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 250 毫秒
1.
d-Galacturonic acid can be obtained by hydrolyzing pectin, which is an abundant and low value raw material. By means of metabolic engineering, we constructed fungal strains for the conversion of d-galacturonate to meso-galactarate (mucate). Galactarate has applications in food, cosmetics, and pharmaceuticals and as a platform chemical. In fungi d-galacturonate is catabolized through a reductive pathway with a d-galacturonate reductase as the first enzyme. Deleting the corresponding gene in the fungi Hypocrea jecorina and Aspergillus niger resulted in strains unable to grow on d-galacturonate. The genes of the pathway for d-galacturonate catabolism were upregulated in the presence of d-galacturonate in A. niger, even when the gene for d-galacturonate reductase was deleted, indicating that d-galacturonate itself is an inducer for the pathway. A bacterial gene coding for a d-galacturonate dehydrogenase catalyzing the NAD-dependent oxidation of d-galacturonate to galactarate was introduced to both strains with disrupted d-galacturonate catabolism. Both strains converted d-galacturonate to galactarate. The resulting H. jecorina strain produced galactarate at high yield. The A. niger strain regained the ability to grow on d-galacturonate when the d-galacturonate dehydrogenase was introduced, suggesting that it has a pathway for galactarate catabolism.d-Galacturonate is the main component of pectin, an abundant and cheap raw material. Sugar beet pulp and citrus peel are both rich in pectin residues. At present, these residues are mainly used as cattle feed. However, since energy-consuming drying and pelletizing of the residues is required to prevent them from rotting, it is not always economical to process the residues, and it is desirable to find alternative uses.Various microbes which live on decaying plant material have the ability to catabolize d-galacturonate using various, completely different pathways (19). Eukaryotic microorganisms use a reductive pathway in which d-galacturonate is first reduced to l-galactonate by an NAD(P)H-dependent reductase (12, 17). In the following steps a dehydratase, aldolase, and reductase convert the l-galactonate to pyruvate and glycerol (9, 11, 14).In Hypocrea jecorina (anamorph Trichoderma reesei) the gar1 gene codes for a strictly NADPH-dependent d-galacturonate reductase. In Aspergillus niger a homologue gene sequence, gar2, exists; however, a different gene, gaaA, is upregulated during growth on d-galacturonate containing medium (16). The gaaA codes for a d-galacturonate reductase with different kinetic properties than the H. jecorina enzyme, having a higher affinity toward d-galacturonate and using either NADH or NADPH as cofactor. It is not known whether gar2 codes for an active protein.Some bacteria, such as Agrobacterium tumefaciens or Pseudomonas syringae, have an oxidative pathway for d-galacturonate catabolism. In this pathway d-galacturonate is first oxidized to meso-galactarate (mucate) by an NAD-utilizing d-galacturonate dehydrogenase. Galactarate is then converted in the following steps to α-ketoglutarate. This route is sometimes called the α-ketoglutarate pathway (20). Galactarate can also be catabolized through the glycerate pathway (20). The products of this pathway are pyruvate and d-glycerate. These pathways have been described in prokaryotes, and it is not certain whether similar pathways also exist in fungi, some of which are able to metabolize galactarate.d-Galacturonate dehydrogenase (EC 1.1.1.203) has been described in Agrobacterium tumefaciens and in Pseudomonas syringae, and the enzymes from these organisms have been purified and characterized (3, 6, 22). Recently, the corresponding genes were also identified (4, 24). Both enzymes are specific for NAD as a cofactor but are not specific for the substrate. They oxidize d-galacturonate and d-glucuronate to meso-galactarate (mucate) and d-glucarate (saccharate), respectively. The reaction product is probably the hexaro-lactone which spontaneously hydrolyzes. The reverse reaction can only be observed at acidic pH where some of the galactarate is in the lactone form (22).We describe here strains of filamentous fungi that have been genetically engineered to produce galactarate by disruption of d-galacturonate reductase and expression of d-galacturonate dehydrogenase (Fig. (Fig.1).1). Galactarate is currently commercially produced from d-galactose by oxidation with nitric acid (1) or from d-galacturonate by electrolytic oxidation (8). Oxidation with nitric acid is expensive and produces toxic wastes. Galactarate is used as a chelator and in skin care products. It was formerly used as a leavening agent in self-rising flour (2) and has potential applications in polymer synthesis (10) and as a platform chemical (for a review, see reference 13).Open in a separate windowFIG. 1.Engineering the d-galacturonic acid pathway in fungi. Deletion of the gene encoding d-galacturonate reductase resulted in strains unable to utilize d-galacturonic acid as a carbon source. The expression of a bacterial udh gene, encoding an NAD-dependent d-galacturonate dehydrogenase, resulted in fungal strains, which were able to oxidize d-galacturonic acid to meso-galactaric acid (mucic acid). d-Galacturonate dehydrogenase forms a galactaro-lactone which spontaneously hydrolyzes.  相似文献   

2.
d-Glucose 6-phosphate cycloaldolase is inhibited 83% by 0.66 mm EDTA and stimulated 1.7-fold by 0.6 mm KCl. Dihydroxyacetone phosphate, an analog of the last three carbons in the proposed intermediate, d-xylo-5-hexulose 6-phosphate, acts as a partially competitive inhibitor. Treatment with NaBH4 in the presence of dihydroxyacetone phosphate does not cause permanent inactivation as would be expected if a Schiff base were being formed. In these properties it resembles a type II, metal-containing aldolase. Photooxidation in the presence of Rose Bengal inactivates this enzyme. NAD+ partially protects against this photooxidation. Cells grown on medium lacking myoinositol had four times as much enzyme activity as cells grown on medium containing 100 mg of myoinositol per liter.  相似文献   

3.
SalM is a short-chain dehydrogenase/reductase enzyme from the marine actinomycete Salinispora tropica that is involved in the biosynthesis of chloroethylmalonyl-CoA, a novel halogenated polyketide synthase extender unit of the proteasome inhibitor salinosporamide A. SalM was heterologously overexpressed in Escherichia coli and characterized in vitro for its substrate specificity, kinetics, and reaction profile. A sensitive real-time 13C NMR assay was developed to visualize the oxidation of 5-chloro-5-deoxy-d-ribose to 5-chloro-5-deoxy-d-ribono-γ-lactone in an NAD+-dependent reaction, followed by spontaneous lactone hydrolysis to 5-chloro-5-deoxy-d-ribonate. Although short-chain dehydrogenase/reductase enzymes are widely regarded as metal-independent, a strong divalent metal cation dependence for Mg2+, Ca2+, or Mn2+ was observed with SalM. Oxidative activity was also measured with the alternative substrates d-erythrose and d-ribose, making SalM the first reported stereospecific non-phosphorylative ribose 1-dehydrogenase.  相似文献   

4.
Several weakly transported sugars were tested for transport by the Na+-dependent sugar carrier with slices of everted hamster intestinal tissue. Sugars were assumed to be transported by this carrier if the accumulation was diminished in the absence of Na+ and in the presence of the competitive inhibitor 1,5-anhydro-d-glucitol. The extent of accumulation was correlated with the number of hydroxyl groups in the d-gluco configuration if the ring oxygen was placed in the normal d-glucose position. 5-Thio-d-glucose, with a sulphur atom in the ring, was transported at about the same rate as d-glucose and had a similar Ki for d-galactose transport, but myoinositol was poorly accumulated. It is suggested that there is no hydrogen bonding at the ring oxygen atom, but that the oxygen atom is found at this position as a result of steric constraints. No sugar without a hydroxyl group in the d-gluco position at C-2 of the sugar, including d-mannose, 2-deoxy-d-glucose, 2-chloro-2-deoxy-d-glucose and 2-deoxy-2-fluoro-d-glucose, was transported by the Na+-dependent carrier, but these sugars and l-fucose weakly and competitively inhibit the Na+-dependent accumulation of l-glucose into slices of everted hamster intestinal tissue. It is concluded that the bond between the carrier and C-2 of the sugar may be covalent, and a possible mechanism for active intestinal transport is proposed.  相似文献   

5.
Lactobacillus casei strains 64H and BL23, but not ATCC 334, are able to ferment d-ribitol (also called d-adonitol). However, a BL23-derived ptsI mutant lacking enzyme I of the phosphoenolpyruvate:carbohydrate phosphotransferase system (PTS) was not able to utilize this pentitol, suggesting that strain BL23 transports and phosphorylates d-ribitol via a PTS. We identified an 11-kb region in the genome sequence of L. casei strain BL23 (LCABL_29160 to LCABL_29270) which is absent from strain ATCC 334 and which contains the genes for a GlpR/IolR-like repressor, the four components of a mannose-type PTS, and six metabolic enzymes potentially involved in d-ribitol metabolism. Deletion of the gene encoding the EIIB component of the presumed ribitol PTS indeed prevented d-ribitol fermentation. In addition, we overexpressed the six catabolic genes, purified the encoded enzymes, and determined the activities of four of them. They encode a d-ribitol-5-phosphate (d-ribitol-5-P) 2-dehydrogenase, a d-ribulose-5-P 3-epimerase, a d-ribose-5-P isomerase, and a d-xylulose-5-P phosphoketolase. In the first catabolic step, the protein d-ribitol-5-P 2-dehydrogenase uses NAD+ to oxidize d-ribitol-5-P formed during PTS-catalyzed transport to d-ribulose-5-P, which, in turn, is converted to d-xylulose-5-P by the enzyme d-ribulose-5-P 3-epimerase. Finally, the resulting d-xylulose-5-P is split by d-xylulose-5-P phosphoketolase in an inorganic phosphate-requiring reaction into acetylphosphate and the glycolytic intermediate d-glyceraldehyde-3-P. The three remaining enzymes, one of which was identified as d-ribose-5-P-isomerase, probably catalyze an alternative ribitol degradation pathway, which might be functional in L. casei strain 64H but not in BL23, because one of the BL23 genes carries a frameshift mutation.  相似文献   

6.
The metabolism of myo-inositol-2-14C, d-glucuronate-1-14C, d-glucuronate-6-14C, and l-methionine-methyl-14C to cell wall polysaccharides was investigated in excised root-tips of 3 day old Zea mays seedlings. From myo-inositol, about one-half of incorporated label was recovered in ethanol insoluble residues. Of this label, about 90% was solubilized by treatment, first with a preparation of pectinase-EDTA, then with dilute hydrochloric acid. The only labeled constituents in these hydrolyzates were d-galacturonic acid, d-glucuronic acid, 4-O-methyl-d-glucuronic acid, d-xylose, and l-arabinose, or larger oligosaccharide fragments containing these units. Medium external to excised root-tips grown under sterile conditions in myo-inositol-2-14C contained labeled polysaccharide.  相似文献   

7.
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   

8.
This study presents evidence for a new enzyme, d-ribose-5-P reductase, which catalyzes the reaction: d-ribose-5-P + NADPH + H+d-ribitol-5-P + NADP+. The enzyme was isolated from Adonis vernalis L. leaves in 38% yield and was purified 71-fold. The reductase was NADPH specific and had a pH optimum in the range of 5.5 to 6.0. The Michaelis constant value for d-ribose-5-P reduction was 1.35 millimolar. The enzyme also reduced d-erythrose-4-P, d-erythrose, dl-glyceraldehyde, and the aromatic aldehyde 3-pyridinecarboxaldehyde. Hexoses, hexose phosphates, pentoses, and dihydroxyacetone did not serve as substrates. d-Ribose-5-P reductase is distinct from the other known ribitol synthesizing enzymes detected in bacteria and yeast, and may be responsible for ribitol synthesis in Adonis vernalis.  相似文献   

9.
Polyol dehydrogenases of Acetobacter melanogenum were investigated. Three polyol dehydrogenases, i. e. NAD+-linked d-mannitol dehydrogenase, NAD+-linked sorbitol dehydrogenase and NADP+-linked d-mannitol dehydrogenase, in the soluble fraction of the organism were purified 12-fold, 8-fold and 88-fold, respectively, by fractionation with ammonium sulfate and DEAE-cellulose column chromatography. NAD+-linked sorbitol dehydrogenase reduced 5-keto-d-fructose (5KF) to l-sorbose in the presence of NADH, whereas NADP+-linked d-mannitol dehydrogenase reduced the same substrate to d-fructose in the presence of NADPH. It was also shown that NAD+-linked d-mannitol dehydrogenase was specific for the interconversion between d-mannitol and d-fructose and that this enzyme was very unstable in alkaline conditions.  相似文献   

10.
The pathway of d-xylose degradation in archaea is unknown. In a previous study we identified in Haloarcula marismortui the first enzyme of xylose degradation, an inducible xylose dehydrogenase (Johnsen, U., and Schönheit, P. (2004) J. Bacteriol. 186, 6198–6207). Here we report a comprehensive study of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. The analyses include the following: (i) identification of the degradation pathway in vivo following 13C-labeling patterns of proteinogenic amino acids after growth on [13C]xylose; (ii) identification of xylose-induced genes by DNA microarray experiments; (iii) characterization of enzymes; and (iv) construction of in-frame deletion mutants and their functional analyses in growth experiments. Together, the data indicate that d-xylose is oxidized exclusively to the tricarboxylic acid cycle intermediate α-ketoglutarate, involving d-xylose dehydrogenase (HVO_B0028), a novel xylonate dehydratase (HVO_B0038A), 2-keto-3-deoxyxylonate dehydratase (HVO_B0027), and α-ketoglutarate semialdehyde dehydrogenase (HVO_B0039). The functional involvement of these enzymes in xylose degradation was proven by growth studies of the corresponding in-frame deletion mutants, which all lost the ability to grow on d-xylose, but growth on glucose was not significantly affected. This is the first report of an archaeal d-xylose degradation pathway that differs from the classical d-xylose pathway in most bacteria involving the formation of xylulose 5-phosphate as an intermediate. However, the pathway shows similarities to proposed oxidative pentose degradation pathways to α-ketoglutarate in few bacteria, e.g. Azospirillum brasilense and Caulobacter crescentus, and in the archaeon Sulfolobus solfataricus.d-Xylose, a constituent of the polymer xylan, is the major component of the hemicellulose plant cell wall material and thus one of the most abundant carbohydrates in nature. The utilization of d-xylose by microorganisms has been described in detail in bacteria and fungi, for which two different catabolic pathways have been reported. In many bacteria, such as Escherichia coli, Bacillus, and Lactobacillus species, xylose is converted by the activities of xylose isomerase and xylulose kinase to xylulose 5-phosphate as an intermediate, which is further degraded mainly by the pentose phosphate cycle or phosphoketolase pathway. Most fungi convert xylose to xylulose 5-phosphate via xylose reductase, xylitol dehydrogenase, and xylulose kinase. Xylulose 5-phosphate is also an intermediate of the most common l-arabinose degradation pathway in bacteria, e.g. of E. coli, via activities of isomerase, kinase, and epimerase (1).Recently, by genetic evidence, a third pathway of xylose degradation was proposed for the bacterium Caulobacter crescentus, in analogy to an alternative catabolic pathway of l-arabinose, reported for some bacteria, including species of Azospirillum, Pseudomonas, Rhizobium, Burkholderia, and Herbasprillum (2, 3). In these organisms l-arabinose is oxidatively degraded to α-ketoglutarate, an intermediate of the tricarboxylic acid cycle, via the activities of l-arabinose dehydrogenase, l-arabinolactonase, and two successive dehydration reactions forming 2-keto-3-deoxy-l-arabinoate and α-ketoglutarate semialdehyde; the latter compound is further oxidized to α-ketoglutarate via NADP+-specific α-ketoglutarate semialdehyde dehydrogenase (KGSADH).2 In a few Pseudomonas and Rhizobium species, a variant of this l-arabinose pathway was described involving aldolase cleavage of the intermediate 2-keto-3-deoxy-l-arabinoate to pyruvate and glycolaldehyde, rather than its dehydration and oxidation to α-ketoglutarate (4). Because of the presence of some analogous enzyme activities in xylose-grown cells of Azosprillum and Rhizobium, the oxidative pathway and its variant was also proposed as a catabolic pathway for d-xylose. Recent genetic analysis of Caulobacter crecentus indicates the presence of an oxidative pathway for d-xylose degradation to α-ketoglutarate. All genes encoding xylose dehydrogenase and putative lactonase, xylonate dehydratase, 2-keto-3-deoxylonate dehydratase, and KGSADH were found to be located on a xylose-inducible operon (5). With exception of xylose dehydrogenase, which has been partially characterized, the other postulated enzymes of the pathway have not been biochemically analyzed.The pathway of d-xylose degradation in the domain of archaea has not been studied so far. First analyses with the halophilic archaeon Haloarcula marismortui indicate that the initial step of d-xylose degradation involves a xylose-inducible xylose dehydrogenase (6) suggesting an oxidative pathway of xylose degradation to α-ketoglutarate, or to pyruvate and glycolaldehyde, in analogy to the proposed oxidative bacterial pentose degradation pathways. Recently, a detailed study of d-arabinose catabolism in the thermoacidophilic crenarchaeon Sulfolobus solfataricus was reported. d-Arabinose was found to be oxidized to α-ketoglutarate involving d-arabinose dehydrogenase, d-arabinoate dehydratase, 2-keto-3-deoxy-d-arabinoate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase (3).In this study, we present a comprehensive analysis of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. This halophilic archaeon was chosen because it exerts several suitable properties for the analyses. For example, it can be cultivated on synthetic media with sugars, e.g. xylose, an advantage for in vivo labeling studies in growing cultures. Furthermore, a shotgun DNA microarray of H. volcanii is available (7) allowing the identification of xylose-inducible genes, and H. volcanii is one of the few archaea for which an efficient protocol was recently described to generate in-frame deletion mutants.Accordingly, the d-xylose degradation pathway was elucidated following in vivo labeling experiments with [13C]xylose, DNA microarray analyses, and the characterization of enzymes involved and their encoding genes. The functional involvement of genes and enzymes was proven by constructing corresponding in-frame deletion mutants and their analysis by selective growth experiments on xylose versus glucose. The data show that d-xylose was exclusively degraded to α-ketoglutarate involving xylose dehydrogenase, a novel xylonate dehydratase, 2-keto-3-deoxyxylonate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase.  相似文献   

11.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

12.
Renal transport of four different categories of organic solutes, namely sugars, neutral amino acids, monocarboxylic acids and dicarboxylic acids, was studied by using the potential-sensitive dye 3,3′-diethyloxadicarbocyanine iodide in purified luminal-membrane and basolateral-membrane vesicles isolated from rabbit kidney cortex. Valinomycin-induced K+ diffusion potentials resulted in concomitant changes in dye–membrane-vesicle absorption spectra. Linear relationships were obtained between these changes and depolarization and hyperpolarization of the vesicles. Addition of d-glucose, l-phenylalanine, succinate or l-lactate to luminal-membrane vesicles, in the presence of an extravesicular>intravesicular Na+ gradient, resulted in rapid transient depolarization. With basolateral-membrane vesicles no electrogenic transport of d-glucose or l-phenylalanine was observed. Spectrophotometric competition studies revealed that d-galactose is electrogenically taken up by the same transport system as that for d-glucose, whereas l-phenylalanine, succinate and l-lactate are transported by different systems in luminal-membrane vesicles. The absorbance changes associated with simultaneous addition of d-glucose and l-phenylalanine were additive. The uptake of these solutes was influenced by the presence of Na+-salt anions of different permeabilities in the order: Cl>SO42−>gluconate. Addition of valinomycin to K+-loaded vesicles enhanced uptake of d-glucose and l-phenylalanine in the presence of an extravesicular>intravesicular Na+ gradient. Gramicidin or valinomycin plus nigericin diminished/abolished electrogenic solute uptake by Na+- or Na++K+-loaded vesicles respectively. These results strongly support the presence of Na+-dependent renal electrogenic transport of d-glucose, l-phenylalanine, succinate and l-lactate in luminal-membrane vesicles.  相似文献   

13.
Peptidoglycan hydrolases (PGHs) are responsible for bacterial cell lysis. Most PGHs have a modular structure comprising a catalytic domain and a cell wall-binding domain (CWBD). PGHs of bacteriophage origin, called endolysins, are involved in bacterial lysis at the end of the infection cycle. We have characterized two endolysins, Lc-Lys and Lc-Lys-2, identified in prophages present in the genome of Lactobacillus casei BL23. These two enzymes have different catalytic domains but similar putative C-terminal CWBDs. By analyzing purified peptidoglycan (PG) degradation products, we showed that Lc-Lys is an N-acetylmuramoyl-l-alanine amidase, whereas Lc-Lys-2 is a γ-d-glutamyl-l-lysyl endopeptidase. Remarkably, both lysins were able to lyse only Gram-positive bacterial strains that possess PG with d-Ala4d-Asx-l-Lys3 in their cross-bridge, such as Lactococcus casei, Lactococcus lactis, and Enterococcus faecium. By testing a panel of L. lactis cell wall mutants, we observed that Lc-Lys and Lc-Lys-2 were not able to lyse mutants with a modified PG cross-bridge, constituting d-Ala4l-Ala-(l-Ala/l-Ser)-l-Lys3; moreover, they do not lyse the L. lactis mutant containing only the nonamidated d-Asp cross-bridge, i.e. d-Ala4d-Asp-l-Lys3. In contrast, Lc-Lys could lyse the ampicillin-resistant E. faecium mutant with 3→3 l-Lys3-d-Asn-l-Lys3 bridges replacing the wild-type 4→3 d-Ala4-d-Asn-l-Lys3 bridges. We showed that the C-terminal CWBD of Lc-Lys binds PG containing mainly d-Asn but not PG with only the nonamidated d-Asp-containing cross-bridge, indicating that the CWBD confers to Lc-Lys its narrow specificity. In conclusion, the CWBD characterized in this study is a novel type of PG-binding domain targeting specifically the d-Asn interpeptide bridge of PG.  相似文献   

14.
A Mg2+-dependent, alkaline phosphatase has been isolated from mature pollen of Lilium longiflorum Thunb., cv. Ace and partially purified. It hydrolyzes 1l- and 1d-myo-inositol 1-phosphate, myo-inositol 2-phosphate, and β-glycerophosphate at rates decreasing in the order named. The affinity of the enzyme for 1l- and 1d-myo-inositol 1-phosphate is approximately 10-fold greater than its affinity for myo-inositol 2-phosphate. Little or no activity is found with phytate, d-glucose 6-phosphate, d-glucose 1-phosphate, d-fructose 1-phosphate, d-fructose 6-phosphate, d-mannose 6-phosphate, or p-nitrophenyl phosphate. 3-Phosphosphoglycerate is a weak competitive inhibitor. myo-Inositol does not inhibit the reaction. Optimal activity is obtained at pH 8.5 and requires the presence of Mg2+. At 4 millimolar, Co2+, Fe2+ or Mn2+ are less effective. Substantial inhibition is obtained with 0.25 molar Li+. With β-glycerophosphate as substrate the Km is 0.06 millimolar and the reaction remains linear at least 2 hours. In 0.1 molar Tris, β-glycerophosphate yields equivalent amounts of glycerol and inorganic phosphate, evidence that transphosphorylation does not occur.  相似文献   

15.
An l-glucose-utilizing bacterium, Paracoccus sp. 43P, was isolated from soil by enrichment cultivation in a minimal medium containing l-glucose as the sole carbon source. In cell-free extracts from this bacterium, NAD+-dependent l-glucose dehydrogenase was detected as having sole activity toward l-glucose. This enzyme, LgdA, was purified, and the lgdA gene was found to be located in a cluster of putative inositol catabolic genes. LgdA showed similar dehydrogenase activity toward scyllo- and myo-inositols. l-Gluconate dehydrogenase activity was also detected in cell-free extracts, which represents the reaction product of LgdA activity toward l-glucose. Enzyme purification and gene cloning revealed that the corresponding gene resides in a nine-gene cluster, the lgn cluster, which may participate in aldonate incorporation and assimilation. Kinetic and reaction product analysis of each gene product in the cluster indicated that they sequentially metabolize l-gluconate to glycolytic intermediates, d-glyceraldehyde-3-phosphate, and pyruvate through reactions of C-5 epimerization by dehydrogenase/reductase, dehydration, phosphorylation, and aldolase reaction, using a pathway similar to l-galactonate catabolism in Escherichia coli. Gene disruption studies indicated that the identified genes are responsible for l-glucose catabolism.  相似文献   

16.
The biosynthesis of apiose was investigated in cell wall polysaccharide of Lemna gibba G3 (duckweed) and in detached leaves of Petroselinum crispum (parsley). Lemna grown either in short days or in continuous light incorporated 14C from a medium containing myo-inositol-2-14C into d-apiosyl and d-xylosyl units of cell wall polysaccharides. Labeled d-apiose was characterized by paper chromatography, by formation of labeled crystalline di-O-isopropylidene d-apiose, and by gas chromatography of trimethylsilyl derivatives of apiose and of its sodium borohydride reduction product, apiitol. Periodate oxidation of labeled apiose revealed 86 to 94% of the 14C was located in formaldehyde fragments corresponding to C3′ and C4. Comparison of this result with work reported by Grisebach and Doebereiner and by Beck and Kandler supports the conclusion that myo-inositol-2-14C was converted to d-apiose labeled specifically at C4.  相似文献   

17.
Particulate preparations from Phaseolus aureus produce a d-mannosyl-lipid when treated with GDP-d-mannose. This lipid complex appears to be an active d-mannose donor, and some investigators have proposed that its role might be an obligatory intermediate in mannan synthesis of higher plants. When the partially purified d-mannosyl-lipids, isotopically labeled in the d-mannose moiety, were treated with particulate enzymes under a variety of conditions, a negligible amount of material was produced that behaved as a polysaccharide. Endogenous, particle-bound d-mannosyl-14C-lipid prepared from P. aureus particles readily transferred d-mannose to GDP to yield GDP-d-mannose and was hydrolyzed to free d-mannose when treated briefly with 0.01 n HCl at 100 C. The d-mannosyl-lipid, therefore, exhibits active d-mannose transfer potential in its endogenous state. When endogenous glycosyl-lipid was incubated in the absence of GDP-d-mannose-14C, little or no polysaccharide was produced. It was, instead, slowly degraded to d-mannose. Addition of several different unlabeled sugar nucleotides had no effect on the results. Our studies to date, therefore, offer no evidence that the mannosyl-lipid is an obligatory precursor of polysaccharide.  相似文献   

18.
Thermotoga maritima is a Gram-negative, hyperthermophilic bacterium whose peptidoglycan contains comparable amounts of l- and d-lysine. We have determined the fine structure of this cell-wall polymer. The muropeptides resulting from the digestion of peptidoglycan by mutanolysin were separated by high-performance liquid chromatography and identified by amino acid analysis after acid hydrolysis, dinitrophenylation, enzymatic determination of the configuration of the chiral amino acids, and mass spectrometry. The high-performance liquid chromatography profile contained four main peaks, two monomers, and two dimers, plus a few minor peaks corresponding to anhydro forms. The first monomer was the d-lysine-containing disaccharide-tripeptide in which the d-Glu-d-Lys bond had the unusual γ→ϵ arrangement (GlcNAc-MurNAc-l-Ala-γ-d-Glu-ϵ-d-Lys). The second monomer was the conventional disaccharide-tetrapeptide (GlcNAc-MurNAc-l-Ala-γ-d-Glu-l-Lys-d-Ala). The first dimer contained a disaccharide-l-Ala as the acyl donor cross-linked to the α-amine of d-Lys in a tripeptide acceptor stem with the sequence of the first monomer. In the second dimer, donor and acceptor stems with the sequences of the second and first monomers, respectively, were connected by a d-Ala4-α-d-Lys3 cross-link. The cross-linking index was 10 with an average chain length of 30 disaccharide units. The structure of the peptidoglycan of T. maritima revealed for the first time the key role of d-Lys in peptidoglycan synthesis, both as a surrogate of l-Lys or meso-diaminopimelic acid at the third position of peptide stems and in the formation of novel cross-links of the l-Ala1(α→α)d-Lys3 and d-Ala4(α→α)d-Lys3 types.Peptidoglycan (or murein) is a giant macromolecule whose main function is the protection of the cytoplasmic membrane against the internal osmotic pressure. It is composed of alternating residues of N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc)2 cross-linked by short peptides (1). The composition of the peptide stem in nascent peptidoglycan is l-Ala1-γ-d-Glu2-X3-d-Ala4-d-Ala5, where X is most often meso-diaminopimelic acid (meso-A2pm) or l-lysine in Gram-negative and Gram-positive species, respectively (2, 3). In the mature macromolecule, the last d-Ala residue is removed. Cross-linking of the glycan chains generally occurs between the carboxyl group of d-Ala at position 4 of a donor peptide stem and the side-chain amino group of the diamino acid at position 3 of an acceptor peptide stem (4→3 cross-links). Cross-linking is either direct or through a short peptide bridge such as pentaglycine in Staphylococcus aureus (2, 3). The enzymes for the formation of the 4→3 cross-links are active-site serine dd- transpeptidases that belong to the penicillin-binding protein (PBP) family and are the essential targets of β-lactam antibiotics in pathogenic bacteria (4). Catalysis involves the cleavage of the d-Ala4-d-Ala5 bond of a donor peptide stem and the formation of an amide bond between the carboxyl of d-Ala4 and the side chain amine at the third position of an acceptor stem. Transpeptidases of the ld specificity are active-site cysteine enzymes that were shown to act as surrogates of the PBPs in mutants of Enterococcus faecium resistant to β-lactam antibiotics (5). They cleave the X3-d-Ala4 bond of a donor stem peptide to form 3→3 cross-links. This alternate mode of cross-linking is usually marginal, although it has recently been shown to predominate in non-replicative “dormant” forms of Mycobacterium tuberculosis (6).Thermotoga maritima is a Gram-negative, extremely thermophilic bacterium isolated from geothermally heated sea floors by Huber et al. (7). A morphological characteristic is the presence of an outer sheath-like envelope called “toga.” Although the organism has received considerable attention for its biotechnological potential, studies about its peptidoglycan are scarce (811), and in particular the fine structure of the macromolecule is still unknown. In their initial work, Huber et al. (7) showed that the composition of its peptidoglycan was unusual for a Gram-negative species, because it contained both isomers of lysine and no A2pm. Recently, we purified and studied the properties of T. maritima MurE (12); this enzyme is responsible for the addition of the amino acid residue at position 3 of the peptide stem (13, 14). We demonstrated that T. maritima MurE added in vitro l- and d-Lys to UDP-MurNAc-l-Ala-d-Glu. Although l-Lys was added in the usual way, yielding the conventional nucleotide UDP-MurNAc-l-Ala-γ-d-Glu-l-Lys containing a d-Glu(γ→α)l-Lys amide bond, the d-isomer was added in an “upside-down” manner, yielding the novel nucleotide UDP-MurNAc-l-Ala-d-Glu(γ→ϵ)d-Lys. We also showed that the d-Lys-containing nucleotide was not a substrate for T. maritima MurF, the subsequent enzyme in the biosynthetic pathway, whereas this ligase catalyzed the addition of dipeptide d-Ala-d-Ala to the l-Lys-containing tripeptide, yielding the conventional UDP-MurNAc-pentapeptide (12).However, both the l-Lys-containing UDP-MurNAc-pentapeptide and d-Lys-containing UDP-MurNAc-tripeptide were used as substrates by T. maritima MraY with comparable efficiencies in vitro (12). This observation implies that the unusual d-Lys-containing peptide stems are likely to be translocated to the periplasmic face of the cytoplasmic membrane and to participate in peptidoglycan polymerization. Therefore, we have determined here the fine structure of T. maritima peptidoglycan and we have shown that l-Lys- and d-Lys-containing peptide stems are both present in the polymer, the latter being involved in the formation of two novel types of peptidoglycan cross-link.  相似文献   

19.
d-Lactate accumulation in Chlamydomonas reinhardtii was dependent on anaerobic conditions. As much as 50% of the 14C after 2 minutes of photosynthetic 14CO2 fixation moved into d-lactate from sugar phosphates if the cells became anaerobic for short time periods. No lactate accumulated in the dark until the O2 concentration decreased to less than 0.1%. Lactate was determined to be of the d-configuration using stereospecific lactate dehydrogenases. d-Lactate produced anaerobically by algae grown on 5% CO2 was only slowly metabolized aerobically in the light or dark, and in the dark, only a trace of the lactate was excreted.  相似文献   

20.
Reduction of FeIIIEDTA by excised roots of soybean seedlings (Glycine max L.) is stimulated by l-malate in the bathing solution. Reduction occurs much more rapidly with roots of seedlings grown in the absence of iron than with roots of seedlings grown with iron. Cell-wall preparations from these roots catalyze reduction of FeIIIEDTA by NADH. They also contain NAD+-dependent l-malate dehydrogenase. Enzymic activity of the cell-wall preparations is not affected by previous iron nutrition of the plants, but the amount of l-malate in the roots is increased when seedlings have been deprived of iron. We propose that reduction of iron before absorption by soybean roots occurs in the cell-wall space, with l-malate secreted from the roots serving as the source of electrons. Part of the iron reductase activity of the cell walls can be solubilized by extraction with 1 molar NaCl. The enzyme has been partially purified.  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