首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Here, the class I polyhydroxyalkanoate synthase (PhaC) from Ralstonia eutropha was investigated regarding the functionality of its conserved C-terminal region and its ability to tolerate translational fusions to its C terminus. MalE, the maltose binding protein, and green fluorescent protein (GFP) were considered reporter proteins to be translationally fused to the C terminus. Interestingly, PhaC remained active only when a linker was inserted between PhaC and MalE, whereas MalE was not functional. However, the extension of the PhaC N terminus by 458 amino acid residues was required to achieve a functionality of MalE. These data suggested a positive interaction of the extended N terminus with the C terminus. To assess whether a linker and/or N-terminal extension is generally required for a functional C-terminal fusion, GFP was fused to the C terminus of PhaC. Both fusion partners were active without the requirement of a linker and/or N-terminal extension. A further reporter protein, the immunoglobulin G binding ZZ domain of protein A, was translationally fused to the N terminus of the fusion protein PhaC-GFP and resulted in a tripartite fusion protein mediating the production of polyester granules displaying two functional protein domains.Polyhydroxyalkanoates (PHAs) are biopolyesters synthesized by many bacteria and some archaea in times of unbalanced nutrient availability (7, 14-16, 22). These polyesters are stored as water-insoluble inclusions inside the cells and serve as energy and carbon storage (11, 29, 30). PHA synthases catalyze the stereoselective conversion of (R)-3-hydroxyacyl-coenzyme A (CoA) to PHAs while CoA is released and intracellular PHA granules are formed (32). The PHA synthase remains covalently attached to the PHA granule surface and has been targeted by protein engineering, i.e., translational fusion to the dispensable and variable N terminus, to enable the display of various protein functions without affecting the synthase activity (8, 26). PHA granules displaying certain functionalities have been considered as biobeads for biotechnological and medical applications (11).PHA synthases can be divided into four classes. Class I and class II enzymes consist of only one subunit (PhaC) (28) and produce short-chain-length PHAs (class I) or medium-chain-length PHAs (class II), respectively (30, 33). Polyester synthases belonging to class III consist of two subunits, PhaC and PhaE, and produce short-chain-length PHAs (20, 21). Class IV PHA synthases are similar to enzymes belonging to class III. The synthases of this class comprise the two subunits PhaC and PhaR (23, 24).It was previously shown that the N terminus of PhaC is a highly variable region and not essential for PHA synthase activity (30, 35). In contrast, the C terminus is a rather conserved region among class I and class II PHA synthases and is essential for enzyme activity (31). Alignments of the amino acid sequences of different PHA synthases revealed that the C terminus of these enzymes is hydrophobic and was therefore suggested to interact with the hydrophobic core of PHA granules (30). The PhaC subunits of class III and class IV PHA synthases do not show a high hydrophobicity for their C- terminal regions. Previous studies showed that the PhaC subunit of the class IV PHA synthase from Bacillus megaterium tolerates fusions to its C terminus without a loss in activity as long as the hydrophobic second subunit, PhaR, is present as well (23).The aim of this study was to assess the effect of the conserved hydrophobic C terminus of PhaC on enzyme activity with regard to the possibility of translationally fusing protein functions for display at the PHA granule surface. This will be of interest for the display of proteins that require their free C terminus for activity.  相似文献   

2.
Subunit b, the peripheral stalk of bacterial F1Fo ATP synthases, is composed of a membrane-spanning and a soluble part. The soluble part is divided into tether, dimerization, and δ-binding domains. The first solution structure of b30-82, including the tether region and part of the dimerization domain, has been solved by nuclear magnetic resonance, revealing an α-helix between residues 39 and 72. In the solution structure, b30-82 has a length of 48.07 Å. The surface charge distribution of b30-82 shows one side with a hydrophobic surface pattern, formed by alanine residues. Alanine residues 61, 68, 70, and 72 were replaced by single cysteines in the soluble part of subunit b, b22-156. The cysteines at positions 61, 68, and 72 showed disulfide formation. In contrast, no cross-link could be formed for the A70C mutant. The patterns of disulfide bonding, together with the circular dichroism spectroscopy data, are indicative of an adjacent arrangement of residues 61, 68, and 72 in both α-helices in b22-156.ATP synthesis by oxidative phosphorylation or photophosphorylation is a multistep membrane-located process that provides the bulk of cellular energy in eukaryotes and many prokaryotes. The majority of ATP synthesis is accomplished by the enzyme ATP synthase (EC 3.6.1.34), also called F1Fo ATP synthase, which, in its simplest form, as in bacteria, is composed of eight different subunits (α3, β3, γ, δ, ɛ, a, b2, and c9-12). This multisubunit complex is divided into the F1 headpiece, α33, and a membrane-embedded ion-translocating part known as Fo, to which F1 is attached by a central and a peripheral stalk (1, 5, 25). ATP is synthesized or hydrolyzed on the α33 hexamer, and the energy provided for or released during that process is transmitted to the membrane-bound Fo sector, consisting of subunits a and c and part of subunit b (30, 31). The energy coupling between the two active domains occurs via the stalk part(s) (6). The central stalk is made of subunits γ and ɛ, and the peripheral stalk is formed by subunits δ and b. The peripheral stalk, which lies at the edge of the multisubunit assembly of the F1Fo ATP synthase, acts as a stator to counter the tendency of the α33 hexamer to follow the rotation of the central stalk and the attached c-ring, and to anchor the membrane-embedded a subunit (17, 36).In Escherichia coli, subunit b with its 156 residues extends with its soluble part (bsol; b21-156) from the top of the F1 sector down, into, and across the membrane, where it is associated with subunit a (2, 15, 32, 34). The 156-residue b subunit has been divided into four functional domains (28). They are, in order from the N to the C terminus; the membrane domain, the tether region, the dimerization domain, and the δ-binding domain. The structure of the synthesized 33-residue peptide comprising the N-terminal membrane-spanning region has been solved by 1H NMR, showing an α-helical feature (14). The crystallographic structure of the major part of the dimerization domain, b62-122, revealed an α-helix with a length of 9.0 nm (12). Most recently, the NMR solution structure of the very C-terminal segment, b140-156, which interacts with the C terminus of subunit δ (δ91-177), has been determined by NMR spectroscopy (26). This molecule adopts a stable helix formation in solution with a flexible tail between amino acid residues 140 and 145. SAXS (26) and analytical ultracentrifuge studies have indicated that the soluble domain of subunit b (b21-156, b22-156) is dimeric in solution (12). So far, no high-resolution structure of the tether domain, including residues 25 to 52, or the N-terminal segment of the dimerization domain, which is formed by residues 53 to 122, is available (14).Here, we have turned our attention to the production and purification of residues 30 to 82 of subunit b (b30-82) from E. coli F1Fo ATP synthase, which forms the remaining unsolved structural segment of subunit b. The structural features of this segment have been determined in solution using NMR spectroscopy. The introduction of a cysteine residue into b22-156 at four positions resulted in different intersubunit disulfide patterns, giving insight into the proximity of the residues.  相似文献   

3.
Linalool production was evaluated in different Saccharomyces cerevisiae strains expressing the Clarkia breweri linalool synthase gene (LIS). The wine strain T73 was shown to produce higher levels of linalool than conventional laboratory strains (i.e., almost three times the amount). The performance of this strain was further enhanced by manipulating the endogenous mevalonate (MVA) pathway: deregulated overexpression of the rate-limiting 3-hydroxy-3-methylglutaryl coenzyme A reductase (HMG-CoA reductase) doubled linalool production. In a haploid laboratory strain, engineering of this key step also improved linalool yield.Monoterpenes are a class of isoprenoids of increasing industrial and clinical interest usually produced by plants. They are used as aromatic additives in the food and cosmetics industries and are also important components in wine aroma. Moreover, certain monoterpenes display antimicrobial, antiparasitic, and antiviral properties as well as a plethora of promising health benefits (for recent reviews, see references 2, 7, 15, 28, and 30 and references cited therein). To date, many studies have focused on plant metabolic engineering of monoterpene production (for selected reviews, see references 1, 14, 19, 29, and 35 and references cited therein), and few studies have been carried out on microorganisms (9, 21, 22, 34, 38). Efficient microbial production of these metabolites could provide an alternative to the current methods of chemical synthesis or extraction from natural sources. In this regard, a considerable number of studies have shown the utility of Saccharomyces cerevisiae as a valuable platform for sesquiterpene, diterpene, triterpene, and carotene production (references 5, 10, 23, 26, 30, 31, 32, and 33 and references cited therein). However, all the efforts dedicated to the improvement of isoprenoid yields in S. cerevisiae have been performed using conventional laboratory strains, and there are no studies concerning natural or industrially relevant isolates.In recent years, many genes that encode plant monoterpene synthases (MTS), a family of enzymes which specifically catalyze the conversion of the ubiquitous C10 intermediate of isoprenoid biosynthesis geranyl pyrophosphate (GPP) to monoterpenes, have been characterized. Such is the case with the LIS gene (codes for S-linalool synthase) of Clarkia breweri, the first MTS-encoding gene to be isolated (13). In contrast to plants, S. cerevisiae cannot produce monoterpenes efficiently, mainly due to the lack of specific pathways involving MTS. However, GPP is formed as a transitory intermediate in the two-step synthesis of farnesyl pyrophosphate (FPP), catalyzed by FPP synthase (FPPS) (Fig. (Fig.1),1), and some natural S. cerevisiae strains have been shown to possess the ability to produce small amounts of monoterpenes (8). Whether this occurs through unspecific dephosphorylation of a more available endogenous pool of GPP and subsequent bioconversions is not known. In addition, it has recently been established that S. cerevisiae has enough free GPP to be used by exogenous monoterpene synthases to produce monoterpenes under laboratory and vinification conditions (22, 34).Open in a separate windowFIG. 1.Simplified isoprenoid pathway in S. cerevisiae, including the branch point to linalool. Dotted arrows indicate that more than one reaction is required to convert the substrate to the product indicated. Dashed arrows indicate the engineered steps. Abbreviations: HMG-CoA, 3-hydroxy-3-methylglutaryl coenzyme A; IPP, isopentenyl pyrophosphate; GPP, geranyl pyrophosphate; FPP, farnesyl pyrophosphate; DMAPP, dimethylallyl pyrophosphate; HMGR, HMG-CoA reductase; FPPS, FPP synthase; LIS, linalool synthase.Here we present the process for selecting and optimizing yeast strains for foreign monoterpene production. We have chosen the C. breweri LIS gene as a prototype because, when heterologously expressed in S. cerevisiae, it specifically results in the production of linalool (3,7-dimethyl-1,6-octadien-3-ol; a floral scent and bioactive acyclic monoterpene identified in numerous fruits and flowers) and no other by-products (22). Two S. cerevisiae strains of different origins have been selected and their endogenous mevalonate (MVA) pathways engineered to enhance the production of linalool. These strategies might be employed to produce any other recombinant monoterpene in S. cerevisiae by expressing the appropriate monoterpene synthase.  相似文献   

4.
The chlorosome envelope of Chlorobaculum tepidum contains 10 proteins that belong to four structural motif families. A previous mutational study (N.-U. Frigaard, H. Li, K. J. Milks, and D. A. Bryant, J. Bacteriol. 186:646-653, 2004) suggested that some of these proteins might have redundant functions. Six multilocus mutants were constructed to test the effects of eliminating the proteins of the CsmC/CsmD and CsmB/CsmF motif families, and the resulting strains were characterized physiologically and biochemically. Mutants lacking all proteins of either motif family still assembled functional chlorosomes, and as measured by growth rates of the mutant strains, light harvesting was affected only at the lowest light intensities tested (9 and 32 μmol photons m−2 s−1). The size, composition, and biogenesis of the mutant chlorosomes differed from those of wild-type chlorosomes. Mutants lacking proteins of the CsmC/CsmD motif family produced smaller chlorosomes than did the wild type, and the Qy absorbance maximum for the bacteriochlorophyll c aggregates in these chlorosomes was strongly blueshifted. Conversely, the chlorosomes of mutants lacking proteins of the CsmB/CsmF motif family were larger than wild-type chlorosomes, and the Qy absorption for their bacteriochlorophyll c aggregates was redshifted. When CsmH was eliminated in addition to other proteins of either motif family, chlorosomes had smaller diameters. These data show that the chlorosome envelope proteins of the CsmB/CsmF and CsmC/CsmD families play important roles in determining chlorosome size as well as the assembly and supramolecular organization of the bacteriochlorophyll c aggregates within the chlorosome.Green sulfur bacteria (GSB; phylum Chlorobi) are obligate photolithoautotrophs that utilize chlorosomes for light harvesting (2, 13). Chlorosomes additionally occur in some green-nonsulfur bacteria, also known as filamentous anoxygenic phototrophs (phylum Chloroflexi), and in a recently discovered chlorophototrophic member of the phylum Acidobacteria, “Candidatus Chloracidobacterium thermophilum” (2, 3). Chlorosomes are the largest known light-harvesting organelles and can contain up to 250,000 bacteriochlorophyll (BChl) molecules (13, 29, 30, 39). They do not have a fixed stoichiometric ratio of the major pigment, which may be BChl c, d, or e, to any protein component, and as a result they are highly variable in size, shape, and composition. In spite of this structural heterogeneity (34), the detailed molecular and supramolecular structures of the BChls in chlorosomes of Chlorobaculum tepidum were recently solved by combining systems biology, solid-state nuclear magnetic resonance (NMR), cryo-electron microscopy, and molecular modeling (22). The fundamental structural units were found to be syn-anti monomer stacks that form coaxial nanotubes, which have a 2.1-nm spacing between the adjacent BChl layers. In addition to the major BChl species, chlorosomes contain carotenoids, isoprenoid quinones, wax esters, and a small quantity of BChl a. BChl a is known to be associated with CsmA, the most highly conserved protein in chlorosomes (13).Although the structural organization of the BChl molecules in all chlorosomes may be similar (4, 22, 25, 37, 38), with the exception of CsmA, the composition and sequences of the envelope proteins of chlorosomes of the phyla Chlorobi, Chloroflexi, and Acidobacteria are not well conserved. Blankenship (1) suggested that lateral gene transfer might have been responsible for the presence of the genes for chlorosome biogenesis among some of these three groups of bacteria. However, because chlorosomes are found in each of three, early-diverging bacterial lineages that contain chlorophototrophs, two of which additionally contain homodimeric type 1 reaction centers (2, 3), it is possible that chlorosomes represent one of the earliest types of photosynthetic antennae and were present in a common ancestor of these phyla.A protein-stabilized, glycolipid envelope surrounds the chlorosome BChls, and this membrane can be considered to be an asymmetric bilayer membrane in which glycolipids form the outer leaflet and the hydrophobic tails of BChls form the inner leaflet (13, 24, 50, 53). In C. tepidum, a genetically tractable model GSB, this envelope contains 10 proteins, which are designated CsmA, CsmB, CsmC, CsmD, CsmE, CsmF, CsmH, CsmI, CsmJ, and CsmX (6-8, 14, 47, 50). The structural organization of these proteins has been studied by cross-linking and immunoblotting, which led to a model for the organization of these proteins in the chlorosome envelope (28, 50, 53). CsmA, the only protein for which any detailed structural information is available, probably binds both BChl a and carotenoids (13, 23, 31, 35, 40) and forms a large, paracrystalline array known as the “baseplate” (8, 23, 28, 35, 36, 42). The structure for apo-CsmA in an organic solvent was recently determined by NMR spectroscopy, and a model for the structural organization of CsmA in the chlorosome baseplate of C. tepidum was proposed (35, 36).Sequence comparisons suggest that the chlorosome envelope proteins can be assigned to four motif families: 1, CsmA/CsmE; 2, CsmB/CsmF (CsmH); 3, CsmC/CsmD (CsmH); and 4, CsmI/CsmJ/CsmX (48, 50). CsmA and CsmE are 49% identical and are both synthesized as precursors, which are proteolytically processed by the removal of ∼20 amino acids at their carboxy termini to generate the mature polypeptides (7, 8). CsmB and CsmF are 29% identical and 63% similar in sequence (6, 50). Moreover, the amino-terminal domain of CsmH is related in sequence to these two proteins (50). The CsmC and CsmD proteins are 26% identical and 45% similar in sequence, and these two proteins additionally share sequence similarity to the carboxyl-terminal region of CsmH. The other three chlorosome proteins (CsmI, CsmJ, and CsmX) share some sequence similarities to the precursor forms of CsmA and CsmE in their carboxyl-terminal regions, while their amino-terminal domains are obviously related to adrenodoxin-type [2Fe-2S] ferredoxins (47-50). These sequence relationships strongly imply that gene duplication and divergence have occurred among a small number of ancestral gene types, and these observations additionally suggest that some of these proteins might be functionally redundant (47, 50). This view was supported by mutational studies that showed that only CsmA was essential for the viability of C. tepidum. Mutants lacking any other single chlorosome protein still assembled functional chlorosomes that were similar in pigment composition and functionality to those of the wild type (14).Because of the possible functional redundancy of chlorosome proteins of the different motif classes, double, triple, and quadruple mutants were constructed to study the roles of the CsmC/CsmD/CsmH and CsmB/CsmF/CsmH protein motif families in chlorosome biogenesis and structure. Mutants lacking CsmI, CsmJ, and CsmX, which form the Fe/S motif family of envelope proteins, were also constructed, and these mutants will be described in detail elsewhere (27; H. Li, N.-U. Frigaard, and D. A. Bryant, unpublished data). The results presented here show that functional chlorosomes assemble in the complete absence of proteins of the CsmC/CsmD or CsmB/CsmF motif families, but the size, shape, and composition of the resulting chlorosomes are altered. The results suggest that the chlorosome envelope proteins may also influence the structural organization of the BChls in chlorosomes and thus help to define chlorosome assembly and shape.  相似文献   

5.
6.
Lipoteichoic acid (LTA) is an important cell wall component of Gram-positive bacteria. The key enzyme responsible for polyglycerolphosphate lipoteichoic acid synthesis in the Gram-positive pathogen Staphylococcus aureus is the membrane-embedded lipoteichoic acid synthase enzyme, LtaS. It is presumed that LtaS hydrolyzes the glycerolphosphate head group of the membrane lipid phosphatidylglycerol (PG) and catalyzes the formation of the polyglycerolphosphate LTA backbone chain. Here we describe an in vitro assay for this new class of enzyme using PG with a fluorescently labeled fatty acid chain (NBD-PG) as the substrate and the recombinant soluble C-terminal enzymatic domain of LtaS (eLtaS). Thin-layer chromatography and mass spectrometry analysis of the lipid reaction products revealed that eLtaS is sufficient to cleave the glycerolphosphate head group from NBD-PG, resulting in the formation of NBD-diacylglycerol. An excess of soluble glycerolphosphate could not compete with the hydrolysis of the fluorescently labeled PG lipid substrate, in contrast to the addition of unlabeled PG. This indicates that the enzyme recognizes and binds other parts of the lipid substrate, besides the glycerolphosphate head group. Furthermore, eLtaS activity was Mn2+ ion dependent; Mg2+ and Ca2+ supported only weak enzyme activity. Addition of Zn2+ or EDTA inhibited enzyme activity even in the presence of Mn2+. The pH optimum of the enzyme was 6.5, characteristic for an enzyme that functions extracellularly. Lastly, we show that the in vitro assay can be used to study the enzyme activities of other members of the lipoteichoic acid synthase enzyme family.Lipoteichoic acid (LTA) is a crucial component of the cell wall envelope in Gram-positive bacteria. Diverse functions have been ascribed to LTA, including regulation of the activity of hydrolytic enzymes (4), an essential role in divalent cation homeostasis (2, 26, 37), and retention of noncovalently attached proteins within the cell wall envelope (20, 41). In addition, functions of LTA in host-pathogen interactions have been reported (44). d-Alanine modifications on LTA protect bacteria from killing by cationic antimicrobial peptides (36, 43) and are critical during the infection and colonization processes (1, 5, 10). On the other hand, LTA may also play a positive role for the host in wound healing, by preventing excessive inflammation (25).In the Gram-positive bacterial pathogen Staphylococcus aureus and in many other bacteria belonging to the Firmicutes, including Bacillus, Listeria, Streptococcus, Enterococcus, and Lactococcus spp., LTA is composed of a linear 1,3-linked polyglycerolphosphate backbone chain that is tethered via a glycolipid anchor to the bacterial membrane (6, 9). Recently, the staphylococcal protein LtaS was identified and shown to be responsible for polyglycerolphosphate LTA synthesis in vivo (14). An S. aureus strain depleted of LtaS is unable to synthesize LTA and shows severe growth and morphological defects (14); an S. aureus ltaS deletion strain is viable at 30°C only in a growth medium containing at least 1% NaCl or at higher temperatures at high salt (7.5%) or high sucrose (40%) concentrations (35). Taken together, these findings provide further evidence for the importance of this abundant cell envelope component for normal cell morphology and physiology.Pulse-chase experiments have provided strong biochemical evidence that the glycerolphosphate subunits of LTA are derived from the head group of the membrane lipid phosphatidylglycerol (PG) (7, 8, 12). A rapid and almost complete turnover of the nonacylated glycerolphosphate group of PG into LTA is observed in S. aureus and other Gram-positive bacteria that synthesize polyglycerolphosphate LTA (23, 24). It is assumed that the LtaS enzyme cleaves the head group of PG and uses this glycerolphosphate subunit to polymerize the LTA backbone chain.One or more LtaS-like enzymes are encoded in the genomes of Gram-positive bacteria that synthesize polyglycerolphosphate LTA (14). S. aureus LtaS and all other members of this enzyme family are predicted to contain five N-terminal transmembrane helices followed by an extracellular C-terminal enzymatic domain (eLtaS) (14, 29). The LtaS enzyme is processed in S. aureus, and the eLtaS domain is released into the culture supernatant as well as partially retained within the cell wall envelope (11, 29, 45). The crystal structure of the S. aureus eLtaS domain, alone and in a complex with soluble glycerolphosphate and the soluble domain of the Bacillus subtilis LtaS (LtaSBs) enzyme (YflE), identified a threonine as the catalytic residue. This is based on the location of the glycerolphosphate head group in the active site for S. aureus LtaS and on threonine phosphorylation in the B. subtilis enzyme structure (29, 37). Replacement of this threonine residue with an alanine renders the S. aureus enzyme inactive and unable to synthesize LTA in vivo (29). In addition, a Mn2+ ion was detected in the active center of the S. aureus LtaS structure, while the B. subtilis enzyme contained a Mg2+ ion.To provide insight into the enzymatic activity of the S. aureus lipoteichoic acid synthase enzyme, we developed an in vitro assay for this enzyme using purified recombinant eLtaS and fluorescently labeled PG as a substrate. Using thin-layer chromatography (TLC) and mass spectrometry analysis of the lipid reaction products, we show that eLtaS protein is sufficient to cleave the glycerolphosphate head group from NBD-PG, resulting in the formation of NBD-diacylglycerol (NBD-DAG). Furthermore, we provide experimental evidence that LtaS requires Mn2+ for enzyme activity, while Zn2+ inhibits enzyme function. Our results suggest that LtaS has a narrow substrate specificity, with PG serving as a substrate while phosphatidylethanolamine (PE), phosphatidylcholine (PC), and phosphatidylserine (PS) do not. Lastly, we show that this in vitro assay can be used to study the enzyme functions of other members of this protein family, such as the Listeria monocytogenes LTA synthase (LtaSLm) and LTA primase (LtaPLm) enzymes. This study is the first in vitro characterization of lipoteichoic acid synthase enzymes and an important first step towards the development of an assay to screen and identify enzyme-specific inhibitors for this new and important class of bacterial enzymes.  相似文献   

7.
Although most inbred mouse strains are highly susceptible to mouse hepatitis virus (MHV) infection, the inbred SJL line of mice is highly resistant to its infection. The principal receptor for MHV is murine CEACAM1 (mCEACAM1). Susceptible strains of mice are homozygous for the 1a allele of mCeacam1, while SJL mice are homozygous for the 1b allele. mCEACAM1a (1a) has a 10- to 100-fold-higher receptor activity than does mCEACAM1b (1b). To explore the hypothesis that MHV susceptibility is due to the different MHV receptor activities of 1a and 1b, we established a chimeric C57BL/6 mouse (cB61ba) in which a part of the N-terminal immunoglobulin (Ig)-like domain of the mCeacam1a (1a) gene, which is responsible for MHV receptor function, is replaced by the corresponding region of mCeacam1b (1b). We compared the MHV susceptibility of these chimeric mice to that of SJL and B6 mice. B6 mice that are homozygous for 1a are highly susceptible to MHV-A59 infection, with a 50% lethal dose (LD50) of 102.5 PFU, while chimeric cB61ba mice and SJL mice homozygous for 1ba and 1b, respectively, survived following inoculation with 105 PFU. Unexpectedly, cB61ba mice were more resistant to MHV-A59 infection than SJL mice as measured by virus replication in target organs, including liver and brain. No infectious virus or viral RNA was detected in the organs of cB61ba mice, while viral RNA and infectious virus were detected in target organs of SJL mice. Furthermore, SJL mice produced antiviral antibodies after MHV-A59 inoculation with 105 PFU, but cB61ba mice did not. Thus, cB61ba mice are apparently completely resistant to MHV-A59 infection, while SJL mice permit low levels of MHV-A59 virus replication during self-limited, asymptomatic infection. When expressed on cultured BHK cells, the mCEACAM1b and mCEACAM1ba proteins had similar levels of MHV-A59 receptor activity. These results strongly support the hypothesis that although alleles of mCEACAM1 are the principal determinants of mouse susceptibility to MHV-A59, other as-yet-unidentified murine genes may also play a role in susceptibility to MHV.Differences in susceptibility to a number of viral infections have been documented among inbred mouse strains (20). These differences have been studied as models for the various degrees of susceptibility of individual humans to some viral infections. Numerous host factors have been found to be involved in such differences (2, 15). For example, allelic variations in the virus receptor and coreceptor for HIV-1 are important host factors influencing susceptibility to HIV-1 infection (36).A virus receptor is a molecule with which the virus interacts at an initial step of infection. Therefore, receptors are crucial host determinants of virus susceptibility (15, 16). A variety of receptor proteins has been identified for many different viruses, including the murine coronavirus mouse hepatitis virus (MHV) (12, 50). The principal receptor for MHV is murine carcinoembryonic antigen-related cell adhesion molecule 1 (mCEACAM1; previously called Bgp or MHVR [3]), which is in the immunoglobulin (Ig) superfamily (12, 50). Four isoforms of mCEACAM1a (1a) are expressed on the plasma membranes of a variety of murine cells and tissues (14). The two mCEACAM1 isoforms with a molecular mass of 100 to 120 kDa are composed of four Ig-like ectodomains, a transmembrane (TM) domain, and either a long or a short cytoplasmic tail (Cy) (3, 22). Two other isoforms consist of two Ig-like domains, with either long or short Cy (3, 22). The N-terminal (N) domain is responsible for virus binding (10, 24), the induction of conformational changes in the viral spike protein (S), and membrane fusion during virus entry and syncytium formation (13, 24). The replacement of the N-terminal domain of mCEACAM1a with that of the murine homolog of the poliovirus receptor (PVR) yields a functional receptor for MHV (10), and Ceacam1a-knockout mice are completely resistant to infection with the hepatotropic A59 strain of MHV (17, 25).Wild mice have two alleles of the mCeacam1 gene, called mCeacam1a and mCeacam1b. Inbred mouse strains that are homozygous for mCeacam1a, including BALB/c, C57BL/6 (B6), C3H, and A/J mice, etc., are highly susceptible to infection with strains of MHV. In contrast, the SJL line of inbred mice, which is resistant to death from MHV infection, is homozygous for the mCeacam1b allele (5, 11, 50). The most extensive differences in amino acid sequence between mCEACAM1a and mCEACAM1b are found in the N-terminal domain, where the virus-binding region is located (21, 22, 32). It was initially reported by Boyle et al. that mCEACAM1a proteins had MHV-A59 virus-binding activity in a virus overlay protein blot, while mCEACAM1b did not (5). Those authors speculated that the different viral affinities of these mCEACAM1 proteins may account for the various MHV-A59 susceptibilities of BALB/c mice compared to those of SJL mice (49). However, Yokomori and Lai (53) and Dveksler et al. (11) previously showed that when recombinant CEACAM1a and CEACAM1b proteins are expressed at high levels on cultured cells, both proteins have MHV-A59 receptor activity. Yokomori and Lai suggested that the difference in MHV susceptibility between BALB/c and SJL mice does not depend solely upon the interaction of the virus with mCEACAM1 proteins (52, 53). Dveksler et al. suggested that small differences in MHV-A59 receptor activity between mCEACAM1a and mCEACAM1b could result in very large biological differences during multiple cycles of infection in in vivo infection (11). We then quantitatively showed that recombinant mCEACAM1a expressed in BHK cells has 10- to 30-times-higher MHV-binding activity than mCEACAM1b (31). Similar results were observed in other laboratories (7, 32). Because the mCeacam1 gene is located on chromosome 7 (34) and the gene controlling MHV-A59 susceptibility and the resistance of BALB/c mice versus SJL mice is also located on chromosome 7 close to the mCeacam1 gene (40), we speculated that the mCeacam1 gene is identical to the gene that determines the susceptibility and/or resistance of mice to MHV-A59 and MHV-JHM infection.To examine the above-described hypothesis, we used progeny mice produced by crossing BALB/c and SJL mice. F2 mice and F1 mice backcrossed to SJL mice were examined for the mCeacam1 genotype and for MHV-JHM susceptibility (30). Mice homozygous for mCeacam1a (1a/1a) and heterozygous mice (1a/1b) were susceptible to lethal MHV-JHM infection, while mice homozygous for mCeacam1b (1b/1b) were not killed by inoculation with MHV-JHM. These data are consistent with the hypothesis that the susceptibility of mice to MHV is determined by the mCeacam1a allele (30). However, this classical genetic analysis could not prove that mCeacam1 alone determines the susceptibility or resistance of mice to MHV-JHM infection, because this methodology cannot rule out the possibility that a different unknown host gene located close to mCeacam1 on chromosome 7 could also affect MHV-JHM susceptibility. Therefore, we used gene replacement in B6 embryonic stem (ES) cells to create a mouse strain in which the exon encoding the N-terminal part of the N-terminal Ig domain of mCeacam1a was replaced with the corresponding region of mCeacam1b from SLJ mice. We bred the chimeric mCeacam1 gene on the B6 background (called B6 chimeric mCeacam1ba, or cB61ba). We compared these mice, wild-type B6 mice, and SJL mice for their susceptibilities to MHV-A59 infection. We confirmed that the expression of mCEACAM1a makes mice susceptible to lethal infection with MHV-A59. However, surprisingly, we found that cB61ba mice were profoundly resistant to MHV-A59 infection, while the virus could replicate at low levels in SJL mice in a self-limited, unapparent infection. Our results suggest that one or more as-yet-unidentified murine genes may also contribute to murine susceptibility and/or resistance to MHV-A59 infection.  相似文献   

8.
9.
In silico analysis of group 4 [NiFe]-hydrogenases from a hyperthermophilic archaeon, Thermococcus onnurineus NA1, revealed a novel tripartite gene cluster consisting of dehydrogenase-hydrogenase-cation/proton antiporter subunits, which may be classified as the new subgroup 4b of [NiFe]-hydrogenases-based on sequence motifs.Hydrogenases are the key enzymes involved in the metabolism of H2, catalyzing the following chemical reaction: 2H+ + 2e ↔ H2. Hydrogenases can be classified into [NiFe]-hydrogenases, [FeFe]-hydrogenases, and [Fe]-hydrogenases, based on their distinctive functional core containing the catalytic metal center (11, 17).The genomic analysis of Thermococcus onnurineus NA1, a hyperthermophilic archaeon isolated from a deep-sea hydrothermal vent area, revealed the presence of several distinct gene clusters encoding seven [NiFe]-hydrogenases and one homolog similar to Mbx (membrane-bound oxidoreductase) from Pyrococcus furiosus (1, 6, 8, 12). According to the classification system of hydrogenases by Vignais et al. (17), three hydrogenases (one F420-reducing and two NADP-reducing hydrogenases) belong to group 3 [NiFe]-hydrogenases, and four hydrogenases belong to group 4 [NiFe]-hydrogenases. The group 4 hydrogenases are widely distributed among bacteria and archaea (17), with Hyc and Hyf (hydrogenase 3 and 4, respectively) from Escherichia coli (19), Coo (CO-induced hydrogenase) from Rhodospirillum rubrum (4), Ech (energy-converting hydrogenase) from Methanosarcina barkeri (7), and Mbh (membrane-bound hydrogenase) from P. furiosus (6, 10, 12) being relatively well-characterized hydrogenases in this group. One of the four group 4 hydrogenases from T. onnurineus NA1 was found to be similar in sequence to that of P. furiosus Mbh (10).  相似文献   

10.
The development of cellular systems in which the enzyme hydrogenase is efficiently coupled to the oxygenic photosynthesis apparatus represents an attractive avenue to produce H2 sustainably from light and water. Here we describe the molecular design of the individual components required for the direct coupling of the O2-tolerant membrane-bound hydrogenase (MBH) from Ralstonia eutropha H16 to the acceptor site of photosystem I (PS I) from Synechocystis sp. PCC 6803. By genetic engineering, the peripheral subunit PsaE of PS I was fused to the MBH, and the resulting hybrid protein was purified from R. eutropha to apparent homogeneity via two independent affinity chromatographical steps. The catalytically active MBH-PsaE (MBHPsaE) hybrid protein could be isolated only from the cytoplasmic fraction. This was surprising, since the MBH is a substrate of the twin-arginine translocation system and was expected to reside in the periplasm. We conclude that the attachment of the additional PsaE domain to the small, electron-transferring subunit of the MBH completely abolished the export competence of the protein. Activity measurements revealed that the H2 production capacity of the purified MBHPsaE fusion protein was very similar to that of wild-type MBH. In order to analyze the specific interaction of MBHPsaE with PS I, His-tagged PS I lacking the PsaE subunit was purified via Ni-nitrilotriacetic acid affinity and subsequent hydrophobic interaction chromatography. Formation of PS I-hydrogenase supercomplexes was demonstrated by blue native gel electrophoresis. The results indicate a vital prerequisite for the quantitative analysis of the MBHPsaE-PS I complex formation and its light-driven H2 production capacity by means of spectroelectrochemistry.Molecular hydrogen (H2) is often discussed as an alternative source of energy (13, 22, 26, 41). It is a highly energetic, renewable, and zero-carbon dioxide emission fuel; however, it is produced mainly from fossil resources. One intriguing possibility for sustainable H2 production is the development of cellular systems in which the light-driven oxygenic photosynthesis is efficiently coupled to hydrogen production by hydrogenase (1, 21, 36).During the process of oxygenic photosynthesis, photosystem II (PS II), a thylakoid membrane (TM)-embedded multiprotein complex, utilizes solar energy to oxidize water into dioxygen (O2), protons, and electrons. The electrons released by PS II are further conducted through an electron transport chain consisting of plastoquinones, the cytochrome b6f complex, and either plastocyanin or cytochrome c6 to the chlorophyll (Chl) dimer P700 in photosystem I (PS I) (20, 48). During light-induced charge separation in PS I, P700 is oxidized, leading to the reduction of the adjacent cofactor A0 (Chl a). From there, the electrons are transmitted to the phylloquinone A1 and subsequently to the Fe4S4 clusters FX, FA, and FB (9) that are located at the acceptor site of PS I. The acceptor site is composed of the PsaC subunit, which harbors the iron-sulfur clusters FA and FB, and the two additional cofactor-free extrinsic subunits PsaD and PsaE. In the final step, the electrons are transferred from FB to the ferredoxin (PetF), which has a midpoint potential of −412 mV (see Fig. Fig.1B)1B) (8, 9).Open in a separate windowFIG. 1.Models of the hydrogenase and photosystem I complexes used in this study. (A) Membrane-bound hydrogenase (MBHwt) of Ralstonia eutropha H16. (B) Wild-type photosystem I (PS I) from Synechocystis sp. PCC 6803. (C) MBHstop protein lacking the C-terminal anchor domain of HoxK. (D) MBHPsaE and PS IΔPsaE.Hydrogenases of the NiFe and FeFe types catalyze the reversible cleavage of H2 into protons and electrons (18, 63). For most hydrogenases, this reaction is highly sensitive to O2 and leads to the reversible or even irreversible inactivation of the enzyme (49, 66, 67). A prominent exception is the oxygen-tolerant membrane-bound [NiFe]-hydrogenase (MBH) from Ralstonia eutropha H16, which catalyzes H2 conversion in the presence of O2 (42, 65). The MBH consists of large subunit HoxG (67 kDa), harboring the NiFe active site, and small subunit HoxK (35 kDa), bearing three FeS clusters (Fig. (Fig.1)1) (32). Both cofactor-containing subunits are completely assembled within the cytoplasm and become subsequently translocated through the cytoplasmic membrane by the twin-arginine translocation (Tat) system. This transport is guided by a specific Tat signal peptide that is located at the N terminus of small subunit HoxK (53). The MBH is then connected to the membrane via the hydrophobic C-terminal “anchor” domain of HoxK, which provides the electronic connection to the diheme cytochrome b, HoxZ (5, 57). All structural, accessory, and regulatory genes for the synthesis of active MBH are arranged in a large, megaplasmid-borne operon (7, 11, 14, 29, 33, 38, 58).The concept of light-driven hydrogen production has been investigated in numerous studies (for reviews, see references 3, 21, and 23), including one involving direct electron transfer from PS I to the free form of hydrogenase in vitro (45). In a preliminary attempt, the MBH from R. eutropha was recently directly fused to PsaE (creating MBHPsaE) (28). The fusion protein was partially purified and subjected to in vitro reconstitution with PS I lacking PsaE (PS IΔPsaE) (54) for light-driven hydrogen production. This concept was based on the previous observation that PS I lacking the peripheral subunit PsaE is fully reconstituted in vitro simply by the addition of independently purified PsaE protein (12).In the present communication, we describe a novel purification procedure for R. eutropha MBHPsaE that yields homogeneous, functionally active MBHPsaE. Additionally, a new method for efficient and fast purification of Synechocystis sp. PCC 6803 (hereafter referred to as Synechocystis) His-tagged PS I was established. Finally, the pure proteins MBHPsaE and PS IΔPsaE were successfully subjected to in vitro reconstitution.  相似文献   

11.
12.
13.
Class III chitin synthases play important roles in tip growth and conidiation in many filamentous fungi. However, little is known about their functions in those processes. To address these issues, we characterized the deletion mutant of a class III chitin synthase-encoding gene of Aspergillus nidulans, chsB, and investigated ChsB localization in the hyphae and conidiophores. Multilayered cell walls and intrahyphal hyphae were observed in the hyphae of the chsB deletion mutant, and wavy septa were also occasionally observed. ChsB tagged with FLAG or enhanced green fluorescent protein (EGFP) localized mainly at the tips of germ tubes, hyphal tips, and forming septa during hyphal growth. EGFP-ChsB predominantly localized at polarized growth sites and between vesicles and metulae, between metulae and phialides, and between phalides and conidia in asexual development. These results strongly suggest that ChsB functions in the formation of normal cell walls of hyphae, as well as in conidiophore and conidia development in A. nidulans.Chitin, a polymer of β-1,4-linked N-acetylglucosmine, is one of the major structural components of the fungal cell wall. Its metabolism, including synthesis, degradation, assembly, and cross-linking to other cell wall components, is thought to be very important for many fungi (5, 22, 24, 36, 45). Fungal chitin synthases have been classified into seven groups, classes I to VII, depending on the structures of their conserved regions (6). The genes encoding the synthases belonging to classes III, V, VI, and VII are only found in fungi with high chitin contents in their cell walls. We have identified six chitin synthase genes from Aspergillus nidulans and designated them chsA, chsB, chsC, chsD, csmA, and csmB; these gene products belong to classes II, III, I, IV, V, and VI, respectively (9, 13, 30, 31, 44, 52). The chsB deletion mutant grew very slowly and formed small colonies with highly branched hyphae, suggesting its important role in hyphal tip growth (3, 52). Repression of chsB expression in the deletion mutant of chsA, chsC, or chsD exaggerated the defects in the formation of aerial hyphae, the production of cell mass, or the growth under high-osmolarity conditions, respectively, compared to each single mutant. These results indicate that chsB functions at various stages of development (15, 16).The deletion of class III chitin synthase-encoding genes leads to severe defects in most of the filamentous fungi thus far investigated. However, their detailed functions are currently unknown. In Neurospora crassa, inactivation of the gene encoding Chs-1, a class III chitin synthase with 63% identity to A. nidulans ChsB, leads to slow growth, aberrant hyphal morphology, and a decrease in chitin synthase activity. The mutant of chs-1 became sensitive to Nikkomycin Z, a chitin synthase inhibitor (53). In Aspergillus fumigatus, two genes encoding class III chitin synthases, chsC and chsG, have been identified. Their gene products showed 66 and 89% identity, respectively, to A. nidulans ChsB. The chsG deletion mutant showed slow growth and defects in conidiation, and its hyphae were highly branched. chsC deletion did not cause any phenotypic change. The chsC chsG double deletion mutant showed almost the same phenotype as the chsG single deletion mutant (28). Class III chitin synthases have been reported to be involved in the virulence of some pathogens. Deletion of Bcchs3a in the phytopathogenic fungus Botrytis cinerea and double deletion of WdCHS3 and class I chitin synthase WdCHS2 in the human pathogen Wangiella dermatitidis both caused a reduction of virulence (40, 48). On the other hand, the deletion mutant of a class III chitin synthase-encoding gene, CgChsIII, of the maize pathogen Colletotrichum graminicola did not exhibit the significant phenotypic difference from the wild-type strain (50). Deletion of a gene, chs1, encoding a class III chitin synthase of the maize pathogenic dimorphic fungi Ustilago maydis caused minor defects in the growth of haploid yeastlike cells and conjugation tube formation (49). These results indicate that the functions of class III chitin synthases has evolutionally diverged.In the present study, we characterized the cytological defects of the A. nidulans chsB deletion mutant and investigated the localization of ChsB using FLAG- or enhanced green fluorescent protein (EGFP)-tagged ChsB. We reveal that the deletion mutant formed hyphae with aberrant cell wall structures and that ChsB tagged with EGFP primarily localized at polarized growth sites during germination, hyphal growth, septation, and conidiation. These findings suggest that ChsB functions at the polarized growth sites and forming septa during the hyphal growth and conidia development.  相似文献   

14.
15.
16.
Nitrate-reducing enrichments, amended with n-hexadecane, were established with petroleum-contaminated sediment from Onondaga Lake. Cultures were serially diluted to yield a sediment-free consortium. Clone libraries and denaturing gradient gel electrophoresis analysis of 16S rRNA gene community PCR products indicated the presence of uncultured alpha- and betaproteobacteria similar to those detected in contaminated, denitrifying environments. Cultures were incubated with H34-hexadecane, fully deuterated hexadecane (d34-hexadecane), or H34-hexadecane and NaH13CO3. Gas chromatography-mass spectrometry analysis of silylated metabolites resulted in the identification of [H29]pentadecanoic acid, [H25]tridecanoic acid, [1-13C]pentadecanoic acid, [3-13C]heptadecanoic acid, [3-13C]10-methylheptadecanoic acid, and d27-pentadecanoic, d25-, and d24-tridecanoic acids. The identification of these metabolites suggests a carbon addition at the C-3 position of hexadecane, with subsequent β-oxidation and transformation reactions (chain elongation and C-10 methylation) that predominantly produce fatty acids with odd numbers of carbons. Mineralization of [1-14C]hexadecane was demonstrated based on the recovery of 14CO2 in active cultures.Linear alkanes account for a large component of crude and refined petroleum products and, therefore, are of environmental significance with respect to their fate and transport (38). The aerobic activation of alkanes is well documented and involves monooxygenase and dioxygenase enzymes in which not only is oxygen required as an electron acceptor but it also serves as a reactant in hydroxylation (2, 16, 17, 32, 34). Alkanes are also degraded under anoxic conditions via novel degradation strategies (34). To date, there are two known pathways of anaerobic n-alkane degradation: (i) alkane addition to fumarate, commonly referred to as fumarate addition, and (ii) a putative pathway, proposed by So et al. (25), involving carboxylation of the alkane. Fumarate addition proceeds via terminal or subterminal addition (C-2 position) of the alkane to the double bond of fumarate, resulting in the formation of an alkylsuccinate. The alkylsuccinate is further degraded via carbon skeleton rearrangement and β-oxidation (4, 6, 8, 12, 13, 21, 37). Alkane addition to fumarate has been documented for a denitrifying isolate (21, 37), sulfate-reducing consortia (4, 8, 12, 13), and five sulfate-reducing isolates (4, 6-8, 12). In addition to being demonstrated in these studies, fumarate addition in a sulfate-reducing enrichment growing on the alicyclic alkane 2-ethylcyclopentane has also been demonstrated (23). In contrast to fumarate addition, which has been shown for both sulfate-reducers and denitrifiers, the putative carboxylation of n-alkanes has been proposed only for the sulfate-reducing isolate strain Hxd3 (25) and for a sulfate-reducing consortium (4). Experiments using NaH13CO3 demonstrated that bicarbonate serves as the source of inorganic carbon for the putative carboxylation reaction (25). Subterminal carboxylation of the alkane at the C-3 position is followed by elimination of the two terminal carbons, to yield a fatty acid that is one carbon shorter than the parent alkane (4, 25). The fatty acids are subject to β-oxidation, chain elongation, and/or C-10 methylation (25).In this study, we characterized an alkane-degrading, nitrate-reducing consortium and surveyed the metabolites of the consortium incubated with either unlabeled or labeled hexadecane in order to elucidate the pathway of n-alkane degradation. We present evidence of a pathway analogous to the proposed carboxylation pathway under nitrate-reducing conditions.  相似文献   

17.
18.
The pyrene-degrading Mycobacterium sp. strain AP1 grew in nutrient-supplemented artificial seawater with a heavy fuel oil as the sole carbon source, causing the complete removal of all linear (C12 to C40) and branched alkanes from the aliphatic fraction, as well as an extensive degradation of the three- and four-ring polycyclic aromatic hydrocarbons (PAHs) phenanthrene (95%), anthracene (80%), fluoranthene (80%), pyrene (75%), and benzo(a)anthracene (30%). Alkylated PAHs, which are more abundant in crude oils than the nonsubstituted compounds, were selectively attacked at extents that varied from more than 90% for dimethylnaphthalenes, methylphenanthrenes, methylfluorenes, and methyldibenzothiophenes to about 30% for monomethylated fluoranthenes/pyrenes and trimethylated phenanthrenes and dibenzothiophenes. Identification of key metabolites indicated the utilization of phenanthrene, pyrene, and fluoranthene by known assimilatory metabolic routes, while other components were cooxidized. Detection of mono- and dimethylated phthalic acids demonstrated ring cleavage and further oxidation of alkyl PAHs. The extensive degradation of the alkanes, the two-, three-, and four-ring PAHs, and their 1-, 2-, and 3-methyl derivatives from a complex mixture of hydrocarbons by Mycobacterium sp. strain AP1 illustrates the great substrate versatility of alkane- and PAH-degrading mycobacteria.Accidental oil spills cause extensive ecological damage to marine shorelines and also have an enormous impact on related economic activities due to the potential risk to public health. One of the most recent examples is the heavy fuel oil spill from the tanker Prestige in 2002, which affected 1,900 km of coast in northwestern Spain. While the light fractions of the oil evaporate in the early stages of a spill, microbial degradation plays a major role in the removal of the heavier fractions. Stimulation of natural biodegradation processes by nutrient and fertilizer addition has proven to enhance oil degradation in a variety of coastal environments (3, 42, 44).Oil is a complex mixture of hundreds of components that can be separated into saturates, aromatics, resins, and asphaltenes. The saturated hydrocarbons are usually the most abundant, while polycyclic aromatic hydrocarbons (PAHs) cause the greatest concern because of their toxic and genotoxic potentials.Most of the available knowledge on the microbial processes involved in PAH biodegradation has been obtained from studies involving bacterial isolates acting on single substrates that serve as the sole source of carbon and energy for growth (7, 20, 22). The pathways for the complete degradation of hydrocarbons containing two and three aromatic rings by gram-negative bacteria are well characterized for such conditions (7, 22). Conversely, degradation of hydrocarbons containing four or more fused aromatic rings, such as pyrene, has been reported only for soil actinomycetes (20, 25, 29, 30, 36, 45), which use multibranched pathways involving both classical dioxygenation and meta-cleavage reactions and novel ortho-cleavage mechanisms uncommon in gram-negative organisms (23). Due to the relaxed specificity of some degradative enzymes, mainly dioxygenases (15, 37), PAH-degrading strains have a wide range of substrates, being able to act simultaneously on a number of structural analogs and to oxidize them to different extents (18, 37). However, the individual processes involved in the degradation of naturally occurring complex mixtures of PAHs (crude oils and coal derivatives) have rarely been addressed (18, 31).Early studies on biodegradation of crude oil were carried out with bacterial strains able to use this mixture for growth. Since PAHs and other components are contained within a predominantly aliphatic matrix in crude oil, most of these studies reported actions of alkane degraders on individual oil components (2, 34, 38, 41, 50). In addition to alkanes, these alkane degraders selectively depleted some alkylated PAHs (2, 41), a process that has been attributed to partial oxidation due to a monooxygenase attack on the methyl groups to produce the corresponding carboxylic acids (35). Recent studies reported the isolation of a number of two- and three-ring-PAH-degrading bacterial strains from coastal sediments affected by crude oil spills. These strains include members of genera commonly isolated from PAH-contaminated soils, such as Pseudomonas (39, 43) and Sphingomonas (49), as well as less common genera, such as Marinobacter (13), Moraxella (43), Vibrio (51), and Cycloclasticus (12). The last genus seems to play a major role in the fate of low-molecular-weight PAHs in the marine environment, as members of this genus have been isolated from several crude oil-contaminated locations (6, 14, 21). When incubated with crude oil, Cycloclasticus strains degraded most of the two- and three-ring PAHs and some of their alkyl derivatives (C0-4 naphthalene, C0-2 dibenzothiophene, C0-2 phenanthrene, and C0-2 fluorene [numerals indicate the number of methyl groups]). However, neither alkanes, trimethyl derivatives of three-ring PAHs, or higher-molecular-weight PAHs were significantly depleted (21). On the other hand, no attempts were made to identify metabolic intermediates indicative of specific degradation or cometabolic pathways.Alkyl-PAH degradation is isomer specific, a feature that has been used in geochemistry to define source recognition and oil weathering ratios (47). For example, given the resistance of 9-methyl phenanthrene to microbial oxidation in relation to the other isomers, the ratio of 3-methylphenanthrene plus 2-methylphenanthrene to 9-methylphenanthrene plus 1-methylphenanthrene has been utilized as a diagnostic ratio (47). These ratios have been defined on the basis of analysis of environmental samples (47) and results of crude oil biodegradation assays with mixed cultures (10, 48) or single strains (2, 41), mainly alkane-degrading pseudomonads. The actions of high-molecular-weight-PAH-degrading mycobacteria on the alkylated families of PAHs present in crude oil and derivatives have not been addressed.Mycobacterium strains isolated by their ability to grow on pyrene have often been shown to also utilize phenanthrene, fluoranthene, and high-molecular-weight alkanes as single carbon sources (8, 45). In a recent study, we showed that when Mycobacterium strain AP1, isolated from an oil-polluted marine beach, was incubated with a mixture of PAHs from creosote, this strain caused a significant depletion of the three-aromatic-ring PAHs but had a limited action on the higher-molecular-weight PAHs fluoranthene and pyrene (31). Given the wide substrate versatility of pyrene-degrading mycobacteria, especially for alkane degradation, their presence in marine environments (16), and their distinctive reactions during PAH degradation (22, 25, 30), in this study we used strain AP1 to investigate the catabolic potential of mycobacteria in the removal of the most abundant hydrocarbon families and their derivatives from crude oil in a marine medium under laboratory conditions. The identification of key metabolites indicative of previously proposed reactions gave insight into the metabolic and cometabolic processes involved. As a model mixture, we used the heavy fuel oil spilled from the Prestige, a Russian M100 fuel oil especially rich in aromatic hydrocarbons (52%) (27).  相似文献   

19.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

20.
Halogenases have been shown to play a significant role in biosynthesis and introducing the bioactivity of many halogenated secondary metabolites. In this study, 54 reduced flavin adenine dinucleotide (FADH2)-dependent halogenase gene-positive strains were identified after the PCR screening of a large collection of 228 reference strains encompassing all major families and genera of filamentous actinomycetes. The wide distribution of this gene was observed to extend to some rare lineages with higher occurrences and large sequence diversity. Subsequent phylogenetic analyses revealed that strains containing highly homologous halogenases tended to produce halometabolites with similar structures, and halogenase genes are likely to propagate by horizontal gene transfer as well as vertical inheritance within actinomycetes. Higher percentages of halogenase gene-positive strains than those of halogenase gene-negative ones contained polyketide synthase genes and/or nonribosomal peptide synthetase genes or displayed antimicrobial activities in the tests applied, indicating their genetic and physiological potentials for producing secondary metabolites. The robustness of this halogenase gene screening strategy for the discovery of particular biosynthetic gene clusters in rare actinomycetes besides streptomycetes was further supported by genome-walking analysis. The described distribution and phylogenetic implications of the FADH2-dependent halogenase gene present a guide for strain selection in the search for novel organohalogen compounds from actinomycetes.It is well known that actinomycetes, notably filamentous actinomycetes, have a remarkable capacity to produce bioactive molecules for drug development (4, 6). However, novel technologies are demanded for the discovery of new bioactive secondary metabolites from these microbes to meet the urgent medical need for drug candidates (5, 9, 31).Genome mining recently has been used to search for new drug leads (7, 20, 42, 51). Based on the hypothesis that secondary metabolites with similar structures are biosynthesized by gene clusters that harbor certain homologous genes, such homologous genes could serve as suitable markers for distinct natural-product gene clusters (26, 51). A wide range of structurally diverse bioactive compounds are synthesized by polyketide synthase (PKS) and nonribosomal peptide synthetase (NRPS) systems in actinomycetes, therefore much attention has been given to revealing a previously unrecognized biosynthetic potential of actinomycetes through the genome mining of these genes (2, 3, 22). However, the broad distribution of PKS and NRPS genes and their high numbers even in a single actinomycete complicate their use (2, 3). To rationally exploit the genetic potential of actinomycetes, more and more special genes, such as tailoring enzyme genes, are being utilized for this sequence-guided genetic screening strategy (20, 38).Tailoring enzymes, which are responsible for the introduction and generation of diversity and bioactivity in several structural classes during or after NRPS, PKS, or NRPS/PKS assembly lines, usually include acyltransferases, aminotransferases, cyclases, glycosyltransferases, halogenases, ketoreductases, methyltransferases, and oxygenases (36, 45). Halogenation, an important feature for the bioactivity of a large number of distinct natural products (16, 18, 30), frequently is introduced by one type of halogenase, called reduced flavin adenine dinucleotide (FADH2)-dependent (or flavin-dependent) halogenase (10, 12, 35). More than 4,000 halometabolites have been discovered (15), including commercially important antibiotics such as chloramphenicol, vancomycin, and teicoplanin (43).Previous investigations of FADH2-dependent halogenase genes were focused largely on related gene clusters in the genera Amycolatopsis (33, 44, 53) and Streptomyces (8, 10, 21, 27, 32, 34, 47-49) and also on those in the genera Actinoplanes (25), Actinosynnema (50), Micromonospora (1), and Nonomuraea (39); however, none of these studies has led to the rest of the major families and genera of actinomycetes. In addition, there is evidence that FADH2-dependent halogenase genes of streptomycetes usually exist in halometabolite biosynthetic gene clusters (20), but we lack knowledge of such genes and clusters in other actinomycetes.In the present study, we show that the distribution of the FADH2-dependent halogenase gene in filamentous actinomycetes does indeed correlate with the potential for halometabolite production based on other genetic or physiological factors. We also showed that genome walking near the halogenase gene locus could be employed to identify closely linked gene clusters that likely encode pathways for organohalogen compound production in actinomycetes other than streptomycetes.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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