首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
During yeast sporulation, a forespore membrane (FSM) initiates at each spindle-pole body and extends to form the spore envelope. We used Schizosaccharomyces pombe to investigate the role of septins during this process. During the prior conjugation of haploid cells, the four vegetatively expressed septins (Spn1, Spn2, Spn3, and Spn4) coassemble at the fusion site and are necessary for its normal morphogenesis. Sporulation involves a different set of four septins (Spn2, Spn5, Spn6, and the atypical Spn7) that does not include the core subunits of the vegetative septin complex. The four sporulation septins form a complex in vitro and colocalize interdependently to a ring-shaped structure along each FSM, and septin mutations result in disoriented FSM extension. The septins and the leading-edge proteins appear to function in parallel to orient FSM extension. Spn2 and Spn7 bind to phosphatidylinositol 4-phosphate [PtdIns(4)P] in vitro, and PtdIns(4)P is enriched in the FSMs, suggesting that septins bind to the FSMs via this lipid. Cells expressing a mutant Spn2 protein unable to bind PtdIns(4)P still form extended septin structures, but these structures fail to associate with the FSMs, which are frequently disoriented. Thus, septins appear to form a scaffold that helps to guide the oriented extension of the FSM.Yeast sporulation is a developmental process that involves multiple, sequential events that need to be tightly coordinated (59, 68). In the fission yeast Schizosaccharomyces pombe, when cells of opposite mating type (h+ and h) are mixed and shifted to conditions of nitrogen starvation, cell fusion and karyogamy occur to form a diploid zygote, which then undergoes premeiotic DNA replication, the two meiotic divisions, formation of the spore envelopes (comprising the plasma membrane and a specialized cell wall), and maturation of the spores (74, 81). At the onset of meiosis II, precursors of the spore envelopes, the forespore membranes (FSMs), are formed by the fusion of vesicles at the cytoplasmic surface of each spindle-pole body (SPB) and then extend to engulf the four nuclear lobes (the nuclear envelope does not break down during meiosis), thus capturing the haploid nuclei, along with associated cytoplasm and organelles, to form the nascent spores (55, 68, 81). How the FSMs recognize and interact with the nuclear envelope, extend in a properly oriented manner, and close to form uniformly sized spherical spores is not understood, and study of this model system should also help to elucidate the more general question of how membranes obtain their shapes in vivo.It has been shown that both the SPB and the vesicle trafficking system play important roles in the formation and development of the FSM and of its counterpart in the budding yeast Saccharomyces cerevisiae, the prospore membrane (PSM). In S. pombe, the SPB changes its shape from a compact dot to a crescent at metaphase of meiosis II (26, 29), and its outer plaque acquires meiosis-specific components such as Spo2, Spo13, and Spo15 (30, 57, 68). This modified outer plaque is required for the initiation of FSM assembly. In S. cerevisiae, it is well established that various secretory (SEC) gene products are required for PSM formation (58, 59). Similarly, proteins presumably involved in the docking and/or fusion of post-Golgi vesicles and organelles in S. pombe, such as the syntaxin-1A Psy1, the SNAP-25 homologue Sec9, and the Rab7 GTPase homologue Ypt7, are also required for proper FSM extension (34, 53, 54). Consistent with this hypothesis, Psy1 disappears from the plasma membrane upon exit from meiosis I and reappears in the nascent FSM.Phosphoinositide-mediated membrane trafficking also contributes to the development of the FSM. Pik3/Vps34 is a phosphatidylinositol 3-kinase whose product is phosphatidylinositol 3-phosphate [PtdIns(3)P] (35, 72). S. pombe cells lacking this protein exhibit defects in various steps of FSM formation, such as aberrant starting positions for extension, disoriented extension and/or failure of closure, and the formation of spore-like bodies near, rather than surrounding, the nuclei, suggesting that Pik3 plays multiple roles during sporulation (61). The targets of PtdIns(3)P during sporulation appear to include two sorting nexins, Vps5 and Vps17, and the FYVE domain-containing protein Sst4/Vps27. vps5Δ and vps17Δ mutant cells share some of the phenotypes of pik3Δ cells (38). sst4Δ cells also share some of the phenotypes of pik3Δ cells but are distinct from vps5Δ and vps17Δ cells, consistent with the hypothesis that Pik3 has multiple roles during sporulation (62).Membrane trafficking processes alone do not seem sufficient to explain how the FSMs and PSMs extend around and engulf the nuclei, suggesting that some other mechanism(s) must regulate and orient FSM/PSM extension. The observation that the FSM is attached to the SPB until formation of the immature spore is complete (68) suggests that the SPB may regulate FSM extension. In addition, the leading edge of the S. cerevisiae PSM is coated with a complex of proteins (the LEPs) that appear to be involved in PSM extension (51, 59). S. pombe Meu14 also localizes to the leading edge of the FSM, and deletion of meu14 causes aberrant FSM formation in addition to a failure in SPB modification (60). However, it has remained unclear whether the SPB- and LEP-based mechanisms are sufficient to account for the formation of closed FSMs and PSMs of proper size and position (relative to the nuclear envelope), and evidence from S. cerevisiae has suggested that the septin proteins may also be involved.The septins are a conserved family of GTP-binding proteins that were first identified in S. cerevisiae by analysis of the cytokinesis-defective cdc3, cdc10, cdc11, and cdc12 mutants (41). Cdc3, Cdc10, Cdc11, and Cdc12 are related to each other in sequence and form an oligomeric complex that localizes to a ring in close apposition to the plasma membrane at the mother-bud neck in vegetative cells (12, 20, 25, 41, 47, 77). The septin ring appears to be filamentous in vivo (12), and indeed, the septins from both yeast (11, 20) and metazoans (31, 36, 69) can form filaments in vitro. The yeast septin ring appears to form a scaffold for the localization and organization of a wide variety of other proteins (8, 22), and it forms a diffusion barrier that constrains movement of membrane proteins through the neck region (7, 8, 73). In metazoan cells, the septins are involved in cytokinesis but are also implicated in a variety of other cellular processes, such as vesicular transport, organization of the actin and microtubule cytoskeletons, and oncogenesis (27, 70).In S. cerevisiae, a fifth septin (Shs1) is also expressed in vegetative cells, but the remaining two septin genes, SPR3 and SPR28, are expressed at detectable levels only during sporulation (15, 17). In addition, at least some of the vegetatively expressed septins are also present in sporulating cells (17, 48), and one of them (Cdc10) is expressed at much higher levels there than in vegetative cells (32). The septins present during sporulation are associated with the PSM (15, 17, 48, 51), and their normal organization there depends on the Gip1-Glc7 protein phosphatase complex (71). However, it has been difficult to gain insight into the precise roles of the septins during sporulation in S. cerevisiae (59), because some septins are essential for viability during vegetative growth, and the viable mutants have only mild phenotypes during sporulation (15, 17), possibly because of functional redundancy among the multiple septins.S. pombe seemed likely to provide a better opportunity for investigating the role of septins during spore formation. There are seven septin genes (spn1+ to spn7+) in this organism (23, 41, 63). Four of these genes (spn1+ to spn4+) are expressed in vegetative cells, and their products form a hetero-oligomeric complex that assembles during cytokinesis into a ring at the division site (2, 3, 10, 76, 79). The septin ring is important for proper targeting of endoglucanases to the division site (44), and septin mutants show a corresponding delay in cell separation (10, 41, 44, 76). However, even the spn1Δ spn2Δ spn3Δ spn4Δ quadruple mutant is viable and grows nearly as rapidly as the wild type (our unpublished results), a circumstance that greatly facilitates studies of the septins'' role during sporulation.spn5+, spn6+, and spn7+ are expressed at detectable levels only during sporulation (1, 45, 78; our unpublished results), and spn2+, like its orthologue CDC10 (see above), is strongly induced (45), but the roles of the S. pombe septins in sporulation have not previously been investigated. In this study, we show that the septins are important for the orientation of FSM extension, suggesting that the septins may have a more general role in dynamic membrane organization and shape determination.  相似文献   

4.
5.
6.
7.
Deficiency in both ATM and the DNA-dependent protein kinase catalytic subunit (DNA-PKcs) is synthetically lethal in developing mouse embryos. Using mice that phenocopy diverse aspects of Atm deficiency, we have analyzed the genetic requirements for embryonic lethality in the absence of functional DNA-PKcs. Similar to the loss of ATM, hypomorphic mutations of Mre11 (Mre11ATLD1) led to synthetic lethality when juxtaposed with DNA-PKcs deficiency (Prkdcscid). In contrast, the more moderate DNA double-strand break response defects associated with the Nbs1ΔB allele permitted viability of some Nbs1ΔB/ΔB Prkdcscid/scid embryos. Cell cultures from Nbs1ΔB/ΔB Prkdcscid/scid embryos displayed severe defects, including premature senescence, mitotic aberrations, sensitivity to ionizing radiation, altered checkpoint responses, and increased chromosome instability. The known functions of DNA-PKcs in the regulation of Artemis nuclease activity or nonhomologous end joining-mediated repair do not appear to underlie the severe genetic interaction. Our results reveal a role for DNA-PKcs in the maintenance of S/G2-phase chromosome stability and in the induction of cell cycle checkpoint responses.The Mre11 complex, consisting of Mre11, Rad50, and Nbs1 (Xrs2 in Saccharomyces cerevisiae), is involved in diverse aspects of DNA double-strand break (DSB) metabolism. The Mre11 complex acts as a DSB sensor, mediates cell cycle checkpoint arrest and apoptosis, and promotes DSB repair (47, 48). The influence of the Mre11 complex on DSB responses is attributable partly to its influence on ataxia-telangiectasia mutated (ATM) kinase activity (29). ATM is a central signal transducer in the response to DSBs and is required for arrest throughout the cell cycle, as well as the efficient execution of apoptosis in response to many types of genotoxic stress (43).The Mre11 complex is required for ATM activation and governs the phosphorylation of ATM substrates such as SMC1, Chk2, and BID (4, 6, 26, 47, 49, 51). The C terminus of Nbs1 interacts with ATM and plays an important role in facilitating a subset of these events, particularly those important for apoptosis (11, 14, 47, 58). However, ATM makes multiple functional contacts with members of the Mre11 complex. Nbs1, Mre11, and Rad50 are all ATM substrates, and many aspects of ATM checkpoint signaling are impaired by hypomorphic Mre11 and Nbs1 mutations that do not affect the ATM binding domain in the C terminus of Nbs1 (32, 36, 52, 54).Several molecular and genetic observations support the view that the Mre11 complex''s role in preserving genome stability is particularly relevant to the S and G2 phases of the cell cycle (3, 56). The complex, predominantly nucleoplasmic in G1 cells, becomes predominantly chromatin associated and colocalizes with PCNA throughout S phase (35, 38). This association is a likely prerequisite for the complex''s influence on DNA damage signaling as well as DNA repair.Cell cultures established with samples from patients with Nijmegen breakage syndrome (NBS1 hypomorphism) and ataxia-telangiectasia-like disorder (MRE11 hypomorphism) exhibit checkpoint defects in S phase and at the G2/M transition, while the G1/S transition is relatively unaffected. These checkpoint defects are correlated with reduced Mre11 complex chromatin association both in human cells and in mouse models of Nijmegen breakage syndrome and ataxia-telangiectasia-like disorder (5, 45, 49, 52). Chromosomal aberrations arising in these cells are predominantly chromatid type breaks, consistent with impaired metabolism of DNA replication-associated DNA breaks (49, 52).Further supporting a predominant role for the Mre11 complex in S phase is the observation that its primary role in DSB repair is the promotion of recombination between sister chromatids (3, 24). Structural and genetic evidence that the Mre11 complex effects molecular bridging between DNA duplexes offers a mechanistic basis for this observation (10, 23, 53). Molecular bridging by the Mre11 complex may also contribute to its influence on nonhomologous end joining (NHEJ) (12, 34, 57). Collectively, these data strongly support the view that the Mre11 complex''s checkpoint and DSB repair functions are manifested predominantly in the S and G2 phases of the cell cycle.Although the Mre11 complex and ATM function in the same arm of the DNA damage response, ATM deficiency is lethal in hypomorphic Mre11 and Nbs1 mutants (Mre11ATLD1/ATLD1 and Nbs1ΔB/ΔB mice, respectively) (49, 52), suggesting that aspects of ATM function are Mre11 complex independent. ATM deficiency is also synthetically lethal with mutations in Prkdc, the gene encoding the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) that is mutated in mice with severe combined immunodeficiency (Prkdcscid mice) (22, 42). DNA-PKcs is an ATM paralog required for NHEJ, which appears to be the predominant mode of DSB repair in G1 cells (16).Defective NHEJ is unlikely to be the basis for the embryonic lethality of Prkdc/ Atm/ or Prkdcscid/scid Atm/ mice, as loss of ATM rescues the late embryonic lethality of both DNA ligase IV (Lig4) and XRCC4 null embryos, which have more severe NHEJ defects than Prkdcscid mice abolished by the Atm/ genotype (31, 42). These observations argue that the DNA-PKcs functions required for viability in the absence of ATM do not include NHEJ.To address this issue, we crossed Mre11ATLD1/ATLD1 and Nbs1ΔB/ΔB mice with Prkdcscid/scid mice. As these Mre11 complex hypomorphs do not completely phenocopy ATM deficiency, we reasoned that double-mutant animals would be viable and thus provide a venue in which to examine the functional relationship between the Mre11 complex/ATM arm of the DNA damage response and DNA-PKcs. Whereas the Mre11ATLD1/ATLD1 mutation was synthetically lethal with the Prkdcscid/scid genotype, some Nbs1ΔB/ΔB Prkdcscid/scid mice were born, consistent with the more moderate DNA damage response defects associated with the Nbs1ΔB allele than with the Mre11ATLD1 allele (48). Nbs1ΔB/ΔB Prkdcscid/scid embryos were born at drastically reduced Mendelian ratios, displayed gross developmental defects, and were severely runted. Nbs1ΔB/ΔB Prkdcscid/scid cell cultures exhibited profound chromosome instability, growth defects, and increased sensitivity to ionizing radiation (IR). DNA repair defects associated with DNA-PKcs deficiency did not appear to underlie the observed phenotypic synergy. Rather, the data suggest a novel regulatory function of DNA-PKcs in the maintenance of chromosomal stability during the S and G2 phases of the cell cycle.  相似文献   

8.
During the course of infection, transmitted HIV-1 isolates that initially use CCR5 can acquire the ability to use CXCR4, which is associated with an accelerated progression to AIDS. Although this coreceptor switch is often associated with mutations in the stem of the viral envelope (Env) V3 loop, domains outside V3 can also play a role, and the underlying mechanisms and structural basis for how X4 tropism is acquired remain unknown. In this study we used a V3 truncated R5-tropic Env as a starting point to derive two X4-tropic Envs, termed ΔV3-X4A.c5 and ΔV3-X4B.c7, which took distinct molecular pathways for this change. The ΔV3-X4A.c5 Env clone acquired a 7-amino-acid insertion in V3 that included three positively charged residues, reestablishing an interaction with the CXCR4 extracellular loops (ECLs) and rendering it highly susceptible to the CXCR4 antagonist AMD3100. In contrast, the ΔV3-X4B.c7 Env maintained the V3 truncation but acquired mutations outside V3 that were critical for X4 tropism. In contrast to ΔV3-X4A.c5, ΔV3-X4B.c7 showed increased dependence on the CXCR4 N terminus (NT) and was completely resistant to AMD3100. These results indicate that HIV-1 X4 coreceptor switching can involve (i) V3 loop mutations that establish interactions with the CXCR4 ECLs, and/or (ii) mutations outside V3 that enhance interactions with the CXCR4 NT. The cooperative contributions of CXCR4 NT and ECL interactions with gp120 in acquiring X4 tropism likely impart flexibility on pathways for viral evolution and suggest novel approaches to isolate these interactions for drug discovery.For human immunodeficiency virus type I (HIV-1) to enter a target cell, the gp120 subunit of the viral envelope glycoprotein (Env) must engage CD4 and a coreceptor on the cell surface. Although numerous coreceptors have been identified in vitro, the two most important coreceptors in vivo are the CCR5 (3, 11, 19, 22, 24) and CXCR4 (27) chemokine receptors. HIV-1 variants that can use only CCR5 (R5 viruses) are critical for HIV-1 transmission and predominate during the early stages of infection (86, 90). The importance of CCR5 for HIV-1 transmission is underscored by the fact that individuals bearing a homozygous 32-bp deletion in the CCR5 gene (ccr5-Δ32) are largely resistant to HIV-1 infection (15, 49, 84). Although R5 viruses typically persist into late disease stages, viruses that can use CXCR4, either alone (X4 viruses) or in addition to CCR5 (R5X4 viruses), emerge in approximately 50% of individuals infected with subtype B or D viruses (12, 39, 44). Although not required for disease progression, the appearance of X4 and/or R5X4 viruses is associated with a more rapid depletion of CD4+ cells in peripheral blood and faster progression to AIDS (12, 44, 77, 86). However, it remains unclear whether these viruses are a cause or a consequence of accelerated CD4+ T cell decline (57). The emergence of CXCR4-using viruses has also complicated the use of small-molecule CCR5 antagonists as anti-HIV-therapeutics as these compounds can select for the outgrowth of X4 or R5X4 escape variants (93).Following triggering by CD4, gp120 binds to a coreceptor via two principal interactions: (i) the bridging sheet, a four-stranded antiparallel beta sheet that connects the inner and outer domains of gp120, together with the base of the V3 loop, engages the coreceptor N terminus (NT); and (ii) more distal regions of V3 interact with the coreceptor extracellular loops (ECLs) (13, 14, 36-38, 43, 59, 60, 78, 79, 88). Although both the NT and ECL interactions are important for coreceptor binding and entry, their relative contributions vary among different HIV-1 strains (23). For example, V3 interactions with the ECLs, particularly ECL2, serve a dominant role in CXCR4 utilization (7, 21, 50, 63, 72), while R5 viruses exhibit a more variable use of CCR5 domains, with the NT interaction being particularly important (4, 6, 20, 67, 83). Although V3 is the primary determinant of coreceptor preference (34), it is unclear how specificity for CCR5 and/or CXCR4 is determined, and, in particular, it is unknown how X4 tropism is acquired. Several reports have shown that the emergence of X4 tropism correlates with the acquisition of positively charged residues in the V3 stem (17, 29, 87), particularly at positions 11, 24, and 25 (8, 17, 28, 29, 42, 75), raising the possibility that these mutations directly or indirectly mediate interactions with negatively charged residues in the CXCR4 ECLs. However, Env domains outside V3, including V1/V2 (9, 32, 45, 46, 61, 64, 65, 80, 95) and even gp41 (40), can also contribute to coreceptor switching, and it is unclear mechanistically or structurally how X4 tropism is determined.We previously derived a replication-competent variant of the R5X4 HIV-1 clone R3A that contained a markedly truncated V3 loop (47). This Env was generated by introducing a mutation termed ΔV3(9,9), which deleted the distal 15 amino acids of V3. The ΔV3(9,9) mutation selectively ablated X4 tropism but left R5 tropism intact, consistent with the view that an interaction between the distal half of V3 and the ECLs is critical for CXCR4 usage (7, 21, 43, 50, 59, 60, 63, 72). This V3-truncated virus provided a unique opportunity to address whether CXCR4 utilization could be regained on a background in which this critical V3-ECL interaction had been ablated and, if so, by what mechanism. Here, we characterize two novel X4 variants of R3A ΔV3(9,9) derived by adapting this virus to replicate in CXCR4+ CCR5 SupT1 cells. We show that R3A ΔV3(9,9) could indeed reacquire X4 tropism but through two markedly different mechanisms. One X4 variant, designated ΔV3-X4A, acquired changes in the V3 remnant that reestablished an interaction with the CXCR4 ECLs; the other, ΔV3-X4B, acquired changes outside V3 that engendered interactions with the CXCR4 NT. These divergent evolutionary pathways led to profound differences in sensitivity to the CXCR4 antagonist AMD3100, with ΔV3-X4A showing increased sensitivity relative to R3A and with ΔV3-X4B becoming completely resistant. These findings demonstrate the contributions that interactions with distinct coreceptor regions have in mediating tropism and drug sensitivity and illustrate how HIV''s remarkable evolutionary plasticity in adapting to selection pressures can be exploited to better understand its biological potential.  相似文献   

9.
10.
11.
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   

12.
The filovirus VP40 protein is capable of budding from mammalian cells in the form of virus-like particles (VLPs) that are morphologically indistinguishable from infectious virions. Ebola virus VP40 (eVP40) contains well-characterized overlapping L domains, which play a key role in mediating efficient virus egress. L domains represent only one component required for efficient budding and, therefore, there is a need to identify and characterize additional domains important for VP40 function. We demonstrate here that the 96LPLGVA101 sequence of eVP40 and the corresponding 84LPLGIM89 sequence of Marburg virus VP40 (mVP40) are critical for efficient release of VP40 VLPs. Indeed, deletion of these motifs essentially abolished the ability of eVP40 and mVP40 to bud as VLPs. To address the mechanism by which the 96LPLGVA101 motif of eVP40 contributes to egress, a series of point mutations were introduced into this motif. These mutants were then compared to the eVP40 wild type in a VLP budding assay to assess budding competency. Confocal microscopy and gel filtration analyses were performed to assess their pattern of intracellular localization and ability to oligomerize, respectively. Our results show that mutations disrupting the 96LPLGVA101 motif resulted in both altered patterns of intracellular localization and self-assembly compared to wild-type controls. Interestingly, coexpression of either Ebola virus GP-WT or mVP40-WT with eVP40-ΔLPLGVA failed to rescue the budding defective eVP40-ΔLPLGVA mutant into VLPs; however, coexpression of eVP40-WT with mVP40-ΔLPLGIM successfully rescued budding of mVP40-ΔLPLGIM into VLPs at mVP40-WT levels. In sum, our findings implicate the LPLGVA and LPLGIM motifs of eVP40 and mVP40, respectively, as being important for VP40 structure/stability and budding.Ebola and Marburg viruses are members of the family Filoviridae. Filoviruses are filamentous, negative-sense, single-stranded RNA viruses that cause lethal hemorrhagic fevers in both humans and nonhuman primates (5). Filoviruses encode seven viral proteins including: NP (major nucleoprotein), VP35 (phosphoprotein), VP40 (matrix protein), GP (glycoprotein), VP30 (minor nucleoprotein), VP24 (secondary matrix protein), and L (RNA-dependent RNA polymerase) (2, 5, 10, 12, 45). Numerous studies have shown that expression of Ebola virus VP40 (eVP40) alone in mammalian cells leads to the production of virus-like particles (VLPs) with filamentous morphology which is indistinguishable from infectious Ebola virus particles (12, 17, 18, 25, 26, 27, 30, 31, 34, 49). Like many enveloped viruses such as rhabdovirus (11) and arenaviruses (44), Ebola virus encodes late-assembly or L domains, which are sequences required for the membrane fission event that separates viral and cellular membranes to release nascent virion particles (1, 5, 7, 10, 12, 18, 25, 27, 34). Thus far, four classes of L domains have been identified which were defined by their conserved amino acid core sequences: the Pro-Thr/Ser-Ala-Pro (PT/SAP) motif (25, 27), the Pro-Pro-x-Tyr (PPxY) motif (11, 12, 18, 19, 41, 53), the Tyr-x-x-Leu (YxxL) motif (3, 15, 27, 37), and the Phe-Pro-Ile-Val (FPIV) motif (39). Both PTAP and the PPxY motifs are essential for efficient particle release for eVP40 (25, 27, 48, 49), whereas mVP40 contains only a PPxY motif. L domains are believed to act as docking sites for the recruitment of cellular proteins involved in endocytic trafficking and multivesicular body biogenesis to facilitate virus-cell separation (8, 13, 14, 16, 28, 29, 33, 36, 43, 50, 51).In addition to L domains, oligomerization, and plasma-membrane localization of VP40 are two functions of the protein that are critical for efficient budding of VLPs and virions. Specific sequences involved in self-assembly and membrane localization have yet to be defined precisely. However, recent reports have attempted to identify regions of VP40 that are important for its overall function in assembly and budding. For example, the amino acid region 212KLR214 located at the C-terminal region was found to be important for efficient release of eVP40 VLPs, with Leu213 being the most critical (30). Mutation of the 212KLR214 region resulted in altered patterns of cellular localization and oligomerization of eVP40 compared to those of the wild-type genotype (30). In addition, the proline at position 53 was also implicated as being essential for eVP40 VLP release and plasma-membrane localization (54).In a more recent study, a YPLGVG motif within the M protein of Nipah virus (NiV) was shown to be important for stability, membrane binding, and budding of NiV VLPs (35). Whether this NiV M motif represents a new class of L domain remains to be determined. However, it is clear that this YPLGVG motif of NiV M is important for budding, perhaps involving a novel mechanism (35). Our rationale for investigating the corresponding, conserved motifs present within the Ebola and Marburg virus VP40 proteins was based primarily on these findings with NiV. In addition, Ebola virus VP40 motif maps close to the hinge region separating the N- and C-terminal domains of VP40 (4). Thus, the 96LPLGVA101 motif of eVP40 is predicted to be important for the overall stability and function of VP40 during egress. Findings presented here indicate that disruption of these filovirus VP40 motifs results in a severe defect in VLP budding, due in part to impairment in overall VP40 structure, stability and/or intracellular localization.  相似文献   

13.
TEL1 is important in Saccharomyces cerevisiae telomere maintenance, and its kinase activity is required. Tel1p associates with telomeres in vivo, is enriched at short telomeres, and enhances the binding of telomerase components to short telomeres. However, it is unclear how the kinase activity and telomere association contribute to Tel1p''s overall function in telomere length maintenance. To investigate this question, we generated a set of single point mutants and a double point mutant (tel1KD) of Tel1p that were kinase deficient and two Xrs2p mutants that failed to bind Tel1p. Using these separation-of-function alleles in a de novo telomere elongation assay, we found, surprisingly, that the tel1KD allele and xrs2 C-terminal mutants were both partially functional. Combining the tel1KD and xrs2 C-terminal mutants had an additive effect and resembled the TEL1 null (tel1Δ) phenotype. These data indicate that Tel1p has two separate functions in telomere maintenance and that the Xrs2p-dependent recruitment of Tel1p to telomeres plays an important role even in the absence of its kinase activity.The telomere is a highly ordered complex of proteins and DNA found at the ends of linear chromosomes that functions to protect the ends and prevents them from being recognized as double-strand DNA breaks (51). Telomeres shorten gradually due to incomplete replication (1, 20), and this shortening is counteracted by telomerase, which elongates telomeres (18, 19).Saccharomyces cerevisiae telomeres are composed of 300 ± 50 bp of the sequence TG1-3/C1-3A. The yeast telomerase complex consists of Est2p (catalytic subunit), the RNA component TLC1, and two accessory proteins, Est1p and Est3p (50). Cells deficient for any of these telomerase components undergo progressive telomere shortening and a simultaneous decrease in growth rate, described as senescence (24, 27). Typically, a small fraction of cells, termed survivors, escape senescence and maintain telomere length by utilizing RAD52-dependent recombination (24, 26).In addition to the telomerase complex, a number of yeast proteins are important in maintaining telomere length and integrity. These include Tel1p and Mec1p, the yeast homologues of mammalian ATM and ATR, respectively (39). While deletion of TEL1 results in short but stable telomeres, MEC1 deletion has little effect on average telomere length. However, cells lacking TEL1 that have a mutant mec1-21 allele undergo senescence, similar to telomerase null cells (36), suggesting that MEC1 plays a minor but essential role in telomere length maintenance in tel1Δ cells. It has been shown that the protein kinase activities of Tel1p and Mec1p are essential in telomere maintenance, since tel1KD cells have short telomeres and tel1Δ mec1KD cells undergo senescence (29).In current models, Tel1p acts to maintain telomere length by regulating the access of telomerase to short telomeres. TEL1 is required for the association of Est1p and Est2p with telomeres in the late S/G2 phase of the cell cycle (16), the time when telomeres are elongated (9, 31). Additionally, in both yeast and mammalian cells, telomerase preferentially elongates the shortest telomeres (22, 30, 47). Therefore, TEL1 seems to be required mainly for the association of telomerase to short telomeres in yeast. Indeed, Tel1p preferentially binds to short telomeres (4, 21, 38) and is essential for the increased association of Est1p and Est2p to short telomeres during late S/G2 (38). However, the kinase activity of Tel1p is not required for the telomere association (21). In addition to its role in telomerase recruitment, TEL1 may also regulate telomere length by enhancing the processivity of telomerase at short telomeres (7).The Mre11p, Rad50p, and Xrs2p (MRX) complex also plays important roles in telomere maintenance. Cells lacking any one of these components (mrxΔ) have short and stable telomeres. Since combining mrxΔ with tel1Δ has no synergistic effect on telomere shortening and mrxΔ mec1Δ cells undergo senescence, it was proposed that the MRX complex and Tel1p function in the same telomere maintenance pathway (37). In agreement with this model, the C-terminal region of Xrs2p is essential in recruiting Tel1p both to double-strand breaks (32) and to short telomeres (38). Interestingly, the mammalian functional homologue of Xrs2p, NBS1, interacts with ATM via its extreme C terminus (13), suggesting that the recruitment of Tel1p to telomeres and the recruitment of ATM to DNA damage sites are conserved.It remains a question what exact roles the kinase activity of Tel1p and its telomere binding play in telomere maintenance. Tel1p''s telomere maintenance function seems to be dependent on its kinase activity, since tel1KD cells have short telomeres (29). It has been proposed that Tel1p may regulate the recruitment of Est1p, and thus the rest of the telomerase complex (12, 23, 54), to telomeres by phosphorylating Cdc13p (3, 48). Other experiments suggest the association of Tel1p to the telomere plays a major role. The preferential binding of Tel1p to short telomeres is lost in xrs2-664 cells (38), which lack the C-terminal 190 amino acids of Xrs2p and have short telomeres, similar to xrs2Δ (41). It has been suggested that the association of Tel1p to telomeres is required for its substrate phosphorylation and, therefore, telomere length maintenance (3, 39).To further analyze the functions of Tel1p in telomere maintenance, we generated a novel kinase-dead allele of TEL1 and new alleles of XRS2 that do not interact with Tel1p. Through these separation-of-function mutants, we show that both sets of alleles are partially active in a de novo telomere elongation assay. However, combining both the tel1KD and either of the Tel1p interaction-deficient xrs2 alleles resulted in a phenotype resembling the tel1Δ phenotype, suggesting that Tel1p has kinase-dependent and kinase-independent, but telomere binding-dependent, functions in telomere maintenance.  相似文献   

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

18.
19.
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