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 62 毫秒
1.
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   

2.
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   

3.
4.
5.
Antibody recognition force microscopy showed that OmcA and MtrC are expressed on the exterior surface of living Shewanella oneidensis MR-1 cells when Fe(III), including solid-phase hematite (Fe2O3), was the terminal electron acceptor. OmcA was localized to the interface between the cell and mineral. MtrC displayed a more uniform distribution across the cell surface. Both cytochromes were associated with an extracellular polymeric substance.Shewanella oneidensis MR-1 is a dissimilatory metal-reducing bacterium that is well known for its ability to use a variety of anaerobic terminal electron acceptors (TEAs), including solid-phase iron oxide minerals, such as goethite and hematite (8, 10). Previous studies suggest that S. oneidensis MR-1 uses outer membrane cytochromes OmcA and MtrC to catalyze the terminal reduction of Fe(III) through direct contact with the extracellular iron oxide mineral (2, 8, 10, 15, 16, 20, 21, 23). However, it has yet to be shown whether OmcA or MtrC is actually targeted to the external surface of live S. oneidensis MR-1 cells when Fe(III) serves as the TEA.In the present study, we used atomic force microscopy (AFM) to probe the surface of live S. oneidensis MR-1 cells, using AFM tips that were functionalized with cytochrome-specific polyclonal antibodies (i.e., anti-OmcA or anti-MtrC). This technique, termed antibody recognition force microscopy (Ig-RFM), detects binding events that occur between antibodies (e.g., anti-OmcA) on an AFM tip and antigens (e.g., OmcA) that are exposed on a cell surface. While this is a relatively new technique, Ig-RFM has been used to map the nanoscale spatial location of single molecules in complex biological structures under physiological conditions (5, 9, 11, 13).Anti-MtrC or anti-OmcA molecules were covalently coupled to silicon nitride (Si3N4) cantilevers (Veeco or Olympus) via a flexible, heterofunctional polyethylene glycol (PEG) linker molecule. The PEG linker consists of an NHS (N-hydroxysuccinimide) group at one end and an aldehyde group at the other end (i.e., NHS-PEG-aldehyde). AFM tips were functionalized with amine groups, using ethanolamine (6, 7). The active NHS ester of the NHS-PEG-aldehyde linker molecule was then used to form a covalent linkage between PEG-aldehyde and the amine groups on the AFM tips (6, 7). Next, anti-MtrC or anti-OmcA molecules were covalently tethered to these tips via the linker molecule''s aldehyde group. This was accomplished by incubating the tips with antibody (0.2 mg/ml) and NaCNBH3 as described previously (7). The cantilevers were purchased from Veeco and had spring constant values between 0.06 and 0.07 N/m, as determined by the thermal method of Hutter and Bechhoefer (12).Prior to conducting the Ig-RFM experiments, the specificity of each polyclonal antibody (i.e., anti-OmcA and anti-MtrC) for OmcA or MtrC was verified by Western blot analysis as described previously (24, 28). Proteins were resolved by both denaturing and nondenaturing polyacrylamide gel electrophoresis (PAGE). Briefly, 2.5 μg of purified OmcA or MtrC (23) was resolved by sodium dodecyl sulfate-PAGE or native PAGE, transferred to a polyvinylidene difluoride membrane, incubated with either anti-OmcA or anti-MtrC, and then visualized using the Amersham ECL Plus Western blotting detection kit. Anti-OmcA bound exclusively to OmcA, anti-MtrC bound exclusively to MtrC, and neither antibody showed cross-reactivity with the other cytochrome. Antibody specificities of anti-OmcA and anti-MtrC were also validated by immunoblot analysis of S. oneidensis whole-cell lysate (28).To determine if MtrC or OmcA was expressed on the external surface of live bacteria when Fe(III) served as the TEA, Ig-RFM was conducted on wild-type versus ΔomcA ΔmtrC double mutant cells. For these experiments, bacteria were cultivated anaerobically with Fe(III), in the form of Fe(III) chelated to nitrilotriacetic acid (NTA), serving as the TEA (19, 23). Growth conditions have been described elsewhere (3, 15) and were based on previous studies (3, 15, 16, 18) that suggest that S. oneidensis MR-1 targets OmcA and MtrC to the cell surface when Fe(III) serves as the TEA.An Asylum Research MFP-3D-BIO AFM or a Digital Instruments Bioscope AFM (16, 17) was used for these experiments. The z-piezoelectric scanners were calibrated as described previously (17). Cells were deposited on a hydrophobic glass coverslip and immersed in imaging buffer (i.e., phosphate-buffered saline [pH 7.4]). The hydrophobic glass coverslips were made as described previously (17) using a self-assembling silane compound called octadecyltrichlorosilane (OTS; Sigma-Aldrich). S. oneidensis MR-1 cells readily adsorbed onto OTS glass coverslips and remained attached to the coverslips during the entire experiment. No lateral cell movement was observed during the experiment, consistent with previous studies that used OTS glass to immobilize bacteria (15, 17, 18, 27).The AFM tip was brought into contact with the surface of a bacterium, and the antibody-functionalized tip was repeatedly brought into and out of contact with the sample, “fishing” for a binding reaction with cytochrome molecules that were exposed on the external cell surface. Binding events were observed upon separating anti-OmcA- or anti-MtrC-functionalized tips from wild-type S. oneidensis MR-1 cells (Fig. (Fig.1).1). For the wild-type cells, we observed both nonspecific and specific interactions (Fig. (Fig.11).Open in a separate windowFIG. 1.Retraction force curves for anti-MtrC-functionalized tips (A) and anti-OmcA-functionalized tips (B) that are being pulled away from the surface of living ΔomcA ΔmtrC double mutant (gray dotted line) or wild-type (solid black line) S. oneidensis MR-1. These bacteria were adsorbed onto OTS glass coverslips. (C) Retraction curves exhibiting nonspecific binding, specific binding, or no binding between the AFM tip and the cell surface.The distinction between “specific” and “nonspecific” adhesion is made by observing the change in slope of the force curve during the retraction process (26). During specific binding (Fig. (Fig.1C),1C), the cantilever is initially relaxed as it is pulled away from the sample. Upon further retraction, the ligand-receptor complex becomes stretched and unravels, resulting in a nonlinear force profile as noted in references 26 and 16. On the other hand, nonspecific adhesion (Fig. (Fig.1C)1C) maintains the same slope during the retraction process because only the cantilever flexes (26).Figure Figure22 summarizes the frequency or probability of observing a binding event for both anti-OmcA and anti-MtrC tips. Each bar in Fig. Fig.22 represents one experiment in which 500 to 1,000 force curves were collected between one AFM tip and two to four live bacterial cells. This figure does not make a distinction between specific and nonspecific binding. It simply shows the frequency of observing an attractive interaction as the antibody-functionalized tip was pulled away from the surface of S. oneidensis MR-1. Binding events occurred with roughly the same frequency when wild-type S. oneidensis MR-1 cells were probed with anti-MtrC-functionalized tips as when they were probed with anti-OmcA-functionalized tips (Fig. (Fig.22).Open in a separate windowFIG. 2.Histograms showing the frequency of observing a binding event for anti-MtrC-functionalized (blue) or anti-OmcA-functionalized (red) AFM tips on live wild-type S. oneidensis MR-1 (solid bars) or ΔomcA ΔmtrC double mutant (diagonally hatched bars) cells. The downward arrows designate injection of free antibody into the imaging buffer. The solid gray bars correspond to results obtained with unbaited AFM tips.A number of control experiments were performed to verify the detection of OmcA and MtrC on the surface of wild-type S. oneidensis MR-1. First, 0.1 μM of free anti-OmcA (or anti-MtrC) was added to the imaging fluid to block binding between the antibody-functionalized AFM tip and surface-exposed cytochromes (11, 16). This decreased the adhesion that was observed between the antibody-functionalized tip and the cell surface (Fig. (Fig.22).Second, we performed force measurements on ΔomcA ΔmtrC double mutant S. oneidensis MR-1 cells. This mutant is deficient in both OmcA and MtrC (19, 23, 24) but produces other proteins native to the outer surface of S. oneidensis MR-1. The resulting force spectra showed a noticeable reduction in binding events for the ΔomcA ΔmtrC double mutant cells (Fig. (Fig.2).2). The binding events that were observed for the double mutant were only nonspecific in nature (Fig. (Fig.1).1). This indicates that the antibodies on the tip do not participate in specific interactions with other proteins on the surface of S. oneidensis MR-1 cells.As a final control experiment, force measurements were conducted on wild-type S. oneidensis MR-1 cells, using Si3N4 tips conjugated with the PEG linker but not functionalized with polyclonal antibody (unbaited tips). Like the results with the double mutant, the unbaited tips were largely unreactive with the surface of the bacteria (Fig. (Fig.2).2). Those binding events that were observed were nonspecific in nature. Taken together, these results demonstrate that the antibody-coated tips have a specific reactivity with OmcA and MtrC molecules. Furthermore, these force measurements show that MtrC and OmcA are present on the external cell surface when Fe(III) serves as the TEA.To map the distribution of cytochromes on living cells, Ig-RFM was conducted on living S. oneidensis MR-1 cells that were growing on a hematite (α-Fe2O3) thin film. The conditions for these experiments were as follows. A hematite film was grown on a 10-mm by 10-mm by 1-mm oxide substrate via oxygen plasma-assisted molecular beam epitaxy (14, 16). The cells were grown anaerobically to mid-log phase with Fe(III)-NTA serving as the TEA. Cells were deposited onto the hematite thin film along with anaerobic growth medium that lacked Fe(III)-NTA. The cells were allowed to attach to the hematite surface (without drying) overnight in an anaerobic chamber. The following day, the liquid was carefully removed and immediately replaced with fresh anaerobic solution (pH 7.4). Ig-RFM was performed on the cells by raster scanning an antibody-functionalized AFM tip across the sample surface, thereby creating an affinity map (1). Force curves were collected for a 32-by-32 array. The raw pixilated force-volume data were deconvoluted using a regularized filter algorithm. The total time to acquire a complete image was approximately 20 min.As noted above, attractive interactions between an antibody tip and cell resulted in relatively short-range, nonspecific and longer-range, specific adhesive forces (Fig. (Fig.1C).1C). To distinguish between these two interactions, we integrated each force curve beginning at >20 nm and ending at the full retraction of the piezoelectric motor (∼1,800 nm). This integration procedure quantifies the work of binding, measured in joules, between the antibody tip and a particular position on the sample. While this integration procedure does not totally exclude nonspecific binding, it does select for those events associated primarily with specific antibody-antigen binding. Figure Figure33 is the antibody-cytochrome recognition images for MtrC and OmcA. The corresponding height (or topography) images of the bacterial cells are also shown in Fig. Fig.33.Open in a separate windowFIG. 3.Ig-RFM of live S. oneidensis MR-1 cells deposited on a hematite (α-Fe2O3) thin film. Height image (A) and corresponding Ig-RFM image (B) for a bare unfunctionalized Si3N4 tip. Height and corresponding Ig-RFM image for a tip functionalized with anti-MtrC (C and D) or anti-OmcA (E and F). Each panel contains a thin white oval showing the approximate location of the bacterium on the hematite surface. A color-coded scale bar is shown on the right (height in micrometers [μm], and the work required to separate the tip from the surface in attojoules [aJ]).OmcA molecules were concentrated at the boundary between the bacterial cell and hematite surface (Fig. 3E and F). MtrC molecules were also detected at the edge of a cell (Fig. 3C and D). Some MtrC, unlike OmcA, was observed on the cell surface distal from the point of contact with the mineral (Fig. 3C and D). Both OmcA and MtrC were also present in an extracellular polymeric substance (EPS) on the hematite surface (Fig. 3D and F), which is consistent with previous results showing MtrC and OmcA in an EPS produced by cells under anaerobic conditions (19, 24). This discovery is interesting in light of the research by Rosso et al. (22) and Bose et al. (4), who found that Shewanella can implement a nonlocal electron transfer strategy to reduce the surface of hematite at locations distant from the point of cell attachment. Rosso et al. (22) proposed that the bacteria utilize unknown extracellular factors to access the most energetically favorable regions of the Fe(III) oxide surface. The Ig-AFM results (Fig. (Fig.3)3) suggest the possibility that MtrC and/or OmcA are the “unknown extracellular factors” that are synthesized by Shewanella to reduce crystalline Fe(III) oxides at points distal from the cell. Additional experiments showing reductive dissolution features coinciding with the extracellular location of MtrC and/or OmcA would need to be performed to test this hypothesis.It is important to note that these affinity maps were collected on only a few cells because it so challenging to produce large numbers of quality images. Future work should be conducted on a population of cells. Until this time, these affinity maps can be used to provide a crude, lowest-order estimate of the number of cytochromes on the outer surface of living S. oneidensis MR-1. For example, there were 236 force curves collected on the bacterium shown in Fig. Fig.3D.3D. Thirty-eight of these curves exhibited a distinct, sawtooth-shaped, antibody-antigen binding event. In other words, MtrC molecules were detected in one out of every six force curves (16%) that were collected on the cell surface.This probability can be compared to other independent studies that estimated the density and size of MtrC and OmcA molecules from S. oneidensis MR-1. Lower et al. (16) estimated that S. oneidensis has 4 × 1015 to 7 × 1015 cytochromes per square meter by comparing AFM measurements for whole cells to force curves on purified MtrC and OmcA molecules. Wigginton et al. (25) used scanning tunneling microscopy to determine that the diameter of an individual cytochrome is 5 to 8 nm. These values can be used to create a simple, geometric, close-packing arrangement of MtrC or OmcA molecules on a surface. Using this approach, cytochromes could occupy 8 to 34% of the cell surface.This estimate is consistent with the observed number of putative MtrC molecules shown in Fig. Fig.3D.3D. Therefore, it appears that these affinity maps can be used as a lowest-order estimate for the number of cytochromes on S. oneidensis MR-1 even though we do not know a priori the exact configuration of the antibody tip (e.g., the concentration of antibody on the tip, the exact shape of the tip, the binding epitopes within the antibody).In summary, the data presented here show that S. oneidensis MR-1 localizes OmcA and MtrC molecules to the exterior cell surface, including an EPS, when Fe(III) is the TEA. Here, the cytochromes presumably serve as terminal reductases that catalyze the reduction of Fe(III) through direct contact with the extracellular iron-oxide mineral.  相似文献   

6.
7.
Paenibacillus macerans is one of the species with the broadest metabolic capabilities in the genus Paenibacillus, able to ferment hexoses, deoxyhexoses, pentoses, cellulose, and hemicellulose. However, little is known about glycerol metabolism in this organism, and some studies have reported that glycerol is not fermented. Despite these reports, we found that several P. macerans strains are capable of anaerobic fermentation of glycerol. One of these strains, P. macerans N234A, grew fermentatively on glycerol at a maximum specific growth rate of 0.40 h−1 and was chosen for further characterization. The use of [U-13C]glycerol and further analysis of extracellular metabolites and proteinogenic amino acids via nuclear magnetic resonance (NMR) spectroscopy allowed identification of ethanol, formate, acetate, succinate, and 1,2-propanediol (1,2-PDO) as fermentation products and demonstrated that glycerol is incorporated into cellular components. A medium formulation with low concentrations of potassium and phosphate, cultivation at acidic pH, and the use of a CO2-enriched atmosphere stimulated glycerol fermentation and are proposed to be environmental determinants of this process. The pathways involved in glycerol utilization and synthesis of fermentation products were identified using NMR spectroscopy in combination with enzyme assays. Based on these studies, the synthesis of ethanol and 1,2-PDO is proposed to be a metabolic determinant of glycerol fermentation in P. macerans N234A. Conversion of glycerol to ethanol fulfills energy requirements by generating one molecule of ATP per molecule of ethanol synthesized. Conversion of glycerol to 1,2-PDO results in the consumption of reducing equivalents, thus facilitating redox balance. Given the availability, low price, and high degree of reduction of glycerol, the high metabolic rates exhibited by P. macerans N234A are of paramount importance for the production of fuels and chemicals.Although many microorganisms can metabolize glycerol in the presence of external electron acceptors (respiratory metabolism), few are able to do so fermentatively (i.e., in the absence of electron acceptors). Fermentative metabolism of glycerol has been reported in species of the genera Klebsiella, Citrobacter, Enterobacter, Clostridium, Lactobacillus, Bacillus, Propionibacterium, and Anaerobiospirillum but has been studied more extensively in a few species of the family Enterobacteriaceae, namely, Citrobacter freundii and Klebsiella pneumoniae (6, 9). Glycerol fermentation in these organisms is mediated by a two-branch pathway, which results in the synthesis of the glycolytic intermediate dihydroxyacetone (DHA) phosphate (DHAP) and the fermentation product 1,3-propanediol (1,3-PDO) (Fig. (Fig.1A)1A) (6). In the oxidative branch, glycerol is dehydrogenated to DHA by a type I NAD-linked glycerol dehydrogenase (glyDH). DHA is then phosphorylated by ATP- or phosphoenolpyruvate (PEP)-dependent DHA kinases (DHAKs) to generate DHAP. In the parallel reductive branch, glycerol is dehydrated by glycerol dehydratase, and 3-hydroxypropionaldehyde (3-HPA) is formed. 3-HPA is then reduced to the major fermentation product 1,3-PDO by an NADH-linked 1,3-PDO dehydrogenase (1,3-PDODH), thereby regenerating NAD+ (Fig. (Fig.1A).1A). Organisms that lack the capacity to synthesize 1,3-PDO have been deemed unable to utilize glycerol in a fermentative manner (6, 9, 10). The metabolism of glycerol in these organisms is thought to require an electron acceptor and takes place through a respiratory pathway that involves a glycerol kinase and two respiratory (aerobic and anaerobic) glycerol-3-phosphate dehydrogenases (G3PDHs) (6, 7, 24, 29, 35, 38) (Fig. (Fig.1B).1B). A recent development in this area is the finding that Escherichia coli, an organism that is unable to produce 1,3-PDO, can indeed ferment glycerol in the absence of external electron acceptors (15, 26). In this model, synthesis of the fermentation products 1,2-PDO and ethanol enables glycerol fermentation by facilitating redox balance and ATP generation, respectively (Fig. (Fig.1C)1C) (15). A type II glyDH and a PEP-dependent DHAK mediate the conversion of glycerol to glycolytic intermediates. glyDH also catalyzes the last step in the synthesis of the key fermentation product 1,2-PDO (Fig. (Fig.1C1C).Open in a separate windowFIG. 1.Glycerol metabolism in bacteria. (A) 1,3-PDO model for the fermentative utilization of glycerol. (B) Respiratory metabolism of glycerol (i.e., metabolism in the presence of an electron acceptor). (C) 1,2-PDO-ethanol model for the fermentative utilization of glycerol. Dashed lines indicate multiple steps. glyD, glycerol dehydratase; glyDH-I, type I glyDH; GK, glycerol kinase; ae-G3PDH, aerobic G3PDH; an-G3PDH, anaerobic G3PDH; QH2, reduced quinone; glyDH-II, type II glyDH; FHL, formate hydrogen lyase; ADH, alcohol/acetaldehyde dehydrogenase.Paenibacillus macerans, previously called Bacillus macerans and Bacillus acetoethylicum, is a gram-positive, spore-forming bacterium belonging to the genus Paenibacillus (17) that is capable of fermentative metabolism of hexoses, deoxyhexoses, pentoses, cellulose, and hemicellulose (33, 39, 40, 41). Glycerol, however, is considered a nonfermentable carbon source for P. macerans. The “nonfermentable status” of glycerol has been used to determine whether certain electron acceptors, such as fumarate, trimethylamine N-oxide, nitrate, and nitrite, can mediate anaerobic respiration in this organism (34).In this study we found that several P. macerans strains are able to ferment glycerol in the absence of external electron acceptors. The fermentation of glycerol by one of these strains, P. macerans N234A, occurred at high metabolic rates and in the absence of an active 1,3-PDO pathway. Therefore, the environmental and metabolic determinants of glycerol fermentation in P. macerans N234A were investigated.  相似文献   

8.
9.
The amino-terminal 290 residues of UL44, the presumed processivity factor of human cytomegalovirus DNA polymerase, possess all of the established biochemical activities of the full-length protein, while the carboxy-terminal 143 residues contain a nuclear localization signal (NLS). We found that although the amino-terminal domain was sufficient for origin-dependent synthesis in a transient-transfection assay, the carboxy-terminal segment was crucial for virus replication and for the formation of DNA replication compartments in infected cells, even when this segment was replaced with a simian virus 40 NLS that ensured nuclear localization. Our results suggest a role for this segment in viral DNA synthesis.Human cytomegalovirus (HCMV) encodes a DNA polymerase which is composed of two subunits, UL54, the catalytic subunit, and UL44, an accessory protein (8, 12, 21). UL44 can be divided into two regions, a 290-residue amino (N)-terminal domain and a 143-residue carboxy (C)-terminal segment. The overall fold of the N-terminal domain is markedly similar to that of processivity factors such as herpes simplex virus type 1 (HSV-1) UL42 and eukaryotic proliferating cell nuclear antigen (6, 22, 41), which function to tether catalytic subunits to DNA to ensure long-chain DNA synthesis. In vitro, the N-terminal domain of UL44 is sufficient for all of the established biochemical activities of full-length UL44, including dimerization, binding to double-stranded DNA, interaction with UL54, and stimulation of long-chain DNA synthesis, consistent with a role as a processivity factor (4, 5, 8, 11, 23, 24, 39). In contrast, little is known about the functions of the C-terminal segment of UL44 other than its having been reported from transfection experiments to be important for downregulation of transactivation of a non-HCMV promoter (7) and to contain a nuclear localization signal (NLS) (3). Neither the importance of this NLS nor the role of the entire C-terminal segment has been investigated in HCMV-infected cells.We first examined whether the N-terminal domain is sufficient to support DNA synthesis from HCMV oriLyt in cells using a previously described cotransfection-replication assay (27, 28). A DpnI-resistant fragment, indicative of oriLyt-dependent DNA synthesis, was detected in the presence of wild-type (WT) UL44 (pSI-UL44) (34) and in the presence of the UL44 N-terminal domain (pSI-UL44ΔC290), but not in the presence of UL44-F121A (6, 34), a mutant form previously shown not to support oriLyt-dependent DNA synthesis (34) (Fig. (Fig.1A).1A). Thus, the N-terminal domain alone is sufficient to support oriLyt-dependent DNA synthesis in a transient-transfection assay.Open in a separate windowFIG. 1.Effects of UL44 C-terminal truncations in various assays. (A) HFF cells were cotransfected with the pSP50 plasmid (containing the oriLyt DNA replication origin), a plasmid expressing WT or mutant UL44 (as indicated at the top of the panel), and plasmids expressing all of the other essential HCMV DNA replication proteins. At 5 days posttransfection, total DNA was extracted and cleaved with DpnI to digest unreplicated DNA and a Southern blot assay was performed to detect replicated pSP50. An arrow indicates DpnI-resistant, newly synthesized pSP50 fragments. (B) FLAG-tagged constructs analyzed in panel C are cartooned as horizontal bars. The names of the constructs are above the bars. The lengths of the constructs in amino acids are indicated by the scale at the bottom of the panel. The positions of residues required but not necessarily sufficient for features of the constructs are designated by shading, as indicated at the bottom of the panel. (C) Vero cells were transfected with plasmids expressing WT UL44 (parts a to c), FLAG-UL44 (parts d to f), FLAG-UL44-290stop (parts g to i), or FLAG-UL44-290NLSstop (parts j to l). At 48 h posttransfection, cells were fixed and stained with 4′,6-diamidino-2-phenylindole (DAPI) to visualize the nucleus (blue) (parts a, d, g, and j) and by IF with anti-UL44 (part b) or anti-FLAG (parts e, h, and k) and a secondary antibody conjugated with Alexa 488 (green). Parts c, f, i, and l are merged from images in the left and middle columns. Magnification: ×1,000. (D) Replication kinetics of rescued viruses. Rescued derivatives of UL44 mutant viruses (UL44-290stop-R and UL44-290NLSstop-R) or WT AD169 viruses were used to infect HFF cells at an MOI of 1 PFU/cell. The supernatants from infected cells were collected every 24 h, and viral titers were determined by plaque assays on HFF cells.These results were somewhat unexpected, as the C-terminal segment contains a functional NLS identified in transfection assays (3). We therefore assayed the intracellular localization of WT and mutant UL44 following transient transfection using pcDNA3-derived expression plasmids. Since the anti-UL44 antibodies that we have tested do not recognize the N-terminal domain of UL44, we constructed UL44 genes to encode N-terminally FLAG-tagged full-length UL44 (FLAG-UL44) or a FLAG-tagged N-terminal domain, the latter by inserting three in-frame tandem stop codons after codon 290 (FLAG-UL44-290stop, Fig. Fig.1B).1B). We also constructed a mutant form encoding a FLAG-tagged N-terminal domain, followed by the simian virus 40 (SV40) T-antigen NLS (15-17), followed by three tandem stop codons (FLAG-UL44-290NLSstop, Fig. Fig.1B).1B). Vero cells were transfected with each construct using Lipofectamine 2000, fixed with 4% formaldehyde at 48 h posttransfection, and assayed by indirect immunofluorescence (IF) using anti-UL44 (Virusys) or anti-FLAG antibody (Sigma). We observed mostly nuclear localization of WT UL44 or FLAG-UL44 with either diffuse or more localized intranuclear distribution (Fig. (Fig.1C,1C, parts a to c and d to f, respectively) and some occasional perinuclear staining, which may be due to protein overexpression. In cells expressing FLAG-UL44-290NLSstop, we observed mostly diffuse nuclear localization with little to no perinuclear staining (Fig. (Fig.1C,1C, parts j to l). In cells expressing FLAG-UL44-290stop, we observed mostly cytoplasmic staining, but with some cells exhibiting some nuclear staining (Fig. (Fig.1C,1C, parts g to i), which may explain the ability of truncated UL44 to support oriLyt-dependent DNA replication in a transient-transfection assay (Fig. (Fig.1A1A).We next investigated whether the C-terminal segment of UL44 is necessary for viral replication. We reasoned that we could investigate whether any requirement for this segment could be due to a requirement for an NLS by testing whether the SV40 NLS could substitute for the loss of the UL44 C terminus. We therefore constructed HCMV UL44 mutant viruses by introducing the UL44-290stop and UL44-290NLSstop mutations into a WT AD169 bacterial artificial chromosome (BAC) using two-step red-mediated recombination as previously described (35, 38). We also constructed the same mutants with a FLAG epitope at the N terminus of UL44 (BAC-FLAG-UL44-290stop and BAC-FLAG-UL44-290NLSstop) to monitor UL44 expression, and we constructed rescued derivatives of the mutant BACs by replacing the mutated sequences with WT UL44 sequences, as described previously (35). We introduced BACs into human foreskin fibroblast (HFF) cells using electroporation (35, 38). In several experiments using at least two independent clones for each mutant, cells electroporated with any of the mutant BACs did not exhibit any cytopathic effect (CPE) within 21 days. In contrast, within 7 to 10 days, cells electroporated with the WT AD169 BAC, a BAC expressing WT UL44 with an N-terminal FLAG tag [AD169-BACF44 (35)], or any of the rescued derivatives began displaying a CPE and yielded infectious virus. The rescued derivatives of the nontagged mutants displayed replication kinetics similar to those of the WT virus following infection at a multiplicity of infection (MOI) of 1 PFU/cell (Fig. (Fig.1D).1D). The rescued derivatives of the FLAG-tagged mutants also replicated to WT levels (data not shown). Thus, the replication defects of the mutants were due to the introduced mutations that result in truncated UL44 either with or without the SV40 NLS. We therefore conclude that the C-terminal segment of UL44 is required for viral replication.To investigate the stage of viral replication at which the UL44 C-terminal segment is important, we first assayed the subcellular localization of immediate-early proteins IE1 and IE2 and FLAG-UL44 in cells electroporated with BAC DNA expressing the FLAG-tagged WT or the two mutant UL44s using IF at 2 days postelectroporation. IE1/IE2 could be detected diffusely distributed in nuclei of cells electroporated with all three BACs (Fig. 2b, f, and j). In cells electroporated with AD169-BACF44 or BAC-FLAG-UL44-290NLSstop, FLAG-UL44 was localized largely within the nucleus (Fig. 2c and k, respectively). In contrast, in cells electroporated with BAC-FLAG-UL44-290stop, the FLAG epitope was mainly localized diffusely in the cytoplasm, with only a small amount diffusely distributed in the nucleus (Fig. (Fig.2g).2g). These data indicate that IE proteins expressed from mutant BACs are properly localized and suggest that without its C-terminal segment, which includes the NLS identified in transfection assays (3), UL44 cannot efficiently localize to the nucleus in HCMV-infected cells. However, addition of the SV40 NLS was sufficient to efficiently localize the N-terminal domain of UL44 to the nucleus. Thus, the requirement for the C-terminal segment of UL44 for viral replication is not due solely to its NLS.Open in a separate windowFIG. 2.Localization of IE1/IE2 and FLAG-UL44 proteins in electroporated cells. HFF cells were electroporated with AD169-BACF44 (panels a to d), BAC-UL44-290stop (panels e to h), or BAC-FLAG-UL44-290NLSstop (panels i to l). At 48 h posttransfection, cells were fixed and probed with anti-IE1/2 (Virusys) or anti-FLAG (Sigma). Secondary antibodies coupled to fluorophores were used for visualization of IE1/2 (anti-mouse Alexa 594; panels b, f, and j) and FLAG (anti-rabbit Alexa 488; panels c, g, and k) antibodies. DAPI was used to counterstain the nucleus (panels a, e, and i). Panels d, h, and l are merged images of the panels in the other columns. Magnification: ×1,000.We next investigated if the block in viral replication due to the loss of the C-terminal segment could be attributed to a defect in viral DNA synthesis. Cells were electroporated with AD169-BACF44 or BAC-FLAG-UL44-290NLSstop, and viral DNA accumulation was assayed by quantitative real-time PCR at various times postelectroporation (Fig. (Fig.3)3) as previously described (32, 35). In HFFs electroporated with AD169-BACF44, viral DNA began to accumulate above the input levels by 8 days postelectroporation and increased over time, with as much as a 350-fold increase over the input DNA level by 18 days postelectroporation. In contrast, levels of viral DNA in cells electroporated with BAC-UL44-290NLSstop did not increase above input levels, even by 18 days postelectroporation. These data are consistent with the notion that the UL44 C-terminal segment is required for viral DNA synthesis, although we caution that the assay did not detect DNA synthesis from AD169-BACF44 until day 8, when viral spread had likely occurred (see below).Open in a separate windowFIG. 3.Quantification of viral DNA accumulation in electroporated cells. HFF cells were electroporated with AD169-BACF44 or BAC-FLAG-UL44-290NLSstop, and total DNA was harvested on the days postelectroporation indicated. Viral DNA accumulation was assessed by real-time PCR by assessing levels of the UL83 gene and normalizing to levels of the cellular β-actin gene (32). The data are presented as the fold increase in normalized viral DNA levels over the amount of input DNA (day 1).We also analyzed the localization patterns of UL44 and UL57, the viral single-stranded DNA binding protein, which is a marker for viral DNA replication compartments (1, 2, 18, 26, 29). At 8 days postelectroporation with AD169-BACF44, UL57 and FLAG-UL44 largely colocalized within a single large intranuclear structure that likely represents a fully formed replication compartment, with some cells containing multiple smaller globular structures within the nucleus that likely represent earlier stages of replication compartments (1, 2, 29) (Fig. 4a to d). Neighboring cells also stained for UL57 and FLAG-UL44, indicative of viral spread. In contrast, in cells electroporated with BAC-FLAG-UL44-290NLSstop, UL57 (Fig. (Fig.4f)4f) was found in either punctate or small globular structures. This pattern of UL57 staining resembled that observed at very early stages of viral DNA synthesis in HCMV-infected cells, but the structures were larger and less numerous than those observed in HCMV-infected cells in the presence of a viral DNA polymerase inhibitor (2, 29). Staining for FLAG-UL44 was nuclear and largely diffuse, with some areas of more concentrated staining (Fig. (Fig.4g),4g), which could also be observed in some cells at day 2 postelectroporation (Fig. (Fig.3k).3k). This pattern of UL44 localization was generally similar to that observed in HCMV-infected cells at very early stages of infection or when HCMV DNA synthesis is blocked and also similar to the pattern in cells transfected with a UL84 null mutant BAC (2, 29, 33, 40). Importantly, little colocalization of UL57 and UL44 was observed, with areas of concentration of UL57 or UL44 occupying separate regions in the nuclei of these cells (Fig. (Fig.4h).4h). We are unaware of any other examples of this pattern of localization of these proteins in HCMV-infected cells and suggest that it may be a result of the loss of the UL44 C-terminal segment. These results indicate that this segment is important for efficient formation of viral DNA replication compartments, again consistent with a requirement for this portion of UL44 for viral DNA synthesis.Open in a separate windowFIG. 4.Localization of UL57 and FLAG-UL44 proteins in electroporated cells. HFF cells were electroporated with AD169-BACF44 (panels a to d) or BAC-FLAG-UL44-290NLSstop (panels e to h). At 8 days posttransfection, cells were fixed and then stained with antibodies specific for UL57 (Virusys) or FLAG (Sigma), followed by a secondary antibody coupled to fluorophores to detect UL57 (anti-mouse Alexa 594; panels b and f) and FLAG (anti-rabbit Alexa 488; panels c and g) antibodies. DAPI stain was used to counterstain the nucleus (panels a and e). Panels d and h are merged images of the panels in the other columns. White arrows identify punctate UL57 staining. Yellow arrows identify areas of concentration of FLAG-UL44 staining. Magnification: ×1,000.Our results, taken together, argue for a role for the C-terminal segment of UL44 in HCMV-infected cells in efficient nuclear localization of UL44 and a role in viral DNA synthesis beyond its role in nuclear localization. It is possible that this segment interacts with host or viral proteins involved in DNA replication. Of the various proteins reported to interact with UL44 (10, 19, 30, 31, 35-37), interesting candidates include the host protein nucleolin, which has been shown to associate with UL44 and be important for viral DNA synthesis (35), and the viral UL112-113 proteins, which in transfection assays were shown to recruit UL44 to early sites of DNA replication (2, 29, 33). After this paper was submitted, Kim and Ahn reported that the C-terminal segment of UL44 is necessary for interaction with a UL112-113 protein and, similar to our findings, crucial for viral replication (19). However, contrary to our findings, they reported that this segment was not necessary for efficient nuclear localization of UL44 (19). It may well be that the C-terminal segment of UL44 also has some other role later in viral replication, perhaps in gene expression, as has been suggested (7, 13, 14).A virus with a deletion of the C-terminal 150 amino acids of the HSV-1 polymerase accessory subunit UL42 displays no obvious defect in replication (9). Thus, it appears that HSV-1 and HCMV exhibit different requirements for the C-terminal segments of their respective accessory proteins. This and many other differences between these functionally and structurally orthologous proteins (5, 6, 20, 24, 25) suggest considerable selection for different features during evolution.  相似文献   

10.
A 3-hydroxypropionate/4-hydroxybutyrate cycle operates during autotrophic CO2 fixation in various members of the Crenarchaea. In this cycle, as determined using Metallosphaera sedula, malonyl-coenzyme A (malonyl-CoA) and succinyl-CoA are reductively converted via their semialdehydes to the corresponding alcohols 3-hydroxypropionate and 4-hydroxybutyrate. Here three missing oxidoreductases of this cycle were purified from M. sedula and studied. Malonic semialdehyde reductase, a member of the 3-hydroxyacyl-CoA dehydrogenase family, reduces malonic semialdehyde with NADPH to 3-hydroxypropionate. The latter compound is converted via propionyl-CoA to succinyl-CoA. Succinyl-CoA reduction to succinic semialdehyde is catalyzed by malonyl-CoA/succinyl-CoA reductase, a promiscuous NADPH-dependent enzyme that is a paralogue of aspartate semialdehyde dehydrogenase. Succinic semialdehyde is then reduced with NADPH to 4-hydroxybutyrate by succinic semialdehyde reductase, an enzyme belonging to the Zn-dependent alcohol dehydrogenase family. Genes highly similar to the Metallosphaera genes were found in other members of the Sulfolobales. Only distantly related genes were found in the genomes of autotrophic marine Crenarchaeota that may use a similar cycle in autotrophic carbon fixation.The thermoacidophilic autotrophic crenarchaeum Metallosphaera sedula uses a 3-hydroxypropionate/4-hydroxybutyrate cycle for CO2 fixation (9, 28, 29, 35) (Fig. (Fig.1).1). A similar cycle may operate in other autotrophic members of the Sulfolobales (31) and in mesophilic marine group I Crenarchaea (Cenarchaeum sp., Nitrosopumilus sp.). This cycle uses elements of the 3-hydroxypropionate cycle that was originally discovered in the phototrophic bacterium Chloroflexus aurantiacus (15, 22-25, 41, 42). It involves the carboxylation of acetyl coenzyme A (acetyl-CoA) to malonyl-CoA by a biotin-dependent acetyl-CoA carboxylase (12, 29). The carboxylation product is reduced to malonic semialdehyde by malonyl-CoA reductase (1). Malonic semialdehyde is further reduced to 3-hydroxypropionate, the characteristic intermediate of the pathway (9, 31, 35). 3-Hydroxypropionate is further reductively converted to propionyl-CoA (3), which is carboxylated to (S)-methylmalonyl-CoA by propionyl-CoA carboxylase. Only one copy of the genes encoding the acetyl-CoA/propionyl-CoA carboxylase subunits is present in most Archaea, indicating that this enzyme is a promiscuous enzyme that acts on both acetyl-CoA and propionyl-CoA (12, 29). (S)-Methylmalonyl-CoA is isomerized to (R)-methylmalonyl-CoA, which is followed by carbon rearrangement to succinyl-CoA catalyzed by coenzyme B12-dependent methylmalonyl-CoA mutase.Open in a separate windowFIG. 1.Proposed 3-hydroxypropionate/4-hydroxybutyrate cycle in M. sedula and other autotrophic Sulfolobales. Enzymes: 1, acetyl-CoA carboxylase; 2, malonyl-CoA reductase (NADPH); 3, malonate semialdehyde reductase (NADPH); 4, 3-hydroxypropionate-CoA ligase (AMP forming); 5, 3-hydroxypropionyl-CoA dehydratase; 6, acryloyl-CoA reductase (NADPH); 7, propionyl-CoA carboxylase, identical to acetyl-CoA carboxylase; 8, (S)-methylmalonyl-CoA epimerase; 9, methylmalonyl-CoA mutase; 10, succinyl-CoA reductase (NADPH), identical to malonyl-CoA reductase; 11, succinic semialdehyde reductase (NADPH); 12, 4-hydroxybutyrate-CoA ligase (AMP forming); 13, 4-hydroxybutyryl-CoA dehydratase; 14, crotonyl-CoA hydratase; 15, (S)-3-hydroxybutyryl-CoA dehydrogenase (NAD+); 16, acetoacetyl-CoA β-ketothiolase. The highlighted steps are catalyzed by the enzymes studied here.Succinyl-CoA is converted via succinic semialdehyde and 4-hydroxybutyrate to two molecules of acetyl-CoA (9), thus regenerating the starting CO2 acceptor molecule and releasing another acetyl-CoA molecule for biosynthesis. Hence, the 3-hydroxypropionate/4-hydroxybutyrate cycle (Fig. (Fig.1)1) can be divided into two parts. The first part transforms one acetyl-CoA molecule and two bicarbonate molecules into succinyl-CoA (Fig. (Fig.1,1, steps 1 to 9), and the second part converts succinyl-CoA to two acetyl-CoA molecules (Fig. (Fig.1,1, steps 10 to 16).The second part of the autotrophic cycle also occurs in the dicarboxylate/4-hydroxybutyrate cycle, which operates in autotrophic CO2 fixation in Desulfurococcales and Thermoproteales (Crenarchaea) (27, 37), raising the question of whether the enzymes in these two lineages have common roots (37). The first part of the cycle also occurs in the 3-hydroxypropionate cycle for autotrophic CO2 fixation in Chloroflexus aurantiacus and a few related green nonsulfur phototrophic bacteria (19, 22, 23, 32, 49).The two-step reduction of malonyl-CoA to 3-hydroxpropionate in Chloroflexus is catalyzed by a single bifunctional 300-kDa enzyme (30). The M. sedula malonyl-CoA reductase is completely unrelated and forms only malonic semialdehyde (1), and the enzyme catalyzing the second malonic semialdehyde reduction step that forms 3-hydroxypropionate is unknown. In the second part of the 3-hydroxypropionate/4-hydroxybutyrate cycle a similar reduction of succinyl-CoA via succinic semialdehyde to 4-hydroxybutyrate takes place. The enzymes responsible for these reactions also have not been characterized.In this work we purified the enzymes malonic semialdehyde reductase, succinyl-CoA reductase, and succinic semialdehyde reductase from M. sedula. The genes coding for these enzymes were identified in the genome, and recombinant proteins were studied in some detail. Interestingly, succinyl-CoA reductase turned out to be identical to malonyl-CoA reductase. We also show here that enzymes that are highly similar to succinyl-CoA reductase in Thermoproteus neutrophilus do not function as succinyl-CoA reductases in M. sedula.  相似文献   

11.
Strains of Salmonella enterica serovar Typhimurium LT2 lacking a functional 2-methylcitric acid cycle (2-MCC) display increased sensitivity to propionate. Previous work from our group indicated that this sensitivity to propionate is in part due to the production of 2-methylcitrate (2-MC) by the Krebs cycle enzyme citrate synthase (GltA). Here we report in vivo and in vitro data which show that a target of the 2-MC isomer produced by GltA (2-MCGltA) is fructose-1,6-bisphosphatase (FBPase), a key enzyme in gluconeogenesis. Lack of growth due to inhibition of FBPase by 2-MCGltA was overcome by increasing the level of FBPase or by micromolar amounts of glucose in the medium. We isolated an fbp allele encoding a single amino acid substitution in FBPase (S123F), which allowed a strain lacking a functional 2-MCC to grow in the presence of propionate. We show that the 2-MCGltA and the 2-MC isomer synthesized by the 2-MC synthase (PrpC; 2-MCPrpC) are not equally toxic to the cell, with 2-MCGltA being significantly more toxic than 2-MCPrpC. This difference in 2-MC toxicity is likely due to the fact that as a si-citrate synthase, GltA may produce multiple isomers of 2-MC, which we propose are not substrates for the 2-MC dehydratase (PrpD) enzyme, accumulate inside the cell, and have deleterious effects on FBPase activity. Our findings may help explain human inborn errors in propionate metabolism.Humans have used fermentation as an effective method of preservation for a wide variety of foods (41). Today, the weak short-chain fatty acids (SCFAs) produced by fermentation, such as acetic, propionic, butyric, and lactic acids, are widely used as food preservatives and in pre- and postharvest agricultural processes (34, 38, 45). Propionate, one of the most abundant SCFAs found in the environment (12), is widely used as a preservative of baked goods in the food industry (38).While SCFAs such as propionate are extensively used as food preservatives, our understanding of how microbial growth is prevented by them is incomplete. Early studies argued that growth inhibition either was caused by dissipation of the proton motive force (4, 48) or was due to decreases in intracellular pH (15, 48) or the intracellular accumulation of the propionate anion (46, 47). More recently, the global affects of SCFAs on gene expression (1, 43, 44) and protein synthesis (8, 37, 52, 56) were reported, revealing wide-ranging effects on gene expression in response to propionate in the environment (43). Evidence also suggests that central metabolic processes may be inhibited by SCFAs or their catabolites. An overview of the effects of propionate on the cell can be seen in Fig. Fig.11.Open in a separate windowFIG. 1.Overview of propionate metabolism and toxicity in Salmonella.Propionyl coenzyme A (Pr-CoA), an intermediate in propionate metabolism, was shown to inhibit pyruvate dehydrogenase in Rhodobacter sphaeroides (40) and Aspergillus niger (10) and competitively inhibit citrate synthase in Escherichia coli (39). 2-Methylcitrate (2-MC), the product of the condensation of oxaloacetate (OAA) and Pr-CoA, was shown to inhibit growth of Salmonella enterica, but the mechanism of action remained unclear (28) (Fig. (Fig.1).1). With such broad negative effects exerted by propionate or its catabolites, the best strategy for microbes to deal with SCFAs such as propionate is to efficiently catabolize them into central metabolites (Fig. (Fig.11).S. enterica, like many other enteric bacteria, is exposed to high levels of propionate in human digestive tracts with total SCFA levels varying from 20 to 300 mM and propionate reaching levels as high as 23.1 mmol/kg (9, 17). To cope with such high concentrations of propionate, this bacterium and other enterobacteria like E. coli utilize the 2-methylcitric acid cycle (2-MCC) to convert propionate to pyruvate (31, 53). In S. enterica, the prpBCDE operon encodes most of the 2-MCC enzymes (30). These genes encode a 2-methylisocitrate lyase (PrpB), a 2-methylcitrate synthase (PrpC), a 2-methylcitrate dehydratase (PrpD), and a propionyl coenzyme A (CoA) synthetase (PrpE) (Fig. (Fig.1).1). Early work with S. enterica showed that insertion elements placed within the prpBCDE operon greatly increased the sensitivity of S. enterica to propionate (23). Strains carrying insertions in prpE, however, were still able to grow on propionate and were not sensitive to propionate because acetyl-CoA synthetase (Acs) compensates for the lack of PrpE (32).The goal of the studies reported here was to identify a target of 2-MC in S. enterica. Our in vivo and in vitro data support the conclusion that 2-MC inhibits fructose-1,6-bisphosphatase (FBPase), a key enzyme of gluconeogenesis. The inhibition of FBPase blocks the synthesis of glucose, with the concomitant broad negative effects on cell function. We show that while both the 2-MC synthase (PrpC) and citrate synthase (GltA) enzymes synthesize 2-MC, the 2-MC made by GltA (2-MCGltA) is more toxic to the cell than the 2-MC made by PrpC (2-MCPrpC), and we suggest that the reason for this toxicity is due to the difference in stereochemistry of the GltA and PrpC reaction products.  相似文献   

12.
For the ornithine fermentation pathway, described more than 70 years ago, genetic and biochemical information are still incomplete. We present here the experimental identification of the last four missing genes of this metabolic pathway. They encode l-ornithine racemase, (2R,4S)-2,4-diaminopentanoate dehydrogenase, and the two subunits of 2-amino-4-ketopentanoate thiolase. While described only for the Clostridiaceae to date, this pathway is shown to be more widespread.The catabolism of ornithine by anaerobic bacteria can be accomplished through the Stickland reaction, the main chemical reaction by which Clostridium sporogenes obtains its energy (16, 17). The Stickland reaction usually involves one amino acid which acts as an electron donor while another acts as an electron acceptor, as described for Clostridium sporogenes (16, 20), Clostridium botulinum (4, 5), and Clostridium sticklandii (6, 7). However, l-ornithine, as a single substrate, is converted into both an electron donor and acceptor and metabolized in a way similar to the Stickland reaction: it is oxidized to acetate, alanine, and ammonia (oxidative pathway) and reduced to 5-aminovalerate through the formation of proline (reductive pathway) (Fig. (Fig.1).1). This study focuses on the oxidative degradation pathway, starting with the conversion of l-ornithine to the d isomer by ornithine racemase (OR) (EC 5.1.1.12) (Fig. (Fig.1,1, step 1) (2). d-Ornithine is next converted to (2R,4S)-2,4-diaminopentanoate (DAP) through the action of d-ornithine aminomutase (OA) (EC 5.4.3.5) (Fig. (Fig.1,1, step 2), an adenosylcobalamine and pyridoxal phosphate (PLP)-dependent enzyme (3, 14). DAP then undergoes a NAD+- or NADP+-dependent oxidative deamination by DAP dehydrogenase (DAPDH) (EC 1.4.1.12) (Fig. (Fig.1,1, step 3), leading to 2-amino-4-ketopentanoate (AKP) (13, 18). This compound is metabolized by AKP thiolase (AKPT), a PLP-dependent enzyme, through a thiolytic cleavage with coenzyme A (CoA) to form acetyl-CoA and alanine (Fig. (Fig.1,1, step 4) (9).Open in a separate windowFIG. 1.The ornithine fermentation pathway. Enzymes involved are OR (encoded by or-5) (EC 5.1.1.12) in step 1, OA (encoded by oraS and oraE) (EC 5.4.3.5) in step 2, DAPDH (encoded by or-1) (EC 1.4.1.12) in step 3, AKPT (encoded by or-2 and or-3) in step 4, ornithine transaminase (EC 2.6.1.13) in step 5, spontaneous in step 6, pyrroline-5-carboxylate reductase (EC 1.5.1.2) in step 7, proline racemase (EC 5.1.1.4) in step 8, d-proline reductase (EC 1.21.4.1) in step 9, and ornithine cyclodeaminase (EC 4.3.1.12) in step 10.Although the proteins of this oxidative pathway were characterized biochemically 30 years ago for C. sticklandii, only the genes corresponding to the two subunits of OA (oraS and oraE) have been identified to date (3). In this article, we present the analysis of genes which are colocalized with oraS and oraE and which are hypothesized to be involved in the conversion of l-ornithine to d-ornithine, the oxidative deamination of DAP, and the thiolytic cleavage of AKP. The proteins encoded by these genes were purified and their enzymatic activity characterized, which made it possible to reconstitute the whole oxidative branch of the l-ornithine fermentation pathway in vitro. The occurrence of this oxidative metabolic pathway in bacterial genomes which have been sequenced to date is discussed.  相似文献   

13.
A pathway toward isobutanol production previously constructed in Escherichia coli involves 2-ketoacid decarboxylase (Kdc) from Lactococcus lactis that decarboxylates 2-ketoisovalerate (KIV) to isobutyraldehyde. Here, we showed that a strain lacking Kdc is still capable of producing isobutanol. We found that acetolactate synthase from Bacillus subtilis (AlsS), which originally catalyzes the condensation of two molecules of pyruvate to form 2-acetolactate, is able to catalyze the decarboxylation of KIV like Kdc both in vivo and in vitro. Mutational studies revealed that the replacement of Q487 with amino acids with small side chains (Ala, Ser, and Gly) diminished only the decarboxylase activity but maintained the synthase activity.We have previously shown that 2-keto acids generated from amino acid biosynthesis can serve as precursors for the Ehrlich degradation pathway (15) to higher alcohols (3). In order to produce isobutanol, the valine biosynthesis pathway was used to generate 2-ketoisovalerate (KIV), the precursor to valine, which was then converted to isobutanol via a decarboxylation and reduction step (Fig. (Fig.1A).1A). The entire pathway to isobutanol from glucose is shown in Fig. Fig.1A.1A. To produce isobutanol, we overexpressed five genes, alsS (Bacillus subtilis), ilvC (Escherichia coli), ilvD (E. coli), kdc (Lactococcus lactis), and ADH2 (Saccharomyces cerevisiae) (Fig. (Fig.1A).1A). This E. coli strain produced 6.8 g/liter isobutanol in 24 h (Fig. (Fig.1B)1B) and more than 20 g/liter in 112 h (3). More recently, we have found that an alcohol dehydrogenase (Adh) encoded by yqhD on the E. coli genome can convert isobutyraldehyde to isobutanol efficiently (5) (Fig. (Fig.1B1B).Open in a separate windowFIG. 1.Schematic representation of the pathway for isobutanol production. (A) The Kdc-dependent synthetic pathway for isobutanol production. (B) Isobutanol production with the Kdc-dependent and -independent synthetic pathways. IlvC, acetohydroxy acid isomeroreductase; IlvD, dihydroxy acid dehydratase. (C) Enzymatic reaction of Als, Ahbs, and Kdc activities.One key reaction in the production of isobutanol is the conversion of KIV to isobutyraldehyde catalyzed by 2-ketoacid decarboxylase (Kdc) (Fig. (Fig.1C).1C). Since E. coli does not have Kdc, kdc from L. lactis was overexpressed. Kdc is a nonoxidative thiamine PPi (TPP)-dependent enzyme and is relatively rare in bacteria, being more frequently found in plants, yeasts, and fungi (8, 19). Several enzymes with Kdc activity have been found, including pyruvate decarboxylase, phenylpyruvate decarboxylase (18), branched-chain Kdc (8, 19), 2-ketoglutarate decarboxylase (10, 17, 20), and indole-3-pyruvate decarboxylase (13).In this work, unexpectedly, we find that Kdc is nonessential for E. coli to produce isobutanol (Fig. (Fig.1).1). An E. coli strain overexpressing only alsS (from B. subtilis), ilvC, and ilvD (both from E. coli) is still able to produce isobutanol. Since E. coli is not a natural producer of isobutanol, it cannot be detected from the culture media in any unmodified strain. We identify that AlsS from B. subtilis, which was introduced in E. coli for acetolactate synthesis (Als), catalyzes the decarboxylation of 2-ketoisovalerate like Kdc both in vivo and in vitro. AlsS is part of the acetoin synthesis pathway and catalyzes the aldo condensation of two molecules of pyruvate to 2-acetolactate (Als activity) (Fig. (Fig.1C)1C) (11). The overall reaction catalyzed by AlsS is irreversible because of CO2 evolution. The first step in catalysis is the ionized thiazolium ring of TPP reacting with the first pyruvate, followed by decarboxylation. This intermediate then reacts with the second pyruvate. Deprotonation followed by C-C bond breakage produces 2-acetolactate. In this work, mutational approaches were used to assess the importance of Q487 in the Kdc activity of AlsS.  相似文献   

14.
15.
Heat activates the dormant spores of certain Bacillus spp., which is reflected in the “activation shoulder” in their survival curves. At the same time, heat also inactivates the already active and just activated spores, as well as those still dormant. A stochastic model based on progressively changing probabilities of activation and inactivation can describe this phenomenon. The model is presented in a fully probabilistic discrete form for individual and small groups of spores and as a semicontinuous deterministic model for large spore populations. The same underlying algorithm applies to both isothermal and dynamic heat treatments. Its construction does not require the assumption of the activation and inactivation kinetics or knowledge of their biophysical and biochemical mechanisms. A simplified version of the semicontinuous model was used to simulate survival curves with the activation shoulder that are reminiscent of experimental curves reported in the literature. The model is not intended to replace current models to predict dynamic inactivation but only to offer a conceptual alternative to their interpretation. Nevertheless, by linking the survival curve''s shape to probabilities of events at the individual spore level, the model explains, and can be used to simulate, the irregular activation and survival patterns of individual and small groups of spores, which might be involved in food poisoning and spoilage.Heat inactivation kinetics of bacterial spores is a well-researched field. Much of the work on its relation to foods has focused on the heat-resistant spores of Clostridia, particularly those of Clostridium botulinum, which to this date serves as the reference organism in sterility calculations of low-acid foods (8, 32). The thermal resistance of Bacilli spores, although also extensively studied, has received less attention in the literature on food preservation. This is primarily because they are unlikely to germinate and produce cells that will survive and divide under the anaerobic conditions in a sterilized food container. Yet the mere possibility of viable Bacillus spores being present in processed foods has become an issue of food safety and a security concern. For this reason, there is a renewed interest in these spores'' heat resistance (2, 3, 6, 7, 16, 30). One of the peculiarities of certain Bacillus spores, like those of Bacillus sporothermodurans or Bacillus stearothermophilus, is that many of them can remain dormant unless activated by heat. The result is a survival curve that exhibits an “activation shoulder,” as shown schematically in Fig. Fig.11 and with published data in Fig. Fig.2.2. Thus, modeling this survival pattern, where the number of spores initially grows rather than declines, must account for the heat''s dual role of being a lethal agent and activator at the same time.Open in a separate windowFIG. 1.A schematic view of a survival curve having an activation shoulder. S(t) is the ratio between the number N(t) of viable spores at time t and the initial number N0. Notice the discrepancy between the two ways to estimate the number of dormant spores, represented by the dashed and dotted gray lines.Open in a separate windowFIG. 2.Demonstration of the fit of equation 1 (solid line) and equation 2 (dashed line) to survival curves of B. stearothermophilus spores at two temperatures. Notice the postpeak concavity of the curves. In such cases, the estimated number of dormant spores reached by the tangent method will depend on the experiment duration. The original experimental data are from Sapru et al. (25).Traditionally, the thermal inactivation of both Clostridia and Bacilli spores has been thought to follow first-order kinetics (9, 12, 31), an assumption that has been frequently challenged in recent years (18, 21, 33, 35). The most publicized model of the simultaneous heat activation and inactivation of Bacillus spores in food is that proposed by Sapru et al. (24, 25), which is an improved version of models proposed earlier by Shull et al. (29) and Rodriguez et al. (23). All of these authors and others (1, 17) assumed that the activation of dormant spores follows first-order kinetics and so does their inactivation before and after activation. The temperature dependence of the corresponding exponential rate constants was assumed to follow the Arrhenius equation.Peleg (18, 20) and van Boekel (33, 35) have shown that none of the above assumptions was necessary and that the same survival data on Bacillus stearothermophilus reported by Sapru et al. (25) and other investigators (5) can be described by different kinds of alternative four-parameter empirical models, which have a slightly better fit. This was evident not only visually (Fig. (Fig.2)2) but also as judged by statistical criteria (34). Fig. Fig.22 shows the fit of the “double Weibullian” model proposed by van Boekel (33). It has the following form: (1) where S(t) = N(t)/N0 is the survival/activation ratio, N0 and N(t) are the initial and momentary number of countable spores, respectively, and b1, b2, n1, and n2 are adjustable temperature-dependent constants. Figure 2 also shows the fit of an ad hoc empirical model, a hybrid between the double Weibullian model and one previously proposed (20) that can be written in the following form: (2) or (3) where a1, b1, tc2, and m2 are adjustable temperature-dependent parameters. According to this model, a1 is the asymptote of the first term on the right, b1 is a time characteristic of the activation, tc2 is a characteristic time of the inactivation, and m2 is a parameter that represents the curve''s postpeak concavity. The structure of equation 2 or 3 dictates that the number of dormant spores must be finite and cannot exceed N0 × 10a1, if the logarithm is base 10, or N0 × exp(a1), if it is base e. (A demonstration that generates realistic-looking activation/inactivation curves using equation 3 as a model is available from Wolfram Research [http://demonstrations.wolfram.com/SurvivalCurvesOfBacilliSporesWithAnActivationShoulder/].)Corradini and Peleg (5) proposed a way to estimate the initial number of dormant spores from survival curves having an activation shoulder using a similar model, which was originally described in Peleg (20). They suggested that the intersection of a tangent to the survival curve drawn at its postpeak region with the time axis (Fig. (Fig.1)1) is not a recommended method to estimate the number of dormant spores and that it can render unrealistically high values if used. Also, where there is no evidence that the survival curve in the postpeak region ever becomes a straight line; the same survival curve will yield different estimates of the dormant spores'' initial number depending on the experiment''s duration. Moreover, if in the postpeak region the survival ratio drop rate progressively increases, as it most probably does (Fig. (Fig.2)2) (20, 33), then the number of dormant spores estimated by the tangent extrapolation method will grow indefinitely, despite the fact that it must be finite (1). Also, since the exponential inactivation rate can be a function of time as well as of temperature, the applicability of the Arrhenius equation as a secondary model might come into question. The same can also be said about the log-linearity of the D value''s temperature dependence if used instead of the Arrhenius equation.The question that arises in light of all the above is whether one can construct a conceptual population dynamic model of the activation/inactivation of spores without assuming any fixed kinetic order. The biochemical and biophysical mechanisms that govern bacterial spore germination, activation, and inactivation have been thoroughly investigated (11, 13, 14, 15, 22, 26-28). Still, it is not clear how processes within an individual spore can be translated into activation and survival patterns at the population level and how their manifestation can be expressed in a mathematical model. Whenever a system has inherent variability and knowledge of its working is incomplete or merely insufficient to develop a model from basic principles, one can, and sometimes must, resort to a probabilistic modeling approach. The general objective of this work has been to explore the merits and limitations of this option by developing a stochastic model of Bacilli spores'' heat activation and inactivation and examining its properties. The goal has not been to develop a new method to predict the spores'' survival under dynamic conditions—rate versions of the existing empirical models such as equation 1, 2, or 3 seem to be quite suitable for that—but to offer an alternative interpretation of the patterns reported and discussed in the literature.  相似文献   

16.
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   

17.
18.
19.
The Rhizobium etli CE3 O antigen is a fixed-length heteropolymer with O methylation being the predominant type of sugar modification. There are two O-methylated residues that occur, on average, once per complete O antigen: a multiply O-methylated terminal fucose and 2-O methylation of a fucose residue within a repeating unit. The amount of the methylated terminal fucose decreases and the amount of 2-O-methylfucose increases when bacteria are grown in the presence of the host plant, Phaseolus vulgaris, or its seed exudates. Insertion mutagenesis was used to identify open reading frames required for the presence of these O-methylated residues. The presence of the methylated terminal fucose required genes wreA, wreB, wreC, wreD, and wreF, whereas 2-O methylation of internal fucoses required the methyltransferase domain of bifunctional gene wreM. Mutants lacking only the methylated terminal fucose, lacking only 2-O methylation, or lacking both the methylated terminal fucose and 2-O methylation exhibited no other lipopolysaccharide structural defects. Thus, neither of these decorations is required for normal O-antigen length, transport, or assembly into the final lipopolysaccharide. This is in contrast to certain enteric bacteria in which the absence of a terminal decoration severely affects O-antigen length and transport. R. etli mutants lacking only the methylated terminal fucose were not altered in symbiosis with host Phaseolus vulgaris, whereas mutants lacking only 2-O-methylfucose exhibited a delay in nodule development during symbiosis. These results support previous conclusions that the methylated terminal fucose is dispensable for symbiosis, whereas 2-O methylation of internal fucoses somehow facilitates early events in symbiosis.O antigens typically constitute the distal portions of lipopolysaccharides (LPS) and help determine the diverse surface characteristics of Gram-negative bacteria. These repeat unit carbohydrate polymers vary tremendously in structure and, as a family, they exhibit all known sugars and sugar modifications, linked in myriad ways forming homopolymers and heteropolymers. Control of polymer length also varies, allowing highly uniform to completely random lengths. Great diversity of O-antigen structures even within a species is well known. Moreover, O antigens of a single strain can vary according to growth and environmental conditions. One such condition is the presence of a multicellular host (5, 18, 36, 40, 42, 44).Rhizobium etli CE3 fixes nitrogen inside root nodules it incites on the common bean Phaseolus vulgaris. The O antigen of its LPS (Fig. (Fig.1)1) is essential for bacterial infection during development of this symbiosis (41). In addition, at least two alterations occur in the O antigen when R. etli CE3 is grown in the presence of either the host plant or plant exudates. The content of the multiply O-methylated terminal fucose is decreased (19, 44), whereas the 2-O methylation of internal fucoses (2OMeFuc) increases twofold (Fig. (Fig.1)1) (15, 44). In addition to the multiply O-methylated terminal fucose and 2OMeFuc, methylation occurs always on 6-deoxytalose and likely on glucuronic acid to yield 3-O-methyl-6-deoxytalose (3OMe6dTal) and methyl-esterified glucuronyl (MeGlcA) residues (Fig. (Fig.1)1) (22); however, the incidence of these methylations is not known to vary with growth condition. The genetics responsible for the variable O methylations and the additions of the residues they modify have not been elucidated.Open in a separate windowFIG. 1.R. etli CE3 O-antigen structure (22). The portion of the LPS conceptually defined as O antigen begins with N-acetyl-quinovosamine (QuiNAc) at the reducing end followed by a mannose (Man) residue and a fucose (Fuc) residue. Attached to this fucose is the repeating unit consisting of one fucose residue, one 3-O-methyl-6-deoxytalose residue (3OMe6dTal), and one glucuronyl methyl ester residue (MeGlcA). The sugars of the repeating unit are added sequentially exactly five times (in most molecules). An O-acetyl group is present in each of the repeating units, but its location is unknown at this time. Growth in TY culture results in one 2-O-methylfucose (2OMeFuc) per O antigen on average (22). The O-antigen backbone is capped with a 2,3-di-O-methylfucose (referred to as the terminal residue in this report) on which additional O methylation at the 4-position is variable as indicated by parentheses. Growth of the bacteria in the presence of the host plant or plant exudates induces the increase of 2-O methylation of internal fucose (2OMeFuc) residues and decreased relative amount of the terminal residue (44).Most mutations affecting the known R. etli CE3 O-antigen structure map to a 28-kb genetic cluster on the chromosome (Fig. (Fig.2)2) (previously referred to as lps region α [8, 19, 37, 40, 45]). Genes and mutations within this cluster previously have been given the designations lps (9) and lpe (19). Recently, the new designation wre has been sanctioned by the Bacterial Polysaccharide Gene Database for this genetic cluster and other genes specifically devoted to the R. etli CE3 O antigen, in keeping with the system of nomenclature for bacterial polysaccharide genes (47).Open in a separate windowFIG. 2.R. etli CE3 O-antigen genetic cluster. (A) The R. etli CE3 chromosomal O antigen genetic cluster spans nucleotides 784527 to 812262 of the genome sequence (28) and consists of 25 putative ORFs. ORFs relevant to the present study are enlarged, and the relative locations of mutations are indicated. White triangles indicate mutations created by insertion of antibiotic cassettes, and black triangles indicate mutations created by Tn5 mutagenesis. The strain numbers carrying these mutations are indicated above the triangles. (B) The solid bars represent the extents of R. etli CE3 DNA cloned for complementation analysis. The scale and positions match those of the lower map in panel A.Duelli et al. (19) identified a 3-kb genetic locus that is required for the presence of the 2,3-di-O-methylfucose or 2,3,4-tri-O-methylfucose at the terminus of the O antigen. Now known to be near one end of the O-antigen genetic cluster (Fig. (Fig.2),2), the DNA sequence reported by Duelli et al. encompasses nucleotides 807701 to 810147 of the subsequently determined genome sequence (28). Sequence and annotation of the 3-kb locus have since been revised. In place of the four open reading frames (ORFs) suggested previously (19), the current annotation predicts two ORFs: wreA and wreC (Fig. (Fig.2).2). The wreA ORF is predicted to encode a methyltransferase (19), but the predicted WreC polypeptide sequence matches no known methyltransferase or glycosyltransferase or any other polypeptide sequence in the database (Fig. (Fig.3).3). When it became clear that this locus was part of the larger O-antigen genetic cluster, the nucleotide sequence suggested that three genes contiguous to wreA also might encode functions needed for synthesis and addition of the terminal fucose. The results to be shown bore out predictions of this hypothesis.Open in a separate windowFIG. 3.Conserved domain predictions. Spanning nucleotides 804817 to 810147 of the genome sequence (28), ORFs RHE_CH00766, RHE_CH00767, RHE_CH00768, RHE_CH00769, and RHE_CH00770 were named wreB, wreD, wreF, wreA, and wreC, respectively. Previously, wreF, wreA, and wreC were referred to as nlpe2, lpeA, and nlpe1, respectively (19). ORF RHE_CH00755, spanning nucleotides 791286 to 794093, was named wreM. Predicted positions of conserved domains are indicated by amino acid positions. Abbreviations: GT, conserved glycosyltransferase domain; MT, conserved methyltransferase domain. Gray boxes indicate the predicted transmembrane domains.The gene responsible for the other conditionally variable O-antigen methylation, the 2-O methylation of internal fucose residues (2OMeFuc), had not been identified in prior published work. However, among mutants isolated by random Tn5 mutagenesis, a few had been shown to lack 2OMeFuc entirely (44). We show here that the transposon insertions were located in the bifunctional gene wreM. Furthermore, results of directed insertion mutagenesis confirm two separate enzymatic domains encoded by this gene, with the α domain being required for the 2-O methylation activity and mutation of the other domain resulting in a truncated O antigen. Mutants from the directed mutagenesis that appeared to have no LPS defects other than the lack of 2OMeFuc served as tools to assess the importance of just this structural feature in the symbiosis with P. vulgaris.  相似文献   

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