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 375 毫秒
1.
2.
The mechanism of hexose transport into plasma membrane vesicles isolated from mature sugarbeet leaves (Beta vulgaris L.) was investigated. The initial rate of glucose uptake into the vesicles was stimulated approximately fivefold by imposing a transmembrane pH gradient (ΔpH), alkaline inside, and approximately fourfold by a negative membrane potential (ΔΨ), generated as a K+-diffusion potential, negative inside. The -fold stimulation was directly related to the relative ΔpH or ΔΨ gradient imposed, which were determined by the uptake of acetate or tetraphenylphosphonium, respectively. ΔΨ- and ΔpH-dependent glucose uptake showed saturation kinetics with a Km of 286 micromolar for glucose. Other hexose molecules (e.g. 2-deoxy-d-glucose, 3-O-methyl-d-glucose, and d-mannose) were also accumulated into plasma membrane vesicles in a ΔpH-dependent manner. Inhibition constants of a number of compounds for glucose uptake were determined. Effective inhibitors of glucose uptake included: 3-O-methyl-d-glucose, 5-thio-d-glucose, d-fructose, d-galactose, and d-mannose, but not 1-O-methyl-d-glucose, d- and l-xylose, l-glucose, d-ribose, and l-sorbose. Under all conditions of proton motive force magnitude and glucose and sucrose concentration tested, there was no effect of sucrose on glucose uptake. Thus, hexose transport on the sugarbeet leaf plasma membrane was by a H+-hexose symporter, and the carrier and possibly the energy source were not shared by the plasma membrane H+-sucrose symporter.  相似文献   

3.
Organisms that overproduced l-cysteine and l-cystine from glucose were constructed by using Escherichia coli K-12 strains. cysE genes coding for altered serine acetyltransferase, which was genetically desensitized to feedback inhibition by l-cysteine, were constructed by replacing the methionine residue at position 256 of the serine acetyltransferase protein with 19 other amino acid residues or the termination codon to truncate the carboxy terminus from amino acid residues 256 to 273 through site-directed mutagenesis by using PCR. A cysteine auxotroph, strain JM39, was transformed with plasmids having these altered cysE genes. The serine acetyltransferase activities of most of the transformants, which were selected based on restored cysteine requirements and ampicillin resistance, were less sensitive than the serine acetyltransferase activity of the wild type to feedback inhibition by l-cysteine. At the same time, these transformants produced approximately 200 mg of l-cysteine plus l-cystine per liter, whereas these amino acids were not detected in the recombinant strain carrying the wild-type serine acetyltransferase gene. However, the production of l-cysteine and l-cystine by the transformants was very unstable, presumably due to a cysteine-degrading enzyme of the host, such as cysteine desulfhydrase. Therefore, mutants that did not utilize cysteine were derived from host strain JM39 by mutagenesis with N-methyl-N′-nitro-N-nitrosoguanidine. When a newly derived host was transformed with plasmids having the altered cysE genes, we found that the production of l-cysteine plus l-cystine was markedly increased compared to production in JM39.l-Cysteine, one of the important amino acids used in the pharmaceutical, food, and cosmetics industries, has been obtained by extracting it from acid hydrolysates of the keratinous proteins in human hair and feathers. The first successful microbial process used for industrial production of l-cysteine involved the asymmetric conversion of dl-2-aminothiazoline-4-carboxylic acid, an intermediate compound in the chemical synthesis of dl-cysteine, to l-cysteine by enzymes from a newly isolated bacterium, Pseudomonas thiazoliniphilum (11). Yamada and Kumagai (13) also described enzymatic synthesis of l-cysteine from beta-chloroalanine and sodium sulfide in which Enterobacter cloacae cysteine desulfhydrase (CD) was used. However, high level production of l-cysteine from glucose with microorganisms has not been studied.Biosynthesis of l-cysteine in wild-type strains of Escherichia coli and Salmonella typhimurium is regulated through feedback inhibition by l-cysteine of serine acetyltransferase (SAT), a key enzyme in l-cysteine biosynthesis, and repression of expression of a series of enzymes used for sulfide reduction from sulfate by l-cysteine (4), as shown in Fig. Fig.1.1. Denk and Böck reported that a small amount of l-cysteine was excreted by a revertant of a cysteine auxotroph of E. coli. In this revertant, SAT encoded by the cysE gene was desensitized to feedback inhibition by l-cysteine, and the methionine residue at position 256 in SAT was replaced by isoleucine (2). These results indicate that it may be possible to construct organisms that produce high levels of l-cysteine by amplifying an altered cysE gene. Although the residue at position 256 is supposedly part of the allosteric site for cysteine binding, no attention has been given to the effect of an amino acid substitution at position 256 in SAT on feedback inhibition by l-cysteine and production of l-cysteine. It is also not known whether isoleucine is the best residue for desensitization to feedback inhibition. Open in a separate windowFIG. 1Biosynthesis and regulation of l-cysteine in E. coli. Abbreviations: APS, adenosine 5′-phosphosulfate; PAPS, phosphoadenosine 5′-phosphosulfate; Acetyl CoA, acetyl coenzyme A. The open arrow indicates feedback inhibition, and the dotted arrows indicate repression.On the other hand, l-cysteine appears to be degraded by E. coli cells. Therefore, in order to obtain l-cysteine producers, a host strain with a lower level of l-cysteine degradation activity must be isolated. In this paper we describe high-level production of l-cysteine plus l-cystine from glucose by E. coli resulting from construction of altered cysE genes. The methionine residue at position 256 in SAT was replaced by other amino acids or the termination codon in order to truncate the carboxy terminus from amino acid residues 256 to 273 by site-directed mutagenesis. A newly derived cysteine-nondegrading E. coli strain with plasmids having the altered cysE genes was used to investigate production of l-cysteine plus l-cystine.  相似文献   

4.
Cystine lyase degrades l-cystine by a β-elimination to form cysteine persulfide, pyruvate, and ammonia. This enzyme is common in Brassica sp. and has been purified to homogeneity from extracts of broccoli (Brassica oleracea var botrytis) buds. Two isozymes were separated on DEAE-Fractogel columns and the first peak, cystine lyase I further purified to homogeneity. The purified enzyme had a narrow range of substrate specificity with l-cystine and S-alkyl-l-cysteine sulfoxides being the primary substrates. The Km for l-cystine was 1.9 millimolar and for S-ethyl-l-cysteine sulfoxide was 15.6 millimolar, suggesting that l-cystine would be preferred in vivo. Using gel filtration and sodium dodecyl sulfate-polyacrylamide gel electrophoresis the molecular weight of the holoenzyme was estimated as 152,000 composed of subunits of approximately 49,000. This strongly suggests the native enzyme is a trimer. The presence of carbohydrate in the native enzyme was detected at the level of 5.8% on a weight basis. Except for the ability to utilize l-cystine as a substrate there are many similarities between cystine lyase I and the alliin lyase of onion (Allium cepa).  相似文献   

5.
The control of sulphate reduction in bacteria   总被引:2,自引:2,他引:0  
1. An enzyme from Escherichia coli 9723 that reduces adenosine 3′-phosphate 5′-sulphatophosphate to inorganic sulphite is described. Extracts of E. coli K12 and Bacillus subtilis 1379 contain a similar enzyme. 2. This reductase and sulphite reductase (EC 1.8.1.2) of E. coli 9723, E. coli K12 and of B. subtilis are repressed by growth in the presence of l-cystine. Cysteine synthase (EC 4.2.1.22) is unaffected. 3. Growth of E. coli 9723 on inorganic sulphite represses the sulphate-activating enzymes (EC 2.7.7.4 and 2.7.1.25) almost completely but has little effect on sulphite reductase. Growth on 0·042–0·056mm-l-cystine gives a similar result. 4. Such differential repression by cyst(e)ine prevents E. coli, when growing on sulphite, from synthesizing unnecessary enzymes.  相似文献   

6.
Escherichia coli K-12 provided with glucose and a mixture of amino acids depletes l-serine more quickly than any other amino acid even in the presence of ammonium sulfate. A mutant without three 4Fe4S l-serine deaminases (SdaA, SdaB, and TdcG) of E. coli K-12 is unable to do this. The high level of l-serine that accumulates when such a mutant is exposed to amino acid mixtures starves the cells for C1 units and interferes with cell wall synthesis. We suggest that at high concentrations, l-serine decreases synthesis of UDP-N-acetylmuramate-l-alanine by the murC-encoded ligase, weakening the cell wall and producing misshapen cells and lysis. The inhibition by high l-serine is overcome in several ways: by a large concentration of l-alanine, by overproducing MurC together with a low concentration of l-alanine, and by overproducing FtsW, thus promoting septal assembly and also by overexpression of the glycine cleavage operon. S-Adenosylmethionine reduces lysis and allows an extensive increase in biomass without improving cell division. This suggests that E. coli has a metabolic trigger for cell division. Without that reaction, if no other inhibition occurs, other metabolic functions can continue and cells can elongate and replicate their DNA, reaching at least 180 times their usual length, but cannot divide.The Escherichia coli genome contains three genes, sdaA, sdaB, and tdcG, specifying three very similar 4Fe4S l-serine deaminases. These enzymes are very specific for l-serine for which they have unusually high Km values (3, 32). Expression of the three genes is regulated so that at least one of the gene products is synthesized under all common growth conditions (25). This suggests an important physiological role for the enzymes. However, why E. coli needs to deaminate l-serine has been a long-standing problem of E. coli physiology, the more so since it cannot use l-serine as the sole carbon source.We showed recently that an E. coli strain devoid of all three l-serine deaminases (l-SDs) loses control over its size, shape, and cell division when faced with complex amino acid mixtures containing l-serine (32). We attributed this to starvation for single-carbon (C1) units and/or S-adenosylmethionine (SAM). C1 units are usually made from serine via serine hydroxymethyl transferase (GlyA) or via glycine cleavage (GCV). The l-SD-deficient triple mutant strain is starved for C1 in the presence of amino acids, because externally provided glycine inhibits GlyA and a very high internal l-serine concentration along with several other amino acids inhibits glycine cleavage. While the parent cell can defend itself by reducing the l-serine level by deamination, this crucial reaction is missing in the ΔsdaA ΔsdaB ΔtdcG triple mutant. We therefore consider these to be “defensive” serine deaminases.The fact that an inability to deaminate l-serine leads to a high concentration of l-serine and inhibition of GlyA is not surprising. However, it is not obvious why a high level of l-serine inhibits cell division and causes swelling, lysis, and filamentation. Serine toxicity due to inhibition of biosynthesis of isoleucine (11) and aromatic amino acids (21) has been reported but is not relevant here, since these amino acids are provided in Casamino Acids.We show here that at high internal concentrations, l-serine also causes problems with peptidoglycan synthesis, thus weakening the cell wall. Peptidoglycan is a polymer of long glycan chains made up of alternating N-acetylglucosamine and N-acetylmuramic acid residues, cross-linked by l-alanyl-γ-d-glutamyl-meso-diaminopimelyl-d-alanine tetrapeptides (1, 28). The glucosamine and muramate residues and the pentapeptide (from which the tetrapeptide is derived) are all synthesized in the cytoplasm and then are exported to be polymerized into extracellular peptidoglycan (2).In this paper, we show that lysis is caused by l-serine interfering with the first step of synthesis of the cross-linking peptide, the addition of l-alanine to uridine diphosphate-N-acetylmuramate. This interference is probably due to a competition between serine and l-alanine for the ligase, MurC, which adds the first l-alanine to UDP-N-acetylmuramate (7, 10, 15). As described here, the weakening of the cell wall by l-serine can be overcome by a variety of methods that reduce the endogenous l-serine pool or counteract the effects of high levels of l-serine.  相似文献   

7.
d-Galactan I is a polysaccharide with the disaccharide repeat unit structure [→3-β-d-Galf-(1→3)-α-d-Galp-(1→]. This glycan represents the lipopolysaccharide O antigen found in many Gram-negative bacteria, including several Klebsiella pneumoniae O serotypes. The polysaccharide is synthesized in the cytoplasm prior to its export via an ATP-binding cassette transporter. Sequence analysis predicts three galactosyltransferases in the d-galactan I genetic locus. They are WbbO (belonging to glycosyltransferase (GT) family 4), WbbM (GT-family 8), and WbbN (GT-family 2). The WbbO and WbbM proteins are each predicted to contain two domains, with the GT modules located toward their C termini. The N-terminal domains of WbbO and WbbM exhibit no similarity to proteins with known function. In vivo complementation assays suggest that all three glycosyltransferases are required for d-galactan I biosynthesis. Using a bacterial two-hybrid system and confirmatory co-purification strategies, evidence is provided for protein-protein interactions among the glycosyltransferases, creating a membrane-located enzyme complex dedicated to d-galactan I biosynthesis.  相似文献   

8.
Transporters of the dicarboxylate amino acid-cation symporter family often mediate uptake of C4-dicarboxylates, such as succinate or l-malate, in bacteria. A member of this family, dicarboxylate transporter A (DctA) from Corynebacterium glutamicum, was characterized to catalyze uptake of the C4-dicarboxylates succinate, fumarate, and l-malate, which was inhibited by oxaloacetate, 2-oxoglutarate, and glyoxylate. DctA activity was not affected by sodium availability but was dependent on the electrochemical proton potential. Efficient growth of C. glutamicum in minimal medium with succinate, fumarate, or l-malate as the sole carbon source required high dctA expression levels due either to a promoter-up mutation identified in a spontaneous mutant or to ectopic overexpression. Mutant analysis indicated that DctA and DccT, a C4-dicarboxylate divalent anion/sodium symporter-type transporter, are the only transporters for succinate, fumarate, and l-malate in C. glutamicum.In bacteria, the uptake of dicarboxylates, such as the tricarboxylic acid (TCA) cycle intermediates succinate, fumarate, and l-malate, is mediated by transporters of different protein families. Whereas Dcu-type transporters facilitate dicarboxylate uptake under anaerobic conditions, the most common aerobic dicarboxylate transporters are members of the dicarboxylate amino acid-cation symporter (DAACS), divalent anion sodium symporter (DASS), tripartite ATP-independent periplasmic (TRAP), and CitMHS transporter families. DAACS transporters are responsible for C4-dicarboxylate uptake under aerobic conditions in various bacteria, e.g., DctA from Escherichia coli, Bacillus subtilis, or Rhizobium leguminosarum, and are involved in different physiological functions (2, 4, 27, 41). The first described member of the TRAP family is the C4-dicarboxylate transporter DctPQM from Rhodobacter capsulatus, which facilitates substrate uptake by the use of an extracytoplasmic solute receptor (8). An example of the DASS family, members of which occur in bacteria, as well in eukaryotes, is the well-characterized transporter SdcS from Staphylococcus aureus (13). Members of the CitHMS family import citrate in symport with the cation Mg2+ or Ca2+. Whereas E. coli possesses one DctA and four different Dcu carriers, no Dcu transporter-encoding genes were found in Corynebacterium glutamicum (16, 19), which is used for the industrial production of amino acids, such as glutamate (33) or l-lysine (39), and is capable of succinate and l-lactate production under oxygen deprivation conditions. A dctA gene was annotated (19); however, C. glutamicum is not able to utilize succinate, malate, or fumarate as a sole carbon source. The uptake systems CitH and TctCBA have been characterized recently as citrate uptake systems (3, 26). Interestingly, we and others have shown that C. glutamicum possesses a DASS family transporter (DccT) for uptake of the C4-dicarboxylates succinate, fumarate, and l-malate (36, 40). Spontaneous mutants showing fast growth in succinate or fumarate minimal medium were isolated and shown to possess promoter-up mutations in the dccT gene (40). In l-malate minimal medium, these spontaneous mutants showed relatively slow growth, and the affinity of DccT for succinate and fumarate was found to be 5- and 12-fold higher than for l-malate, respectively (40). These findings prompted us to search for other uptake systems for l-malate in C. glutamicum. Here, we describe the identification and characterization of the DAACS family protein DctA from C. glutamicum as a proton motive force-driven uptake system for C4-dicarboxylate intermediates of the TCA cycle. Additionally, we compare both uptake systems, DccT and DctA, from C. glutamicum.  相似文献   

9.
10.
11.
The serP1 and serP2 genes found adjacently on the chromosome of Lactococcus lactis strains encode two members of the amino acid-polyamine-organocation (APC) superfamily of secondary transporters that share 61% sequence identity. SerP1 transports l-serine, l-threonine, and l-cysteine with high affinity. Affinity constants (Km) are in the 20 to 40 μM range. SerP2 is a dl-alanine/dl-serine/glycine transporter. The preferred substrate appears to be dl-alanine for which the affinities were found to be 38 and 20 μM for the d and l isomers, respectively. The common substrate l-serine is a high-affinity substrate of SerP1 and a low-affinity substrate of SerP2 with affinity constants of 18 and 356 μM, respectively. Growth experiments demonstrate that SerP1 is the main l-serine transporter responsible for optimal growth in media containing free amino acids as the sole source of amino acids. SerP2 is able to replace SerP1 in this role only in medium lacking the high-affinity substrates l-alanine and glycine. SerP2 plays an adverse role for the cell by being solely responsible for the uptake of toxic d-serine. The main function of SerP2 is in cell wall biosynthesis through the uptake of d-alanine, an essential precursor in peptidoglycan synthesis. SerP2 has overlapping substrate specificity and shares 42% sequence identity with CycA of Escherichia coli, a transporter whose involvement in peptidoglycan synthesis is well established. No evidence was obtained for a role of SerP1 and SerP2 in the excretion of excess amino acids during growth of L. lactis on protein/peptide-rich media.  相似文献   

12.
Homyk, Rodriguez and Weil (1976) have described T4 mutants, called sip, that partially suppress the inability of T4rII mutants to grow in λ lysogens. We have found that mutants sip1 and sip2 are resistant to folate analogs and overproduce FH2 reductase. The results of recombination and complementation studies indicate that sip mutations are in the mot gene. Like other mot mutations (Mattson, Richardson and Goodin 1974; Chace and Hall 1975; Sauerbier, Hercules and Hall 1976), the sip2 mutation affects the expression of many genes and appears to affect promoter utilization. The mot gene function is not required for T4 growth on most hosts, but we have found that it is required for good growth on E. coli CTr5X. Homyk, Rodriguez and Weil (1976) also described L mutations that reverse the effects of sip mutations. L2 decreases the folate analog resistance and the inability of sip2 to grow on CTr5X. L2 itself is partially resistant to a folate analog, and appears to reverse the effects of sip2 on gene expression. These results suggest that L2 affects another regulatory gene related to the mot gene.  相似文献   

13.
Diketopiperazines (DKPs) are naturally-occurring cyclic dipeptides with a small structure and are found in many organisms and in large amounts in some foods and beverages. We found that a chicken essence beverage, which is popular among Southeast Asians as a traditional remedy and a rich source of DKPs, inhibited the serotonin transporter (SERT) and suppressed serotonin uptake from rat brain synaptosomes, which prompted us to isolate and identify the active substance(s). We purified a SERT inhibitor from the chicken essence beverage and identified it as the DKP cyclo(L-Phe-L-Phe). Interestingly, it was a naturally occurring dual inhibitor that inhibited both SERT and acetylcholinesterase (AChE) in vitro. The DKP increased extracellular levels of the cerebral monoamines serotonin, norepinephrine, and dopamine in the medial prefrontal cortex and acetylcholine in the ventral hippocampus of freely moving rats when administered orally. Moreover, cyclo(L-Phe-L-Phe) significantly shortened escape latency in the water maze test in depressed mice previously subjected to a repeated open-space swimming task, which induces a depression-like state. Cyclo(L-Phe-L-Phe) also significantly improved accuracy rates in a radial maze test in rats and increased step-through latencies in a passive avoidance test in mice with scopolamine-induced amnesia. These animal test results suggest that cyclo(L-Phe-L-Phe), which is present abundantly in some foods such as chicken essence, may abrogate the onset of depression and, thus, contribute to preventing the development of Alzheimer’s disease and other dementia, because senile depression is a risk factor for dementia.  相似文献   

14.
15.
Escherichia coli that is unable to metabolize d-glucose (with knockouts in ptsG, manZ, and glk) accumulates a small amount of d-glucose (yield of about 0.01 g/g) during growth on the pentoses d-xylose or l-arabinose as a sole carbon source. Additional knockouts in the zwf and pfkA genes, encoding, respectively, d-glucose-6-phosphate 1-dehydrogenase and 6-phosphofructokinase I (E. coli MEC143), increased accumulation to greater than 1 g/liter d-glucose and 100 mg/liter d-mannose from 5 g/liter d-xylose or l-arabinose. Knockouts of other genes associated with interconversions of d-glucose-phosphates demonstrate that d-glucose is formed primarily by the dephosphorylation of d-glucose-6-phosphate. Under controlled batch conditions with 20 g/liter d-xylose, MEC143 generated 4.4 g/liter d-glucose and 0.6 g/liter d-mannose. The results establish a direct link between pentoses and hexoses and provide a novel strategy to increase carbon backbone length from five to six carbons by directing flux through the pentose phosphate pathway.  相似文献   

16.
L-arginine (L-Arg) deficiency results in decreased T-cell proliferation and impaired T-cell function. Here we have found that L-Arg depletion inhibited expression of different membrane antigens, including CD247 (CD3ζ), and led to an ER stress response, as well as cell cycle arrest at G0/G1 in both human Jurkat and peripheral blood mitogen-activated T cells, without undergoing apoptosis. By genetic and biochemical approaches, we found that L-Arg depletion also induced autophagy. Deprivation of L-Arg induced EIF2S1 (eIF2α), MAPK8 (JNK), BCL2 (Bcl-2) phosphorylation, and displacement of BECN1 (Beclin 1) binding to BCL2, leading to autophagosome formation. Silencing of ERN1 (IRE1α) prevented the induction of autophagy as well as MAPK8 activation, BCL2 phosphorylation and XBP1 splicing, whereas led T lymphocytes to apoptosis under L-Arg starvation, suggesting that the ERN1-MAPK8 pathway plays a major role in the activation of autophagy following L-Arg depletion. Autophagy was required for survival of T lymphocytes in the absence of L-Arg, and resulted in a reversible process. Replenishment of L-Arg made T lymphocytes to regain the normal cell cycle profile and proliferate, whereas autophagy was inhibited. Inhibition of autophagy by ERN1, BECN1 and ATG7 silencing, or by pharmacological inhibitors, promoted cell death of T lymphocytes incubated in the absence of L-Arg. Our data indicate for the first time that depletion of L-Arg in T lymphocytes leads to a reversible response that preserves T lymphocytes through ER stress and autophagy, while remaining arrested at G0/G1. Our data also show that the L-Arg depletion-induced ER stress response could lead to apoptosis when autophagy is blocked.  相似文献   

17.
Membrane proteins of the amino acid-polyamine-organocation (APC) superfamily transport amino acids and amines across membranes and play an important role in the regulation of cellular processes. We report the heterologous production of the LysP-related transporter STM2200 from Salmonella typhimurium in Escherichia coli, its purification, and functional characterization. STM2200 is assumed to be a proton-dependent APC transporter of l-lysine. The functional interaction between basic amino acids and STM2200 was investigated by thermoanalytical methods, i.e. differential scanning and isothermal titration calorimetry. Binding of l-lysine to STM2200 in its solubilized monomer form is entropy-driven. It is characterized by a dissociation constant of 40 μm at pH 5.9 and is highly selective; no evidence was found for the binding of l-arginine, l-ornithine, l-2,4-diaminobutyric acid, and l-alanine. d-Lysine is bound 45 times more weakly than its l-chiral form. We thus postulate that STM2200 functions as a specific transport protein. Based on the crystal structure of ApcT (Shaffer, P. L., Goehring, A., Shankaranarayanan, A., and Gouaux, E. (2009) Science 325, 1010–1014), a proton-dependent amino acid transporter of the APC superfamily, a homology model of STM2200 was created. Docking studies allowed identification of possible ligand binding sites. The resulting predictions indicated that Glu-222 and Arg-395 of STM2200 are markedly involved in ligand binding, whereas Lys-163 is suggested to be of structural and functional relevance. Selected variants of STM2200 where these three amino acid residues were substituted using single site-directed mutagenesis showed no evidence for l-lysine binding by isothermal titration calorimetry, which confirmed the predictions. Molecular aspects of the observed ligand specificity are discussed.  相似文献   

18.
Previously, we successfully cloned a d-cycloserine (d-CS) biosynthetic gene cluster consisting of 10 open reading frames (designated dcsA to dcsJ) from d-CS-producing Streptomyces lavendulae ATCC 11924. In this study, we put four d-CS biosynthetic genes (dcsC, dcsD, dcsE, and dcsG) in tandem under the control of the T7 promoter in an Escherichia coli host. SDS-PAGE analysis demonstrated that the 4 gene products were simultaneously expressed in host cells. When l-serine and hydroxyurea (HU), the precursors of d-CS, were incubated together with the E. coli resting cell suspension, the cells produced significant amounts of d-CS (350 ± 20 μM). To increase the productivity of d-CS, the dcsJ gene, which might be responsible for the d-CS excretion, was connected downstream of the four genes. The E. coli resting cells harboring the five genes produced d-CS at 660 ± 31 μM. The dcsD gene product, DcsD, forms O-ureido-l-serine from O-acetyl-l-serine (OAS) and HU, which are intermediates in d-CS biosynthesis. DcsD also catalyzes the formation of l-cysteine from OAS and H2S. To repress the side catalytic activity of DcsD, the E. coli chromosomal cysJ and cysK genes, encoding the sulfite reductase α subunit and OAS sulfhydrylase, respectively, were disrupted. When resting cells of the double-knockout mutant harboring the four d-CS biosynthetic genes, together with dcsJ, were incubated with l-serine and HU, the d-CS production was 980 ± 57 μM, which is comparable to that of d-CS-producing S. lavendulae ATCC 11924 (930 ± 36 μM).  相似文献   

19.
We developed a novel process for efficient synthesis of l-threo-3-hydroxyaspartic acid (l-THA) using microbial hydroxylase and hydrolase. A well-characterized mutant of asparagine hydroxylase (AsnO-D241N) and its homologous enzyme (SCO2693-D246N) were adaptable to the direct hydroxylation of l-aspartic acid; however, the yields were strictly low. Therefore, the highly stable and efficient wild-type asparagine hydroxylases AsnO and SCO2693 were employed to synthesize l-THA. By using these recombinant enzymes, l-THA was obtained by l-asparagine hydroxylation by AsnO followed by amide hydrolysis by asparaginase via 3-hydroxyasparagine. Subsequently, the two-step reaction was adapted to one-pot bioconversion in a test tube. l-THA was obtained in a small amount with a molar yield of 0.076% by using intact Escherichia coli expressing the asnO gene, and thus, two asparaginase-deficient mutants of E. coli were investigated. A remarkably increased l-THA yield of 8.2% was obtained with the asparaginase I-deficient mutant. When the expression level of the asnO gene was enhanced by using the T7 promoter in E. coli instead of the lac promoter, the l-THA yield was significantly increased to 92%. By using a combination of the E. coli asparaginase I-deficient mutant and the T7 expression system, a whole-cell reaction in a jar fermentor was conducted, and consequently, l-THA was successfully obtained from l-asparagine with a maximum yield of 96% in less time than with test tube-scale production. These results indicate that asparagine hydroxylation followed by hydrolysis would be applicable to the efficient production of l-THA.  相似文献   

20.
The oxidative d-xylose catabolic pathway of Caulobacter crescentus, encoded by the xylXABCD operon, was expressed in the gram-negative bacterium Pseudomonas putida S12. This engineered transformant strain was able to grow on d-xylose as a sole carbon source with a biomass yield of 53% (based on g [dry weight] g d-xylose−1) and a maximum growth rate of 0.21 h−1. Remarkably, most of the genes of the xylXABCD operon appeared to be dispensable for growth on d-xylose. Only the xylD gene, encoding d-xylonate dehydratase, proved to be essential for establishing an oxidative d-xylose catabolic pathway in P. putida S12. The growth performance on d-xylose was, however, greatly improved by coexpression of xylXA, encoding 2-keto-3-deoxy-d-xylonate dehydratase and α-ketoglutaric semialdehyde dehydrogenase, respectively. The endogenous periplasmic glucose dehydrogenase (Gcd) of P. putida S12 was found to play a key role in efficient oxidative d-xylose utilization. Gcd activity not only contributes to d-xylose oxidation but also prevents the intracellular accumulation of toxic catabolic intermediates which delays or even eliminates growth on d-xylose.The requirement for renewable alternatives to replace oil-based chemicals and fuels necessitates development of novel technologies. Lignocellulose provides a promising alternative feedstock. However, since the pentose sugar fraction may account for up to 25% of lignocellulosic biomass (12), it is essential that this fraction is utilized efficiently to obtain cost-effective biochemical production. In a previous study, the solvent-tolerant bacterium Pseudomonas putida S12, known for its use as a platform host for the production of aromatic compounds (15, 16, 19, 22), was engineered to use d-xylose as a sole carbon source. This was achieved by introducing genes encoding the phosphorylative d-xylose metabolic pathway of Escherichia coli, followed by laboratory evolution (14). Prior to evolutionary improvement, extensive oxidation of d-xylose to d-xylonate occurred, resulting in a very low biomass-for-substrate yield as d-xylonate is a metabolic dead-end product in P. putida. The evolution approach resulted in elimination of the activity of periplasmic glucose dehydrogenase (Gcd), the enzyme responsible for d-xylose oxidation, which turned out to be a critical step in optimizing phosphorylative d-xylose utilization in P. putida S12.Instead of prevention of endogenous oxidation of d-xylose, this oxidation may be used to our advantage when it is combined with an oxidative d-xylose metabolic pathway, such as the pathways described for several Pseudomonas species, Caulobacter crescentus, and Haloarcula marismortui (7, 11, 18, 20). In these pathways, d-xylonate is dehydrated to 2-keto-3-deoxy-d-xylonate. This intermediate either can be cleaved into pyruvate and glycolaldehyde (7) or is further dehydrated to α-ketoglutaric semialdehyde (α-KGSA). In the final step of the latter pathway, α-KGSA is oxidized to the tricarboxylic acid (TCA) cycle intermediate α-ketoglutarate (18, 20).In addition to Gcd (PP1444), some of the enzymes required for oxidative d-xylose metabolism are expected to be endogenous in P. putida S12. Transport of d-xylonate into the cytoplasm likely occurs through the gluconate transporter (encoded by gntP [PP3417]). The enzyme catalyzing the final step of the pathway, α-KGSA dehydrogenase, is also likely to be present (presumably PP1256 and/or PP3602) because of the requirement for metabolism of 4-hydroxyproline (1), a compound that is efficiently utilized by P. putida S12. In view of these properties, the most obvious approach for constructing d-xylose-utilizing P. putida S12 is reconstruction of a complete oxidative d-xylose metabolic pathway by introducing the parts of such a pathway that complement the endogenous activities. Recently, the genetic information for one such oxidative d-xylose pathway has become available (18), enabling the approach used in the present study, i.e., expression of the oxidative d-xylose metabolic pathway of C. crescentus in P. putida S12 and investigation of the contribution of endogenous enzyme activities.  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