首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Intracellular thiols like L-cystine and L-cystine play a critical role in the regulation of cellular processes. Here we show that Escherichia coli has two L-cystine transporters, the symporter YdjN and the ATP-binding cassette importer FliY-YecSC. These proteins import L-cystine, an oxidized product of L-cystine from the periplasm to the cytoplasm. The symporter YdjN, which is expected to be a new member of the L-cystine regulon, is a low affinity L-cystine transporter (K m = 1.1 μM) that is mainly involved in L-cystine uptake from outside as a nutrient. E. coli has only two L-cystine importers because ΔydjNΔyecS mutant cells are not capable of growing in the minimal medium containing L-cystine as a sole sulfur source. Another protein YecSC is the FliY-dependent L-cystine transporter that functions cooperatively with the L-cystine transporter YdeD, which exports L-cystine as reducing equivalents from the cytoplasm to the periplasm, to prevent E. coli cells from oxidative stress. The exported L-cystine can reduce the periplasmic hydrogen peroxide to water, and then generated L-cystine is imported back into the cytoplasm via the ATP-binding cassette transporter YecSC with a high affinity to L-cystine (K m = 110 nM) in a manner dependent on FliY, the periplasmic L-cystine-binding protein. The double disruption of ydeD and fliY increased cellular levels of lipid peroxides. From these findings, we propose that the hydrogen peroxide-inducible L-cystine/L-cystine shuttle system plays a role of detoxification of hydrogen peroxide before lipid peroxidation occurs, and then might specific prevent damage to membrane lipids.  相似文献   

2.
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   

3.
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   

4.
5.
Emission of Hydrogen Sulfide by Leaf Tissue in Response to l-Cysteine   总被引:6,自引:2,他引:4  
Leaf discs and detached leaves exposed to l-cysteine emitted a volatile sulfur compound which was proven by gas chromatography to be H2S. This phenomenon was demonstrated in all nine species tested (Cucumis sativus, Cucurbita pepo, Nicotiana tabacum, Coleus blumei, Beta vulgaris, Phaseolus vulgaris, Medicago sativa, Hordeum vulgare, and Gossypium hirsutum). The emission of volatile sulfur by cucumber leaves occurred in the dark at a similar rate to that in the light. The emission of leaf discs reached the maximal rate, more than 40 picomoles per minute per square centimeter, 2 to 4 hours after starting exposure to l-cysteine; then it decreased. In the case of detached leaves, the maximum occurred 5 to 10 h after starting exposure. The average emission rate of H2S during the first 4 hours from leaf discs of cucurbits in response to 10 millimolar l-cysteine, was usually more than 40 picomoles per minute per square centimeter, i.e. 0.24 micromoles per hour per square decimeter. Leaf discs exposed to 1 millimolar l-cysteine emitted only 2% as much as did the discs exposed to 10 millimolar l-cysteine. The emission from leaf discs and from detached leaves lasted for at least 5 and 15 hours, respectively. However, several hours after the maximal emission, injury of the leaves, manifested as chlorosis, was evident. H2S emission was a specific consequence of exposure to l-cysteine; neither d-cysteine nor l-cystine elicited H2S emission. Aminooxyacetic acid, an inhibitor of pyridoxal phosphate dependent enzymes, inhibited the emission. In a cell free system from cucumber leaves, H2S formation and its release occurred in response to l-cysteine. Feeding experiments with [35S]l-cysteine showed that most of the sulfur in H2S was derived from sulfur in the l-cysteine supplied and that the H2S emitted for 9 hours accounted for 7 to 10% of l-cysteine taken up. 35S-labeled SO32− and SO42− were found in the tissue extract in addition to internal soluble S2−. These findings suggest the existence of a sulfur cycle which converts l-cysteine to SO42− through cysteine desulfhydration.  相似文献   

6.
Quinto G 《Applied microbiology》1966,14(6):1022-1026
Nutritional studies were performed on nine Bacteroides strains, by use of the methodology and media of anaerobic rumen microbiology. Ristella perfoetens CCI required l-arginine hydrochloride, l-tryptophan, l-leucine, l-histidine hydrochloride, l-cysteine hydrochloride, dl-valine, dl-tyrosine, and the vitamin calcium-d-pantothenate, since scant turbidity developed in media without these nutrients. R. perfoetens was stimulated by glycine, dl-lysine hydrochloride, dl-isoleucine, l-proline, l-glutamic acid, dl-alanine, dl-phenylalanine, dl-methionine, and the vitamins nicotinamide and p-aminobenzoic acid, since maximal turbidity developed more slowly in media without these nutrients than in complete medium. Medium A-23, which was devised for R. perfoetens, contained salts, 0.0002% nicotinamide and calcium d-pantothenate, 0.00001% p-aminobenzoic acid, 0.044% l-tryptophan, 0.09% l-glutamic acid, and 0.1% of the other 13 amino acids listed above. Zuberella clostridiformis and seven strains of R. pseudoinsolita did not require vitamins, and showed no absolute requirement for any one amino acid. Various strains produced maximal turbidity more slowly in media deficient in l-proline, glycine, l-glutamic acid, dl-serine, l-histidine hydrochloride, dl-alanine, or l-cysteine hydrochloride, than in complete medium. These eight strains grew optimally in medium A-23 plus 0.1% dl-serine but without vitamins.  相似文献   

7.
Promotion of seed germination by cyanide   总被引:2,自引:2,他引:0  
Potassium cyanide at 3 μm to 10 mm promotes germination of Amaranthus albus, Lactuca sativa, and Lepidium virginicum seeds. l-Cysteine hydrogen sulfide lyase, which catalyzes the reaction of HCN with l-cysteine to form β-l cyanoalanine, is active in the seeds. β-l-Cyanoalanine is the most effective of the 23 α-amino acids tested for promoting germination of A. albus seeds. Aspartate, which is produced by enzymatic hydrolysis of asparagine formed by hydrolysis from β-cyanoalanine, is the second most effective of the 23 amino acids. Uptake of aspartate-4-14C is much lower than of cyanide.  相似文献   

8.
Dihydrodipicolinate synthase (EC 4.2.1.52), the first enzyme unique to lysine biosynthesis in bacteria and higher plants, has been purified to homogeneity from etiolated pea (Pisum sativum) seedlings using a combination of conventional and affinity chromatographic steps. This is the first report on a homogeneous preparation of native dihydrodipicolinate synthase from a plant source. The pea dihydrodipicolinate synthase has an apparent molecular weight of 127,000 and is composed of three identical subunits of 43,000 as determined by gel filtration and cross-linking experiments. The trimeric quaternary structure resembles the trimeric structure of other aldolases, such as 2-keto-3-deoxy-6-phosphogluconic acid aldolase, which catalyze similar aldol condensations. The amino acid compositions of dihydrodipicolinate synthase from pea and Escherichia coli are similar, the most significant difference concerns the methionine content: dihydrodipicolinate synthase from pea contains 22 moles of methionine residue per mole of native protein, contrary to the E. coli enzyme, which does not contain this amino acid at all. Dihydrodipicolinate synthase from pea is highly specific for the substrates pyruvate and l-aspartate-β-semialdehyde; it follows Michaelis-Menten kinetics for both substrates. The pyruvate and l-aspartate-β-semialdehyde have Michaelis constant values of 1.70 and 0.40 millimolar, respectively. l-Lysine, S-(2-aminoethyl)-l-cysteine, and l-α-(2-aminoethoxyvinyl)glycine are strong allosteric inhibitors of the enzyme with 50% inhibitory values of 20, 160, and 155 millimolar, respectively. The inhibition by l-lysine and l-α-(2-aminoethoxyvinyl)glycine is noncompetitive towards l-aspartate-β-semialdehyde, whereas S-(2-aminoethyl)-l-cysteine inhibits dihydrodipicolinate synthase competitively with respect to l-aspartate-β-semialdehyde. Furthermore, the addition of (2R,3S,6S)-2,6-diamino-3-hydroxy-heptandioic acid (1.2 millimolar) and (2S,6R/S)-2,6-diamino-6-phosphono-hexanic acid (1.2 millimolar) activates dihydrodipicolinate synthase from pea by a factor of 1.4 and 1.2, respectively. This is the first reported activation process found for dihydrodipicolinate synthase.  相似文献   

9.
The metabolism of S-methyl-l-cysteine   总被引:1,自引:1,他引:0  
1. Methylsulphinylacetic acid, 2-hydroxy-3-methylsulphinylpropionic acid and methylmercapturic acid sulphoxide (N-acetyl-S-methyl-l-cysteine S-oxide) were isolated as their dicyclohexylammonium salts from the urine of rats after they had been dosed with S-methyl-l-cysteine. 2. A fourth sulphoxide was isolated but not identified. 3. The excretion of sulphate in the urine of rats dosed with S-methyl-l-cysteine was measured. 4. The metabolism of S-methyl-l-cysteine by the hamster and guinea pig was examined chromatographically. 5. The preparation of the following compounds is reported: (−)-dicyclohexylammonium methyl-mercapturate sulphoxide; the dicyclohexylammonium salts of the optically inactive forms of 2-hydroxy-3-methylthiopropionic acid, 2-hydroxy-3-methyl-sulphinylpropionic acid and methylsulphinylacetic acid.  相似文献   

10.
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   

11.
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   

12.
Early studies revealed that chicken embryos incubated with a rare analog of l-proline, 4-oxo-l-proline, showed increased levels of the metabolite 4-hydroxy-l-proline. In 1962, 4-oxo-l-proline reductase, an enzyme responsible for the reduction of 4-oxo-l-proline, was partially purified from rabbit kidneys and characterized biochemically. However, only recently was the molecular identity of this enzyme solved. Here, we report the purification from rat kidneys, identification, and biochemical characterization of 4-oxo-l-proline reductase. Following mass spectrometry analysis of the purified protein preparation, the previously annotated mammalian cytosolic type 2 (R)-β-hydroxybutyrate dehydrogenase (BDH2) emerged as the only candidate for the reductase. We subsequently expressed rat and human BDH2 in Escherichia coli, then purified it, and showed that it catalyzed the reversible reduction of 4-oxo-l-proline to cis-4-hydroxy-l-proline via chromatographic and tandem mass spectrometry analysis. Specificity studies with an array of compounds carried out on both enzymes showed that 4-oxo-l-proline was the best substrate, and the human enzyme acted with 12,500-fold higher catalytic efficiency on 4-oxo-l-proline than on (R)-β-hydroxybutyrate. In addition, human embryonic kidney 293T (HEK293T) cells efficiently metabolized 4-oxo-l-proline to cis-4-hydroxy-l-proline, whereas HEK293T BDH2 KO cells were incapable of producing cis-4-hydroxy-l-proline. Both WT and KO HEK293T cells also produced trans-4-hydroxy-l-proline in the presence of 4-oxo-l-proline, suggesting that the latter compound might interfere with the trans-4-hydroxy-l-proline breakdown in human cells. We conclude that BDH2 is a mammalian 4-oxo-l-proline reductase that converts 4-oxo-l-proline to cis-4-hydroxy-l-proline and not to trans-4-hydroxy-l-proline, as originally thought. We also hypothesize that this enzyme may be a potential source of cis-4-hydroxy-l-proline in mammalian tissues.  相似文献   

13.
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   

14.
The human mitochondrial outer membrane protein mitoNEET is a novel target of the type II diabetes drug pioglitazone. The C-terminal cytosolic domain of mitoNEET hosts a redox-active [2Fe-2S] cluster via an unusual ligand arrangement of three cysteine residues and one histidine residue. Here we report that human mitoNEET [2Fe-2S] clusters are fully reduced when expressed in Escherichia coli cells. In vitro studies show that purified mitoNEET [2Fe-2S] clusters can be partially reduced by monothiols such as reduced glutathione, l-cysteine or N-acetyl-l-cysteine and fully reduced by dithiothreitol or the E. coli thioredoxin/thioredoxin reductase system under anaerobic conditions. Importantly, thiol-reduced mitoNEET [2Fe-2S] clusters can be reversibly oxidized by hydrogen peroxide without disruption of the clusters in vitro and in E. coli cells, indicating that mitoNEET may act as a sensor of oxidative signals to regulate mitochondrial functions via its [2Fe-2S] clusters. Furthermore, the binding of the type II diabetes drug pioglitazone in mitoNEET effectively inhibits the thiol-mediated reduction of [2Fe-2S] clusters, suggesting that pioglitazone may modulate the function of mitoNEET by blocking the thiol-mediated reduction of [2Fe-2S] clusters in the protein.  相似文献   

15.
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   

16.
Glutathione (GSH) is an essential antioxidant responsible for the maintenance of intracellular redox homeostasis. As tumors outgrow their blood supply and become hypoxic, their redox homeostasis is challenged by the production of nitric oxide and reactive oxygen species (ROS). In gliomas, the sustained import of l-cystine via the l-cystine/l-glutamate exchanger, system xc, is rate-limiting for the synthesis of GSH. We show that hypoxia causes a significant increase in NO and ROS but without affecting glioma cell growth. This is explained by a concomitant increase in the utilization of GSH, which is accompanied by an increase in the cell-surface expression of xCT, the catalytic subunit of system xc, and l-cystine uptake. Growth was inhibited when GSH synthesis was blocked by buthionine sulfoximine (BSO), an inhibitor of the enzyme required for GSH synthesis, or when cells were deprived of l-cystine. These findings suggest that glioma cells show an increased requirement for GSH to maintain growth under hypoxic conditions. Therefore, approaches that limit GSH synthesis such as blocking system xc may be considered as an adjuvant to radiation or chemotherapy.  相似文献   

17.
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   

18.
Several esters of the α-N-toluene-p-sulphonyl and α-N-benzoyl derivatives of S-(3-aminopropyl)-l-cysteine and the methyl ester of S-(4-aminobutyl)-N-toluene-p-sulphonyl-l-cysteine were synthesized. The kinetics of hydrolysis of these and esters of the α-N-toluene-p-sulphonyl and α-N-benzoyl derivatives of l-arginine, l-lysine, S-(2-aminoethyl)-l-cysteine and esters of γ-guanidino-l-α-toluene-p-sulphonamidobutyric acid and α-N-toluene-p-sulphonyl-l-homoarginine by α- and β-trypsin were compared. On the basis of values of the specificity constants (kcat./Km), the two enzymes display similar catalytic efficiency towards some substrates. In other cases α-trypsin is less efficient than β-trypsin. It is possible that α-trypsin possesses greater molecular flexibility than β-trypsin.  相似文献   

19.
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   

20.
The property of loose stereochemical control at aldol products from aldolases helped to synthesize multiple polyhydroxylated compounds with nonnatural stereoconfiguration. In this study, we discovered for the first time that some fructose 1,6-diphosphate aldolases (FruA) and tagatose 1,6-diphosphate (TagA) aldolases lost their strict stereoselectivity when using l-glyceraldehyde and synthesized not only l-sorbose but also a high proportion of l-psicose. Among the aldolases tested, TagA from Bacillus licheniformis (BGatY) showed the highest enzyme activity with l-glyceraldehyde. Subsequently, a “one-pot” reaction based on BGatY and fructose-1-phosphatase (YqaB) generated 378 mg/liter l-psicose and 199 mg/liter l-sorbose from dihydroxyacetone-phosphate (DHAP) and l-glyceraldehyde. Because of the high cost and instability of DHAP, a microbial fermentation strategy was used further to produce l-sorbose/l-psicose from glucose and l-glyceraldehyde, in which DHAP was obtained from glucose through the glycolytic pathway, and some recombination pathways based on FruA or TagA and YqaB were constructed in Escherichia coli and Corynebacterium glutamicum strains. After evaluation of different host cells and combinations of FruA or TagA with YqaB and optimization of gene expression, recombinant C. glutamicum strain WT(pXFTY) was selected and produced 2.53 g/liter total ketoses, with a yield of 0.50 g/g l-glyceraldehyde. Moreover, deletion of gene cgl0331, encoding the Zn-dependent alcohol dehydrogenase in C. glutamicum, was confirmed for the first time to significantly decrease conversion of l-glyceraldehyde to glycerol and to increase yield of target products. Finally, fed-batch culture of strain SY14(pXFTY) produced 3.5 g/liter l-sorbose and 2.3 g/liter l-psicose, with a yield of 0.61 g/g l-glyceraldehyde. This microbial fermentation strategy also could be applied to efficiently synthesize other l-sugars.  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