首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
A lysine racemase (lyr) gene was isolated from a soil metagenome by functional complementation for the first time by using Escherichia coli BCRC 51734 cells as the host and d-lysine as the selection agent. The lyr gene consisted of a 1,182-bp nucleotide sequence encoding a protein of 393 amino acids with a molecular mass of about 42.7 kDa. The enzyme exhibited higher specific activity toward lysine in the l-lysine-to-d-lysine direction than in the reverse reaction.Amino acids are the building blocks of proteins and play an important role in the regulation of the metabolism of living organisms. Among two enantiomers of naturally occurring amino acids, l-amino acids are predominant in living organisms, while d-amino acids are found in both free and bound states in various organisms like bacteria (36), yeasts (35), plants (47), insects (11), mammals (17), bivalves (39), and fish (28). The d-amino acids are mostly endogenous and produced by racemization from their counterparts by the action of a racemase. Thus, the amino acid racemases are involved in d-amino acid metabolism (29, 46). Since the discovery of alanine racemase in 1951 (42), several racemases toward amino acids, such as those for glutamate, threonine, serine, aspartate, methionine, proline, arginine, and phenylalanine, have been reported in bacteria, archaea, and eukaryotes, including mammals (1, 2, 15, 30, 31, 44). They are classified into two groups: pyridoxal 5′-phosphate (PLP)-dependent and PLP-independent enzymes (9, 36).Lysine racemase (Lyr, EC 5.1.1.5) was first reported in Proteus vulgaris ATCC 4669 (19) and proposed to be involved in the lysine degradation of bacterial cells (5, 19). Catabolism of lysine occurs via two parallel pathways. In one of the pathways, δ-aminovalerate is the key metabolite, whereas in the other l-lysine is racemized to d-lysine, and l-pipecolate and α-aminoadipate (AMA) are the key metabolites (5). d-Lysine catabolism proceeds through a series of cyclized intermediates which are necessary to regenerate an α-amino acid and comprise the following metabolites (AMA pathway): d-lysine→α-keto-ɛ-amino caproate→Δ1-piperideine-2-carboxylate→pipecolate→Δ1-piperideine-6-carboxylate→α-amino-δ-formylcaproate→α-AMA→α-ketoadipate (6, 7, 12, 27). The final product is converted to α-ketoglutarate via a series of coenzyme A derivatives and subsequently participates as an intermediate in the Krebs cycle. This pathway suggests that the biological function of d-lysine in the bacteria is that of a carbon or nitrogen source. Racemization of added l-lysine to d-lysine by whole cells of Proteus spp. and Escherichia spp. (19) and by the cell extract of Pseudomonas putida ATCC 15070 (5, 20) has been found. However, the enzyme has not been purified to homogeneity, and thus, its molecular and catalytic characteristics, including its gene structure, have not been elucidated. In this study, we explored a metagenomic library constructed from a garden soil to isolate a novel Lyr enzyme. After expression in Escherichia coli, the purified enzyme was characterized in terms of optimal pH and temperature, thermal stability, and racemization activity.  相似文献   

3.
The uncharacterized gene previously proposed as a mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in Escherichia coli. The maximal activity of the recombinant enzyme was observed at pH 7.5 and 40°C in the presence of 0.5 mM Co2+. The isomerization activity was specific for aldose substrates possessing hydroxyl groups oriented in the same direction at the C-2 and C-3 positions, such as the d and l forms of ribose, lyxose, talose, mannose, and allose. The enzyme exhibited the highest activity for l-ribulose among all pentoses and hexoses. Thus, l-ribose, as a potential starting material for many l-nucleoside-based pharmaceutical compounds, was produced at 213 g/liter from 300-g/liter l-ribulose by mannose-6-phosphate isomerase at 40°C for 3 h, with a conversion yield of 71% and a volumetric productivity of 71 g liter−1 h−1.l-Ribose is a potential starting material for the synthesis of many l-nucleoside-based pharmaceutical compounds, and it is not abundant in nature (5, 19). l-Ribose has been produced mainly by chemical synthesis from l-arabinose, l-xylose, d-glucose, d-galactose, d-ribose, or d-mannono-1,4-lactone (2, 17, 23). Biological l-ribose manufacture has been investigated using ribitol or l-ribulose. Recently, l-ribose was produced from ribitol by a recombinant Escherichia coli containing an NAD-dependent mannitol-1-dehydrogenase (MDH) with a 55% conversion yield when 100 g/liter ribitol was used in a 72-h fermentation (18). However, the volumetric productivity of l-ribose in the fermentation is 28-fold lower than that of the chemical method synthesized from l-arabinose (8). l-Ribulose has been biochemically converted from l-ribose using an l-ribose isomerase from an Acinetobacter sp. (9), an l-arabinose isomerase mutant from Escherichia coli (4), a d-xylose isomerase mutant from Actinoplanes missouriensis (14), and a d-lyxose isomerase from Cohnella laeviribosi (3), indicating that l-ribose can be produced from l-ribulose by these enzymes. However, the enzymatic production of l-ribulose is slow, and the enzymatic production of l-ribose from l-ribulose has been not reported.Sugar phosphate isomerases, such as ribose-5-phosphate isomerase, glucose-6-phosphate isomerase, and galactose-6-phosphate isomerase, work as general aldose-ketose isomerases and are useful tools for producing rare sugars, because they convert the substrate sugar phosphates and the substrate sugars without phosphate to have a similar configuration (11, 12, 21, 22). l-Ribose isomerase from an Acinetobacter sp. (9) and d-lyxose isomerase from C. laeviribosi (3) had activity with l-ribose, d-lyxose, and d-mannose. Thus, we can apply mannose-6-phosphate (EC 5.3.1.8) isomerase to the production of l-ribose, because there are no sugar phosphate isomerases relating to l-ribose and d-lyxose. The production of the expensive sugar l-ribose (bulk price, $1,000/kg) from the rare sugar l-ribulose by mannose-6-phosphate isomerase may prove to be a valuable industrial process, because we have produced l-ribulose from the cheap sugar l-arabinose (bulk price, $50/kg) using the l-arabinose isomerase from Geobacillus thermodenitrificans (20) (Fig. (Fig.11).Open in a separate windowFIG. 1.Schematic representation for the production of l-ribulose from l-arabinose by G. thermodenitrificans l-arabinose isomerase and the production of l-ribose from l-ribulose by B. subtilis mannose-6-phosphate isomerase.In this study, the gene encoding mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in E. coli. The substrate specificity of the recombinant enzyme for various aldoses and ketoses was investigated, and l-ribulose exhibited the highest activity among all pentoses and hexoses. Therefore, mannose-6-phosphate isomerase was applied to the production of l-ribose from l-ribulose.  相似文献   

4.
5.
6.
7.
8.
Clostridium sordellii is a spore-forming, obligately anaerobic, Gram-positive bacterium that can cause toxic shock syndrome after gynecological procedures. Although the incidence of C. sordellii infection is low, it is fatal in most cases. Since spore germination is believed to be the first step in the establishment of Bacilli and Clostridia infections, we analyzed the requirements for C. sordellii spore germination in vitro. Our data showed that C. sordellii spores require three structurally different amino acids and bicarbonate for maximum germination. Unlike the case for Bacilli species, d-alanine had no effect on C. sordellii spore germination. C. sordellii spores germinated only in a narrow pH range between 5.7 and 6.5. In contrast, C. sordellii spore germination was significantly less sensitive to temperature changes than that of the Bacilli. The analysis of the kinetics of C. sordellii spore germination showed strong allosteric behavior in the binding of l-phenylalanine and l-alanine but not in that of bicarbonate or l-arginine. By comparing germinant apparent binding affinities to their known in vivo concentrations, we postulated a mechanism for differential C. sordellii spore activation in the female reproductive tract.Clostridium sordellii is an anaerobic, Gram-positive, spore-forming bacterium that is commonly found in soil and in the intestines of animals (4). Many C. sordellii strains are nonpathogenic; however, virulent strains cause lethal infections in several animal species, such as hemorrhagic enteritis in foals, sheep, and cattle (5, 10, 16, 28), omphalitis in foals (43), and wound infection in humans (4, 35).C. sordellii also can cause life-threatening necrotizing infections after gynecological procedures (4). In addition, fatal cases of C. sordellii endometritis following medical abortion with a mifepristone-misoprostol combination have been reported recently (13, 19, 56). The increased use of mifepristone-misoprostol for medical abortion may result in larger numbers of C. sordellii infections (38, 40).Although C. sordellii rarely has been identified in the genital tract, a correlation between gynecological procedures and C. sordellii-mediated toxic shock syndrome is apparent (19). Pregnancy, childbirth, or abortion may predispose some women to acquire C. sordellii in the vaginal tract (19). Under these conditions, C. sordellii infections result in an almost 100% mortality rate.Since there is no national system for tracking and reporting complications associated with gynecological procedures, the identification of the true rates of reproductive tract infections in women is not readily available (8). Therefore, the number of known C. sordellii-associated infections, although low, may be underreported (19, 29). Furthermore, unsafe abortion practices in developing countries cause large mortality rates due to complicating infections (24, 34). In many cases, however, the causative agent of the abortion-associated sepsis have not been characterized (24). Thus, the worldwide morbidity and mortality associated with C. sordellii infections is not currently known.C. sordellii produces several virulence factors. The two major toxins are the lethal toxin (TcsL) and the hemorrhagic toxin (37, 46). The lethal toxin produced by C. sordellii is causally involved in enteritis of domestic animals and in systemic toxicity following infections of humans (46). Furthermore, TcsL is associated with rapid mortality in C. sordellii endometritis rodent models (26). Interestingly, TcsL cytopathic effects are increased at low pH, a characteristic found in the vaginal tract (48). The hemorrhagic toxin is not well characterized, but it has been reported to cause dermal and intestinal necrosis in guinea pigs (6, 52).C. sordellii, like other Bacilli and Clostridia species, has the ability to form metabolically dormant spores that are extremely resistant to environmental stresses, such as heat, radiation, and toxic chemicals (42, 55). Upon encountering a suitable environment, spores germinate into vegetative cells, the form that is responsible for toxin production and disease onset (39, 54).In most cases, the germination process initially is triggered by the detection of low-molecular-weight germinants by a sensitive biosensor (39, 54). This sensor consists of a proteinaceous germination (Ger) receptor encoded, in general, by a tricistronic operon. Spore germination requirements have been studied most extensively for Bacilli and can be initiated by a variety of factors, including amino acids, sugars, and nucleosides (20, 30).Spore germination in the Clostridia generally requires combinations of multiple germinants. The germination of spores of proteolytic Clostridium botulinum types A and B was triggered by a defined three-component mixture comprised of l-alanine (or l-cysteine), l-lactate (or sodium thioglycolate), and sodium bicarbonate (3). In contrast, the optimum germination of spores of nonproteolytic C. botulinum types B, E, and F required binary combinations of l-alanine-l-lactate, l-cysteine-l-lactate, and l-serine-l-lactate (45).Clostridium difficile is a human pathogen that can cause fulminant colitis (11). Interestingly, C. difficile does not encode any known Ger receptors (53). However, it is likely that germination receptors exist, because C. difficile spores must germinate in order to complete their life cycle. While C. difficile germination receptors remain elusive, the spores of C. difficile germinate in rich medium supplemented with bile salts (62). More recently, taurocholate (a bile salt) and glycine (an amino acid) were shown to act as cogerminants for C. difficile spore germination (57, 61).Clostridium bifermentans is a close relative of C. sordellii (14). The minimum requirement for C. bifermentans spore germination was the presence of l-alanine, l-phenylalanine, and l-lactate (59). In addition, an unknown factor present in yeast extract was suggested to enhance germination (59). However, the Ger receptors involved in C. bifermentans spore germination are not known.Even though many Bacilli and Clostridia species use similar metabolites as germinants, the mechanisms of germinant recognition remain to be elucidated. Unfortunately, the multimeric interactions of Ger receptor complexes and the hydrophobic nature of the Ger receptor subunits have hindered our understanding of the mechanism of germinant recognition.To understand the molecular determinants of germinant recognition, we recently applied kinetic methods to study bacterial spore germination (1, 2, 18). Spore germination can be analyzed quantitatively by fitting optical density (OD) decreases to the Michaelis-Menten equation (2). The kinetic parameters obtained allow the determination of the apparent binding affinity (Km) of spores for the different cogerminants and the maximum rate of spore germination (Vmax). In these instances, Km refers to the concentration of substrate required to reach half of the maximal germination rate. These parameters can, in turn, be used to determine the mechanism of germination and potential interactions between germination receptors. Furthermore, by comparing apparent Km values to germinant concentrations in vivo, models for spore-germinant complex distribution can be proposed, and rate-limiting steps for the germination process can be derived. Thus, kinetic analysis can yield information on spore activation even if the identities of the germination receptors are not known.Using this procedure, we were able to determine the mechanism for Bacillus anthracis germination with inosine and l-alanine. In turn, this information was used to design nucleoside analogs that inhibit B. anthracis spore germination in vitro and protect macrophages from anthrax cytotoxicity (2).Since C. sordellii germination receptors have not been identified, we used chemical probes and kinetic methods to investigate the conditions necessary for spore germination. We found that C. sordellii spores germinate better at slightly acidic pH. Furthermore, germination rates varied slightly from 25 to 40°C. We also found that C. sordellii spores have an absolute requirement for a small amino acid, a basic amino acid, an aromatic amino acid, and bicarbonate (NaHCO3) for efficient germination. Kinetic analysis showed allosteric interaction for the putative l-phenylalanine and l-alanine germination receptors. In contrast, l-arginine or bicarbonate recognition followed typical Michaelis-Menten kinetics. The implication of germinant recognition and host environment is discussed.  相似文献   

9.
In order to achieve direct and efficient fermentation of optically pure d-lactic acid from raw corn starch, we constructed l-lactate dehydrogenase gene (ldhL1)-deficient Lactobacillus plantarum and introduced a plasmid encoding Streptococcus bovis 148 α-amylase (AmyA). The resulting strain produced only d-lactic acid from glucose and successfully expressed amyA. With the aid of secreting AmyA, direct d-lactic acid fermentation from raw corn starch was accomplished. After 48 h of fermentation, 73.2 g/liter of lactic acid was produced with a high yield (0.85 g per g of consumed sugar) and an optical purity of 99.6%. Moreover, a strain replacing the ldhL1 gene with an amyA-secreting expression cassette was constructed. Using this strain, direct d-lactic acid fermentation from raw corn starch was accomplished in the absence of selective pressure by antibiotics. This is the first report of direct d-lactic acid fermentation from raw starch.Poly-lactic acid (PLA) is an important agro-based plastic that can be produced from inexpensive, renewable, and abundantly available biomass resources, including starchy materials. These resources have advantages over limited oil- and fossil-based sources, as they do not result in any net carbon dioxide release to the atmosphere (7). Recently, stereocomplex PLA, which is composed of both poly-l- and -d-lactic acid, has been attracting much attention due to its high thermostability. Stereocomplex-type polymers show a melting point (ca. 230°C) that is approximately 50°C higher than that of the respective single polymers (8). Therefore, d-lactic acid, in addition to l-lactic acid, which has been the focus of production to date, is of significant importance.Lactic acid bacteria (LAB) are promising microorganisms for the efficient production of lactic acid from various sugars, such as glucose, sucrose, and lactose. However, when starchy materials are used as a carbon source, they must be saccharified by physicochemical and enzymatic treatment because most LAB cannot utilize starchy materials directly (13). This makes the whole process less economically viable. Therefore, many researchers have examined the direct production of lactic acid from starchy materials by using wild amylolytic LAB (ALAB) (6, 24, 25) or genetically modified amylase-producing LAB (15, 16). Although d-lactic acid has been produced by fermentation from pretreated substrates such as rice starch (5) and by simultaneous saccharification and fermentation from cellulose (23), there have been no reports on the direct production of d-lactic acid from starchy materials. This is due to a lack of d-lactic acid-producing ALAB and difficulties in gene manipulation of d-lactic acid-producing LAB, such as Lactobacillus delbrueckii (22).We focused on Lactobacillus plantarum, which is an industrially important strain due to its environmental flexibility and its ability to assimilate a wide range of carbohydrates (9). In recent years, several gene manipulation methods for Lactobacillus plantarum have been established (18, 19). Moreover, the complete genome sequence has been decoded for L. plantarum NCIMB 8826 (9). Based on whole-genome analysis, L. plantarum possesses two types of lactate dehydrogenase (LDH), l-LDH and d-LDH, which convert pyruvate into l- and d-lactic acid, respectively. Ferain et al. (4) reported that chromosomal deletion in the ldhL1 gene of L. plantarum NCIMB 8826 provoked an absence of l-LDH activity and produced d-lactic acid from glucose.In the present study, to produce d-lactic acid directly from starch, we constructed an l-LDH-deficient, α-amylase-secreting L. plantarum strain. The engineered strain expressed α-amylase from Streptococcus bovis 148 (AmyA) (20) and efficiently degraded raw starch with the aid of a C-terminal starch-binding domain (11). Using this strain, we achieved the direct and efficient fermentation of optically pure d-lactic acid from raw corn starch.  相似文献   

10.
Laboratory adaptive evolution studies can provide key information to address a wide range of issues in evolutionary biology. Such studies have been limited thus far by the inability of workers to readily detect mutations in evolved microbial strains on a genome scale. This limitation has now been overcome by recently developed genome sequencing technology that allows workers to identify all accumulated mutations that appear during laboratory adaptive evolution. In this study, we evolved Escherichia coli K-12 MG1655 with a nonnative carbon source, l-1,2-propanediol (l-1,2-PDO), for ∼700 generations. We found that (i) experimental evolution of E. coli for ∼700 generations in 1,2-PDO-supplemented minimal medium resulted in acquisition of the ability to use l-1,2-PDO as a sole carbon and energy source so that the organism changed from an organism that did not grow at all initially to an organism that had a growth rate of 0.35 h−1; (ii) six mutations detected by whole-genome resequencing accumulated in the evolved E. coli mutant over the course of adaptive evolution on l-1,2-PDO; (iii) five of the six mutations were within coding regions, and IS5 was inserted between two fuc regulons; (iv) two major mutations (mutations in fucO and its promoter) involved in l-1,2-PDO catabolism appeared early during adaptive evolution; and (v) multiple defined knock-in mutant strains with all of the mutations had growth rates essentially matching that of the evolved strain. These results provide insight into the genetic basis underlying microbial evolution for growth on a nonnative substrate.Evolution of microorganisms in the laboratory offers the possibility of relating acquired mutations to increased fitness of the organism under the conditions used. Complete identification of mutations over defined evolutionary periods is necessary to fully understand the evolutionary change because spontaneous mutation is the foundational biological source of phenotypic variation (52). Since microbes grow rapidly and have large population sizes and since ancestors can be preserved by freezing them for later direct comparison of evolved types, laboratory evolution using microorganisms provides a powerful context for studying the genetics of evolutionary adaptation (5, 12, 14, 19, 43) due to the advent of new technologies for genome-wide detection of mutations (30, 33). A large number of studies of experimental evolution with various microbes have been carried out using natural carbon sources, especially glucose (12, 19, 47, 55), since glucose is the preferred carbon and energy source for most bacteria and eukaryotic cells (4, 50). Recently, a few studies have investigated the adaptive evolution of Escherichia coli at the genetic and metabolic levels with gluconeogenic carbon sources, including lactate (34) and glycerol (20). Compared to experimental evolution with native carbon sources, microorganisms might be more capable of adapting to various nonnative carbon compounds because microorganisms are able to adapt to environmental changes by using a number of strategies to meet their growth requirements and to achieve optimal overall performance in the new conditions (20, 21, 34). However, a comprehensive analysis of the genetic basis of adaptation to nonnative carbon sources has not been performed.The K-12 MG1655 strain of E. coli is not able to utilize l-1,2-propanediol (l-1,2-PDO) as a sole carbon and energy source. However, E. coli has an enzyme, l-1,2-PDO oxidoreductase (POR), which is involved in fermentative l-fucose metabolism and catalyzes the oxidation of l-1,2-PDO to l-lactaldehyde (Fig. (Fig.11 A). The E. coli POR is encoded by the fucO gene of the fucose regulon (11, 23), which consists of two divergent operons (fucAO and fucPIKUR) under positive control of FucR (Fig. (Fig.1B)1B) (9). FucR is activated by fuculose-1-phosphate, which is the inducer of the fuc regulon (3). In E. coli, fucose metabolism is initiated by the sequential actions of a permease (encoded by fucP), an isomerase (encoded by fucI), a kinase (encoded by fucK), and an aldolase (encoded by fucA). The aldolase catalyzes the cleavage of fuculose-1-phosphate to dihydroxyacetone phosphate and l-lactaldehyde. Under aerobic respiratory conditions, l-lactaldehyde is oxidized to l-lactate by an NAD-linked aldehyde dehydrogenase with broad functions (encoded by aldA). l-Lactate is then oxidized to pyruvate by a flavin adenine dinucleotide (FAD)-dependent l-lactate dehydrogenase (encoded by the lldD gene of the lldPRD operon [formerly the lctPRD operon]). Under anaerobic fermentative conditions, however, redox balance requires sacrifice of the l-lactaldehyde as a hydrogen acceptor at the expense of NADH (Fig. (Fig.1A).1A). This reaction is catalyzed by the POR. The terminal fermentation product, l-1,2-PDO, is then released by a permease (57). Although the POR catalyzes the oxidation of l-1,2-PDO to l-lactaldehyde, l-1,2-PDO cannot be utilized by wild-type (WT) E. coli as a sole carbon source under aerobic conditions because this compound cannot induce expression of the fuc regulon (11). Indeed, the fuc regulon was not expressed under any conditions when a database of 213 expression profiles produced in our laboratory was examined (38). Furthermore, even if the POR is expressed, it is oxidatively inactivated by a metal-catalyzed oxidation (MCO) mechanism (7).Open in a separate windowFIG. 1.Metabolic pathway and fuc regulon for l-fucose and l-1,2-PDO. (A) Metabolic pathway for l-fucose and l-1,2-PDO. In E. coli, fucose metabolism is initiated by the sequential actions of a permease (encoded by fucP), an isomerase (encoded by fucI), a kinase (encoded by fucK), and an aldolase (encoded by fucA). The aldolase catalyzes cleavage of fuculose-1-phosphate to dihydroxyacetone phosphate and l-lactaldehyde. Under aerobic respiratory conditions, the l-lactaldehyde is further oxidized by a series of enzymes to pyruvate, which subsequently enters central metabolism. Under anaerobic fermentative conditions, the l-lactaldehyde is reduced to l-1,2-PDO by oxidoreductase (encoded by fucO). (B) Genetic organization of the fuc regulon. The fuc regulon for l-fucose uptake and metabolism consists of two divergent operons, fucAO and fucPIKUR.Sridhara et al. (48) previously described E. coli mutants that were isolated from an E. coli K-12 derivative treated with the mutagen ethyl methanesulfonate and were able to grow aerobically on l-1,2-PDO as a sole carbon source. Previous studies showed that an IS5 insertion between the fucAO and fucPIKUR operons caused constitutive expression of the fucAO operon (9, 41) at a level that enabled the E. coli mutant to grow on l-1,2-PDO. In addition, mutations resulting in increased resistance to MCO under aerobic conditions were found in the N-terminal domain of POR (39). However, at present, little is known about the accumulated genome-wide mutations and their effects on the fitness in E. coli that has acquired the ability to use l-1,2-PDO because previous studies have focused on mutations in POR and its regulatory region.In an attempt to investigate the genetic basis of adaptive evolution of E. coli during growth on l-1,2-PDO, we first isolated an E. coli mutant able to use l-1,2-PDO using experimental evolution without a mutagen, and we then characterized this evolved E. coli mutant. Using whole-genome sequencing, we identified all accumulated mutations of the evolved E. coli mutant related to the known ancestor and also determined the fitness benefits and phenotypic behaviors of the mutations discovered. Our results offer a systematic view of the genetic basis underlying microbial adaptation to a nonnative substrate.  相似文献   

11.
Prolyl dipeptide synthesis by S9 aminopeptidase from Streptomyces thermocyaneoviolaceus (S9AP-St) has been demonstrated. In the synthesis, S9AP-St preferentially used l-Pro-OBzl as the acyl donor, yielding synthesized dipeptides having an l-Pro-Xaa structure. In addition, S9AP-St showed broad specificity toward the acyl acceptor. Furthermore, S9AP-St produced cyclo (l-Pro-l-His) with a conversion ratio of substrate to cyclo (l-Pro-l-His) higher than 40%.Some proline-containing dipeptides and their cyclic analogs exhibit biological activity. For example, cyclo (l-arginyl-d-proline) [c(lR-dP)] is known to act as a specific inhibitor of family 18 chitinase (4, 10). A cyclic peptide, c(lP-lH), produced by the cleavage of thyrotropin-releasing hormone protects against oxidative stress, promotes cytoprotection (6, 7), and exhibits antihyperglycemic activity (11).Some serine peptidases exhibit peptide bond formation (i.e., aminolysis of esters, thioesters, and amides) in accordance with their hydrolytic activity (2, 14). The exchange of catalytic Ser for Cys to engineer the serine endopeptidase into “transpeptidase” for peptide bond formation has been well characterized (3, 5). Our recent approach confirmed the wide distribution of family S9 aminopeptidases that have catalytic Ser in actinomycetes (12). Of them, we obtained S9 aminopeptidase from Streptomyces thermocyaneoviolaceus NBRC14271 (S9AP-St). The enzyme was engineered into “transaminopeptidase” by exchange of catalytic Ser for Cys, and its aminolytic activity was evaluated (13). The engineered enzyme, designated as aminolysin-S, can synthesize hydrophobic dipeptides through an aminolysis reaction. However, aminolysin-S was unable to synthesize peptides containing proline. Although the report of aminolysin-S demonstrated that S9AP-St shows no aminolysis reaction toward limited substrates, details of its characteristics remain unknown. This study verified the peptide synthetic activity of S9AP-St, demonstrating that S9AP-St can synthesize widely varied prolyl dipeptides through an aminolysis reaction. The report also shows that S9AP-St is applicable to the synthesis of a biologically active peptide—c(lP-lH).  相似文献   

12.
Halophilic archaea were found to contain in their cytoplasm millimolar concentrations of γ-glutamylcysteine (γGC) instead of glutathione. Previous analysis of the genome sequence of the archaeon Halobacterium sp. strain NRC-1 has indicated the presence of a sequence homologous to sequences known to encode the glutamate-cysteine ligase GshA. We report here the identification of the gshA gene in the extremely halophilic archaeon Haloferax volcanii and show that H. volcanii gshA directs in vivo the synthesis and accumulation of γGC. We also show that the H. volcanii gene when expressed in an Escherichia coli strain lacking functional GshA is able to restore synthesis of glutathione.Many organisms contain millimolar concentrations of low-molecular-weight thiol compounds that participate in a number of important biological functions involving thiol-disulfide exchanges (7). In particular, they serve to maintain an intracellular reducing environment, to provide reducing power for key reductive enzymes, to combat the effects of oxidative and disulfide stress, and to detoxify xenobiotic compounds (7). Glutathione (GSH), a cysteine-containing tripeptide, l-γ-glutamyl-l-cysteinylglycine, is the best-characterized low-molecular-weight thiol (7, 19, 21). GSH is made in a highly conserved two-step ATP-dependent process by two unrelated peptide bond-forming enzymes (3, 21). The γ-carboxyl group of l-glutamate and the amino group of l-cysteine are ligated by the enzyme glutamylcysteine (GC) ligase EC 6.3.2.2 (GshA, encoded by gshA), which is then condensed with glycine in a reaction catalyzed by GSH synthetase (GshB, encoded by gshB) to form GSH (10, 38). GSH is found primarily in gram-negative bacteria and eukaryotes and only rarely in gram-positive bacteria (26). Fahey and coworkers showed that GSH is absent from the high-GC gram-positive actinomycetes which produce, as the major low-molecular-weight thiol, mycothiol, 1-d-myo-inosityl-2-(N-acetyl-l-cysteinyl)-amido-2-deoxy-α-d-glucopyranoside (13, 26-28, 35). GSH is also absent in Archaea. In Pyrococcus furiosus, coenzyme A SH (CoASH) is the main thiol (11), whereas in Halobacterium salinarum, γGC is the predominant thiol and the organism possesses bis-γGC reductase activity (30, 36). Similarly, Leuconostoc kimchi and Leuconostoc mesenteroides, gram-positive lactic acid bacterial species, were recently found to contain γGC rather than GSH (15). To date, these are the sole procaryotic species reported to naturally produce γGC but not GSH (6, 30). In this report, we describe the identification of the gshA gene in the extremely halophilic archaeon Haloferax volcanii. Copley and Dhillon (6) previously identified, using bioinformatic tools, an open reading frame (ORF) (gene VNG1397C) in Halobacterium sp. strain NRC-1 with limited sequence relatedness to known GshA proteins (6). However, no genetic or biochemical evidence was presented to substantiate their conclusion. Here, we show that Haloferax volcanii strain DS2 (1, 25) contains an ORF that directs in vivo the synthesis and accumulation of γGC. We also show that the H. volcanii ORF, when expressed in Escherichia coli lacking functional GshA, is able to restore synthesis of GSH.  相似文献   

13.
14.
In comparison to other pseudomonads, Pseudomonas aeruginosa grows poorly in l-lysine as a sole source of nutrient. In this study, the ldcA gene (lysine decarboxylase A; PA1818), previously identified as a member of the ArgR regulon of l-arginine metabolism, was found essential for l-lysine catabolism in this organism. LdcA was purified to homogeneity from a recombinant strain of Escherichia coli, and the results of enzyme characterization revealed that this pyridoxal-5-phosphate-dependent decarboxylase takes l-lysine, but not l-arginine, as a substrate. At an optimal pH of 8.5, cooperative substrate activation by l-lysine was depicted from kinetics studies, with calculated Km and Vmax values of 0.73 mM and 2.2 μmole/mg/min, respectively. Contrarily, the ldcA promoter was induced by exogenous l-arginine but not by l-lysine in the wild-type strain PAO1, and the binding of ArgR to this promoter region was demonstrated by electromobility shift assays. This peculiar arginine control on lysine utilization was also noted from uptake experiments in which incorporation of radioactively labeled l-lysine was enhanced in cells grown in the presence of l-arginine but not l-lysine. Rapid growth on l-lysine was detected in a mutant devoid of the main arginine catabolic pathway and with a higher basal level of the intracellular l-arginine pool and hence elevated ArgR-responsive regulons, including ldcA. Growth on l-lysine as a nitrogen source can also be enhanced when the aruH gene encoding an arginine/lysine:pyruvate transaminase was expressed constitutively from plasmids; however, no growth of the ldcA mutant on l-lysine suggests a minor role of this transaminase in l-lysine catabolism. In summary, this study reveals a tight connection of lysine catabolism to the arginine regulatory network, and the lack of lysine-responsive control on lysine uptake and decarboxylation provides an explanation of l-lysine as a poor nutrient for P. aeruginosa.Decarboxylation of amino acids, including lysine, arginine, and glutamate, is important for bacterial survival under low pH (2, 7, 19). Lysine is abundant in the rhizosphere where fluorescent Pseudomonas preferentially resides, and serves as a nitrogen and carbon source to these organisms (28). In microbes, lysine catabolism can be initiated either through monooxygenase, decarboxylase, or transaminase activities. The monooxygenase pathway has been considered the major route for l-lysine utilization in Pseudomonas putida, and davBATD encoding enzymes for the first four steps of the pathway have been characterized (25, 26). In contrast, Pseudomonas aeruginosa cannot use exogenous l-lysine efficiently for growth (5, 24). It has been reported that enzymatic activities for the first two steps of the monooxygenase pathway are not detectable in P. aeruginosa, and no davBA orthologs can be identified from this organism (24, 25).Mutants of P. aeruginosa with improved growth on l-lysine and a high level of lysine decarboxylase activity can be isolated by repeated subcultures in l-lysine (5). This suggests that in P. aeruginosa, l-lysine utilization might be mediated by the lysine decarboxylase pathway with cadaverine and 5-aminovalerate as intermediates (Fig. (Fig.1).1). Alternatively, conversion of l-lysine into 5-aminovalerate may also be accomplished by a coupled reaction catalyzed by AruH and AruI. The AruH and AruI enzymes were reported as arginine:pyruvate transaminase and 2-ketoarginine decarboxylase, respectively (36). Interestingly, transamination by AruH using l-lysine as an amino group donor can also be detected in vitro (35). The reaction product α-keto-ɛ-aminohexanonate can potentially be decarboxylated into 5-aminovalerate by AruI, providing an alternative route for lysine degradation.Open in a separate windowFIG. 1.Lysine catabolic pathways. l-lysine decarboxylase pathway is shown at center. Broken arrows represent lysine monooxygenase pathway from P. putida which is not present in P. aeruginosa.In this study, we showed that the lysine decarboxylase pathway is the main route for lysine utilization under arginine control. Expression of the ldcAB operon encoding l-lysine decarboxylase and a putative lysine/cadaverine antiporter was analyzed regarding its response to l-lysine, l-arginine, and the arginine-responsive regulator ArgR. Enzyme characterization was performed to verify the function of LdcA as l-lysine decarboxylase. Arginine control on lysine incorporation was also investigated by genetic studies and uptake experiments. The peculiar role of ArgR controlling arginine and lysine uptake and catabolism provides the explanation for poor growth in lysine, and it implies a higher level of complexity in metabolic networks of pseudomonads.  相似文献   

15.
The pathway of lysine biosynthesis in the methanococci has not been identified previously. A variant of the diaminopimelic acid (DAP) pathway uses diaminopimelate aminotransferase (DapL) to catalyze the direct conversion of tetrahydrodipicolinate (THDPA) to ll-DAP. Recently, the enzyme DapL (MTH52) was identified in Methanothermobacter thermautotrophicus and shown to belong to the DapL1 group. Although the Methanococcus maripaludis genome lacks a gene that can be unambiguously assigned a DapL function based on sequence similarity, the open reading frame MMP1527 product shares 30% amino acid sequence identity with MTH52. A Δmmp1527 deletion mutant was constructed and found to be a lysine auxotroph, suggesting that this DapL homolog in methanococci is required for lysine biosynthesis. In cell extracts of the M. maripaludis wild-type strain, the specific activity of DapL using ll-DAP and α-ketoglutarate as substrates was 24.3 ± 2.0 nmol min−1 mg of protein−1. The gene encoding the DapL homolog in Methanocaldococcus jannaschii (MJ1391) was cloned and expressed in Escherichia coli, and the protein was purified. The maximum activity of MJ1391 was observed at 70°C and pH 8.0 to 9.0. The apparent Kms of MJ1391 for ll-DAP and α-ketoglutarate were 82.8 ± 10 μM and 0.42 ± 0.02 mM, respectively. MJ1391 was not able to use succinyl-DAP or acetyl-DAP as a substrate. Phylogenetic analyses suggested that two lateral gene transfers occurred in the DapL genes, one from the archaea to the bacteria in the DapL2 group and one from the bacteria to the archaea in the DapL1 group. These results demonstrated that the DapL pathway is present in marine methanogens belonging to the Methanococcales.Two lysine biosynthesis pathways evolved separately in organisms, the diaminopimelic acid (DAP) and aminoadipic acid (AAA) pathways. The DAP pathway synthesizes l-lysine from aspartate and pyruvate, and diaminopimelic acid is an intermediate. This pathway is utilized by most bacteria, some archaea, some fungi, some algae, and plants (28, 29). The AAA pathway synthesizes l-lysine from α-ketoglutarate and acetyl coenzyme A (acetyl-CoA), and α-aminoadipic acid is an intermediate. This pathway is utilized by most fungi, some algae, the bacterium Thermus thermophilus, and probably some archaea, such as Sulfolobus, Thermoproteus, and Pyrococcus (27, 36). No organism is known to possess both pathways.There are four known variations of the DAP pathway in bacteria: the succinylase, acetylase, aminotransferase, and dehydrogenase pathways (Fig. (Fig.1).1). These pathways share the steps converting l-aspartate to l-2,3,4,5-tetrahydrodipicolinate (THDPA), but the subsequent steps leading to the production of meso-diaminopimelate, the immediate precursor of l-lysine, are different. The succinylase pathway acylates THDPA with succinyl-CoA to generate N-succinyl-ll-2-amino-6-ketopimelate and forms meso-DAP by subsequent transamination, desuccinylation, and epimerization. This pathway is utilized by proteobacteria and many firmicutes and actinobacteria (12, 14, 20, 29). The acetylase pathway is analogous to the succinylase pathway but uses N-acetyl intermediates. This pathway is limited to certain Bacillus species, in which the corresponding genes have not been identified (33, 39). The aminotransferase pathway converts THDPA directly to ll-DAP by diaminopimelate aminotransferase (DapL) without acylation. This pathway is shared by cyanobacteria (19), chlamydia (24), the archaeon Methanothermobacter thermautotrophicus (15, 18), and the plant Arabidopsis thaliana (19). The dehydrogenase pathway forms meso-DAP directly from THDPA, NADPH, and NH4+ by using diaminopimelate dehydrogenase (Ddh). This pathway is utilized by some Bacillus and Brevibacterium species and Corynebacterium glutamicum (25, 26, 40). Most bacteria use only one of the four variants, although certain bacteria, such as C. glutamicum and Bacillus macerans, possess both the succinylase and dehydrogenase pathways (3, 30).Open in a separate windowFIG. 1.Variations in the DAP pathway for lysine biosynthesis. 1, succinylase pathway; 2, acetylase pathway; 3, aminotransferase pathway; 4, dehydrogenase pathway. Abbreviations and designations: THDPA, l-2,3,4,5-tetrahydrodipicolinate; l,l-DAP, ll-2,6-diaminopimelate; meso-DAP, meso-2,6-diaminopimelate; LysC, aspartate kinase; Asd, aspartate semialdehyde dehydrogenase; DapA, dihydrodipicolinate synthase; DapB, dihydrodipicolinate reductase; DapD, THDPA succinylase; DapC, succinyl-DAP aminotransferase; DapE, succinyl-DAP desuccinylase; DapF, DAP epimerase; LysA, DAP decarboxylase; DapL, ll-DAP aminotransferase; Ddh, DAP dehydrogenase.The diaminopimelate aminotransferase (DapL) catalyzes the transfer of an amino group from l-glutamate to THDPA, forming ll-DAP (19, 24). It uses pyridoxal 5′-phosphate (PLP) as a coenzyme and has constrained substrate specificity. DapL is not closely related to the DapC/ArgD aminotransferase, which functions in the succinylase pathway. Comparative genomic analysis identified dapL homologs in both bacterial and archaeal genomes. Homologs of dapD and dapE have not been found in genomes with dapL homologs, suggesting that transamination of THDPA does not require succinylation in these organisms (18). Phylogenetic analysis also suggested classification of DapL into two groups, DapL1 and DapL2, which share ∼30% amino acid sequence identity (18). These two groups both exhibit DapL activity, and they cannot be differentiated by kinetic properties (18, 37). The distribution of the two groups is not obviously associated with specific prokaryotic lineages (18).Methanogens are strictly anaerobic archaea that obtain all or most of their energy for growth from the production of large quantities of methane. All methanogens belong to the Euryarchaeota and are currently classified in six orders: Methanobacteriales, Methanococcales, Methanomicrobiales, Methanosarcinales, Methanopyrales, and Methanocellales (23, 41, 42). Biochemical studies of Methanocaldococcus jannaschii and Methanococcus voltae belonging to Methanococcales, Methanospirillum hungatei belonging to Methanomicrobiales, and Methanothermobacter thermautotrophicus belonging to Methanobacteriales suggested that these organisms derive their l-lysine from a DAP pathway, but the studies did not discriminate among the four DAP pathway variations (2, 9, 10, 32). Genome sequence analysis also suggested a DAP pathway in Methanosarcina mazei belonging to Methanosarcinales (8). Recent studies identified a dapL homolog belonging to the DapL1 group in M. thermautotrophicus. The gene product complemented an Escherichia coli dapD dapE double mutant and catalyzed the transamination of DAP to THDPA, suggesting that Methanobacteriales use the DapL pathway for l-lysine biosynthesis (15, 18). Homologs of asd, dapA, dapB, dapF, and lysA have been identified in the genomes of M. maripaludis and M. jannaschii belonging to the Methanococcales, but homologs responsible for the conversion of THDPA to ll-DAP have not been annotated (4, 17). Here we identified methanococcal DapL homologs and demonstrated that the DapL pathway is present in Methanococcales.  相似文献   

16.
17.
Glycopeptidolipids (GPLs) are one of the major glycolipid components present on the surface of Mycobacterium avium complex (MAC) that belong to opportunistic pathogens distributed in the natural environment. The serovars of MAC, up to around 30 types, are defined by the variable oligosaccharide portions of the GPLs. Epidemiological studies show that serovar 4 is the most prevalent type, and the prognosis of pulmonary disease caused by serovar 4 is significantly worse than that caused by other serovars. However, little is known about the biosynthesis of serovar 4-specific GPL, particularly the formation of the oligosaccharide portion that determines the properties of serovar 4. To investigate the biosynthesis of serovar 4-specific GPL, we focused on one segment that included functionally unknown genes in the GPL biosynthetic gene cluster of a serovar 4 strain. In this segment, a putative hemolytic protein gene, hlpA, and its downstream gene were found to be responsible for the formation of the 4-O-methyl-rhamnose residue, which is unique to serovar 4-specific GPL. Moreover, functional characterization of the hlpA gene revealed that it encodes a rhamnosyltransferase that transfers a rhamnose residue via 1→4 linkage to a fucose residue of serovar 2-specific GPL, which is a key pathway leading to the synthesis of oligosaccharide of serovar 4-specific GPL. These findings may provide clues to understanding the biological role of serovar 4-specific GPL in MAC pathogenicity and may also provide new insights into glycosyltransferase, which generates structural and functional diversity of GPLs.The genus Mycobacterium has a unique feature in the cell envelope that contains a multilayered structure consisting of peptidoglycan, mycolyl-arabinogalactan complex, and surface glycolipids (8, 12). It is known that these components play a role in protection from environmental stresses, such as antimicrobial agents and host immune responses (8, 12). Some of them are recognized as pathogenic factors related to mycobacterial diseases, such as tuberculosis and leprosy (8, 12). In case of nontuberculous mycobacteria that are widely distributed in the natural environment as opportunistic pathogens, glycopeptidolipids (GPLs) are abundantly present on the cell envelope as surface glycolipids (34). GPLs have a core structure in which a fatty acyl-tetrapeptide is glycosylated with 6-deoxy-talose (6-d-Tal) and O-methyl-rhamnose (O-Me-Rha) (2, 5, 13). This structure is common to all types of GPLs, and GPLs with this structure that have not undergone further glycosylation are termed non-serovar-specific GPLs (nsGPLs) (2, 5, 13). Structural diversity generated by further glycosylations, such as rhamnosylation, fucosylation, and glucosylation, is observed for the oligosaccharide portion linked to the 6-d-Tal residue of nsGPLs from Mycobacterium avium complex (MAC), a member of the nontuberculous mycobacteria consisting of two species, M. avium and M. intracellulare (2, 5, 34). Consequently, these nsGPLs with varied oligosaccharides lead to the formation of the serovar-specific GPLs (ssGPLs) that define around 30 types of MAC serovars (10).The properties of MAC serovars are known to be notably different from each other and also to be closely associated with the pathogenicity of MAC (3, 6, 18, 30, 31, 32). Various epidemiological studies indicate that serovar 4 is the most prevalent type and is also one of the serovars frequently isolated from AIDS patients (1, 20, 33, 36). Additionally, pulmonary MAC disease caused by serovar 4 is shown to exhibit a poorer prognosis than that caused by other serovars (23). With respect to host immune responses to MAC infection, serovar 4-specific GPL is reported to have characteristic features that are in contrast to those of other ssGPLs (21, 30). Structurally, serovar 4-specific GPL contains a unique oligosaccharide in which the oligosaccharide of serovar 2-specific GPL is further glycosylated with 4-O-methyl-rhamnose (4-O-Me-Rha) residue through a 1→4 linkage (Table (Table1)1) (24). Therefore, it is thought that the presence of 4-O-Me-Rha and its linkage position are important in exhibiting the specificity of biological activities. The biosynthesis of the oligosaccharide portion in several ssGPLs is currently being clarified (15, 16, 17, 25, 26), while that of serovar 4-specific GPL is still unresolved. In this study, we have focused on the genomic region predicted to be associated with GPL biosynthesis in the serovar 4 strain and explored the key genes responsible for the formation of 4-O-Me-Rha that might determine the specific properties of MAC serovar 4.

TABLE 1.

Oligosaccharide structures of serovar 2- and 4-specific GPLs
SerovarOligosaccharideReference
22,3-di-O-Me-α-l-Fuc-(1→3)-α-l-Rha-(1→2)-l-6-d-Tal9
44-O-Me-α-l-Rha-(1→4)-2-O-Me-α-l-Fuc-(1→3)-α-l-Rha-(1→2)-l-6-d-Tal24
Open in a separate window  相似文献   

18.
Spores of Bacillus species are said to be committed when they continue through nutrient germination even when germinants are removed or their binding to spores'' nutrient germinant receptors (GRs) is both reversed and inhibited. Measurement of commitment and the subsequent release of dipicolinic acid (DPA) during nutrient germination of spores of Bacillus cereus and Bacillus subtilis showed that heat activation, increased nutrient germinant concentrations, and higher average levels of GRs/spore significantly decreased the times needed for commitment, as well as lag times between commitment and DPA release. These lag times were also decreased dramatically by the action of one of the spores'' two redundant cortex lytic enzymes (CLEs), CwlJ, but not by the other CLE, SleB, and CwlJ action did not affect the timing of commitment. The timing of commitment and the lag time between commitment and DPA release were also dependent on the specific GR activated to cause spore germination. For spore populations, the lag times between commitment and DPA release were increased significantly in spores that germinated late compared to those that germinated early, and individual spores that germinated late may have had lower appropriate GR levels/spore than spores that germinated early. These findings together provide new insight into the commitment step in spore germination and suggest several factors that may contribute to the large heterogeneity among the timings of various events in the germination of individual spores in spore populations.Spores of Bacillus species can remain dormant for long times and are extremely resistant to a variety of environmental stresses (26). However, under appropriate conditions, normally upon the binding of specific nutrients to spores'' nutrient germinant receptors (GRs), spores can come back to active growth through a process called germination followed by outgrowth (19, 20, 25, 26). Germination of Bacillus subtilis spores can be triggered by l-alanine or l-valine or a combination of l-asparagine, d-glucose, d-fructose, and K+ (AGFK). These nutrient germinants trigger germination by binding to and interacting with GRs that have been localized to the spore''s inner membrane (12, 20). l-Alanine and l-valine bind to the GerA GR, while the AGFK mixture triggers germination by interacting with both the GerB and GerK GRs (25). Normally, l-asparagine alone does not trigger B. subtilis spore germination. However, a mutant form of the GerB GR, termed GerB*, displays altered germinant specificity such that l-asparagine alone will trigger the germination of gerB* mutant spores (1, 18).A number of events occur in a defined sequence during spore germination. Initially, exposure of spores to nutrient germinants causes a reaction that commits spores to germinate, even if the germinant is removed or displaced from its cognate GR (7, 10, 21, 27, 28). This commitment step is followed by release of monovalent cations, as well as the spore core''s large pool of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) along with divalent cations, predominantly Ca2+, that are chelated with DPA (Ca-DPA). In Bacillus spores, the release of Ca-DPA triggers the hydrolysis of spores'' peptidoglycan cortex by either of two cortex lytic enzymes (CLEs), CwlJ and SleB (11, 16, 23). CwlJ is activated during germination by Ca-DPA as it is being released from individual spores, while SleB activation requires that most Ca-DPA be released (14, 16, 17). Cortex hydrolysis, in turn, allows the spore core to expand and fully hydrate, which leads to activation of enzymes and initiation of metabolism in the spore core (21, 25).As noted above, commitment is the first event that can be assessed during spore germination, although the precise mechanism of commitment is not known. Since much has been learned about proteins important in spore germination in the many years since commitment was last studied (25, 26), it seemed worth reexamining commitment, with the goal of determining those factors that influence this step in the germination process. Knowledge of factors important in determining kinetics of commitment could then lead to an understanding of what is involved in this reaction.Kinetic analysis of spore germination, as well as commitment, has mostly been based on the decrease in optical density at 600 nm (OD600) of spore suspensions, which monitors a combination of events that occur well after commitment, including DPA release, cortex hydrolysis, and core swelling (25-27). In the current work, we have used a germination assay that measures DPA release, an early event in spore germination, and have automated this assay to allow routine measurement of commitment, as well as DPA release from large numbers of spore samples simultaneously. This assay has allowed comparison of the kinetics of DPA release and commitment during germination and study of the effects of heat activation, germinant concentration, GR levels, and CLEs on commitment.  相似文献   

19.
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   

20.
A major question about cytokinesis concerns the role of the septin proteins, which localize to the division site in all animal and fungal cells but are essential for cytokinesis only in some cell types. For example, in Schizosaccharomyces pombe, four septins localize to the division site, but deletion of the four genes produces only a modest delay in cell separation. To ask if the S. pombe septins function redundantly in cytokinesis, we conducted a synthetic-lethal screen in a septin-deficient strain and identified seven mutations. One mutation affects Cdc4, a myosin light chain that is an essential component of the cytokinetic actomyosin ring. Five others cause frequent cell lysis during cell separation and map to two loci. These mutations and their dosage suppressors define a signaling pathway (including Rho1 and a novel arrestin) for repairing cell-wall damage. The seventh mutation affects the poorly understood RNA-binding protein Scw1 and severely delays cell separation when combined either with a septin mutation or with a mutation affecting the septin-interacting, anillin-like protein Mid2, suggesting that Scw1 functions in a pathway parallel to that of the septins. Taken together, our results suggest that the S. pombe septins participate redundantly in one or more pathways that cooperate with the actomyosin ring during cytokinesis and that a septin defect causes septum defects that can be repaired effectively only when the cell-integrity pathway is intact.THE fission yeast Schizosaccharomyces pombe provides an outstanding model system for studies of cytokinesis (McCollum and Gould 2001; Balasubramanian et al. 2004; Pollard and Wu 2010). As in most animal cells, successful cytokinesis in S. pombe requires an actomyosin ring (AMR). The AMR begins to assemble at the G2/M transition and involves the type II myosin heavy chains Myo2 and Myp2 and the light chains Cdc4 and Rlc1 (Wu et al. 2003). Myo2 and Cdc4 are essential for cytokinesis under all known conditions, Rlc1 is important at all temperatures but essential only at low temperatures, and Myp2 is essential only under stress conditions. As the AMR constricts, a septum of cell wall is formed between the daughter cells. The primary septum is sandwiched by secondary septa and subsequently digested to allow cell separation (Humbel et al. 2001; Sipiczki 2007). Because of the internal turgor pressure of the cells, the proper assembly and structural integrity of the septal layers are essential for cell survival.Septum formation involves the β-glucan synthases Bgs1/Cps1/Drc1, Bgs3, and Bgs4 (Ishiguro et al. 1997; Le Goff et al. 1999; Liu et al. 1999, 2002; Martín et al. 2003; Cortés et al. 2005) and the α-glucan synthase Ags1/Mok1 (Hochstenbach et al. 1998; Katayama et al. 1999). These synthases are regulated by the Rho GTPases Rho1 and Rho2 and the protein kinase C isoforms Pck1 and Pck2 (Arellano et al. 1996, 1997, 1999; Nakano et al. 1997; Hirata et al. 1998; Calonge et al. 2000; Sayers et al. 2000; Ma et al. 2006; Barba et al. 2008; García et al. 2009b). The Rho GTPases themselves appear to be regulated by both GTPase-activating proteins (GAPs) and guanine-nucleotide-exchange factors (GEFs) (Nakano et al. 2001; Calonge et al. 2003; Iwaki et al. 2003; Tajadura et al. 2004; Morrell-Falvey et al. 2005; Mutoh et al. 2005; García et al. 2006, 2009a,b). In addition, septum formation and AMR function appear to be interdependent. In the absence of a normal AMR, cells form aberrant septa and/or deposit septal materials at random locations, whereas a mutant defective in septum formation (bgs1) is also defective in AMR constriction (Gould and Simanis 1997; Le Goff et al. 1999; Liu et al. 1999, 2000). Both AMR constriction and septum formation also depend on the septation initiation network involving the small GTPase Spg1 (McCollum and Gould 2001; Krapp and Simanis 2008). Despite this considerable progress, many questions remain about the mechanisms and regulation of septum formation and its relationships to the function of the AMR.One major question concerns the role(s) of the septins. Proteins of this family are ubiquitous in fungal and animal cells and typically localize to the cell cortex, where they appear to serve as scaffolds and diffusion barriers for other proteins that participate in a wide variety of cellular processes (Longtine et al. 1996; Gladfelter et al. 2001; Hall et al. 2008; Caudron and Barral 2009). Despite the recent progress in elucidating the mechanisms of septin assembly (John et al. 2007; Sirajuddin et al. 2007; Bertin et al. 2008; McMurray and Thorner 2008), the details of septin function remain obscure. However, one prominent role of the septins and associated proteins is in cytokinesis. Septins concentrate at the division site in every cell type that has been examined, and in Saccharomyces cerevisiae (Hartwell 1971; Longtine et al. 1996; Lippincott et al. 2001; Dobbelaere and Barral 2004) and at least some Drosophila (Neufeld and Rubin 1994; Adam et al. 2000) and mammalian (Kinoshita et al. 1997; Surka et al. 2002) cell types, the septins are essential for cytokinesis. In S. cerevisiae, the septins are required for formation of the AMR (Bi et al. 1998; Lippincott and Li 1998). However, this cannot be their only role, because the AMR itself is not essential for cytokinesis in this organism (Bi et al. 1998; Korinek et al. 2000; Schmidt et al. 2002). Moreover, there is no evidence that the septins are necessary for AMR formation or function in any other organism. A further complication is that in some cell types, including most Caenorhabditis elegans cells (Nguyen et al. 2000; Maddox et al. 2007) and some Drosophila cells (Adam et al. 2000; Field et al. 2008), the septins do not appear to be essential for cytokinesis even though they localize to the division site.S. pombe has seven septins, four of which (Spn1, Spn2, Spn3, and Spn4) are expressed in vegetative cells and localize to the division site shortly before AMR constriction and septum formation (Longtine et al. 1996; Berlin et al. 2003; Tasto et al. 2003; Wu et al. 2003; An et al. 2004; Petit et al. 2005; Pan et al. 2007; Onishi et al. 2010). Spn1 and Spn4 appear to be the core members of the septin complex (An et al. 2004; McMurray and Thorner 2008), and mutants lacking either of these proteins do not assemble the others at the division site. Assembly of a normal septin ring also depends on the anillin-like protein Mid2, which colocalizes with the septins (Berlin et al. 2003; Tasto et al. 2003). Surprisingly, mutants lacking the septins are viable and form seemingly complete septa with approximately normal timing. These mutants do, however, display a variable delay in separation of the daughter cells, suggesting that the septins play some role(s) in the proper completion of the septum or in subsequent processes necessary for cell separation (Longtine et al. 1996; An et al. 2004; Martín-Cuadrado et al. 2005).It is possible that the septins localize to the division site and yet are nonessential for division in some cell types because their role is redundant with that of some other protein(s) or pathway(s). To explore this possibility in S. pombe, we screened for mutations that were lethal in combination with a lack of septins. The results suggest that the septins cooperate with the AMR during cytokinesis and that, in the absence of septin function, the septum is not formed properly, so that an intact system for recognizing and repairing cell-wall damage becomes critical for cell survival.  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