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.
Complex N-glycans flank the receptor binding sites of the outer domain of HIV-1 gp120, ostensibly forming a protective “fence” against antibodies. Here, we investigated the effects of rebuilding this fence with smaller glycoforms by expressing HIV-1 pseudovirions from a primary isolate in a human cell line lacking N-acetylglucosamine transferase I (GnTI), the enzyme that initiates the conversion of oligomannose N-glycans into complex N-glycans. Thus, complex glycans, including those that surround the receptor binding sites, are replaced by fully trimmed oligomannose stumps. Conversely, the untrimmed oligomannoses of the silent domain of gp120 are likely to remain unchanged. For comparison, we produced a mutant virus lacking a complex N-glycan of the V3 loop (N301Q). Both variants exhibited increased sensitivities to V3 loop-specific monoclonal antibodies (MAbs) and soluble CD4. The N301Q virus was also sensitive to “nonneutralizing” MAbs targeting the primary and secondary receptor binding sites. Endoglycosidase H treatment resulted in the removal of outer domain glycans from the GnTI- but not the parent Env trimers, and this was associated with a rapid and complete loss in infectivity. Nevertheless, the glycan-depleted trimers could still bind to soluble receptor and coreceptor analogs, suggesting a block in post-receptor binding conformational changes necessary for fusion. Collectively, our data show that the antennae of complex N-glycans serve to protect the V3 loop and CD4 binding site, while N-glycan stems regulate native trimer conformation, such that their removal can lead to global changes in neutralization sensitivity and, in extreme cases, an inability to complete the conformational rearrangements necessary for infection.The intriguing results of a recent clinical trial suggest that an effective HIV-1 vaccine may be possible (97). Optimal efficacy may require a component that induces broadly neutralizing antibodies (BNAbs) that can block virus infection by their exclusive ability to recognize the trimeric envelope glycoprotein (Env) spikes on particle surfaces (43, 50, 87, 90). Env is therefore at the center of vaccine design programs aiming to elicit effective humoral immune responses.The amino acid sequence variability of Env presents a significant challenge for researchers seeking to elicit broadly effective NAbs. Early sequence comparisons revealed, however, that the surface gp120 subunit can be divided into discrete variable and conserved domains (Fig. (Fig.1A)1A) (110), the latter providing some hope for broadly effective NAb-based vaccines. Indeed, the constraints on variability in the conserved domains of gp120 responsible for binding the host cell receptor CD4, and coreceptor, generally CCR5, provide potential sites of vulnerability. However, viral defense strategies, such as the conformational masking of conserved epitopes (57), have made the task of eliciting bNAbs extremely difficult.Open in a separate windowFIG. 1.Glycan biosynthesis and distribution on gp120 and gp41. (A) Putative carbohydrate modifications are shown on gp120 and gp41 secondary structures, based on various published works (26, 42, 63, 74, 119, 128). The gp120 outer domain is indicated, as are residues that form the SOS gp120-gp41 disulfide bridge. The outer domain is divided into neutralizing and silent faces. Symbols distinguish complex, oligomannose, and unknown glycans. Generally, the complex glycans of the outer domain line the receptor binding sites of the neutralizing face, while the oligomannose glycans of the outer domain protect the silent domain (105). Asterisks denote sequons that are unlikely to be utilized, including position 139 (42), position 189 (26, 42), position 406 (42, 74), and position 637 (42). Glycans shown in gray indicate when sequon clustering may lead to some remaining unused, e.g., positions 156 and 160 (42, 119), positions 386, 392, and 397 (42), and positions 611 and 616 (42). There is also uncertainty regarding some glycan identities: glycans at positions 188, 355, 397, and 448 are not classified as predominantly complex or oligomannose (26, 42, 63, 128). The number of mannose moieties on oligomannose glycans can vary, as can the number of antennae and sialic acids on complex glycans (77). The glycan at position 301 appears to be predominantly a tetra-antennary complex glycan, as is the glycan at position 88, while most other complex glycans are biantennary (26, 128). (B) Schematic of essential steps of glycan biosynthesis from the Man9GlcNAc2 precursor to a mature multiantennary complex glycan. Mannosidase I progressively removes mannose moieties from the precursor, in a process that can be inhibited by the drug kifunensine. GnTI then transfers a GlcNAc moiety to the D1 arm of the resulting Man5GlcNAc2 intermediate, creating a hybrid glycan. Mannose trimming of the D2 and D3 arms then allows additional GlcNAc moieties to be added by a series of GnT family enzymes to form multiantennary complexes. This process can be inhibited by swainsonine. The antennae are ultimately capped and decorated by galactose and sialic acid. Hybrid and complex glycans are usually fucosylated at the basal GlcNAc, rendering them resistant to endo H digestion. However, NgF is able to remove all types of glycan.Carbohydrates provide a layer of protection against NAb attack (Fig. (Fig.1A).1A). As glycans are considered self, antibody responses against them are thought to be regulated by tolerance mechanisms. Thus, a glycan network forms a nonimmunogenic “cloak,” protecting the underlying protein from antibodies (3, 13, 20, 29, 39, 54, 65, 67, 74, 85, 96, 98, 117, 119, 120). The extent of this protection can be illustrated by considering the ways in which glycans differ from typical amino acid side chains. First, N-linked glycans are much larger, with an average mass more than 20 times that of a typical amino acid R-group. They are also usually more flexible and may therefore affect a greater volume of surrounding space. In the more densely populated parts of gp120, the carbohydrate field may even be stabilized by sugar-sugar hydrogen bonds, providing even greater coverage (18, 75, 125).The process of N-linked glycosylation can result in diverse structures that may be divided into three categories: oligomannose, hybrid, and complex (56). Each category shares a common Man3GlcNAc2 pentasaccharide stem (where Man is mannose and GlcNAc is N-acetylglucosamine), to which up to six mannose residues are attached in oligomannose N-glycans, while complex N-glycans are usually larger and may bear various sizes and numbers of antennae (Fig. (Fig.1B).1B). Glycan synthesis begins in the endoplasmic reticulum, where N-linked oligomannose precursors (Glc3Man9GlcNAc2; Glc is glucose) are transferred cotranslationally to the free amide of the asparagine in a sequon Asn-X-Thr/Ser, where X is not Pro (40). Terminal glucose and mannose moieties are then trimmed to yield Man5GlcNAc2 (Fig. (Fig.1B).1B). Conversion to a hybrid glycan is then initiated by N-acetylglucosamine transferase I (GnTI), which transfers a GlcNAc moiety to the D1 arm of the Man5GlcNAc2 substrate (19) (Fig. (Fig.1B).1B). This hybrid glycoform is then a substrate for modification into complex glycans, in which the D2 and D3 arm mannose residues are replaced by complex antennae (19, 40, 56). Further enzymatic action catalyzes the addition of α-1-6-linked fucose moiety to the lower GlcNAc of complex glycan stems, but usually not to oligomannose glycan stems (Fig. (Fig.1B)1B) (21, 113).Most glycoproteins exhibit only fully mature complex glycans. However, the steric limitations imposed by the high density of glycans on some parts of gp120 lead to incomplete trimming, leaving “immature” oligomannose glycans (22, 26, 128). Spatial competition between neighboring sequons can sometimes lead to one or the other remaining unutilized, further distancing the final Env product from what might be expected based on its primary sequence (42, 48, 74, 119). An attempt to assign JR-FL gp120 and gp41 sequon use and types, based on various studies, is shown in Fig. Fig.1A1A (6, 26, 34, 35, 42, 63, 71, 74, 119, 128). At some positions, the glycan type is conserved. For example, the glycan at residue N301 has consistently been found to be complex (26, 63, 128). At other positions, considerable heterogeneity exists in the glycan populations, in some cases to the point where it is difficult to unequivocally assign them as predominantly complex or oligomannose. The reasons for these uncertainties might include incomplete trimming (42), interstrain sequence variability, the form of Env (e.g., gp120 or gp140), and the producer cell. The glycans of native Env trimers and monomeric gp120 may differ due to the constraints imposed by oligomerization (32, 41, 77). Thus, although all the potential sequons of HXB2 gp120 were found to be occupied in one study (63), some are unutilized or variably utilized on functional trimers, presumably due to steric limitations (42, 48, 75, 96, 119).The distribution of complex and oligomannose glycans on gp120 largely conforms with an antigenic map derived from structural models (59, 60, 102, 120), in which the outer domain is divided into a neutralizing face and an immunologically silent face. Oligomannose glycans cluster tightly on the silent face of gp120 (18, 128), while complex glycans flank the gp120 receptor binding sites of the neutralizing face, ostensibly forming a protective “fence” against NAbs (105). The relatively sparse clustering of complex glycans that form this fence may reflect a trade-off between protecting the underlying functional domains from NAbs by virtue of large antennae while at the same time permitting sufficient flexibility for the refolding events associated with receptor binding and fusion (29, 39, 67, 75, 98, 117). Conversely, the dense clustering of oligomannose glycans on the silent domain may be important for ensuring immune protection and/or in creating binding sites for lectins such as DC-SIGN (9, 44).The few available broadly neutralizing monoclonal antibodies (MAbs) define sites of vulnerability on Env trimers (reviewed in reference 52). They appear to fall into two general categories: those that access conserved sites by overcoming Env''s various evasion strategies and, intriguingly, those that exploit these very defensive mechanisms. Regarding the first category, MAb b12 recognizes an epitope that overlaps the CD4 binding site of gp120 (14), and MAbs 2F5 and 4E10 (84, 129) recognize adjacent epitopes of the membrane-proximal external region (MPER) at the C-terminal ectodomain of gp41. The variable neutralizing potencies of these MAbs against primary isolates that contain their core epitopes illustrate how conformational masking can dramatically regulate their exposure (11, 118). Conformational masking also limits the activities of MAbs directed to the V3 loop and MAbs whose epitopes overlap the coreceptor binding site (11, 62, 121).A second category of MAbs includes MAb 2G12, which recognizes a tight cluster of glycans in the silent domain of gp120 (16, 101, 103, 112). This epitope has recently sparked considerable interest in exploiting glycan clusters as possible carbohydrate-based vaccines (2, 15, 31, 70, 102, 116). Two recently described MAbs, PG9 and PG16 (L. M. Walker and D. R. Burton, unpublished data), also target epitopes regulated by the presence of glycans that involve conserved elements of the second and third variable loops and depend largely on the quaternary trimer structure and its in situ presentation on membranes. Their impressive breadth and potency may come from the fact that they target the very mechanisms (variable loops and glycans) that are generally thought to protect the virus from neutralization. Like 2G12, these epitopes are likely to be constitutively exposed and thus may not be subject to conformational masking (11, 118).The above findings reveal the importance of N-glycans both as a means of protection against neutralization as well as in directly contributing to unique neutralizing epitopes. Clearly, further studies on the nature and function of glycans in native Env trimers are warranted. Possible approaches may be divided into four categories, namely, (i) targeted mutation, (ii) enzymatic removal, (iii) expression in the presence of glycosylation inhibitors, and (iv) expression in mutant cell lines with engineered blocks in the glycosylation pathway. Much of the available information on the functional roles of glycans in HIV-1 and simian immunodeficiency virus (SIV) infection has come from the study of mutants that eliminate glycans either singly or in combination (20, 54, 66, 71, 74, 91, 95, 96). Most mutants of this type remain at least partially functional (74, 95, 96). In some cases these mutants have little effect on neutralization sensitivity, while in others they can lead to increased sensitivity to MAbs specific for the V3 loop and CD4 binding site (CD4bs) (54, 71, 72, 74, 106). In exceptional cases, increased sensitivity to MAbs targeting the coreceptor binding site and/or the gp41 MPER has been observed (54, 66, 72, 74).Of the remaining approaches for studying the roles of glycans, enzymatic removal is constrained by the extreme resistance of native Env trimers to many common glycosidases, contrasting with the relative sensitivity of soluble gp120 (67, 76, 101). Alternatively, drugs can be used to inhibit various stages of mammalian glycan biosynthesis. Notable examples are imino sugars, such as N-butyldeoxynojirimycin (NB-DNJ), that inhibit the early trimming of the glucose moieties from Glc3Man9GlcNAc2 precursors in the endoplasmic reticulum (28, 38, 51). Viruses produced in the presence of these drugs may fail to undergo proper gp160 processing or fusion (37, 51). Other classes of inhibitor include kifunensine and swainsonine, which, respectively, inhibit the trimming of the Man9GlcNAc2 precursor into Man5GlcNAc2 or inhibit the removal of remaining D2 and D3 arm mannoses from the hybrid glycans, thus preventing the construction of complex glycan antennae (Fig. (Fig.1B)1B) (17, 33, 76, 104, 119). Unlike NB-DNJ, viruses produced in the presence of these drugs remain infectious (36, 76, 79, 100).Yet another approach is to express virus in insect cells that can only modify proteins with paucimannose N-glycans (58). However, the inefficient gp120/gp41 processing by furin-like proteases in these cells prevents their utility in functional studies (123). Another option is provided by ricin-selected GnTI-deficient cell lines that cannot transfer GlcNAc onto the mannosidase-trimmed Man5GlcNAc2 substrate, preventing the formation of hybrid and complex carbohydrates (Fig. (Fig.1B)1B) (17, 32, 36, 94). This arrests glycan processing at a well-defined point, leading to the substitution of complex glycans with Man5GlcNAc2 rather than with the larger Man9GlcNAc2 precursors typically obtained with kifunensine treatment (17, 32, 33, 104). With this in mind, here we produced HIV-1 pseudoviruses in GnTI-deficient cells to investigate the role of complex glycan antennae in viral resistance neutralization. By replacing complex glycans with smaller Man5GlcNAc2 we can determine the effect of “lowering the glycan fence” that surrounds the receptor binding sites, compared to the above-mentioned studies of individual glycan deletion mutants, whose effects are analogous to removing a fence post. Furthermore, since oligomannose glycans are sensitive to certain enzymes, such as endoglycosidase H (endo H), we investigated the effect of dismantling the glycan fence on Env function and stability. Our results suggest that the antennae of complex glycans protect against certain specificities but that glycan stems regulate trimer conformation with often more dramatic consequences for neutralization sensitivity and in extreme cases, infectious function.  相似文献   

2.
Alphavirus particles are covered by 80 glycoprotein spikes that are essential for viral entry. Spikes consist of the E2 receptor binding protein and the E1 fusion protein. Spike assembly occurs in the endoplasmic reticulum, where E1 associates with pE2, a precursor containing E3 and E2 proteins. E3 is a small, cysteine-rich, extracellular glycoprotein that mediates proper folding of pE2 and its subsequent association with E1. In addition, cleavage of E3 from the assembled spike is required to make the virus particles efficiently fusion competent. We have found that the E3 protein in Sindbis virus contains one disulfide bond between residues Cys19 and Cys25. Replacing either of these two critical cysteines resulted in mutants with attenuated titers. Replacing both cysteines with either alanine or serine resulted in double mutants that were lethal. Insertion of additional cysteines based on E3 proteins from other alphaviruses resulted in either sequential or nested disulfide bond patterns. E3 sequences that formed sequential disulfides yielded virus with near-wild-type titers, while those that contained nested disulfide bonds had attenuated activity. Our data indicate that the role of the cysteine residues in E3 is not primarily structural. We hypothesize that E3 has an enzymatic or functional role in virus assembly, and these possibilities are further discussed.Alphaviruses are members of the Togaviradae family and are single-stranded, positive-sense RNA, enveloped viruses (17). The lipid membranes of the viruses have 80 glycoprotein spikes which are required for viral entry. Each spike is comprised of three copies of a heterodimer which consists of the E2 and E1 proteins (22, 54). E2 and E1 are glycoproteins with a single transmembrane helix that traverses the host-derived lipid bilayer. E2 interacts with the nucleocapsid core at the C terminus (12, 16, 27, 43) and contains the receptor binding site at the N terminus (5, 21, 45). E1 is the viral fusion protein responsible for mediating fusion between the virus membrane and the host cell membrane during an infection (13, 39, 47). Specific interactions in both the ectodomain and transmembrane regions are critical for heterodimer formation (30, 35, 46, 54). The assembly of each heterodimer, its subsequent assembly into a spike, and the interaction of the cytoplasmic tail of the spike with the nucleocapsid core are all essential for the efficient production of infectious particles.Glycoprotein spike assembly requires four structural proteins, E3, E2, 6K, and E1, which are expressed as a single polyprotein. E3 is a small, 64-amino-acid protein (Sindbis virus [SINV] numbering) and contains a signal sequence that translocates the protein into the endoplasmic reticulum (ER) (3, 4, 15). Early in translation, glycosylation of N14 (SINV numbering) occurs and this promotes E3''s release from the ER membrane into the lumen. As a result, the signal sequence is not cleaved from the E3 protein (14). Cellular enzymes cleave the polyprotein to yield pE2 (an uncleaved protein consisting of E3 and E2), 6K, and E1 (23, 55) proteins. In the ER, E1 is found in several conformations, only one of which will form a functional heterodimer with pE2, allowing its transport to the Golgi apparatus (1, 2, 6, 7, 36). After pE2-E1 heterodimerization, self-association between three heterodimers occurs and each individual spike is formed (25, 26, 36). As observed with Semliki Forest virus, disulfide bonds reshuffle within pE2 during protein folding (34), possibly forming intermolecular disulfide bonds between E3 and E2 residues. However, no intermolecular disulfide bonds between pE2 and E1 have been identified (34). Once the viral spikes have been assembled, they are transported to the plasma membrane (11) and are thus exposed to subcellular changes of pH, from pH 7.2 in the ER to pH 5.7 in the vesicles constitutively transporting the spikes to the plasma membrane. In the trans-Golgi network, the E3 protein is cleaved from pE2 by the cellular protein furin (18, 44, 55). E3 remains noncovalently attached to the released virus particle, while in other species E3 is found in the medium of virus-infected cells (32, 49).E3 is required for efficient particle assembly, both in mediating spike folding and in spike activation for viral entry. When an ER signal sequence was substituted for the E3 protein, heterodimerization of pE2 and E1 was abolished (26). Furthermore, when E2 and E1 were expressed individually, low levels of E2 were transported to the cell surface while E1 remained in the ER, suggesting that heterodimerization with pE2 is necessary for E1 to be transported to the cell surface (24, 26, 46). These results are consistent with E3 playing a critical role in mediating the folding of pE2 and the association of pE2 and E1 proteins during spike assembly (7, 38). In viruses where the furin cleavage site was mutated, the virus particles were correctly assembled but severely reduced in infectivity, presumably because the fusion protein was unable to dissociate from pE2 and initiate fusion (44, 55).A comparison of an amino acid sequence alignment of E3 proteins from different alphaviruses (Fig. (Fig.1)1) shows that the E3 protein is a small protein with four conserved cysteine (Cys) residues. A subset of E3 proteins contains an additional two Cys residues in a narrow cysteine/proline-rich region, PPCXPCC (Fig. (Fig.1).1). We have purified recombinant E3 protein from SINV and have determined that a disulfide bond is present and, furthermore, that these Cys residues are important in virus assembly. Within the alphavirus E3 proteins, we have identified a region that is important for mediating spike transport to the plasma membrane and thus is critical for spike assembly.Open in a separate windowFIG. 1.E3 amino acid sequence alignment from a representative group of alphaviruses. The cysteines marked with asterisks are conserved in all alphavirus species. The ⋄ indicates the conserved but nonessential glycosylation site. The PPCXPCC motif present in ∼50% of alphaviruses is underlined. SFV, Semliki Forest virus; RRV, Ross River virus; BFV, Barmah Forest virus; EEE, eastern equine encephalitis virus; ONN, O''nyong nyong virus; IGB, Igbo Ora virus; OCK, Ockelbo virus; WEE, western equine encephalitis virus; AUR, Aura virus; VEE, Venezuelan equine encephalitis virus.  相似文献   

3.
4.
Flaviviruses are a group of single-stranded, positive-sense RNA viruses causing ∼100 million infections per year. We have recently shown that flaviviruses produce a unique, small, noncoding RNA (∼0.5 kb) derived from the 3′ untranslated region (UTR) of the genomic RNA (gRNA), which is required for flavivirus-induced cytopathicity and pathogenicity (G. P. Pijlman et al., Cell Host Microbe, 4: 579-591, 2008). This RNA (subgenomic flavivirus RNA [sfRNA]) is a product of incomplete degradation of gRNA presumably by the cellular 5′-3′ exoribonuclease XRN1, which stalls on the rigid secondary structure stem-loop II (SL-II) located at the beginning of the 3′ UTR. Mutations or deletions of various secondary structures in the 3′ UTR resulted in the loss of full-length sfRNA (sfRNA1) and production of smaller and less abundant sfRNAs (sfRNA2 and sfRNA3). Here, we investigated in detail the importance of West Nile virus Kunjin (WNVKUN) 3′ UTR secondary structures as well as tertiary interactions for sfRNA formation. We show that secondary structures SL-IV and dumbbell 1 (DB1) downstream of SL-II are able to prevent further degradation of gRNA when the SL-II structure is deleted, leading to production of sfRNA2 and sfRNA3, respectively. We also show that a number of pseudoknot (PK) interactions, in particular PK1 stabilizing SL-II and PK3 stabilizing DB1, are required for protection of gRNA from nuclease degradation and production of sfRNA. Our results show that PK interactions play a vital role in the production of nuclease-resistant sfRNA, which is essential for viral cytopathicity in cells and pathogenicity in mice.Arthropod-borne flaviviruses such as West Nile virus (WNV), dengue virus (DENV), and Japanese encephalitis virus (JEV) cause major outbreaks of potentially fatal disease and affect over 50 million people every year. The highly pathogenic North American strain of WNV (WNVNY99) has already claimed more than 1,000 lives with over 27,000 cases reported since its emergence in New York in 1999 and has raised global public health concerns (9). In contrast, the closely related Australian strain of WNV, WNVKUN, is highly attenuated and does not cause overt disease in humans and animals (11). WNVKUN has been used extensively as a model virus to study flavivirus replication and flavivirus-host interactions (13, 14, 16-19, 26, 38, 39).The ∼11-kb positive-stranded, capped WNV genomic RNA (gRNA) lacks a poly(A) tail and consists of 5′ and 3′ untranslated regions (UTRs) flanking one open reading frame, which encodes the viral proteins required for the viral life cycle (6, 15, 38, 39). Flavivirus UTRs are involved in translation and initiation of RNA replication and likely determine genome packaging (13, 14, 16, 21, 30, 39-41). Both the 5′ UTR (∼100 nucleotides [nt] in size) and the 3′ UTR (from ∼400 to 700 nucleotides) can form secondary and tertiary structures which are highly conserved among mosquito-borne flaviviruses (1, 8, 10, 14, 29, 32, 34). More specifically, the WNVKUN 3′ UTR consists of several conserved regions and secondary structures (Fig. (Fig.1A)1A) which were previously predicted or shown to exist in various flaviviruses by computational and chemical analyses, respectively (4, 10, 25, 26, 29-32). The 5′ end of the 3′ UTR starts with an AU-rich region which can form stem-loop structure I (SL-I) followed by SL-II, which we previously showed to be vitally important for subgenomic flavivirus RNA (sfRNA) production (26; see also below). SL-II is followed by a short, repeated conserved hairpin (RCS3) and SL-III (26). Further downstream of SL-III are the SL-IV and CS3 structures, which are remarkably similar to the preceding SL-II-RCS3 structure (26, 29). Further downstream of the SL-IV-CS3 structure are dumbbells 1 and 2 (DB1 and DB2, respectively) followed by a short SL and the 3′ SL (25, 26).Open in a separate windowFIG. 1.(A) Model of the WNVKUN 3′ UTR RNA structure. Highlighted in bold are the secondary structures investigated here. Dashed lines indicate putative PKs. The two sites of the putative PK interactions are shown in open boxes. sfRNA1, -2, -3, and -4 start sites are indicated by arrows. (R)CS, (repeated) conserved sequence; DB, dumbbell structure; PK, pseudoknot; SL, stem-loop. (B) Structural model of PK1 in SL-II with disruptive mutations. Nucleotide numbering is from the end of the 3′ UTR. The sfRNA1 start is indicated by an arrow. Nucleotides forming PK1 are on a gray background, and mutated nucleotides are white on a black background. (C) Sequences mutated in the different constructs. Nucleotides in the wt PK sequences used for mutations are bold and underlined. Introduced mutations are shown under the corresponding nucleotides in the wt sequence.The described structures have been investigated in some detail for their requirement in RNA replication and translation. Generally, a progressive negative effect on viral growth was shown with progressive deletions into the 3′-proximal region of the JEV 3′ UTR (41). However, only a relatively short region of the JEV 3′ UTR, consisting of the 3′-terminal 193 nt, was shown to be absolutely essential for gRNA replication (41). The minimal region for DENV replication was reported to be even shorter (23). Extensive analysis has shown that the most 3′-terminal, essential regions of the 3′ UTR include the cyclization sequence and 3′ SL, which are required for efficient RNA replication (2, 14, 16, 23, 35). As we showed, deletion of SL-II or SL-I did not overtly affect WNVKUN replication (26). However, deletion of CS2, RCS2, CS3, or RCS3 in WNV replicon RNA significantly reduced RNA replication but not translation (20), indicating that these elements facilitate but are not essential for RNA replication. In addition, it was shown that deletion of DB1 or DB2 resulted in a viable mutant virus that was reduced in growth efficiency, while deletion of both DB structures resulted in a nonviable mutant (23).In addition to the above-mentioned secondary stem-loop structures, computational and chemical analysis of the flavivirus 3′ UTR suggested the presence of 5 pseudoknot (PK) interactions (Fig. (Fig.1A)1A) (25, 26, 32). A PK is a structure formed upon base pairing of a single-stranded region of RNA in the loop of a hairpin to a stretch of complementary nucleotides elsewhere in the RNA chain (Fig. (Fig.1B).1B). These structures are referred to as hairpin type (H-type) PKs (3), and they usually stabilize secondary RNA structures. Typically, the final tertiary structure does not significantly alter the preformed secondary structure (5). In general, PK interactions have been shown to be important in biological processes such as initiation and/or elongation of translation, initiation of gRNA replication, and ribosomal frameshifting for a number of different viruses, including flaviviruses (reviewed in references 3 and 22). The first PK in the WNV 3′ UTR was predicted to form in SL-II, followed by a similar PK in SL-IV (26) (PK1 and PK2 in Fig. Fig.1A).1A). For the DENV, yellow fever virus (YFV), and JEV subgroup of flaviviruses, two PKs further downstream were predicted to form between DB1 and DB2 and corresponding single-stranded RNA regions located further downstream (25) (PK3 and PK4 in Fig. Fig.1A).1A). The formation of these structures is supported by covariations in the WNV RNAs. In addition, a PK was proposed to form between a short SL and the 3′ SL at the 3′ terminus of the viral genome (32) (PK5 in Fig. Fig.1A1A).Importantly, in addition to its role in viral replication and translation, we have shown that the WNVKUN 3′ UTR is important for the production of a small noncoding RNA fragment designated sfRNA (26). This short RNA fragment of ∼0.5 kb is derived from the 3′ UTR of the gRNA and exclusively produced by the members of the Flavivirus genus of the Flaviviridae family, where it is required for efficient viral replication, cytopathicity, and pathogenicity (26). Our studies suggested that sfRNA is a product of incomplete degradation of the gRNA presumably by the cellular 5′-3′ exoribonuclease XRN1, resulting from XRN1 stalling on the rigid secondary/tertiary structures located at the beginning of the 3′ UTR (26). XRN1 is an exoribonuclease which usually degrades mRNA from the 5′ to the 3′ end as part of cellular mRNA decay and turnover (33), and it was shown previously that XRN1 can be stalled by SL structures (28). Mutations or deletions of WNV 3′ UTR secondary structures resulted in the loss of full-length sfRNA (sfRNA1) and production of smaller and less abundant sfRNAs (sfRNA2 and sfRNA3) (26). In particular, SL-II (Fig. (Fig.1A)1A) was shown to be important for sfRNA1 production; deletion of this structure either alone or in conjunction with other structures located downstream of SL-II abolished sfRNA1 production, leading to the production of the smaller RNA fragments sfRNA2 and sfRNA3.Here, we extended our investigation and studied the importance of several predicted 3′ UTR secondary structures and PK interactions for the production of sfRNA. To further understand the generation mechanism of sfRNA and its requirements, we deleted or mutated a number of RNA structures in the WNVKUN 3′ UTR and investigated the size and amount of sfRNA generated from these mutant RNAs. The results show that not only SLs but also PK interactions play a vital role in stabilizing the 3′ UTR RNA and preventing complete degradation of viral gRNA to produce nuclease-resistant sfRNA, which is required for efficient virus replication and cytopathicity in cells and virulence in mice.  相似文献   

5.
The biofilm matrix contributes to the chemistry, structure, and function of biofilms. Biofilm-derived membrane vesicles (MVs) and DNA, both matrix components, demonstrated concentration-, pH-, and cation-dependent interactions. Furthermore, MV-DNA association influenced MV surface properties. This bears consequences for the reactivity and availability for interaction of matrix polymers and other constituents.The biofilm matrix contributes to the chemistry, structure, and function of biofilms and is crucial for the development of fundamental biofilm properties (46, 47). Early studies defined polysaccharides as the matrix component, but proteins, lipids, and nucleic acids are all now acknowledged as important contributors (7, 15). Indeed, DNA has emerged as a vital participant, fulfilling structural and functional roles (1, 5, 6, 19, 31, 34, 36, 41, 43, 44). The phosphodiester bond of DNA renders this polyanionic at a physiological pH, undoubtedly contributing to interactions with cations, humic substances, fine-dispersed minerals, and matrix entities (25, 41, 49).In addition to particulates such as flagella and pili, membrane vesicles (MVs) are also found within the matrices of gram-negative and mixed biofilms (3, 16, 40). MVs are multifunctional bilayered structures that bleb from the outer membranes of gram-negative bacteria (reviewed in references 4, 24, 27, 28, and 30) and are chemically heterogeneous, combining the known chemistries of the biofilm matrix. Examination of biofilm samples by transmission electron microscopy (TEM) has suggested that matrix material interacts with MVs (Fig. (Fig.1).1). Since MVs produced in planktonic culture have associated DNA (11, 12, 13, 20, 21, 30, 39, 48), could biofilm-derived MVs incorporate DNA (1, 39, 40, 44)?Open in a separate windowFIG. 1.Possible interactions between matrix polymers and particulate structures. Shown is an electron micrograph of a thin section through a P. aeruginosa PAO1 biofilm. During processing, some dehydration occurred, resulting in collapse of matrix material into fibrillate arrangements (black filled arrows). There is a suggestion of interactions occurring with particulate structures such as MVs (hollow white arrow) and flagella (filled white arrows) (identified by the appearance and cross-dimension of these highly ordered structures when viewed at high magnification), which was consistently observed with other embedded samples and also with whole-mount preparations of gently disrupted biofilms (data not shown). The scale bar represents 200 nm.  相似文献   

6.
7.
The flavivirus genome comprises a single strand of positive-sense RNA, which is translated into a polyprotein and cleaved by a combination of viral and host proteases to yield functional proteins. One of these, nonstructural protein 3 (NS3), is an enzyme with both serine protease and NTPase/helicase activities. NS3 plays a central role in the flavivirus life cycle: the NS3 N-terminal serine protease together with its essential cofactor NS2B is involved in the processing of the polyprotein, whereas the NS3 C-terminal NTPase/helicase is responsible for ATP-dependent RNA strand separation during replication. An unresolved question remains regarding why NS3 appears to encode two apparently disconnected functionalities within one protein. Here we report the 2.75-Å-resolution crystal structure of full-length Murray Valley encephalitis virus NS3 fused with the protease activation peptide of NS2B. The biochemical characterization of this construct suggests that the protease has little influence on the helicase activity and vice versa. This finding is in agreement with the structural data, revealing a single protein with two essentially segregated globular domains. Comparison of the structure with that of dengue virus type 4 NS2B-NS3 reveals a relative orientation of the two domains that is radically different between the two structures. Our analysis suggests that the relative domain-domain orientation in NS3 is highly variable and dictated by a flexible interdomain linker. The possible implications of this conformational flexibility for the function of NS3 are discussed.Flaviviruses such as dengue virus (DENV), yellow fever virus (YFV), West Nile virus (WNV), and Japanese encephalitis virus (JEV) belong to the family Flaviviridae and are the causative agents of a range of serious human diseases including hemorrhagic fever, meningitis, and encephalitis (37). They remain a global health priority, as many viruses are endemic in large parts of the Americas, Africa, Australia, and Asia, and vaccines remain unavailable for most members (31, 46, 57).Flaviviruses have a positive-sense single-stranded RNA (ssRNA) genome (approximately 11 kb) that encodes one large open reading frame containing a 5′ type 1 cap and conserved RNA structures at both the 5′ and 3′ untranslated regions that are important for viral genome translation and replication. The genomic RNA is translated into a single polyprotein precursor (11) consisting of three structural (C [capsid], prM [membrane], and E [envelope]) and seven nonstructural (NS1, NS2a, NS2b, NS3, NS4a, NS4b, and NS5) proteins arranged in the order C-prM-E-NS1-NS2a-NS2b-NS3-NS4a-NS4b-NS5 (reviewed in reference 33) (Fig. (Fig.1).1). Only the structural proteins become part of the mature, infectious virion, whereas the nonstructural proteins are involved in polyprotein processing, viral RNA synthesis, and virus morphogenesis (33, 43). The precursor protein is directed by signal sequences into the host endoplasmic reticulum (ER), where NS1 and the exogenous domains of prM and E face the lumen, while C, NS3, and NS5 are cytoplasmic. NS2A, NS2B, NS4A, and NS4B are largely hydrophobic transmembrane proteins with small hydrophilic segments (Fig. (Fig.1).1). The post- and cotranslational cleavage of the polyprotein is performed by NS3 in the cytoplasm and by host proteases in the ER lumen to yield the mature proteins (Fig. (Fig.1)1) (33, 43). Of the nonstructural proteins, NS3 and NS5 are the best characterized, and both are essential for viral replication (23, 27, 41). Both proteins are multifunctional. The N-terminal one-third of NS3 contains the viral protease (NS3pro), which requires a portion of NS2B for its activity, while the remaining portion codes for the RNA helicase/NTPase/RTPase domain (NS3hel) (21, 22, 32, 55). NS5, however, contains both an N-terminal methyltransferase and a C-terminal RNA-dependent RNA polymerase (16, 51). The functions of NS1, NS2A, NS4A, and NS4B are not well understood, but they appear to play important roles in replication and virus assembly/maturation and have been found to bind to NS3 and NS5, possibly modulating their activity (33, 36).Open in a separate windowFIG. 1.Schematic diagram of flavivirus polyprotein organization and processing. (Top) Linear organization of the structural and nonstructural proteins within the polyprotein. (Middle) Putative membrane topology of the polyprotein predicted from biochemical and cellular analyses, which is then processed by cellular and viral proteases (indicated by arrows). (Bottom) Different complexes that are thought to arise in different cellular compartments during and following polyprotein processing.Because of its enzymatic activities and its critical role in viral replication and polyprotein processing, NS3 constitutes a promising drug target for antiviral therapy (31). NS3pro (residues 1 to 169) is a trypsin-like serine protease with the characteristic catalytic triad (Asp-His-Ser) and a highly specific substrate recognition sequence, conserved in all flaviviruses, consisting of two basic residues in P2 and P1 followed by a small unbranched amino acid in P1′ (11). NS3pro has an aberrant fold compared to the canonical trypsin structure, and its folding and protease activity are dependent on a noncovalent association with a central 47-amino-acid hydrophilic domain of NS2B (19, 21). The remainder of NS2B contains three transmembrane helices involved in membrane associations. NS3 mediates cleavages at the C-terminal side of the highly conserved dibasic residue located at the coding junctions NS2A/NS2B, NS2B/NS3, NS3/NS4A, and NS4B/NS5 and also between the C terminus of C and NS4A (11, 33) (Fig. (Fig.11).The C-terminal portion of NS3 (NS3hel, residues 170 to 619) performs several catalytically related activities, namely, RNA strand separation and (poly)nucleotide hydrolysis (5, 22, 32, 55) at a common, RecA-like NTPase catalytic center that couples the energy released from the hydrolysis of the triphosphate moieties of nucleotides to RNA unwinding. Although the precise role of NS3 in replication has not been established, its helicase activity is thought to separate nascent RNA strands from the template strands and to assist replication initiation by unwinding RNA secondary structure in the 3′ untranslated region (11, 13, 15, 33). NS3 is a member of the DEAH/D box family within helicase superfamily 2 (SF2) and is characterized by seven conserved sequence motifs involved in nucleic acid binding and hydrolysis (45). In addition, its RNA triphosphatase activity is thought to be involved in the capping of the viral RNA. In the process of replication, NS3 interacts, most likely via its C-terminal domain, with NS5 (13, 15, 24, 26, 58, 62). The NS3 5′ triphosphatase and NS5 methyltransferase activities probably cooperate in cap formation by removing the terminal γ-phosphate and performing sequential N7 and 2′ O methylations, respectively (16, 28, 46, 56). The guanylyltransferase activity required for cap formation remains elusive at present, although recent evidence suggests that it may be present in NS5 (8, 17). In addition, the interaction between NS3 and NS5 can stimulate NS3 helicase/NTPase activity (15, 62).The atomic structures of NS3pro in the presence and absence of ligands and/or the NS2B activating domain (2, 19, 47) and NS3hel (35, 38, 39, 49, 58-60) are known, and recently, the structure of full-length DENV4 (one of four dengue virus serotypes) NS3 fused to an 18-residue NS2B cofactor (NS2B18NS3) was reported (34). This structure revealed an elongated conformation, with the protease domain interfacing with the NTP binding pocket and being separated from NS3hel by a relatively flexible linker, which suggested that the protease domain may have a positive effect on the activity of the NTPase/helicase domain. However, other reports suggested that NS3pro has no or a very limited effect on the activity of NS3hel (32, 62). In addition, since current evidence suggests that NS2B is not part of the replication complex (Fig. (Fig.1)1) (36), and it is known that in the absence of the NS2B cofactor, NS3pro is unfolded and inactive, it becomes hard to envisage what effect the NS3 protease domain may have on the helicase domain in a biologically relevant context. Equally, it is still not clear what role the helicase domain plays during polyprotein processing by NS3pro and, in general, why these two apparently distinct and unrelated catalytic activities are harbored within a single polypeptide.In order to gain further insights into these questions, we report the biochemical analysis and crystallographic structure at a 2.75-Å resolution of full-length NS3 from Murray Valley encephalitis virus (MVEV), a member of the JEV group of flaviviruses, fused to the entire protease activation peptide of the NS2B cofactor (NS2B45NS3). The structure reveals the protease and helicase domains to be structurally independent and differs dramatically from the structure observed for DENV4 NS2B18NS3. We discuss the implications of this unexpectedly different configuration of the NS3 protein and argue that the structural flexibility observed is likely to be crucial for its multifunctional nature.  相似文献   

8.
9.
The assembly of retroviruses is driven by oligomerization of the Gag polyprotein. We have used cryo-electron tomography together with subtomogram averaging to describe the three-dimensional structure of in vitro-assembled Gag particles from human immunodeficiency virus, Mason-Pfizer monkey virus, and Rous sarcoma virus. These represent three different retroviral genera: the lentiviruses, betaretroviruses and alpharetroviruses. Comparison of the three structures reveals the features of the supramolecular organization of Gag that are conserved between genera and therefore reflect general principles of Gag-Gag interactions and the features that are specific to certain genera. All three Gag proteins assemble to form approximately spherical hexameric lattices with irregular defects. In all three genera, the N-terminal domain of CA is arranged in hexameric rings around large holes. Where the rings meet, 2-fold densities, assigned to the C-terminal domain of CA, extend between adjacent rings, and link together at the 6-fold symmetry axis with a density, which extends toward the center of the particle into the nucleic acid layer. Although this general arrangement is conserved, differences can be seen throughout the CA and spacer peptide regions. These differences can be related to sequence differences among the genera. We conclude that the arrangement of the structural domains of CA is well conserved across genera, whereas the relationship between CA, the spacer peptide region, and the nucleic acid is more specific to each genus.Retrovirus assembly is driven by the oligomerization of Gag, a multidomain protein, including an N-terminal membrane binding domain (MA), a two-domain structural component (CA), and an RNA binding domain (NC). The Gag proteins of all orthoretroviruses, including the alpha-, beta-, and lentiretroviruses discussed here, share this conserved modular architecture (Fig. (Fig.1).1). Despite very weak sequence conservation, the tertiary structures of MA, CA, and NC are conserved among retroviruses. Outside these conserved domains the Gag proteins of different retroviruses exhibit substantial variability. Other domains may be present or absent, and the length and sequence of linker peptides may also vary (12) (Fig. (Fig.11).Open in a separate windowFIG. 1.Modular architecture of the full-length Gag proteins of HIV, M-PMV, and RSV. White rectangles illustrate Gag polyprotein cleavage products. The extent of the constructs used in the electron microscopic analysis is specified under each protein as a black rectangle. Gray triangles specify cleavage sites. Residue numbers are counted from the beginning of Gag.Oligomerization of Gag in an infected cell leads to the formation of roughly spherical immature virus particles, where Gag is arranged in a radial fashion with the N-terminal MA domain associated with a surrounding lipid bilayer, and the more C-terminal NC pointing toward the center of the particle (15, 44, 46). Subsequent multiple cleavages of Gag by the viral protease lead to a rearrangement of the virus. NC and the RNA condense in the center of the particle, CA assembles into a capsid or shell around the nucleoprotein, and MA remains associated with the viral membrane. This proteolytic maturation is required to generate an infectious virion (2). In contrast to the mature CA lattice, which has been extensively studied (11, 16, 36), the Gag lattice in immature particles is incompletely understood.Gag itself contains all of the necessary determinants for particle assembly. For example, the expression of Gag alone in an insect cell expression system is sufficient to generate viruslike particles (3, 17, 22, 38). Retroviral Gag proteins also can be assembled in vitro in the presence of nucleic acids to form spherical particles (9, 19, 39, 43, 47). The arrangement of Gag within these in vitro-assembled Gag particles is indistinguishable from that found in immature virus particles (6), and the in vitro assembly systems have proved valuable for unraveling the principles of virus assembly (18, 28, 29, 39). Multiple layers of interaction promote the assembly of Gag in vivo, including MA-membrane-MA interactions, CA-CA interactions, and NC-RNA-NC interactions. An extensive body of literature has explored which regions of Gag are required for assembly and which can be replaced or deleted without compromising assembly. MA-membrane-MA interactions contribute but are not essential. NC-RNA-NC interactions appear to function to nonspecifically link Gag molecules together and can be replaced both in vivo and in vitro by other interaction domains such as leucine zippers (4, 13, 20, 32, 48). The C-terminal domain of CA (referred to here as C-CA) and the stretch of amino acids immediately following this domain (termed the spacer peptide [SP] region) are critical for assembly and sensitive to mutation (1, 22, 27, 30).We set out to understand how the substantial sequence variation among Gag proteins in different retroviruses is manifested in structural differences in the immature Gag lattice. To do this, we studied three retroviruses from different genera: the lentivirus human immunodeficiency virus type 1 (HIV-1), the betaretrovirus Mason-Pfizer monkey virus (M-PMV), and the alpharetrovirus Rous sarcoma virus (RSV). These retroviruses are those for which in vitro assembly was first established and has been most extensively studied (6, 19, 24, 28, 29, 35, 43, 47).The domain structures of the three retroviruses differ most substantially upstream of CA. Both M-PMV and RSV have domains located between MA and CA that are absent in HIV (Fig. (Fig.1).1). In M-PMV there are 198 residues forming the pp24 and p12 domains; in RSV there are 84 residues forming the p2a, p2b, and p10 domains. The three retroviruses have different requirements for regions upstream of CA during assembly. The C-terminal 25 residues of p10 are essential for proper immature RSV assembly, both in vitro and in vivo, and these residues are inferred to interact directly with N-CA to stabilize the hexamer by forming contacts between adjacent N-CA domains (35). An equivalent assembly domain has not been described for other retroviruses. Within M-PMV p12 is the so-called internal scaffolding domain that is not essential for assembly in vitro (43) but is required for particle assembly when the precursor is expressed under the control of the M-PMV promoter (41). It is a key domain for the membrane-independent assembly of immature capsids (40).In HIV, five residues upstream of CA must be present for assembly of immature virus-like spherical particles in vitro, although larger upstream extensions, including part of MA, are required for efficient assembly of regular particles, both for HIV and RSV. For HIV, if the entire MA domain is included, in vitro assembly requires the presence of inositol penta- or hexakis phosphate (8). If no sequences upstream of CA are present, the in vitro particles in both of these viruses adopt a mature-type tubular morphology (10, 18). It has been hypothesized that cleavage at the N terminus of N-CA during maturation leads to the N-terminal residues of CA folding back into the N-CA structure to form a β-hairpin. The β-hairpin is important for assembly of the mature CA lattice, whereas its absence is important for immature assembly (23, 42). These requirements explain why, in HIV and RSV, immature Gag lattice-like structures are formed only if regions upstream of CA are present (18). In M-PMV, an immature Gag lattice can be produced when the regions upstream of CA are deleted if this is combined with mutations (such as deleting the initial proline of CA), which prevent β-hairpin formation (43).During maturation, HIV and RSV Gag proteins are cleaved twice between CA and NC to release a small peptide called SP1 or SP. In RSV the most N-terminal of these two cleavages can occur at one of two possible positions such that the released peptide is either 9 or 12 amino acids long (33). In M-PMV only one cleavage occurs between CA and NC, and no short peptide is produced. The region between the final helix of CA and the Zn fingers has been proposed to adopt a helical bundle architecture in HIV and RSV based on bioinformatic prediction, on mutational analysis, and on structural studies (1, 22, 27, 45). In all three viruses, C-CA and the residues immediately downstream are critical for assembly and are sensitive to mutation. C-CA contains the major homology region, a group of residues that are highly conserved across the retroviruses.Cryo-electron tomography (cET) studies of immature virus particles (6, 45) have resolved the electron density of the HIV Gag lattice in three dimensions at low resolution. Using these methods, we have also described the three-dimensional architecture of in vitro-assembled HIV Gag particles (6). In immature viruses and in vitro-assembled particles, Gag is seen to adopt an 8 nm hexameric lattice, as was predicted from previous Fourier analysis of two-dimensional images (7, 46). The hexameric lattice is interrupted by irregularly shaped holes and cracks in the lattice (6, 45). A similar observation has been made using AFM of in vitro-assembled particles of M-PMV Gag (26). These holes and cracks allow an otherwise planar hexameric lattice to form the surface of an approximately spherical particle.The radial positions of the MA, CA, and NC domains had been assigned previously from cryo-electron micrographs (44, 46). Based on these assignments and the shape of the density, the position and relative orientations of CA domains can be modeled into the low-resolution structure of the HIV lattice (6, 45). Density ascribed to the N-terminal domain of CA (N-CA) forms rings around large holes at the 6-fold symmetry positions in the lattice. Below this layer, at the expected radius of the C-CA, are 2-fold densities, interpreted as corresponding to dimers of C-CA. These densities are linked by rodlike densities, which descend into the NC-nucleic acid layer.HIV is the only retrovirus for which the arrangement of Gag in the immature particle has been described in three dimensions. Prior to this work, important open questions were therefore: which features of the arrangement of Gag are conserved between genera and therefore reflect general principles of Gag-Gag interactions, and which features are specific to certain genera? We have applied subtomogram averaging of cryo-electron tomograms to generate reconstructions of in vitro-assembled Gag particles from HIV, M-PMV, and RSV. These allow identification of the general and variable features of the arrangement of Gag and the architecture of immature retroviruses.  相似文献   

10.
The Env protein from gibbon ape leukemia virus (GaLV) has been shown to be incompatible with human immunodeficiency virus type 1 (HIV-1) in the production of infectious pseudotyped particles. This incompatibility has been mapped to the C-terminal cytoplasmic tail of GaLV Env. Surprisingly, we found that the HIV-1 accessory protein Vpu modulates this incompatibility. The infectivity of HIV-1 pseudotyped with murine leukemia virus (MLV) Env was not affected by Vpu. However, the infectivity of HIV-1 pseudotyped with an MLV Env with the cytoplasmic tail from GaLV Env (MLV/GaLV Env) was restricted 50- to 100-fold by Vpu. A Vpu mutant containing a scrambled membrane-spanning domain, VpuRD, was still able to restrict MLV/GaLV Env, but mutation of the serine residues at positions 52 and 56 completely alleviated the restriction. Loss of infectivity appeared to be caused by reduced MLV/GaLV Env incorporation into viral particles. The mechanism of this downmodulation appears to be distinct from Vpu-mediated CD4 downmodulation because Vpu-expressing cells that failed to produce infectious HIV-1 particles nonetheless continued to display robust surface MLV/GaLV Env expression. In addition, if MLV and HIV-1 were simultaneously introduced into the same cells, only the HIV-1 particle infectivity was restricted by Vpu. Collectively, these data suggest that Vpu modulates the cellular distribution of MLV/GaLV Env, preventing its recruitment to HIV-1 budding sites.The gammaretrovirus gibbon ape leukemia virus (GaLV) has been widely used for gene therapy because of its wide host cell tropism and nonpathogenicity (1, 6, 10, 12, 13, 20). The host cell receptor for GaLV Env has been cloned and identified as a sodium-dependent phosphate transporter protein (25, 26). Like other retroviruses, GaLV encodes a single transmembrane surface glycoprotein (GaLV Env), which is cleaved into surface (SU) and transmembrane (TM) subunits (Fig. (Fig.1).1). The TM domain of GaLV Env contains a short 30-amino-acid C-terminal cytoplasmic tail. Although GaLV Env functions well when coupled (pseudotyped) with murine leukemia virus (MLV)-based retroviral vectors, it has been shown to be completely incompatible with HIV-1 (4, 35). When GaLV Env is expressed with HIV-1, essentially no infectious HIV-1 particles are produced (4, 35). The mechanism for this infectivity downmodulation is unknown, but the component of GaLV Env responsible for the restriction has been mapped to the cytoplasmic tail. Replacing the cytoplasmic tail of GaLV Env with the equivalent sequence from MLV Env ameliorates the restriction. Likewise, replacing the cytoplasmic tail of MLV Env with that from GaLV Env confers the restriction (4).Open in a separate windowFIG. 1.Schematic of MLV Env protein. Sequences are the C-terminal cytoplasmic tails of MLV Env, GaLV Env, and human CD4. GaLV sequences in boldface are residues that have been shown to modulate the HIV-1 incompatibility (4). Underlined sequences in CD4 are amino acids required for Vpu-mediated downmodulation (2, 15). Arrows denote the location of MLV/GaLV tail substitution. SU, surface domain; TM, transmembrane domain.Vpu is an 81-amino-acid HIV-1 accessory protein produced from the same mRNA as the HIV-1 Env gene. The N terminus of Vpu contains a membrane-spanning domain, followed by a 50-amino-acid cytoplasmic domain. Vpu is unique to HIV-1 and a few closely related SIV strains. The best-characterized roles for Vpu in the HIV-1 life cycle are modulation of host proteins CD4 and tetherin (also known as BST-2, CD317, and HM1.24) (24, 38, 39). Vpu promotes the degradation of CD4 in the endoplasmic reticulum through a proteasome-dependent mechanism (29). The cytoplasmic tail of Vpu physically interacts with the cytoplasmic tail of CD4 and recruits the human β-transducing repeat-containing protein (β-TrCP) and E3 ubiquitin ligase components to polyubiquitinate and ultimately trigger the degradation of CD4 (18). Two serine residues at positions 52 and 56 of Vpu are phosphorylated by casein kinase-2 and are required for CD4 degradation (31, 32). The membrane-spanning domain of Vpu is not specifically required for CD4 degradation. A mutant protein containing a scrambled membrane-spanning sequence, VpuRD, is still able to trigger the degradation of CD4 (32). The region of CD4 that is targeted by Vpu is approximately 17 to 13 amino acids from the C terminus in the cytoplasmic tail (Fig. (Fig.1)1) (2, 15).In addition to degrading CD4, Vpu has also long been known to result in enhanced viral release (EVR) in certain cell lines (14, 36). Recently, the type I interferon-induced host protein tetherin was identified as being responsible for this Vpu-modulated restriction (24, 38). In the absence of Vpu, tetherin causes particles to remain tethered (hence the name) to the host cell postfission. Although Vpu counteracts the function of tetherin, the exact mechanism has not been fully elucidated. However, the mechanism for tetherin antagonism appears to be distinct from that for modulating CD4. Mutation of the serines 52 and 56 of Vpu abolish CD4 degradation, but only reduce EVR activity (5, 17, 21, 32). Some EVR activity remains even when much of the Vpu cytoplasmic tail is deleted (30). In addition, many mutations in the membrane-spanning domain, such as VpuRD, do not affect CD4 degradation and yet completely abolish EVR activity (27, 30, 37). The critical residues in tetherin for recognition by Vpu appear to be in the membrane-spanning domain and not the cytoplasmic tail (9, 19, 28). Although β-TrCP is required for complete EVR activity, there is no consensus whether the degradation of tetherin is proteasome or lysosome mediated (5, 7, 21) or whether degradation is required at all. In some cases there can be some EVR activity in the absence of tetherin degradation (17, 22).We demonstrate here that Vpu is responsible for the incompatibility between HIV-1 and GaLV Env. Glycoproteins containing the cytoplasmic tail from GaLV Env are prevented from being incorporated into HIV-1 particles by Vpu, effectively reducing infectious particle production by 50- to 100-fold. The serines at positions 52 and 56 are required for this restriction, but the membrane-spanning domain is not. Although the mechanism for this restriction appears similar to CD4 degradation, there are apparent differences. Vpu does not prevent surface expression, and it does not prevent its incorporation into MLV particles. Therefore, the mechanism of restriction appears to involve a system that does not rely directly on global protein degradation.  相似文献   

11.
12.
Cyanophycin (multi-l-arginyl-poly-l-aspartic acid; also known as cyanophycin grana peptide [CGP]) is a putative precursor for numerous biodegradable technically used chemicals. Therefore, the biosynthesis and production of the polymer in recombinant organisms is of special interest. The synthesis of cyanophycin derivatives consisting of a wider range of constituents would broaden the applications of this polymer. We applied recombinant Saccharomyces cerevisiae strains defective in arginine metabolism and expressing the cyanophycin synthetase of Synechocystis sp. strain PCC 6308 in order to synthesize CGP with citrulline and ornithine as constituents. Strains defective in arginine degradation (Car1 and Car2) accumulated up to 4% (wt/wt) CGP, whereas strains defective in arginine synthesis (Arg1, Arg3, and Arg4) accumulated up to 15.3% (wt/wt) of CGP, which is more than twofold higher than the previously content reported in yeast and the highest content ever reported in eukaryotes. Characterization of the isolated polymers by different analytical methods indicated that CGP synthesized by strain Arg1 (with argininosuccinate synthetase deleted) consisted of up to 20 mol% of citrulline, whereas CGP from strain Arg3 (with ornithine carbamoyltransferase deleted) consisted of up to 8 mol% of ornithine, and CGP isolated from strain Arg4 (with argininosuccinate lyase deleted) consisted of up to 16 mol% lysine. Cultivation experiments indicated that the incorporation of citrulline or ornithine is enhanced by the addition of low amounts of arginine (2 mM) and also by the addition of ornithine or citrulline (10 to 40 mM), respectively, to the medium.Cyanophycin (multi-l-arginyl-poly-[l-aspartic acid]), also referred to as cyanophycin grana peptide (CGP), represents a polydisperse nonribosomally synthesized polypeptide consisting of poly(aspartic acid) as backbone and arginine residues bound to each aspartate (49) (Fig. (Fig.1).1). One enzyme only, referred to as cyanophycin synthetase (CphA), catalyzes the synthesis of the polymer from amino acids (55). Several CphAs originating from different bacteria exhibit specific features (2, 7, 5, 32, 50, 51). CphAs from the cyanobacteria Synechocystis sp. strain PCC 6308 and Anabaena variabilis ATCC 29413, respectively, exhibit a wide substrate range in vitro (2, 7), whereas CphA from Acinetobacter baylyi or Nostoc ellipsosporum incorporates only aspartate and arginine (23, 24, 32). CphA from Thermosynechococcus elongatus catalyzes the synthesis of CGP primer independently (5); CphA from Synechococcus sp. strain MA19 exhibits high thermostability (22). Furthermore, two types of CGP were observed concerning its solubility behavior: (i) a water-insoluble type that becomes soluble at high or low pH (34, 48) and (ii) a water-soluble type that was only recently observed in recombinant organisms (19, 26, 42, 50, 56). In the past, bacteria were mainly applied for the synthesis of CGP (3, 14, 18, 53), whereas recently there has been greater interest in synthesis in eukaryotes (26, 42, 50). CGP was accumulated to almost 7% (wt/wt) of dry matter in recombinant Nicotiana tabacum and Saccharomyces cerevisiae (26, 50).Open in a separate windowFIG. 1.Chemical structures of dipeptide building blocks of CGP variants detected in vivo. Structure: 1, aspartate-arginine; 2, aspartate-lysine; 3, aspartate-citrulline; 4, aspartate-ornithine. Aspartic acid is presented in black; the second amino acid of the dipeptide building blocks is shown in gray. The nomenclature of the carbon atoms is given.In S. cerevisiae the arginine metabolism is well understood and has been investigated (30) (see Fig. Fig.2).2). Arginine is synthesized from glutamate via ornithine and citrulline in eight successive steps. The enzymes acetylglutamate synthase, acetylglutamate kinase, N-acetyl-γ-glutamylphosphate reductase, and acetylornithine aminotransferase are involved in the formation of N-α-acetylornithine. The latter is converted to ornithine by acetylornithine acetyltransferase. In the next step, ornithine carbamoyltransferase (ARG3) condenses ornithine with carbamoylphosphate, yielding citrulline. Citrulline is then converted to l-argininosuccinate by argininosuccinate synthetase. The latter is in the final step cleaved into fumarate and arginine by argininosuccinate lyase (ARG4). The first five steps occur in the mitochondria, whereas the last three reactions occur in the cytosol (28, 54). Arginine degradation is initiated by arginase (CAR1) and ornithine aminotransferase (CAR2) (10, 11, 38, 39).Open in a separate windowFIG. 2.Schematic overview of the arginine metabolism in S. cerevisiae. Reactions shown in the shaded area occur in the mitochondria, while the other reactions are catalyzed in the cytosol. Abbreviations: ARG2, acetylglutamate synthase; ARG6, acetylglutamate kinase; ARG5, N-acetyl-γ-glutamyl-phosphate reductase; ARG8, acetylornithine aminotransferase; ECM40, acetylornithine acetyltransferase; ARG1, argininosuccinate synthetase; ARG3, ornithine carbamoyltransferase; ARG4, argininosuccinate lyase; CAR1, arginase; CAR2, ornithine aminotransferase.A multitude of putative applications for CGP derivatives are available (29, 41, 45, 47), thus indicating a need for efficient biotechnological production and for further investigations concerning the synthesis of CGP with alternative properties and different constituents. It is not only the putative application of the polymer as a precursor for poly(aspartic acid), which is used as biodegradable alternative for poly(acrylic acid) or for bulk chemicals, that makes CGP interesting (29, 45-47). In addition, a recently developed process for the production of dipeptides from CGP as a precursor makes the synthesis of CGP variants worthwhile (43). Dipeptides play an important role in medicine and pharmacy, e.g., as additives for malnourished patients, as treatments against liver diseases, or as aids for muscle proliferation (43). Because dipeptides are synthesized chemically (40) or enzymatically (6), novel biotechnological production processes are welcome.  相似文献   

13.
14.
15.
16.
17.
18.
Fusogenic reoviruses utilize the FAST proteins, a novel family of nonstructural viral membrane fusion proteins, to induce cell-cell fusion and syncytium formation. Unlike the paradigmatic enveloped virus fusion proteins, the FAST proteins position the majority of their mass within and internal to the membrane in which they reside, resulting in extended C-terminal cytoplasmic tails (CTs). Using tail truncations, we demonstrate that the last 8 residues of the 36-residue CT of the avian reovirus p10 FAST protein and the last 20 residues of the 68-residue CT of the reptilian reovirus p14 FAST protein enhance, but are not required for, pore expansion and syncytium formation. Further truncations indicate that the membrane-distal 12 residues of the p10 and 47 residues of the p14 CTs are essential for pore formation and that a residual tail of 21 to 24 residues that includes a conserved, membrane-proximal polybasic region present in all FAST proteins is insufficient to maintain FAST protein fusion activity. Unexpectedly, a reextension of the tail-truncated, nonfusogenic p10 and p14 constructs with scrambled versions of the deleted sequences restored pore formation and syncytiogenesis, while reextensions with heterologous sequences partially restored pore formation but failed to rescue syncytiogenesis. The membrane-distal regions of the FAST protein CTs therefore exert multiple effects on the membrane fusion reaction, serving in both sequence-dependent and sequence-independent manners as positive effectors of pore formation, pore expansion, and syncytiogenesis.The only examples of nonenveloped viruses that induce cell-cell fusion and syncytium formation occur within the family Orthoreoviridae, an extremely diverse group of viruses containing segmented double-stranded RNA genomes (9). In recent years, the viral proteins responsible for the syncytiogenic phenotype of the fusogenic orthoreoviruses and aquareoviruses have been identified and characterized (14, 18, 41, 46). These fusion-associated small transmembrane (FAST) proteins define a new family of viral fusogens with several unique biological and biophysical properties. Unlike the well-characterized enveloped virus fusion proteins, reovirus FAST proteins are nonstructural viral proteins and are therefore not involved in mediating virus-cell fusion and virus entry (18, 21, 46). The FAST proteins are instead dedicated to inducing cell-cell fusion and syncytium formation following their expression and trafficking to the plasma membrane of virus-infected or transfected cells (14, 17, 46). Data from previously reported studies also suggest that the FAST proteins serve as virulence factors for the fusogenic reoviruses, promoting virus dissemination and increased tissue destruction (6, 43). How this atypical family of viral fusogens functions to mediate cell-cell membrane fusion remains unclear.The unusual biological role of the FAST proteins as nonstructural, virus-encoded, “cellular” fusogens is embodied in structural features that clearly distinguish the FAST proteins from the membrane fusion proteins of enveloped viruses. There are currently four distinct members of the FAST protein family, named according to their molecular masses: the homologous p10 proteins of avian reovirus (ARV) and Nelson Bay reovirus and the unrelated p14, p15, and p22 proteins of reptilian reovirus (RRV), baboon reovirus, and Atlantic salmon aquareovirus, respectively (14, 18, 41, 46). These proteins are the smallest known fusogens, ranging from 95 to 198 amino acids in size, and assume an asymmetric topology in the plasma membrane, with a single transmembrane domain that separates small N-terminal ectodomains of ∼20 to 41 residues from equal-sized or considerably larger C-terminal endodomains of ∼36 to 141 residues (Fig. (Fig.1A).1A). A number of structural motifs in both the ecto- and endodomains of the FAST proteins have been identified, including sites of acylation, hydrophobic patches, a membrane-proximal polybasic region, and regions rich in proline, cysteine, or arginine, proline, and histidine. Each of the FAST proteins has its own signature repertoire and arrangement of these motifs. Determining how these various motifs contribute to the fusogenic activity of the FAST proteins remains an area of active investigation.Open in a separate windowFIG. 1.ARV p10 and RRV p14 FAST protein topologies and tail truncations. (A) Diagrammatic representation of the p10 and p14 FAST proteins showing their topology in the plasma membrane. Both are single-pass transmembrane proteins with N-terminal ectodomains on the surface of cells and C-terminal endodomains in the cytoplasm. Structural motifs include hydrophobic patches (HP), polybasic motifs (PB), fatty acid modifications (indicated by squiggly lines) that are either the N-terminal myristoylation or palmitoylation of a dicysteine motif (CC), and a polyproline motif (PP). The total number of residues in each protein is indicated by the numbers. (B) The amino acid sequences of the p10 and p14 endodomains are shown, along with the motifs described above. Progressive truncations of the CTs were constructed (arrows), with the numbers indicating the last amino acid present in the full-length proteins or each truncation.Numerous studies of diverse fusion processes define five general steps of the pathway for membrane fusion and syncytium formation: membrane binding, close membrane apposition, hemifusion (i.e., the mixing of the outer leaflets of the two bilayers), stable pore formation, and pore expansion (12, 13, 44). The well-characterized enveloped virus fusion proteins utilize extensive structural rearrangement of their complex ectodomains to provide mechanical energy to draw membranes into close proximity and promote membrane merger (21, 53). The limited size of the FAST protein ectodomains precludes such a mechanical model for membrane fusion, necessitating the development of alternate models to explain how the diminutive FAST proteins breach the thermodynamic barriers that prevent the spontaneous merger of biological membranes. The FAST proteins are both necessary and sufficient to mediate membrane fusion (51). However, data from recent studies indicate that for maximal cell fusion activity, the FAST proteins rely on surrogate adhesins to mediate close membrane apposition (42). Data from recent studies also indicate that a small percentage of the p14 FAST protein expressed in virus-infected or transfected cells is proteolytically processed to generate a bioactive, soluble endodomain that recruits cellular pathways to drive the expansion of stable fusion pores into the extended fusion apertures needed for syncytium formation (50). The FAST proteins therefore utilize accessory proteins to mediate the prefusion (membrane binding and apposition) and postfusion (pore expansion) stages of syncytiogenesis, retaining within their rudimentary structures all that is required to mediate the actual process of membrane merger. This subdivision of the multistep process of syncytium formation is reflected in, and is perfectly suited to, the evolution of the FAST proteins as virus-encoded cellular fusogens.The small size of the FAST protein ectodomains and their donor membrane-focused topology contrast markedly with enveloped virus fusion proteins that position the majority of their mass external to the membrane. While the complex ectodomains of the enveloped virus fusion proteins clearly play an essential role in the fusion reaction, the involvement of their cytoplasmic tails (CTs) is far less certain, and no consistent picture of the role of these C-terminal tails has emerged. The CTs of many enveloped viral fusion proteins, including baculovirus (31), severe acute respiratory syndrome coronavirus (5), vesicular stomatitis virus (36), parainfluenza virus type 2 (56), and influenza A virus subtype H3 (10), play no role in the membrane fusion reaction. Of the fusion protein tails that do modulate the fusion reaction, the majority serve inhibitory roles, including the F proteins of measles virus and parainfluenza virus type 5 SER (7, 45, 52), glycoprotein B from several herpesviruses (22, 24, 28), and the fusion proteins of numerous retroviruses (1, 8, 30, 32, 34, 47, 48). These inhibitory cytoplasmic domains alter the conformation of the fusion protein ectodomains, thereby coupling virion maturation to fusion competence (1, 2, 35, 52, 54). In the few cases where extensive tail truncations adversely affect fusion, these truncations generally decrease but do not eliminate syncytiogenesis, and it is the membrane-proximal portion of the tail that promotes pore formation or pore expansion (20, 25, 26, 32).Since the FAST proteins are nonstructural viral proteins, their CTs (also referred to as endodomains) are not required to suppress fusion activity until after virus particle assembly. At the same time, the disproportionate size of their endodomains strongly suggests that these CTs play an important role in membrane fusion activity. Although one such role of the p14 CT is the generation of a soluble endodomain that recruits cellular factors involved in pore expansion, the majority of p14 is not proteolytically processed, suggesting that FAST protein CTs may serve additional roles as components of the intact protein (50). We now show that C-terminal truncations of the p10 and p14 FAST proteins reduced and eventually eliminated cell-cell fusion. Fluorescence-based pore formation assays coupled with tail reextension studies further revealed that FAST protein CTs drive fusion pore formation and expansion in both sequence-dependent and sequence-independent manners. The membrane-distal regions of FAST protein CTs therefore exert multiple effects on the mechanism of membrane fusion.  相似文献   

19.
Coronavirus membrane (M) proteins play key roles in virus assembly, through M-M, M-spike (S), and M-nucleocapsid (N) protein interactions. The M carboxy-terminal endodomain contains a conserved domain (CD) following the third transmembrane (TM) domain. The importance of the CD (SWWSFNPETNNL) in mouse hepatitis virus was investigated with a panel of mutant proteins, using genetic analysis and transient-expression assays. A charge reversal for negatively charged E121 was not tolerated. Lysine (K) and arginine (R) substitutions were replaced in recovered viruses by neutrally charged glutamine (Q) and leucine (L), respectively, after only one passage. E121Q and E121L M proteins were capable of forming virus-like particles (VLPs) when coexpressed with E, whereas E121R and E121K proteins were not. Alanine substitutions for the first four or the last four residues resulted in viruses with significantly crippled phenotypes and proteins that failed to assemble VLPs or to be rescued into the envelope. All recovered viruses with alanine substitutions in place of SWWS residues had second-site, partially compensating, changes in the first TM of M. Alanine substitution for proline had little impact on the virus. N protein coexpression with some M mutants increased VLP production. The results overall suggest that the CD is important for formation of the viral envelope by helping mediate fundamental M-M interactions and that the presence of the N protein may help stabilize M complexes during virus assembly.Coronaviruses are widespread, medically important respiratory and enteric pathogens of humans and a wide range of animals. New human coronaviruses (HCoV), including severe acute respiratory syndrome CoV (SARS-CoV), HCoV-NL63, and HCoV-HKU1, were recently identified (40, 47). The potential for emergence of other new viruses and the zoonotic nature of some coronaviruses strongly warrants understanding old and new viruses. Understanding vital interactions that take place during virus assembly and conserved domains (CDs) that mediate these interactions can provide insight toward identification of targets for development of antiviral therapeutics and vaccines.Coronaviruses are enveloped positive-stranded RNA viruses that belong to the Coronaviridae family in the Nidovirales order. The virion envelope contains at least three structural proteins, the membrane (M), spike (S), and envelope (E) proteins. The genomic RNA is encapsidated by the N phosphoprotein to form a helical nucleocapsid. The S glycoprotein is the viral attachment protein that facilitates infection through fusion of viral and cellular membranes and is the major target of neutralizing antibodies (13). The M glycoprotein is the most abundant component of the viral envelope and plays required, key roles in virus assembly (9, 20, 31, 33, 41). The E protein is a minor component of the viral envelope that plays an important, not clearly defined, role(s) during virus assembly and release (2, 5, 41).Coronavirus M proteins are divergent in their amino acid content, but all share the same overall basic structural characteristics. The proteins have three TM domains, flanked by a short amino-terminal glycosylated domain and a long carboxy-terminal tail located outside and inside the virion, respectively (14) (Fig. (Fig.11 A). M localizes in the Golgi region when expressed alone (20, 22). M molecules interact with each other and also with the spike and nucleocapsid during virus assembly (8-10, 23, 31, 33). M-M interactions constitute the overall scaffold for the viral envelope. The S protein and a small number of E molecules are interspersed in the M protein lattice in mature virions. Previous studies from a number of labs implicated multiple M domains and residues as being important for coronavirus assembly (6, 8, 9, 17, 43). Coronaviruses assemble and bud at intracellular membranes in the region of the endoplasmic reticulum (ER) Golgi intermediate compartment (ERGIC) (22, 39). Coexpression of only the M and the E proteins is sufficient for virus-like particle (VLP) assembly for most coronaviruses (2, 41).Open in a separate windowFIG. 1.M protein conserved domain and mutants. (A) A linear schematic of the M protein illustrating the relative positions of the three TM domains (black boxes) and the position of the CD in the tail. (B) Alignment of CDs from representative coronaviruses. Full-length amino acid sequences from transmissible gastroenteritis virus (TGEV), feline coronavirus (FeCoV), human coronavirus 229E, human coronavirus NL63, mouse hepatitis virus (MHV), bovine coronavirus (BCoV), human coronavirus OC43, porcine hemagglutinating encephalomyelitis virus (HEV), human coronavirus HKU1, SARS-CoV, infectious bronchitis virus (IBV), and turkey coronavirus (TCoV) were aligned by using CLUSTAL W (25). (C) Mutations introduced into the MHV CD, with + and − symbols used to indicate VLP production and virus recovery for each mutant.The long intravirion (cytoplasmic) tail of M consists of an amphipathic domain following the third TM and a short hydrophilic region at the carboxyl end of the tail (Fig. (Fig.1A).1A). The amphipathic domain appears to be closely associated with the membrane (34). At the amino terminus of the amphipathic domain, there is a highly conserved 12-amino-acid domain (SWWSFNPETNNL), consisting of residues 114 to 125 in the mouse hepatitis virus (MHV) A59 M protein (Fig. (Fig.1B)1B) (19). These residues are almost identically conserved across the entire Coronaviridae family. Because of the crucial role that M plays in virus assembly and the high conservation of this domain, we hypothesized that it is functionally important for virus assembly. To test this, a series of changes were introduced in the CD. The functional impact of the changes was studied in the context of the virus by genetic analysis and the ability of the mutant M proteins to participate in VLP assembly. The results show that the CD is functionally important for M protein to participate in virus assembly. The domain may help mediate important lateral interactions between M molecules. The results suggest that the N protein helps stabilize M complexes during virus assembly.  相似文献   

20.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  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