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.
The pseudorabies virus (PrV) gene homologous to herpes simplex virus type 1 (HSV-1) UL53, which encodes HSV-1 glycoprotein K (gK), has recently been sequenced (J. Baumeister, B. G. Klupp, and T. C. Mettenleiter, J. Virol. 69:5560–5567, 1995). To identify the corresponding protein, a rabbit antiserum was raised against a 40-kDa glutathione S-transferase–gK fusion protein expressed in Escherichia coli. In Western blot analysis, this serum detected a 32-kDa polypeptide in PrV-infected cell lysates as well as a 36-kDa protein in purified virion preparations, demonstrating that PrV gK is a structural component of virions. After treatment of purified virions with endoglycosidase H, a 34-kDa protein was detected, while after incubation with N-glycosidase F, a 32-kDa protein was specifically recognized. This finding indicates that virion gK is modified by N-linked glycans of complex as well as high-mannose type. For functional analysis, the UL53 open reading frame was interrupted after codon 164 by insertion of a gG-lacZ expression cassette into the wild-type PrV genome (PrV-gKβ) or by insertion of the bovine herpesvirus 1 gB gene into a PrV gB genome (PrV-gKgB). Infectious mutant virus progeny was obtained only on complementing gK-expressing cells, suggesting that gK has an important function in the replication cycle. After infection of Vero cells with either gK mutant, only single infected cells or small foci of infected cells were visible. In addition, virus yield was reduced approximately 30-fold, and penetration kinetics showed a delay in entry which could be compensated for by phenotypic gK complementation. Interestingly, the plating efficiency of PrV-gKβ was similar to that of wild-type PrV on complementing and noncomplementing cells, pointing to an essential function of gK in virus egress but not entry. Ultrastructurally, virus assembly and morphogenesis of PrV gK mutants in noncomplementing cells were similar to wild-type virus. However, late in infection, numerous nucleocapsids were found directly underneath the plasma membrane in stages typical for the entry process, a phenomenon not observed after wild-type virus infection and also not visible after infection of gK-complementing cells. Thus, we postulate that presence of gK is important to inhibit immediate reinfection.Herpesvirions are complex structures consisting of a nucleoprotein core, capsid, tegument, and envelope. They comprise at least 30 structural proteins (35). Pseudorabies virus (PrV), a member of the Alphaherpesvirinae, is an economically important animal pathogen, causing Aujeszky’s disease in swine. It is also highly pathogenic for most other mammals except higher primates, including humans (28, 45), and a wide range of cultured cells from different species support productive virus replication, reflecting the wide in vivo host range. Envelope glycoproteins play major roles in the early and late interactions between virion and host cell. They are required for virus entry and participate in release of free virions and viral spread by direct cell-to-cell transmission (27, 37). For PrV, 10 glycoproteins, designated gB, gC, gD, gE, gG, gH, gI, gL, gM, and gN, have been characterized (20, 27); these glycoproteins are involved in the attachment of virion to host cell (gC and gD), fusion of viral envelope and cellular cytoplasmic membrane (gB, gD, gH, and gL), spread from infected to noninfected cells (gB, gE, gH, gI, gL, and gM), and egress (gC, gE, and gI) (27, 37). Homologs of these glycoproteins are also present in other alphaherpesviruses (37). The gene coding for a potential 11th PrV glycoprotein, gK, has been described recently (3), but the protein and its function have not been identified.The product of the homologous UL53 open reading frame (ORF) of herpes simplex virus type 1 (HSV-1) is gK (13, 32). gK was detected in nuclear membranes and in membranes of the endoplasmic reticulum but was not observed in the plasma membrane (14). Also, it did not appear to be present in purified virion preparations (15). The latter result was surprising since earlier studies identified several mutations in HSV-1 gK resulting in syncytium-inducing phenotypes (7, 14), which indicates participation of gK in membrane fusion events during HSV-1 infection. Moreover, HSV-1 mutants in gK exhibited a delayed entry into noncomplementing cells, which is difficult to reconcile with absence of gK from virions (31). Mutants deficient for gK expression have been isolated and investigated by different groups (16, 17). Mutant F-gKβ carries a lacZ gene insertion in the HSV-1 strain F gK gene, which interrupts the ORF after codon 112 (16). In mutant ΔgK, derived from HSV-1 KOS, almost all of the UL53 gene was deleted (17). Both mutants formed small plaques on Vero cells, and virus yield was reduced to an extent which varied with the different confluencies of the infected cells, cell types, and mutants used for infection. However, both HSV-1 gK mutants showed a defect in efficient translocation of virions from the cytoplasm to the extracellular space, and only a few enveloped virions were present in the extracellular space after infection of Vero cells (16, 17). The authors therefore suggested that HSV-1 gK plays a role in virion transport during egress.Different routes of final envelopment and egress of alphaherpesvirions are discussed. It has been suggested that HSV-1 nucleocapsids acquire their envelope at the inner nuclear membrane and are transported as enveloped particles through the endoplasmic reticulum to the Golgi stacks, where glycoproteins are modified in situ during transport (5, 6, 19, 39), although other potential egress pathways cannot be excluded (4). In contrast, maturation of varicella-zoster virus and PrV involves primary envelopment at the nuclear membrane, followed by release of nucleocapsids into the cytoplasm and secondary envelopment in the trans-Golgi area (10, 12, 43). Final egress of virions appears to occur via transport vesicles containing one or more virus particles by fusion of vesicle and cell membrane. The possibility of different routes of virion egress is supported by studies of other proteins involved in egress, e.g., the UL20 proteins of HSV-1 and PrV and the PrV UL3.5 protein, which lacks a homolog in the HSV-1 genome (1, 8, 9). In UL20-negative HSV-1, virions accumulated in the perinuclear cisterna of Vero cells (1), while PrV UL20 virions accumulated and were retained in cytoplasmic vesicles (9). PrV UL3.5 is important for budding of nucleocapsids into Golgi-derived vesicles during secondary envelopment (8). Thus, there appear to be profound differences in the egress pathways. Since HSV-1 gK was also implicated in egress, we were interested in identifying the PrV homolog and analyzing its function.  相似文献   

5.
In order to analyze bonding contacts that stabilize the virion or promote capsid assembly, bovine papillomavirus (BPV) virions were subjected to buffer conditions known to disrupt polyomavirus virions. At physiologic ionic strength, incubation with dithiothreitol (DTT), EGTA, or DTT plus EGTA did not disrupt BPV virions as determined by electron microscopy. However, incubation of virions with DTT rendered the BPV L1 protein susceptible to trypsin cleavage at its carboxy terminus and rendered the genome susceptible to digestion with DNase I. When DTT-treated BPV virions were analyzed by analytical ultracentrifugation, they sedimented at 230S compared with 273S for untreated virions, suggesting a capsid shell expansion. Incubation with EGTA had no effect on trypsin or DNase I sensitivity and only a small effect upon the virion S value. A single cysteine residue conserved among BPV and human papillomavirus (HPV) L1 proteins resides within the trypsin-sensitive carboxy terminus of L1, which is required for capsid assembly. A recombinant HPV type 11 L1 protein, which was purified after expression in Escherichia coli and which has a Cys-to-Gly change at this position (Cys424), formed pentamers; however, unlike the wild-type protein, these mutant pentamers could no longer assemble in vitro into capsid-like structures. These results indicate an important role for interpentamer disulfide bonds in papillomavirus capsid assembly and disassembly and suggest a mechanism of virus uncoating in the reducing environment of the cytoplasm.Papovaviruses are small nonenveloped DNA viruses which include murine polyomavirus, simian virus 40 (SV40), and the papillomaviruses. These viruses have a common capsid structure composed of 72 capsomeric subunits arranged in a T=7d icosahedral lattice. For polyomaviruses, the capsomeres are pentamers of the VP1 structural protein, while for the papillomaviruses the capsomeres are pentamers of the L1 protein (1, 15). Although atomic-resolution structures for murine polyomavirus and SV40 (14, 24, 25) and high-resolution cryoelectron microscopic structures for several papillomaviruses (1, 2, 4, 26) are now known, much less is known about the biology of virion assembly and disassembly.For polyomaviruses, virion disruption experiments initially indicated the importance of disulfide bonds and bound calcium in capsid stability (57, 28). In agreement with these observations, the in vitro self-assembly of the recombinant polyomavirus VP1 protein supported the importance of calcium and ionic bonds in the formation of capsids (19, 20). Most recently, crystallographic data have explicitly identified the calcium binding sites, disulfide-bonded residues, and other interactions which link capsomeres in virions (14, 24, 25). A crystallographic structure is not yet available for any papillomavirus. Virus-like particles (VLPs) generated by expressing the L1 protein in insect cells with recombinant baculovirus vectors (12, 17, 27) or in mammalian cells with vaccinia virus vectors (11, 29) have been useful for both cryoelectron microscopic structural analysis and to assess the assembly and disassembly properties for L1-containing capsids. For example, Sapp et al. (21) determined that human papillomavirus type 33 (HPV33) VLPs were disassembled by dithiothreitol (DTT) treatment, suggesting disulfide bonds between a subset of capsomeres in agreement with the observations of Doorbar and Gallimore with native HPV1 virions (8).The recombinant expression of HPV11 L1 in Escherichia coli (13) now permits a similar analysis of capsid assembly for papillomaviruses as previously carried out for polyomaviruses (19, 20, 23). In order to assess their biological significance, in vitro assembly conditions for the recombinant L1 protein require correlation with conditions that affect the stability of native papillomavirus virions. However, although VLP disruption experiments have been performed, a systematic analysis of papillomavirus virion disruption similar to that carried out for polyomavirus has not been described. Therefore, we have used purified bovine papillomavirus type 1 (BPV) virions to test conditions affecting virion stability. Based on these results, disulfide bonds were found to be critical for virion integrity, and their reduction led to a substantial conformational change in the capsid which may be a precursor state for disassembly in vivo.  相似文献   

6.
7.
8.
The membrane-interacting abilities of three sequences representing the putative fusogenic subdomain of the Ebola virus transmembrane protein have been investigated. In the presence of calcium, the sequence EBOGE (GAAIGLAWIPYFGPAAE) efficiently fused unilamellar vesicles composed of phosphatidylcholine, phosphatidylethanolamine, cholesterol, and phosphatidylinositol (molar ratio, 2:1:1:0.5), a mixture that roughly resembles the lipid composition of the hepatocyte plasma membrane. Analysis of the lipid dependence of the process demonstrated that the fusion activity of EBOGE was promoted by phosphatidylinositol but not by other acidic phospholipids. In comparison, EBOEA (EGAAIGLAWIPYFGPAA) and EBOEE (EGAAIGLAWIPYFGPAAE) sequences, which are similar to EBOGE except that they bear the negatively charged glutamate residue at the N terminus and at both the N and C termini, respectively, induced fusion to a lesser extent. As revealed by binding experiments, the glutamate residue at the N terminus severely impaired peptide-vesicle interaction. In addition, the fusion-competent EBOGE sequence did not associate significantly with vesicles lacking phosphatidylinositol. Tryptophan fluorescence quenching by vesicles containing brominated phospholipids indicated that the EBOGE peptide penetrated to the acyl chain level only when the membranes contained phosphatidylinositol. We conclude that binding and further penetration of the Ebola virus putative fusion peptide into membranes might be governed by the nature of the N-terminal residue and by the presence of phosphatidylinositol in the target membrane. Moreover, since insertion of such a peptide leads to membrane destabilization and fusion, the present data would be compatible with the involvement of this sequence in Ebola virus fusion.Ebola virus belongs to the Filoviridae family (23). This human pathogen occasionally causes epidemics of African hemorrhagic fever with a high rate of mortality (8, 23, 37). Little is known about the viral infectivity mechanism, and there is no specific treatment for Ebola virus hemorrhagic fever as yet. The most prominent pathology of Ebola virus infection includes necrosis of liver parenchyma as a direct consequence of virus replication (23). Ebola virus virions are composed of a helical nucleocapsid containing one linear, negative-sense, single-stranded RNA and surrounded by a lipidic envelope derived from the host cell plasma membrane (8, 23). The envelope contains solely one type of highly glycosylated protein (Ebola GP) arranged into oligomers, most probably trimers, which constitute the spikes that protrude from the virion surface (8, 30, 38, 39).The mode of entry of Ebola virus into target cells remains unknown. However it seems likely that the single surface protein Ebola GP is responsible for both receptor binding and membrane fusion during entry into the host cells. Homology analysis of its coding gene-derived sequence has identified several structural features that Ebola GP shares with other envelope fusion proteins derived from oncogenic retroviruses (12, 39). Just recently a detailed analysis has detected a high degree of structural homology between Ebola GP and the Rous sarcoma virus transmembrane protein (12). Several structural elements that might be involved in the ectodomain fusogenic function are shared by these viruses. In particular, there exists in both viruses an amino acid region bounded by cysteines that has at its center a sequence of approximately 16 uncharged and hydrophobic residues. Its location with respect to the viral membrane, the presence of a canonical fusion tripeptide (YFG in Ebola virus), and the fact that this sequence exhibits a high degree of identity among the Filoviridae members suggest that this region might constitute in Ebola virus the fusion peptide that is critical for virion-membrane fusion in the Retroviridae and other families (11, 40, 41).According to the most widely accepted mechanistic model proposed for the initial phase of the viral fusion process, activation of the viral spikes induces the exposure of previously buried hydrophobic fusion peptides in the vicinity of the target cell (5, 43). Further interaction of the viral fusion peptides with the cell membrane would depend mainly on the capacity for binding of these peptides to the membrane lipid components and could eventually trigger the process that brings about the actual merging of the viral and cell membranes via a currently unknown mechanism (41). This fact has justified the development of in vitro studies on the membrane-destabilizing effects of fusion peptides by using representative synthetic peptides of different viruses and model membranes (7, 15, 19, 29).The membrane environment into which the fusion peptide should partition obviously plays an important role in the process. Previous work from this laboratory has focused on the effect of the target membrane composition on viral fusion. Reports from this and other laboratories indicate the existence of conformational changes induced by lipidic components in the membrane-bound human immunodeficiency virus type 1 (HIV-1) fusion peptide (25, 28, 29), and we have identified a fusogenic conformation of the peptide represented by an extended β-type structure (25, 26, 28). The fusogenic interaction of the HIV-1 fusion peptide is, moreover, sensitive to factors that affect gp41 activity in vivo (27). Modulation of viral fusion by lipids has also been observed for complete virions and reconstituted systems fusing with model membranes (6, 24, 42). These observations indicate that enveloped viruses may optimize host interactions during the entry process, not only at the level of the selective binding to cell receptors but also at the level of the envelope fusion and subsequent capsid penetration.Our primary objective in this study was to confirm that the proposed fusogenic sequence for Ebola virus might interact with membranes, destabilize them, and eventually induce fusion. Because Ebola virus infects and replicates very efficiently in the liver, we initially employed as target membranes large unilamellar vesicles (LUV) made of a lipidic mixture that represents the hepatocyte plasma membrane composition (18). Our results demonstrate that this Ebola virus peptide interacts with phosphatidylinositol (PI)-containing membranes and induces vesicle fusion. Moreover, we show that the sequence lacking the negatively charged Glu residue at the N terminus interacts more efficiently with membranes. These data suggest that, similarly to the HIV-1 fusion peptide (2628), the Ebola virus peptide segment under study may be important in viral fusion in vivo.  相似文献   

9.
10.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

11.
HIV-1 Gag can assemble and generate virions at the plasma membrane, but it is also present in endosomes where its role remains incompletely characterized. Here, we show that HIV-1 RNAs and Gag are transported on endosomal vesicles positive for TiVamp, a v-SNARE involved in fusion events with the plasma membrane. Inhibition of endosomal traffic did not prevent viral release. However, inhibiting lysosomal degradation induced an accumulation of Gag in endosomes and increased viral production 7-fold, indicating that transport of Gag to lysosomes negatively regulates budding. This also suggested that endosomal Gag-RNA complexes could access retrograde pathways to the cell surface and indeed, depleting cells of TiVamp-reduced viral production. Moreover, inhibition of endosomal transport prevented the accumulation of Gag at sites of cellular contact. HIV-1 Gag could thus generate virions using two pathways, either directly from the plasma membrane or through an endosome-dependent route. Endosomal Gag-RNA complexes may be delivered at specific sites to facilitate cell-to-cell viral transmission.The production of infectious retroviral particles is an ordered process that includes many steps (for review see Refs. 13). In particular, three major viral components, Gag, the envelope, and genomic RNAs have to traffic inside the cell to reach their assembly site. Viral biogenesis is driven by the polyprotein Gag, which is able to make viral-like particles when expressed alone (4). Upon release, HIV-14 Gag is processed by the viral protease into matrix (MA(p17)), capsid (CA(p24)), nucleocapsid (NC(p7)), p6, and smaller peptides SP1 and SP2. Gag contains several domains that are essential for viral assembly: a membrane binding domain (M) in MA; a Gag-Gag interaction domain in CA; an assembly domain (I) in NC; and a late domain (L) in p6, which recruits the cellular budding machinery. Genomic RNAs are specifically recognized by NC, and they play fundamental roles in viral biogenesis by acting as a scaffold for Gag multimerization (5).It has been demonstrated that retroviruses bud by hijacking the endosomal machinery that sorts proteins into internal vesicles of multivesicular bodies (for review, see Refs. 6, 7). Indeed, these vesicles bud with the same topology as viral particles. Proteins sorted into this pathway are usually destined for degradation in lysosomes, but some can also recycle to the plasma membrane (for review see Refs. 8, 9). They are also frequently ubiquitinated on their cytoplasmic domain (10, 11), allowing their recognition by ESCRT complexes. ESCRT-0 and ESCRT-I recognize ubiquitinated cargo present at the surface of endosomes and recruit other ESCRT complexes (1214). ESCRT-III is believed to function directly in the formation of multivesicular body intralumenal vesicles (12), even though its mechanism of action is currently not understood. Remarkably, Gag L domains interact directly with components of the multivesicular body-sorting machinery (for review see Ref. 15). HIV-1 Gag uses a PTAP motif to bind Tsg101, a component of ESCRT-I (1619), and a YPLTSL motif to interact with Alix, a protein linked to ESCRT-I and -III (2022). Finally, various ubiquitin ligases are also required directly or indirectly during HIV-1 biogenesis (23, 24; for review see Ref. 25).In many cell lines, Gag is found both at the plasma membrane and in endosomes. This has led to the hypothesis that there are several assembly sites for HIV-1 (1, 3). First, Gag can initiate and complete assembly at the plasma membrane. This is thought to occur predominantly in T lymphocytes, and this process is supported by several lines of evidences: (i) disruption of endosomal trafficking with drugs does not prevent viral production (26, 27); (ii) ESCRT complexes can be recruited at the plasma membrane, at sites where Gag accumulates (2830); (iii) Gag can be seen multimerizing and budding from the plasma membrane in live cells (31). Second, Gag could initiate assembly in endosomes, and then traffic to the cell surface to be released. This is mainly supported by the presence of Gag in endosomes in several cell lines (3234), including T cells and more strikingly macrophages (32, 35, 3639). However, we are currently lacking functional experiments addressing the role of this endosomal pool of Gag, and it is still not clear to what extent it contributes to the production of viral particles. Nevertheless, the presence of Gag in endosomes might facilitate recruitment of ESCRT complexes (34, 40), packaging of viral genomic RNAs (32, 41), and incorporation of the envelope (42). It may also be important for polarized budding (43, 44) and to create a viral reservoir in infected cells (45, 46).Despite great progress, the traffic of HIV-1 components is still not fully elucidated. In particular, the transport of the genomic RNAs is poorly understood. In this study, we have used single molecule techniques to investigate the trafficking of HIV-1 RNAs in fixed and live cells, and we show that they are transported on endosomal vesicles. We also obtained functional evidence that Gag and viral RNAs can use at least two trafficking pathways to produce virions, one going directly from the plasma membrane and another one passing through endosomes.  相似文献   

12.
13.
14.
Previous studies have shown that protein-protein interactions among splicing factors may play an important role in pre-mRNA splicing. We report here identification and functional characterization of a new splicing factor, Sip1 (SC35-interacting protein 1). Sip1 was initially identified by virtue of its interaction with SC35, a splicing factor of the SR family. Sip1 interacts with not only several SR proteins but also with U1-70K and U2AF65, proteins associated with 5′ and 3′ splice sites, respectively. The predicted Sip1 sequence contains an arginine-serine-rich (RS) domain but does not have any known RNA-binding motifs, indicating that it is not a member of the SR family. Sip1 also contains a region with weak sequence similarity to the Drosophila splicing regulator suppressor of white apricot (SWAP). An essential role for Sip1 in pre-mRNA splicing was suggested by the observation that anti-Sip1 antibodies depleted splicing activity from HeLa nuclear extract. Purified recombinant Sip1 protein, but not other RS domain-containing proteins such as SC35, ASF/SF2, and U2AF65, restored the splicing activity of the Sip1-immunodepleted extract. Addition of U2AF65 protein further enhanced the splicing reconstitution by the Sip1 protein. Deficiency in the formation of both A and B splicing complexes in the Sip1-depleted nuclear extract indicates an important role of Sip1 in spliceosome assembly. Together, these results demonstrate that Sip1 is a novel RS domain-containing protein required for pre-mRNA splicing and that the functional role of Sip1 in splicing is distinct from those of known RS domain-containing splicing factors.Pre-mRNA splicing takes place in spliceosomes, the large RNA-protein complexes containing pre-mRNA, U1, U2, U4/6, and U5 small nuclear ribonucleoprotein particles (snRNPs), and a large number of accessory protein factors (for reviews, see references 21, 22, 37, 44, and 48). It is increasingly clear that the protein factors are important for pre-mRNA splicing and that studies of these factors are essential for further understanding of molecular mechanisms of pre-mRNA splicing.Most mammalian splicing factors have been identified by biochemical fractionation and purification (3, 15, 19, 3136, 45, 6971, 73), by using antibodies recognizing splicing factors (8, 9, 16, 17, 61, 66, 67, 74), and by sequence homology (25, 52, 74).Splicing factors containing arginine-serine-rich (RS) domains have emerged as important players in pre-mRNA splicing. These include members of the SR family, both subunits of U2 auxiliary factor (U2AF), and the U1 snRNP protein U1-70K (for reviews, see references 18, 41, and 59). Drosophila alternative splicing regulators transformer (Tra), transformer 2 (Tra2), and suppressor of white apricot (SWAP) also contain RS domains (20, 40, 42). RS domains in these proteins play important roles in pre-mRNA splicing (7, 71, 75), in nuclear localization of these splicing proteins (23, 40), and in protein-RNA interactions (56, 60, 64). Previous studies by us and others have demonstrated that one mechanism whereby SR proteins function in splicing is to mediate specific protein-protein interactions among spliceosomal components and between general splicing factors and alternative splicing regulators (1, 1a, 6, 10, 27, 63, 74, 77). Such protein-protein interactions may play critical roles in splice site recognition and association (for reviews, see references 4, 18, 37, 41, 47 and 59). Specific interactions among the splicing factors also suggest that it is possible to identify new splicing factors by their interactions with known splicing factors.Here we report identification of a new splicing factor, Sip1, by its interaction with the essential splicing factor SC35. The predicted Sip1 protein sequence contains an RS domain and a region with sequence similarity to the Drosophila splicing regulator, SWAP. We have expressed and purified recombinant Sip1 protein and raised polyclonal antibodies against the recombinant Sip1 protein. The anti-Sip1 antibodies specifically recognize a protein migrating at a molecular mass of approximately 210 kDa in HeLa nuclear extract. The anti-Sip1 antibodies sufficiently deplete Sip1 protein from the nuclear extract, and the Sip1-depleted extract is inactive in pre-mRNA splicing. Addition of recombinant Sip1 protein can partially restore splicing activity to the Sip1-depleted nuclear extract, indicating an essential role of Sip1 in pre-mRNA splicing. Other RS domain-containing proteins, including SC35, ASF/SF2, and U2AF65, cannot substitute for Sip1 in reconstituting splicing activity of the Sip1-depleted nuclear extract. However, addition of U2AF65 further increases splicing activity of Sip1-reconstituted nuclear extract, suggesting that there may be a functional interaction between Sip1 and U2AF65 in nuclear extract.  相似文献   

15.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

16.
The capsid (CA) and nucleocapsid domains of the human immunodeficiency virus type 1 Gag polyprotein are separated by the p2 spacer peptide, which is essential for virus replication. Previous studies have revealed that p2 has an important role in virus morphogenesis. In this paper, we show that a crucial assembly determinant maps to the highly conserved N terminus of p2, which is predicted to form part of an α-helix that begins in CA. A mutational analysis indicates that the ability of the N terminus of p2 to adopt an α-helical structure is essential for its function during virus assembly. To prevent CA-p2 processing, it was necessary to mutate both the CA-p2 cleavage site and an internal cleavage site within p2. Virions produced by the double mutant lacked a conical core shell and instead contained a thin electron-dense shell about 10 nm underneath the virion membrane. These results suggest that p2 is transiently required for proper assembly, but needs to be removed from the C terminus of CA to weaken CA-CA interactions and allow the rearrangement of the virion core shell during virus maturation.The internal structural proteins of the human immunodeficiency virus type 1 (HIV-1) virion are synthesized in the form of a polyprotein (Pr55gag) which can efficiently form enveloped virus-like particles even when expressed alone (17). Pr55gag is modified by N-terminal myristylation, which is required for its stable association with the inner leaflet of the plasma membrane, where virus assembly occurs (4, 21). During or after the release of an immature particle from the plasma membrane, Pr55gag is cleaved by the viral protease. The major Gag cleavage products are matrix (MA), capsid (CA), nucleocapsid (NC), and p6 (25, 34). MA, which has a crucial role in the incorporation of the viral surface glycoproteins (10, 52), remains associated with the host cell-derived lipid envelope of the virion (16). CA forms the shell of the characteristic cone-shaped core of the mature virion which encloses the viral genomic RNA (16, 27). NC is essential for the encapsidation of the viral genome and is believed to coat the viral RNA within the core of the virion (2, 19, 30). The C-terminal p6 domain of Pr55gag facilitates the release of assembled viral particles from the cell surface (20) and is also needed for the incorporation of the regulatory viral protein Vpr (31, 39).Within the context of Pr55gag, two spacer peptides, p2 and p1, are located between CA and NC and between NC and p6, respectively (24, 25). Cleavage between CA and p2 is much slower than that between p2 and NC or between MA and CA (41). As a consequence, a CA-p2 protein (p25) accumulates in virus-producing cells (34). However, CA-p2 is normally found only in trace amounts in virions. In addition to p2, which comprises 14 amino acids (Ala-363 through Met-376) of the HIV-1HXB2 Gag precursor, a 10-amino-acid p2 fragment which extends from Ser-367 through Met-376 has been isolated from HIV-1 virions, indicating that the viral protease can also cleave within p2 (24, 25).Genetic analyses indicate that the region surrounding the CA-p2 boundary has an important role in particle assembly (21, 28, 50). Within CA, the N-terminal two-thirds forms a domain which appears dispensable for particle assembly but is required for the formation of the cone-shaped core of the mature virion (8, 44, 51). Recent structure determinations have revealed that the N-terminal HIV-1 CA domain is largely α-helical (18, 35). An exposed loop region between two α-helices interacts with the prolyl isomerase cyclophilin A (14), which leads to the incorporation of the cellular enzyme into virions (13, 48). The C-terminal third of CA forms a distinct domain which is essential for Gag oligomerization and particle assembly (8, 12, 44). While genetic and structural studies indicate that the N-terminal boundary of the CA assembly domain coincides with a uniquely conserved sequence, termed the major homology region (8, 15, 18, 32), its C-terminal boundary remains less well defined.The replacement of the scissile dipeptide Leu-Ala at the CA-p2 boundary with Ser-Arg in a mutant designated SVC-C2 led to the formation of grossly distorted capsid structures and caused a significant reduction in particle yield, indicating that the very C terminus of CA and/or p2 is crucial for HIV-1 morphogenesis (21). The possibility that the CA assembly domain extends into p2 is also suggested by the finding that the precise deletion of p2 from Pr55gag markedly reduced particle production (28). Electron microscopy revealed an accumulation of large electron-dense plaques underneath the plasma membrane in the absence of p2 (28), a phenotype which is similar to that observed for the SVC-C2 cleavage site mutant (21). However, the role of p2 in virus assembly remains controversial, because its removal appeared to have no effect on particle release in another study (41).In the present study, we focused on the N-terminal portion of p2, since it is considerably more conserved than the C terminus and because it is predicted to be part of an α-helix which begins in CA. The analysis of a panel of single-amino-acid changes shows that the conserved N terminus of p2 is essential for virus replication and indicates that its predicted α-helical conformation is crucial for virus assembly. In contrast, a deletion which removed 5 out of 10 amino acids between a previously reported cleavage site within p2 and NC delayed but did not abolish virus replication, demonstrating that this relatively variable region of p2 has no essential function in the viral life cycle. We also show that processing of CA-p2 can be essentially prevented by disrupting both the CA-p2 cleavage site and the reported Met-Ser site (25) within p2. Interestingly, the mutant particles often contained a prominent circular structure underneath the viral membrane, indicating that the presence of p2 at the C terminus of CA prevented the rearrangement of the core into a conical tube.  相似文献   

17.
The human immunodeficiency virus type 1 (HIV-1) vpu gene encodes a type I anchored integral membrane phosphoprotein with two independent functions. First, it regulates virus release from a post-endoplasmic reticulum (ER) compartment by an ion channel activity mediated by its transmembrane anchor. Second, it induces the selective down regulation of host cell receptor proteins (CD4 and major histocompatibility complex class I molecules) in a process involving its phosphorylated cytoplasmic tail. In the present work, we show that the Vpu-induced proteolysis of nascent CD4 can be completely blocked by peptide aldehydes that act as competitive inhibitors of proteasome function and also by lactacystin, which blocks proteasome activity by covalently binding to the catalytic β subunits of proteasomes. The sensitivity of Vpu-induced CD4 degradation to proteasome inhibitors paralleled the inhibition of proteasome degradation of a model ubiquitinated substrate. Characterization of CD4-associated oligosaccharides indicated that CD4 rescued from Vpu-induced degradation by proteasome inhibitors is exported from the ER to the Golgi complex. This finding suggests that retranslocation of CD4 from the ER to the cytosol may be coupled to its proteasomal degradation. CD4 degradation mediated by Vpu does not require the ER chaperone calnexin and is dependent on an intact ubiquitin-conjugating system. This was demonstrated by inhibition of CD4 degradation (i) in cells expressing a thermally inactivated form of the ubiquitin-activating enzyme E1 or (ii) following expression of a mutant form of ubiquitin (Lys48 mutated to Arg48) known to compromise ubiquitin targeting by interfering with the formation of polyubiquitin complexes. CD4 degradation was also prevented by altering the four Lys residues in its cytosolic domain to Arg, suggesting a role for ubiquitination of one or more of these residues in the process of degradation. The results clearly demonstrate a role for the cytosolic ubiquitin-proteasome pathway in the process of Vpu-induced CD4 degradation. In contrast to other viral proteins (human cytomegalovirus US2 and US11), however, whose translocation of host ER molecules into the cytosol occurs in the presence of proteasome inhibitors, Vpu-targeted CD4 remains in the ER in a transport-competent form when proteasome activity is blocked.

The human immunodeficiency virus type 1 (HIV-1)-specific accessory protein Vpu performs two distinct functions in the viral life cycle (11, 12, 29, 34, 46, 47, 5052; reviewed in references 31 and 55): enhancement of virus particle release from the cell surface, and the selective induction of proteolysis of newly synthesized membrane proteins. Known targets for Vpu include the primary virus receptor CD4 (63, 64) and major histocompatibility complex (MHC) class I molecules (28). Vpu is an oligomeric class I integral membrane phosphoprotein (35, 48, 49) with a structurally and functionally defined domain architecture: an N-terminal transmembrane anchor and C-terminal cytoplasmic tail (20, 34, 45, 47, 50, 65). Vpu-induced degradation of endoplasmic reticulum (ER) membrane proteins involves the phosphorylated cytoplasmic tail of the protein (50), whereas the virion release function is mediated by a cation-selective ion channel activity associated with the membrane anchor (19, 31, 45, 47).CD4 is a 55-kDa class I integral membrane glycoprotein that serves as the primary coreceptor for HIV entry into cells. CD4 consists of a large lumenal domain, a transmembrane peptide, and a 38-residue cytoplasmic tail. It is expressed on the surface of a subset of T lymphocytes that recognize MHC class II-associated peptides, and it plays a pivotal role in the development and maintenance of the immune system (reviewed in reference 30). Down regulation of CD4 in HIV-1-infected cells is mediated through several independent mechanisms (reviewed in references 5 and 55): intracellular complex formation of CD4 with the HIV envelope protein gp160 (8, 14), endocytosis of cell surface CD4 induced by the HIV-1 nef gene product (1, 2), and ER degradation induced by the HIV-1 vpu gene product (63, 64).Vpu-induced degradation of CD4 is an example of ER-associated protein degradation (ERAD). ERAD is a common outcome when proteins in the secretory pathway are unable to acquire their native structure (4). Although it was thought that ERAD occurs exclusively inside membrane vesicles of the ER or other related secretory compartments, this has gained little direct experimental support. Indeed, there are several recent reports that ERAD may actually represent export of the target protein to the cytosol, where it is degraded by cytosolic proteases. It was found that in yeast, a secreted protein, prepro-α-factor (pαF), is exported from microsomes and degraded in the cytosol in a proteasome-dependent manner (36). This process was dependent on the presence of calnexin, an ER-resident molecular chaperone that interacts with N-linked oligosaccharides containing terminal glucose residues (3). In mammalian cells, two human cytomegalovirus (HCMV) proteins, US2 and US11, were found to cause the retranslocation of MHC class I molecules from the ER to the cytosol, where they are destroyed by proteasomes (61, 62). In the case of US2, class I molecules were found to associate with a protein (Sec61) present in the channel normally used to translocate newly synthesized proteins into the ER (termed the translocon), leading to the suggestion that the ERAD substrates are delivered to the cytosol by retrograde transport through the Sec61-containing pore (61). Fujita et al. (24) reported that, similar to these findings, the proteasome-specific inhibitor lactacystin (LC) partially blocked CD4 degradation in transfected HeLa cells coexpressing CD4, Vpu, and HIV-1 Env glycoproteins. In the present study, we show that Vpu-induced CD4 degradation can be completely blocked by proteasome inhibitors, does not require the ER chaperone calnexin, but requires the function of the cytosolic polyubiquitination machinery which apparently targets potential ubiquitination sites within the CD4 cytoplasmic tail. Our findings point to differences between the mechanism of Vpu-mediated CD4 degradation and ERAD processes induced by the HCMV proteins US2 and US11 (61, 62).  相似文献   

18.
Following infection of mice with lymphocytic choriomeningitis virus (LCMV), virus-neutralizing antibodies appear late, after 30 to 60 days. Such neutralizing antibodies play an important role in protection against reinfection. To analyze whether a neutralizing antibody response which developed earlier could contribute to LCMV clearance during the acute phase of infection, we generated transgenic mice expressing LCMV-neutralizing antibodies. Transgenic mice expressing the immunoglobulin μ heavy chain of the LCMV-neutralizing monoclonal antibody KL25 (H25 transgenic mice) mounted LCMV-neutralizing immunoglobulin M (IgM) serum titers within 8 days after infection. This early inducible LCMV-neutralizing antibody response significantly improved the host’s capacity to clear the infection and did not cause an enhancement of disease after intracerebral (i.c.) LCMV infection. In contrast, mice which had been passively administered LCMV-neutralizing antibodies and transgenic mice exhibiting spontaneous LCMV-neutralizing IgM serum titers (HL25 transgenic mice expressing the immunoglobulin μ heavy and the κ light chain) showed an enhancement of disease after i.c. LCMV infection. Thus, early-inducible LCMV-neutralizing antibodies can contribute to viral clearance in the acute phase of the infection and do not cause antibody-dependent enhancement of disease.Against many cytopathic viruses such as poliovirus, influenza virus, rabies virus, and vesicular stomatitis virus, protective virus-neutralizing antibodies are generated early, within 1 week after infection (3, 31, 36, 44, 49). In contrast, several noncytopathic viruses (e.g., human immunodeficiency virus and hepatitis viruses B and C in humans or lymphocytic choriomeningitis virus [LCMV] in mice) elicit poor and delayed virus-neutralizing antibody responses (1, 7, 20, 24, 27, 35, 45, 48).In the mouse, the natural host of LCMV, the acute LCMV infection is predominantly controlled by cytotoxic T lymphocytes (CTLs) in an obligatory perforin-dependent manner (13, 18, 28, 50). In addition to the CTL response, LCMV-specific antibodies are generated. Early after infection (by day 8), a strong antibody response specific for the internal viral nucleoprotein (NP) is mounted (7, 19, 23, 28). These early LCMV NP-specific antibodies exhibit no virus-neutralizing capacity (7, 10). Results from studies of B-cell-depleted mice and B-cell-deficient mice implied that the early LCMV NP-specific antibodies are not involved in the clearance of LCMV (8, 11, 12, 40). Late after infection (between days 30 and day 60), LCMV-neutralizing antibodies develop (7, 19, 22, 28, 33); these antibodies are directed against the surface glycoprotein (GP) of LCMV (9, 10). LCMV-neutralizing antibodies have an important function in protection against reinfection (4, 6, 38, 41, 47).In some viral infections, subprotective virus-neutralizing antibody titers can enhance disease rather than promote host recovery (i.e., exhibit antibody-dependent enhancement of disease [ADE] [14, 15, 21, 46]). For example, neutralizing antibodies are involved in the resolution of a primary dengue virus infection and in the protection against reinfection. However, if subprotective neutralizing antibody titers are present at the time of reinfection, a severe form of the disease (dengue hemorrhagic fever/dengue shock syndrome [15, 21]), which might be caused by Fc receptor-mediated uptake of virus-antibody complexes leading to an enhanced infection of monocytes (15, 16, 25, 39), can develop. Similarly, an enhancement of disease after intracerebral (i.c.) LCMV infection was observed in mice which had been treated with virus-neutralizing antibodies before the virus challenge (6). ADE in LCMV-infected mice was either due to an enhanced infection of monocytes by Fc receptor-mediated uptake of antibody-virus complexes or due to CTL-mediated immunopathology caused by an imbalanced virus spread and CTL response.To analyze whether LCMV-neutralizing antibodies generated early after infection improve the host’s capacity to clear the virus or enhance immunopathological disease, immunoglobulin (Ig)-transgenic mice expressing LCMV-neutralizing IgM antibodies were generated. After LCMV infection of transgenic mice expressing the Ig heavy chain (H25 transgenic mice), LCMV-neutralizing serum antibodies were mounted within 8 days, which significantly improved the host’s capacity to eliminate LCMV. H25 transgenic mice did not show any signs of ADE after i.c. LCMV infection.Transgenic mice expressing the Ig heavy and light chains (HL25 transgenic mice) exhibited spontaneous LCMV-neutralizing serum antibodies and confirmed the protective role of preexisting LCMV-neutralizing antibodies, even though the neutralizing serum antibodies were of the IgM isotype. Similar to mice which had been treated with LCMV-neutralizing antibodies, HL25 transgenic mice developed an enhanced disease after i.c. LCMV infection, which indicated that ADE was due to an imbalance between virus spread and CTL response. Thus, the early-inducible LCMV-neutralizing antibody response significantly enhanced clearance of the acute infection without any risk of causing ADE.  相似文献   

19.
The Epstein-Barr virus gH-gL complex includes a third glycoprotein, gp42, which is the product of the BZLF2 open reading frame (ORF). gp42 has been implicated as critical to infection of the B lymphocyte by virtue of its interaction with HLA class II on the B-cell surface. A neutralizing antibody that reacts with gp42 inhibits virus-cell fusion and blocks binding of gp42 to HLA class II; antibody to HLA class II can inhibit infection, and B cells that lack HLA class II can only be infected if HLA class II expression is restored. To confirm whether gp42 is an essential component of the virion, we derived a recombinant virus with a selectable marker inserted into the BZLF2 ORF to interrupt expression of the protein. A complex of gH and gL was expressed by the recombinant virus in the absence of gp42. Recombinant virus egressed from the cell normally and could bind to receptor-positive cells. It had, however, lost the ability to infect or transform B lymphocytes. Treatment with polyethylene glycol restored the infectivity of recombinant virus, confirming that gp42 is essential for penetration of the B-cell membrane.Entry of enveloped viruses into mammalian cells requires that the virion envelope fuse with the cell membrane after attachment to the cell surface. Herpesviruses require the functions of multiple protein species to mediate this event, and in keeping with the common origin and diverse habitats of these viruses, some of the proteins involved in penetration appear to be conserved throughout the family and some appear to be restricted to individual members or more closely related members with similar tropism. The two glycoproteins gH and gL fall into the first category of conserved proteins. Glycoprotein gH has been implicated as a major player in virus-cell fusion in many herpesviruses (8, 10, 11, 22, 28, 32, 34), and gL is an essential partner which is required for folding and transport of gH out of the endoplasmic reticulum (6, 19, 21, 27, 28, 35, 38, 45). The gH and gL homologs of Epstein-Barr virus (EBV) are gp85, the product of the BXLF2 open reading frame (ORF) (13, 31), and gp25, the product of the BKRF2 ORF (45), and these homologs appear to behave much as their counterparts in other herpesviruses do (45). However, a third glycoprotein, gp42, associates with the EBV gH-gL complex and falls into the second category of proteins, those with a more restricted distribution.Glycoprotein gp42 is the product of the BZLF2 ORF (26), and although there may be a functionally similar protein in cytomegalovirus (18, 24), it is not predicted to have a homolog in other human herpesviruses. It does, however, have a homolog in ORF51 of equine herpes virus 2 (43). Both EBV and equine herpes virus 2 infect B lymphocytes (1), and several lines of evidence suggest that, at least in the case of EBV, gp42 is critical to the infection of this cell type. A monoclonal antibody (MAb) called F-2-1 that reacts with gp42 has no affect on EBV attachment to its receptor, complement receptor type 2 (CR2) (CD21), but inhibits fusion of the virus with the B-cell membrane and neutralizes infection (29). Glycoprotein gp42 interacts with the β1 domain of the HLA class II protein HLA-DR (39), and MAb F-2-1 interferes with this interaction (25). Like F-2-1, a MAb to HLA-DR or a soluble form of gp42 can block B-cell transformation, and B-cell lines which lack expression of HLA class II are not susceptible to superinfection with EBV unless expression of HLA class II is restored (25). Collectively these observations suggest that gp42, probably by virtue of its interaction with HLA class II, is essential to infection of the B lymphocyte. To answer directly the question of whether gp42 is an indispensable glycoprotein, we derived a virus that could be definitively shown to lack expression of the molecule and examined its ability to infect normal resting B lymphocytes. We report here that virus with expression of gp42 blocked can exit cells normally and can bind to receptor-positive target cells. However, it is unable to penetrate into cells and initiate infection.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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