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.
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   

2.
3.
4.
5.
6.
Nonstructural protein 1 (nsp1), a 28-kDa protein in the bovine coronavirus (BCoV) and closely related mouse hepatitis coronavirus, is the first protein cleaved from the open reading frame 1 (ORF 1) polyprotein product of genome translation. Recently, a 30-nucleotide (nt) cis-replication stem-loop VI (SLVI) has been mapped at nt 101 to 130 within a 288-nt 5′-terminal segment of the 738-nt nsp1 cistron in a BCoV defective interfering (DI) RNA. Since a similar nsp1 coding region appears in all characterized groups 1 and 2 coronavirus DI RNAs and must be translated in cis for BCoV DI RNA replication, we hypothesized that nsp1 might regulate ORF 1 expression by binding this intra-nsp1 cistronic element. Here, we (i) establish by mutation analysis that the 72-nt intracistronic SLV immediately upstream of SLVI is also a DI RNA cis-replication signal, (ii) show by gel shift and UV-cross-linking analyses that cellular proteins of ∼60 and 100 kDa, but not viral proteins, bind SLV and SLVI, (SLV-VI) and (iii) demonstrate by gel shift analysis that nsp1 purified from Escherichia coli does not bind SLV-VI but does bind three 5′ untranslated region (UTR)- and one 3′ UTR-located cis-replication SLs. Notably, nsp1 specifically binds SLIII and its flanking sequences in the 5′ UTR with ∼2.5 μM affinity. Additionally, under conditions enabling expression of nsp1 from DI RNA-encoded subgenomic mRNA, DI RNA levels were greatly reduced, but there was only a slight transient reduction in viral RNA levels. These results together indicate that nsp1 is an RNA-binding protein that may function to regulate viral genome translation or replication but not by binding SLV-VI within its own coding region.Coronaviruses (CoVs) (59) cause primarily respiratory and gastroenteric diseases in birds and mammals (35, 71). In humans, they most commonly cause mild upper respiratory disease, but the recently discovered human CoVs (HCoVs), HCoV-NL63 (65), HCoV-HKU1 (73), and severe acute respiratory syndrome (SARS)-CoV (40) cause serious diseases in the upper and lower respiratory tracts. The SARS-CoV causes pneumonia with an accompanying high (∼10%) mortality rate (69). The ∼30-kb positive-strand CoV genome, the largest known among RNA viruses, is 5′ capped and 3′ polyadenylated and replicates in the cytoplasm (41). As with other characterized cytoplasmically replicating positive-strand RNA viruses (3), translation of the CoV genome is an early step in replication, and terminally located cis-acting RNA signals regulate translation and direct genome replication (41). How these happen mechanistically in CoVs is only beginning to be understood.In the highly studied group 2 mouse hepatitis coronavirus model (MHV A59 strain) and its close relative the bovine CoV (BCoV Mebus strain), five higher-order cis-replication signals have been identified in the 5′ and 3′ untranslated regions (UTRs). These include two in the 5′ UTR required for BCoV defective interfering (DI) RNA replication (Fig. (Fig.1A)1A) described as stem-loop III (SLIII) (50) and SLIV (51). Recently, the SLI region in BCoV (15) has been reanalyzed along with the homologous region in MHV and is now described as comprising SL1 and SL2 (Fig. (Fig.1A),1A), of which SL2 has been shown to be a cis-replication structure in the context of the MHV genome (38). In the 3′ UTR, two higher-order cis-replication structures have been identified that function in both DI RNA and the MHV genome. These are a 5′-proximal bulged SL and adjacent pseudoknot that potentially act together as a unit (23, 27, 28, 72) and a 3′-proximal octamer-associated bulged SL (39, 76) (Fig. (Fig.1A).1A). In addition, the 5′-terminal 65-nucleotide (nt) leader and the 3′-terminal poly(A) tail have been shown to be cis-replication signals for BCoV DI RNA (15, 60).Open in a separate windowFIG. 1.RNA structures in the BCoV genome tested for nsp1 binding. (A) BCoV 5′-terminal and 3′-terminal cis-acting RNA SL structures and flanking sequences identified for BCoV DI RNA replication. Regions of the genome are identified and SL cis-replication elements are identified schematically. Open boxes at nt 100 and 211 identify AUG start codons for the short upstream ORF and ORF 1, respectively. A closed box at nt 124 identifies the UAG stop codon for the short upstream ORF. Shown below the SL structures are the RNA segments used as 32P-labeled probes in the gel shift assays. BSL-PK, bulged SL-pseudoknot; 8mer-BSL, octamer-associated bulged SL. (B) Gel shift assays for probes when used with purified nsp1. Protein-RNA complexes identifying a shifted probe are labeled C.In CoVs, the 5′-proximal open reading frame (ORF) of ∼20 kb (called ORF 1) comprising the 5′ two-thirds of the genome is translated to overlapping polyproteins of ∼500 and ∼700 kDa, named pp1a and pp1ab (41). pp1ab is formed by a −1 ribosomal frameshift event at the ORF1a-ORF1b junction during translation (41). pp1a and pp1ab are proteolytically processed into potentially 16 nonstructural protein (nsp) end products or partial end products that are proposed to function together as the replicase (24). ORF 1a encodes nsps 1 to 11 which include papain-like proteases (nsp3), a 3C-like main protease (nsp5), membrane-anchoring proteins (nsps 4 and 6), a potential primase (nsp8), and RNA-binding proteins (nsp 7/nsp 8 complex and nsps 9 and 10) of imprecisely understood function (19, 20, 24, 25, 29, 43, 49, 77). ORF 1b encodes nsps 12 to 16 which function as an RNA-dependent RNA polymerase, a helicase, an exonuclease, an endonuclease, and a 2′-O-methyltransferase, respectively (6, 17, 24, 44). 3′ Proximal genomic ORFs encoding structural and accessory proteins are translated from a 3′-nested set of subgenomic mRNAs (sgmRNAs) (41).The N-terminal ORF 1a protein, nsp1, in the case of BCoV and MHV is also named p28 to identify the cleaved 28-kDa product (18). The precise role of nsp1 in virus replication has not been determined, but it is known that a sequence encoding an N-proximal nsp1 region in MHV (nt 255 to 369 in the 738-nt coding sequence) cannot be deleted from the genome without loss of productive infection (10). nsp1 also directly binds nsp7 and nsp10 (11) and by confocal microscopy is found associated with the membranous replication complex (10, 66) and virus assembly sites (11). The amino acid sequence of nsp1 is poorly conserved among CoVs, indicating that it may be a protein that interacts with cellular components (1, 58). In the absence of other viral proteins, MHV nsp1 induces general host mRNA degradation (79) and cell cycle arrest (16). The SARS-CoV nsp1 homolog, a 20-kDa protein, has been reported to cause mRNA degradation (30, 45), inhibition of host protein synthesis (30, 45, 70), inhibition of interferon signaling (70, 79), and cytokine dysregulation in lung cells (36).In this study, we examine the RNA-binding properties of BCoV nsp1 with the hypothesis that it is a potential regulator of translation or replication through its binding of SLVI mapping within its coding region. The rationale for this hypothesis stems from five observations. (i) In the BCoV DI RNA, the 5′-terminal one-third (approximately) of the nsp1 cistron and the entire nucleocapsid (N) protein cistron together comprise the single contiguous ORF in the DI RNA, and most of both coding regions appear required for DI RNA replication (15). (ii) The partial nsp1 cistron in the DI RNA must be translated in cis for DI RNA replication in helper virus-infected cells (12, 14). (iii) A similar part of the nsp1 cistron is found in the genome of all characterized naturally occurring group 1 and 2 CoV DI RNAs described to date (7, 8). (iv) A cis-acting SL named SLVI is found within the partial nsp1 cistron in the BCoV DI RNA (12). (v) Translation, which involves a 5′→3′ transit of ribosomes, and negative-strand synthesis, which involves a 3′→5′ transit of the RNA-dependent RNA polymerase, cannot simultaneously occur on the same molecule with a single ORF (4, 31). Thus, to enable genome replication an inhibition of translation at least early in infection for cytoplasmically replicating positive-strand RNA viruses is required (4, 5, 22, 32). Mechanisms of translation inhibition have been described for the Qβ viral genome, wherein the viral replicase autoregulates translation by binding an intracistronic cis-replication element (32), and for the polio virus genome, wherein genome circularization inhibits the early translation step (5, 22). Therefore, since nsp1 is synthesized early and also contains an intracistronic cis-replication element, we postulated that it is autoregulatory with RNA binding properties.Here, we do the following: (i) demonstrate by mutagenesis analysis that the 72-nt SLV, mapping immediately upstream of SLVI and within the partial nsp1 cistron, is also a cis-acting DI RNA replication element; (ii) show by gel shift and UV cross-linking analyses that there is likely no binding of an intracellular viral protein to SLV and SLVI (SLV-VI), but there is binding of unidentified cellular proteins of ∼60 and 100 kDa; and (iii) show by gel shift analysis that recombinant nsp1 purified from Escherichia coli does not bind SLV-VI but does bind SLs I to IV in the 5′ UTR and also the 3′-terminal bulged SL in the 3′ UTR, suggesting a possible regulatory role at these sites. Notably, specific binding with ∼2.5 μM affinity of nsp1 to SLIII and its flanking regions in the 5′ UTR was observed. Additionally, we show that, under conditions that would express nsp1 from a DI RNA-encoded sgmRNA, DI RNA levels are greatly reduced; viral RNA species levels, however, are reduced only slightly, and this reduction is transient. These results together indicate that nsp1 is an RNA-binding protein that may function as a regulator of viral translation or replication but not through its binding of cis-acting SLs V and VI within its own cistron.  相似文献   

7.
8.
9.
10.
11.
The human immunodeficiency virus type 1 (HIV-1) matrix (MA) protein targets HIV-1 precursor Gag (PrGag) proteins to assembly sites at plasma membrane (PM) sites that are enriched in cholesterol and phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2]. MA is myristoylated, which enhances membrane binding, and specifically binds PI(4,5)P2 through headgroup and 2′ acyl chain contacts. MA also binds nucleic acids, although the significance of this association with regard to the viral life cycle is unclear. We have devised a novel MA binding assay and used it to examine MA interactions with membranes and nucleic acids. Our results indicate that cholesterol increases the selectivity of MA for PI(4,5)P2-containing membranes, that PI(4,5)P2 binding tolerates 2′ acyl chain variation, and that the MA myristate enhances membrane binding efficiency but not selectivity. We also observed that soluble PI(4,5)P2 analogues do not compete effectively with PI(4,5)P2-containing liposomes for MA binding but surprisingly do increase nonspecific binding to liposomes. Finally, we have demonstrated that PI(4,5)P2-containing liposomes successfully outcompete nucleic acids for MA binding, whereas other liposomes do not. These results support a model in which RNA binding protects MA from associating with inappropriate cellular membranes prior to PrGag delivery to PM assembly sites.The matrix (MA) domain of the human immunodeficiency virus type 1 (HIV-1) precursor Gag (PrGag) protein serves several functions in the viral replication cycle. One essential function is to target PrGag proteins to their assembly sites at the plasma membranes (PMs) of infected cells (4, 5, 11, 16, 25, 29, 30, 33, 35, 39, 43-45, 47, 50, 54, 56, 57). A second function is the recruitment of the viral surface/transmembrane (SU/TM; also referred to as gp120/gp41) envelope (Env) protein complex into virions (14, 15, 18, 19, 27, 51-53). In addition to these activities, numerous reports have attributed nucleic acid binding properties to retroviral MAs (24, 38, 47), and with some viruses MA appears to serve in an encapsidation capacity (24). While no encapsidation role has been assigned for HIV-1 MA, experiments have shown that MA can substitute for the HIV-1 nucleocapsid (NC) protein assembly function (38) under some circumstances, presumably by virtue of its facility to concentrate PrGag proteins by binding them to RNAs (38).A number of structural studies have been conducted on HIV-1 MA (1, 22, 41, 42, 49). The protein is N terminally myristoylated and composed of six α helices, capped by a three-strand β sheet (7, 22, 41, 42, 49). The protein trimerizes in solution and in crystals (22, 28, 49) and recently has been shown to organize as hexamers of trimers on lipid membranes (1). The membrane binding face of HIV-1 MA is basic, fostering its ability to associate with negatively charged phospholipid headgroups (1, 22, 30, 41, 42, 49). The importance of such an interaction has been underscored in molecular genetic experiments which demonstrated that depletion of PM phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2] reduced the assembly efficiency of HIV-1 (9, 36). Consistent with these observations, HIV-1 MA preferentially binds to soluble PI(4,5)P2 mimics through contacts with the headgroup and 2′ acyl chain, and binding promotes exposure of the MA myristate group and protein oligomerization (17, 21, 40-43, 46). However, PI(4,5)P2 is not the only lipid to demonstrate an association with HIV-1. In particular, HIV-1 appears to assemble at cholesterol-rich PM sites, cholesterol is highly enriched in HIV-1 virions, and cholesterol depletion reduces viral infectivity (2, 6, 8, 20, 23, 26, 31, 34, 37). The HIV-1 lipidome shows additional differences from the PM lipids of infected cells (2, 5, 8), suggesting that other lipids could affect PrGag-membrane binding or virus assembly site selection.To gain a better understanding of the functions and interactions of HIV-1 MA, we have examined the liposome and nucleic acid binding properties of purified myristoylated MA. Using liposome flotation assays and a novel liposome bead binding assay, we have demonstrated that the PI(4,5)P2 binding specificity of MA is enhanced by cholesterol, that protein myristoylation increases membrane binding efficiency but not specificity, and that 2′ acyl chain variation is compatible with PI(4,5)P2 binding. We also examined whether soluble PI(4,5)P2 mimics could compete with liposomes for MA binding. Surprisingly, we found that soluble mimics not only failed to compete with PI(4,5)P2 liposomes but also increased MA binding to membranes that do not contain acidic phospholipids. Finally, we have observed that while MA does bind nucleic acids, nucleic acid binding is outcompeted by PI(4,5)P2-containing liposomes. Our results suggest models for PrGag-membrane and RNA association and the HIV-1 assembly pathway.  相似文献   

12.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

13.
The positive-strand RNA genome of Japanese encephalitis virus (JEV) terminates in a highly conserved 3′-noncoding region (3′NCR) of six domains (V, X, I, II-1, II-2, and III in the 5′-to-3′ direction). By manipulating the JEV genomic RNA, we have identified important roles for RNA elements present within the 574-nucleotide 3′NCR in viral replication. The two 3′-proximal domains (II-2 and III) were sufficient for RNA replication and virus production, whereas the remaining four (V, X, I, and II-1) were dispensable for RNA replication competence but required for maximal replication efficiency. Surprisingly, a lethal mutant lacking all of the 3′NCR except domain III regained viability through pseudoreversion by duplicating an 83-nucleotide sequence from the 3′-terminal region of the viral open reading frame. Also, two viable mutants displayed severe genetic instability; these two mutants rapidly developed 12 point mutations in domain II-2 in the mutant lacking domains V, X, I, and II-1 and showed the duplication of seven upstream sequences of various sizes at the junction between domains II-1 and II-2 in the mutant lacking domains V, X, and I. In all cases, the introduction of these spontaneous mutations led to an increase in RNA production that paralleled the level of protein accumulation and virus yield. Interestingly, the mutant lacking domains V, X, I, and II-1 was able to replicate in hamster BHK-21 and human neuroblastoma SH-SY5Y cells but not in mosquito C6/36 cells, indicating a cell type-specific restriction of its viral replication. Thus, our findings provide the basis for a detailed map of the 3′ cis-acting elements in JEV genomic RNA, which play an essential role in viral replication. They also provide experimental evidence for the function of 3′ direct repeat sequences and suggest possible mechanisms for the emergence of these sequences in the 3′NCR of JEV and perhaps in other flaviviruses.Japanese encephalitis virus (JEV), a mosquito-borne flavivirus of the family Flaviviridae, is serologically related to several significant human pathogens, including West Nile virus (WNV), Kunjin virus (KUNV), St. Louis encephalitis virus, and Murray Valley encephalitis virus. It is also phylogenetically close to other clinically important human pathogens, including yellow fever virus (YFV) and dengue virus (DENV) (11, 67). JEV is the leading cause of viral encephalitis in Southeast Asia, including China, Japan, Korea, the Philippines, Thailand, and India, and it has begun to expand throughout the Indonesian archipelago and as far as Australia (21, 43). Despite the fact that JEV is generally asymptomatic, ∼50,000 cases are reported annually, and the disease has a mortality rate of ∼25%, mainly in children and young adults (29, 63). Thus, the geographic expansion and clinical importance of JEV infection have drawn increasing attention from the international public health community (44, 71).Like other flaviviruses, JEV is a spherical enveloped virus (∼50 nm diameter) with a single-stranded positive-sense RNA genome that contains a 5′ cap structure but lacks a 3′ polyadenylated tail. Its genomic RNA of ∼11,000 nucleotides (nt) consists of a single long open reading frame (ORF) with two noncoding regions (NCRs) at the 5′ and 3′ ends (41, 84). The ORF is translated into an ∼3,400-amino acid polyprotein precursor, which is co- or posttranslationally cleaved by a cellular protease(s) or a viral protease complex into 10 mature proteins: (i) three structural proteins, the capsid (C), premembrane (prM; which is further processed into pr and M), and envelope (E) proteins; and (ii) seven nonstructural proteins, NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5, as arranged in the genome (13, 41, 84). The nonstructural proteins, together with cellular factors, form a viral replicase complex that directs the replication of the genomic RNA in the cytoplasm of the host cell in association with perinuclear membranes (40, 74). For the synthesis of the genomic RNA to take place, this replicase complex must specifically recognize viral cis-acting RNA elements, defined by primary sequences or secondary/tertiary structures. These RNA elements are found in various locations within the genome but most frequently are located in the 5′- and 3′NCRs (23, 47). The identification and characterization of these cis-acting RNA elements is critical for understanding the complete cycle of JEV genome replication.The availability of the complete nucleotide sequence of YFV genomic RNA (57) has led to the identification of three major conserved elements in the 5′- and 3′-terminal regions of the genomic RNA that contain the short primary sequences and secondary structures required for flavivirus RNA replication. (i) Both ends of the genomic RNA terminate with the conserved dinucleotides 5′-AG and CU-3′ (9, 10, 32, 45, 57, 72, 73) in all flaviviruses except an insect cell fusing agent virus (12). Mutations substituting another nucleotide for one of these four nucleotides in KUNV or WNV replicon RNA are known to abolish or compromise RNA replication (35, 69). (ii) A 3′ stem-loop structure (3′SL) has been recognized in all flaviviruses within the ∼90-nt 3′-terminal region of the genomic RNA (9, 45, 57). The structural and functional importance of this 3′SL in RNA replication has been demonstrated in several flaviviruses (9, 18, 49, 50, 61, 70, 82, 86). (iii) The presence of short 5′ and 3′ cyclization sequences (5′CYC and 3′CYC, respectively) in all mosquito-borne flaviviruses suggests that flavivirus genomes can cyclize via 5′-3′ long-range base-pairing interaction, since the 3′CYC upstream of the 3′SL is complementary to the 5′CYC in the 5′ coding region of the C protein (30). The role of these CYC motifs in RNA replication has been well characterized via cell-based assays in many mosquito-borne flaviviruses, including KUNV (34), WNV (42), YFV (8, 14), and DENV (2, 22, 49), and in cell-free systems in the case of WNV (51) and DENV (1, 3, 79, 80). Other RNA elements that have recently been shown to be important for RNA replication in DENV and WNV include an additional pair of complementary sequences (designated 5′- and 3′UARs) that participate in genome cyclization (3, 4, 17, 87) and a 5′ stem-loop structure (designated 5′SLA) present within the 5′NCR that promotes RNA synthesis in association with the 3′NCR (22).In all flaviviruses, the 3′NCR of the genomic RNA is relatively long (∼400 to ∼800 nt), with an array of conserved primary sequences and secondary structures. Although significant progress has been made in identifying cis-acting elements within the 3′NCRs that are essential for RNA replication, most of these elements (i.e., the 3′CYC, 3′SL, and CU-3′) are limited to the ∼100-nt 3′-terminal region that is highly conserved in these viruses (see recent reviews in references 23 and 47). However, the functional importance of the remaining 5′-proximal region of the 3′NCR, which differs in sequence between the various serological groups, is poorly understood. In particular, comparative sequence analyses and genetic algorithm-based computer modeling have suggested that in addition to the well-studied ∼100-nt 3′-proximal region, the remaining ∼474-nt 5′-proximal region of the 574-nt JEV 3′NCR also contains several RNA elements that may play critical roles in the viral life cycle (52, 55, 56, 68). To date, however, experimental evidence for the functional importance of these potential RNA elements in JEV genomic RNA replication is lacking.In the present study, we have identified and characterized the 3′ cis-acting RNA elements within the JEV 3′NCR and shown that they play an essential and/or regulatory role in genomic RNA replication. In particular, we have constructed and functionally characterized genome-length JEV mutant cDNAs with a series of 5′-to-3′ or 3′-to-5′ progressive deletions within the 3′NCR. In addition to identifying particular mutations within this region that affect either the competence or efficiency of genomic RNA replication, we found that the serial passaging of these mutants in susceptible BHK-21 cells produced a large number of pseudorevertants bearing a wide variety of spontaneous point mutations and sequence duplications, some of which were capable of restoring the replication competence of the defective mutants or enhancing replication efficiency. In addition, we assessed the replication of these mutants in three different cell types (BHK-21, SH-SY5Y, and C6/36 cells). Collectively, these data offer new insights into the functional importance of 3′ cis-acting RNA elements that regulate the cell type-dependent replication of JEV and perhaps other closely related mosquito-borne flaviviruses. Our findings also provide experimental evidence for the emergence of functional 3′ direct repeat sequences that are duplicated from the coding region and 3′NCR of JEV genomic RNA.  相似文献   

14.
15.
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   

16.
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   

17.
18.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

19.
20.
The p6 region of HIV-1 Gag contains two late (L) domains, PTAP and LYPXnL, that bind Tsg101 and Alix, respectively. Interactions with these two cellular proteins recruit members of the host''s fission machinery (ESCRT) to facilitate HIV-1 release. Other retroviruses gain access to the host ESCRT components by utilizing a PPXY-type L domain that interacts with cellular Nedd4-like ubiquitin ligases. Despite the absence of a PPXY motif in HIV-1 Gag, interaction with the ubiquitin ligase Nedd4-2 was recently shown to stimulate HIV-1 release. We show here that another Nedd4-like ubiquitin ligase, Nedd4-1, corrected release defects resulting from the disruption of PTAP (PTAP), suggesting that HIV-1 Gag also recruits Nedd4-1 to facilitate virus release. Notably, Nedd4-1 remediation of HIV-1 PTAP budding defects is independent of cellular Tsg101, implying that Nedd4-1''s function in HIV-1 release does not involve ESCRT-I components and is therefore distinct from that of Nedd4-2. Consistent with this finding, deletion of the p6 region decreased Nedd4-1-Gag interaction, and disruption of the LYPXnL motif eliminated Nedd4-1-mediated restoration of HIV-1 PTAP. This result indicated that both Nedd4-1 interaction with Gag and function in virus release occur through the Alix-binding LYPXnL motif. Mutations of basic residues located in the NC domain of Gag that are critical for Alix''s facilitation of HIV-1 release, also disrupted release mediated by Nedd4-1, further confirming a Nedd4-1-Alix functional interdependence. In fact we found that Nedd4-1 binds Alix in both immunoprecipitation and yeast-two-hybrid assays. In addition, Nedd4-1 requires its catalytic activity to promote virus release. Remarkably, RNAi knockdown of cellular Nedd4-1 eliminated Alix ubiquitination in the cell and impeded its ability to function in HIV-1 release. Together our data support a model in which Alix recruits Nedd4-1 to facilitate HIV-1 release mediated through the LYPXnL/Alix budding pathway via a mechanism that involves Alix ubiquitination.Retroviral Gag polyproteins bear short conserved sequences that control virus budding and release. As such, these motifs have been dubbed late or L domains (49). Three types of L domains have thus far been characterized: PT/SAP, LYPXnL, and PPPY motifs (5, 9, 32). They recruit host proteins known to function in the vacuolar protein sorting (vps) of cargo proteins and the generation of multivesicular bodies (MVB) compartments (2). It is currently accepted that budding of vesicles into MVB involves the sequential recruitment of endosomal sorting complexes required for transport (ESCRT-I, -II, and -III) and the activity of the VPS4 AAA-ATPase (22). These sorting events are believed to be triggered by recognition of ubiquitin molecules conjugated to cargo proteins (20, 24, 41). For retrovirus budding, L domain motifs are the primary signals in Gag that elicit the recruitment of ESCRT components to facilitate viral budding. Consequently, mutations in L domain motifs or dominant-negative interference with the function of ESCRT-III members or the VPS4 ATPase adversely affect virus release. This indicates that Gag interactions with the ESCRT machinery are necessary for virus budding and separation from the cell (7, 10, 15, 16, 21, 28, 44).Two late domains have been identified within the p6 region of human immunodeficiency virus type 1 (HIV-1) Gag protein: the PTAP and LYPXnL motifs. The PTAP motif binds the cellular protein Tsg101 (15, 39, 40, 47), whereas the LYPXnL motif is the docking site for Alix (44). Tsg101 functions in HIV-1 budding (15) as a member of ESCRT-I (30, 48), a soluble complex required for the generation of MVB. This process is topologically similar to HIV-1 budding and requires the recruitment of ESCRT-III members called the charged-multivesicular body proteins (3, 29, 48) and the activity of the VPS4 AAA-ATPase (4, 48). In addition to binding the LYPXnL motif, Alix also interacts with the nucleocapsid (NC) domain of HIV-1 Gag (13, 38), thus linking Gag to components of ESCRT-III that are critical for virus release (13).Other retroviruses, including the human T-cell leukemia virus (HTLV) and the Moloney murine leukemia virus (MoMLV), utilize the PPPY-type L domain to efficiently release virus (7, 26, 51). The PPPY motif binds members of the Nedd4-like ubiquitin ligase family (6, 7, 16, 19, 25, 43), whose normal cellular function is to ubiquitinate cargo proteins and target them into the MVB sorting pathway (11, 12, 20). Members of the Nedd4-like ubiquitin ligase family include Nedd4-1, Nedd4-2 (also known as Nedd4L), WWP-1/2, and Itch. They contain three distinct domains: an N-terminal membrane binding C2 domain (12), a central PPPY-interacting WW domain (43), and a C-terminal HECT domain that contains the ubiquitin ligase active site (42). The functional requirement for the binding of Nedd4-like ubiquitin ligases to the PPPY motif in virus budding has been demonstrated (7, 16, 18, 19, 25, 26, 28, 50, 51). Overexpression of dominant-negative mutants of Nedd4-like ligases, ESCRT-III components, or VPS4 cause a potent inhibition of PPPY-dependent virus release (7, 19, 29, 31, 52) and induce assembly and budding defects similar to those observed after perturbation of the PPPY motif (26, 51). These observations demonstrated that Nedd4-like ligases connect Gag encoding PPPY motif to ESCRT-III and VPS4 proteins to facilitate virus release.Whereas the role of Nedd4-like ubiquitin ligases in virus budding has been established, the protein interactions that link them to the cell''s ESCRT-III pathway are still unknown. Evidence for associations of Nedd4-like ligases with ESCRT proteins have been previously reported and include: the binding of Nedd4-like ubiquitin ligases LD1 and Nedd4-1 to ESCRT-I member Tsg101 (6, 31), the colocalization of multiple Nedd4-like ubiquitin ligases with endosomal compartments (1, 28), the requirement of the cell''s ESCRT pathway for Itch mediated L domain independent stimulation of MoMLV release (23), and the ubiquitination of ESCRT-I components with a shorter isoform, Nedd4-2s (8). Therefore, Nedd4-like ubiquitin ligase interactions with members of the cell''s ESCRT pathway may provide retroviral Gag with access to the host budding machinery required for virus release.Although HIV-1 Gag does not carry the PPPY canonical sequence known to interact with Nedd4-like ubiquitin ligases, both Nedd4-1 and Nedd4-2 were shown to restore the release of the HIV-1 PTAP mutant, albeit Nedd4-1 with less efficiency than Nedd4-2 (8, 46). These findings suggested that HIV-1 might utilize cellular Nedd4-like ubiquitin ligases to increase virus release. We present here evidence demonstrating that Nedd4-1 interacts with Gag and enhances HIV-1 PTAP virus release. Furthermore, we show that Nedd4-1''s function in HIV-1 release is distinct from that of Nedd4-2 in both its viral and cellular requirements. Notably, we found that Nedd4-1 enhancement of HIV-1 release requires the Alix-binding LYPXnL L domain motif in the p6 region and basic residues in the NC domain. In addition, Alix''s facilitation of HIV-1 release requires cellular Nedd4-1, since mutations in NC that prevented Alix-mediated HIV-1 release also eliminated release by overexpression of Nedd4-1. This suggested a Nedd4-1-Alix physical and functional interdependence. In agreement with this, we found Nedd4-1 to bind and ubiquitinate Alix in the cell. Taken together, these results support a model in which Alix recruits Nedd4-1 to facilitate late steps of HIV-1 release through the LYPXnL L domain motif via a mechanism that involves Alix ubiquitination.  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