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 531 毫秒
1.
2.
Invasion of the intestinal epithelium is a critical step in Salmonella enterica infection and requires functions encoded in the gene cluster known as Salmonella Pathogenicity Island 1 (SPI-1). Expression of SPI-1 genes is repressed by l-arabinose, and not by other pentoses. Transport of l-arabinose is necessary to repress SPI-1; however, repression is independent of l-arabinose metabolism and of the l-arabinose-responsive regulator AraC. SPI-1 repression by l-arabinose is exerted at a single target, HilD, and the mechanism appears to be post-translational. As a consequence of SPI-1 repression, l-arabinose reduces translocation of SPI-1 effectors to epithelial cells and decreases Salmonella invasion in vitro. These observations reveal a hitherto unknown role of l-arabinose in gene expression control and raise the possibility that Salmonella may use L-arabinose as an environmental signal.  相似文献   

3.
4.
The first enzyme in the pathway for l-arabinose catabolism in eukaryotic microorganisms is a reductase, reducing l-arabinose to l-arabitol. The enzymes catalyzing this reduction are in general nonspecific and would also reduce d-xylose to xylitol, the first step in eukaryotic d-xylose catabolism. It is not clear whether microorganisms use different enzymes depending on the carbon source. Here we show that Aspergillus niger makes use of two different enzymes. We identified, cloned, and characterized an l-arabinose reductase, larA, that is different from the d-xylose reductase, xyrA. The larA is up-regulated on l-arabinose, while the xyrA is up-regulated on d-xylose. There is however an initial up-regulation of larA also on d-xylose but that fades away after about 4 h. The deletion of the larA gene in A. niger results in a slow growth phenotype on l-arabinose, whereas the growth on d-xylose is unaffected. The l-arabinose reductase can convert l-arabinose and d-xylose to their corresponding sugar alcohols but has a higher affinity for l-arabinose. The Km for l-arabinose is 54 ± 6 mm and for d-xylose 155 ± 15 mm.  相似文献   

5.
Two enzymes, l-arabinose isomerase and mannose-6-phosphate isomerase, from Geobacillus thermodenitrificans produced 118 g/liter l-ribose from 500 g/liter l-arabinose at pH 7.0, 70°C, and 1 mM Co2+ for 3 h, with a conversion yield of 23.6% and a volumetric productivity of 39.3 g liter−1 h−1.l-Ribose, a potential starting material for the synthesis of many l-nucleoside-based pharmaceutical compounds, is not abundant in nature (4, 15, 20). l-Ribose has been synthesized primarily from l-arabinose, l-xylose, d-glucose, d-galactose, d-ribose, and d-mannono-1,4-lactone (1, 13, 20). Recombinant cells containing a NAD-dependent mannitol-1-dehydrogenase produced 52 g/liter l-ribose from 100 g/liter ribitol after fermentation for 72 h (14). However, the volumetric productivity of l-ribose was 26-fold lower than that of the chemical synthetic method starting from l-arabinose (6). l-Ribose isomerase from an Acinetobacter sp., which is most active with l-ribose, showed poor efficiency in the conversion of l-ribulose to l-ribose (9). Recently, l-ribulose was produced with a conversion yield of 19% from the inexpensive sugar l-arabinose using l-arabinose isomerase (AI) from Geobacillus thermodenitrificans (18). l-Ribose has been produced from l-ribulose using mannose-6-phosphate isomerase (MPI) from Bacillus subtilis with a conversion yield of 70% (17). In this study, the production of l-ribose from l-arabinose was demonstrated via a two-enzyme system from G. thermodenitrificans, in which l-ribulose was first produced from l-arabinose by AI and subsequently converted to l-ribose by MPI.The analysis of monosaccharides and the purification and thermostability of AI and MPI from G. thermodenitrificans (2) isolated from compost were performed as described previously (7, 18, 19). The cross-linked enzymes were obtained from the treatment of 0.5% glutaraldehyde (10, 16). The reaction was performed by replacing the reaction solution with 100 g/liter l-arabinose and 1 mM Co2+ every 6 h at 70°C and pH 7.0. The reaction volume of 10 ml contained 5 g of the cross-linked enzymes with 8 U/ml AI and 20 U/ml MPI. One unit of AI or MPI activity, which corresponded to 0.0625 or 2.5 mg protein, respectively, was defined as the amount of enzyme required to produce 1 μmol of l-ribulose or l-ribose, respectively, per min at 70°C, pH 7.0, and 1 mM Co2+. Unless otherwise stated, the reaction was carried out in 50 mM piperazine-N,N′-bis(2-ethanesulfonic acid) (PIPES) buffer (pH 7.0) in the presence of 1 mM Co2+ at 70°C for 4 h. All experiments were performed in triplicate.The recombinant Escherichia coli ER2566 (New England Biolabs, Ipswich, MA) containing pTrc99A plasmid (Pharmacia Biotech, Piscataway, NJ) and the AI or MPI gene was cultivated in a 7-liter fermentor containing 3 liters of chemically defined medium (11). When the cell mass reached 2 g/liter, 10 g/liter lactose was added for enzyme induction. After 14 h, 40 g/liter cells with 13,400 U/liter of AI or 34 g/liter cells with 630 U/liter of MPI was obtained. The enzyme was purified by heat treatment and Hi-Trap anion-exchange chromatography. The purification yields of AI and MPI were 21 and 78%, respectively, and the levels of purity for the concentrated AI and MPI by gene scanning were 48 and 92%, respectively. Maximum l-ribose production from l-arabinose by AI and by MPI in 10 ml of total volume was observed at pH 7.0, 70°C, and 1 mM Co2+ (data not shown). Half-lives for the two-enzyme system containing 10 mM l-arabinose, 0.2 U/ml AI, and 0.5 U/ml MPI at 60, 65, 70, 75, and 80°C were 1,216, 235, 48, 26, and 12 h, respectively. The use of Co2+ may be disadvantageous, as it is fairly toxic. This problem can be solved by using Mn2+ instead of Co2+. When Mn2+ was used in the reaction with the same amounts of enzymes, the conversion yield was the same as that obtained with Co2+, even though the volumetric productivity was lower than that with Co2+ (data not shown).The effect of the ratio of AI to MPI in the two-step enzymatic production of l-ribose from l-arabinose was investigated by mixing the enzyme solutions (8 U/ml AI and 20 U/ml MPI) to obtain AI/MPI ratios ranging from 10:90 to 90:10 (vol/vol) (Fig. (Fig.1).1). The reactions were run with 300 g/liter l-arabinose. Maximum l-ribose production was observed at a volume ratio of 50:50 of the enzyme solutions. The effects of enzyme concentration on l-ribose production were investigated at the optimal unit ratio (AI/MPI ratio, 1:2.5) with 500 g/liter l-arabinose and AI and MPI concentrations from 0.4 and 1.0 U/ml, respectively, to 9.2 and 23.0 U/ml, respectively (Fig. (Fig.2A).2A). l-Ribose production increased with increasing amounts of enzymes until reaching a plateau at 8 U/ml AI and 20 U/ml MPI. The effect of substrate concentration on l-ribose production was evaluated at l-arabinose concentrations ranging from 15 to 500 g/liter with 8 U/ml AI and 20 U/ml MPI (Fig. (Fig.2B).2B). The production of both l-ribose and l-ribulose, an intermediate, increased with increasing substrate level. The results suggest that concentrations of substrate above 500 g/liter l-arabinose might cause the increased production. The conversion yields of l-ribose and l-ribulose from l-arabinose were constant at 32% and 14%, respectively, within an initial concentration of 100 g/liter l-arabinose, indicating that the reactions reached equilibrium at an l-arabinose/l-ribulose/l-ribose ratio of 54:14:32, which was in agreement with the calculated equilibrium (17). However, at l-arabinose concentrations above 100 g/liter, the conversion yields of l-ribose and l-ribulose from l-arabinose decreased with increasing l-arabinose concentration. The l-arabinose/l-ribulose/l-ribose ratio, with an initial l-arabinose concentration of 300 g/liter, was 71:6:23 after 4 h of reaction. To obtain near-equilibrium (54:14:32) at this high concentration of l-arabinose, more effective enzymes are required.Open in a separate windowFIG. 1.Effect of the ratio of AI to MPI on l-ribose production from l-arabinose by the purified AI and MPI from G. thermodenitrificans. Data are the means for three separate experiments, and error bars represent standard deviations. Symbols: •, l-ribose; ▪, l-ribulose.Open in a separate windowFIG. 2.(A) Effect of enzyme concentration on l-ribose production from l-arabinose at the optimal unit ratio (AI/MPI ratio, 1:2.5). Symbols: •, l-ribose; ▪, l-ribulose; ○, l-arabinose. (B) Effect of l-arabinose concentration on l-ribose production. Symbols: •, l-ribose; ▪, l-ribulose. Data are the means for three separate experiments, and error bars represent standard deviations.A time course reaction of l-ribose production from l-arabinose was monitored for 3 h with 8 U/ml AI and 20 U/ml MPI (Fig. (Fig.3).3). As a result, 118 g/liter l-ribose was obtained from an initial l-arabinose concentration of 500 g/liter after 3 h, with a conversion yield of 23.6% and a productivity of 39.3 g liter−1 h−1. Recombinant E. coli containing MDH yielded 52 g/liter l-ribose from an initial ribitol concentration of 100 g/liter after 72 h, with a productivity of 0.72 g liter−1 h−1 (14). The production and productivity obtained in the current study using AI and MPI from G. thermodenitrificans were 2.3- and 55-fold higher, respectively, than those obtained from ribitol and 17- and 21-fold higher than those obtained with the production of l-ribose from l-arabinose using resting cells of recombinant Lactobacillus plantarum (5). The chemical synthetic method is capable of producing 56.5 g/liter l-ribose from 250 g/liter l-arabinose after 3 h, corresponding to a productivity of 18.8 g liter−1 h−1 (6). Still, both the production and productivity of l-ribose using the method described herein were 2.1-fold higher. Thus, the method of production of l-ribose in the present study exhibited the highest productivity and production, compared to other fermentation methods and chemical syntheses.Open in a separate windowFIG. 3.Time course of l-ribose production from l-arabinose by purified AI and MPI from G. thermodenitrificans. Data are the means for three separate experiments, and error bars represent standard deviations. Symbols: •, l-ribose; ▪, l-ribulose; ○, l-arabinose.Several rounds of conversion reusing the cross-linked enzymes were performed (Fig. (Fig.4).4). The immobilized enzymes showed more than 20% conversion of l-ribose from l-arabinose for the 9th batch, and the concentration of l-ribose was reduced to 43% after the 20th batch. These results suggest that the immobilization of enzyme facilitates separation of product and enzyme, and it enables the enzyme to function continuously, as reported previously (3, 8, 12). Thus, the reuse of enzyme by immobilization improves the economic viability of this enzymatic process.Open in a separate windowFIG. 4.Reuse of immobilized AI and MPI from G. thermodenitrificans for l-ribose production from 100 g/liter l-arabinose. Data are the means for three separate experiments, and error bars represent standard deviations.  相似文献   

6.
Functional expression in heterologous hosts is often less successful for integral membrane proteins than for soluble proteins. Here, two Ambrosiozyma monospora transporters were successfully expressed in Saccharomyces cerevisiae as tagged proteins. Growth of A. monospora on l-arabinose instead of glucose caused transport activities of l-arabinose, l-arabitol, and ribitol, measured using l-[1-3H]arabinose, l-[14C]arabitol, and [14C]ribitol of demonstrated purity. A. monospora LAT1 and LAT2 genes were cloned earlier by using their ability to improve the growth of genetically engineered Saccharomyces cerevisiae on l-arabinose. However, the l-arabinose and pentitol transport activities of S. cerevisiae carrying LAT1 or LAT2 are only slightly greater than those of control strains. S. cerevisiae carrying the LAT1 or LAT2 gene fused in frame to the genes for green fluorescent protein (GFP) or red fluorescent protein (mCherry) or adenylate kinase (AK) exhibited large (>3-fold for LAT1; >20-fold for LAT2) increases in transport activities. Lat1-mCherry transported l-arabinose with high affinity (Km ≈ 0.03 mM) and l-arabitol and ribitol with very low affinity (Km ≥ 75 mM). The Lat2-GFP, Lat2-mCherry, and Lat2-AK fusion proteins could not transport l-arabinose but were high-affinity pentitol transporters (Kms ≈ 0.2 mM). The l-arabinose and pentitol transport activities of A. monospora could not be completely explained by any combination of the observed properties of tagged Lat1 and Lat2, suggesting either that tagging and expression in a foreign membrane alters the transport kinetics of Lat1 and/or Lat2 or that A. monospora contains at least one more l-arabinose transporter.  相似文献   

7.
8.
Limited uptake is one of the bottlenecks for l-arabinose fermentation from lignocellulosic hydrolysates in engineered Saccharomyces cerevisiae. This study characterized two novel l-arabinose transporters, LAT-1 from Neurospora crassa and MtLAT-1 from Myceliophthora thermophila. Although the two proteins share high identity (about 83%), they display different substrate specificities. Sugar transport assays using the S. cerevisiae strain EBY.VW4000 indicated that LAT-1 accepts a broad substrate spectrum. In contrast, MtLAT-1 appeared much more specific for l-arabinose. Determination of the kinetic properties of both transporters revealed that the Km values of LAT-1 and MtLAT-1 for l-arabinose were 58.12 ± 4.06 mM and 29.39 ± 3.60 mM, respectively, with corresponding Vmax values of 116.7 ± 3.0 mmol/h/g dry cell weight (DCW) and 10.29 ± 0.35 mmol/h/g DCW, respectively. In addition, both transporters were found to use a proton-coupled symport mechanism and showed only partial inhibition by d-glucose during l-arabinose uptake. Moreover, LAT-1 and MtLAT-1 were expressed in the S. cerevisiae strain BSW2AP containing an l-arabinose metabolic pathway. Both recombinant strains exhibited much faster l-arabinose utilization, greater biomass accumulation, and higher ethanol production than the control strain. In conclusion, because of higher maximum velocities and reduced inhibition by d-glucose, the genes for the two characterized transporters are promising targets for improved l-arabinose utilization and fermentation in S. cerevisiae.  相似文献   

9.
10.
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   

11.
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   

12.
Bacillus licheniformis l-arabinose isomerase (l-AI) is distinguished from other l-AIs by its high degree of substrate specificity for l-arabinose and its high turnover rate. A systematic strategy that included a sequence alignment-based first screening of residues and a homology model-based second screening, followed by site-directed mutagenesis to alter individual screened residues, was used to study the molecular determinants for the catalytic efficiency of B. licheniformis l-AI. One conserved amino acid, Y333, in the substrate binding pocket of the wild-type B. licheniformis l-AI was identified as an important residue affecting the catalytic efficiency of B. licheniformis l-AI. Further insights into the function of residue Y333 were obtained by replacing it with other aromatic, nonpolar hydrophobic amino acids or polar amino acids. Replacing Y333 with the aromatic amino acid Phe did not alter catalytic efficiency toward l-arabinose. In contrast, the activities of mutants containing a hydrophobic amino acid (Ala, Val, or Leu) at position 333 decreased as the size of the hydrophobic side chain of the amino acid decreased. However, mutants containing hydrophilic and charged amino acids, such as Asp, Glu, and Lys, showed almost no activity with l-arabinose. These data and a molecular dynamics simulation suggest that Y333 is involved in the catalytic efficiency of B. licheniformis l-AI.l-Arabinose isomerase (l-AI) is an enzyme that mediates in vivo isomerization between l-arabinose and l-ribulose as well as in vitro isomerization of d-galactose and d-tagatose (20). l-Ribulose (l-erythro-pentulose) is a rare and expensive ketopentose sugar (1) that can be used as a precursor for the production of other rare sugars of high market value, such as l-ribose. Despite being a common metabolic intermediate in different organisms, l-ribulose is scarce in nature. The market for rare and unnatural sugars has been growing, especially in the sweetener and pharmaceutical industries. For example, several modified nucleosides derived from l-sugars have been shown to act as potent antiviral agents and are also useful in antigen therapy. Derivatives of rare sugars have also been used as agents against hepatitis B virus and human immunodeficiency virus (2, 22).For these reasons, interest in the enzymology of rare sugars has also been increasing. Various forms of l-AI from a variety of organisms have been characterized, and some have shown potential for industrial use. Several highly thermotolerant enzyme forms from Thermotoga maritima (12), Thermotoga neapolitana (10), Bacillus stearothermophilus (18), Thermoanaerobacter mathranii (9), and Lactobacillus plantarum (5) have been characterized previously. All of these reported l-AIs tend to have broad specificity, although a few l-AIs with high degrees of substrate specificity for l-arabinose have also been documented.The enzyme properties of l-AIs have been examined by engineering several forms by error-prone PCR and site-directed mutagenesis. Galactose conversion was reportedly enhanced 20% following site-directed introduction of a double mutation (C450S-N475K) into l-AI (16). Error-prone PCR manipulation of l-AI from Geobacillus stearothermophilus resulted in a shift in temperature specificity from 60 to 65°C and increased isomerization activity (11). All of these previously reported mutational studies have been aimed at improving enzymatic properties for industrial application. However, even though the three-dimensional (3D) structure of Escherichia coli l-AI has been determined previously (15), few new structural studies have been performed to decipher the reaction mechanism of this enzyme. Rhimi et al. (19) have reported an important role for D308, F329, E351, and H446 in catalysis, as indicated by findings from site-directed mutagenesis. Nonetheless, detailed analysis of the important molecular determinants controlling the catalytic activities of the l-AIs is still lacking.Previously, we have reported the cloning and characterization of a novel l-AI from Bacillus licheniformis (17). This enzyme can be distinguished from other l-AIs by its wide pH range, high degree of substrate specificity for l-arabinose, and extremely high turnover rate. In the present paper, we report the identification of an important amino acid residue responsible for the catalytic efficiency of l-AIs, as determined by a systematic screening process composed of sequence alignment and molecular dynamics (MD) simulation, followed by site-directed mutagenesis. Using the crystal structure of E. coli l-AI as a template, we have built a 3D model of B. licheniformis l-AI. Analysis of the 3D model of B. licheniformis l-AI docked with l-arabinose, followed by a systematic screening process, showed that Y333 interacted with the substrate, suggesting that this residue in B. licheniformis l-AI may be essential for catalysis. We further characterized the role of Y333 in B. licheniformis l-AI binding of and catalytic efficiency for l-arabinose.  相似文献   

13.
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   

14.
Xylan-debranching enzymes facilitate the complete hydrolysis of xylan and can be used to alter xylan chemistry. Here, the family GH62 α-l-arabinofuranosidase from Streptomyces thermoviolaceus (SthAbf62A) was shown to have a half-life of 60 min at 60°C and the ability to cleave α-1,3 l-arabinofuranose (l-Araf) from singly substituted xylopyranosyl (Xylp) backbone residues in wheat arabinoxylan; low levels of activity on arabinan as well as 4-nitrophenyl α-l-arabinofuranoside were also detected. After selective removal of α-1,3 l-Araf substituents from disubstituted Xylp residues present in wheat arabinoxylan, SthAbf62A could also cleave the remaining α-1,2 l-Araf substituents, confirming the ability of SthAbf62A to remove α-l-Araf residues that are (1→2) and (1→3) linked to monosubstituted β-d-Xylp sugars. Three-dimensional structures of SthAbf62A and its complex with xylotetraose and l-arabinose confirmed a five-bladed β-propeller fold and revealed a molecular Velcro in blade V between the β1 and β21 strands, a disulfide bond between Cys27 and Cys297, and a calcium ion coordinated in the central channel of the fold. The enzyme-arabinose complex structure further revealed a narrow and seemingly rigid l-arabinose binding pocket situated at the center of one side of the β propeller, which stabilized the arabinofuranosyl substituent through several hydrogen-bonding and hydrophobic interactions. The predicted catalytic amino acids were oriented toward this binding pocket, and the catalytic essentiality of Asp53 and Glu213 was confirmed by site-specific mutagenesis. Complex structures with xylotetraose revealed a shallow cleft for xylan backbone binding that is open at both ends and comprises multiple binding subsites above and flanking the l-arabinose binding pocket.  相似文献   

15.
16.
17.
α-l-Arabinofuranosidases I and II were purified from the culture filtrate of Aspergillus awamori IFO 4033 and had molecular weights of 81,000 and 62,000 and pIs of 3.3 and 3.6, respectively. Both enzymes had an optimum pH of 4.0 and an optimum temperature of 60°C and exhibited stability at pH values from 3 to 7 and at temperatures up to 60°C. The enzymes released arabinose from p-nitrophenyl-α-l-arabinofuranoside, O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose, and arabinose-containing polysaccharides but not from O-β-d-xylopyranosyl-(1→2)-O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. α-l-Arabinofuranosidase I also released arabinose from O-β-d-xylopy-ranosyl-(1→4)-[O-α-l-arabinofuranosyl-(1→3)]-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. However, α-l-arabinofuranosidase II did not readily catalyze this hydrolysis reaction. α-l-Arabinofuranosidase I hydrolyzed all linkages that can occur between two α-l-arabinofuranosyl residues in the following order: (1→5) linkage > (1→3) linkage > (1→2) linkage. α-l-Arabinofuranosidase II hydrolyzed the linkages in the following order: (1→5) linkage > (1→2) linkage > (1→3) linkage. α-l-Arabinofuranosidase I preferentially hydrolyzed the (1→5) linkage of branched arabinotrisaccharide. On the other hand, α-l-arabinofuranosidase II preferentially hydrolyzed the (1→3) linkage in the same substrate. α-l-Arabinofuranosidase I released arabinose from the nonreducing terminus of arabinan, whereas α-l-arabinofuranosidase II preferentially hydrolyzed the arabinosyl side chain linkage of arabinan.Recently, it has been proven that l-arabinose selectively inhibits intestinal sucrase in a noncompetitive manner and reduces the glycemic response after sucrose ingestion in animals (33). Based on this observation, l-arabinose can be used as a physiologically functional sugar that inhibits sucrose digestion. Effective l-arabinose production is therefore important in the food industry. l-Arabinosyl residues are widely distributed in hemicelluloses, such as arabinan, arabinoxylan, gum arabic, and arabinogalactan, and the α-l-arabinofuranosidases (α-l-AFases) (EC 3.2.1.55) have proven to be essential tools for enzymatic degradation of hemicelluloses and structural studies of these compounds.α-l-AFases have been classified into two families of glycanases (families 51 and 54) on the basis of amino acid sequence similarities (11). The two families of α-l-AFases also differ in substrate specificity for arabinose-containing polysaccharides. Beldman et al. summarized the α-l-AFase classification based on substrate specificities (3). One group contains the Arafur A (family 51) enzymes, which exhibit very little or no activity with arabinose-containing polysaccharides. The other group contains the Arafur B (family 54) enzymes, which cleave arabinosyl side chains from polymers. However, this classification is too broad to define the substrate specificities of α-l-AFases. There have been many studies of the α-l-AFases (3, 12), especially the α-l-AFases of Aspergillus species (28, 1215, 17, 22, 23, 2832, 3639, 4143, 46). However, there have been only a few studies of the precise specificities of these α-l-AFases. In previous work, we elucidated the substrate specificities of α-l-AFases from Aspergillus niger 5-16 (17) and Bacillus subtilis 3-6 (16, 18), which should be classified in the Arafur A group and exhibit activity with arabinoxylooligosaccharides, synthetic methyl 2-O-, 3-O-, and 5-O-arabinofuranosyl-α-l-arabinofuranosides (arabinofuranobiosides) (20), and methyl 3,5-di-O-α-l-arabinofuranosyl-α-l-arabinofuranoside (arabinofuranotrioside) (19).In the present work, we purified two α-l-AFases from a culture filtrate of Aspergillus awamori IFO 4033 and determined the substrate specificities of these α-l-AFases by using arabinose-containing polysaccharides and the core oligosaccharides of arabinoxylan and arabinan.  相似文献   

18.
19.
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   

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