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 865 毫秒
1.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

2.
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   

3.
4.
The hepatitis C virus (HCV) RNA-dependent RNA polymerase (RdRp) has been proposed to change conformations in association with RNA synthesis and to interact with cellular proteins. In vitro, the RdRp can initiate de novo from the ends of single-stranded RNA or extend a primed RNA template. The interactions between the Δ1 loop and thumb domain in NS5B are required for de novo initiation, although it is unclear whether these interactions are within an NS5B monomer or are part of a higher-order NS5B oligomeric complex. This work seeks to address how polymerase conformation and/or oligomerization affects de novo initiation. We have shown that an increasing enzyme concentration increases de novo initiation by the genotype 1b and 2a RdRps while primer extension reactions are not affected or inhibited under similar conditions. Initiation-defective mutants of the HCV polymerase can increase de novo initiation by the wild-type (WT) polymerase. GTP was also found to stimulate de novo initiation. Our results support a model in which the de novo initiation-competent conformation of the RdRp is stimulated by oligomeric contacts between individual subunits. Using electron microscopy and single-molecule reconstruction, we attempted to visualize the low-resolution conformations of a dimer of a de novo initiation-competent HCV RdRp.Polymerases undergo a series of conformational changes at different stages of nucleic acid synthesis (14). Of the template-dependent polymerases, the RNA-dependent RNA polymerases (RdRps) are the least understood in terms of their mechanism of action. RdRps are of increasing interest since cellular RdRps play important roles in the defense against nonself RNAs (44). In addition, virus-encoded RdRps are important targets for the development of antivirals. A better understanding of RNA-dependent RNA polymerases is thus important for both basic and applied science.Several model systems for biochemical study of viral RNA-dependent RNA synthesis exist (4, 19, 20, 25, 37, 42). Well-characterized RdRps include those from the hepatitis C virus (HCV) and poliovirus (5, 17). In the host, the RdRps are complexed with other viral and/or cellular proteins that are usually associated with membranous intracellular structures. The replicases are usually difficult to study biochemically, but the catalytic RdRp subunits of several viruses can be purified for functional and structural analyses (53). These recombinant proteins can reproduce some of the activities of the replicases, including the ability to initiate RNA synthesis by a de novo mechanism (22, 47-49). Furthermore, recombinant RdRps can affect the activities of other replicase subunits in vitro, suggesting that the recombinant RdRp is useful for an in-depth understanding of RNA synthesis by HCV (45, 60).RdRps form a right-hand-like structure with thumb, finger, and palm subdomains. The metal-coordinating residues important for nucleotide binding are positioned within the palm subdomain (26). An interesting feature of viral RdRps is that they tend to exist in a closed conformation, even in the absence of template, in contrast to DNA-dependent RNA polymerases, which transition from open to closed complexes upon template recognition (13). The closed form of the phage φ6 RdRp has been proposed to allow specific recognition of the single-stranded viral RNA (7). The template channel formed by the closed structure, however, is too narrow to accommodate the partially duplexed RNA that forms during RNA synthesis, and hence, the closed conformation needs to undergo significant rearrangements in the ternary complex. Biswal et al. (3) have captured an X-ray crystallographic structure of a partially open conformation of the HCV RdRp. Bovine viral diarrhea virus (BVDV) RdRp was also shown to exist in a partially open conformation (11). Ranjith-Kumar and Kao (49) demonstrated that the HCV RdRp could initiate RNA synthesis from a circular RNA template, and thus, the threading of a single-stranded RNA into the template channel is not required for de novo-initiated RNA synthesis. Altogether, these results raise the possibility that the HCV RdRp can undergo rearrangements from the closed conformation seen in the crystal structure prior to de novo initiation.A secondary structure that extends from the finger to the thumb subdomains, named the Δ1 loop, has been proposed to serve as a gate to cover the template channel and regulate the switch from de novo initiation to elongation (5, 10). Mutations that affect the interaction between the Δ1 loop and the mostly hydrophobic residues that it contacts have resulted in polymerases that are defective for de novo initiation but can bind to partially duplexed RNA and can extend from the 3′ terminus of an RNA primer (10).Two general models for RNA synthesis by the HCV RdRp can be proposed (Fig. (Fig.1).1). The first posits that the HCV RdRp functions as a monomer at least during de novo initiation because the closed template channel is needed for specific recognition of the template (5, 7, 10). It was presumed that the Δ1 loop and thumb domain interaction in the HCV RdRp is stable and mutations that disrupted this interaction would render the enzyme catalytically inactive (5, 24). However, a deletion of five residues in the tip of the Δ1 loop did not prevent RNA synthesis from a primed template by the polymerase (10). Furthermore, a genotype 2a RdRp was crystallized in a form with altered interaction between the Δ1 loop and thumb domain in comparison to the 1b RdRp (3). Interestingly, a low-affinity GTP binding site exists on the thumb domain close to the base of the Δ1 loop binding pocket. GTP binding at this site has been proposed to stabilize the Δ1 loop and thumb domain interactions, favoring the closed monomer model (6). A second model is based on the reports that HCV RdRp can oligomerize and that oligomerization increases its activity (12, 16, 46, 54). The dimer could be active due to either the second subunit increasing the stability of the Δ1 loop and thumb interactions in the first subunit to increase de novo initiation or the two subunits forming a common template-binding domain (Fig. (Fig.1).1). Here we have attempted to determine whether monomers or oligomers of the HCV RdRp can better perform de novo initiation using biochemical and biophysical analyses.Open in a separate windowFIG. 1.Models for RNA synthesis by the HCV RdRp. The monomer model is based on the central tenet that intramolecular interactions within an RdRp molecule regulate the modes of RNA synthesis. The curved arrow represents the possible orientation of the template RNA. The oligomer model is an adaptation from the dimer model of the norovirus RdRp (18). T, P, and F represent the thumb, palm, and finger domains, respectively, in different shades of gray, and the thick black line connecting the thumb and finger domains represents the Δ1 loop.  相似文献   

5.
In the nitrate-responsive, homodimeric NarX sensor, two cytoplasmic membrane α-helices delimit the periplasmic ligand-binding domain. The HAMP domain, a four-helix parallel coiled-coil built from two α-helices (HD1 and HD2), immediately follows the second transmembrane helix. Previous computational studies identified a likely coiled-coil-forming α-helix, the signaling helix (S helix), in a range of signaling proteins, including eucaryal receptor guanylyl cyclases, but its function remains obscure. In NarX, the HAMP HD2 and S-helix regions overlap and apparently form a continuous coiled-coil marked by a heptad repeat stutter discontinuity at the distal boundary of HD2. Similar composite HD2-S-helix elements are present in other sensors, such as Sln1p from Saccharomyces cerevisiae. We constructed deletions and missense substitutions in the NarX S helix. Most caused constitutive signaling phenotypes. However, strongly impaired induction phenotypes were conferred by heptad deletions within the S-helix conserved core and also by deletions that remove the heptad stutter. The latter observation illuminates a key element of the dynamic bundle hypothesis for signaling across the heptad stutter adjacent to the HAMP domain in methyl-accepting chemotaxis proteins (Q. Zhou, P. Ames, and J. S. Parkinson, Mol. Microbiol. 73:801-814, 2009). Sequence comparisons identified other examples of heptad stutters between a HAMP domain and a contiguous coiled-coil-like heptad repeat sequence in conventional sensors, such as CpxA, EnvZ, PhoQ, and QseC; other S-helix-containing sensors, such as BarA and TorS; and the Neurospora crassa Nik-1 (Os-1) sensor that contains a tandem array of alternating HAMP and HAMP-like elements. Therefore, stutter elements may be broadly important for HAMP function.Transmembrane signaling in homodimeric bacterial sensors initiates upon signal ligand binding to the extracytoplasmic domain. In methyl-accepting chemotaxis proteins (MCPs), the resulting conformational change causes a displacement of one transmembrane α-helix (TM α-helix) relative to the other. This motion is conducted by the HAMP domain to control output domain activity (reviewed in references 33 and 39).Certain sensors of two-component regulatory systems share topological organization with MCPs. For example, the paralogous nitrate sensors NarX and NarQ contain an amino-terminal transmembrane signaling module similar to those in MCPs, in which a pair of TM α-helices delimit the periplasmic ligand-binding domain (Fig. (Fig.1)1) (24) (reviewed in references 32 and 62). The second TM α-helix connects to the HAMP domain. Hybrid proteins in which the NarX transmembrane signaling module regulates the kinase control modules of the MCPs Tar, DifA, and FrzCD demonstrate that NarX and MCPs share a mechanism for transmembrane signaling (73, 74, 81, 82).Open in a separate windowFIG. 1.NarX modular structure. Linear representation of the NarX protein sequence, from the amino (N) to carboxyl (C) termini, drawn to scale. The four modules are indicated at the top of the figure and shown in bold typeface, whereas domains within each module are labeled with standard (lightface) typeface. The nomenclature for modules follows that devised by Swain and Falke (67) for MCPs. Overlap between the HAMP domain HD2 and S-helix elements is indicated in gray. The three conserved Cys residues within the central module (62) are indicated. TM1 and TM2 denote the two transmembrane helices. Helices H1 to H4 of the periplasmic domain (24), and the transmitter domain H, N, D, G (79), and X (41) boxes, are labeled. The HPK 7 family of transmitter sequences, including NarX, have no F box and an unconventional G box (79). The scale bar at the bottom of the figure shows the number of aminoacyl residues.The HAMP domain functions as a signal conversion module in a variety of homodimeric proteins, including histidine protein kinases, adenylyl cyclases, MCPs, and certain phosphatases (12, 20, 77). This roughly 50-residue domain consists of a pair of amphiphilic α-helices, termed HD1 and HD2 (formerly AS1 and AS2) (67), joined by a connector (Fig. (Fig.2A).2A). Results from nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR) spectroscopy, Cys and disulfide scanning, and mutational analysis converge on a model in which the HD1 and HD2 α-helices form a four-helix parallel coiled-coil (7, 20, 30, 42, 67, 75, 84). The mechanisms through which HAMP domains mediate signal conduction remain to be established (30, 42, 67, 84) (for commentary, see references 43, 49, and 50).Open in a separate windowFIG. 2.HAMP domain extensions. (A) Sequences from representative MCPs (E. coli Tsr and Salmonella enterica serovar Typhimurium Tar) and S-helix-containing sensors (E. coli NarX, NarQ, and BarA, and S. cerevisiae Sln1p). The HAMP domain, S-helix element, and the initial sequence of the MCP adaptation region are indicated. Flanking numbers denote positions of the terminal residues within the overall sequence. Sequential heptad repeats are indicated in alternating bold and standard (lightface) typeface. Numbering for heptad repeats in the methylation region and S-helix sequences has been described previously (4, 8). Numbers within the HD1 and HD2 helices indicate interactions within the HAMP domain (42). Residues at heptad positions a and d are enclosed within boxes, residues at the stutter position a/d are enclosed within a thickly outlined box, and residues in the S-helix ERT signature are in bold typeface. (B) NarX mutational alterations. Deletions are depicted as boxes, and missense substitutions are shown above the sequence. Many of these deletions were reported previously (10) and are presented here for comparison. The phenotypes conferred by the alterations are indicated as follows: impaired induction, black box; constitutive and elevated basal, light gray box; reversed response, dark gray box; wild-type, white box; null, striped box.Coiled-coils result from packing of two or more α-helices (27). The primary sequence of coiled-coils exhibits a characteristic heptad repeat pattern, denoted as a-b-c-d-e-f-g (52, 61), in which positions a and d are usually occupied by nonpolar residues (reviewed in references 1, 47, and 80). For example, the coiled-coil nature of the HAMP domain can be seen in the heptad repeat patterns within the HD1 and HD2 sequences (Fig. (Fig.2A2A).Coiled-coil elements adjacent to the HAMP domain have been identified in several sensors, including Saccharomyces cerevisiae Sln1p (69) and Escherichia coli NarX (60). Recently, this element was defined as a specific type of dimeric parallel coiled-coil, termed the signaling helix (S helix), present in a wide range of signaling proteins (8). Sequence comparisons delimit a roughly 40-residue element with a conserved heptad repeat pattern (Fig. (Fig.2A).2A). Based on mutational analyses of Sln1p and other proteins, the S helix is suggested to function as a switch that prevents constitutive activation of adjacent output domains (8).The term “signaling helix” previously was used to define the α4-TM2 extended helix in MCPs (23, 33). Here, we use the term S helix to denote the element described by Anantharaman et al. (8).The NarX and NarQ sensors encompass four distinct modules (Fig. (Fig.1):1): the amino-terminal transmembrane signaling module, the signal conversion module (including the HAMP domain and S-helix element), the central module of unknown function, and the carboxyl-terminal transmitter module (62). The S-helix element presumably functions together with the HAMP domain in conducting ligand-responsive motions from the transmembrane signaling module to the central module, ultimately regulating transmitter module activity.Regulatory output by two-component sensors reflects opposing transmitter activities (reviewed in reference 55). Positive function results from transmitter autokinase activity; the resulting phosphosensor serves as a substrate for response regulator autophosphorylation. Negative function results from transmitter phosphatase activity, which accelerates phosphoresponse regulator autodephosphorylation (reviewed in references 64 and 65). We envision a homogeneous two-state model for NarX (17), in which the equilibrium between these mutually exclusive conformations is modulated by ligand-responsive signaling.Previous work from our laboratory concerned the NarX and other HAMP domains (9, 10, 26, 77) and separately identified a conserved sequence in NarX and NarQ sensors, the Y box, that roughly corresponds to the S helix (62). Therefore, we were interested to explore the NarX S helix and to test some of the predictions made for its function. Results show that the S helix is critical for signal conduction and suggest that it functions as an extension of the HAMP HD2 α-helix in a subset of sensors exemplified by Sln1p and NarX. Moreover, a stutter discontinuity in the heptad repeat pattern was found to be essential for the NarX response to signal and to be conserved in several distinct classes of HAMP-containing sensors.  相似文献   

6.
7.
8.
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   

9.
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   

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

12.
13.
14.
Phosphoprotein ppUL44 of the human cytomegalovirus (HCMV) DNA polymerase plays an essential role in viral replication, conferring processivity to the DNA polymerase catalytic subunit pUL54 by tethering it to the DNA. Here, for the first time, we examine in living cells the function of the highly flexible loop of ppUL44 (UL44-FL; residues 162 to 174 [PHTRVKRNVKKAP174]), which has been proposed to be directly involved in ppUL44''s interaction with DNA. In particular, we use a variety of approaches in transfected cells to characterize in detail the behavior of ppUL44Δloop, a mutant derivative in which three of the five basic residues within UL44-FL are replaced by nonbasic amino acids. Our results indicate that ppUL44Δloop is functional in dimerization and binding to pUL54 but strongly impaired in binding nuclear structures within the nucleus, as shown by its inability to form nuclear speckles, reduced nuclear accumulation, and increased intranuclear mobility compared to wild-type ppUL44. Moreover, analysis of cellular fractions after detergent and DNase treatment indicates that ppUL44Δloop is strongly reduced in DNA-binding ability, in similar fashion to ppUL44-L86A/L87A, a point mutant derivative impaired in dimerization. Finally, ppUL44Δloop fails to transcomplement HCMV oriLyt-dependent DNA replication in cells and also inhibits replication in the presence of wild-type ppUL44, possibly via formation of heterodimers defective for double-stranded DNA binding. UL44-FL thus emerges for the first time as an important determinant for HCMV replication in cells, with potential implications for the development of novel antiviral approaches by targeting HCMV replication.The Betaherpesviridae subfamily member human cytomegalovirus (HCMV) is a major human pathogen, causing serious disease in newborns following congenital infection and in immunocompromised individuals (28, 42). Replication of its double-stranded DNA (dsDNA) genome occurs in the nuclei of infected cells via a rolling-circle process mediated by 11 virally encoded proteins (32, 33), including a viral DNA polymerase holoenzyme, comprising a catalytic subunit, pUL54, and a proposed processivity factor, ppUL44 (14). ppUL44 is readily detectable in virus-infected cells as a 52-kDa phosphoprotein of 433 amino acids with strong dsDNA-binding ability (30, 45). Defined as a “polymerase accessory protein” (PAP) whose function is highly conserved among herpesviruses, ppUL44 is an essential factor for viral replication in cultured cells and hence represents a potential therapeutic target to combat HCMV infection (39). It is a multifunctional protein capable of self-associating (5, 10), as well as interacting with a plethora of viral and host cell proteins, including the viral kinase pUL97 (29), the viral transactivating protein pUL84 (15), the viral uracil DNA glycosylase ppUL114 (37), and the host cell importin α/β (IMPα/β) heterodimer, which is responsible for its transport into the nucleus (4). The activities of ppUL44 as a processivity factor, including the ability to dimerize, as well as bind to, pUL54 and DNA, reside in the N-terminal portion (26, 45), whereas the C terminus is essential for phosphorylation-regulated, IMPα/β-dependent nuclear targeting of ppUL44 monomers and dimers (4-6). Once within the nucleus, ppUL44 is thought to tether the DNA polymerase holoenzyme to the DNA, thus increasing its processivity (14).Recent studies have identified specific residues responsible for ppUL44 interaction with pUL54, as well as for the interaction with IMPα/β and homodimerization (4, 10, 27, 41). The crystal structure of ppUL44''s N-terminal domain (Fig. (Fig.1A)1A) reveals striking similarity to that of other processivity factors, such as proliferating cell nuclear antigen (PCNA) and its herpes simplex virus type 1 (HSV-1) homologue UL42 (10, 46). Unlike the PCNA trimeric ring, however, both ppUL44 and UL42, which bind to dsDNA as dimers and monomers, respectively, have an open structure, which is believed to be the basis for their ability to bind to dsDNA in the absence of clamp loaders and ATP (9, 10, 46). Both ppUL44 and UL42 share a very basic “back” face, which appears to be directly involved in DNA binding via electrostatic interactions (19, 22, 23, 38, 46). One striking difference between ppUL44 and UL42 is the presence on the former of an extremely basic flexible loop (UL44-FL, PHTRVKRNVKKAP174) protruding from the basic back face of the protein (Fig. (Fig.1A).1A). Comparison of ppUL44 homologues from different betaherpesviruses, including human herpesvirus 6 (HHV-6) and 7 (HHV-7), showed that all possess similar sequences in the same position (44) (Fig. (Fig.1B),1B), implying functional significance.Open in a separate windowFIG. 1.The highly conserved flexible loop (residues 162 to 174) within ppUL44 protrudes from ppUL44 basic face and is important for efficient nuclear accumulation and localization in nuclear speckles. (A) Schematic representation of ppUL44 N-terminal domain (residues 9 to 270, protein data bank accession no. 1T6L) generated using the Chimera software based on the published crystal structure (10, 35). Color: yellow, β-sheets; red, α-helices. Residues involved in ppUL44 dimerization (P85, L86, L87, L93, F121, and M123), as well as basic residues potentially involved in DNA binding (K21, R28, K32, K35, K128, K158, K224, and K237), are represented as spacefill in orange and green, respectively. Residues P162 and C175, in black, are indicated by arrowheads, while residues 163 to 174 are not visible in the electron density maps and could potentially extend in the cavity formed by ppUL44''s basic face to directly contact DNA. Residues forming ppUL44 connector loop (128-142) are in blue. (B) Sequence alignment between HCMVUL44-FL and the corresponding region of several betaherpesvirus ppUL44 homologues. The single-letter amino acid code is used, with basic residues in boldface. (C) COS-7 cells were transfected to express the indicated GFP fusion proteins and imaged live 16 h after transfection using CLSM and a 40× water immersion objective lens. (D) Quantitative results for the Fn/c and speckle formation for GFP-UL44 fusion proteins. The data for the Fn/c ratios represent the mean Fn/c relative to each protein indicated as a percentage of the mean Fn/c relative to GFP-UL44wt ± the standard error of the mean, with the number of analyzed cells in parentheses. (E) HEK 293 cells expressing the indicated GFP-UL44 fusion proteins were lysed, separated by PAGE, and analyzed by Western blotting as described in Materials and Methods, using either the anti-GFP or the anti-α-tubulin MAbs.A recent study revealed that substitution of UL44-FL basic residues with alanine residues strongly impairs the ability of a bacterially expressed N-terminal fragment of UL44 to bind 30-bp dsDNA oligonucleotides in vitro, suggesting that UL44-FL could be involved in dsDNA-binding during viral replication (22). However, the role of UL44-FL in mediating the binding of full-length UL44 to dsDNA in cells and its role in DNA replication have not been investigated. We use here a variety of approaches to delineate the role of UL44-FL in living cells, our data revealing that UL44-FL is not required for ppUL44 dimerization or binding to the catalytic subunit pUL54 but is crucial for HCMV oriLyt-dependent DNA replication, being required for the formation of nuclear aggregates, nuclear accumulation/retention, and DNA binding of ppUL44. Importantly, ppUL44Δloop exhibits a transdominant-negative phenotype, inhibiting HCMV oriLyt-dependent DNA replication in the presence of wild-type ppUL44, possibly via formation of heterodimers defective for dsDNA binding. This underlines ppUL44-FL as an important determinant for HCMV replication in a cellular context for the first time, with potential implications for the development of novel antiviral approaches.  相似文献   

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

18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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