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.
Exponentially growing recA mutant cells of Escherichia coli display pronounced DNA degradation that starts at the sites of DNA damage and depends on RecBCD nuclease (ExoV) activity. As a consequence of this “reckless” DNA degradation, populations of recA mutants contain a large proportion of anucleate cells. We have found that both DNA degradation and anucleate-cell production are efficiently suppressed by mutations in the xonA (sbcB) and sbcD genes. The suppressive effects of these mutations were observed in normally grown, as well as in UV-irradiated, recA cells. The products of the xonA and sbcD genes are known to code for the ExoI and SbcCD nucleases, respectively. Since both xonA and sbcD mutations are required for strong suppression of DNA degradation while individual mutations have only a weak suppressive effect, we infer that ExoI and SbcCD play partially redundant roles in regulating DNA degradation in recA cells. We suggest that their roles might be in processing (blunting) DNA ends, thereby producing suitable substrates for RecBCD binding.The RecA protein plays a central role in homologous recombination and recombinational DNA repair in Escherichia coli, as well as in other bacterial species. It catalyzes the key stages of the recombination process—homologous pairing and DNA strand exchange. Cells carrying null mutations in the recA gene are completely deficient for homologous recombination and are extremely sensitive to DNA-damaging agents (for a review, see references 21, 24, and 25). Populations of recA null mutants contain a large proportion (50 to 60%) of nonviable cells, reflecting the inability of these mutants to repair spontaneously occurring DNA damage (31). Also, exponentially growing recA cells display pronounced spontaneous DNA degradation that presumably starts at the sites of DNA damage and that depends on RecBCD nuclease (ExoV) activity (5, 48). This phenotype of recA cells is aggravated after DNA-damaging treatment, such as UV irradiation (48).According to the present data, the majority of RecA-catalyzed DNA transactions in E. coli start with binding of the RecA protein onto single-stranded DNA (ssDNA) substrates. This binding is mediated by the RecBCD and/or RecFOR protein, which helps RecA to overcome hindrance imposed by the SSB protein during competition for the DNA substrate. The RecBCD and RecFOR proteins begin RecA polymerization on ssDNA, giving rise to a nucleoprotein filament that is indispensable for further recombination reactions (3, 33; reviewed in reference 44).The RecBCD enzyme is crucial for initiation of recombinational processes at double-stranded DNA (dsDNA) ends (or breaks [DSBs]) in wild-type E. coli (a set of reactions known as the RecBCD pathway) (9, 43, 44). Upon recognizing a blunt or nearly blunt dsDNA end and binding to it, RecBCD acts as a combination of powerful helicase and nuclease, thus unwinding and simultaneously degrading both strands of the DNA duplex. After encountering a specific octanucleotide sequence designated Chi, the strong 3′-5′ nuclease activity of the enzyme is attenuated and a weaker 5′-3′ nuclease activity is upregulated (1). This Chi-dependent modification allows RecBCD to create a long 3′ ssDNA tail and to direct the loading of RecA protein onto it (2, 3). In vivo data suggest that this transition of RecBCD from a nuclease to a recombinase mode of action requires the presence of the RecA protein, suggesting that the two proteins might interact (27).In wild-type E. coli cells, the RecFOR protein complex works predominantly on DNA gaps, which may arise in chromosomes due to replication forks passing over the noncoding lesions (e.g., UV-induced pyrimidine dimers) or may be present in replication forks stalled at different obstacles in DNA (44). On the other hand, the RecFOR complex has an important role in recBC sbcBC(D) mutant cells, replacing the RecA-loading activity of RecBCD during recombination reactions starting from dsDNA ends. Recombination reactions mediated by RecFOR proteins are termed the RecF (or RecFOR) pathway (44).Cells mutated in the recB and/or recC gene exhibit strong deficiency in conjugational and transductional recombination, as well as in the repair of DSBs (8, 21). These defects can be rectified by extragenic sbcB and sbcC(D) suppressor mutations that inactivate two nucleases, thus enabling full efficiency of the RecF pathway on dsDNA ends (21, 44). The sbcB gene (also designated xonA) encodes exonuclease I (ExoI), the enzyme that digests ssDNA in the 3′-5′ direction (23). The sbcC and sbcD genes encode subunits of the SbcCD nuclease, which acts both as an endonuclease that cleaves hairpin structures and as an exonuclease that degrades linear dsDNA molecules (10, 11). Inactivation of either of the two subunits leads to the loss of SbcCD enzyme activity (18).The exact mechanism of activation of the RecF pathway by sbc mutations is not completely understood. A plausible explanation is that inactivation of ExoI and SbcCD nucleases is necessary to prevent the degradation of recombinogenic 3′ DNA ends created in a RecBCD-independent manner (8, 23, 38, 45, 46). It was recently shown that the sbcB15 mutant allele (encoding a protein without nucleolytic activity) (37) is a better suppressor of the RecBCD phenotype than an sbcB deletion (50), suggesting that some nonnucleolytic activity of ExoI may also contribute to the efficiency of the RecF pathway (46, 50).ExoI and SbcCD are usually viewed as enzymes with inhibitory roles in recombination due to their deleterious actions on the RecF pathway. However, some results suggest that these enzymes could also have stimulatory roles in recombination reactions proceeding on the RecBCD pathway. Genetic experiments with UV-irradiated E. coli cells indicated that ExoI and SbcCD might be involved in blunting radiation-induced DNA ends prior to RecBC(D) action (38, 45, 46). Such a role of ExoI and SbcCD seems to be particularly critical in recD recF mutants, in which the majority of DSB repair depends on the RecBC enzyme (38). It was also suggested that the blunting roles of the two nucleases may be required during conjugational recombination (16, 46).In this work, we studied the effects of sbcB (xonA) and sbcD mutations on DNA degradation occurring spontaneously in exponentially growing recA mutant cells, as well as on DNA degradation induced in recA mutants by UV irradiation. We have demonstrated that in both cases DNA degradation is strongly reduced in recA mutants that carry in addition a combination of xonA and sbcD null mutations. The results described in this paper suggest that ExoI and SbcCD play partially redundant roles in regulating DNA degradation in recA cells.  相似文献   

2.
Helicobacter pylori infection of the human stomach is associated with disease-causing inflammation that elicits DNA damage in both bacterial and host cells. Bacteria must repair their DNA to persist. The H. pylori AddAB helicase-exonuclease is required for DNA repair and efficient stomach colonization. To dissect the role of each activity in DNA repair and infectivity, we altered the AddA and AddB nuclease (NUC) domains and the AddA helicase (HEL) domain by site-directed mutagenesis. Extracts of Escherichia coli expressing H. pylori addANUCB or addABNUC mutants unwound DNA but had approximately half of the exonuclease activity of wild-type AddAB; the addANUCBNUC double mutant lacked detectable nuclease activity but retained helicase activity. Extracts with AddAHELB lacked detectable helicase and nuclease activity. H. pylori with the single nuclease domain mutations were somewhat less sensitive to the DNA-damaging agent ciprofloxacin than the corresponding deletion mutant, suggesting that residual nuclease activity promotes limited DNA repair. The addANUC and addAHEL mutants colonized the stomach less efficiently than the wild type; addBNUC showed partial attenuation. E. coli ΔrecBCD expressing H. pylori addAB was recombination-deficient unless H. pylori recA was also expressed, suggesting a species-specific interaction between AddAB and RecA and also that H. pylori AddAB participates in both DNA repair and recombination. These results support a role for both the AddAB nuclease and helicase in DNA repair and promoting infectivity.Infection of the stomach with Helicobacter pylori causes a variety of diseases including gastritis, peptic ulcers, and gastric cancer (1). A central feature of the pathology of these conditions is the establishment of a chronic inflammatory response that acts both on the host and the infecting bacteria (2). Both epithelial (3, 4) and lymphoid (5, 6) cells in the gastric mucosa of infected individuals release DNA-damaging agents that can introduce double-stranded (ds)2 breaks into the bacterial chromosome (7). The ds breaks must be repaired for the bacteria to survive and establish chronic colonization of the stomach. Homologous recombination is required for the faithful repair of DNA damage and bacterial survival. Alteration of the expression of one of a series of cell surface proteins on H. pylori occurs by an apparent gene conversion of babA, the frequency of which is reduced in repair-deficient strains (8, 9). This change in the cell surface, which may allow H. pylori to evade the host immune response, is a second means by which recombination can promote efficient colonization of the stomach by H. pylori.The initiation or presynaptic steps of recombination at dsDNA breaks in most bacteria involves the coordinated action of nuclease and helicase activities provided by one of two multisubunit enzymes, the AddAB and RecBCD enzymes (10). Escherichia coli recBCD null mutants have reduced cell viability, are hypersensitive to DNA-damaging agents, and are homologous recombination-deficient (1114). Similarly, H. pylori addA and addB null mutants are hypersensitive to DNA-damaging agents, have reduced frequencies of babA gene conversion, and colonize the stomach of mice less efficiently than wild-type strains (8).The activities of RecBCD enzyme from E. coli (1519) and AddAB from H. pylori (8) or Bacillus subtilis (2023) indicate some common general features of the presynaptic steps of DNA repair. In the case of E. coli, repair begins when the RecBCD enzyme binds to a dsDNA end and unwinds the DNA using its ATP-dependent helicase activities (17, 24). Single-stranded (ss) DNA produced during unwinding, with or without accompanying nuclease, is coated with RecA protein (16, 25). This recombinogenic substrate engages in strand exchange with a homologous intact duplex to form a joint molecule. Joint molecules are thought to be converted into intact, recombinant DNA either by replication or by cutting and ligation of exchanged strands (26).Although the AddAB and RecBCD enzymes appear to play similar roles in promoting recombination and DNA repair, they differ in several ways. RecBCD is a heterotrimer, composed of one copy of the RecB, RecC, and RecD gene products (27), whereas AddAB has two subunits, encoded by the addA and addB genes (21, 28). The enzyme subunit(s) responsible for helicase activity can be inferred from the presence of conserved protein domains or the activity of purified proteins. AddA, RecB, and RecD are superfamily I helicases with six highly conserved helicase motifs, including the conserved Walker A box found in many enzymes that bind ATP (2932). A Walker A box is defined by the consensus sequence (G/A)XXGXGKT (X is any amino acid (29). RecBCD enzymes in which the conserved Lys in this motif is changed to Gln have a reduced affinity for ATP binding (33, 34) and altered helicase activity (17, 3537).A nuclease domain with the conserved amino acid sequence LDYK is found in RecB, AddA, AddB, and many other nucleases (38). The conserved Asp plays a role in Mg2+ binding at the active site; Mg2+ is required for nuclease activity (39). The recB1080 mutation, which changes codon 1080 from the conserved Asp in this motif to Ala, eliminates nuclease activity (39).We have recently shown that addA and addB deletion mutants are hypersensitive to DNA-damaging agents and impaired in colonization of the mouse stomach compared with wild-type strains (8). To determine the roles of the individual helicase and nuclease activities of H. pylori AddAB in DNA repair and infectivity, we used site-directed mutagenesis to inactivate the conserved nuclease domains of addA and addB and the conserved ATPase (helicase) domain of AddA. Here, we report that loss of the AddAB helicase is sufficient to impair H. pylori DNA repair and infectivity and, when the genes are expressed in E. coli, homologous recombination. AddAB retains partial activity in biochemical and genetic assays when either of the two nuclease domains is inactivated but loses all detectable nuclease activity when both domains are inactivated. Remarkably, H. pylori AddAB can produce recombinants in E. coli only in the presence of H. pylori RecA, suggesting a species-specific interaction in which AddAB facilitates the production of ssDNA-coated with RecA protein. Our results show that both the helicase and nuclease activities are required for the biological roles of H. pylori AddAB.  相似文献   

3.
4.
5.
6.
7.
8.
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   

9.
The herpes simplex virus (HSV) virion host shutoff protein (vhs) encoded by gene UL41 is an mRNA-specific RNase that triggers accelerated degradation of host and viral mRNAs in infected cells. We report here that vhs is also able to modulate reporter gene expression without greatly altering the levels of the target mRNA in transient-transfection assays conducted in HeLa cells. We monitored the effects of vhs on a panel of bicistronic reporter constructs bearing a variety of internal ribosome entry sites (IRESs) located between two test cistrons. As expected, vhs inhibited the expression of the 5′ cistrons of all of these constructs; however, the response of the 3′ cistron varied with the IRES: expression driven from the wild-type EMCV IRES was strongly suppressed, while expression controlled by a mutant EMCV IRES and the cellular ApaF1, BiP, and DAP5 IRES elements was strongly activated. In addition, several HSV type 1 (HSV-1) 5′ untranslated region (5′ UTR) sequences also served as positive vhs response elements in this assay. IRES activation was also observed in 293 and HepG2 cells, but no such response was observed in Vero cells. Mutational analysis has yet to uncouple the ability of vhs to activate 3′ cistron expression from its shutoff activity. Remarkably, repression of 5′ cistron expression could be observed under conditions where the levels of the reporter RNA were not correspondingly reduced. These data provide strong evidence that vhs can modulate gene expression at the level of translation and that it is able to activate cap-independent translation through specific cis-acting elements.The virion host shutoff protein (vhs) encoded by herpes simplex virus (HSV) gene UL41 is an endoribonuclease that is packaged into the tegument of mature HSV virions. Once delivered into the cytoplasm of newly infected cells, vhs triggers shutoff of host protein synthesis, disruption of preexisting polysomes, and degradation of host mRNAs (reviewed in reference 62). The vhs-dependent shutoff system destabilizes many cellular and viral mRNAs (36, 46, 67). The rapid decline in host mRNA levels presumably helps viral mRNAs gain access to the cellular translational apparatus. In addition, the relatively short half-lives of viral mRNAs contribute to the sharp transitions between the successive phases of viral protein synthesis by tightly coupling changes in the rates of synthesis of viral mRNAs to altered mRNA levels (46). These effects enhance virus replication and may account for the modest reduction in virus yield displayed by vhs mutants in cultured Vero cells (55, 61).vhs also plays a critical role in HSV pathogenesis: vhs mutants are severely impaired for replication in the corneas and central nervous systems of mice and cannot efficiently establish or reactivate from latency (63, 65, 66). Mounting evidence indicates that this attenuation stems at least in part from an impaired ability to disarm elements of the innate and adaptive host immune responses (reviewed in reference 62). For example, vhs suppresses certain innate cellular antiviral responses, including production of proinflammatory cytokines and chemokines (68); dampens the type I interferon system (11, 45, 49, 78); and blocks activation of dendritic cells (58). Moreover, vhs mutants display enhanced virulence in knockout mice lacking type I interferon (IFN) receptors (37, 45) or Stat1 (48) and are hypersensitive to the antiviral effects of IFN in some cells in tissue culture (11, 49, 68). Thus, vhs is arguably a bona fide virulence factor.vhs present in extracts of HSV virions or purified from bacteria has nonspecific RNase activity capable of degrading all RNA substrates (15, 70, 71, 79). However, vhs is highly selective in vivo, targeting mRNAs and sparing other cytoplasmic RNAs (36, 46). In vivo and in mammalian whole-cell extracts, vhs-induced decay of at least some mRNAs initiates near regions of translation initiation and proceeds in an overall 5′-to-3′ direction (12, 13, 29, 52). Moreover, vhs binds to the translation initiation factors eIF4H, eIF4B, and eIF4A II, all components of the cap recognition factor eIF4F (10, 16, 17). Thus, it has been proposed that vhs selectively targets actively translated mRNAs through interactions with eIF4F components (17). Consistent with this hypothesis, recent data document that eIF4H is required for vhs activity in vivo (59).A previous report from this laboratory documented that the internal ribosome entry sites (IRESs) of the picornaviruses poliovirus and encephalomyocarditis virus (EMCV) strongly target vhs-induced RNA cleavage events to sequences immediately 3′ to the IRES in an in vitro translation system derived from rabbit reticulocyte lysates (RRL) (13). IRES elements are highly structured RNA sequences that are able to direct cap-independent translational initiation (reviewed in references 21, 25, 30, and 64). In the case of the poliovirus and EMCV elements, this is achieved by directly recruiting the eIF4F scaffolding protein eIF4G, thus bypassing the requirement for the cap-binding eIF4F subunit, eIF4E (reviewed in reference 30). Based on these data, we suggested that vhs is strongly targeted to the picornavirus IRES elements via interactions with eIF4 factors.A growing number of cellular mRNAs have been proposed to bear IRES elements in their 5′ untranslated regions (5′ UTRs). These include many that are involved in cellular stress responses, apoptosis, and cell cycle progression (24, 64, 74). Given the striking ability of picornavirus IRES elements to target vhs RNase activity in vitro, we asked whether viral and cellular IRES elements are able to modify the susceptibility of mRNAs to vhs in vivo. During the course of preliminary experiments designed to test this hypothesis, we unexpectedly discovered that vhs is able to strongly activate gene expression controlled by some cellular IRES elements and HSV 5′ UTR sequences in in vivo bicistronic reporter assays. These observations are the subject of the present report.  相似文献   

10.
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   

11.
12.
13.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

14.
15.
16.
17.
18.
Porcine circovirus type 1 (PCV1), originally isolated as a contaminant of PK-15 cells, is nonpathogenic, whereas porcine circovirus type 2 (PCV2) causes an economically important disease in pigs. To determine the factors affecting virus replication, we constructed chimeric viruses by swapping open reading frame 1 (ORF1) (rep) or the origin of replication (Ori) between PCV1 and PCV2 and compared the replication efficiencies of the chimeric viruses in PK-15 cells. The results showed that the replication factors of PCV1 and PCV2 are fully exchangeable and, most importantly, that both the Ori and rep of PCV1 enhance the virus replication efficiencies of the chimeric viruses with the PCV2 backbone.Porcine circovirus (PCV) is a single-stranded DNA virus in the family Circoviridae (34). Type 1 PCV (PCV1) was discovered in 1974 as a contaminant of porcine kidney cell line PK-15 and is nonpathogenic in pigs (31-33). Type 2 PCV (PCV2) was discovered in piglets with postweaning multisystemic wasting syndrome (PMWS) in the mid-1990s and causes porcine circovirus-associated disease (PCVAD) (1, 9, 10, 25). PCV1 and PCV2 have similar genomic organizations, with two major ambisense open reading frames (ORFs) (16). ORF1 (rep) encodes two viral replication-associated proteins, Rep and Rep′, by differential splicing (4, 6, 21, 22). The Rep and Rep′ proteins bind to specific sequences within the origin of replication (Ori) located in the intergenic region, and both are responsible for viral replication (5, 7, 8, 21, 23, 28, 29). ORF2 (cap) encodes the immunogenic capsid protein (Cap) (26). PCV1 and PCV2 share approximately 80%, 82%, and 62% nucleotide sequence identity in the Ori, rep, and cap, respectively (19).In vitro studies using a reporter gene-based assay system showed that the replication factors of PCV1 and PCV2 are functionally interchangeable (2-6, 22), although this finding has not yet been validated in a live infectious-virus system. We have previously shown that chimeras of PCV in which cap has been exchanged between PCV1 and PCV2 are infectious both in vitro and in vivo (15), and an inactivated vaccine based on the PCV1-PCV2 cap (PCV1-cap2) chimera is used in the vaccination program against PCVAD (13, 15, 18, 27).PCV1 replicates more efficiently than PCV2 in PK-15 cells (14, 15); thus, we hypothesized that the Ori or rep is directly responsible for the differences in replication efficiencies. The objectives of this study were to demonstrate that the Ori and rep are interchangeable between PCV1 and PCV2 in a live-virus system and to determine the effects of swapped heterologous replication factors on virus replication efficiency in vitro.  相似文献   

19.
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