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.
2.
Lipoic acid is an essential cofactor required for the function of key metabolic pathways in most organisms. We report the characterization of a Bacillus subtilis mutant obtained by disruption of the lipA (yutB) gene, which encodes lipoyl synthase (LipA), the enzyme that catalyzes the final step in the de novo biosynthesis of this cofactor. The function of lipA was inferred from the results of genetic and physiological experiments, and this study investigated its role in B. subtilis fatty acid metabolism. Interrupting lipoate-dependent reactions strongly inhibits growth in minimal medium, impairing the generation of branched-chain fatty acids and leading to accumulation of copious amounts of straight-chain saturated fatty acids in B. subtilis membranes. Although depletion of LipA induces the expression of the Δ5 desaturase, controlled by a two-component system that senses changes in membrane properties, the synthesis of unsaturated fatty acids is insufficient to support growth in the absence of precursors for branched-chain fatty acids. However, unsaturated fatty acids generated by deregulated overexpression of the Δ5 desaturase functionally replaces lipoic acid-dependent synthesis of branched-chain fatty acids. Furthermore, we show that the cold-sensitive phenotype of a B. subtilis strain deficient in Δ5 desaturase is suppressed by isoleucine only if LipA is present.Lipoic acid (LA; 6,8-thioctic acid or 1,2-dithiolane-3-pentanoic acid) is a sulfur-containing cofactor required for the function of several key enzymes involved in oxidative and single-carbon metabolism, including pyruvate dehydrogenase, 2-oxoglutarate dehydrogenase, branched-chain 2-oxoacid dehydrogenase (BCKADH), acetoin dehydrogenase, and the glycine cleavage system (10). Lipoate-requiring complexes typically contain three protein subunits, E1, E2, and E3. LA is linked through an amide bond to lysine residues in the E2 subunits (42) and acts as a swinging arm, transferring covalently attached reaction intermediates among the active sites of the enzyme complexes (40).Although the general role of LA as a bound cofactor has been known for decades, the mechanisms by which LA is synthesized and becomes linked to its cognate proteins in different organisms continue to be elucidated. The reactions whereby LA-modified proteins are produced are best understood in Escherichia coli. In this organism, lipoylation is mediated by two separate enzymes, lipoyl protein ligase A (LplA) and octanoyl-acyl carrier protein-protein transferase (LipB) (30, 31). While LplA uses exogenous LA, LipB transfers endogenous octanoic acid to the target proteins (19). These octanoylated domains are then converted into lipoylated derivatives by the S-adenosyl-l-methionine-dependent enzyme lipoyl synthase (LipA), which catalyzes the insertion of sulfur atoms into the carbon-6 and -8 positions of the corresponding fatty acids (29). This process bypasses the requirement for an exogenous supply of LA.In contrast to the wealth of knowledge available on LA synthesis and utilization in E. coli, the existing information about these pathways in gram-positive bacteria is scarce. It has been found that Listeria monocytogenes mutants defective in proteins homologous to the E. coli LplA enzymes are unable to scavenge exogenous LA for modification of lipoyl domains (22, 23, 38). However, L. monocytogenes is a natural lipoate auxotroph since it does not encode the enzymes necessary for lipoate biosynthesis (15, 55). Bacillus subtilis synthesizes LA, but the biosynthesis, attachment, and function of this essential nutrient in this model gram-positive organism have not yet been studied in detail (50). Analysis of the genome sequence of B. subtilis (25) revealed that it contains an open reading frame, yutB, encoding a protein with a high degree of homology to E. coli LipA and two open reading frames encoding proteins slightly similar to LplA, while no LipB homolog was detected.LA is a critical cofactor of BCKADH, the enzyme involved in the formation of the primer carbons for the initiation of branched-chain fatty acid (BCFA) synthesis (21). Early work indicated that a bfmB mutant of B. subtilis, defective in both BCKADH and pyruvate dehydrogenase, requires short-branched-chain carboxilic acids for growth (56). However, in our hands, this mutant presented a high percentage of reversion, precluding its use in the study of lipid metabolism. Since BCFAs are the dominant acyl chains found in membrane phospholipids of B. subtilis, the goal of this study was to employ a genetic approach to investigate the role of yutB in the physiology of this organism, in particular in fatty acid metabolism. In addition, we provide compelling evidence showing that Δ5 unsaturated fatty acids (UFA), the products of the B. subtilis desaturase, can fully replace the function of BCFAs. Furthermore, we demonstrate that UFA are essential to provide cryoprotective properties in strains depleted of LipA. This work reports the first characterization of a gram-positive mutant deficient in LA synthesis and its use to study the interplay between BCFAs and UFA metabolism.  相似文献   

3.
Lipoic acid is a sulfur-containing cofactor required for the function of several multienzyme complexes involved in the oxidative decarboxylation of α-keto acids and glycine. Mechanistic details of lipoic acid metabolism are unclear in eukaryotes, despite two well defined pathways for synthesis and covalent attachment of lipoic acid in prokaryotes. We report here the involvement of four genes in the synthesis and attachment of lipoic acid in Saccharomyces cerevisiae. LIP2 and LIP5 are required for lipoylation of all three mitochondrial target proteins: Lat1 and Kgd2, the respective E2 subunits of pyruvate dehydrogenase and α-ketoglutarate dehydrogenase, and Gcv3, the H protein of the glycine cleavage enzyme. LIP3, which encodes a lipoate-protein ligase homolog, is necessary for lipoylation of Lat1 and Kgd2, and the enzymatic activity of Lip3 is essential for this function. Finally, GCV3, encoding the H protein target of lipoylation, is itself absolutely required for lipoylation of Lat1 and Kgd2. We show that lipoylated Gcv3, and not glycine cleavage activity per se, is responsible for this function. Demonstration that a target of lipoylation is required for lipoylation is a novel result. Through analysis of the role of these genes in protein lipoylation, we conclude that only one pathway for de novo synthesis and attachment of lipoic acid exists in yeast. We propose a model for protein lipoylation in which Lip2, Lip3, Lip5, and Gcv3 function in a complex, which may be regulated by the availability of acetyl-CoA, and which in turn may regulate mitochondrial gene expression.Several oxidative decarboxylation reactions are carried out in prokaryotes and eukaryotes by multienzyme complexes. The function of these complexes requires the action of a sulfur-containing cofactor, lipoic acid (6,8-thioctic acid) (1, 2). Lipoic acid is covalently attached via an amide linkage to a specific lysine residue on the surface of the conserved lipoyl domain of the E2 subunits of pyruvate dehydrogenase (PDH),3 α-ketoglutarate dehydrogenase (α-KDH), the branched chain α-keto acid dehydrogenase complexes, and the H protein of the glycine cleavage (GC) enzyme (3). The lipoyl moiety serves as a swinging arm that shuttles reaction intermediates between active sites within the complexes (1). Despite the well characterized function of lipoic acid as a prosthetic group, the mechanisms of its synthesis and attachment to proteins are the subject of ongoing investigations (47).These reactions are best understood in Escherichia coli, which has two well defined pathways for lipoic acid synthesis and attachment: a de novo pathway and a salvage pathway (8). Octanoic acid, synthesized on the acyl carrier protein (ACP) (9), is the substrate for the de novo pathway. Lipoyl synthase (LipA) catalyzes the addition of two sulfur atoms to form lipoic acid from octanoic acid either before or after transfer to the target protein (10) by lipoyl(octanoyl)-ACP:protein transferase (LipB) (11, 12). The preferred order of these two reactions is attachment of octanoic acid by LipB, followed by addition of sulfur by LipA (13). By contrast, in the salvage pathway, lipoate-protein ligase (LplA) attaches free lipoic acid to proteins in a two-step reaction. Lipoic acid, which can be scavenged from the medium, is first activated to lipoyl-AMP and then the lipoyl group is transferred to the proteins (14).Lipoic acid synthesis and attachment to target proteins are less well understood in eukaryotes. Homologs of the E. coli enzymes have been found in fungi, plants, protists, and mammals, but many mechanistic details are unclear (1517). In Saccharomyces cerevisiae, the mitochondrial type II fatty acid biosynthetic pathway (FAS II) synthesizes octanoyl-ACP, which is the substrate for de novo lipoic acid synthesis (18). Lip2 and Lip5, the respective yeast homologs of E. coli LipB and LipA, were shown to be required for respiratory growth on glycerol medium, PDH activity (19), and lipoic acid synthesis (20), indicating functional roles in de novo lipoic acid synthesis and attachment. However, there has been no previous report of an LplA-like lipoate-protein ligase homolog in yeast. Furthermore, lip2 and lip5 mutant strains cannot grow on medium containing lipoic acid (19, 20), suggesting that yeast either cannot use exogenously supplied lipoic acid or there is no yeast equivalent of the E. coli LplA-driven salvage pathway.Here we report the involvement of two additional enzymes in protein lipoylation in yeast mitochondria. The first, Lip3, is a lipoate-protein ligase homolog and is required with Lip2 and Lip5 for lipoylation of the E2 subunits of PDH (Lat1) and α-KDH (Kgd2). The second enzyme, Gcv3, the H protein of the GC enzyme, is absolutely required for lipoylation of all proteins in yeast.  相似文献   

4.
Lipoic acid is a covalently bound cofactor found throughout the domains of life that is required for aerobic metabolism of 2-oxoacids and for C1 metabolism. Utilization of exogenous lipoate is catalyzed by a ligation reaction that proceeds via a lipoyl-adenylate intermediate to attach the cofactor to the ϵ-amino group of a conserved lysine residue of protein lipoyl domains. The lipoyl ligases of demonstrated function have a large N-terminal catalytic domain and a small C-terminal accessory domain. Half of the members of the LplA family detected in silico have only the large catalytic domain. Two x-ray structures of the Thermoplasma acidophilum LplA structure have been reported, although the protein was reported to lack ligase activity. McManus et al. (McManus, E., Luisi, B. F., and Perham, R. N. (2006) J. Mol. Biol. 356, 625–637) hypothesized that the product of an adjacent gene was also required for ligase activity. We have shown this to be the case and have named the second protein, LplB. We found that complementation of Escherichia coli strains lacking lipoate ligase with T. acidophilum LplA was possible only when LplB was also present. LplA had no detectable ligase activity in vitro in the absence of LplB. Moreover LplA and LplB were shown to interact and were purified as a heterodimer. LplB was required for lipoyl-adenylate formation but was not required for transfer of the lipoyl moiety of lipoyl-adenylate to acceptor proteins. Surveys of sequenced genomes show that most lipoyl ligases of the kingdom Archaea are heterodimeric. We propose that the presence of an accessory domain provides a diagnostic to distinguish lipoyl ligase homologues from other members of the lipoate/biotin attachment enzyme family.Lipoic acid is a covalently bound cofactor that conveys activated reaction intermediates between different active sites of multienzyme complexes (1). Lipoate is essential for aerobic metabolism of 2-oxoacids and for glycine cleavage. In its active form lipoate is attached to the ϵ-amino group of a small (∼80-residue) well conserved lipoyl domain (LD)2 lysine residue via an amide bond. LDs are typically found at the N termini of the E2 subunits of 2-oxoacid dehydrogenase complexes. In the 2-oxoacid complexes, lipoylated LD receives the decarboxylated acid from the E1 subunit active site in thioester linkage to a lipoate thiol. The acyl thioester is then converted to the corresponding CoA thioester by thioester exchange catalyzed by the E2 subunit active site. The dihydrolipoamide dehydrogenase subunit (E3) then oxidizes the dihydrolipoyl-LD back to the lipoyl-LD to reset the catalytic cycle. In the glycine cleavage (also called glycine decarboxylase and glycine dehydrogenase) system, the lipoyl domain exists as a free protein designated H. Lipoyl-LD receives the product of glycine decarboxylation, methylamine, from the P protein. The methylamine is then transferred to the T protein to produce methylenetetrahydrofolate that is typically used to synthesize serine from a second molecule of glycine. In Escherichia coli, lipoic acid is essential for aerobic growth because of the need for pyruvate dehydrogenase and 2-oxoglutarate dehydrogenase. The glycine cleavage system is not required for growth of wild type E. coli strains but is required for growth of Arabidopsis where the H protein can be present at millimolar concentrations in photosynthetic cells (2).The reactions whereby lipoic acid-modified proteins are produced are best understood in E. coli. The most straightforward pathway is via lipoate-protein ligase, an activity first described by L. J. Reed et al. (3) (see Fig. 1). These workers postulated that lipoate was attached to protein by a two-step ATP-dependent reaction with lipoyl-AMP as an activated intermediate (Fig. 1). Although the lipoate-protein ligases were key reagents in demonstration of the role of lipoic acid in the 2-oxoacid dehydrogenase reactions (3, 4), the protein was not purified to homogeneity, and thus the proposed mechanism could not be considered proved. The E. coli lplA gene was the first gene encoding a lipoate-protein ligase isolated, and LplA was the first such ligase purified to homogeneity (5, 6). The isolation of null mutants in lplA showed that LplA does not play a role in de novo lipoic acid synthesis but rather acts to scavenge lipoic acid from the environment (6, 7).Open in a separate windowFIGURE 1.Lipoic acid metabolism in E. coli. Panel A, the lipoyl ligase (LplA) reaction that proceeds through the lipoyl-adenylate intermediate. In E. coli LplA acts to scavenge lipoic acid from the growth medium. Panel B, schematic of lipoic acid synthesis in E. coli. LipB transfers an octanoyl moiety from the fatty acid biosynthetic intermediate, octanoyl-acyl carrier protein, to the LD domain of a lipoate-accepting protein (in this case the E2 subunit of a 2-oxoacid dehydrogenase). The octanoylated LD domain is the substrate of LipA, an S-adenosylmethionine radical enzyme that replaces one hydrogen atom on each of octanoate carbons 6 and 8 with sulfur atoms. Panel C, the differing arrangements of genes and domains found in lipoate ligases in T. acidophilum, E. coli, and Streptomyces coelicolor. Only a single nucleotide lies between the T. acidophilum LplB and LplA coding sequences.LplA is a 38-kDa monomeric protein (5). Assays with a fully defined system have demonstrated that LplA plus lipoate and Mg-ATP are sufficient to reconstitute lipoylation in vitro and that lipoyl-AMP is a reaction intermediate (5, 6, 8, 9). Thus, it is clear that LplA catalyzes both the ATP-dependent activation of lipoate to lipoyl-AMP as well as the transfer of this activated lipoyl species to apoprotein with concomitant release of AMP. The E. coli LplA enzyme has been shown to be capable of utilizing lipoate and several lipoate analogues such as octanoate as donors for the post-translational modification of E2 apoproteins in vivo (5, 6).Recently crystal structures of E. coli LplA and of LplA homologues have been reported including an E. coli LplA-lipoic acid complex (1012). The reported structures of the unliganded proteins agree well and show E. coli LplA to be a two-domain protein consisting of a large N-terminal domain and a small C-terminal domain (Figs. 1 and and2).2). However, the E. coli LplA-lipoic acid complex is difficult to interpret because lipoic acid molecules were heterogeneously bound to LplA molecules within the crystals and were poorly resolved. In one case the lipoic acid carboxyl was hydrogen-bonded to Ser-72, whereas in another case Arg-140 was the hydrogen bond donor (10). Because enzymes rarely show such plasticity and lipoic acid is a hydrophobic molecule, it seemed possible that the observed association of the cofactor with a hydrophobic LplA surface in the interdomain cavity was artifactual. Moreover in prior work K. E. Reed et al. (13) had isolated LplA mutants resistant to inhibition by an analogue of lipoic acid in which the sulfur atoms had been replaced with selenium. Because this is a very discrete modification of the LplA substrate, the mutant protein would be expected to have an alteration close to the pocket that binds the lipoic acid thiolane ring. However, the site of this mutation (Gly-76 to serine (7)) was distal from the lipoate binding site reported. This dilemma was resolved by two lipoic acid-containing structures of an LplA homologue from the archaeon Thermoplasma acidophilum (11, 12) that can be readily superimposed on the E. coli LplA structure except that the T. acidophilum protein lacks the E. coli LplA C-terminal domain (Fig. 2). In both T. acidophilum structures the lipoate thiolane ring was adjacent to the glycine residue that corresponds to E. coli Gly-76, the residue giving resistance to the selenium analogue, and a plausible reorganization of the molecule to prevent binding of the slightly larger analogue was proposed (12). Moreover addition of lipoic acid to a complex of the T. acidophilum LplA with ATP gave lipoyl-AMP showing that the lipoic acid was bound in a physiologically meaningful manner (11). The lipoyl-AMP was bound in a U-shaped pocket and was well shielded from solvent. Thus, it seems that the locations of the lipoate moieties in the two T. acidophilum LplA structures indicate that these represent catalytically competent lipoate binding sites (rather than the sites of E. coli LplA where lipoate bound). A caveat was that the T. acidophilum LplA was inactive in catalysis of the overall LplA reaction (12). Because T. acidophilum LplA lacks the C-terminal domain (CTD) of E. coli LplA (11, 12), this suggested that the missing domain was required for activity, and a second protein was proposed to interact with T. acidophilum LplA to allow the complete reaction (12). If this were the case, the T. acidophilum lipoyl ligase would provide an unusually facile system to investigate the role of the CTD in lipoate-protein ligases.Open in a separate windowFIGURE 2.Structural alignments of LplA and LipB structures. Previously published crystal structures were aligned using DeepView (37). Panel A, E. coli LplA (Protein Data Bank code 1X2H in green) aligned with T. acidophilum LplA (Protein Data Bank code 2ART in orange). The lipoyl-adenylate intermediate bound to T. acidophilum LplA is shown in purple. The adenylate binding loop is indicated with an arrow. Panel B, M. tuberculosis LipB (Protein Data Bank code 1W66 in gray) is aligned with the E. coli LplA structure of panel A. The purple line denotes the covalent decanoate adduct present in the M. tuberculosis LipB structure. The substrate binding pocket is conserved among members of the protein family. The accessory domain is not part of the binding pocket and appears to play an indirect role in catalysis.If the lipoate-protein ligase reaction can be catalyzed by a heteromeric protein this may allow better discrimination of lipoyl ligases from acyl carrier protein:protein octanoyltransferases. In E. coli de novo lipoic acid biosynthesis is accomplished by two enzymes, the LipB octanoyltransferase and the LipA lipoyl synthase (14) (Fig. 1). LipB transfers the octanoate moiety from the octanoyl-acyl carrier protein intermediate of fatty acid biosynthesis to the ϵ-amine of the conserved LD lysine residue resulting in amide-linked octanoate (Fig. 1). LipA then catalyzes replacement of a hydrogen atom on each of octanoate carbons 6 and 8 with sulfur atoms derived from a LipA iron-sulfur center via an S-adenosylmethionine-dependent radical mechanism (14, 15). That is, lipoic acid is assembled on its cognate proteins (16).Although the two classes of LD-modifying enzymes, LplA and LipB, show very low amino acid sequence conservation and utilize different chemistries, the proteins surprisingly show structural conservation and have related active site architectures (17, 18) (Fig. 2). The Mycobacterium tuberculosis LipB and T. acidophilum LplA can be superimposed by using all matching Cα positions with a root mean square deviation of ≈2.5 Å with good topological matching of most secondary structural elements (18). Hence in length and structure LipBs resemble LplAs that lack the C-terminal domain. Although the E. coli LipB and LplA sequences align very poorly, a large number of proteins in the data bases have similarities to both proteins, and therefore annotation of a given protein as a ligase or octanoyltransferase is not straightforward. If an LplA CTD can be a separate protein, an additional criterion to distinguish lipoate ligases and octanoyltransferases would be available. It should be noted that biotin ligases also show structural (but not sequence) conservation with LipB and LplA, and this group of proteins comprises the Pfam family PF03099 (19). However, all known biotin ligases have a C-terminal domain that greatly aids in their annotation. We report that, as predicted by McManus et al. (12), the CTD function essential for lipoate-protein ligase activity is encoded by a gene located immediately upstream of T. acidophilum lplA that we call lplB.  相似文献   

5.
Here, we report a fluorescence in situ hybridization (FISH) method for rapid detection of Cronobacter strains in powdered infant formula (PIF) using a novel peptide nucleic acid (PNA) probe. Laboratory tests with several Enterobacteriaceae species showed that the specificity and sensitivity of the method were 100%. FISH using PNA could detect as few as 1 CFU per 10 g of Cronobacter in PIF after an 8-h enrichment step, even in a mixed population containing bacterial contaminants.Cronobacter strains were originally described as Enterobacter sakazakii (12), but they are now known to comprise a novel genus consisting of six separate genomospecies (20, 21). These opportunistic pathogens are ubiquitous in the environment and various types of food and are occasionally found in the normal human flora (11, 12, 16, 32, 47). Based on case reports, Cronobacter infections in adults are generally less severe than Cronobacter infections in newborn infants, with which a high fatality rate is associated (24).The ability to detect Cronobacter and trace possible sources of infection is essential as a means of limiting the impact of these organisms on neonatal health and maintaining consumer confidence in powdered infant formula (PIF). Conventional methods, involving isolation of individual colonies followed by biochemical identification, are more time-consuming than molecular methods, and the reliability of some currently proposed culture-based methods has been questioned (28). Recently, several PCR-based techniques have been described (23, 26, 28-31, 38). These techniques are reported to be efficient even when low levels of Cronobacter cells are found in a sample (0.36 to 66 CFU/100 g). However, PCR requires DNA extraction and does not allow direct, in situ visualization of the bacterium in a sample.Fluorescence in situ hybridization (FISH) is a method that is commonly used for bacterial identification and localization in samples. This method is based on specific binding of nucleic acid probes to particular DNA or RNA target regions (1, 2). rRNA has been regarded as the most suitable target for bacterial FISH, allowing differentiation of potentially viable cells. Traditionally, FISH methods are based on the use of conventional DNA oligonucleotide probes, and a commercial system, VIT-E sakazakii (Vermicon A.G., Munich, Germany), has been developed based on this technology (25). However, a recently developed synthetic DNA analogue, peptide nucleic acid (PNA), has been shown to provide improved hybridization performance compared to DNA probes, making FISH procedures easier and more efficient (41). Taking advantage of the PNA properties, FISH using PNA has been successfully used for detection of several clinically relevant microorganisms (5, 15, 17, 27, 34-36).  相似文献   

6.
The nanATEK-yhcH, yjhATS, and yjhBC operons in Escherichia coli are coregulated by environmental N-acetylneuraminic acid, the most prevalent sialic acid in nature. Here we show that YjhS (NanS) is a probable 9-O-acetyl N-acetylneuraminic acid esterase required for E. coli to grow on this alternative sialic acid, which is commonly found in mammalian host mucosal sites.The coregulated nanATEK-yhcH, yjhATS, and yjhBC operons involved in sialic acid catabolism in Escherichia coli are thought to be induced by the most common sialic acid, N-acetylneuraminic acid (Neu5Ac), through reversible inactivation of the NanR repressor encoded by nanR mapping immediately upstream of nanA (15, 27, 28; http://vetmed.illinois.edu/path/sialobiology/). Sialic acids are a family of over 40 naturally occurring 9-carbon keto sugar acids found mainly in metazoans of the deuterostome (starfish to human) developmental lineage and in some, mostly pathogenic, bacteria, where sialic acids expressed at the microbial cell surface inhibit host innate immunity (27). By contrast, most bacterial commensals and pathogens catabolize sialic acids as sole carbon and nitrogen sources, indicating exploitation of the sialic acid-rich host mucosal environment by a wide range of species (2, 27, 28). Interestingly, in vivo experimental evidence further indicates that sialic acid catabolism functions directly (nutrition) or indirectly (surface decoration and cell signaling) in host-microbe commensal and pathogenic interactions in organisms such as E. coli, Haemophilus influenzae, Pasteurella multocida, Salmonella enterica serovar Typhi, Streptococcus pneumoniae, Vibrio vulnificus, and Vibrio cholerae (1, 3, 5, 6, 10, 14, 23, 24, 26, 29). The animal species used for these studies include rodent models and natural hosts such as cattle and turkeys. The structural diversity of sialic acids at the terminal positions on glycoconjugates (glycoproteins and glycolipids) of mucosal surfaces of these hosts requires sialidases, acetyl esterases, and probably other enzymes that convert alternative or at least minor sialic acids to the more digestible Neu5Ac form (8, 9). We have previously demonstrated that E. coli has an epicurean propensity for metabolizing alternative sialic acids (30, 31). In the current communication, we show that YjhS is required for growth of E. coli on 9-O-acetyl-N-acetylneuraminic acid (Neu5,9Ac2).Because most sialic acids are bound to other sugars, including other sialic acids, as part of the oligosaccharide chains on glycoconjugates, either microbial or endogenous (host) sialidases (NanH, or N-acylneuraminate hydrolases) are needed to release free sugar, which is then transported by NanT in E. coli (15, 16, 26, 31). Once internalized, sialic acid is cleaved by an nanA-encoded aldolase or lyase to yield the 6-carbon hexosamine, N-acetylmannosamine (ManNAc), and pyruvate, with the latter entering the tricarboxylic acid cycle or gluconeogenesis. ManNAc is converted to its 6-phosphate derivative by a specific kinase encoded by nanK and epimerized by NanE to yield N-acetylglucosamine 6-phosphate, which is converted to fructose 6-phosphate by products of the nag operon (15, 17, 31, 32). The functions of the coregulated yjhS, yjhB, yjhC, and yhcH gene products are unknown but are not required for growth on Neu5Ac (15). However, YjhA (NanC) is an outer membrane porin required for diffusion of Neu5Ac in the absence of the major porins (7), while YjhT (NanM) is a mutarotase that catalyzes the conversion of the alpha sialic acid isomer to the more thermodynamically stable beta form (21). Neither nanC nor nanM is required for growth on Neu5Ac (15), suggesting that yjhS, yjhBC, and yhcH are involved in reactions that convert alternative sialic acids to Neu5Ac (22, 23). YhcH was crystallized and has been suggested to be an isomerase or epimerase involved in processing N-glycolylneuraminic acid (Neu5Gc) (25), but deletion of yhcH did not affect growth on this sialic acid as a sole carbon source (16).Computer-assisted analysis indicated that YjhB is a permease similar to NanT (16) whereas YjhC is a likely oxidoreductase or dehydrogenase. Orthologs of yhcH, nanC, nanM, and yjhBC are found in most bacterial species with intact Neu5Ac utilization systems, while yjhS is confined to E. coli and shigellae, either as part of the chromosomes in these strains or integrated with phages or phage remnants. However, a significant match (E value = 0.0007) was found between YjhS and AxeA in Rhodopirellula baltica, where AxeA is an acetyl xylan esterase (11), suggesting YjhS might be a sialate esterase. We propose that YjhS should be designated NanS to indicate its direct participation in utilization of an alternative sialic acid.  相似文献   

7.
8.
To cause disease, Clostridium difficile spores must germinate in the host gastrointestinal tract. Germination is initiated upon exposure to glycine and certain bile acids, e.g., taurocholate. Chenodeoxycholate, another bile acid, inhibits taurocholate-mediated germination. By applying Michaelis-Menten kinetic analysis to C. difficile spore germination, we found that chenodeoxycholate is a competitive inhibitor of taurocholate-mediated germination and appears to interact with the spores with greater apparent affinity than does taurocholate. We also report that several analogs of chenodeoxycholate are even more effective inhibitors. Some of these compounds resist 7α-dehydroxylation by Clostridium scindens, a core member of the normal human colonic microbiota, suggesting that they are more stable than chenodeoxycholate in the colonic environment.Clostridium difficile is a Gram-positive, spore-forming, anaerobic bacterium that is pathogenic for both humans and animals (33, 44). Infections caused by C. difficile range from mild diarrhea to more life-threatening conditions, such as pseudomembranous colitis (33). In the classic case, prior antibiotic treatment that disrupts the normally protective colonic flora makes patients susceptible to C. difficile infection (CDI) (35, 53). Other antibiotics, such as vancomycin and metronidazole, are the most commonly used treatments for CDI (54). However, because these antibiotics also disrupt the colonic flora, 10 to 40% of patients whose symptoms have been ameliorated suffer from relapsing CDI (15, 24). The annual treatment-associated cost for CDI in the United States is estimated to be between $750 million and $3.2 billion (8, 9, 16, 31). Moreover, the number of fatal cases of CDI has been increasing rapidly (14, 39). Thus, there is an urgent need to find alternative therapies for CDI.C. difficile infection likely is initiated by infection with the spore form of C. difficile (12). C. difficile elicits disease through the actions of two secreted toxins, TcdA and TcdB (48). TcdB was recently shown to be critical for pathogenesis in an animal model of disease (18). Since the toxins are produced by vegetative cells, not by spores (17), germination and outgrowth are prerequisites for pathogenesis.Spore germination is triggered by the interaction of small molecules, called germinants, with receptors within the spore inner membrane. These germinants vary by bacterial species and can include ions, amino acids, sugars, nucleotides, surfactants, or combinations thereof (43). The recognition of germinants triggers irreversible germination, leading to Ca2+-dipicolinic acid release, the uptake of water, the degradation of the cortex, and, eventually, the outgrowth of the vegetative bacterium (43). The germination receptors that C. difficile uses to sense the environment have not been identified. Based on homology searches, C. difficile germination receptors must be very different from known germination receptors (42), but they appear to be proteinaceous (13).Taurocholate, a primary bile acid, has been used for approximately 30 years by researchers and clinical microbiologists to increase colony formation by C. difficile spores from patient and environmental samples (3, 49, 51, 52). This suggested that C. difficile spores interact with bile acids along the gastrointestinal (GI) tract and that spores use a host-derived signal to initiate germination.The liver synthesizes the two major primary bile acids, cholate and chenodeoxycholate (40). These compounds are modified by conjugation with either taurine (to give taurocholate or taurochenodeoxycholate) or glycine (producing glycocholate or glycochenodeoxycholate). Upon secretion into the digestive tract, bile aids in the absorption of fat and cholesterol; much of the secreted bile is actively absorbed and recycled back to the liver for reutilization (40). Though efficient, enterohepatic recirculation is not complete; bile enters the cecum of the large intestine at a concentration of approximately 2 mM (30).In the cecum, bile is modified by the normal, benign colonic flora. First, bile salt hydrolases found on the surfaces of many bacterial species remove the conjugated amino acid, producing the deconjugated primary bile acids cholate and chenodeoxycholate (40). These deconjugated primary bile acids are further metabolized by only a few species of intestinal bacteria, including Clostridium scindens. C. scindens actively transports unconjugated primary bile acids into the cytoplasm, where they are 7α-dehydroxylated, converting cholate to deoxycholate and chenodeoxycholate to lithocholate (21, 40). The disruption of the colonic flora by antibiotic treatment abolishes 7α-dehydroxylation activity (41).Building upon the work on Wilson and others (51, 52), we demonstrated that taurocholate and glycine, acting together, trigger the loss of the birefringence of C. difficile spores (45). All cholate derivatives (taurocholate, glycocholate, cholate, and deoxycholate) stimulate the germination of C. difficile spores (45). Recently it was shown that taurocholate binding is prerequisite to glycine binding (37). At physiologically relevant concentrations, chenodeoxycholate inhibits taurocholate-mediated germination (46) and also inhibits C. difficile vegetative growth, as does deoxycholate (45). In fact, C. difficile spores use the relative concentrations of the various bile acids as cues for germination within the host (10).Since chenodeoxycholate is absorbed by the colonic epithelium and metabolized to lithocholate by the colonic flora (25, 40), the use of chenodeoxycholate as a therapy against C. difficile disease might be hindered by its absorption and conversion to lithocholate.Here, we further characterize the interaction of C. difficile spores with various bile acids and demonstrate that chenodeoxycholate is a competitive inhibitor of taurocholate-mediated germination. Further, we identify chemical analogs of chenodeoxycholate that are more potent inhibitors of germination and that resist 7α-dehydroxylation by the colonic flora, potentially increasing their stability and effectiveness as inhibitors of C. difficile spore germination in the colonic environment.  相似文献   

9.
Phenoxyalkanoic acid (PAA) herbicides are widely used in agriculture. Biotic degradation of such herbicides occurs in soils and is initiated by α-ketoglutarate- and Fe2+-dependent dioxygenases encoded by tfdA-like genes (i.e., tfdA and tfdAα). Novel primers and quantitative kinetic PCR (qPCR) assays were developed to analyze the diversity and abundance of tfdA-like genes in soil. Five primer sets targeting tfdA-like genes were designed and evaluated. Primer sets 3 to 5 specifically amplified tfdA-like genes from soil, and a total of 437 sequences were retrieved. Coverages of gene libraries were 62 to 100%, up to 122 genotypes were detected, and up to 389 genotypes were predicted to occur in the gene libraries as indicated by the richness estimator Chao1. Phylogenetic analysis of in silico-translated tfdA-like genes indicated that soil tfdA-like genes were related to those of group 2 and 3 Bradyrhizobium spp., Sphingomonas spp., and uncultured soil bacteria. Soil-derived tfdA-like genes were assigned to 11 clusters, 4 of which were composed of novel sequences from this study, indicating that soil harbors novel and diverse tfdA-like genes. Correlation analysis of 16S rRNA and tfdA-like gene similarity indicated that any two bacteria with D > 20% of group 2 tfdA-like gene-derived protein sequences belong to different species. Thus, data indicate that the soil analyzed harbors at least 48 novel bacterial species containing group 2 tfdA-like genes. Novel qPCR assays were established to quantify such new tfdA-like genes. Copy numbers of tfdA-like genes were 1.0 × 106 to 65 × 106 per gram (dry weight) soil in four different soils, indicating that hitherto-unknown, diverse tfdA-like genes are abundant in soils.Phenoxyalkanoic acid (PAA) herbicides such as MCPA (4-chloro-2-methyl-phenoxyacetic acid) and 2,4-D (2,4-dichlorophenoxyacetic acid) are widely used to control broad-leaf weeds in agricultural as well as nonagricultural areas (19, 77). Degradation occurs primarily under oxic conditions in soil, and microorganisms play a key role in the degradation of such herbicides in soil (62, 64). Although relatively rapidly degraded in soil (32, 45), both MCPA and 2,4-D are potential groundwater contaminants (10, 56, 70), accentuating the importance of bacterial PAA herbicide-degrading bacteria in soils (e.g., references 3, 5, 6, 20, 41, 59, and 78).Degradation can occur cometabolically or be associated with energy conservation (15, 54). The first step in the degradation of 2,4-D and MCPA is initiated by the product of cadAB or tfdA-like genes (29, 30, 35, 67), which constitutes an α-ketoglutarate (α-KG)- and Fe2+-dependent dioxygenase. TfdA removes the acetate side chain of 2,4-D and MCPA to produce 2,4-dichlorophenol and 4-chloro-2-methylphenol, respectively, and glyoxylate while oxidizing α-ketoglutarate to CO2 and succinate (16, 17).Organisms capable of PAA herbicide degradation are phylogenetically diverse and belong to the Alpha-, Beta-, and Gammproteobacteria and the Bacteroidetes/Chlorobi group (e.g., references 2, 14, 29-34, 39, 60, 68, and 71). These bacteria harbor tfdA-like genes (i.e., tfdA or tfdAα) and are categorized into three groups on an evolutionary and physiological basis (34). The first group consists of beta- and gammaproteobacteria and can be further divided into three distinct classes based on their tfdA genes (30, 46). Class I tfdA genes are closely related to those of Cupriavidus necator JMP134 (formerly Ralstonia eutropha). Class II tfdA genes consist of those of Burkholderia sp. strain RASC and a few strains that are 76% identical to class I tfdA genes. Class III tfdA genes are 77% identical to class I and 80% identical to class II tfdA genes and linked to MCPA degradation in soil (3). The second group consists of alphaproteobacteria, which are closely related to Bradyrhizobium spp. with tfdAα genes having 60% identity to tfdA of group 1 (18, 29, 34). The third group also harbors the tfdAα genes and consists of Sphingomonas spp. within the alphaproteobacteria (30).Diverse PAA herbicide degraders of all three groups were identified in soil by cultivation-dependent studies (32, 34, 41, 78). Besides CadAB, TfdA and certain TfdAα proteins catalyze the conversion of PAA herbicides (29, 30, 35). All groups of tfdA-like genes are potentially linked to the degradation of PAA herbicides, although alternative primary functions of group 2 and 3 TfdAs have been proposed (30, 35). However, recent cultivation-independent studies focused on 16S rRNA genes or solely on group 1 tfdA sequences in soil (e.g., references 3-5, 13, and 41). Whether group 2 and 3 tfdA-like genes are also quantitatively linked to the degradation of PAA herbicides in soils is unknown. Thus, tools to target a broad range of tfdA-like genes are needed to resolve such an issue. Primers used to assess the diversity of tfdA-like sequences used in previous studies were based on the alignment of approximately 50% or less of available sequences to date (3, 20, 29, 32, 39, 47, 58, 73). Primers specifically targeting all major groups of tfdA-like genes to assess and quantify a broad diversity of potential PAA degraders in soil are unavailable. Thus, the objectives of this study were (i) to develop primers specific for all three groups of tfdA-like genes, (ii) to establish quantitative kinetic PCR (qPCR) assays based on such primers for different soil samples, and (iii) to assess the diversity and abundance of tfdA-like genes in soil.  相似文献   

10.
11.
Deleting individual genes for outer surface c-type cytochromes in Geobacter sulfurreducens partially inhibited the reduction of humic substances and anthraquinone-2,6,-disulfonate. Complete inhibition was obtained only when five of these genes were simultaneously deleted, suggesting that diverse outer surface cytochromes can contribute to the reduction of humic substances and other extracellular quinones.Humic substances can play an important role in the reduction of Fe(III), and possibly other metals, in sedimentary environments (6, 34). Diverse dissimilatory Fe(III)-reducing microorganisms (3, 5, 7, 9, 11, 19-22, 25) can transfer electrons onto the quinone moieties of humic substances (38) or the model compound anthraquinone-2,6-disulfonate (AQDS). Reduced humic substances or AQDS abiotically reduces Fe(III) to Fe(II), regenerating the quinone. Electron shuttling in this manner can greatly increase the rate of electron transfer to insoluble Fe(III) oxides, presumably because soluble quinone-containing molecules are more accessible for microbial reduction than insoluble Fe(III) oxides (19, 22). Thus, catalytic amounts of humic substances have the potential to dramatically influence rates of Fe(III) reduction in soils and sediments and can promote more rapid degradation of organic contaminants coupled to Fe(III) reduction (1, 2, 4, 10, 24).To our knowledge, the mechanisms by which Fe(III)-reducing microorganisms transfer electrons to humic substances have not been investigated previously for any microorganism. However, reduction of AQDS has been studied using Shewanella oneidensis (17, 40). Disruption of the gene for MtrB, an outer membrane protein required for proper localization of outer membrane cytochromes (31), inhibited reduction of AQDS, as did disruption of the gene for the outer membrane c-type cytochrome, MtrC (17). However, in each case inhibition was incomplete, and it was suggested that there was a possibility of some periplasmic reduction (17), which would be consistent with the ability of AQDS to enter the cell (40).The mechanisms for electron transfer to humic substances in Geobacter species are of interest because molecular studies have frequently demonstrated that Geobacter species are the predominant Fe(III)-reducing microorganisms in sedimentary environments in which Fe(III) reduction is an important process (references 20, 32, and 42 and references therein). Geobacter sulfurreducens has routinely been used for investigations of the physiology of Geobacter species because of the availability of its genome sequence (29), a genetic system (8), and a genome-scale metabolic model (26) has made it possible to take a systems biology approach to understanding the growth of this organism in sedimentary environments (23).  相似文献   

12.
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   

13.
14.
Sulfate-reducing bacteria (SRB) play a major role in the coupled biogeochemical cycling of sulfur and chalcophilic metal(loid)s. By implication, they can exert a strong influence on the speciation and mobility of multiple metal(loid) contaminants. In this study, we combined DsrAB gene sequencing and sulfur isotopic profiling to identify the phylogeny and distribution of SRB and to assess their metabolic activity in salt marsh sediments exposed to acid mine drainage (AMD) for over 100 years. Recovered dsrAB sequences from three sites sampled along an AMD flow path indicated the dominance of a single Desulfovibrio species. Other major sequence clades were related most closely to Desulfosarcina, Desulfococcus, Desulfobulbus, and Desulfosporosinus species. The presence of metal sulfides with low δ34S values relative to δ34S values of pore water sulfate showed that sediment SRB populations were actively reducing sulfate under ambient conditions (pH of ∼2), although possibly within less acidic microenvironments. Interestingly, δ34S values for pore water sulfate were lower than those for sulfate delivered during tidal inundation of marsh sediments. 16S rRNA gene sequence data from sediments and sulfur isotope data confirmed that sulfur-oxidizing bacteria drove the reoxidation of biogenic sulfide coupled to oxygen or nitrate reduction over a timescale of hours. Collectively, these findings imply a highly dynamic microbially mediated cycling of sulfate and sulfide, and thus the speciation and mobility of chalcophilic contaminant metal(loid)s, in AMD-impacted marsh sediments.Salt marshes exhibit high primary production rates (1, 101) and form biogeochemical “transition zones” for nutrient production, transport, and cycling between terrestrial and coastal marine environments (41, 66, 100). These zones also serve to reduce the flux of potentially toxic metals in contaminated groundwater to estuaries (12, 99, 106). Both functions depend strongly on microbial activity, especially that of sulfate-reducing bacteria (SRB) (42, 62, 67). SRB recycle much of the sedimentary organic carbon pool in marsh sediments (42-44) and indirectly inhibit production of the greenhouse gas methane (37, 71). They can restrict the mobility of dissolved contaminant metals by inducing precipitation of poorly soluble metal sulfides, and studies have examined their use in constructed wetlands to bioremediate acid mine drainage (AMD) and other metalliferous waste streams (11, 35, 40, 46, 50, 76, 90, 94, 104). However, the high acidity and metal concentrations inherent to AMD can inhibit SRB growth (15, 88, 98), and preferential growth of iron- and sulfur-oxidizing bacteria over SRB has been observed in some treatment wetlands (39).For natural salt marshes, 16S ribosomal nucleic acid- and phospholipid fatty acid (PLFA)-based analyses have shown that SRB commonly comprise a significant fraction of the microbial community (13, 24, 31, 34, 51, 58). Studies of salt marsh dissimilatory sulfite reductase genes (dsrAB), a highly conserved functional phylogenetic marker of prokaryotic sulfate reducers (49, 57, 102, 103, 107), have revealed both novel and deeply branching clades (3). Studies of mining-impacted sites at pH 2.0 to 7.8 (5, 7, 39, 70, 72, 77, 84), of soils and geothermal settings at a pH of ∼4 (55, 68), of metal-contaminated estuaries at pH 6.8 to 7.2 (65), and of hypersaline lakes at pH 7.5 (56) further outline the distribution and tolerance of specific groups and species of SRB under geochemically stringent conditions. Other findings point toward the existence of deltaproteobacteria in environments at a pH of ∼1 (10), although it is unknown if these include SRB. SRB diversity in salt marshes under long-term contamination by AMD has not been well investigated. Such studies may provide useful information for bioremediation projects in estuarine environments, as well as general insights into relationships between SRB physiology and the geochemistry of AMD.We studied the diversity of SRB, based on phylogenetic analysis of recovered DsrAB gene sequences (∼1.9 kb), in natural salt marsh sediments of the San Francisco Bay impacted by AMD for over 100 years. Sulfur isotope ratio and concentration measurements of pore water sulfate and metal sulfide minerals provided information about the spatial and temporal extent of active bacterial sulfate reduction (BSR) in sediment cores taken from specific sites along an AMD flow path. Collectively, the results revealed a tidal marsh system characterized by rapidly cycling bacterial sulfate reduction and sulfide reoxidation associated with oscillating tidal inundation and groundwater infiltration.  相似文献   

15.
16.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

17.
Glutamate contributes to the acid tolerance response (ATR) of many Gram-negative and Gram-positive bacteria, but its role in the ATR of the oral bacterium Streptococcus mutans is unknown. This study describes the discovery and characterization of a glutamate transporter operon designated glnQHMP (Smu.1519 to Smu.1522) and investigates its potential role in acid tolerance. Deletion of glnQHMP resulted in a 95% reduction in transport of radiolabeled glutamate compared to the wild-type UA159 strain. The addition of glutamate to metabolizing UA159 cells resulted in an increased production of acidic end products, whereas the glnQHMP mutant produced less lactic acid than UA159, suggesting a link between glutamate metabolism and acid production and possible acid tolerance. To investigate this possibility, we conducted a microarray analysis with glutamate and under pH 5.5 and pH 7.5 conditions which showed that expression of the glnQHMP operon was downregulated by both glutamate and mild acid. We also measured the growth kinetics of UA159 and its glnQHMP-negative derivative at pH 5.5 and found that the mutant doubled at a much slower rate than the parent strain but survived at pH 3.5 significantly better than the wild type. Taken together, these findings support the involvement of the glutamate transporter operon glnQHMP in the acid tolerance response in S. mutans.Streptococcus mutans is 1 of over 700 bacterial species commonly found in the oral environment (1). Its ability to rapidly metabolize dietary carbohydrates to acid end products causes demineralization of the tooth enamel, leading to caries formation (19). Acidogenicity (the ability to produce acid end products via glycolysis) and aciduricity (the ability to survive and grow in acidic environments) are two important virulence factors of S. mutans. Maintenance of a pH gradient across the cell membrane by increasing intracellular pH by 0.5 to 1.0 relative to the extracellular pH (ΔpH) when exposed to a low pH environment is critical for the survival of S. mutans at low pH. This is primarily accomplished by acid-induced mechanisms that facilitate proton extrusion via the proton-translocating ATPase (5, 20) and by acid end product efflux (8, 12). S. mutans also possesses an acid tolerance response (ATR) mechanism, whereby preexposure to sublethal pH environments (e.g., pH 5.5) affords protection from killing under lethal pH values as low as pH 3.0 (7). This adaptive process is characterized by increased acid resistance (4), increased glycolytic capacities (20), and increased proton-translocating enzyme F1F0-ATPase activity (44). The ATR is enhanced by sugar starvation and the addition of amino acids (48), the addition of potassium ions (12), growth in biofilms, and activity of multiple two-component signal transduction systems that include the ComDE, HK11/RR11 (also designated LiaS/LiaR), VicKR, CiaHR, LevSR, ScnKR, and HK1037/RR1038 (6, 17, 31, 32, 46).Previously, Noji et al. and Sato et al. described a glutamate/aspartate transporter in S. mutans (38, 45). Those researchers showed that the presence of potassium ions was required for transport and that, in environments of pH 6.0 or below, the activity of the H+-ATPase system was required (38, 45). Potassium ions are the main cations in plaque (50), and potassium uptake is associated with intracellular pH homeostasis in S. mutans (24, 35). In addition, expression of several genes involved in the glutamate synthesis pathway (icd, citZ, and acn) are downregulated under low pH (10), suggesting a link between glutamate metabolism, potassium levels, and aciduricity in S. mutans. Since acid tolerance is an important virulence property of S. mutans, we aimed to investigate a possible link between glutamate uptake and acid resistance in this oral pathogen. In bacteria, intracellular glutamate and glutamine levels are closely linked with nitrogen metabolism of the cell. Glutamine is synthesized from glutamate and ammonium, which is a major way for cells to assimilate the nitrogen required for biosynthesis of all amino acids, thus affecting protein synthesis and the structural and functional integrity of the cell. Notably, nitrogen metabolism, especially glutamine metabolism, has been linked to virulence in a number of microorganisms, including Streptococcus pneumoniae (26, 42), Staphylococcus aureus (41), Candida albicans (33), and Pseudomonas aeruginosa (51). Glutamate uptake and metabolism are known to be involved in the ATR of Gram-negative bacteria such as Escherichia coli via the use of glutamate decarboxylase and the glutamate/gamma-amino butyrate (glutamate/GABA) antiporter (9). Similarly, the homologous proteins of these systems in Lactococcus lactis, encoded by the gadBC genes, were shown to assist in a glutamate-dependent acid-resistance mechanism in that Gram-positive bacterium (44).In this study, we searched the S. mutans UA159 genome for potential glutamine transporter operons. We constructed a deletion mutant (SmuGLT) of the glnQHMP operon (Smu.1519 to Smu.1522) and confirmed its role as a glutamate transporter. The inability of SmuGLT to take up glutamate resulted in a general growth deficiency, especially at pH 5.5, as well as an increased tolerance to acid. Results from this study provide insight into the ATR of S. mutans, including a potential link between glutamate metabolism and acid resistance in S. mutans.  相似文献   

18.
The acid-dependent disassembly of foot-and-mouth disease virus (FMDV) is required for viral RNA release from endosomes to initiate replication. Although the FMDV capsid disassembles at acid pH, mutants escaping inhibition by NH4Cl of endosomal acidification were found to constitute about 10% of the viruses recovered from BHK-21 cells infected with FMDV C-S8c1. For three of these mutants, the degree of NH4Cl resistance correlated with the sensitivity of the virion to acid-induced inactivation of its infectivity. Capsid sequencing revealed the presence in each of these mutants of a different amino acid substitution (VP3 A123T, VP3 A118V, and VP2 D106G) that affected a highly conserved residue among FMDVs located close to the capsid interpentameric interfaces. These residues may be involved in the modulation of the acid-induced dissociation of the FMDV capsid. The substitution VP3 A118V present in mutant c2 was sufficient to confer full resistance to NH4Cl and concanamycin A (a V-ATPase inhibitor that blocks endosomal acidification) as well as to increase the acid sensitivity of the virion to an extent similar to that exhibited by mutant c2 relative to the sensitivity of the parental virus C-S8c1. In addition, the increased propensity to dissociation into pentameric subunits of virions bearing substitution VP3 A118V indicates that this replacement also facilitates the dissociation of the FMDV capsid.Foot-and-mouth disease virus (FMDV) is a member of the Aphthovirus genus in the family Picornaviridae. FMDV displays epithelial tropism and is responsible for a highly contagious disease of cloven-hoofed animals (23, 60). FMDV populations are quasispecies and exhibit a high potential for variation and adaptation, one consequence of which is the extensive antigenic diversity of this virus, reflected in the existence of seven serotypes and multiple antigenic variants (reviewed in references 17 and 60). Different cellular receptors, including αvβ integrins and heparan sulfate (HS) glycosaminoglycans, have been described for natural isolates and tissue culture-adapted FMDVs (3, 4, 6, 28-31, 56). However, viruses that are infectious in vivo use integrins as receptors (28). The interaction between FMDV and the integrin molecule is mediated by an Arg-Gly-Asp (RGD) triplet located at the G-H loop of capsid protein VP1 (9, 47). FMDV isolates interacting with integrins gain entry into the cell following clathrin-mediated endocytosis (8, 39, 52). On the other hand, it has been described that a genetically engineered HS-binding mutant uses caveolae to enter into cultured cells (51). After internalization, FMDV must release its genomic RNA molecule of positive polarity into the host cell cytoplasm to establish a productive infection. Early work showed that a variety of lysosomotropic agents, such as weak bases and ionophores that block acidification of endosomes, inhibit FMDV infection (5, 11-13), indicating that genome release is dependent on endosomal acidification. In addition, internalized FMDV particles colocalize with markers from early and recycling endosomes (8, 51, 52) and FMDV infection is reduced by expression of a dominant negative mutant of Rab5 (33), suggesting that FMDV may release its genome from these compartments.The FMDV capsid comprises 60 copies of each of the four structural proteins (VP1 to VP4) arranged in an icosahedral lattice of 12 pentameric subunits. FMDV particles are highly acid labile and disassemble at pH values slightly below neutrality (13). Acid lability is not a feature of the capsids of other picornaviruses, such as Enterovirus. Pentameric subunits are intermediates of FMDV assembly and disassembly (64). A high density of His residues is found close to the interpentameric interface. Protonation of these residues at the acidic pH in the endosomes has been proposed to trigger acid-induced capsid disassembly by electrostatic repulsion between the protonated His side chains (1). His 142 (H142) in VP3 of type A FMDV is involved in a His-α-helix dipole interaction, which is likely to influence the acid lability of FMDV (13). In silico predictions suggested that H142 and H145 in VP3 may have the greatest effect on this process (63). Experimental evidence of the involvement of H142 of VP3 in acid-induced disassembly of FMDV has also been reported (20). Concomitantly with capsid disassembly into pentameric intermediates, internal protein VP4 and viral RNA are released. VP4 is a highly hydrophobic and myristoylated protein (7) whose release has been suggested to mediate membrane permeabilization and ion channel formation, thus facilitating the endosomal exit of viral RNA (15, 16, 34).Besides providing information about the endosomal pH requirements for the release of virus genomes, drugs modifying endosomal acidification can reveal the molecular changes associated with viral resistance to their action. These analyses may also address whether the balance between acid lability and capsid stability required for completion of virus replication allows FMDV, which disassembles at a pH close to neutrality, to escape inhibition by drugs raising the endosomal pH. In this work, we have isolated and characterized FMDV mutants that are able to escape from the inhibition of endosomal acidification exerted by NH4Cl, a lysosomotropic weak base that raises endolysosomal pH and impairs uncoating and infection of viruses that require transit through acidic endosomal compartments for penetration (5, 26, 53). These mutants showed an increased acid lability, which is likely to allow them to uncoat at more-alkaline pH values. A single amino acid substitution close to the interpentameric interfaces in the capsid of one of these mutants was responsible for a total resistance to the elevation in endosomal pH caused by NH4Cl treatment and for the acid-labile phenotype.  相似文献   

19.
20.
Tripartite efflux pumps found in Gram-negative bacteria are involved in antibiotic resistance and toxic-protein secretion. In this study, we show, using site-directed mutational analyses, that the conserved residues located in the tip region of the α-hairpin of the membrane fusion protein (MFP) AcrA play an essential role in the action of the tripartite efflux pump AcrAB-TolC. In addition, we provide in vivo functional data showing that both the length and the amino acid sequence of the α-hairpin of AcrA can be flexible for the formation of a functional AcrAB-TolC pump. Genetic-complementation experiments further indicated functional interrelationships between the AcrA hairpin tip region and the TolC aperture tip region. Our findings may offer a molecular basis for understanding the multidrug resistance of pathogenic bacteria.The tripartite efflux pumps that are found in Gram-negative bacteria have been implicated in their intrinsic resistance to diverse antibiotics, as well as their secretion of protein toxins (10, 12, 24, 31). The bacterial efflux pump is typically assembled from three essential components: an inner membrane transporter (IMT), an outer membrane factor (OMF), and a periplasmic membrane fusion protein (MFP) (10, 12, 24, 31). The IMT provides energy for transporters, like the resistance nodulation cell division (RND) type and the ATP-binding cassette (ABC) type (18). The OMF connects to the IMT in the periplasm, providing a continuous conduit to the external medium. This conduit uses the central channel, which is opened only when in complex with other components (11, 18). The third essential component of the pump is the MFP, which is an adapter protein for the direct interaction between the IMT and OMF in the periplasm (32). The MFP consists of four linearly arranged domains: the membrane-proximal (MP) domain, the β-barrel domain, the lipoyl domain, and the α-hairpin domain (1, 6, 16, 22, 30). The MFP α-hairpin domain is known to interact with OMF, while the other domains are related to interaction with the IMT (15, 22).The Escherichia coli AcrAB-TolC pump, comprised of RND-type IMT-AcrB, MFP-AcrA, and OMF-TolC, is the major contributor to the multidrug resistance phenotype of the bacteria (7, 8, 25). The AcrAB-TolC pump, together with its homolog, the Pseudomonas aeruginosa MexAB-OprM pump (7, 13), has primarily been studied in order to elucidate the molecular mechanisms underlying the actions of the tripartite efflux pumps. Whereas the crystal structures of these proteins have revealed that RND-type IMTs (AcrB and MexB) and OMFs (TolC and OprM) are homotrimeric in their functional states (1, 6, 11, 16, 22, 30), the oligomeric state of MFP remains a topic of debate, despite the presence of crystal structures (3, 5, 17, 18, 22, 27, 30).MacAB-TolC, which was identified as a macrolide-specific extrusion pump (9), has also been implicated in E. coli enterotoxin secretion (29). While MFP-MacA shares high sequence similarity with AcrA and MexA, IMT-MacB is a homodimeric ABC transporter that uses ATP hydrolysis as the driving force (9, 14). MacA forms hexamers, and the funnel-like hexameric structure of MacA is physiologically relevant for the formation of a functional MacAB-TolC pump (30). Although the α-hairpins from AcrA and MacA are commonly involved in the interaction with TolC (30, 32), the interaction mode between AcrA and TolC remains to be elucidated. In this study, we provide experimental evidence showing that the conserved amino acid residues in the AcrA hairpin tip region is important for the action of the AcrAB-TolC efflux pump and is functionally related to the TolC aperture tip region.  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