 102). Thus, we subjected HIV-1 to repeated incubations at temperatures between 42°C and 56°C, followed by expansion and analysis of the remaining replication-competent virus fraction. In this manner, we identified an envelope glycoprotein variant, H66N, in which histidine 66 in the gp120 N-terminal segment was altered to asparagine. The resistance of HIV-1 bearing the H66N envelope glycoproteins to changes in temperature has been reported elsewhere (37). Here, we examine the effect of the H66N change on the ability of the HIV-1 envelope glycoproteins to negotiate conformational transitions, either spontaneously or in the presence of sCD4. The H66N phenotype was studied in the context of both CD4-dependent and CD4-independent HIV-1 variants.  相似文献   

16.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

17.
18.
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   

19.
20.
The Nef protein of human immunodeficiency virus type 1 downregulates the CD4 coreceptor from the surface of host cells by accelerating the rate of CD4 endocytosis through a clathrin/AP-2 pathway. Herein, we report that Nef has the additional function of targeting CD4 to the multivesicular body (MVB) pathway for eventual delivery to lysosomes. This targeting involves the endosomal sorting complex required for transport (ESCRT) machinery. Perturbation of this machinery does not prevent removal of CD4 from the cell surface but precludes its lysosomal degradation, indicating that accelerated endocytosis and targeting to the MVB pathway are separate functions of Nef. We also show that both CD4 and Nef are ubiquitinated on lysine residues, but this modification is dispensable for Nef-induced targeting of CD4 to the MVB pathway.Primate immunodeficiency viruses infect helper T lymphocytes and cells of the macrophage/monocyte lineage by binding of their viral envelope glycoprotein, Env, to a combination of two host cell-specific surface proteins, CD4 and either the CCR5 or CXCR4 chemokine receptors (reviewed in reference 62). Ensuing fusion of the viral envelope with the host cell plasma membrane delivers the viral genetic material into the cytoplasm. Remarkably, the most highly transcribed viral gene in the early phase of infection does not encode an enzyme or structural protein but an accessory protein named Nef. Early expression of Nef is thought to reprogram the host cell for optimal replication of the virus. Indeed, Nef has been shown to enhance virus production (19, 24, 59, 74) and to promote progression to AIDS (23, 47, 48), making it an attractive candidate for pharmacologic intervention.Nef is an N-terminally myristoylated protein with a molecular mass of 27 kDa for human immunodeficiency virus type 1 (HIV-1) and 35 kDa for HIV-2 and simian immunodeficiency virus (27, 29, 50, 65). Nef has been ascribed many functions, the best characterized of which is the downregulation of the CD4 coreceptor from the surface of infected cells (28, 35, 57). CD4 downregulation is believed to prevent superinfection (8, 52) and to preclude the cellular retention of newly synthesized Env (8, 49), thus allowing the establishment of a robust infection (30, 71).The molecular mechanism by which Nef downregulates CD4 has been extensively studied. A consensus has emerged that Nef accelerates the endocytosis of cell surface CD4 (2, 64) by linking the cytosolic tail of CD4 to the heterotetrameric (α-β2-μ2-σ2) adaptor protein-2 (AP-2) complex (17, 25, 34, 45, 67). Determinants in the CD4 tail bind to a hydrophobic pocket comprising tryptophan-57 and leucine-58 on the folded core domain of Nef (34). On the other hand, a dileucine motif (i.e., ENTSLL, residues 160 to 165) (14, 22, 32) and a diacidic motif (i.e., DD, residues 174 and 175) (3) (residues correspond to the NL4-3 clone of HIV-1) within a C-terminal, flexible loop of Nef bind to the α and σ2 subunits of AP-2 (17, 18, 25, 51). AP-2, in turn, binds to clathrin, leading to the concentration of CD4 within clathrin-coated pits (15, 33). These pits eventually bud from the plasma membrane as clathrin-coated vesicles that deliver internalized CD4 to endosomes. In essence, then, Nef acts as a connector that confers on CD4 the ability to be rapidly internalized in a manner similar to endocytic receptors (75).Unlike typical endocytic recycling receptors like the transferrin receptor or the low-density lipoprotein receptor, however, CD4 that is forcibly internalized by Nef does not return to the cell surface but is delivered to lysosomes for degradation (4, 64, 68). Thus, expression of Nef decreases both the surface and total levels of CD4. What keeps internalized CD4 from returning to the plasma membrane? We hypothesized that Nef might additionally act on endosomes to direct CD4 to lysosomes. This is precisely the fate followed by signaling receptors, transporters, and other transmembrane proteins that undergo ubiquitination-mediated internalization and targeting to the multivesicular body (MVB) pathway (40, 46). This targeting involves the endosomal sorting complex required for transport (ESCRT), including the ESCRT-0, -I, -II, and -III complexes, which function to sort ubiquitinated cargoes into intraluminal vesicles of MVBs for eventual degradation in lysosomes (40, 46). Herein, we show that Nef indeed plays a novel role in targeting internalized CD4 from endosomes to the MVB pathway in an ESCRT-dependent manner. We also show that both Nef and CD4 undergo ubiquitination on lysine residues, but, strikingly, this modification is not required for CD4 targeting to the MVB pathway.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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