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 15 毫秒
1.
Torula corallina, a strain presently being used for the industrial production of erythritol, has the highest erythritol yield ever reported for an erythritol-producing microorganism. The increased production of erythritol by Torula corallina with trace elements such as Cu2+ has been thoroughly reported, but the mechanism by which Cu2+ increases the production of erythritol has not been studied. This study demonstrated that supplemental Cu2+ enhanced the production of erythritol, while it significantly decreased the production of a major by-product that accumulates during erythritol fermentation, which was identified as fumarate by instrumental analyses. Erythrose reductase, a key enzyme that converts erythrose to erythritol in T. corallina, was purified to homogeneity by chromatographic methods, including ion-exchange and affinity chromatography. In vitro, purified erythrose reductase was significantly inhibited noncompetitively by increasing the fumarate concentration. In contrast, the enzyme activity remained almost constant regardless of Cu2+ concentration. This suggests that supplemental Cu2+ reduced the production of fumarate, a strong inhibitor of erythrose reductase, which led to less inhibition of erythrose reductase and a high yield of erythritol. This is the first report that suggests catabolite repression by a tricarboxylic acid cycle intermediate in T. corallina.  相似文献   

2.
The yeast Torula corallina is a strong erythritol producer that is used in the industrial production of erythritol. However, melanin accumulation during culture represents a serious problem for the purification of erythritol from the fermentation broth. Melanin biosynthesis inhibitors such as 3,4-dihydroxyphenylalanine and 1,8-dihydroxynaphthalene (DHN)-melanin inhibitors were added to the T. corallina cultures. Only the DHN-melanin inhibitors showed an effect on melanin production, which suggests that the melanin formed during the culturing of T. corallina is derived from DHN. This finding was confirmed by the detection of a shunt product of the pentaketide pathway, flaviolin, and elemental analysis. Among the DHN-melanin inhibitors, tricyclazole was the most effective. Supplementation with tricyclazole enhanced the production of erythritol while significantly inhibiting the production of DHN-melanin and DHN-melanin biosynthetic enzymes, such as trihydroxynaphthalene reductase. The erythrose reductase from T. corallina was purified to homogeneity by ion-exchange and affinity chromatography. Purified erythrose reductase was significantly inhibited in vitro in a noncompetitive manner by elevated levels of DHN-melanin. In contrast, the level of erythrose reductase activity was unaffected by increasing concentrations of tricyclazole. These results suggest that supplemental tricyclazole reduces the production of DHN-melanin, which may lead to a reduction in the inhibition of erythrose reductase and a higher yield of erythritol. This is the first report to demonstrate that melanin biosynthesis inhibitors increase the production of a sugar alcohol in T. corallina.  相似文献   

3.
Torula corallina (KCCM-10171) is a yeast strain that is currently used for the industrial production of erythritol and has the highest erythritol yield ever reported for an erythritol-producing microorganism. Production of erythritol in T. corallina is catalyzed by erythrose reductase, an enzyme that converts erythrose to erythritol using NADPH as a cofactor. In this study, NADPH-dependent erythrose reductase was purified to homogeneity from the newly isolated T. corallina. The relative molecular weight of the erythrose reductase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and size exclusion chromatography was 35.4 and 71.0 kDa, respectively, indicating that the enzyme is dimeric. This enzyme catalyzed both erythrose reduction and erythritol oxidation; both enzyme activities required NADP(H). The pH and temperature optima for erythrose reduction and erythritol oxidation were 6.0, 40 degrees C and 8.0, 45 degrees C, respectively. The sequence of the first 10 amino acids of this enzyme was N-V-K-N-F-Y-Q-P-N-D. The affinity (K(m)( )()= 7.12 mM) of the enzyme for erythrose was comparable to that of other known erythrose reductases, and the specificity for erythrose was very high, resulting in no production of other polyols, which may explain the high erythritol yield observed in this strain.  相似文献   

4.
The yeast Torula corallina is a strong erythritol producer that is used in the industrial production of erythritol. However, melanin accumulation during culture represents a serious problem for the purification of erythritol from the fermentation broth. Melanin biosynthesis inhibitors such as 3,4-dihydroxyphenylalanine and 1,8-dihydroxynaphthalene (DHN)-melanin inhibitors were added to the T. corallina cultures. Only the DHN-melanin inhibitors showed an effect on melanin production, which suggests that the melanin formed during the culturing of T. corallina is derived from DHN. This finding was confirmed by the detection of a shunt product of the pentaketide pathway, flaviolin, and elemental analysis. Among the DHN-melanin inhibitors, tricyclazole was the most effective. Supplementation with tricyclazole enhanced the production of erythritol while significantly inhibiting the production of DHN-melanin and DHN-melanin biosynthetic enzymes, such as trihydroxynaphthalene reductase. The erythrose reductase from T. corallina was purified to homogeneity by ion-exchange and affinity chromatography. Purified erythrose reductase was significantly inhibited in vitro in a noncompetitive manner by elevated levels of DHN-melanin. In contrast, the level of erythrose reductase activity was unaffected by increasing concentrations of tricyclazole. These results suggest that supplemental tricyclazole reduces the production of DHN-melanin, which may lead to a reduction in the inhibition of erythrose reductase and a higher yield of erythritol. This is the first report to demonstrate that melanin biosynthesis inhibitors increase the production of a sugar alcohol in T. corallina.  相似文献   

5.
Erythritol biosynthesis is catalyzed by erythrose reductase, which converts erythrose to erythritol. Erythrose reductase, however, has never been characterized in terms of amino acid sequence and kinetics. In this study, NAD(P)H-dependent erythrose reductase was purified to homogeneity from Candida magnoliae KFCC 11023 by ion exchange, gel filtration, affinity chromatography, and preparative electrophoresis. The molecular weights of erythrose reductase determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration chromatography were 38,800 and 79,000, respectively, suggesting that the enzyme is homodimeric. Partial amino acid sequence analysis indicates that the enzyme is closely related to other yeast aldose reductases. C. magnoliae erythrose reductase catalyzes the reduction of various aldehydes. Among aldoses, erythrose was the preferred substrate (K(m) = 7.9 mM; k(cat)/K(m) = 0.73 mM(-1) s(-1)). This enzyme had a dual coenzyme specificity with greater catalytic efficiency with NADH (k(cat)/K(m) = 450 mM(-1) s(-1)) than with NADPH (k(cat)/K(m) = 5.5 mM(-1) s(-1)), unlike previously characterized aldose reductases, and is specific for transferring the 4-pro-R hydrogen of NADH, which is typical of members of the aldo/keto reductase superfamily. Initial velocity and product inhibition studies are consistent with the hypothesis that the reduction proceeds via a sequential ordered mechanism. The enzyme required sulfhydryl compounds for optimal activity and was strongly inhibited by Cu(2+) and quercetin, a strong aldose reductase inhibitor, but was not inhibited by aldehyde reductase inhibitors and did not catalyze the reduction of the substrates for carbonyl reductase. These data indicate that the C. magnoliae erythrose reductase is an NAD(P)H-dependent homodimeric aldose reductase with an unusual dual coenzyme specificity.  相似文献   

6.

Background  

Erythrose reductase (ER) catalyzes the final step of erythritol production, which is reducing erythrose to erythritol using NAD(P)H as a cofactor. ER has gained interest because of its importance in the production of erythritol, which has extremely low digestibility and approved safety for diabetics. Although ERs were purified and characterized from microbial sources, the entire primary structure and the corresponding DNA for ER still remain unknown in most of erythritol-producing yeasts. Candida magnoliae JH110 isolated from honeycombs produces a significant amount of erythritol, suggesting the presence of erythrose metabolizing enzymes. Here we provide the genetic sequence and functional characteristics of a novel NADPH-dependent ER from C. magnoliae JH110.  相似文献   

7.
Of the vitamins tested, inositol was the most effective for erythritol production. To increase erythritol production by Torula sp., inositol and a related compound, phytic acid (myoinositol hexaphosphate), were added to the culture media. Erythritol production in the presence of phytic acid was greater than that in the presence of inositol, due to the synergistic effects of phosphate and inositol. Supplementation with phosphate and inositol increased cell growth, erythritol production, and the activity of erythrose reductase in cells. Inositol was a more effective stimulator of cell growth and erythritol production than was phosphate.  相似文献   

8.
Erythritol biosynthesis is catalyzed by erythrose reductase, which converts erythrose to erythritol. Erythrose reductase, however, has never been characterized in terms of amino acid sequence and kinetics. In this study, NAD(P)H-dependent erythrose reductase was purified to homogeneity from Candida magnoliae KFCC 11023 by ion exchange, gel filtration, affinity chromatography, and preparative electrophoresis. The molecular weights of erythrose reductase determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration chromatography were 38,800 and 79,000, respectively, suggesting that the enzyme is homodimeric. Partial amino acid sequence analysis indicates that the enzyme is closely related to other yeast aldose reductases. C. magnoliae erythrose reductase catalyzes the reduction of various aldehydes. Among aldoses, erythrose was the preferred substrate (Km = 7.9 mM; kcat/Km = 0.73 mM−1 s−1). This enzyme had a dual coenzyme specificity with greater catalytic efficiency with NADH (kcat/Km = 450 mM−1 s−1) than with NADPH (kcat/Km = 5.5 mM−1 s−1), unlike previously characterized aldose reductases, and is specific for transferring the 4-pro-R hydrogen of NADH, which is typical of members of the aldo/keto reductase superfamily. Initial velocity and product inhibition studies are consistent with the hypothesis that the reduction proceeds via a sequential ordered mechanism. The enzyme required sulfhydryl compounds for optimal activity and was strongly inhibited by Cu2+ and quercetin, a strong aldose reductase inhibitor, but was not inhibited by aldehyde reductase inhibitors and did not catalyze the reduction of the substrates for carbonyl reductase. These data indicate that the C. magnoliae erythrose reductase is an NAD(P)H-dependent homodimeric aldose reductase with an unusual dual coenzyme specificity.  相似文献   

9.
Fumarate reductase has been purified 100-fold to 95% homogeneity from the cytoplasmic membrane of Escherichia coli, grown anaerobically on a defined medium containing glycerol plus fumarate. Optimal solubilization of total membrane protein and fumarate reductase activity occurred with nonionic detergents having a hydrophobic-lipophilic balance (HLB) number near 13 and we routinely solubilized the enzyme with Triton X-100 (HLB number = 13.5). Membrane enzyme extracts were fractionated by hydrophobic-exchange chromatography on phenyl Sepharose CL-4B to yield purified enzyme. The enzyme whether membrane bound, in Triton extracts, or purified, had an apparent Km near 0.42 mM. Two peptides with molecular weights of 70 000 and 24 000, predent in 1:1 molar ratios, were identified by sodium dodecyl sulfate polyacrylamide slab-gel electrophoresis to coincide with enzyme activity. A minimal native molecular weight of 100 000 was calculated for fumarate reductase by Stephacryl S-200 gel filtration in the presence of sodium cholate. This would indicate that the enzyme is a dimer. The purified enzyme has low, but measurable, succinate dehydrogenase activity.  相似文献   

10.
In order to enhance erythritol production, mutants of Candida magnoliae DSM70638 were generated by ultraviolet and chemical mutagenesis. Erythritol productivity of samples was analyzed by TLC and HPLC with the refractive index detector. One of the mutants named mutant 12-2 gave a 2.4-fold increase in erythritol (20.32 g/L) and a 5.5-fold decrease in glycerol production compared to the wild strain. A sequence-based map of erythrose reductase gene in this mutant showed a replacement of the A321 by G321 that did not cause any amino acid exchange in protein structure. Therefore, the reason of higher erythritol production in C. magnoliae mutant 12-2 is probably the increase in expression of the open reading frame gene. This study revealed that a mutation or minor change in the sequence of genes involved in a production pathway can lead to a significant increase in protein translation.  相似文献   

11.
An unusual fumarate reductase was purified from cell extracts of Methanobacterium thermoautotrophicum and partially characterized. Two coenzymes previously isolated from cell extracts, 2-mercaptoethane-sulfonic acid (HS-CoM) and N-(7-mercaptoheptanoyl)threonine-O3-phosphate (HS-HTP), were established as direct electron donors for fumarate reductase. By measuring the consumption of free thiol, we determined that fumarate reductase catalyzed the oxidation of HS-CoM and HS-HTP; by the direct measurement of succinate and the heterodisulfide of HS-CoM and HS-HTP (CoM-S-S-HTP), we established that these compounds were products of the fumarate reductase reaction. A number of thiol-containing compounds did not function as substrates for fumarate reductase, but this enzyme had high specific activity when HS-CoM and HS-HTP were used as electron donors. HS-CoM and HS-HTP were quantitatively oxidized by the fumarate reductase reaction, and results indicated that this reaction was irreversible. Additionally, by measuring formylmethanofuran, we demonstrated that the addition of fumarate to cell extracts activated CO2 fixation for the formation of formylmethanofuran. Results indicated that this activation resulted from the production of CoM-S-S-HTP (a compound known to be involved in the activation of formylmethanofuran synthesis) by the fumarate reductase reaction.  相似文献   

12.
The enzyme NADH-fumarate reductase is not found in mammalian cells but it is present in several parasitic protozoa including Trypanosoma cruzi, the parasite that causes Chagas' disease. This study shows that the drug 2-mercaptopyridine-N-oxide (MPNO) inhibits NADH-fumarate reductase purified from T. cruzi (ID50 = 35 microM). When added to intact cells, MPNO inhibited the growth of T. cruzi epimastigotes in culture (ID50 = 0.08 microM) as well as the infection of mammalian myoblasts by T. cruzi trypomastigotes (ID50 = 20 microM). At a concentration of 2.4 microM, MPNO also inhibited the growth of amastigotes (intracellular dividing forms) in cultured mammalian myoblasts. Supplementation of culture media with 5 mM succinate, the product of fumarate reductase, partially protected against the inhibition of the growth of epimastigotes by MPNO. Moreover, MPNO inhibited the accumulation of succinate in cultures of epimastigotes, as measured by high performance liquid chromatography. Although MPNO may have other intracellular targets in addition to fumarate reductase, these results support the hypothesis that compounds which inhibit the enzyme fumarate reductase may be potential chemotherapeutic agents against Chagas' disease.  相似文献   

13.
Erythritol uptake and metabolism were compared in wild-type mycelium and a dome morphological mutant of the wood-rotting mushroom Schizophyllum commune. Wild-type mycelium utilized glucose, certain hexitols, and pentitols including ribitol, as well as d-erythrose, erythritol, and glycerol as sole carbon sources for growth. The dome mutant utilized all of these compounds except d-erythrose and erythritol. Erythritol- or glycerol-grown wild-type mycelium incorporated erythritol into various cellular constituents, whereas glucose-grown cells lagged considerably before initiation of erythritol uptake. This acquisition was inhibited by cycloheximide. Dome mycelium showed behavior similar to wild-type in uptake of erythritol after growth on glucose or glycerol, except that erythritol was not further catabolized. Enzymes of carbohydrate metabolism were compared in cell extracts of glucose-cultured wild-type mycelium and dome. Enzymes of hexose monophosphate catabolism, nicotinamide adenine dinucleotide (NAD)-dependent sugar alcohol dehydrogenases, and reduced nicotinamide adenine dinucleotide phosphate (NADPH)-coupled erythrose reductase were demonstrated in both. The occurrence of erythrose reductase was unaffected by the nature of the growth carbon source, showed optimal activity at pH 7, and generated NAD phosphate and erythritol as products of the reaction. Glycerol-, d-erythrose-, or erythritol-grown wild-type mycelium contained an NAD-dependent erythritol dehydrogenase absent in glucose cells. Erythritol dehydrogenase activity was optimal at pH 8.8 and produced erythrulose during NAD reduction. Glycerol-growth of dome mycelium induced the erythritol uptake system, but a functional erythritol dehydrogenase could not be demonstrated. Neither wild-type nor dome mycelium produced erythritol dehydrogenase during growth on ribitol. Erythritol metabolism in wild-type cells of S. commune, therefore, involves an NADPH-dependent reduction of d-erythrose to produce erythritol, followed by induction of an NAD-coupled erythritol dehydrogenase to form erythrulose. A deficiency in erythritol dehydrogenase rather than permeability barriers explains why dome cannot employ erythritol as sole carbon source for mycelial growth.  相似文献   

14.
Abstract

The rapid and effective purification of soluble fumarate reductase from baker's yeast achieved by Blue Sepharose CL–6B chromatography. Cibacron Blue F3GA, the chromophore of Blue Sepharose, inhibited the activity of fumarate reductase. The enzyme bound to the column was selectively eluted by flavin adenine dinucleotide (FAD), flavin mononucleotide (FMN) or riboflavin. The purified enzyme was essentially homogeneous as indicated by polyacrylamide gel electrophoresis under non-denaturing conditions and under denaturing conditions in sodium dodecylsulfate. By this procedure, the enzyme could be rapidly purified with high yield from yeast cells.  相似文献   

15.
The purpose of the present investigation was to produce erythritol by Yarrowia lipolytica mutant without any by-products. Mutants of Y. lipolytica were generated by ultra-violet for enhancing erythrose reductase (ER) activity and erythritol production. The mutants showing the highest ER activity were screened by triphenyl tetrazolium chloride agar plate assay. Productivity of samples was analyzed by thin-layer chromatography and high-performance liquid chromatography equipped with the refractive index detector. One of the mutants named as mutant 49 gave maximum erythritol production without any other by-products (particularly glycerol). Erythritol production and specific ER activity in mutant 49 increased to 1.65 and 1.47 times, respectively, in comparison with wild-type strain. The ER gene of wild and mutant strains was sequenced and analyzed. A general comparison of wild and mutant gene sequences showed the replacement of Asp270 with Glu270 in ER protein. In order to enhance erythritol production, we used a three component-three level-one response Box–Behnken of response surface methodology model. The optimum medium composition for erythritol production was found to be (g/l) glucose 279.49, ammonium sulfate 9.28, and pH 5.41 with 39.76 erythritol production.  相似文献   

16.
17.
18.
In all trypanosomatids, including Trypanosoma brucei, glycolysis takes place in peroxisome-like organelles called glycosomes. These are closed compartments wherein the energy and redox (NAD(+)/NADH) balances need to be maintained. We have characterized a T. brucei gene called FRDg encoding a protein 35% identical to Saccharomyces cerevisiae fumarate reductases. Microsequencing of FRDg purified from glycosome preparations, immunofluorescence, and Western blot analyses clearly identified this enzyme as a glycosomal protein that is only expressed in the procyclic form of T. brucei but is present in all the other trypanosomatids studied, i.e. Trypanosoma congolense, Crithidia fasciculata and Leishmania amazonensis. The specific inactivation of FRDg gene expression by RNA interference showed that FRDg is responsible for the NADH-dependent fumarate reductase activity detected in glycosomal fractions and that at least 60% of the succinate secreted by the T. brucei procyclic form (in the presence of d-glucose as the sole carbon source) is produced in the glycosome by FRDg. We conclude that FRDg plays a key role in the energy metabolism by participating in the maintenance of the glycosomal NAD(+)/NADH balance. We have also detected a significant pyruvate kinase activity in the cytosol of the T. brucei procyclic cells that was not observed previously. Consequently, we propose a revised model of glucose metabolism in procyclic trypanosomes that may also be valid for all other trypanosomatids except the T. brucei bloodstream form. Interestingly, H. Gest has hypothesized previously (Gest, H. (1980) FEMS Microbiol. Lett. 7, 73-77) that a soluble NADH-dependent fumarate reductase has been present in primitive organisms and evolved into the present day fumarate reductases, which are quinol-dependent. FRDg may have the characteristics of such an ancestral enzyme and is the only NADH-dependent fumarate reductase characterized to date.  相似文献   

19.
For expression of Bacillus stearothermophilus NCIB 8924 leucine aminopeptidase II (LAP II) in Escherichia coli regulated by a T5 promoter, the gene was amplified by polymerase chain reaction and cloned into expression vector pQE-32 to generate pQE-LAPII. The His(6)-tagged enzyme was overexpressed in IPTG-induced E. coli M15 (pQE-LAPII) as a soluble protein and was purified to homogeneity by nickel-chelate chromatography to a specific activity of 425 U/mg protein with a final yield of 76%. The subunit molecular mass of the purified protein was estimated to be 44.5 kDa by SDS-PAGE. The temperature and pH optima for the purified protein were 60 degrees C and 8.0, respectively. Under optimal condition, the purified enzyme showed a marked preference for Leu- p-nitroanilide, followed by Arg- and Lys-derivatives. The His(6)-tagged enzyme was stimulated by Co(2+) ions, but was strongly inhibited by Cu(2+) and Hg(2+) and by the chelating agents, DTT and EDTA. The EDTA-treated enzyme could be reactivated with Co(2+) ions, indicating that it is a cobalt-dependent exopeptidase. Taking the biochemical characteristics together, we found that the recombinant LAP II exhibits no important differences from those properties described for the native enzyme.  相似文献   

20.
In this study, we investigated inhibitory effects of some metal ions on human erythrocyte glutathione reductase. For this purpose, initially human erythrocyte glutathione reductase was purified 1051-fold in a yield of 41% by using 2', 5'-ADP Sepharose 4B affinity gel and Sephadex G-200 gel filtration chromatography. SDS polyacrylamide gel electrophoresis was done in order to control the purification of enzyme. SDS polyacrylamide gel electrophoresis showed a single band for enzyme. A constant temperature (4 degrees C) was maintained during the purification process. Enzyme activity was determined with the Beutler method by using a spectrophotometer at 340 nm. Hg(2+), Cd(2+), Pb(2+), Cu(2+), Fe(3+) and Al3+ exhibited inhibitory effects on the enzyme in vitro. K(i) constants and IC(50) values for metal ions were determined by Lineweaver-Burk graphs and plotting activity % vs. [I]. IC(50) values of Pb(2+), Hg(2+), Cu(2+), Cd(2+), Fe(3+) and Al(3+) were 0.011, 0.020, 0.0252, 0.0373, 0.209 and 0.229 mM, and the Ki constants 0.0254+/-0.0027, 0.0378+/-0.0043, 0.0409+/-0.0048, 0.0558+/-0.0083, 0.403+/-0.043 and 1.137+/-0.2 mM, respectively. While Pb(2+), Hg(2+), Cd(2+) and Fe(3+) showed competitive inhibition, others displayed noncompetitive inhibition.  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