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 46 毫秒
1.
2.
3.
The hepatitis B virus (HBV) envelope proteins bear two determinants of viral entry: a receptor-binding site (RBS) in the pre-S1 domain of the large envelope protein and a conformation-dependent determinant, of unknown function, in the antigenic loop (AGL) of the small, middle, and large envelope proteins. Using an in vitro infection assay consisting of susceptible HepaRG cells and the hepatitis delta virus (HDV) as a surrogate of HBV, we first investigated whether subelements of the pre-S1 determinant (amino acids 2 to 75), i.e., the N-terminal myristoyl anchor, subdomain 2-48 (RBS), and subdomain 49-75, were functionally separable. In transcomplementation experiments, coexpression of two distinct infectivity-deficient pre-S1 mutants at the surface of HDV virions failed to restore infectivity, indicating that the myristoyl anchor, the 2-48 RBS, and the 49-75 sequence, likely cooperate in cis at viral entry. Furthermore, we showed that as much as 52% of total pre-S1 in the HDV envelope could bear infectivity-deficient lesions without affecting entry, indicating that a small number of pre-S1 polypeptides—estimated at three to four per virion—is sufficient for infectivity. We next investigated the AGL activity in the small or large envelope protein background (S- and L-AGL, respectively) and found that lesions in S-AGL were more deleterious to infectivity than in L-AGL, a difference that reflects the relative stoichiometry of the small and large envelope proteins in the viral envelope. Finally, we showed that C147S, an AGL infectivity-deficient substitution, exerted a dominant-negative effect on infectivity, likely reflecting an involvement of C147 in intermolecular disulfide bonds.Hepatitis B virus (HBV) remains a major public health concern worldwide, affecting more than 350 millions of chronically infected individuals. Since the discovery of HBV, substantial information has been gathered on the viral replication cycle, but our understanding of the viral entry mechanism remains limited, and the identity of the receptor(s) for HBV is still unknown (15). HBV displays a very narrow host range, which is likely determined at viral entry by a highly specific interaction between the HBV envelope proteins and receptors at the surface of human hepatocytes. The envelope proteins designated large (L-HBsAg), middle (M-HBsAg), and small (S-HBsAg) are membrane-spanning glycoproteins that differ from each other by the size of their N-terminal ectodomain (21). L-HBsAg contains a N-terminal pre-S1, central pre-S2, and C-terminal S domains. M-HBsAg is shorter than L-HBsAg in lacking pre-S1, whereas S-HBsAg consists of the S domain only (Fig. (Fig.1).1). Envelope protein synthesis occurs at the endoplasmic reticulum (ER) membrane. Empty subviral particles (SVPs) assemble from aggregates at a pre-Golgi membrane and exit the cell through the secretory pathway (36). Assembly of mature HBV virions requires, in addition to S-HBsAg, the presence of L-HBsAg as a matrix protein for nucleocapsid envelopment (6). Recent findings indicate that HBV virions and SVPs follow distinct pathways for budding: the late endosomal multivesicular bodies (MVBs) for HBV virions, and the MVB-independent secretory pathway for SVPs (26, 28, 46). The HBV envelope proteins can also package the hepatitis delta virus (HDV) ribonucleoprotein (RNP), in case of HBV/HDV coinfection (5, 45), leading to the formation of HDV virions. Whether HDV uses the SVP secretion pathway rather than an MVB-dependent route is uncertain.Open in a separate windowFIG. 1.Schematic representation of HBV envelope proteins. The topology of the L-, M-, and S-HBsAg proteins at the viral membrane is represented. The pre-S2 domain of L- and M-HBsAg, and the determinants of viral entry, pre-S1 and AGL, are indicated. The M-HBsAg protein, represented in gray, is dispensable for infectivity. The myristic acid (Myr) linked to the L-HBsAg N terminus is indicated (closed box). Subdomains 2-48 and 49-75 of the pre-S1 infectivity determinant are indicated. Open boxes represent transmembrane regions in the S domain.L-HBsAg, but not M-HBsAg, is crucial to infectivity of both HBV and HDV particles (13, 31, 41, 42). L-HBsAg contains a major infectivity determinant located between amino acid residues 2 and 75 of its N-terminal pre-S1 domain (4, 30), including a myristoyl anchor linked to glycine-2 (1, 8, 18), a putative receptor binding site (RBS) between positions 2 and 48, and a domain of unknown function between amino acids 49 and 75. To date, the most compelling evidence that pre-S1 mediates receptor binding comes from studies demonstrating that myristoylated synthetic peptides specific for the N-terminal 2-to-48 pre-S1 domain can bind to hepatocyte plasma membranes and block infection in vitro (3, 16, 17) and in vivo (37). Beside pre-S1, a second determinant was recently identified in the antigenic loop (AGL) borne by the three HBV envelope proteins (Fig. (Fig.1).1). The AGL participation in viral entry was first established in the HDV model (23) and more recently directly in the HBV model (39). Interestingly, serine substitutions for the AGL cysteine residues, which prove detrimental to the conserved immunodominant “a” determinant, could also block viral entry. Note that the “a” determinant consists in conformational epitopes, which elicit highly neutralizing antibodies (22). Infectivity and the “a” determinant were also lost when virions were treated with membrane-impermeable inhibitors of thiol/disulfide isomerization (2). These findings clearly established a correlation between the AGL cysteine disulfide bonds network, the conformation of the “a” determinant, and infectivity. Hence, the strict conservation of the “a” determinant among all HBV genotypes is related to the AGL function at viral entry. The AGL determinant may operate in association with, or independently of pre-S1, in binding to receptors at the early step of entry and/or in the mechanism of envelope disassembly postentry.In the present study, we investigated the pre-S1 determinant by performing transcomplementation experiments between mutants of 3 pre-S1 subelements: the myristoyl anchor, subdomain 2-48, and subdomain 49-75. We analyzed the activity of the AGL determinant in the S- or L-HBsAg background (S- and L-AGL, respectively), and we examined the effect of introducing increasing amounts of infectivity-deficient pre-S1, or AGL, in the virion''s envelope on infectivity.  相似文献   

4.
The catalytic subunit of herpes simplex virus DNA polymerase (Pol), a member of the B family polymerases, possesses both polymerase and exonuclease activities. We previously demonstrated that a recombinant virus (YD12) containing a double mutation within conserved exonuclease motif III of the Pol was highly mutagenic and rapidly evolved to contain an additional leucine-to-phenylalanine mutation at residue 774 (L774F), which is located within the finger subdomain of the polymerase domain. We further demonstrated that the recombinant L774F virus replicated DNA with increased fidelity and that the L774F mutant Pol exhibited altered enzyme kinetics and impaired polymerase activity to extension from mismatched primer termini. In this study, we demonstrated that addition of the L774F mutation to the YD12 Pol did not restore the exonuclease deficiency. However, the polymerase activity of the YD12 Pol to extension from mismatched primer termini and on the nucleotide incorporation pattern was altered upon addition of the L774F mutation. The L774F mutation-containing YD12 Pol also supported the growth of viral progeny and replicated DNA more efficiently and more accurately than did the YD12 Pol. Together, these studies demonstrate that a herpes simplex virus Pol mutant with a highly mutagenic ability can rapidly acquire additional mutations, which may be selected for their survival and outgrowth. Furthermore, the studies demonstrate that the polymerase activity of HSV-1 Pol on primer extension is influenced by sequence context and that herpes simplex virus type 1 Pol may dissociate more frequently at G·C sites during the polymerization reaction. The implications of the findings are discussed.Herpes simplex virus (HSV) DNA polymerase consists of the catalytic subunit of the polymerase (Pol) and the processivity factor UL42. The Pol subunit contains three well-defined activities: polymerization (replication), exonuclease proofreading (editing), and UL42 binding (5, 6, 28). The UL42 binding activity is mediated by amino acid residues located at the C terminus (5, 6). Although the UL42 binding residues are unique to certain alphaherpesvirus DNA polymerases, the sequences comprising the polymerase and exonuclease domains are conserved among the B family (or the α-like) polymerases (2-4). The exonuclease domain of the HSV type 1 (HSV-1) Pol contains conserved exonuclease I (Exo I), II, and III motifs, whereas the polymerase domain contains seven conserved regions (I to VII); conserved region IV overlaps with the Exo II motif. The Exo III motif is located within the δ region C, which is highly conserved among the B family polymerases (Fig. (Fig.1A).1A). These conserved regions are located within the palm, the thumb, and the finger subdomains, which comprise the structural components of the polymerase domain. The crystal structure of the HSV-1 Pol subunit revealed three grooves that form the putative polymerase, exonuclease, and DNA binding sites. The putative exonuclease site is defined as a groove formed between the exonuclease domain and the tip of the thumb subdomain. The palm and thumb subdomains form a groove proposed to be the putative duplex DNA binding site for both the editing and the polymerization complexes (23). Thus, the polymerase and exonuclease domains of HSV-1 are structurally and functionally interconnected (1, 7, 16, 21, 23, 27, 28), although they are organized into two different domains.Open in a separate windowFIG. 1.(A) Schematic diagram of the conserved regions and motifs within HSV-1 Pol. The relative locations of the conserved regions of HSV-1 Pol are shown at the top; regions I to VII and δ region C are represented by open boxes. The conserved exonuclease motifs I, II, and III are indicated with closed boxes. The functional and structural domains (determined by crystal structure analysis [23]) of the HSV-1 Pol are shown below. The N-terminal domain (N domain) is composed of two regions separated by the 3′-to-5′ exonuclease domain (23). (B) Schematic diagram of wild-type and mutant YDL Pol. The BamHI fragment of the wild-type pol from the plasmid pHC629 is shown at the top. The relative location of conserved region VI and the Exo III motif are shown below, with corresponding wild-type and mutant (YDL) amino acid sequences. B, BamHI; M, MstI; N, NotI.The high fidelity of DNA replication is achieved by three different mechanisms: nucleotide discrimination during the polymerization reaction, editing immediately after the polymerization reaction, and postreplication repair. HSV-1 mutant Pol containing mutations within the conserved regions of the polymerase domain can result in altered enzyme kinetics and DNA replication fidelity (8, 9, 11, 12, 18, 26). Similarly, mutation of conserved Exo domain residues can lead to the loss of exonuclease activity and to altered nucleotide selection and incorporation kinetics as well as the mutator phenotype (1, 10, 13, 14, 21, 25). Our previous studies demonstrated that a mutant Pol (YD12) containing a tyrosine-to-histidine substitution at residue 577 and an aspartic acid-to-alanine substitution at residue 581 (Y577H/D581A) is exonuclease deficient (exo) and that recombinant virus expressing the mutant Pol exhibits a mutator phenotype in vivo (14). However, this recombinant virus rapidly evolved to contain an additional leucine-to-phenylalanine substitution at residue 774 (L774F), which is located within conserved region VI of the polymerase domain (18). Interestingly, a recombinant virus containing the L774F Pol mutation exhibits increased fidelity of DNA replication (18). Our recent study also demonstrated that the mutant L774F Pol exhibits altered enzyme kinetics (26). These results led to the hypothesis that the emerged L774F mutation in the context of the YD12 Pol mutant may also affect enzyme activity, DNA replication, and fidelity.  相似文献   

5.
6.
To test a previously coined “charge balance hypothesis” of human hepatitis B virus (HBV) capsid stability, we established an in vitro disassembly and reassembly system using bacterially expressed HBV capsids. Capsid disassembly can be induced by micrococcal nuclease digestion of encapsidated RNA. HBV core protein (HBc) mutants containing various amounts of arginine were constructed by serial truncations at the C terminus. Capsids containing smaller amounts of arginine (HBc 149, 154, and 157) remained intact after micrococcal nuclease digestion by native gel electrophoresis. Capsids containing larger amounts of arginine (HBc 159, 164, 169, and 171) exhibited reduced and more diffuse banding intensity and slightly upshifted mobility (HBc 159 and 164). Capsids containing the largest amounts of arginine (HBc 173, 175, and 183), as well as HBc 167, exhibited no detectable banding signal, indicating loss of capsid integrity or stability. Interestingly, capsid reassembly can be induced by polyanions, including oligonucleotides, poly-glutamic acid, and nonbiological polymer (polyacrylic acid). In contrast, polycations (polylysine and polyethylenimine) and low-molecular-weight anions (inositol triphosphate) induced no capsid reassembly. Results obtained by gel assay were confirmed by electron microscopy. Reassembled capsids comigrated with undigested parental capsids on agarose gels and cosedimented with undigested capsids by sucrose gradient ultracentrifugation. Taken together, the results indicate that HBV capsid assembly and integrity depend on polyanions, which probably can help minimize intersubunit charge repulsion caused mainly by arginine-rich domain III or IV in close contact. The exact structure of polyanions is not important for in vitro capsid reassembly. A large amount of independent experimental evidence for this newly coined “electrostatic interaction hypothesis” is discussed.Chronic infection with hepatitis B virus (HBV) leads to the development of cirrhosis and hepatocellular carcinoma (6, 31, 36). HBV core protein (HBc) consists of the assembly domain (HBc amino acids 1 to 149) at the N terminus and the arginine-rich domain (ARD) at the C terminus (HBc amino acids 150 to 183) (33, 34). Escherichia coli-expressed HBc can spontaneously self-assemble into 28-nm capsid particles with a spherical appearance indistinguishable from that of human liver-derived capsid particles (7). Such capsid particles have been shown to package RNAs transcribed in E. coli (5, 8, 11, 28, 37). The four-helix bundle structure of HBV capsid particles is based on cryo-electron microscopy and X-ray crystallography using C-terminally truncated HBc (34, 38). At present, there is no known structure at the C terminus of HBc capsids (34, 41). HBc amino acids 150 to 183 contains four stretches of clustering arginine residues (ARD-I, -II, -III, and -IV) (Fig. (Fig.1).1). When the C-terminal domain of hepadnaviral core protein was serially truncated, a viral replication defect was observed (4, 19, 22, 27, 39). To date, it remains to be elucidated why the C terminus is so important for diverse biological activities, including RNA encapsidation and DNA replication. To address this issue, we proposed previously a so-called “charge balance hypothesis” (22), which highlights the importance of adequate electrostatic interactions between positive charge (basic residues) from HBc and negative charge from encapsidated RNA or DNA.Open in a separate windowFIG. 1.Effects of micrococcal nuclease treatment on HBV nucleocapsids with serially truncated C termini of core proteins. A series of HBV core expression vectors with different lengths and arginine contents were constructed in pET-Blue-1. The truncations are illustrated in the top panel, and the four ARDs (ARD-I, -II, -III, and -IV) are underlined. All constructs self-assembled into capsids when expressed in E. coli. The capsids run as a distinct band on 1% native agarose gels and can be stained with EtBr (upper panel) and Coomassie blue (middle panel). Untreated controls without micrococcal nuclease were incubated with the same buffer and conditions as for micrococcal nuclease-digested samples. In the upper panel, note the loss of the EtBr signal when the encapsidated RNAs were digested by micrococcal nuclease. In the middle panel, we observed three different groups of mutants with three different CBS patterns. Group 1 mutants, including HBc mutants 149, 154, and 157, exhibited no significant change in CBS banding pattern before or after micrococcal nuclease digestion. In group 2 capsids 159, 164, 169, and 171 (#), the CBS banding pattern became more diffuse, less intense, and slightly (yet reproducibly) upshifted. * indicates the near-complete loss of CBS banding in group 3 capsids 167, 173, 175, and 183. To confirm that this loss of CBS probably results from a loss of structural integrity of the capsid particle, rather than from a loss of the core protein per se, aliquots of samples were also run on denaturing SDS-PAGE (bottom panel).The first clue that such an electrostatic interaction could play an important role in RNA encapsidation is from the study of an engineered mutant, HBc 164. Despite its reduced arginine content relative to the wild-type (WT) full-length HBc 183 (Fig. (Fig.1),1), HBc mutant 164 can encapsidate, as efficiently as WT HBV, both 3.5-kb pregenomic RNA (pgRNA) and a spliced 2.2-kb subgenomic RNA (sgRNA) (19, 22). When WT HBV nucleocapsids (capsids) were treated with micrococcal nuclease, the encapsidated 3.5-kb pgRNA and its reverse-transcribing RNA template were resistant to micrococcal nuclease treatment. In contrast, when mutant 164 capsids were treated with micrococcal nuclease, the encapsidated 3.5-kb RNA was highly nuclease sensitive, while the encapsidated 2.2-kb sgRNA appeared to be nuclease resistant (19, 22).To further elucidate the mechanism behind this phenomenon, we hypothesized that this result could be due to an abnormal or less stable capsid structure generated by a charge imbalance (insufficient positive charge or excessive negative charge) when arginine-deficient mutant 164 encapsidates the full-length 3.5-kb pgRNA. In contrast, a more normal or stable capsid structure can be generated when arginine-deficient mutant 164 encapsidates the 2.2-kb sgRNA (with a reduced negative charge content relative to the 3.5-kb pgRNA) (22). In addition to RNA encapsidation and capsid stability, electrostatic interaction could also play a role in HBV DNA synthesis and genome maturation. For example, when the truncated C terminus of HBc 164 was progressively restored, the core-associated viral DNA gradually increased in both size and signal intensity (22).In this study, we designed an in vitro capsid disassembly/reassembly experiment which is complementary to the in vivo approach in tissue culture (22; P. K. Chua, F. M. Tang, J. Y. Huang, C. S. Suen, and C. Shih, submitted for publication). We investigated the relationship between the arginine content of HBV capsids and the encapsidated nucleic acid in maintaining capsid stability in vitro. In addition, we demonstrated that HBV capsids that are disassembled by depletion of encapsidated RNA can be efficiently reassembled in the presence of exogenous nucleic acid and non-nucleic acid polyanions but not in the presence of polycations or low-molecular-weight (low-MW) anions. Capsid disassembly induced by RNA digestion appears to be related to intersubunit charge repulsion between ARD-III or ARD-IV in close contact.  相似文献   

7.
8.
9.
Streptomyces development was analyzed under conditions resembling those in soil. The mycelial growth rate was much lower than that in standard laboratory cultures, and the life span of the previously named first compartmentalized mycelium was remarkably increased.Streptomycetes are gram-positive, mycelium-forming, soil bacteria that play an important role in mineralization processes in nature and are abundant producers of secondary metabolites. Since the discovery of the ability of these microorganisms to produce clinically useful antibiotics (2, 15), they have received tremendous scientific attention (12). Furthermore, its remarkably complex developmental features make Streptomyces an interesting subject to study. Our research group has extended our knowledge about the developmental cycle of streptomycetes, describing new aspects, such as the existence of young, fully compartmentalized mycelia (5-7). Laboratory culture conditions (dense inocula, rich culture media, and relatively elevated temperatures [28 to 30°C]) result in high growth rates and an orderly-death process affecting these mycelia (first death round), which is observed at early time points (5, 7).In this work, we analyzed Streptomyces development under conditions resembling those found in nature. Single colonies and soil cultures of Streptomyces antibioticus ATCC 11891 and Streptomyces coelicolor M145 were used for this analysis. For single-colony studies, suitable dilutions of spores of these species were prepared before inoculation of plates containing GYM medium (glucose, yeast extract, malt extract) (11) or GAE medium (glucose, asparagine, yeast extract) (10). Approximately 20 colonies per plate were obtained. Soil cultures were grown in petri dishes with autoclaved oak forest soil (11.5 g per plate). Plates were inoculated directly with 5 ml of a spore suspension (1.5 × 107 viable spores ml−1; two independent cultures for each species). Coverslips were inserted into the soil at an angle, and the plates were incubated at 30°C. To maintain a humid environment and facilitate spore germination, the cultures were irrigated with 3 ml of sterile liquid GAE medium each week.The development of S. coelicolor M145 single colonies growing on GYM medium is shown in Fig. Fig.1.1. Samples were collected and examined by confocal microscopy after different incubation times, as previously described (5, 6). After spore germination, a viable mycelium develops, forming clumps which progressively extend along the horizontal (Fig. 1a and b) and vertical (Fig. 1c and d) axes of a plate. This mycelium is fully compartmentalized and corresponds to the first compartmentalized hyphae previously described for confluent surface cultures (Fig. 1e, f, and j) (see below) (5); 36 h later, death occurs, affecting the compartmentalized hyphae (Fig. 1e and f) in the center of the colony (Fig. (Fig.1g)1g) and in the mycelial layers below the mycelial surface (Fig. 1d and k). This death causes the characteristic appearance of the variegated first mycelium, in which alternating live and dead segments are observed (Fig. 1f and j) (5). The live segments show a decrease in fluorescence, like the decrease in fluorescence that occurs in solid confluent cultures (Fig. (Fig.11 h and i) (5, 9). As the cycle proceeds, the intensity of the fluorescence in these segments returns, and the segments begin to enlarge asynchronously to form a new, multinucleated mycelium, consisting of islands or sectors on the colony surfaces (Fig. 1m to o). Finally, death of the deeper layers of the colony (Fig. (Fig.1q)1q) and sporulation (Fig. (Fig.1r)1r) take place. Interestingly, some of the spores formed germinate (Fig. (Fig.1s),1s), giving rise to a new round of mycelial growth, cell death, and sporulation. This process is repeated several times, and typical, morphologically heterogeneous Streptomyces colonies grow (not shown). The same process was observed for S. antibioticus ATCC 11891, with minor differences mainly in the developmental time (not shown).Open in a separate windowFIG. 1.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death of S. coelicolor M145 in surface cultures containing single colonies. Developmental culture times (in hours) are indicated. The images in panels l and n were obtained in differential interference contrast mode and correspond to the same fields as in panels k and m, respectively. The others are culture sections stained with SYTO 9 and propidium iodide. Panels c, d, k, l, p, and q are cross sections; the other images are longitudinal sections (see the methods). Panels h and i are images of the same field taken with different laser intensities, showing low-fluorescence viable hyphae in the center of the colonies that develop into a multinucleated mycelium. The arrows in panels e and s indicate septa (e) and germinated spores (s). See the text for details.Figure Figure22 shows the different types of mycelia present in S. coelicolor cultures under the conditions described above, depending on the compartmentalization status. Hyphae were treated with different fluorescent stains (SYTO 9 plus propidium iodide for nucleic acids, CellMask plus FM4-64 for cell membranes, and wheat germ agglutinin [WGA] for cell walls). Samples were processed as previously described (5). The young initial mycelia are fully compartmentalized and have membranous septa (Fig. 2b to c) with little associated cell wall material that is barely visible with WGA (Fig. (Fig.2d).2d). In contrast, the second mycelium is a multinucleated structure with fewer membrane-cell wall septa (Fig. 2e to h). At the end of the developmental cycle, multinucleated hyphae begin to undergo the segmentation which precedes the formation of spore chains (Fig. 2i to m). Similar results were obtained for S. antibioticus (not shown), but there were some differences in the numbers of spores formed. Samples of young and late mycelia were freeze-substituted using the methodology described by Porta and Lopez-Iglesias (13) and were examined with a transmission electron microscope (Fig. 2n and o). The septal structure of the first mycelium (Fig. (Fig.2n)2n) lacks the complexity of the septal structure in the second mycelium, in which a membrane with a thick cell wall is clearly visible (Fig. (Fig.2o).2o). These data coincide with those previously described for solid confluent cultures (4).Open in a separate windowFIG. 2.Analysis of S. coelicolor hyphal compartmentalization with several fluorescent indicators (single colonies). Developmental culture times (in hours) are indicated. (a, e, and i) Mycelium stained with SYTO 9 and propidium iodide (viability). (b, f, and j) Hyphae stained with Cell Mask (a membrane stain). (c, g, and l) Hyphae stained with FM 4-64 (a membrane stain). (d, h, and m) Hyphae stained with WGA (cell wall stain). Septa in all the images in panels a to j, l, and m are indicated by arrows. (k) Image of the same field as panel j obtained in differential interference contrast mode. (n and o) Transmission electron micrographs of S. coelicolor hyphae at different developmental phases. The first-mycelium septa (n) are comprised of two membranes separated by a thin cell wall; in contrast, second-mycelium septa have thick cell walls (o). See the text for details. IP, propidium iodide.The main features of S. coelicolor growing in soils are shown in Fig. Fig.3.3. Under these conditions, spore germination is a very slow, nonsynchronous process that commences at about 7 days (Fig. 3c and d) and lasts for at least 21 days (Fig. 3i to l), peaking at around 14 days (Fig. 3e to h). Mycelium does not clump to form dense pellets, as it does in colonies; instead, it remains in the first-compartmentalized-mycelium phase during the time analyzed. Like the membrane septa in single colonies, the membrane septa of the hyphae are stained with FM4-64 (Fig. 3j and k), although only some of them are associated with thick cell walls (WGA staining) (Fig. (Fig.3l).3l). Similar results were obtained for S. antibioticus cultures (not shown).Open in a separate windowFIG. 3.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death and hyphal compartmentalization of S. coelicolor M145 growing in soil. Developmental culture times (in days) are indicated. The images in panels b, f, and h were obtained in differential interference contrast mode and correspond to the same fields as the images in panels a, e, and g, respectively. The dark zone in panel h corresponds to a particle of soil containing hyphae. (a, c, d, e, g, i, j, and k) Hyphae stained with SYTO 9, propidium iodide (viability stain), and FM4-64 (membrane stain) simultaneously. (i) SYTO 9 and propidium iodide staining. (j) FM4-64 staining. The image in panel k is an overlay of the images in panels i and j and illustrates that first-mycelium membranous septa are not always apparent when they are stained with nucleic acid stains (SYTO 9 and propidium iodide). (l) Hyphae stained with WGA (cell wall stain), showing the few septa with thick cell walls present in the cells. Septa are indicated by arrows. IP, propidium iodide.In previous work (8), we have shown that the mycelium currently called the substrate mycelium corresponds to the early second multinucleated mycelium, according to our nomenclature, which still lacks the hydrophobic layers characteristic of the aerial mycelium. The aerial mycelium therefore corresponds to the late second mycelium which has acquired hydrophobic covers. This multinucleated mycelium as a whole should be considered the reproductive structure, since it is destined to sporulate (Fig. (Fig.4)4) (8). The time course of lysine 6-aminotransferase activity during cephamycin C biosynthesis has been analyzed by other workers using isolated colonies of Streptomyces clavuligerus and confocal microscopy with green fluorescent protein as a reporter (4). A complex medium and a temperature of 29°C were used, conditions which can be considered similar to the conditions used in our work. Interestingly, expression did not occur during the development of the early mycelium and was observed in the mycelium only after 80 h of growth. This suggests that the second mycelium is the antibiotic-producing mycelium, a hypothesis previously confirmed using submerged-growth cultures of S. coelicolor (9).Open in a separate windowFIG. 4.Cell cycle features of Streptomyces growing under natural conditions. Mycelial structures (MI, first mycelium; MII, second mycelium) and cell death are indicated. The postulated vegetative and reproductive phases are also indicated (see text).The significance of the first compartmentalized mycelium has been obscured by its short life span under typical laboratory culture conditions (5, 6, 8). In previous work (3, 7), we postulated that this structure is the vegetative phase of the bacterium, an hypothesis that has been recently corroborated by proteomic analysis (data not shown). Death in confluent cultures begins shortly after germination (4 h) and continues asynchronously for 15 h. The second multinucleated mycelium emerges after this early programmed cell death and is the predominant structure under these conditions. In contrast, as our results here show, the first mycelium lives for a long time in isolated colonies and soil cultures. As suggested in our previous work (5, 6, 8), if we assume that the compartmentalized mycelium is the Streptomyces vegetative growth phase, then this phase is the predominant phase in individual colonies (where it remains for at least 36 h), soils (21 days), and submerged cultures (around 20 h) (9). The differences in the life span of the vegetative phase could be attributable to the extremely high cell densities attained under ordinary laboratory culture conditions, which provoke massive differentiation and sporulation (5-7, 8).But just exactly what are “natural conditions”? Some authors have developed soil cultures of Streptomyces to study survival (16, 17), genetic transfer (14, 17-19), phage-bacterium interactions (3), and antibiotic production (1). Most of these studies were carried out using amended soils (supplemented with chitin and starch), conditions under which growth and sporulation were observed during the first few days (1, 17). These conditions, in fact, might resemble environments that are particularly rich in organic matter where Streptomyces could conceivably develop. However, natural growth conditions imply discontinuous growth and limited colony development (20, 21). To mimic such conditions, we chose relatively poor but more balanced carbon-nitrogen soil cultures (GAE medium-amended soil) and less dense spore inocula, conditions that allow longer mycelium growth times. Other conditions assayed, such as those obtained by irrigating the soil with water alone, did not result in spore germination and mycelial growth (not shown). We were unable to detect death, the second multinucleated mycelium described above, or sporulation, even after 1 month of incubation at 30°C. It is clear that in nature, cell death and sporulation must take place at the end of the long vegetative phase (1, 17) when the imbalance of nutrients results in bacterial differentiation.In summary, the developmental kinetics of Streptomyces under conditions resembling conditions in nature differs substantially from the developmental kinetics observed in ordinary laboratory cultures, a fact that should be born in mind when the significance of development-associated phenomena is analyzed.  相似文献   

10.
11.
l-2-Amino-4-methoxy-trans-3-butenoic acid (AMB) is a potent antibiotic and toxin produced by Pseudomonas aeruginosa. Using a novel biochemical assay combined with site-directed mutagenesis in strain PAO1, we have identified a five-gene cluster specifying AMB biosynthesis, probably involving a thiotemplate mechanism. Overexpression of this cluster in strain PA7, a natural AMB-negative isolate, led to AMB overproduction.The Gram-negative bacterium Pseudomonas aeruginosa is an opportunistic pathogen that causes a wide range of human infections and is considered the main pathogen responsible for chronic pneumonia in cystic fibrosis patients (7, 23). P. aeruginosa also infects other organisms, such as insects (4), nematodes (6), plants (18), and amoebae (20). Its ability to thrive as a pathogen and to compete in aquatic and soil environments can be partly attributed to the production and interplay of secreted virulence factors and secondary metabolites. While the importance of many of these exoproducts has been studied, the antimetabolite l-2-amino-4-methoxy-trans-3-butenoic acid (AMB; methoxyvinylglycine) (Fig. (Fig.1)1) has received only limited attention. Identified during a search for new antibiotics, AMB was found to reversibly inhibit the growth of Bacillus spp. (26) and Escherichia coli (25) and was later shown to inhibit the growth and metabolism of cultured Walker carcinosarcoma cells (28). AMB is a γ-substituted vinylglycine, a naturally occurring amino acid with a β,γ-C=C double bond. Other members of this family are aminoethoxyvinylglycine from Streptomyces spp. (19) and rhizobitoxine, made by Bradyrhizobium japonicum (16) and Pseudomonas andropogonis (15) (Fig. (Fig.1).1). As inhibitors of pyridoxal phosphate-dependent enzymes (13, 17, 21, 22), γ-substituted vinylglycines have multiple targets in bacteria, animals, and plants (3, 5, 10, 21, 22, 29). However, the importance of AMB as a toxin in biological interactions with P. aeruginosa has not been addressed, as AMB biosynthesis and the genes involved have not been elucidated.Open in a separate windowFIG. 1.Chemical structures of the γ-substituted vinylglycines AMB, aminoethoxyvinylglycine, and rhizobitoxine.  相似文献   

12.
It was recently shown that Myxococcus xanthus harbors an alternative and reversible biosynthetic pathway to isovaleryl coenzyme A (CoA) branching from 3-hydroxy-3-methylglutaryl-CoA. Analyses of various mutants in these pathways for fatty acid profiles and fruiting body formation revealed for the first time the importance of isoprenoids for myxobacterial development.Myxobacteria are unique among the prokaryotes as (i) they can form highly complex fruiting bodies under starvation conditions, even up to microscopic tree-like structures (28); (ii) they can move on solid surfaces using different motility mechanisms (16); (iii) they produce some of the most cytotoxic secondary metabolites, with epothilone already in clinical use against cancer (2, 3); and (iv) they harbor the largest prokaryotic genomes found so far (15, 27). The large genome might be directly related to their complex life-style and the diverse secondary (3) and primary (9) metabolisms. Already in 2002 we found that myxobacteria are able to produce isovaleryl coenzyme A (IV-CoA) and compounds derived thereof via a new pathway that branches from 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA), which is the central intermediate of the well-known mevalonate-dependent isoprenoid biosynthesis (Fig. (Fig.1)1) (22, 23). Usually IV-CoA is derived from leucine degradation via the branched-chain keto acid dehydrogenase (BKD) complex (24), which is also the preferred pathway to IV-CoA in the myxobacteria Myxococcus xanthus and Stigmatella aurantiaca (Fig. (Fig.2A).2A). However, in bkd mutants, where no or only residual leucine degradation is possible (30), the alternative pathway is induced (Fig. (Fig.2B),2B), presumably to ensure the production of iso-fatty acids (iso-FAs) (5). A possible reason for this alternative pathway is the importance of IV-CoA-derived compounds in the complex myxobacterial life cycle, which is the starvation-induced formation of fruiting bodies in which the cells differentiate into myxospores. We showed that this pathway is induced during fruiting body formation in M. xanthus when leucine is limited. Under these conditions, this pathway might be more important for protein synthesis than for lipid remodeling, as lipids are present in excess during development due to the surface reduction from vegetative rods to round myxospores as described previously (29). Examples of IV-CoA-derived compounds are the unusual iso-branched ether lipids, which are almost exclusively produced in the developing myxospores. They might serve as structural lipids and signaling compounds during fruiting body formation (26).Open in a separate windowFIG. 1.Biosynthesis of IV-CoA and compounds derived thereof and biosynthesis of isoprenoids in M. xanthus. Broken arrows indicate multistep reactions; supplementation (double-lined arrows) with MVL and IVA can be used to complement selected mutants.Open in a separate windowFIG. 2.Short representations of proposed metabolic fluxes through the IV-CoA/isoprenoid network. Broken arrows indicate no metabolic flux. (A) DK1622 (wild type); (B) DK5643 (Δbkd); (C) DK5624 (Δbkd mvaS::kan); (D) HB002 (Δbkd liuC::kan); (E) HB002 with 1 mM IVA; (F) HB002 with 1 mM MVL. Ac-CoA, acetyl-CoA; MVA, mevalonic acid.In M. xanthus, we could recently identify candidate genes involved in the alternative pathway from HMG-CoA to IV-CoA. We also described the genes required for the degradation pathway of leucine and subsequently also those involved in the transformation of IV-CoA to HMG-CoA (4). In myxobacteria leucine is an important precursor for isoprenoid biosynthesis, as was already shown elsewhere for the biosynthesis of steroids (7) and prenylated secondary metabolites like aurachin (22) or leupyrrins (6), as well as volatiles like geosmin or germacradienol in M. xanthus and S. aurantiaca (11, 13). The interconnection of iso-FAs and isoprenoid biosynthesis made it difficult to assign functions to these compound classes during fruiting body formation in M. xanthus because it cannot be excluded that reduced leucine degradation also impairs isoprenoid biosynthesis. A mutant strain of M. xanthus that was blocked in the degradation of leucine and the alternative pathway had a deletion in the bkd locus as well as a plasmid insertion in the mvaS gene encoding the HMG-CoA synthase (strain DK5624). This double mutation severely affected isoprenoid biosynthesis (5), and cultures of DK5624 must be supplemented with mevalonolactone (MVL; the cyclized form of mevalonic acid) in order to enable growth (Fig. (Fig.2C).2C). Since we have identified the genes involved in IV-CoA biosynthesis and the mevalonate pathway (4), we can now start to identify differences between strains that show deficiencies in iso-FAs and strains that show deficiencies in isoprenoids via simple analysis of the FA profile and analysis of the myxobacterial development of selected mutants.All mutants used in this study (HB002 [Δbkd liuC::kan], HB015 [Δbkd MXAN_4265::kan], DK5624 [Δbkd mvaS::kan], HB019 [Δbkd mvaS::kan mvaS+], and HB020 [Δbkd MXAN_4265::kan mvaS+]) have been published previously (4), and FA analysis as well as myxobacterial fruiting body formation has also been described previously (26).M. xanthus HB002 (Δbkd liuC) shows only residual amounts of iso-FAs, as both leucine degradation and the alternative pathway to IV-CoA are blocked (Fig. (Fig.2D)2D) and its capability to form fruiting bodies is strongly reduced (Fig. (Fig.3).3). The residual amount of iso-FAs results from a second BKD activity in M. xanthus that has been identified by residual leucine incorporation as well as by residual enzymatic activity in bkd mutants (23, 30). This second BKD activity might be a side activity of the pyruvate dehydrogenase or a related chemical oxidative decarboxylation, as no second bkd locus could be identified in the genome (unpublished results). Moreover, growth of HB002 is not MVL dependent because the block in the alternative pathway does not affect isoprenoid biosynthesis, as liuC encodes a dehydratase/hydratase that is involved in the conversion of HMG-CoA to 3-methylglutaconyl-CoA and vice versa (4). As expected, the FA profile (4) as well as the developmental phenotype (data not shown) can be complemented (Fig. (Fig.2E)2E) by the addition of isovaleric acid (IVA), the free acid of IV-CoA, indicating the importance of iso-branched compounds for development in M. xanthus. Unexpectedly, addition of MVL (Fig. (Fig.2F)2F) also partially restored fruiting body formation without restoring the FA profile (Fig. (Fig.3).3). Similarly, M. xanthus HB015 (Δbkd MXAN_4265::kan) can produce only traces of iso-FAs, as both pathways to IV-CoA are blocked. MXAN_4265 encodes a protein with similarity to a glutaconyl-CoA transferase subunit, but from our previous results, we postulated it to be involved in the alternative pathway to IV-CoA (Fig. (Fig.1)1) (4). The respective mutant shows a severely impaired developmental phenotype, which can be complemented not only by the addition of IVA (not shown) but also by the addition of MVL (Fig. (Fig.3).3). Again, no change in the FA profile was observed after the addition of MVL. However, a plasmid insertion into MXAN_4265 has a polar effect on mvaS, which is the last gene in this five-gene operon and which is crucial for HMG-CoA formation from acetoacetyl-CoA and acetyl-CoA. Therefore, we assume that both pathways to HMG-CoA are blocked in HB015: no HMG-CoA can be made from acetyl-CoA and hardly any can be made via leucine degradation. In order to prove this hypothesis, we complemented HB015 with an additional copy of mvaS under the constitutive T7A1 promoter as described previously, using the plasmid pCK4267exp (4). The resulting strain, HB020 (Δbkd MXAN_4265::kan mvaS+), showed a restored developmental phenotype but still produced only trace amounts of iso-FAs.Open in a separate windowFIG. 3.Fruiting body formation on TPM agar in selected mutants at 24, 48, and 72 h after starvation. Numbers refer to the relative amounts (in percentages) of the most abundant iso-FA, iso-15:0, which is indicative of iso-FAs in general. Strains were DK1622 (wild type), HB002 (Δbkd liuC::kan), HB015 (Δbkd MXAN_4265::kan), DK5624 (Δbkd mvaS::kan), HB019 (Δbkd mvaS::kan mvaS+), and HB020 (Δbkd MXAN_4265::kan mvaS+). DK5624 was grown with 0.3 mM MVL prior to starvation, and the cells were washed and plated on TPM with or without 1 mM of MVL.The data from HB002, HB015, and HB020 indicate an important function of the mevalonate-dependent isoprenoid pathway for fruiting body formation in M. xanthus. Therefore, MVL addition can at least partially complement the developmental phenotype of DK5624, which cannot form fruiting bodies without MVL (Fig. (Fig.3).3). However, genetic complementation with mvaS in HB019 resulted in the expected complementation of the fruiting body formation and the FA profile (Fig. (Fig.3,3, bottom row).Leucine is one of the most abundant proteinogenic amino acids. It is also an essential amino acid for M. xanthus (8), which has a predatory life-style (1), as it lives on other bacteria and fungi that contain a lot of leucine. Moreover, leucine is very efficiently incorporated into isoprenoids like geosmin and aurachin (10, 22). Thus, one can conclude that in fact leucine degradation is the major pathway for HMG-CoA biosynthesis instead of the usual formation via acetoacetyl-CoA and acetyl-CoA by the HMG-CoA synthase MvaS as indicated in Fig. Fig.2A.2A. No difference in growth was observed between culture with and culture without MVL for HB002 (Δbkd liuC::kan) and HB015 (Δbkd MXAN_4265::kan) in rich medium (data not shown), probably due to the complete MvaS activity (in HB002) or residual BKD activity (in HB002 and HB015), resulting in all precursors for the mevalonate-dependent isoprenoid biosynthesis still being present in excess under these conditions. However, under starvation conditions a small reduction in HMG-CoA biosynthesis caused by completely blocked leucine degradation (as in HB002 due to the mutation in liuC [Fig. [Fig.2D])2D]) or reduced leucine degradation and a mutation in mvaS (as in HB015) might each result in a reduced isoprenoid level, which can be complemented at least partially by the addition of MVL. This would also explain the difference in the developmental phenotypes of HB002 and HB015, with the phenotype being more severe in HB002 (Fig. (Fig.3).3). The fact that complementation with IVA is in all cases more efficient than that with MVL can be explained by the role of the already-mentioned isolipids. They can be produced only after IVA addition, which also complements the (developmental) phenotype of some of these mutants (26).As isoprenoids represent probably the most diverse class of natural products (14), it is very hard to predict which particular isoprenoids might be responsible for the observed effects. Several isoprenoids (7, 11-13), prenylated secondary metabolites (6, 22), and carotenoids (18-21) are known from myxobacteria in general, and a major volatile compound from M. xanthus is the terpenoid geosmin (13). In order to test whether geosmin might be required for fruiting body formation, we constructed a plasmid insertion mutant in MXAN_6247, which is involved in the cyclization of farnesyl diphosphate to geosmin, following published procedures (4, 5). The resulting strain, HB022, showed the expected loss in geosmin production but no developmental phenotype (data not shown).Additionally, it cannot be excluded that prenylated proteins, sugars, or quinones from the respiratory chain are important for fruiting body formation. Moreover, stigmolone has been described as a pheromone involved in fruiting body formation in S. aurantiaca (25). Although its biosynthesis has not been elucidated yet, stigmolone could be an isoprenoid as well, which is deducible from the two iso-branched residues within its chemical structure (17). Nevertheless, the importance of isoprenoids for M. xanthus is evident from the data presented, and clearly more work is needed to identify the compound(s) involved.  相似文献   

13.
14.
15.
In the present work, lysine production by Corynebacterium glutamicum was improved by metabolic engineering of the tricarboxylic acid (TCA) cycle. The 70% decreased activity of isocitrate dehydrogenase, achieved by start codon exchange, resulted in a >40% improved lysine production. By flux analysis, this could be correlated to a flux shift from the TCA cycle toward anaplerotic carboxylation.With an annual market volume of more than 1,000,000 tons, lysine is one of the dominating products in biotechnology. In recent years, rational metabolic engineering has emerged as a powerful tool for lysine production (16, 18, 22). Hereby, different target enzymes and pathways in the central metabolism could be identified and successfully modified to create superior production strains (1, 2, 5, 8, 10, 17-20). The tricarboxylic acid (TCA) cycle has not been rationally engineered so far, despite its major role in Corynebacterium glutamicum (6). From metabolic flux studies, however, it seems that the TCA cycle might offer a great potential for optimization (Fig. (Fig.1),1), which is also predicted from in silico pathway analysis (13, 22). Experimental evidence comes from studies with Brevibacterium flavum exhibiting increased lysine production due to an induced bottleneck toward the TCA cycle (21). In the present work, we performed TCA cycle engineering by downregulation of isocitrate dehydrogenase (ICD). ICD is the highest expressed TCA cycle enzyme in C. glutamicum (7). Downregulation was achieved by start codon exchange, controlling ICD expression on the level of translation.Open in a separate windowFIG. 1.Stoichiometric correlation of lysine yield (%), biomass yield (g/mol) and TCA cycle flux (%; entry flux through citrate synthase) determined by 13C metabolic flux analysis achieved by paraboloid fitting of the data set (parameters were determined with Y0 = 105.1, a = −1.27, b = 0.35, c = −9.35 × 10−3, d = −11.16 × 10−3). The data displayed represent values from 18 independent experiments with different C. glutamicum strains taken from previous studies (1-3, 11, 12, 15, 23).  相似文献   

16.
17.
18.
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   

19.
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   

20.
Homologous recombination is an error-free mechanism for the repair of DNA double-strand breaks (DSBs). Most DSB repair events occur by gene conversion limiting loss of heterozygosity (LOH) for markers downstream of the site of repair and restricting deleterious chromosome rearrangements. DSBs with only one end available for repair undergo strand invasion into a homologous duplex DNA, followed by replication to the chromosome end (break-induced replication [BIR]), leading to LOH for all markers downstream of the site of strand invasion. Using a transformation-based assay system, we show that most of the apparent BIR events that arise in diploid Saccharomyces cerevisiae rad51Δ mutants are due to half crossovers instead of BIR. These events lead to extensive LOH because one arm of chromosome III is deleted. This outcome is also observed in pol32Δ and pol3-ct mutants, defective for components of the DNA polymerase δ (Pol δ) complex. The half crossovers formed in Pol δ complex mutants show evidence of limited homology-dependent DNA synthesis and are partially Mus81 dependent, suggesting that strand invasion occurs and the stalled intermediate is subsequently cleaved. In contrast to rad51Δ mutants, the Pol δ complex mutants are proficient for repair of a 238-bp gap by gene conversion. Thus, the BIR defect observed for rad51 mutants is due to strand invasion failure, whereas the Pol δ complex mutants are proficient for strand invasion but unable to complete extensive tracts of recombination-initiated DNA synthesis.DNA double-strand breaks (DSBs) are potentially lethal lesions that can occur spontaneously during normal cell metabolism, by treatment of cells with DNA-damaging agents, or during programmed recombination processes (54). There are two major pathways to repair DSBs: nonhomologous end joining (NHEJ) and homologous recombination (HR). NHEJ involves the religation of the two ends of the broken chromosome and can occur with high fidelity or be accompanied by a gain or loss of nucleotides at the junction (9). Repair of two-ended DSBs by HR generally occurs by gene conversion resulting from a transfer of information from the intact donor duplex to the broken chromosome (Fig. (Fig.1).1). HR occurs preferentially during S and G2 when a sister chromatid is available to template repair (2, 19, 22). Sister-chromatid recombination events are genetically silent, whereas gene conversion between nonsister chromatids associated with an exchange of flanking markers can result in extensive loss of heterozygosity (LOH) or chromosome rearrangements (3, 21). One-ended DSBs that arise by replication fork collapse or by erosion of uncapped telomeres are thought to repair by strand invasion into homologous duplex DNA followed by replication to the end of the chromosome, a process referred to as break-induced replication (BIR) (35). BIR appears to be suppressed at two-ended breaks, presumably because it can lead to extensive LOH if it occurs between homologues or to chromosome translocations when strand invasion initiates within dispersed repeated sequences (5, 28, 31, 50, 52, 55).Open in a separate windowFIG. 1.Models for gene conversion and BIR. After formation of a DSB, the ends are resected to generate 3′ single-strand DNA tails. One end undergoes Rad51-dependent strand invasion to prime DNA synthesis from the invading 3′ end templated by the donor duplex. For gene conversion by the synthesis-dependent strand annealing model, the extended invading end is displaced and can anneal to the other side of the break; completion of repair requires DNA synthesis primed from the noninvading 3′ end. For a one-ended break, or if the other side of the break lacks homology to the donor duplex, DNA synthesis proceeds to the end of the chromosome. Centromeres are shown as solid ovals and a heterozygous marker centromere distal to the site of repair as A/a.The strand invasion step of BIR is assumed to be the same as that for gene conversion based on the requirement for the same HR proteins: Rad51, Rad52, Rad54, Rad55, and Rad57 (10). However, subsequent steps in BIR are less well defined. Recent studies of the fate of the invading end during BIR in diploid strains with polymorphic chromosome III homologues using a plasmid-based assay have shown that following strand invasion, the invading end is capable of dissociating from the initial homologous template. Following dissociation, the displaced end subsequently reinvades into the same or a different chromosome III homologue by a process termed template switching (52). One of the interesting features of the template switching events is that they occur over a region of about 10 kb downstream of the site of strand invasion and do not extend over the entire left arm of chromosome III. There are a number of possible mechanisms that could account for this apparent change in the processivity of BIR. First, it is possible that the strand invasion intermediate is cleaved by a structure-specific nuclease and once the invading strand is covalently joined to one of the template strands, the strand invasion process is irreversible. Recent studies of Schizosaccharomyces pombe have shown an essential role for Mus81, a structure-specific nuclease, in resolution of sister chromatid recombination intermediates during repair of collapsed replication forks (48). Another possibility is that there could be a switch between a translesion DNA polymerase and a highly processive DNA polymerase during BIR. The translesion polymerases in budding yeast, polymerase ζ (Pol ζ) and Pol η, are encoded by REV3-REV7 and RAD30, respectively (34, 40, 43). Deletion of REV3 has been shown to increase the fidelity of DNA synthesis associated with HR but has no effect on the overall frequency of DSB-induced HR (16). Deletion of POLη in chicken DT40 cells reduces the frequency of DSB-induced gene conversion, and human POL η has been shown to extend the invading 3′ end of D-loop intermediates in vitro (23, 36). However, this same preference for Pol η is not found for Saccharomyces cerevisiae. Instead, DNA synthesis during meiotic and mitotic recombination appears to be carried out by Pol δ, one of the three nuclear replicative polymerases, which normally functions with Pol α in Okazaki fragment synthesis (13, 32, 33, 44). Pol ɛ is thought to be the primary leading-strand polymerase (47), but in the absence of the Pol ɛ catalytic domain, Pol δ is presumed to carry out leading-strand synthesis (24). Recent studies by Lydeard et al. (30) have shown a requirement for the lagging-strand polymerases, Pol δ and Pol α, to form the initial primer extension product during BIR, and Pol ɛ is required to complete replication to the end of the chromosome. In contrast, repair of DSBs by gene conversion does not require Pol α, and there appears to be functional redundancy between Pol δ and Pol ɛ (56).To address the roles of Mus81, Pol δ, and Pol η in BIR and in particular template switching, we used the transformation-based BIR assay with diploids with polymorphic chromosome III homologues. Because the transformation assay can only be used with strains with viable mutations of replication factors, we used a null allele of POL32, encoding a nonessential subunit of the Pol δ complex (14), and a point mutation in the gene encoding the essential catalytic subunit, POL3. The pol3-ct allele results in a truncation removing the last four amino acids of the Pol3 protein; the C-terminal region of Pol3 is implicated in interaction with the other essential subunit of the Pol δ complex, Pol31 (15, 49). The interesting feature of the pol3-ct allele is that it decreases the length of gene conversion tracts during mitotic and meiotic recombination, presumably by affecting the processivity of Pol δ, but confers no apparent defect in normal DNA synthesis (32, 33). Because BIR requires more-extensive tracts of DNA synthesis than gene conversion, we expected the pol3-ct mutant to exhibit a BIR defect. We found that in the absence of a fully functional Pol δ complex, chromosome fragment (CF) formation proceeds by a half-crossover mechanism associated with loss of the template chromosome, an event with potentially catastrophic consequences (6, 57). This was also found to occur in rad51 mutants, suggesting nonreciprocal translocations arise by failure to undergo strand invasion or because replication following strand invasion is inefficient. In contrast to rad51 mutants, the Pol δ complex mutants are proficient for repair of a 238-bp gap by gene conversion and fully resistant to ionizing radiation, suggesting there is a unique requirement for Pol δ to complete BIR. Consistent with studies of gene conversion in S. cerevisiae (33), we found no role for Pol η in BIR or the process of template switching.  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