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.
4.
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   

5.
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   

6.
JC virus (JCV) is a human polyomavirus and the causative agent of the fatal demyelinating disease progressive multifocal leukoencephalopathy (PML). JCV infection of host cells is dependent on interactions with cell surface asparagine (N)-linked sialic acids and the serotonin 5-hydroxytryptamine2A receptor (5-HT2AR). The 5-HT2AR contains five potential N-linked glycosylation sites on the extracellular N terminus. Glycosylation of other serotonin receptors is essential for expression, ligand binding, and receptor function. Also, glycosylation of cellular receptors has been reported to be important for JCV infection. Therefore, we hypothesized that the 5-HT2AR N-linked glycosylation sites are required for JCV infection. Treatment of 5-HT2AR-expressing cells with tunicamycin, an inhibitor of N-linked glycosylation, reduced JCV infection. Individual mutation of each of the five N-linked glycosylation sites did not affect the capacity of 5-HT2AR to support JCV infection and did not alter the cell surface expression of the receptor. However, mutation of all five N-linked glycosylation sites simultaneously reduced the capacity of 5-HT2AR to support infection and altered the cell surface expression. Similarly, tunicamycin treatment reduced the cell surface expression of 5-HT2AR. Mutation of all five N-linked glycosylation sites or tunicamycin treatment of cells expressing wild-type 5-HT2AR resulted in an altered electrophoretic mobility profile of the receptor. Treatment of cells with PNGase F, to remove N-linked oligosaccharides from the cell surface, did not affect JCV infection in 5-HT2AR-expressing cells. These data affirm the importance of 5-HT2AR as a JCV receptor and demonstrate that the sialic acid component of the receptor is not directly linked to 5-HT2AR.The initial interaction between virus and host occurs via molecular interactions of viral attachment proteins and receptors on host cells. Therefore, receptor recognition is a critical host cell determinant and may play a key regulatory role in viral pathogenesis. The polyomavirus JC virus (JCV) is a ubiquitous human pathogen (21, 25, 32) that is initially subclinical yet establishes a persistent infection in the kidney (11). In immunosuppressed individuals JCV can become reactivated, leading to infection in the central nervous system (CNS) (13-15, 20), where the virus specifically targets glial cells, including astrocytes and the myelin-producing cells, oligodendrocytes (40, 48). JCV infection and cytolytic destruction of oligodendroglia cause the fatal disease progressive multifocal leukoencephalopathy (PML) (1, 22). The most common cause of PML is associated with human immunodeficiency virus (HIV) and AIDS (10, 23). However, in recent years PML has been reported in patients receiving immunosuppressive therapies for autoimmune diseases such as Crohn''s disease (44), multiple sclerosis (MS) (24, 26, 28, 47), systemic lupus erythematosus (5, 33), and rheumatoid arthritis (5, 19, 37). The prognosis of PML is bleak, as the disease progresses rapidly and usually proves fatal within 1 year of the onset of symptoms. While current treatment options for PML are limited (23), recent studies suggest that mirtazapine, a serotonin receptor antagonist, may be capable of slowing the progression of PML (6, 27, 45, 46).JCV has a nonenveloped, icosahedral capsid that encapsidates a circular double-stranded DNA (dsDNA) genome (39). JCV attachment to cells is mediated by an N-linked glycoprotein with either α(2,3)- or α(2,6)-linked sialic acid (16, 31), suggesting that N-linked glycosylation of cellular receptors is important for JCV infection. N-linked glycosylation is a posttranslational process by which oligosaccharides are added to asparagine residues, and this modification is important for protein processing, folding, expression, and function (43). Previous studies from our laboratory revealed that the JCV also requires the serotonin 5-hydroxytryptamine2A receptor (5-HT2AR) to mediate JCV infection (18, 35, 38), while others report that JCV infection can occur in the absence of 5-HT2AR (7, 8). 5-HT2AR is a seven-transmembrane-spanning G-protein-coupled receptor that belongs to a large family of 5-HT serotonin receptors. 5-HT2AR is abundantly expressed on cells in the brain (4), including glial cells (3), and in the kidney (4), which parallels the sites of JCV infection. N-linked glycosylation plays a key regulatory role in the function of serotonin receptors. Mutation of N-linked glycosylation sites in human 5-HT3AR and 5-HT5AR results in decreased expression at the plasma membrane, which is critical for receptor function (17, 34). N-linked glycosylation of murine 5-HT3AR regulates plasma membrane targeting, ligand binding, Ca2+ flux, and receptor trafficking (36), suggesting that glycosylation is essential for expression and function of serotonin receptors.While previous studies have concluded that JCV utilizes an N-linked glycoprotein with α(2,3)-linked sialic acid (31) or α(2,6)-linked sialic acid (16) and 5-HT2AR (18) to initiate infection in host cells, the mechanism(s) by which JCV engages its cellular receptors and the importance of receptor glycosylation remain unclear. 5-HT2AR contains potential asparagine (N)-linked glycosylation sites, five of which are predicted to be expressed in the extracellular amino-terminal region, where they could be accessible to the virus (2). The goal of this study was to determine whether potential N-linked glycosylation sites expressed in 5-HT2AR are required for JCV infection. We found that N-linked glycosylation of 5-HT2AR is important for receptor expression but not necessary for JCV infection.  相似文献   

7.
An attenuated derivative of simian immunodeficiency virus strain 239 deleted of V1-V2 sequences in the envelope gene (SIV239ΔV1-V2) was used for vaccine/challenge experiments in rhesus monkeys. Peak levels of viral RNA in plasma of 104 to 106.5 copies/ml in the weeks immediately following inoculation of SIV239ΔV1-V2 were 10- to 1,000-fold lower than those observed with parental SIV239 (∼107.3 copies/ml). Viral loads consistently remained below 200 copies/ml after 8 weeks of infection by the attenuated SIV239ΔV1-V2 strain. Viral localization experiments revealed large numbers of infected cells within organized lymphoid nodules of the colonic gut-associated lymphoid tissue at 14 days; double-labeling experiments indicated that 93.5% of the virally infected cells at this site were positive for the macrophage marker CD68. Cellular and humoral immune responses measured principally by gamma interferon enzyme-linked immunospot and neutralization assays were variable in the five vaccinated monkeys. One monkey had responses in these assays comparable to or only slightly less than those observed in monkeys infected with parental, wild-type SIV239. Four of the vaccinated monkeys, however, had low, marginal, or undetectable responses in these same assays. These five vaccinated monkeys and three naïve control monkeys were subsequently challenged intravenously with wild-type SIV239. Three of the five vaccinated monkeys, including the one with strong anti-SIV immune responses, were strongly protected against the challenge on the basis of viral load measurements. Surprisingly, two of the vaccinated monkeys were strongly protected against SIV239 challenge despite the presence of cellular anti-SIV responses of low-frequency and low-titer anti-SIV antibody responses. These results indicate that high-titer anti-SIV antibody responses and high-frequency anti-SIV cellular immune responses measurable by standard assays from the peripheral blood are not needed to achieve strong vaccine protection, even against a difficult, neutralization-resistant strain such as SIV239.The characteristics of human immunodeficiency virus type 1 (HIV-1) infection suggest major difficulty for the development of a preventive vaccine (19, 23). Pessimism regarding the prospects for a vaccine is derived at least in part from the ability of HIV-1 to continually replicate in the face of apparently strong host immune responses, resistance to antibody-mediated neutralization, and the extensive sequence diversity in field strains of the virus. Lack of knowledge regarding the key components of a protective immune response also remains a major scientific obstacle. Vaccine/challenge experiments with macaque monkeys have been used to evaluate the properties and relative effectiveness of different vaccine approaches and to gauge the formidable nature of these difficulties.One lesson that has been learned from vaccine/challenge experiments with macaque monkeys is the importance of challenge strain on outcome. Vaccinated monkeys that have been challenged with strains of simian immunodeficiency virus (SIV) with an HIV-1 envelope (SHIV) have almost invariably exhibited strong, long-term protection against disease, irrespective of the nature of the vaccine. Even peptide immunogens have protected against SHIV-induced disease (6, 12, 38). Vaccine approaches that have protected against SHIV challenge include DNA (5, 13), recombinant poxvirus (4), recombinant adenovirus (57), other viral recombinants (18, 55), prime and boost protocols (3, 53, 65), and purified protein (10, 64). Vaccine protection against pathogenic SIV strains such as SIV239, SIV251, and SIV-E660 has been much more difficult to achieve (2, 11, 27, 63). The identical replication-defective gag-recombinant adenovirus that provided strong protection against SHIV challenge (57) provided little or no protection against SIV239 challenge (11). Disappointing levels of protection against SIV have often been observed in the face of apparently robust vaccine-induced immune responses (see, for example, Vogel et al. [63] and Casimiro et al. [11]). Some partial vaccine protections against these SIV strains have been achieved by recombinant poxvirus (7, 50), replication-competent recombinant adenovirus (51), replication-defective adenovirus (66), recombinant poliovirus (15), recombinant Venezuelan equine encephalitis virus (18), and recombinant Sendai virus (44).Differences between the biological properties of the SIV strains and those of the SHIV strains used for the above-mentioned studies provide clues as to what may be responsible for the differences in outcome. These SIV strains are difficult to neutralize (26, 34), use CCR5 as a coreceptor for entry into cells (21, 52), and induce a chronic, progressive disease course (17), and this course is independent of the infectious dose (17). The SHIV strains used for the above-mentioned studies are easier to neutralize, use CXCR4 for entry, and induce an acute decline in CD4 counts, and the disease course is dose dependent (29, 30, 48, 54). These SIV strains, like HIV-1 in humans, exhibit a marked preference for CD4+ CCR5+ memory cells, in contrast to the acutely pathogenic SHIV strains which principally target naïve cells (48).Live, attenuated strains of SIV have provided the strongest vaccine protection by far against SIV challenge. Although clinical use of a live, attenuated HIV vaccine is not being considered, understanding the basis of the strong protection afforded by live, attenuated SIV strains remains an important research objective for the insights that can be provided. Most of the attenuated SIV strains that have been used lack a functional nef gene (16, 31, 58, 67). Shacklett et al. (56) used an attenuated SIV strain with modifications in the gp41 transmembrane protein for protection. Here, we describe strong vaccine protection by a replication-competent SIV strain lacking 100 amino acids from the essential gp120 envelope protein in the absence of overtly robust immune responses.  相似文献   

8.
9.
10.
11.
12.
13.
14.
15.
16.
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   

17.
There are two protein primers involved in picornavirus RNA replication, VPg, the viral protein of the genome, and VPgpUpUOH. A cis-acting replication element (CRE) within the open reading frame of poliovirus (PV) RNA allows the viral RNA-dependent RNA polymerase 3DPol to catalyze the conversion of VPg into VPgpUpUOH. In this study, we used preinitiation RNA replication complexes (PIRCs) to determine when CRE-dependent VPg uridylylation occurs relative to the sequential synthesis of negative- and positive-strand RNA. Guanidine HCl (2 mM), a reversible inhibitor of PV 2CATPase, prevented CRE-dependent VPgpUpUOH synthesis and the initiation of negative-strand RNA synthesis. VPgpUpUOH and nascent negative-strand RNA molecules were synthesized coincident in time following the removal of guanidine, consistent with PV RNA functioning simultaneously as a template for CRE-dependent VPgpUpUOH synthesis and negative-strand RNA synthesis. The amounts of [32P]UMP incorporated into VPgpUpUOH and negative-strand RNA products indicated that 100 to 400 VPgpUpUOH molecules were made coincident in time with each negative-strand RNA. 3′-dCTP inhibited the elongation of nascent negative-strand RNAs without affecting CRE-dependent VPg uridylylation. A 3′ nontranslated region mutation which inhibited negative-strand RNA synthesis did not inhibit CRE-dependent VPg uridylylation. Together, the data implicate 2CATPase in the mechanisms whereby PV RNA functions as a template for reiterative CRE-dependent VPg uridylylation before and during negative-strand RNA synthesis.A common feature of positive-strand RNA viruses is the asymmetric replication of viral RNA. Poliovirus (PV) RNA replication has been studied extensively; however, it remains to be determined exactly how the synthesis of negative-strand RNA and that of positive-strand RNA are mechanistically distinct, culminating in the synthesis of greater amounts of positive-strand than negative-strand RNA (2). A cis-acting replication element (CRE) within the 2C open reading frame of PV RNA functions as a template for the conversion of the viral protein of the genome (VPg) into VPgpUpUOH (24, 26, 37). 3D polymerase (3DPol), in concert with other viral proteins, catalyzes the conversion of VPg into VPgpUpUOH on CRE RNA templates (22). It remains to be determined whether the tyrosine hydroxyl of VPg (14, 20, 21), the 3′ hydroxyl of VPgpUpUOH (22, 23, 43), or both (38) are used to prime negative-strand RNA synthesis. It would be informative to know whether VPg is converted into VPgpUpUOH before, during, and/or after the initiation of viral negative-strand RNA synthesis. Conversion of VPg into VPgpUpUOH before the initiation of negative-strand RNA synthesis would be consistent with the possibility that it primes the initiation of negative-strand RNA synthesis. Conversely, if VPg were not converted into VPgpUpUOH until after the initiation of negative-strand RNA synthesis, VPgpUpUOH could not possibly participate in the initiation of negative-strand RNA synthesis. Also, because multiple copies of VPgpUpUOH are necessary to prime reiterative initiation of positive-strand RNA synthesis (35), VPg is most likely converted into abundant amounts of VPgpUpUOH before the initiation of positive-strand RNA synthesis.PV preinitiation RNA replication complexes (PIRCs) were used in this study to examine when VPg is converted into VPgpUpUOH. PIRCs assemble and accumulate when PV mRNA is translated in reaction mixtures containing cytoplasmic extracts from uninfected HeLa cells and 2 mM guanidine HCl, a reversible inhibitor of viral RNA replication (5). The viral replication proteins expressed from the viral mRNA interact with lipid membranes in the cytoplasmic extracts to form RNA replication complexes (RCs) similar to those in infected cells (12). PIRCs convert VPg into VPgpUpUOH and initiate viral RNA replication when they are isolated from reaction mixtures containing guanidine and resuspended in reaction mixtures lacking guanidine (6, 19). Guanidine HCl functions as a reversible inhibitor of PV RNA replication, both in cells (11) and in cell-free translation-replication reactions (6). In cells, PV RNA RCs fail to immediately initiate RNA replication following the removal of guanidine HCl (11). Rather, PV RCs formed in the presence of guanidine in cells appear to be translocated to a region of the cytoplasm where the RCs and their contents may be recycled and/or destroyed (11), possibly by autophagic vesicles (17). Recycling and/or destruction of RCs by autophagic vesicles would preclude their function upon the removal of guanidine. PIRCs, which form in the presence of guanidine during the translation of PV mRNA in cytoplasmic extracts of HeLa cells, immediately initiate both negative-strand RNA synthesis and CRE-dependent VPg uridylylation upon the removal of guanidine (6, 19). Viral RNA replication and VPgpUpUOH synthesis are monitored by the incorporation of radiolabeled UTP (19-21). It is important to note that RNA replication in the context of PIRCs is artificial in that the PIRCs are stalled with guanidine and purified and then the guanidine block is removed. Despite this artificiality, the mechanisms of RNA replication within PIRCs appear to reliably represent the mechanisms of RNA replication in cells. There are several advantageous features of the PIRC experimental system: viral RNA replication is synchronous and sequential, with negative-strand RNA being made before positive-strand RNA (6); viral RNA replication is asymmetric, with an excess of positive-strand RNA being made from each negative-strand template; VPg is converted into VPgpUpUOH in a CRE-dependent manner (20, 21); and the reaction conditions, including nucleoside triphosphate concentrations, are easily manipulated (38). Importantly, PIRCs contain all of the viral proteins associated with RNA replication and RNA replication by PIRCs faithfully mimics the asymmetric replication of PV RNA observed in cells.PV protein 2C, the target of guanidine HCl (30), is a critical but poorly understood component of PIRCs and RNA RCs in cells. PV protein 2C has an NH-terminal amphipathic helix which interacts with cellular membranes (40), a central ATPase domain where guanidine-resistant and guanidine-dependent mutations arise (31, 32), a cysteine-rich zinc binding motif (29), and a COOH-terminal RNA binding domain (34) which appears to work in concert with amino acid residues at the NH terminus to bind RNA. 2CATPase can oligomerize (1, 41), anchoring viral replication proteins and RNA templates within membranous RCs (4). The ability of guanidine HCl to reversibly inhibit both CRE-dependent VPg uridylylation and negative-strand RNA synthesis implicates 2CATPase in the mechanisms by which PV RNA functions coordinately as a template for both RNA replication and CRE-dependent VPgpUpUOH synthesis (19, 21).In this study, we found that VPg was converted into VPgpUpUOH before and during negative-strand RNA synthesis and that 2CATPase activity, in the context of membranous PIRCs, allowed PV RNA to function simultaneously as a template for CRE-dependent VPg uridylylation and as a template for negative-strand RNA synthesis. We discuss how picornaviruses coordinate the synthesis of nucleotidylylated protein primers with other steps of viral RNA replication.  相似文献   

18.
19.
20.
The recently discovered Canis familiaris papillomavirus (PV) type 2 (CfPV2) provides a unique opportunity to study PV gene functions in vitro and in vivo. Unlike the previously characterized canine oral PV, CfPV2 contains an E5 open reading frame and is associated with progression to squamous cell carcinoma. In the current study, we have expressed and characterized the CfPV2-encoded E5 protein, a small, hydrophobic, 41-amino-acid polypeptide. We demonstrate that, similar to the E5 protein from high-risk human PV type 16, the CfPV2 E5 protein is localized in the endoplasmic reticulum (ER) and that its expression decreases keratinocyte proliferation and cell life span. E5 expression also increases the percentage of cells in the G1 phase of the cell cycle, with a concomitant decrease in the percentage of cells in S phase. To identify a potential mechanism for E5-mediated growth inhibition from the ER, we developed a real-time PCR method to quantify the splicing of XBP1 mRNA as a measure of ER stress. We found that the CfPV2 E5 protein induced ER stress and that this, as well as the observed growth inhibition, is tempered significantly by coexpression of the CfPV2 E6 and E7 genes. It is possible that the spatial/temporal regulation of E6/E7 gene expression during keratinocyte differentiation might therefore modulate E5 activity and ER stress.Papillomaviruses (PVs) are a large group of DNA tumor viruses that infect differentiated cutaneous and mucosal epithelia in a wide variety of mammalian species. There are nearly 200 types of human PVs (HPVs) (61), some of which are termed high risk (e.g., HPV type 16 [HPV-16]) and have the potential to immortalize primary cells and facilitate malignant progression to cervical cancer (52). An estimated 20 million cases of HPV infection occur each year in the United States alone, and cervical cancer is the second most common cause of cancer deaths among women worldwide. In general, PV infections are species specific, making it impossible to study the in vivo life cycle of HPV and the roles of its encoded proteins in viral replication and tumorigenesis. However, a few animal models do exist and the canine oral PV (COPV) has been helpful in mimicking certain biological properties of the high-risk mucosatropic HPVs, leading to the development of highly effective prophylactic vaccines (39, 49, 56). Although COPV mimics the mucosal tropism of the high-risk HPVs, it rarely progresses to cancer and lacks one of the early viral genes that may play an important role in tumorigenesis, E5. Recently, a new canine PV (Canis familiaris PV type 2 [CfPV2]) was isolated from the footpads of dogs (43). Unlike COPV, CfPV2 induces epidermal tumors and, when persistent, these benign infections progress to squamous cell carcinoma and metastasize widely. CfPV2 also encodes an E5 protein. In general, PV E5 proteins are small hydrophobic oncoproteins that localize to the endoplasmic reticulum (ER) or Golgi membranes (11, 16) but have limited amino acid sequence homology. Numerous cellular binding partners have been described for HPV-16 E5 proteins, including the V-ATPase 16-kDa subunit (1, 16), the nuclear import protein karyopherin beta 3 (25), the ER-resident protein Bap31 (40), proteins involved in zinc transport (ZnT1, EVER1, and EVER2) (27, 35), erbB4 (24), and HLA I (2). The HPV-16 E5 protein alters signaling pathways, predominantly the epidermal growth factor receptor (EGFR) pathway (17, 21, 46, 58); induces koilocytosis in cooperation with the E6 protein (26); and alters the plasma membrane expression of caveolin (47), HLA (3), and ganglioside GM1 (47). The last two changes might explain the ability of HPV-16-infected cells to circumvent detection by the host immune response and initiate tumor formation (3, 4, 21, 36, 46, 47).To provide a foundation for future in vivo studies, we initiated a series of in vitro experiments to define the intracellular localization and biological activity of CfPV2 E5. The current study demonstrates that CfPV2 E5 exhibits several properties of the HPV-16 E5 protein, including ER localization and inhibition of cell proliferation. A novel finding is that CfPV2 E5 activates the ER stress-signaling pathway, which may explain some of E5''s growth-related activities.  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