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.
2.
The NADPH-dependent L-sorbose reductase (SR) of L-sorbose-producing Gluconobacter suboxydans IFO 3291 contributes to intracellular L-sorbose assimilation. The gene disruptant showed no SR activity and did not assimilate the once-produced L-sorbose, indicating that the SR functions mainly as an L-sorbose-reducing enzyme in vivo and not as a D-sorbitol-oxidizing enzyme.  相似文献   

3.
The optimization of L-sorbose synthesis by regiospecific dehydrogenation of D-sorbitol using Gluconobacter oxydans is reported. The current L-sorbose production processes that are based on G. oxydans and other bacterial strains are suboptimal as to yield and rate of L-sorbose synthesis. One reason for these problems is the toxicity that is induced by the substrate D-sorbitol when used in concentrations of >10% (w/v). This phenomenon significantly limits the potentials of L-sorbose production from an industrial point of view. The goal of this study was to develop a fast production process that yields L-sorbose in stoichiometric amounts starting from D-sorbitol concentrations that exceed 10% (w/v). A gradual improvement of the inoculum build-up procedure, culture medium composition, and process parameters ultimately led to a theoretically maximal L-sorbose productivity (200 g L(-1) of L-sorbose from 200 g L(-1) of D-sorbitol in 28 h of fermentation) using a Gluconobacter oxydans mutant strain that was selected under conditions of substrate inhibition. Because the D-sorbitol/L‐sorbose bioconversion is used to mass-produce vitamin C, the procedure reported here will contribute to a more efficient and more economic synthesis of vitamin C.  相似文献   

4.
Two different membrane-bound enzymes oxidizing D-sorbitol are found in Gluconobacter frateurii THD32: pyroloquinoline quinone-dependent glycerol dehydrogenase (PQQ-GLDH) and FAD-dependent D-sorbitol dehydrogenase (FAD-SLDH). In this study, FAD-SLDH appeared to be induced by L-sorbose. A mutant defective in both enzymes grew as well as the wild-type strain did, indicating that both enzymes are dispensable for growth on D-sorbitol. The strain defective in PQQ-GLDH exhibited delayed L-sorbose production, and lower accumulation of it, corresponding to decreased oxidase activity for D-sorbitol in spite of high D-sorbitol dehydrogenase activity, was observed. In the mutant strain defective in PQQ-GLDH, oxidase activity with D-sorbitol was much more resistant to cyanide, and the H(+)/O ratio was lower than in either the wild-type strain or the mutant strain defective in FAD-SLDH. These results suggest that PQQ-GLDH connects efficiently to cytochrome bo(3) terminal oxidase and that it plays a major role in L-sorbose production. On the other hand, FAD-SLDH linked preferably to the cyanide-insensitive terminal oxidase, CIO.  相似文献   

5.
There are two types of membrane-bound D-sorbitol dehydrogenase (SLDH) reported: PQQ-SLDH, having pyrroloquinoline quinone (PQQ), and FAD-SLDH, containing FAD and heme c as the prosthetic groups. FAD-SLDH was purified and characterized from the PQQ-SLDH mutant strain of a thermotolerant Gluconobacter frateurii, having molecular mass of 61.5 kDa, 52 kDa, and 22 kDa. The enzyme properties were quite similar to those of the enzyme from mesophilic G. oxydans IFO 3254. This enzyme was shown to be inducible by D-sorbitol, but not PQQ-SLDH. The oxidation product of FAD-SLDH from D-sorbitol was identified as L-sorbose. The cloned gene of FAD-SLDH had three open reading frames (sldSLC) corresponding to the small, the large, and cytochrome c subunits of FAD-SLDH respectively. The deduced amino acid sequences showed high identity to those from G. oxydans IFO 3254: SldL showed to other FAD-enzymes, and SldC having three heme c binding motives to cytochrome c subunits of other membrane-bound dehydrogenases.  相似文献   

6.
Gluconobacter strains effectively produce L-sorbose from D-sorbitol because of strong activity of the D-sorbitol dehydrogenase (SLDH). L-sorbose is one of the important intermediates in the industrial vitamin C production process. Two kinds of membrane-bound SLDHs, which consist of three subunits, were reportedly found in Gluconobacter strains [Agric. Biol. Chem. 46 (1982) 135,FEMS Microbiol. Lett. 125 (1995) 45]. We purified a one-subunit-type SLDH (80 kDa) from the membrane fraction of Gluconobacter suboxydans IFO 3255 solubilized with Triton X-100 in the presence of D-sorbitol, but the cofactor could not be identified from the purified enzyme. The SLDH was active on mannitol, glycerol and other sugar alcohols as well as on D-sorbitol to produce respective keto-aldoses. Then, the SLDH gene (sldA) was cloned and sequenced. It encodes the polypeptide of 740 residues, which contains a signal sequence of 24 residues. SLDH had 35-37% identity to those of membrane-bound quinoprotein glucose dehydrogenases (GDHs) from Escherichia coli, Gluconobacter oxydans and Acinetobacter calcoaceticus except the N-terminal hydrophobic region of GDH. Additionally, the sldB gene located just upstream of sldA was found to encode the polypeptide consisting of 126 very hydrophobic residues that is similar to the one-sixth N-terminal region of the GDH. Development of the SLDH activity in E. coli required co-expression of the sldA and sldB genes and the presence of PQQ. The sldA gene disruptant showed undetectable oxidation activities on D-sorbitol in growing culture, and resting-cell reaction (pH 4.5 and 7); in addition, they showed undetectable activities on D-mannitol and glycerol. The disruption of the sldB gene by a gene cassette with a downward promoter to express the sldA gene resulted in formation of a larger size of the SLDH protein and in undetectable oxidation of the polyols. In conclusion, the SLDH of the strain 3255 functions as the main polyol dehydrogenase in vivo. The sldB polypeptide possibly has a chaperone-like function to process the SLDH polypeptide into a mature and active form.  相似文献   

7.
The D-sorbitol dehydrogenase gene, sldA, and an upstream gene, sldB, encoding a hydrophobic polypeptide, SldB, of Gluconobacter suboxydans IFO 3255 were disrupted in a check of their biological functions. The bacterial cells with the sldA gene disrupted did not produce L-sorbose by oxidation of D-sorbitol in resting-cell reactions at pHs 4.5 and 7.0, indicating that the dehydrogenase was the main D-sorbitol-oxidizing enzyme in this bacterium. The cells did not produce D-fructose from D-mannitol or dihydroxyacetone from glycerol. The disruption of the sldB gene resulted in undetectable oxidation of D-sorbitol, D-mannitol, or glycerol, although the cells produced the dehydrogenase. The cells with the sldB gene disrupted produced more of what might be signal-unprocessed SldA than the wild-type cells did. SldB may be a chaperone-like component that assists signal processing and folding of the SldA polypeptide to form active D-sorbitol dehydrogenase.  相似文献   

8.
Thermotolerant acetic acid bacteria belonging to the genus Gluconobacter were isolated from various kinds of fruits and flowers from Thailand and Japan. The screening strategy was built up to exclude Acetobacter strains by adding gluconic acid to a culture medium in the presence of 1% D-sorbitol or 1% D-mannitol. Eight strains of thermotolerant Gluconobacter were isolated and screened for D-fructose and L-sorbose production. They grew at wide range of temperatures from 10 degrees C to 37 degrees C and had average optimum growth temperature between 30-33 degrees C. All strains were able to produce L-sorbose and D-fructose at higher temperatures such as 37 degrees C. The 16S rRNA sequences analysis showed that the isolated strains were almost identical to G. frateurii with scores of 99.36-99.79%. Among these eight strains, especially strains CHM16 and CHM54 had high oxidase activity for D-mannitol and D-sorbitol, converting it to D-fructose and L-sorbose at 37 degrees C, respectively. Sugar alcohols oxidation proceeded without a lag time, but Gluconobacter frateurii IFO 3264T was unable to do such fermentation at 37 degrees C. Fermentation efficiency and fermentation rate of the strains CHM16 and CHM54 were quite high and they rapidly oxidized D-mannitol and D-sorbitol to D-fructose and L-sorbose at almost 100% within 24 h at 30 degrees C. Even oxidative fermentation of D-fructose done at 37 degrees C, the strain CHM16 still accumulated D-fructose at 80% within 24 h. The efficiency of L-sorbose fermentation by the strain CHM54 at 37 degrees C was superior to that observed at 30 degrees C. Thus, the eight strains were finally classified as thermotolerant members of G. frateurii.  相似文献   

9.
Cloning and expression of the gene encoding Acetobacter liquefaciens IFO 12258 membrane-bound L-sorbosone dehydrogenase (SNDH) were studied. A genomic library of A. liquefaciens IFO 12258 was constructed with the mobilizable cosmid vector pVK102 (mob+) in Escherichia coli S17-1 (Tra+). The library was transferred by conjugal mating into Gluconobacter oxydans OX4, a mutant of G. oxydans IFO 3293 that accumulates L-sorbosone in the presence of L-sorbose. The transconjugants were screened for SNDH activity by performing a direct expression assay. One clone harboring plasmid p7A6 converted L-sorbosone to 2-keto-L-gulonic acid (2KGA) more rapidly than its host did and also converted L-sorbose to 2KGA with no accumulation of L-sorbosone. The insert (25 kb) of p7A6 was shortened to a 3.1-kb fragment, in which one open reading frame (1,347 bp) was found and was shown to encode a polypeptide with a molecular weight of 48,222. The SNDH gene was introduced into the 2KGA-producing strain G. oxydans IFO 3293 and its derivatives, which contained membrane-bound L-sorbose dehydrogenase. The cloned SNDH was correctly located in the membrane of the host. The membrane fraction of the clone exhibited almost stoichiometric formation of 2KGA from L-sorbosone and L-sorbose. Resting cells of the clones produced 2KGA very efficiently from L-sorbosone and L-sorbose, but not from D-sorbitol; the conversion yield from L-sorbosone was improved from approximately 25 to 83%, whereas the yield from L-sorbose was increased from 68 to 81%. Under fermentation conditions, cloning did not obviously improve the yield of 2KGA from L-sorbose.  相似文献   

10.
11.
Summary The effect of permeabilization of G. suboxydans cells with toluene on the oxidation of D-sorbitol to L-sorbose was investigated. Treatment of the cells with 10% toluene resulted in a three fold increase in the specific sorbitol dehydrogenase activity and a two fold increase in the efficiency of D-sorbitol conversion to L-sorbose of the free cell suspension. When the permeabilized cells were immobilized in calcium alginate, the operational stability during air-lift reactor operation was also found to increase with up to three times longer half-life(44 days) of catalytic activity compared with immobilized intact cells.  相似文献   

12.
We have purified L-sorbose dehydrogenase (SDH) and L-sorbosone dehydrogenase (SNDH) from Gluconobacter oxydans T-100 that showed an ability to convert D-sorbitol to 2-keto-L-gulonate (2-KLGA). A genomic library of Gluconobacter oxydans T-100 was screened with a probe, a 180-bp PCR product which was obtained from degenerate oligodeoxyribonucleotides based on the elucidated sequence of the purified SDH (used as primers) and the genomic DNA of G. oxydans T-100 (used as a template). From sequencing of the DNA from a clone positive to the probe, the SNDH and the SDH were estimated to be coded in sequential open reading frames with 1,497 and 1,599 nucleotides, respectively, which was confirmed by expression of the DNA in Escherichia coli that showed both enzymatic activities. The DNA was introduced to a shuttle vector which was prepared from a plasmid of G. oxydans T-100 and pHSG298 to obtain an expression vector designated pSDH155. The production of 2-KLGA by pSDH155 in G. oxydans G624, an L-sorbose-accumulating strain, was improved to 230% compared to that of G. oxydans T-100. Chemical mutation of the host strain to suppress the L-idonate pathway and replacement of the original promoter with that of E. coli tufB resulted in improving the production of 2-KLGA. Consequently, high-level production from D-sorbitol to 2-KLGA (130 mg/ml) was achieved by simple fermentation of the recombinant Gluconobacter.  相似文献   

13.
Gluconobacter melanogenus IFO 3293 cells capable of converting L-sorbose to L-sorbosone were immobilized in polyacrylamide gel. The preferred polymer composition for high activity and stability was determined to contain a total monomer concentration of 7.2% and 16.6% crosslinking agent. No significant differences in optimal conditions for conversion, e.g., pH and temperature, were found in comparison with free cell suspensions. However, in the absence of L-sorbose, the thermal stability of immobilized cells was lower. After the initial loss, the conversion activity of immobilized cells increased, possibly due to lysis, and this increase was related to the polymerization conditions and the incubation temperature for the L-sorbose conversion. The enzymatic activity and stability of the immobilized cells also depended on the physical form of the gel and the aeration levels. Addition of electron acceptors or addition of L-sorbosone to the medium reduced, while addition of neomycin, ampicillin, chloramphenicol, and tetracycline increased the stability of the enzymatic activity.  相似文献   

14.
Ketogulonicigenium vulgare DSM 4025, known as a 2-keto-L-gulonic acid producing strain from L-sorbose via L-sorbosone, surprisingly produced L-ascorbic acid from D-sorbitol, L-sorbose, L-gulose, and L-sorbosone as the substrate under a growing or resting condition. As the best result, K. vulgare DSM 4025 produced 1.37 g per liter of L-AA from 5.00 g per liter of L-sorbosone during 4 h incubation time at 30 degrees C under the resting cell condition having 5.70 g per liter of wet cells. The precursor of L-AA formation from D-sorbitol and L-sorbose, except for L-gulose, was thought to be the putative furanose form of L-sorbosone. This is the first time it is reported that bacteria can produce vitamin C via L-sorbosone.  相似文献   

15.
Summary It is shown, that bacteria of the strain Gluconobacter suboxydans entrapped in a polyacrylamide gel are capable to convert D-sorbitol to L-sorbose with a sufficiently high reaction rate. The kinetics of the studied process remains the same as it has been found for the case of free cells. Both semicontinuous and continuous patterns are accomplished.  相似文献   

16.
Oxidative fermentations have been well established for a long time, especially in vinegar and in L-sorbose production. Recently, information on the enzyme systems involved in these oxidative fermentations has accumulated and new developments are possible based on these findings. We have recently isolated several thermotolerant acetic acid bacteria, which also seem to be useful for new developments in oxidative fermentation. Two different types of membrane-bound enzymes, quinoproteins and flavoproteins, are involved in oxidative fermentation, and sometimes work with the same substrate but produce different oxidation products. Recently, there have been new developments in two different oxidative fermentations, D-gluconate and D-sorbitol oxidations. Flavoproteins, D-gluconate dehydrogenase, and D-sorbitol dehydrogenase were isolated almost 2 decades ago, while the enzyme involved in the same oxidation reaction for D-gluconate and D-sorbitol has been recently isolated and shown to be a quinoprotein. Thus, these flavoproteins and a quinoprotein have been re-assessed for the oxidation reaction. Flavoprotein D-gluconate dehydrogenase and D-sorbitol dehydrogenase were shown to produce 2-keto- D-gluconate and D-fructose, respectively, whereas the quinoprotein was shown to produce 5-keto- D-gluconate and L-sorbose from D-gluconate and D-sorbitol, respectively. In addition to the quinoproteins described above, a new quinoprotein for quinate oxidation has been recently isolated from Gluconobacter strains. The quinate dehydrogenase is also a membrane-bound quinoprotein that produces 3-dehydroquinate. This enzyme can be useful for the production of shikimate, which is a convenient salvage synthesis system for many antibiotics, herbicides, and aromatic amino acids synthesis. In order to reduce energy costs of oxidative fermentation in industry, several thermotolerant acetic acid bacteria that can grow up to 40 degrees C have been isolated. Of such isolated strains, some thermotolerant Acetobacter species were found to be useful for vinegar fermentation at a high temperature such 38-40 degrees C, where mesophilic strains showed no growth. They oxidized higher concentrations of ethanol up to 9% without any appreciable lag time, while alcohol oxidation with mesophilic strains was delayed or became almost impossible under such conditions. Several useful Gluconobacter species of thermotolerant acetic acid bacteria are also found, especially L-erythrulose-producing strains and cyclic alcohol-oxidizing strains. Gluconobacter frateurii CHM 43 is able to rapidly oxidize meso-erythritol at 37 degrees C leading to the accumulation of L-erythrulose, which may replace dihydroxyacetone in cosmetics. G. frateuriiCHM 9 is able to oxidize cyclic alcohols to their corresponding cyclic ketones or aliphatic ketones, which are known to be useful for preparing many different physiologically active compounds such as oxidized steroids or oxidized bicyclic ketones. The enzymes involved in these meso-erythritol and cyclic alcohol oxidations have been purified and shown to be a similar type of membrane-bound quinoproteins, consisting of a high molecular weight single peptide. This is completely different from another quinoprotein, alcohol dehydrogenase of acetic acid bacteria, which consists of three subunits including hemoproteins.  相似文献   

17.
Gluconate:NADP 5-oxidoreductase (GNO) from the acetic acid bacterium Gluconobacter oxydans subsp. oxydans DSM3503 was purified to homogeneity. This enzyme is involved in the nonphosphorylative, ketogenic oxidation of glucose and oxidizes gluconate to 5-ketogluconate. GNO was localized in the cytoplasm, had an isoelectric point of 4.3, and showed an apparent molecular weight of 75,000. In sodium dodecyl sulfate gel electrophoresis, a single band appeared corresponding to a molecular weight of 33,000, which indicated that the enzyme was composed of two identical subunits. The pH optimum of gluconate oxidation was pH 10, and apparent Km values were 20.6 mM for the substrate gluconate and 73 microM for the cosubstrate NADP. The enzyme was almost inactive with NAD as a cofactor and was very specific for the substrates gluconate and 5-ketogluconate. D-Glucose, D-sorbitol, and D-mannitol were not oxidized, and 2-ketogluconate and L-sorbose were not reduced. Only D-fructose was accepted, with a rate that was 10% of the rate of 5-ketogluconate reduction. The gno gene encoding GNO was identified by hybridization with a gene probe complementary to the DNA sequence encoding the first 20 N-terminal amino acids of the enzyme. The gno gene was cloned on a 3.4-kb DNA fragment and expressed in Escherichia coli. Sequencing of the gene revealed an open reading frame of 771 bp, encoding a protein of 257 amino acids with a predicted relative molecular mass of 27.3 kDa. Plasmid-encoded gno was functionally expressed, with 6.04 U/mg of cell-free protein in E. coli and with 6.80 U/mg of cell-free protein in G. oxydans, which corresponded to 85-fold overexpression of the G. oxydans wild-type GNO activity. Multiple sequence alignments showed that GNO was affiliated with the group II alcohol dehydrogenases, or short-chain dehydrogenases, which display a typical pattern of six strictly conserved amino acid residues.  相似文献   

18.
New quinoproteins in oxidative fermentation   总被引:1,自引:0,他引:1  
Several quinoproteins have been newly indicated in acetic acid bacteria, all of which can be applied to fermentative or enzymatic production of useful materials by means of oxidative fermentation. (1) D-Arabitol dehydrogenase from Gluconobacter suboxydans IFO 3257 was purified from the bacterial membrane and found to be a versatile enzyme for oxidation of various substrates to the corresponding oxidation products. It is worthy of notice that the enzyme catalyzes D-gluconate oxidation to 5-keto-D-gluconate, whereas 2-keto-D-gluconate is produced by a flavoprotein D-gluconate dehydrogenase. (2) Membrane-bound cyclic alcohol dehydrogenase was solubilized and purified for the first time from Gluconobacter frateurii CHM 9. When compared with the cytosolic NAD-dependent cyclic alcohol dehydrogenase crystallized from the same strain, the reaction rate in cyclic alcohol oxidation by the membrane enzyme was 100 times stronger than the cytosolic NAD-dependent enzyme. The NAD-dependent enzyme makes no contribution to cyclic alcohol oxidation but contributes to the reduction of cyclic ketones to cyclic alcohols. (3) Meso-erythritol dehydrogenase has been purified from the membrane fraction of G. frateurii CHM 43. The typical properties of quinoproteins were indicated in many respects with the enzyme. It was found that the enzyme, growing cells and also the resting cells of the organism are very effective in producing L-erythrulose. Dihydroxyacetone can be replaced by L-erythrulose for cosmetics for those who are sensitive to dihydroxyacetone. (4) Two different membrane-bound D-sorbitol dehydrogenases were indicated in acetic acid bacteria. One enzyme contributing to L-sorbose production has been identified to be a quinoprotein, while another FAD-containing D-sorbitol dehydrogenase catalyzes D-sorbitol oxidation to D-fructose. D-Fructose production by the oxidative fermentation would be possible by the latter enzyme and it is superior to the well-established D-glucose isomerase, because the oxidative fermentation catalyzes irreversible one-way oxidation of D-sorbitol to D-fructose without any reaction equilibrium, unlike D-glucose isomerase. (5) Quinate dehydrogenase was found in several Gluconobacter strains and other aerobic bacteria like Pseudomonas and Acinetobacter strains. It has become possible to produce dehydroquinate, dehydroshikimate, and shikimate by oxidative fermentation. Quinate dehydrogenase was readily solubilized from the membrane fraction by alkylglucoside in the presence of 0.1 M KCl. A simple purification by hydrophobic chromatography gave a highly purified quinate dehydrogenase that was monodispersed and showed sufficient purity. When quinate dehydrogenase purification was done with Acinetobacter calcoaceticus AC3, which is unable to synthesize PQQ, purified inactive apo-quinate dehydrogenase appeared to be a dimer and it was converted to the monomeric active holo-quinate dehydrogenase by the addition of PQQ.  相似文献   

19.
The fungal L-arabinose pathway consists of five enzymes, aldose reductase, L-arabinitol 4-dehydrogenase, L-xylulose reductase, xylitol dehydrogenase, and xylulokinase. All the genes encoding the enzymes of this pathway are known except for that of L-xylulose reductase (EC 1.1.1.10). We identified a gene encoding this enzyme from the filamentous fungus Trichoderma reesei (Hypocrea jecorina). The gene was named lxr1. It was overexpressed in the yeast Saccharomyces cerevisiae, and the enzyme activity was confirmed in a yeast cell extract. Overexpression of all enzymes of the L-arabinose pathway in S. cerevisiae led to growth of S. cerevisiae on L-arabinose; i.e., we could show that the pathway is active in a heterologous host. The lxr1 gene encoded a protein with 266 amino acids and a calculated molecular mass of 28 428 Da. The LXRI protein is an NADPH-specific reductase. It has activity with L-xylulose, D-xylulose, D-fructose, and L-sorbose. The highest affinity is toward L-xylulose (K(m) = 16 mM). In the reverse direction, we found activity with xylitol, D-arabinitol, D-mannitol, and D-sorbitol. It requires a bivalent cation for activity. It belongs to the protein family of short chain dehydrogenases. The enzyme is catalytically similar and homologous in sequence to a D-mannitol:NADP 2-dehydrogenase (EC 1.1.1.138).  相似文献   

20.
We developed a novel <50-microm thick nano-porous bi-layer latex coating for preserving Gluconobacter oxydans, a strict aerobe, as a whole cell biocatalyst. G. oxydans was entrapped in an acrylate/vinyl acetate co-polymer matrix (T (g) approximately 10 degrees C) and cast into 12.7-mm diameter patch coatings (cellcoat) containing approximately 10(9) CFU covered by a nano-porous topcoat. The oxidation of D-sorbitol to L-sorbose was used to investigate the coating catalytic properties. Intrinsic kinetics was studied in microbioreactors using a pH 6.0 D-sorbitol, phosphate, pyruvate (SPP) non-growth medium at 30 degrees C, and the Michaelis-Menten constants determined. By using a diffusion cell, cellcoat and topcoat diffusivities, optimized by arresting polymer particle coalescence by glycerol and/or sucrose addition, were determined. Cryo-FESEM images revealed a two-layer structure with G. oxydans surrounded by <40-nm pores. Viable cell density, cell leakage, and oxidation kinetics in SPP medium for >150 h were investigated. Even though the coatings were optimized for permeability, approximately 50% of G. oxydans viability was lost during cellcoat drying and further reduction was observed as the topcoat was added. High reaction rates per unit volume of coating (80-100 g/L x h) were observed which agreed with predictions of a diffusion-reaction model using parameters estimated by independent experiments. Cellcoat effectiveness factors of 0.22-0.49 were observed which are 20-fold greater than any previously reported for this G. oxydans oxidation. These nano-structured coatings and the possibility of improving their ability to preserve G. oxydans viability may be useful for engineering highly reactive adhesive coatings for multi-phase micro-channel and membrane bioreactors to dramatically increase the intensity of whole-cell oxidations.  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