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.
The fermentative metabolism of Escherichia coli was reengineered to efficiently convert glycerol to succinate under anaerobic conditions without the use of foreign genes. Formate and ethanol were the dominant fermentation products from glycerol in wild-type Escherichia coli ATCC 8739, followed by succinate and acetate. Inactivation of pyruvate formate-lyase (pflB) in the wild-type strain eliminated the production of formate and ethanol and reduced the production of acetate. However, this deletion slowed growth and decreased cell yields due to either insufficient energy production or insufficient levels of electron acceptors. Reversing the direction of the gluconeogenic phosphoenolpyruvate carboxykinase reaction offered an approach to solve both problems, conserving energy as an additional ATP and increasing the pool of electron acceptors (fumarate and malate). Recruiting this enzyme through a promoter mutation (pck*) to increase expression also increased the rate of growth, cell yield, and succinate production. Presumably, the high NADH/NAD+ ratio served to establish the direction of carbon flow. Additional mutations were also beneficial. Glycerol dehydrogenase and the phosphotransferase-dependent dihydroxyacetone kinase are regarded as the primary route for glycerol metabolism under anaerobic conditions. However, this is not true for succinate production by engineered strains. Deletion of the ptsI gene or any other gene essential for the phosphotranferase system was found to increase succinate yield. Deletion of pflB in this background provided a further increase in the succinate yield. Together, these three core mutations (pck*, ptsI, and pflB) effectively redirected carbon flow from glycerol to succinate at 80% of the maximum theoretical yield during anaerobic fermentation in mineral salts medium.Renewable bioenergy offers the potential to solve many environmental problems associated with petroleum-based fuels and chemicals. Biodiesel is produced by reacting vegetable oil or animal fat with alcohol (methanol or ethanol) and used as a transportation fuel in many countries (33). Glycerol is formed as an abundant waste product with limited commercial uses. As the worldwide production of biodiesel continues to increase, the development of effective uses for glycerol may prove essential for the economics and competitiveness of the biodiesel industry. The value of glycerol waste from biodiesel is similar to that of sugars currently used to produce fuel ethanol. Bioconversion of glycerol to higher-value products that replace petroleum, such as polymers, surfactants, solvents, and chemical intermediates, represents an opportunity to decrease waste and improve the economics of the biodiesel industry (5).Many previous investigations have focused on the fermentative production of 1,3-propanediol (1,3-PD) from glycerol (2, 26, 35). Microorganisms including Klebsiella (14), Citrobacter (6), Enterobacter (1), Lactobacillus (29), and Clostridium (10, 28) have the native ability to ferment glycerol into this product. Dupont and Genencor have commercialized a 1,3-PD-based polyester, a condensation product of 1,3-PD and terephthalic acid using glucose as the feedstock. Potential demand for this polymer is estimated to be 1 billion to 2 billion pounds per year over the next 10 years (26). Other investigations of glycerol fermentation have described the production of hydrogen and ethanol (15), polyhydroxyalkanoates (PHAs) (20, 27), glyceric acid (13), and small amounts of succinate (21).Succinic acid is currently used as a specialty chemical in the agricultural, food, and pharmaceutical industries (24, 34). It has also been identified by the U.S. Department of Energy as one of the top 12 building block chemicals (31) because it can be converted into a wide variety of products, including green solvents, pharmaceutical products, and biodegradable plastics (24, 34). Succinate is primarily produced from petroleum-derived maleic anhydride. Recent increases in the petroleum price have generated considerable interest in the fermentative production of succinate from sugars using either natural succinate-producing rumen bacteria or metabolically engineered Escherichia coli strains (24, 36, 38). Succinate can also be produced from glycerol by rumen bacteria, such as Anaerobiospirillum succiniciproducens (21). However, these strains require complex nutrients that increase costs of production, purification, and waste treatment.E. coli has been previously engineered for the commercial production of 1,3-PD from sugars by Dupont and Genecor (26). It is an excellent organism for biotechnology applications but was long thought incapable of anaerobic growth on glycerol (23). Recent studies demonstrated that E. coli can ferment glycerol anaerobically (8, 11, 25, 33), and a new model was proposed for glycerol fermentation (11). In this model, glycerol is metabolized through the glycerol dehydrogenase (encoded by gldA) and dihydroxyacetone kinase (encoded by dhaKLM) pathway with the production of ethanol and acetate as primary fermentation products (11). Small amounts of succinate and 1,2-propanediol were also produced. Native genes encoding glycerol dehydrogenase and dihydroxyacetone kinase were expressed from a plasmid to increase the rates of glycerol metabolism and ethanol production (32). Succinate production has also been increased by expressing Clostridium freundii dihydroxyacetone kinase (encoded by dhaKL) (11). However, neither of these enhanced pathways would appear suitable for efficient succinate production due to the absence of net ATP production and the requirement for phosphoenolpyruvate as a phosphoryl donor for dihydroxyacetone, limiting the carboxylation of this intermediate (Fig. (Fig.11).Open in a separate windowFIG. 1.Glycerol uptake and fermentation by E. coli. (A) Native E. coli pathways. Bold black arrows represent dominant fermentation reactions prior to engineering; thin black arrows represent minor fermentation reactions. GlpK and GlpD are thought to function primarily during aerobic metabolism. Pathways are based on current reviews in EcoSal (3, 4, 22), data available in Ecocyc (19), and primary literature (11, 12, 18, 25, 30). (B) Engineered pathway for the fermentative metabolism of glycerol to succinate. Bold black arrows represent the engineered reactions for glycerol fermentation to succinate as the dominant product; thin black arrows represent minor fermentation reactions in the engineered strain. Dashed arrows represent reactions that are not functional due to deletions in ptsI and pflB. Deleted genes are marked with a black X. In native E. coli strains, phosphoenolpyruvate carboxykinase functions during gluconeogenesis to produce phosphoenolpyruvate. Mutational activation of the pck gene (denoted pck*) allows this enzyme to function in the reverse direction and to serve as the dominant carboxylation step, conserving energy as ATP. With this engineered pathway, competing needs for PEP have been eliminated and net ATP production has been increased. PEP is boxed to indicate a common pool. Abbreviations: DHA, dihydroxyacetone; DHAP, dihydroxyacetone 3-phosphate; PEP, phosphoenolpyruvate; G3P, glycerol 3-phosphate; GA3P, glyceraldehydes 3-phosphate.Previous studies in our laboratory (16, 17, 36, 38) have engineered E. coli ATCC 8739 for the efficient production of succinate from glucose by recruiting genes from alternative pathways (36, 38). In this paper, we report the use of a similar approach to engineer strains for succinate production from glycerol in mineral salts medium.  相似文献   

3.
Given its availability and low price, glycerol has become an ideal feedstock for the production of fuels and chemicals. We recently reported the pathways mediating the metabolism of glycerol in Escherichia coli under anaerobic and microaerobic conditions. In this work, we engineer E. coli for the efficient conversion of glycerol to d-lactic acid (d-lactate), a negligible product of glycerol metabolism in wild-type strains. A homofermentative route for d-lactate production was engineered by overexpressing pathways involved in the conversion of glycerol to this product and blocking those leading to the synthesis of competing by-products. The former included the overexpression of the enzymes involved in the conversion of glycerol to glycolytic intermediates (GlpK-GlpD and GldA-DHAK pathways) and the synthesis of d-lactate from pyruvate (d-lactate dehydrogenase). On the other hand, the synthesis of succinate, acetate, and ethanol was minimized through two strategies: (i) inactivation of pyruvate-formate lyase (ΔpflB) and fumarate reductase (ΔfrdA) (strain LA01) and (ii) inactivation of fumarate reductase (ΔfrdA), phosphate acetyltransferase (Δpta), and alcohol/acetaldehyde dehydrogenase (ΔadhE) (strain LA02). A mutation that blocked the aerobic d-lactate dehydrogenase (Δdld) also was introduced in both LA01 and LA02 to prevent the utilization of d-lactate. The most efficient strain (LA02Δdld, with GlpK-GlpD overexpressed) produced 32 g/liter of d-lactate from 40 g/liter of glycerol at a yield of 85% of the theoretical maximum and with a chiral purity higher than 99.9%. This strain exhibited maximum volumetric and specific productivities for d-lactate production of 1.5 g/liter/h and 1.25 g/g cell mass/h, respectively. The engineered homolactic route generates 1 to 2 mol of ATP per mol of d-lactate and is redox balanced, thus representing a viable metabolic pathway.Lactic acid (lactate) and its derivatives have many applications in the food, pharmaceutical, and polymer industries (13, 30). An example is polylactic acid, a renewable, biodegradable, and environmentally friendly polymer produced from d- and l-lactate (19). In this context, biological processes have the advantage of being able to produce chirally pure lactate from inexpensive media containing only the carbon source and mineral salts (43). While lactic acid bacteria traditionally have been used in the production of d-lactate from carbohydrate-rich feedstocks, several laboratories recently have reported alternative biocatalysts (13, 30), many of which are engineered Escherichia coli strains that produce d- or l-lactate (4, 8, 50, 51, 52).Unlike the aforementioned reports, which have dealt with the use of carbohydrates, our work focuses on the use of glycerol as a carbon source for the production of d-lactate. Glycerol has become an inexpensive and abundant substrate due to its generation in large amounts as a by-product of biodiesel and bioethanol production (18, 32, 47). The conversion of glycerol to higher-value products has been proposed as a path to economic viability for the biofuels industry (47). One such product is lactate, whose production could be readily integrated into existing biodiesel and bioethanol facilities, thus establishing true biorefineries.Although many microorganisms are able to metabolize glycerol (25), the use of industrial microbes such as E. coli could greatly accelerate the development of platforms to produce fuels and chemicals from this carbon source. We recently reported on the ability of E. coli to metabolize glycerol under either anaerobic or microaerobic conditions and identified the environmental and metabolic determinants of these processes (9, 11, 28). In one of the studies, the pathways involved in the microaerobic utilization of glycerol were elucidated, and they are shown in Fig. Fig.11 (9). A common characteristic of glycerol metabolism under either anaerobic or microaerobic conditions is the generation of ethanol as the primary product and the negligible production of lactate (6, 9, 11, 28). In the work reported here, the knowledge base created by the aforementioned studies was used to engineer E. coli for the efficient conversion of glycerol to d-lactate in minimal medium. The engineered strains hold great promise as potential biocatalysts for the conversion of low-value glycerol streams to a higher-value product like d-lactate.Open in a separate windowFIG. 1.Pathways involved in the microaerobic utilization of glycerol in E. coli (9). Genetic modifications supporting the metabolic engineering strategies employed in this work are illustrated by thicker lines (overexpression of gldA-dhaKLM, glpK-glpD, and ldhA) or cross bars (disruption of pflB, pta, adhE, frdA, and dld). Broken lines illustrate multiple steps. Relevant reactions are represented by the names of the gene(s) coding for the enzymes: aceEF-lpdA, pyruvate dehydrogenase complex; adhE, acetaldehyde/alcohol dehydrogenase; ackA, acetate kinase; dhaKLM, dihydroxyacetone kinase; dld, respiratory d-lactate dehydrogenase; fdhF, formate dehydrogenase, part of the formate hydrogenlyase complex; frdABCD, fumarate reductase; gldA, glycerol dehydrogenase; glpD, aerobic glycerol-3-phosphate dehydrogenase; glpK, glycerol kinase; hycB-I, hydrogenase 3, part of the formate hydrogenlyase complex; ldhA, fermentative d-lactate dehydrogenase; pflB, pyruvate formate-lyase; pta, phosphate acetyltransferase; pykF, pyruvate kinase. Abbreviations: DHA, dihydroxyacetone; DHAP, DHA phosphate; G-3-P, glycerol-3-phosphate; PEP, phosphoenolpyruvate; PYR, pyruvate; P/O, amount of ATP produced in the oxidative phosphorylation per pair of electrons transferred through the electron transport system; QH2, reduced quinones.  相似文献   

4.
Although the genome of Haloferax volcanii contains genes (flgA1-flgA2) that encode flagellins and others that encode proteins involved in flagellar assembly, previous reports have concluded that H. volcanii is nonmotile. Contrary to these reports, we have now identified conditions under which H. volcanii is motile. Moreover, we have determined that an H. volcanii deletion mutant lacking flagellin genes is not motile. However, unlike flagella characterized in other prokaryotes, including other archaea, the H. volcanii flagella do not appear to play a significant role in surface adhesion. While flagella often play similar functional roles in bacteria and archaea, the processes involved in the biosynthesis of archaeal flagella do not resemble those involved in assembling bacterial flagella but, instead, are similar to those involved in producing bacterial type IV pili. Consistent with this observation, we have determined that, in addition to disrupting preflagellin processing, deleting pibD, which encodes the preflagellin peptidase, prevents the maturation of other H. volcanii type IV pilin-like proteins. Moreover, in addition to abolishing swimming motility, and unlike the flgA1-flgA2 deletion, deleting pibD eliminates the ability of H. volcanii to adhere to a glass surface, indicating that a nonflagellar type IV pilus-like structure plays a critical role in H. volcanii surface adhesion.To escape toxic conditions or to acquire new sources of nutrients, prokaryotes often depend on some form of motility. Swimming motility, a common means by which many bacteria move from one place to another, usually depends on flagellar rotation to propel cells through liquid medium (24, 26, 34). These motility structures are also critical for the effective attachment of bacteria to surfaces.As in bacteria, rotating flagella are responsible for swimming motility in archaea, and recent studies suggest that archaea, like bacteria, also require flagella for efficient surface attachment (37, 58). However, in contrast to bacterial flagellar subunits, which are translocated via a specialized type III secretion apparatus, archaeal flagellin secretion and flagellum assembly resemble the processes used to translocate and assemble the subunits of bacterial type IV pili (34, 38, 54).Type IV pili are typically composed of major pilins, the primary structural components of the pilus, and several minor pilin-like proteins that play important roles in pilus assembly or function (15, 17, 46). Pilin precursor proteins are transported across the cytoplasmic membrane via the Sec translocation pathway (7, 20). Most Sec substrates contain either a class I or a class II signal peptide that is cleaved at a recognition site that lies subsequent to the hydrophobic portion of the signal peptide (18, 43). However, the precursors of type IV pilins contain class III signal peptides, which are processed at recognition sites that precede the hydrophobic domain by a prepilin-specific peptidase (SPase III) (38, 43, 45). Similarly, archaeal flagellin precursors contain a class III signal peptide that is processed by a prepilin-specific peptidase homolog (FlaK/PibD) (3, 8, 10, 11). Moreover, flagellar assembly involves homologs of components involved in the biosynthesis of bacterial type IV pili, including FlaI, an ATPase homologous to PilB, and FlaJ, a multispanning membrane protein that may provide a platform for flagellar assembly, similar to the proposed role for PilC in pilus assembly (38, 44, 53, 54). These genes, as well as a number of others that encode proteins often required for either flagellar assembly or function (flaCDEFG and flaH), are frequently coregulated with the flg genes (11, 26, 44, 54).Interestingly, most sequenced archaeal genomes also contain diverse sets of genes that encode type IV pilin-like proteins with little or no homology to archaeal flagellins (3, 39, 52). While often coregulated with pilB and pilC homologs, these genes are never found in clusters containing the motility-specific flaCDEFG and flaH homologs; however, the proteins they encode do contain class III signal peptides (52). Several of these proteins have been shown to be processed by an SPase III (4, 52). Moreover, in Sulfolobus solfataricus and Methanococcus maripaludis, some of these archaeal type IV pilin-like proteins were confirmed to form surface filaments that are distinct from the flagella (21, 22, 56). These findings strongly suggest that the genes encode subunits of pilus-like surface structures that are involved in functions other than swimming motility.In bacteria, type IV pili are multifunctional filamentous protein complexes that, in addition to facilitating twitching motility, mediate adherence to abiotic surfaces and make close intercellular associations possible (15, 17, 46). For instance, mating between Escherichia coli in liquid medium has been shown to require type IV pili (often referred to as thin sex pili), which bring cells into close proximity (29, 30, 57). Recent work has shown that the S. solfataricus pilus, Ups, is required not only for efficient adhesion to surfaces of these crenarchaeal cells but also for UV-induced aggregation (21, 22, 58). Frols et al. postulate that autoaggregation is required for DNA exchange under these highly mutagenic conditions (22). Halobacterium salinarum has also been shown to form Ca2+-induced aggregates (27, 28). Furthermore, conjugation has been observed in H. volcanii, which likely requires that cells be held in close proximity for a sustained period to allow time for the cells to construct the cytoplasmic bridges that facilitate DNA transfer between them (35).To determine the roles played by haloarchaeal flagella and other putative type IV pilus-like structures in swimming and surface motility, surface adhesion, autoaggregation, and conjugation, we constructed and characterized two mutant strains of H. volcanii, one lacking the genes that encode the flagellins and the other lacking pibD. Our analyses indicate that although this archaeon was previously thought to be nonmotile (14, 36), wild-type (wt) H. volcanii can swim in a flagellum-dependent manner. Consistent with the involvement of PibD in processing flagellins, the peptidase mutant is nonmotile. Unlike nonhalophilic archaea, however, the flagellum mutant can adhere to glass as effectively as the wild type. Conversely, the ΔpibD strain fails to adhere to glass surfaces, strongly suggesting that in H. volcanii surface adhesion involves nonflagellar, type IV pilus-like structures.  相似文献   

5.
Glycerol is one of the few carbon sources that can be utilized by Mycoplasma pneumoniae. Glycerol metabolism involves uptake by facilitated diffusion, phosphorylation, and the oxidation of glycerol 3-phosphate to dihydroxyacetone phosphate, a glycolytic intermediate. We have analyzed the expression of the genes involved in glycerol metabolism and observed constitutive expression irrespective of the presence of glycerol or preferred carbon sources. Similarly, the enzymatic activity of glycerol kinase is not modulated by HPr-dependent phosphorylation. This lack of regulation is unique among the bacteria for which glycerol metabolism has been studied so far. Two types of enzymes catalyze the oxidation of glycerol 3-phosphate: oxidases and dehydrogenases. Here, we demonstrate that the enzyme encoded by the M. pneumoniae glpD gene is a glycerol 3-phosphate oxidase that forms hydrogen peroxide rather than NADH2. The formation of hydrogen peroxide by GlpD is crucial for cytotoxic effects of M. pneumoniae. A glpD mutant exhibited a significantly reduced formation of hydrogen peroxide and a severely reduced cytotoxicity. Attempts to isolate mutants affected in the genes of glycerol metabolism revealed that only the glpD gene, encoding the glycerol 3-phosphate oxidase, is dispensable. In contrast, the glpF and glpK genes, encoding the glycerol facilitator and the glycerol kinase, respectively, are essential in M. pneumoniae. Thus, the enzymes of glycerol metabolism are crucial for the pathogenicity of M. pneumoniae but also for other essential, yet-to-be-identified functions in the M. pneumoniae cell.Mycoplasma pneumoniae causes infections of the upper and lower respiratory tracts. These bacteria are responsible for a large fraction of community-acquired pneumonias. Although usually harmless for adult patients, M. pneumoniae may cause severe disease in children or elderly people. In addition, M. pneumoniae is involved in extrapulmonary complications such as pediatric encephalitis and erythema multiforme (for reviews, see references 15, 21, and 34).M. pneumoniae and its relatives, the Mollicutes, are all characterized by the lack of a cell wall and a very close adaptation to a life within a eukaryotic host. This close adaptation is reflected by degenerative genome evolution that resulted in an extreme genome reduction. As a result, the Mollicutes are the organisms that are capable of independent life with the smallest known genome. M. pneumoniae has a genome of 816 kb and encodes only 688 proteins (18). This genome reduction is taken even further in the close relative Mycoplasma genitalium, which has only 482 protein-coding genes (18). Thus, the analysis of the Mollicutes allows us to study a minimal form of natural life. This question has recently attracted much interest and resulted in the determination of the essential gene sets of M. pneumoniae, M. genitalium, and, more recently, Mycoplasma pulmonis (6, 20). In M. genitalium, with the most reduced genomes, only 100 out of the 482 protein-coding genes are dispensable, suggesting that the remaining 382 genes form the essential gene set (7).Reductive genome evolution in M. pneumoniae is still under way: the genes for the utilization of mannitol as a carbon source seem to be present in M. pneumoniae; however, this substrate cannot be used by the bacteria. M. genitalium, which is further advanced in genome reduction, has lost the genes for mannitol transport and oxidation. It was therefore suggested that the genes for mannitol utilization in M. pneumoniae either are not expressed or encode inactive proteins (12).In M. pneumoniae as well as in other Mollicutes, pathogenicity is closely linked to carbon metabolism (13). M. pneumoniae can use glucose, fructose, and glycerol as the only carbon sources (12). Studies with Mycoplasma mycoides revealed that glycerol metabolism has a major impact on the pathogenicity of these bacteria. Oxidation of glycerol involves the glycerol 3-phosphate oxidase, which produces hydrogen peroxide rather than NADH2, which is generated by the glycerol 3-phosphate dehydrogenase in most other bacteria (28). In addition to the induction of autoimmune responses, the formation of hydrogen peroxide is the only established mechanism by which mycoplasmas cause damage to their hosts (31, 34). Pathogenic strains of M. mycoides possess a highly active ABC transport system for glycerol in addition to the ubiquitous glycerol facilitator (33). The efficient formation of hydrogen peroxide by the membrane-bound glycerol 3-phosphate oxidase is the major virulence factor of the highly pathogenic strains of M. mycoides (28).M. pneumoniae possesses the complete set of genes for glycerol utilization, and the bacteria do indeed use this carbon source (12). The first component in glycerol metabolism is the glycerol facilitator encoded by the glpF gene. The transported glycerol is then phosphorylated by the glycerol kinase (product of glpK), and glycerol 3-phosphate is subsequently oxidized to dihydroxyacetone phosphate, a glycolytic intermediate. The relevant enzyme is annotated as glycerol 3-phosphate dehydrogenase (encoded by the gene glpD) in M. pneumoniae (17).In all organisms studied so far, glycerol metabolism is under dual control: the genes involved in glycerol utilization are expressed only if glycerol or glycerol 3-phosphate is present in the medium, and they are not expressed in the presence of glucose, the preferred carbon source (3, 4). This second mode of regulation, carbon catabolite repression, involves two distinct mechanisms in the Firmicutes, from which the Mollicutes evolved. In the presence of preferred sugars, the CcpA repressor protein binds in the promoter regions of glycerol utilization genes and prevents their expression. Moreover, the molecular inducer of the system, glycerol 3-phosphate, is formed only in the absence of glucose. This results from the low activity of the glycerol kinase. This enzyme is activated upon phosphorylation by HPr, a protein of the phosphoenolpyruvate:sugar phosphotransferase system (PTS). HPr can phosphorylate other proteins only in the absence of glucose, thus providing a link between glucose availability, the activity of the glycerol kinase, and the induction of the glycerol utilization genes (3). Nothing is known about the regulation of glycerol utilization in any member of the Mollicutes; however, regulatory events seem to be rare in these organisms due to the lack of regulatory proteins, among them CcpA.In this work, we studied the mechanisms of glycerol utilization in M. pneumoniae, its regulation, and its contribution to cytotoxicity. We demonstrate constitutive expression of the genes for glycerol utilization in M. pneumoniae. As observed in M. mycoides, glycerol 3-phosphate oxidation involves the formation of hydrogen peroxide and is important for damaging the host cells.  相似文献   

6.
Bioreactor cultures of Escherichia coli recombinants carrying phaBAC and phaP of Azotobacter sp. FA8 grown on glycerol under low-agitation conditions accumulated more poly(3-hydroxybutyrate) (PHB) and ethanol than at high agitation, while in glucose cultures, low agitation led to a decrease in PHB formation. Cells produced smaller amounts of acids from glycerol than from glucose. Glycerol batch cultures stirred at 125 rpm accumulated, in 24 h, 30.1% (wt/wt) PHB with a relative molecular mass of 1.9 MDa, close to that of PHB obtained using glucose.Polyhydroxyalkanoates (PHAs), accumulated as intracellular granules by many bacteria under unfavorable conditions (5, 8), are carbon and energy reserves and also act as electron sinks, enhancing the fitness of bacteria and contributing to redox balance (9, 11, 19). PHAs have thermoplastic properties, are totally biodegradable by microorganisms present in most environments, and can be produced from different renewable carbon sources (8).Poly(3-hydroxybutyrate) (PHB) is the best known PHA, and its accumulation in recombinant Escherichia coli from several carbon sources has been studied (1, 13). In the last few years, increasing production of biodiesel has caused a sharp fall in the cost of its main by-product, glycerol (22). Its use for microbial PHA synthesis has been analyzed for natural PHA producers, such as Methylobacterium rhodesianum, Cupriavidus necator (formerly called Ralstonia eutropha) (3), several Pseudomonas strains (22), the recently described bacterium Zobellella denitrificans (7), and a Bacillus sp. (18), among others. Glycerol has also been used for PHB synthesis in recombinant E. coli (12, 15). PHAs obtained from glycerol were reported to have a significantly lower molecular weight than polymer synthesized from other substrates, such as glucose or lactose (10, 23).Apart from the genes that catalyze polymer biosynthesis, natural PHA producers have several genes that are involved in granule formation and/or have regulatory functions, such as phasins, granule-associated proteins that have been shown to enhance polymer synthesis and the number and size of PHA granules (17, 24). The phasin PhaP has been shown to exert a beneficial effect on bacterial growth and PHB accumulation from glycerol in bioreactor cultures of strain K24KP, a recombinant E. coli that carries phaBAC and phaP of Azotobacter sp. FA8 (6).Because the redox state of the cells is known to affect the synthesis of PHB (1, 4, 14), the present study investigates the behavior of this recombinant strain under different aeration conditions, by using two substrates, glucose and glycerol, with different oxidation states.  相似文献   

7.
Research into archaea will not achieve its full potential until systems are in place to carry out genetics and biochemistry in the same species. Haloferax volcanii is widely regarded as the best-equipped organism for archaeal genetics, but the development of tools for the expression and purification of H. volcanii proteins has been neglected. We have developed a series of plasmid vectors and host strains for conditional overexpression of halophilic proteins in H. volcanii. The plasmids feature the tryptophan-inducible p.tnaA promoter and a 6×His tag for protein purification by metal affinity chromatography. Purification is facilitated by host strains, where pitA is replaced by the ortholog from Natronomonas pharaonis. The latter lacks the histidine-rich linker region found in H. volcanii PitA and does not copurify with His-tagged recombinant proteins. We also deleted the mrr restriction endonuclease gene, thereby allowing direct transformation without the need to passage DNA through an Escherichia coli dam mutant.Over the past century, our understanding of fundamental biological processes has grown exponentially, and this would have been impossible without the use of organisms that are amenable to experimental manipulation. Model species, such as Escherichia coli, Saccharomyces cerevisiae, Caenorhabditis elegans, Drosophila melanogaster, Mus musculus, and Arabidopsis thaliana, have become a byword for scientific progress (15). The rational choice of a model organism is critically important, and certain features are taken for granted, such as ease of cultivation, a short generation time, and systems for genetic manipulation. This list has now grown to include a genome sequence and methods for biochemical analysis of purified proteins in vitro.Research into archaea has lagged behind work on bacteria and eukaryotes but has nonetheless yielded profound insights (2). One hurdle has been the paucity of archaeal organisms suitable for both biochemistry and genetics. For example, Methanothermobacter thermautotrophicus is a stalwart of archaeal biochemistry but has proved resistant to even the most rudimentary genetic manipulation (2). Progress has recently been made with another biochemical workhorse, Sulfolobus spp., and a few genetic tools are now available (6, 13, 37). Methanosarcina spp. and Thermococcus kodakaraensis offer alternative systems with an increasing array of techniques (16, 35, 36), but sophisticated genetics has traditionally been the preserve of haloarchaea, of which Haloferax volcanii is the organism of choice (39). It is easy to culture, the genome has been sequenced (19), and there are several selectable markers and plasmids for transformation and gene knockout (3, 7, 31), including a Gateway system (14), as well as reporter genes (20, 33) and a tightly controlled inducible promoter (26).The genetic prowess of H. volcanii is not yet fully matched by corresponding systems for protein overexpression and purification. Like other haloarchaea, H. volcanii grows in high salt concentrations (2 to 5 M NaCl), and to cope with the osmotic potential of such environments, it accumulates high intracellular concentrations of potassium ions (12). Consequently, halophilic proteins are adapted to function at high salt concentrations and commonly feature a large excess of acidic amino acids; the negative surface charge is thought to be critical to solubility (28). This can pose problems for expression in heterologous hosts, such as E. coli, since halophilic proteins can misfold and aggregate under conditions of low ionic strength. The purification of misfolded halophilic enzymes from E. coli has relied on the recovery of insoluble protein from inclusion bodies, followed by denaturation and refolding in hypersaline solutions (8, 11). This approach is feasible only where the protein is well characterized and reconstitution of the active form can be monitored (for example, by an enzymatic assay). Furthermore, archaeal proteins expressed in heterologous bacterial hosts lack posttranslational modifications, such as acetylation or ubiquitination (4, 22), which are critical to understanding their biological function.Systems for expression of halophilic proteins in a native haloarchaeal host are therefore required. A number of studies have successfully purified recombinant proteins with a variety of affinity tags after overexpression in H. volcanii. For example, Humbard et al. employed tandem affinity tagging to purify 20S proteasomal core particles from the native host (23). However, the protein expression constructs used in these studies were custom made and somewhat tailored to the application in question. We report here the development of “generic” plasmid vectors and host strains for conditional overexpression of halophilic proteins in H. volcanii. The plasmids feature a tryptophan-inducible promoter derived from the tnaA gene of H. volcanii (26). We demonstrate the utility of these vectors by overexpressing a hexahistidine-tagged recombinant version of the H. volcanii RadA protein. Purification was greatly facilitated by a host strain in which the endogenous pitA gene was replaced by an ortholog from Natronomonas pharaonis. The latter protein lacks the histidine-rich linker region found in H. volcanii PitA (5) and therefore does not copurify with His-tagged recombinant proteins. Finally, we deleted the mrr gene of H. volcanii, which encodes a restriction enzyme that cleaves foreign DNA methylated at GATC residues. The mrr deletion strain allows direct transformation of H. volcanii without the need to passage plasmid DNA through an E. coli dam mutant (21).  相似文献   

8.
9.
Based on elementary mode analysis, an Escherichia coli strain was designed for efficient conversion of glycerol to ethanol. By using nine gene knockout mutations, the functional space of the central metabolism of E. coli was reduced from over 15,000 possible pathways to a total of 28 glycerol-utilizing pathways that support cell function. Among these pathways are eight aerobic and eight anaerobic pathways that do not support cell growth but convert glycerol into ethanol with a theoretical yield of 0.50 g ethanol/g glycerol. The remaining 12 pathways aerobically coproduce biomass and ethanol from glycerol. The optimal ethanol production depends on the oxygen availability that regulates the two competing pathways for biomass and ethanol production. The coupling between cell growth and ethanol production enabled metabolic evolution of the designed strain through serial dilution that resulted in strains with improved ethanol yields and productivities. In defined medium, the evolved strain can convert 40 g/liter of glycerol to ethanol in 48 h with 90% of the theoretical ethanol yield. The performance of the designed strain is predicted by the property space of remaining elementary modes.With the recent rising prices of fossil fuels, development of alternative renewable fuels, such as biodiesel, has become attractive. However, the increase in biodiesel production generates a surplus of crude glycerol since this compound is an inevitable waste by-product resulting directly from the transesterification of vegetable oils or animal fats. For every 3 mol of biodiesel produced, 1 mol of glycerol (about 5 to 10% weight equivalent of biodiesel) is generated. To maximize the full economic potential of the biodiesel process, it is important to convert crude glycerol into useful chemicals (9, 25).Both chemical conversion and biological conversion of crude glycerol into value-added products have been considered. For instance, chemical conversion based on the etherification of glycerol with either alcohols (methanol or ethanol) or alkenes (isobutene or 2-methylpropene) can produce useful fuels or solvents, or steam reforming of glycerol can result in methanol and hydrogen production (7). Biological conversion utilizes species belonging to the Enterobacteriaceae family, such as Klebsiella pneumoniae (11), Citrobacter freundii (5), Clostridium butyricum (4), and Pantoea agglomerans (3), to convert glycerol to 1,3-propanediol by fermentation.Even though Escherichia coli belongs to the family Enterobacteriaceae, it cannot ferment glycerol due to a lack of the dha regulon encoding glycerol dehydratase (dhaB) and 1,3-propanediol oxidoreductase (dhaT), which constitute the fermentative 1,3-propanediol-producing pathway (18). However, introduction of this pathway from K. pneumoniae into E. coli can facilitate glycerol fermentation (18), presumably because the redox potential is balanced. In fact, it is well documented that a wild-type E. coli strain cannot grow on glycerol anaerobically in defined medium except after addition of specific electron acceptors (18). Such electron acceptors can come from external sources, including nitrate, nitrite, fumarate, dimethyl sulfoxide, or trimethylalamine-N-oxide, or from internal sources through added fermentative pathways, such as the 1,3-propanediol-producing pathway.A recent study suggested the feasibility of fermenting glycerol into fuels and other reduced chemicals by inducing the dormant, native 1,2-propanediol fermentative pathway in E. coli without using external electron acceptors (10, 14). This approach, however, faces several critical challenges, such as low specific growth rates resulting in low chemical productivities and coproduction of unavoidable by-products, such as 1,2-propanediol. The reported specific growth rate appears to limit any practical application since a minimal doubling time of about 17 h results in consumption of only 8 to 10 g/liter glycerol after 110 h (10, 14). In addition, glycerol fermentation apparently worked only when complex components, such as yeast extract, tryptone, or amino acids that are required for biomass synthesis, were added to the medium (10, 14).Here, we took a different approach by employing oxygen as the electron acceptor in well-defined microaerobic growth conditions. We used elementary mode (EM) analysis to rationally design an E. coli cell with minimized metabolic functionality tailored to efficiently convert glycerol to ethanol under these microaerobic growth conditions. EM analysis is a metabolic pathway analysis tool that identifies all pathway options in a metabolic network (16). Knowledge of these pathway options allows rational implementation of only the efficient pathways of interest by removing the inefficient pathways, resulting in a cell with minimal but specialized functionality (20-22, 24). Furthermore, the cell developed is engineered to tightly couple cell growth and ethanol production. This unique characteristic facilitates metabolic evolution of the minimal cell to improve ethanol productivity because faster-growing cells also produce ethanol at a higher rate. We demonstrate here that the performance of the designed strain falls into the range defined by the EMs that are possible.  相似文献   

10.
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   

11.
12.
Rickettsia prowazekii is an obligate intracellular pathogen that possesses a small genome and a highly refined repertoire of biochemical pathways compared to those of free-living bacteria. Here we describe a novel biochemical pathway that relies on rickettsial transport of host cytosolic dihydroxyacetone phosphate (DHAP) and its subsequent conversion to sn-glycerol-3-phosphate (G3P) for synthesis of phospholipids. This rickettsial pathway compensates for the evolutionary loss of rickettsial glycolysis/gluconeogenesis, the typical endogenous source of G3P. One of the components of this pathway is R. prowazekii open reading frame RP442, which is annotated GpsA, a G3P dehydrogenase (G3PDH). Purified recombinant rickettsial GpsA was shown to specifically catalyze the conversion of DHAP to G3P in vitro. The products of the GpsA assay were monitored spectrophotometrically, and the identity of the reaction product was verified by paper chromatography. In addition, heterologous expression of the R. prowazekii gpsA gene functioned to complement an Escherichia coli gpsA mutant. Furthermore, gpsA mRNA was detected in R. prowazekii purified from hen egg yolk sacs, and G3PDH activity was assayable in R. prowazekii lysed-cell extracts. Together, these data strongly suggested that R. prowazekii encodes and synthesizes a functional GpsA enzyme, yet R. prowazekii is unable to synthesize DHAP as a substrate for the GpsA enzymatic reaction. On the basis of the fact that intracellular organisms often avail themselves of resources in the host cell cytosol via the activity of novel carrier-mediated transport systems, we reasoned that R. prowazekii transports DHAP to supply substrate for GpsA. In support of this hypothesis, we show that purified R. prowazekii transported and incorporated DHAP into phospholipids, thus implicating a role for GpsA in vivo as part of a novel rickettsial G3P acquisition pathway for phospholipid biosynthesis.Rickettsia prowazekii is an obligate intracellular pathogen, a select agent, and the etiologic cause of epidemic typhus fever in humans. R. prowazekii is transmitted by the human body louse, Pediculus humanus, and is prevalent in areas of low socioeconomic conditions where hygiene and sanitation are lacking (3, 12, 35). Known reservoirs include humans with recrudescent Brill-Zinsser disease and flying squirrels (17, 29). If they are misdiagnosed and/or improperly treated, typhus infections result in substantial morbidity and mortality (36).Upon entering a eukaryotic host cell, R. prowazekii quickly escapes the endosomal vesicle and replicates directly within the cytoplasm (21, 45, 48). As an obligate intracellular pathogen, R. prowazekii has evolved to exploit the nutrient-rich cytoplasmic growth niche by transporting the end products of various host cell metabolic pathways (7, 10, 39, 42, 50, 52). As a presumed consequence of its reliance on the transport of host cell metabolites, R. prowazekii has lost many of its biosynthetic pathway genes through the process of reductive evolution (2, 4-6). In short, essential metabolites that cannot be synthesized by the rickettsiae de novo are transported from the eukaryotic host cell cytosol. The ongoing loss of R. prowazekii biochemical pathways from the genome is evidenced by the identification of pseudogenes, biochemical pathway remnants that have acquired mutations and that encode nonfunctional proteins (2, 4, 5, 22).Interestingly, there exists another class of rickettsial biochemical pathway remnants that may be mistaken as pseudogenes. We have termed members of this class “functional orphaned enzymes” because the purified enzyme, unlike the product of a pseudogene, exhibits its annotated biochemical activity. However, the other enzymes in the corresponding pathway are absent from the R. prowazekii genome; thus, there is no endogenous enzymatic synthesis of substrate for the functional orphaned enzyme. We hypothesize that rickettsial functional orphaned enzymes are fed their substrate from the host cell cytosol via the activity of novel rickettsial transport systems. Together, functional orphaned enzymes and their cognate transport systems serve as streamlined biosynthetic pathways for required metabolites.Here we present the characterization of a R. prowazekii functional orphaned enzyme, an sn-glycerol-3-phosphate (G3P) dehydrogenase (G3PDH) enzyme, RP442 (or GpsA), that catalyzes the conversion of dihydroxyacetone phosphate (DHAP) to G3P for phospholipid biosynthesis. GpsA appears to be a pathway remnant because R. prowazekii lacks glycolytic and gluconeogenic pathways and thus does not possess the capacity to synthesize DHAP (6, 30, 56). Evidence is presented to suggest that GpsA is a functional enzyme and that R. prowazekii transports and incorporates DHAP into phospholipid, indicating that DHAP transport and GpsA are part of a novel G3P acquisition pathway. This is the first report of DHAP transport by a bacterium.  相似文献   

13.
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   

14.
15.
16.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  相似文献   

17.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

18.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

19.
Glycation, or nonenzymatic glycosylation, is a chemical reaction between reactive carbonyl-containing compounds and biomolecules containing free amino groups. Carbonyl-containing compounds include reducing sugars such as glucose or fructose, carbohydrate-derived compounds such as methylglyoxal and glyoxal, and nonsugars such as polyunsaturated fatty acids. The latter group includes molecules such as proteins, DNA, and amino lipids. Glycation-induced damage to these biomolecules has been shown to be a contributing factor in human disorders such as Alzheimer''s disease, atherosclerosis, and cataracts and in diabetic complications. Glycation also affects Escherichia coli under standard laboratory conditions, leading to a decline in bacterial population density and long-term survival. Here we have shown that as E. coli aged in batch culture, the amount of carboxymethyl lysine, an advanced glycation end product, accumulated over time and that this accumulation was affected by the addition of glucose to the culture medium. The addition of excess glucose or methylglyoxal to the culture medium resulted in a dose-dependent loss of cell viability. We have also demonstrated that glyoxylase enzyme GloA plays a role in cell survival during glycation stress. In addition, we have provided evidence that carnosine, folic acid, and aminoguanidine inhibit glycation in prokaryotes. These agents may also prove to be beneficial to eukaryotes since the chemical processes of glycation are similar in these two domains of life.One factor that may affect the long-term survival of bacterial cells in a population is the level of damage incurred by macromolecules via the nonenzymatic process of glycation, first described by Louis-Camille Maillard (16). The Maillard reaction is responsible for the formation of several compounds identified as advanced glycation end products (AGEs) (9). In vivo this reaction appears to play a role in the aging process, as it leads to slow degradation of molecules. The principal mechanisms of glycation-related damage involve cross-links between proteins and/or DNA, modifying or destroying their functional properties (2, 8, 38). Most studies of glycation have been performed with eukaryotes because of its relationship to aging and disorders such as Alzheimer''s disease and diabetes (6, 21, 30, 42). However, several studies (32, 33) have shown that glycation also takes place in Escherichia coli, affecting protein and DNA of this prokaryote.Many biochemical pathways produce reactive dicarbonyl intermediates, such as glyoxal and methylglyoxal (MG), which can further react with DNA, proteins, or other biomolecules to form AGEs (8, 36). Reaction of glucose with amino groups of proteins and subsequent formation of reactive dicarbonyls via a series of reactions involving Schiff base and Amadori product intermediates have been well documented (40). Methylglyoxal can be formed by spontaneous decomposition of glycolytic triose phosphates such as dihydroxyacetone phosphate (DHAP) (1) or can be produced enzymatically from DHAP by the E. coli enzyme methylglyoxal synthase (MgsA) (12). MG synthesis usually requires an environment low in phosphate and high in DHAP, a situation that occurs most frequently under high-glucose conditions (25, 26). If MG is not degraded, MG accumulation will lead to cell death (12). E. coli maintains pathways for the detoxification of methylglyoxal, including glyoxalase enzymes I and II (encoded by gloA and gloB, respectively), which convert MG to S-lactoyl glutathione and then to d-lactate (12). This system has been proposed to be the predominant MG detoxification system in E. coli (12, 29).Glyoxal is also a toxic dicarbonyl compound capable of damaging cells via AGE formation. One of the AGEs formed in the presence of glyoxal is carboxymethyl lysine (CML), which has been used extensively as a biomarker for aging (11, 20, 31, 39). CML can be formed by different pathways: glucose can be oxidized to glyoxal, which can react with protein to form CML (1, 17); glucose can also react with protein to form fructoselysine (an Amadori product), which can undergo oxidative cleavage to form CML (1). In this study, we investigated CML formation in E. coli growing under standard and glycation-prone laboratory conditions. Since AGE formation may negatively affect cell survival and reproduction during long-term batch culture (35), we hypothesized that CML would accumulate in these cultures as cells progress through stationary phase.One product that may interfere with AGE formation is carnosine (β-alanyl-l-histidine), a naturally occurring dipeptide in many organisms. Although its mechanism of action has not been fully determined, there is evidence that both the free amino group derived from the β-alanine and the imidazole ring of histidine compete with amino groups of proteins in the presence of reactive dicarbonyl compounds (7, 24). In this study we designed assays to determine the effect of carnosine (and other compounds) on survival of cultures of E. coli under a variety of experimental conditions. Additionally, since strains lacking glyoxalase enzymes I and II have a reduced ability to detoxify methylglyoxal, we hypothesized that gloA and/or gloB mutants would require larger amounts of carnosine than would wild-type strains to survive in the presence of this toxic electrophile.  相似文献   

20.
Paenibacillus macerans is one of the species with the broadest metabolic capabilities in the genus Paenibacillus, able to ferment hexoses, deoxyhexoses, pentoses, cellulose, and hemicellulose. However, little is known about glycerol metabolism in this organism, and some studies have reported that glycerol is not fermented. Despite these reports, we found that several P. macerans strains are capable of anaerobic fermentation of glycerol. One of these strains, P. macerans N234A, grew fermentatively on glycerol at a maximum specific growth rate of 0.40 h−1 and was chosen for further characterization. The use of [U-13C]glycerol and further analysis of extracellular metabolites and proteinogenic amino acids via nuclear magnetic resonance (NMR) spectroscopy allowed identification of ethanol, formate, acetate, succinate, and 1,2-propanediol (1,2-PDO) as fermentation products and demonstrated that glycerol is incorporated into cellular components. A medium formulation with low concentrations of potassium and phosphate, cultivation at acidic pH, and the use of a CO2-enriched atmosphere stimulated glycerol fermentation and are proposed to be environmental determinants of this process. The pathways involved in glycerol utilization and synthesis of fermentation products were identified using NMR spectroscopy in combination with enzyme assays. Based on these studies, the synthesis of ethanol and 1,2-PDO is proposed to be a metabolic determinant of glycerol fermentation in P. macerans N234A. Conversion of glycerol to ethanol fulfills energy requirements by generating one molecule of ATP per molecule of ethanol synthesized. Conversion of glycerol to 1,2-PDO results in the consumption of reducing equivalents, thus facilitating redox balance. Given the availability, low price, and high degree of reduction of glycerol, the high metabolic rates exhibited by P. macerans N234A are of paramount importance for the production of fuels and chemicals.Although many microorganisms can metabolize glycerol in the presence of external electron acceptors (respiratory metabolism), few are able to do so fermentatively (i.e., in the absence of electron acceptors). Fermentative metabolism of glycerol has been reported in species of the genera Klebsiella, Citrobacter, Enterobacter, Clostridium, Lactobacillus, Bacillus, Propionibacterium, and Anaerobiospirillum but has been studied more extensively in a few species of the family Enterobacteriaceae, namely, Citrobacter freundii and Klebsiella pneumoniae (6, 9). Glycerol fermentation in these organisms is mediated by a two-branch pathway, which results in the synthesis of the glycolytic intermediate dihydroxyacetone (DHA) phosphate (DHAP) and the fermentation product 1,3-propanediol (1,3-PDO) (Fig. (Fig.1A)1A) (6). In the oxidative branch, glycerol is dehydrogenated to DHA by a type I NAD-linked glycerol dehydrogenase (glyDH). DHA is then phosphorylated by ATP- or phosphoenolpyruvate (PEP)-dependent DHA kinases (DHAKs) to generate DHAP. In the parallel reductive branch, glycerol is dehydrated by glycerol dehydratase, and 3-hydroxypropionaldehyde (3-HPA) is formed. 3-HPA is then reduced to the major fermentation product 1,3-PDO by an NADH-linked 1,3-PDO dehydrogenase (1,3-PDODH), thereby regenerating NAD+ (Fig. (Fig.1A).1A). Organisms that lack the capacity to synthesize 1,3-PDO have been deemed unable to utilize glycerol in a fermentative manner (6, 9, 10). The metabolism of glycerol in these organisms is thought to require an electron acceptor and takes place through a respiratory pathway that involves a glycerol kinase and two respiratory (aerobic and anaerobic) glycerol-3-phosphate dehydrogenases (G3PDHs) (6, 7, 24, 29, 35, 38) (Fig. (Fig.1B).1B). A recent development in this area is the finding that Escherichia coli, an organism that is unable to produce 1,3-PDO, can indeed ferment glycerol in the absence of external electron acceptors (15, 26). In this model, synthesis of the fermentation products 1,2-PDO and ethanol enables glycerol fermentation by facilitating redox balance and ATP generation, respectively (Fig. (Fig.1C)1C) (15). A type II glyDH and a PEP-dependent DHAK mediate the conversion of glycerol to glycolytic intermediates. glyDH also catalyzes the last step in the synthesis of the key fermentation product 1,2-PDO (Fig. (Fig.1C1C).Open in a separate windowFIG. 1.Glycerol metabolism in bacteria. (A) 1,3-PDO model for the fermentative utilization of glycerol. (B) Respiratory metabolism of glycerol (i.e., metabolism in the presence of an electron acceptor). (C) 1,2-PDO-ethanol model for the fermentative utilization of glycerol. Dashed lines indicate multiple steps. glyD, glycerol dehydratase; glyDH-I, type I glyDH; GK, glycerol kinase; ae-G3PDH, aerobic G3PDH; an-G3PDH, anaerobic G3PDH; QH2, reduced quinone; glyDH-II, type II glyDH; FHL, formate hydrogen lyase; ADH, alcohol/acetaldehyde dehydrogenase.Paenibacillus macerans, previously called Bacillus macerans and Bacillus acetoethylicum, is a gram-positive, spore-forming bacterium belonging to the genus Paenibacillus (17) that is capable of fermentative metabolism of hexoses, deoxyhexoses, pentoses, cellulose, and hemicellulose (33, 39, 40, 41). Glycerol, however, is considered a nonfermentable carbon source for P. macerans. The “nonfermentable status” of glycerol has been used to determine whether certain electron acceptors, such as fumarate, trimethylamine N-oxide, nitrate, and nitrite, can mediate anaerobic respiration in this organism (34).In this study we found that several P. macerans strains are able to ferment glycerol in the absence of external electron acceptors. The fermentation of glycerol by one of these strains, P. macerans N234A, occurred at high metabolic rates and in the absence of an active 1,3-PDO pathway. Therefore, the environmental and metabolic determinants of glycerol fermentation in P. macerans N234A were investigated.  相似文献   

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